arXiv:2012.15303v1 [math.GR] 30 Dec 2020 mi address Email mi address Email mi address Email M M I SIU FÜR NSTITUT ATHEMATICS ATHEMATICS onain fgoercapoiaegoptheory group approximate geometric of Foundations : : : [email protected] [email protected] [email protected] A D D GBAUND LGEBRA EPARTMENT EPARTMENT T Z ETH , U , G EOMETRIE IEST OF NIVERSITY ate Cordes Matthew oisHartnick Tobias ÜRICH eaToni´cVera I,G KIT, , S , R WITZERLAND IJEKA ERMANY C , ROATIA 2020 Subject Classification. Primary: 20N99; Secondary: 20F65, 20F67, 20F69 Key words and phrases. Approximate group, geometric group theory, limit set, asymptotic dimension, Morse boundary

ABSTRACT. We develop the foundations of a geometric theory of countably-infinite approximate groups, extending work of Björklund and the second-named author. Our theory is based on the notion of a quasi- isometric quasi-action (qiqac) of an approximate group on a . More specifically, we introduce a geometric notion of finite generation for approximate group and prove that every geometrically finitely-generated approximate group admits a geometric qiqac on a proper geodesic metric space. We then show that all such spaces are quasi-isometric, hence can be used to associate a canonical QI type with every geometrically finitely-generated approximate group. This in turn allows us to define geometric invariants of approximate groups using QI invariants of metric spaces. Among the invariants we consider are asymptotic dimension, finiteness properties, numbers of ends and growth type. A particular focus is on qiqacs on hyperbolic spaces. Our strongest results are obtained for approxi- mate groups which admit a geometric qiqac on a proper geodesic hyperbolic space. For such “hyperbolic approximate groups” we establish a number of fundamental properties in analogy with the case of hyper- bolic groups. For example, we show that their asymptotic dimension is one larger than the topological dimension of their Gromov boundary and that - under some mild assumption of being “non-elementary” - they have exponential growth and act minimally on their Gromov boundary. We also show that every non-elementary hyperbolic approximate group of asymptotic dimension 1 is quasi-isometric to a finitely- generated non-abelian free group. We also study proper qiqacs of approximate group on hyperbolic spaces, which are not necessarily cobounded. Here one focus is on convex cocompact qiqacs, for which we provide several geometric char- acterizations à la Swenson. In particular we show that all limit points of a convex cocompact qiqac are conical, whereas in general the conical limit set contains additional information. Finally, using the theory of Morse boundaries, we extend some of our results concerning qiqacs on hyperbolic spaces to qiqacs on proper geodesic metric spaces with non-trivial Morse boundary. Throughout this book we emphasize different ways in which definitions and results from geometric group theory can be extended to approximate groups. In cases where there is more than one way to extend a given definition, the relations between different possible generalizations are carefully explored. Contents

Chapter 1. Introduction 5 1. A historical perspective 5 2. A mantra for the reader 7 3. Three approaches to geometric group theory 8 4. Fundamental notions of geometric approximate group theory 9 5. Quasi-isometric quasi-actions on hyperbolic spaces 11 Acknowledgements 13

Chapter 2. Basic constructions of approximate groups 15 1. Categories of filtered groups 15 2. Approximate groups and related notions 16 3. Morphisms between approximate groups 19 4. Intersections and images under local (quasi-)morphisms 21 5. Pre-images and (partial) kernels 22 6. Extensions of approximate groups 26 7. Model sets and approximate lattices 27

Chapter 3. Large-scale of approximate groups 29 1. The coarse space associated with a countable approximate group 29 2. Coarse invariants of countable approximate groups 31 3. External and internal QI type 35 4. Distortion 38 5. Growth of approximate groups 39

Chapter 4. Geometric quasi-actions of approximate groups 43 1. Isometric actions of approximate groups 43 2. Quasi-isometric quasi-actions (qiqacs) 46 3. Geometric quasi-actions 49 4. Left-regular quasi-actions 51 5. The Milnor-Schwarz lemma and its variants 54

Chapter 5. Hyperbolic approximate groups and their boundaries 59 1. Hyperbolic approximate groups 59 2. Hyperbolic approximate groups are visual 60 3. Boundaries of hyperbolic approximate groups 61 4. Embeddings of hyperbolic approximate groups 65 5. Non-elementary hyperbolic approximate groups 68

Chapter 6. Limit sets of quasi-isometric quasi-actions 71 1. Morse and Gromov limit sets of approximate groups 71 2. Conical limit points 74 3. Qiqacs with stable quasi-orbits 78 4. Convex cocompact qiqacs 78

3 4 CONTENTS

Chapter 7. Asymptotic dimension 83 1. Summary of results 83 2. Background on dimension theory 84 3. An asymptotic Hurewicz mapping theorem for countable approximate groups 85 4. Asymptotic dimension of non-finitely generated countable approximate groups 87 5. Asymptotic dimension vs. dimension of the boundary 89 6. Preliminaries for the proof of Theorem 7.21 90 7. Proof of Theorem 7.21 92

Chapter 8. Growth in non-elementary hyperbolic approximate groups 99 1. The case of asymptotic dimension 1 99 2. The case of higher asymptotic dimension 100

Appendix A. Background from large-scale geometry 105 1. Notations concerning (pseudo-)metric spaces 105 2. Coarse maps and coarse equivalences 106 3. Quasi-isometries, bi-Lipschitz maps and rough isometries 107 4. Geodesics and quasi-geodesics in pseudo-metric spaces 108 5. Coarsely connected and large-scale geodesic spaces 110 6. Spaces of coarse bounded geometry 111 7. Coarse ends of metric spaces 112

Appendix B. Some notions from geometric group theory 115 1. Left-invariant (pseudo-)metrics on groups 115 2. The coarse class of a countable group 116 3. The canonical QI class of a finitely-generated group 117 4. Free groups and of words 119 5. Quasimorphisms and quasi-homomorphisms 124 6. Reconstruction theorems for homogeneous quasimorphisms 127

Appendix C. Morse hyperbolic spaces and their Gromov boundaries 129 1. Gromov, Rips and Morse hyperbolic spaces 129 2. Boundaries of Gromov and Morse hyperbolic spaces 131 3. Visual metrics 133

Appendix D. Morse boundaries of proper geodesic metric spaces 135 1. Definition and basic properties 135 2. Ideal Morse triangles 138 3. Limit sets and stable subspaces 138 4. Weak hulls 140

Appendix. Index 143

Appendix. Bibliography 149 CHAPTER 1

Introduction

The goal of this book is to extend a number of foundational results in geometric group theory from finitely-generated groups to “finitely-generated” approximate groups, and thus to carry out a program that was suggested by Michael Björklund and the second-named author in [BH18a, Chapter 3]. In this introduction we first explain some of the origins of this program. We then point out some of the difficulties in its implication and survey a few sample results from the new theory.

1. A historical perspective

REMARK 1.1 (From algebra to quasi-algebra). From the syntactic point of view (as developed in the early 20th century), algebra is concerned with “sets with operations” (e.g. group, rings, fields). These operations transform a number of elements of the underlying set to another element of this set. To relate different sets with operations (of the same type) one then studies maps between the underlying sets which preserve the given operations. It was already pointed out by Ulam in the 1940s (and later popularized in his book [Ula60]) that this setting may be too rigid to encompass many phenomena of interest, both within mathematics and in mathematical applications. Instead of studying functions which satisfy strict functional equations such as “f(g + h) = f(g)+ f(h)”, Ulam suggested to investigate functions which only satisfy such functional equations “approximately”. This idea has since manifested itself in many different areas of mathematics, from stability problems in functional analysis to the study of group quasimorphisms and bounded cohomology - see [BOT13, FK16, HS16] for some recent examples. An even more radical idea is to consider structures, i.e. sets with operations, in which the oper- ations themselves do not produce elements of the underlying set, but rather elements of a larger set which are in some sense “close” to the original set. It was this latter step of “quasifying not only the morphisms, but also the objects”, which (in the early 21st century) lead to the creation of a new kind of quasi-algebra. Approximate groups are among the main protagonists of this new world, and can be defined as follows (see Definition 2.8).

DEFINITION 1.2. Let G be a group and k N. A subset Λ G is called a k-approximate subgroup ∈ ⊂ of G if it is symmetric (i.e. x Λ = x−1 Λ) and contains the identity, and if there is a finite set ∈ ⇒ ∈ F G of cardinality k such that for all x, y Λ there exist z Λ and f F with xy = zf. ⊂ ∈ ∈ ∈ In this case we denote by Λ∞ the subgroup of G generated by Λ and call the pair (Λ, Λ∞) a k- approximate group. It is interesting to compare the genesis and history of approximate groups to the genesis and history of groups:

REMARK 1.3 (Some milestones in the history of group theory). While the modern definition of a group is less than 150 years old, specific classes of groups have been studied since antiquity. Even after the definition of abstract groups, the focus of the theory has always been on specific classes of examples, often arising from other mathematical areas. Some important historical milestones in the development of modern group theory can be listed as follows1:

1We admit that these milestones are chosen with a certain bias towards concepts relevant to approximate groups and geometric group theory.

5 6 1. INTRODUCTION

Finite groups and permutation groups became more and more important in • and algebra during the 18th and 19th century, due to work of Gauß, Abel, Galois, Kummer, Dedekind and many others. Infinite abelian groups, in particular lattices in Euclidean spaces (of low dimensions), have been • studied due to their connection with crystals; an early highlight was the theory of Bravais lattices towards the middle of the 19th century. Lie groups and more generally locally compact groups became the basis for the modern formu- • lation of geometry, due to Riemann, Lie, Klein, Poincaré and others. Linear groups, in particular discrete subgroups of Lie groups (such as Kleinian and Fuchsian • groups), were among the first countably-infinite non-abelian groups that were studied sys- tematically. General (i.e. potentially non-linear) finitely generated groups were studied in the first half of the • 20th century by Dehn, von Neumann, Tarski, Freudenthal and others, but became dominant in group theory only during the second half of the 20th century, following pioneering work of Gromov, Rips and others: this was the “geometric group theory revolution”. More recently, the geometric group theory revolution has also attacked compactly-generated • locally compact groups; the book [CdlH16] has popularized this second wave of the revolution enormously.

REMARK 1.4 (Some milestones in the history of approximate group theory). Just as in the group case, the theory of approximate groups is much older than the actual definition, which was coined by Terry Tao and first appeared in print in 2008 [Tao08]. As in the group case, approximate groups arose in at least three different areas of mathematics simultaneously:

Finite approximate groups play a central role in modern additive combinatorics; implicitly they • already appear in the pioneering works of Plünnecke [Plü69], Fre˘ıman [Fre73], Erdös and Szemerédi [ES83], Ruzsa [Ruz99] and others. More recent works by Bourgain and Gamburd [BG08], respectively Helfgott [Hel08], relate finite approximate groups to superstrong ap- proximation, respectively expansion in finite simple groups. The general structure theory of finite approximate groups due to Breuillard, Green and Tao [BGT12] is one of the corner- stones of the modern theory. Independently, approximate groups occurred in the theory of mathematical quasi-crystals. Ori- • ginally developed by Meyer [Mey72] with a view towards application in number theory and harmonic analysis, they rose to prominence after the discovery of materials with quasi- crystalline structure in the 1980s (see [BG13] for a bibliography with hundreds of references). In the language of the present book, mathematical quasi-crystals are (translates of) uniform approximate lattices in abelian locally compact groups. The connection between approximate groups and locally compact groups goes back even fur- • ther. Compact symmetric identity neighborhoods in locally compact groups and Lie groups play a central role in connection with the solution to Hilbert’s fifth problem [MZ52, Gle52]. These neighborhoods are approximate subgroups, and their study has greatly influenced the modern theory of approximate groups (see [Tao14]). More recently, a structure theory of locally compact approximate groups has been developed by Carolino [Car15], and algebraic approximate groups have been classified by Björklund, Hartnick and Stulemeijer [BHS19]. Discrete approximate subgroups of Lie groups appear implicitly already in work of Chifan and • Ioana on relative Property (T) [CI11]. The theory of model sets in Lie groups (and more general lcsc groups), due to Björklund, Hartnick and Pogorzelski [BHP18] was a major motivation for the introduction of approximate lattices by Björklund and Hartnick [BH18a], which started a whole new line of research [BH21, BH18b, BHS19]. Arithmetic model sets have recently received much attention due to arithmeticity theorems of Machado [Mac20a] and Hrushovski [Hru20]. 2. A MANTRA FOR THE READER 7

Comparing this list to the list in Remark 1.3, what is still missing is a general geometric theory of finitely- generated approximate groups. The first steps towards such a theory were taken in Chapter 3 of[BH18a], but the focus there was mostly on the connection between isometric actions of approximate groups and approximate lattices. It was then suggested that a more general version of “geometric approximate group theory” should be developed. The present book is meant as a first approximation of what such a theory might look like.

2. A mantra for the reader Given that this book is primarily concerned with “finitely-generated” approximate groups, it is important to point out that even defining what a “finitely-generated” approximate group is, is a non- trivial task: Proposition B.21 below lists eight different characterizations of finitely-generated groups, all of which can also be formulated for approximate groups (some in several ways). In this wider con- text, the conditions will no longer be equivalent, and in fact there does not seem to be a single “correct” definition of a finitely-generated approximate group - depending on which results one would like to prove, one or the other condition may be preferable. This situation is somewhat prototypical for this whole book: All of our definitions and even the majority of our theorems and their proofs specialize to well-known classical definitions, theorems and proofs in the group case. At a first glance it might thus seem that geometric group theory generalizes in a straight-forward way to approximate groups. There are, however, at least three reasons why the generalizations presented here are more involved than meets the eye. Firstly, there are of course many classical results in geometric group theory which do not generalize to approximate groups; the fact that these are not mentioned in this book leads to a kind of “survivor bias” that the reader should keep in mind. We certainly do not claim that every result in geometric group theory has a counterpart for approximate groups. Secondly, almost every definition in the group case admits more than one reasonable generaliza- tion to the approximate group case. In fact, it was one of the main tasks in writing this book to find the right combination of definitions to develop a non-trivial theory. If our results look very similar to classical results in geometric group theory, then this is partly due to the fact that we have chosen our definitions accordingly. With a different, but equally “valid” choice of definitions, many of the classical results considered below would not generalize. Finally, one has to keep in mind that certain basic group theoretic arguments are not available to us. A shocking example is the fact that the intersection of two approximate subgroups need not be an approximate subgroup. In many of these situations there is a work-around (e.g. the intersection of the squares of two approximate subgroups is an approximate subgroup), but these work-arounds may interact with each other in unpleasant ways. Therefore, for most of our proofs, we did not start out from a known group-theoretic proof and then work by adding a number of work-arounds. Rather, we would often try to find group-theoretic arguments which generalize well to approximate groups and then try to reprove classical results using only (or mostly) these arguments. In other cases we were able to find the “correct” group-theoretical proof, which “generalizes easily” to approximate groups - usually after first considering different proofs, which do not generalize. For all of these reasons, the step from geometric group theory to geometric approximate group theory is larger than a first glance at our results (and even their proofs) may suggest. Throughout this book, we will choose the old words of Garret Birkhoff as our mantra:

“In carrying out the program [. . . ], it has been found essential to proceed with great care. For at every step we have a variety of definitions to choose from, and it is only after a detailed examination that we can judge of their relative fruitfulness.” –[Bir34] 8 1. INTRODUCTION

3. Three approaches to geometric group theory What is geometric approximate group theory? Following our (i.e. Birkhoff’s) mantra we should look at different formulations of geometric group theory and then decide, which of these formulations can be adapted most “fruitfully” to approximate groups. What all formulations of geometric group theory have in common is that, with every group from a certain class (typically finitely-generated, or at least countable, but possibly also just locally compact) one associates a class of metric spaces. Any invariant of this class of metric spaces is then considered to be a “geometric invariant” of the group. REMARK 1.5 (Geometric group theory via Cayley graphs). The most concrete approach, favoured by most elementary textbooks, is to start from a finitely-generated group G. With every finite generating set S of G one can then associate the corresponding Cayley graph (or “Dehn Gruppenbild”) Cay(G, S), which is a locally finite connected graph and hence can be considered as a proper geodesic metric space.2 Cayley graphs with respect to different finite generating systems are quasi-isometric to each other (in fact, even bi-Lipschitz), and hence quasi-isometry invariants of their Cayley graphs yield geometric invariants of finitely-generated groups. REMARK 1.6 (Geometric group theory via geometric actions). While the Cayley graph approach to geometric group theory is the most elementary one, historically geometric group theory did not arise from the study of group actions on graphs, but rather from the actions of Fuchsian and Kleinian groups on the corresponding hyperbolic spaces. To include such actions into the theory, one observes that both the action of a finitely-generated group on its Cayley graph and, say, the action of a surface group on the hyperbolic plane are proper and cobounded; such actions are sometimes called geometric. It is a fundamental result of geometric group theory, discovered by Schwarz and re-discovered by Milnor, that all proper geodesic metric spaces on which a given group G acts geometrically are quasi- isometric to each other. Quasi-isometry invariants of such spaces thus yield geometric invariants of the acting group, and this provides a second approach to geometric group theory. It turns out a posteriori that the class of groups which admit geometric actions on proper geodesic metric spaces coincides with the class of finitely-generated groups, hence the scope of this second approach is not larger than the scope of the first approach. However, from the approximate group point of view this has the advantage that it does not require the notion of a generating set, which is problematic for approximate groups. REMARK 1.7 (The pseudo-metric approach to geometric group theory). There is actually a third approach to geometric group theory, which is less well-known, but works in larger generality. This approach is based on the observation that if G is a countable group and d, d′ are any two left-invariant proper pseudo-metrics on G, then (G, d) and (G, d′) are coarsely equivalent, hence any coarse invariant of (G, d) is a geometric invariant of G. Note that if G acts geometrically on a metric space (X, dX ) and o X is an arbitrary basepoint, then we can define a proper left-invariant metric d on G by ∈ d(g,h) := dX (g.o, h.o). Then (G, d) is coarsely equivalent to (X, d ) via the orbit map g g.o, and hence the coarse invariants X 7→ of (G, d) coincide with the coarse invariants of (X, dX ). Thus, even in this third approach, geometric actions of G can be used to compute geometric invariants of G, as in the second approach. The main disadvantage of this approach is that for different proper left-invariant metrics d, d′ the spaces (G, d) and (G, d′) will only be coarsely equivalent, rather than quasi-isometric, hence the geometric invariants which can be defined for general countable groups by this approach are rather limited. This problem can be remedied using the basic observation (first recorded by Gromov in [Gro93], who calls it a “trivial lemma”) that every coarse equivalence between geodesic metric spaces is auto- matically a quasi-isometry. Using this observation, we may thus proceed as follows:

2See Appendix A for our terminology concerning metric spaces and their large-scale geometry. 4. FUNDAMENTAL NOTIONS OF GEOMETRIC APPROXIMATE GROUP THEORY 9

Denote by [G]c the class of all metric spaces which are coarsely equivalent to (G, d) for some (hence any) proper left-invariant metric on G and define a subclass [G] [G] by ⊂ c [G] := X [G] X is geodesic . { ∈ c | } Then all representatives of [G] are mutually quasi-isometric by Gromov’s trivial lemma, and hence quasi-isometry invariants of such representatives yield geometric invariants of G, provided [G] = . 6 ∅ Somewhat surprisingly, it turns out that the condition [G] = is also equivalent to G being finitely- 6 ∅ generated, and in this case every Cayley graph of G and every proper geodesic metric space X on which G acts geometrically is contained in the class [G]. This shows that the third approach yields the same geometric invariants for finitely-generated groups as the other two approaches. However, as we will see, it is much better adapted to possible generalizations to approximate groups.

4. Fundamental notions of geometric approximate group theory We now explain how the pseudo-metric approach to geometric group theory carries over to ap- proximate groups. (We will see later that the other two approaches can also be made to work, but this requires additional efforts.) Throughout this section let G be a group and let Λ G be a countable approximate subgroup. ⊂ We denote by Λ∞ < G the (countable) subgroup generated by Λ so that (Λ, Λ∞) is an approximate ∞ ∞ group. If H is any countable subgroup of G containing Λ (for example, Λ itself) and dH is a proper, left-invariant pseudo-metric on H, then we refer to the restriction of dH to Λ as an admissible metric on Λ. We then have the following basic observation (see Lemma 3.1):

LEMMA 1.8. If d, d′ are any two admissible metrics on Λ, then (Λ, d) and (Λ, d′) are coarsely equivalent. 

From now on we fix an admissible metric do on Λ. We introduce (see Definition 3.2):

DEFINITION 1.9. The coarse class of the approximate group (Λ, Λ∞) is defined as [Λ] := X metric space X is coarsely-equivalent to (Λ, d ) . c { | o } REMARK 1.10 (Invariance properties of the coarse class). By Lemma 1.8, the coarse class [Λ]c is ∞ independent of the choice of do. In fact, it depends only on the isomorphism class of (Λ, Λ ) in a suitable sense. Defining isomorphisms of approximate groups is actually a non-trivial task. The most naive no- tion is that of “global isomorphisms” (i.e. isomorphisms of the ambient groups preserving the approx- imate subgroups), and [Λ]c is certainly invariant under such global isomorphisms. However, there are also much weaker notions of isomorphism for approximate groups, and [Λ]c is invariant under most of these. In fact, anticipating the terminology of Chapter 2, we are going to show that [Λ]c is invari- ant under all 2-local quasi-isomorphisms, in particular under all 2-local isomorphisms and Freiman 3-isomorphisms (see Corollary 3.7). If now is any coarse invariant of metric spaces, then we can define the corresponding geometric I invariant of Λ (see Remark 3.10), by (Λ) := (X), where X [Λ] is arbitrary. (4.1) I I ∈ c EXAMPLE 1.11 (Asymptotic dimension of approximate groups). A numerical coarse invariant of metric spaces is Gromov’s asymptotic dimension ([Gro93], see Definition 3.11 below). Using (4.1) we may thus define the asymptotic dimension asdim Λ of an approximate group (Λ, Λ∞) (see Definition 3.13). By Remark 1.10, asymptotic dimension is invariant under 2-local quasi-isomorphisms. As in the group case, most of the finer invariants of geometric group theory require us to consider geodesic representatives of [Λ]c. For technical reasons, it is more convenient to actually work with large-scale geodesic representatives, i.e. representatives which are quasi-isometric to a geodesic metric space. This leads to the following definition (see Definition 3.30): 10 1. INTRODUCTION

DEFINITION 1.12. The approximate group (Λ, Λ∞) is called geometrically finitely-generated if [Λ] := X [Λ] X large-scale geodesic = . int { ∈ c | } 6 ∅ ∞ In this case, [Λ]int is called the internal QI type of (Λ, Λ ).

REMARK 1.13 (Apogees and generalized Cayley graphs). Assume that (Λ, Λ∞) is geometrically finitely-generated. It then follows from Gromov’s trivial lemma that any two representatives of [Λ]int are quasi- isometric to each other. Any quasi-isometry invariant of geodesic metric spaces can thus be used ∞ to define a geometric invariant of (Λ, Λ ) by applying it to a representative from [Λ]int. For practical computations it is then important to find good representatives. One can show that [Λ]int can be represented by a connected proper metric space, and we call any such space an apogee for (Λ, Λ∞) (see Definition 3.33). In fact, every geometrically finitely-generated group admits an apogee which is a locally finite graph. Such graphs are then the natural generaliza- tions of Cayley graphs in our context. It is an interesting question which spaces can arise as apogees of approximate groups. We will see that any such space must be quasi-cobounded and of coarse bounded geometry, and actually most of our results concerning approximate groups apply in the generality of quasi-cobounded proper geodesic spaces of coarse bounded geometry.

REMARK 1.14 (Internal distortion). A new phenomenon in geometric approximate group theory, which has no counterpart in geometric group theory, is internal distortion. Let us say that an approxi- mate group (Λ, Λ∞) is algebraically finitely-generated if Λ∞ is finitely-generated as a group. In this case, every finite-generating set S for Λ∞ defines a word metric d on Λ∞, and we refer to d as an ex- S S|Λ×Λ ternal metric on Λ. Any two external metrics are quasi-isometric, and hence we can define the external QI type of Λ (see Definition 3.27), as [Λ] = X [Λ] X is quasi-isometric to (Λ, d ) . ext { ∈ c | S|Λ×Λ } If Λ is a countable group, then it is geometrically finitely generated if and only if it is (algebraically) finitely-generated, and in this case [Λ]int = [Λ]ext. However, while every geometrically finitely gener- ated approximate group is algebraically finitely generated (see Corollary 4.39), the converse does not hold (see Example 3.37). Moreover, even if (Λ, Λ∞) is geometrically (and hence algebraically) finitely- generated, it may happen that [Λ] = [Λ] (see Example 3.39). We then say that (Λ, Λ∞) is distorted; int 6 ext otherwise it is called undistorted (see Definition 3.38). The geometric theory of undistorted approximate groups mirrors closely the geometric theory of finitely-generated groups. Examples of such approximate groups are given by uniform model sets, or more generally uniform approximate lattices. However, many interesting examples of approximate groups (in particular, quasikernels of many naturally occurring quasimorphisms) are distorted, hence it is important to develop the basic theory without this assumption.

REMARK 1.15 (Isometric actions of approximate groups). In view of Remark 1.6, one could say that geometric group theory is chiefly concerned with certain isometric actions on proper geodesic metric spaces. It turns out, however, that this point of view generalizes poorly to approximate groups. Let us elaborate on this point: There is no problem in defining an isometric action of an approximate group (Λ, Λ∞) on a metric space X - it is just an isometric action of Λ∞ on X. There is also no problem in defining what it means for such an action to be proper, cobounded or geometric: properness means that the map Λ X X X, (λ, x) (x, λ.x) is proper and coboundedness means that the “quasi-orbit” λ.x × → × 7→ { | λ Λ, x X is relatively dense in X for some (hence any) x X. ∈ ∈ } ∈ However, while every finitely-generated group admits a geometric isometric action on a proper geodesic space (e.g. any Cayley graph), this is not the case for approximate groups. In fact, internal distortion turns out to be an obstruction (see Corollary 4.40). 5. QUASI-ISOMETRIC QUASI-ACTIONS ON HYPERBOLIC SPACES 11

PROPOSITION 1.16. Every geometrically finitely-generated approximate group which admits a geometric isometric action on a proper geodesic space is undistorted.  For undistorted approximate groups one has the usual Milnor-Schwarz lemma (see Theorem 4.37):

PROPOSITION 1.17. If an undistorted approximate group (Λ, Λ∞) acts geometrically and isometrically on a proper geodesic space X, then X is an apogee for (Λ, Λ∞). However, in order to include distorted approximate groups into the theory, one has to leave the realm of isometric actions and consider the larger class of “quasi-isometric quasi-actions” (qiqacs).

REMARK 1.18 (Quasi-isometric quasi-actions). If G is a group and X is a metric space, then a qiqac of G on X is a map G X X, (g, x) g.x such that the maps (x g.x) are quasi- × → 7→ 7→ g∈G isometries with uniform quasi-isometry constants and such that there exists a constant C satisfying d(g.(h.x), gh.x) C for all g,h G and x X. ≤ ∈ ∈ Such qiqacs arise naturally in geometric group theory (see e.g. [MSW00]): If Γ is a finitely-genera- ted group and X [Γ], then there may or may not exist a geometric action of Γ on X (this is the ∈ famous QI rigidity problem), but there always exists a geometric (i.e. proper and cobounded) qiqac of Γ on X. The reason that qiqacs do not feature more prominently in geometric group theory (outside of QI rigidity) is that every geometric qiqac is quasi-conjugate to an actual geometric isometric action (albeit on a different space). As Proposition 1.16 shows, this is no longer the case for approximate groups, and as a consequence, qiqacs play a central role in geometric approximate group theory. Because of the various QI constants involved, the precise definition of a qiqac of an approximate group is somewhat involved - see Definition 4.12 below. Our definition is certainly not the only pos- sible definition of a qiqac of an approximate group – certain uniformity conditions could be strength- ened or weakened in the definition. However, as the following result shows, our definition leads to a reasonable theory (cf. Proposition 4.34 and Theorem 4.37):

THEOREM 1.19 (Generalized Milnor-Schwarz lemma). Let (Λ, Λ∞) be a geometrically finitely-genera- ted approximate group. (i) There exists a geometric qiqac of Λ on a proper geodesic metric space X. In fact, X can be chosen to be any apogee of (Λ, Λ∞). (ii) Conversely, if Λ admits a geometric qiqac on a proper geodesic metric space X, then X is an apogee of (Λ, Λ∞). 

5. Quasi-isometric quasi-actions on hyperbolic spaces The results of the previous section indicate that geometric approximate group theory should be concerned with the study of quasi-isometric quasi-actions (qiqacs) of approximate groups on “nice” metric spaces. The precise techniques will necessarily depend on the class of metric spaces under consideration.

REMARK 1.20 (Hyperbolic qiqacs). Historically, one of the early successes of geometric group theory was the theory of hyperbolic groups, i.e. groups acting geometrically on proper geodesic hy- perbolic spaces [Gro93, GdlH90]. This theory has been extended in many different directions, notably to include finitely-generated groups with only very weak hyperbolicity properties (see [Osi16] for an example). For the remainder of this introduction we will thus focus on qiqacs of approximate groups on hyperbolic spaces and compare our results to similar results in the group case. We hasten to point out that a substantial part of this book is concerned with qiqacs on non-hyperbolic proper geodesic metric spaces. This, however, requires additional technical machinery (such as the use of Morse boundaries instead of Gromov boundaries), hence we will not mention the more general results in the introduc- tion. 12 1. INTRODUCTION

We will be interested in approximate groups which admit proper qiqacs on proper geodesic hy- perbolic spaces. We first consider the case of geometric actions (cf. Definition 5.3):

DEFINITION 1.21. A geometrically finitely-generated approximate group is called hyperbolic if it admits a geometric qiqac on a proper geodesic hyperbolic space.

The following theorem summarizes some of our main results concerning hyperbolic approximate groups (see Proposition 5.7, Theorem 5.21, Theorem 5.19, Lemma 5.16, Theorem 7.1, Corollary 5.42, Proposition 5.40, Theorem 8.9 and Theorem 8.1). Concerning the statement we observe that If X is an apogee of a hyperbolic approximate group, then there is a canonical qiqac of (Λ, Λ∞) on X, and this extends to an action of (Λ, Λ∞) on the Gromov boundary ∂X of X by homeomorphism called the boundary action. We then say that (Λ, Λ∞) is non-elementary if this action is fixpoint free.

THEOREM 1.22 (Hyperbolic approximate groups). Let (Λ, Λ∞) be a hyperbolic approximate group and let X be an apogee for (Λ, Λ∞). (i) X is a visual proper geodesic hyperbolic spaces which is quasi-isometric to a closed convex subset of some real hyperbolic space Hn. (ii) ∂X is doubling and locally quasi-self similar. (iii) asdim X = dim ∂X +1. If (Λ, Λ∞) is non-elementary, then moreover the following results hold: (iv) ∂X is infinite and homeomorphic to a perfect compact subset of a sphere. (v) The boundary action of (Λ, Λ∞) on ∂X is minimal. (vi) X has exponential growth. (vii) If asdim X =1, then X is quasi-isometric to a finitely-generated non-abelian free group.

All of these results are classical for hyperbolic groups: (i) is the celebrated Bonk–Schramm em- bedding theorem [BS00], (iii) was conjectured by Gromov [Gro93] and established (after a lot of work by many people) by Buyalo and Lebedeva [BL07]. The proof of (vi) in the group case is based on the ping-pong lemma, which ensures that a non-elementary hyperbolic group contains a free semigroup; no such result is true in our general setting. Instead, if asdim Λ > 1 then (vi) can be proved by bound- ing the exponential growth rate of Λ from below by the dimension of ∂X, which is stricitly positive by (iii). The case asdim Λ = 1 then has to be dealt with separately, and this is where (vii) is needed. The latter is proved using an embedding theorem of Buyalo and Schroeder [BS07a].

REMARK 1.23 (Limit sets of proper qiqacs). If one considers a proper qiqac of an approximate groups (Λ, Λ∞) on a proper geodesic hyperbolic space X which is not cobounded, then there is still a boundary action of (Λ, Λ∞) by homeomorphisms on the Gromov boundary ∂X of X. In this situation, an invariant subset of the boundary is given by the limit set

(Λ) := λ.o λ Λ ∂X, L { | ∈ } ∩ where o X is an arbitrary basepoint (see Definition 6.2). Elements of (Λ) are called limit points of the ∈ L qiqac. Note that, by definition, every limit point ξ is the limit of a sequence of the form λ .o with λ n n ∈ Λ. However, it is not clear whether this sequence can be chosen to be uniformly close to a geodesic representing ξ. If this is the case, then we call ξ a conical limit point of the qiqac (see Definition 6.18). The subset con(Λ) (Λ) of conical limit points often contains additional information compared to L ⊂ L (Λ). L A striking illustration of the difference between limit points and conical limit points can be given as follows: Consider a free group F of rank r 3. With every cyclically reduced non-self-overlapping r ≥ word u of length 2 in F one can associated an approximate subgroup Λ of F , the quasi-kernel of ≥ r u r the associated cyclic counting quasimorphism. This approximate subgroup then admits an isometric action on the Cayley tree T2r of Fr, and we have (see Theorem 6.22): ACKNOWLEDGEMENTS 13

THEOREM 1.24 (Reconstruction from conical limit points). Let u, v F be two cyclically reduced ∈ r non-self-overlapping words of length 2. Then (Λ ) = (Λ ), whereas we have con(Λ ) = con(Λ ) if ≥ L u L v L u L v and only if u = v±1.  On the other hand, there are also interesting qiqacs for which every limit point is conical.

REMARK 1.25 (Weak hulls and convex cocompact qiqacs). Let X be a proper geodesic hyperbolic space. Given a closed subset Z ∂X of the Gromov boundary, we define its weak hull H(Z) as the set ⊂ of all geodesics which connect points in Z (see Definition D.27). In particular, given a proper qiqac of an an approximate group (Λ, Λ∞) with unbounded orbits, we can consider the weak hull H( (Λ)) of the limit set. This weak hull is not quite invariant under L Λ∞, but it is “quasi-invariant” under Λ. This is enough to define a “restricted qiqac” of Λ on H( (Λ)), L and we say that the qiqac is convex cocompact if this restricted qiqac is cobounded. We then have the following characterization (see Corollary 6.33):

THEOREM 1.26 (Convex cocompact qiqacs). For a qiqac of an infinite approximate group (Λ, Λ∞) on a proper geodesic hyperbolic space X the following are equivalent: (i) The qiqac is convex cocompact. (ii) The qiqac is proper has quasi-convex quasi-orbits In this case, all limit points of the qiqac are conical. In the group case, this reduces to a classical result of Swenson [Swe01].

Acknowledgements Just like its authors, this book has been growing slowly but steadily over the last five years. The authors would thus like to express their gratitude to all the people and institutions which have sup- ported the project over this long period of time. In particular, the authors are indebted to the following people: To Michael Björklund, without whom this whole theory would not exist. • To Nina Lebedeva for providing a list of corrections for the proof of [BL07, Thm. 1.1], on • which the proof of Theorem 7.21 is based. To Alexey Talambutsa for help with the proof of Theorem B.29. • To Stefan Witzel, who suggested the definition of finiteness properties for countable approx- • imate groups used in this book; the second half of Section 3.2 paraphrases joint work with him [HW21]. To Gabi Ben Simon, Laura Bonn, Sergei Buyalo, James Farre, Slava Grigorchuk, Anton Hase, • Nir Lazarovich, Leonard Rubin, Michah Sageev, Viktor Schroeder, Alessandro Sisto and Emily Stark for various comments, corrections and suggestions. To the members of the Geometry and Seminar at the Technion for providing a stim- • ulating and collaborative atmosphere over the years; this book can be seen as a continuation of the discussions which arose from various talks in this seminar. Moreover, the authors would like to thank the following institutions for providing them with excellent working conditions during various stages of their collaboration: the Technion Faculty of Mathematics, Haifa, the Department of Mathematics at the University of , the Institut für Algebra und Geome- trie at the Karlsruher Institut für Technologie and the Mathematische Forschungsinsitut Oberwolfach. M.C. was supported in part by a Zuckerman STEM Leadership Fellowship, an ETH Postdoctoral Fellowship cofunded by a Marie Curie Actions for People COFUND Program, and an SNSF Am- bizione Fellowship. T.H. and V.T. were supported by the Erasmus+ KA107 program, through the 2016-2018 inter-institutional agreement between the Technion and the . T.H. was moreover supported through the program“Research in Pairs” by the Mathematische Forschungsinsti- tut Oberwolfach in 2020.

CHAPTER 2

Basic constructions of approximate groups

In this chapter we introduce approximate groups and various variants thereof. We then discuss in some detail various possible notions of morphisms of approximate groups. Finally, we survey various known constructions of approximate groups. These constructions will be used throughout this book to provide examples and illustrate the general theory.

1. Categories of filtered groups Before we begin our discussion of approximate group, we describe a general framework of filtered groups. We will see later that approximate groups fall into this framework. The main advantage of this abstract framework is that it allows for certain categorical constructions which cannot be defined directly for approximate groups, notably kernels. We will need the following notations.

NOTATION 2.1 (Notations concerning subsets of groups). Given a group Γ and subsets A, B Γ ⊂ we denote by AB the product set AB := ab a A, b B { | ∈ ∈ } and by A−1 the inverse set A−1 := a−1 a A . { | ∈ } For k 2 we will denote by ≥ Ak := AA A ··· k times the kth iterated product set. For distinction, we will denote its k-fold Cartesian product by | {z } A×k := A A A. × ×···× k times We observe that for all k N we have (A−1)k = (Ak)−1, and we occasionally denote this set by A−k. ∈ | {z } We call a subset A Γ symmetric provided A = A−1 and unital provided e A. Amap f : A B ⊂ ∈ → between symmetric subsets of groups is called symmetric if f(a−1)= f(a)−1 for all a A. If A is unital ∈ this implies f(e)= e.

REMARK 2.2 (Some properties of products of subsets of a group). We may occasionally need the following basic facts concerning subsets of groups. (1) If A, B are subsets of a group Γ, we have (A B)k Ak Bk for all k N. ∩ ⊂ ∩ ∈ (2) If C is a symmetric subset and A, B are subsets of a group Γ, and l, k N, we have ∈ (CkA) (ClB) Cl((Cl+kA) B). ∩ ⊂ ∩ DEFINITION 2.3. An N-filtered group is a pair (Γ, (Γk)k∈N) where Γ is a group and Γ Γ Γ ... 1 ⊂ 2 ⊂ 3 ⊂ is an ascending sequence of subsets of Γ such that Γ= Γ , e Γ and for all k,l N, k≥1 k ∈ 1 ∈ Γ Γ Γ . k l ⊂ k+Sl The group Γ is called the ambient group and Γk is called the kth filtration step. Since we will only consider N-filtered groups below, we will use the term filtered group as a syn- onym for N-filtered group.

15 16 2. BASIC CONSTRUCTIONS OF APPROXIMATE GROUPS

EXAMPLE 2.4. We have the following basic examples: (1) Every group Γ can be considered as a filtered group by setting Γ := Γ for all k N. k ∈ (2) If Γ is a group and Λ Γ is a unital subset, then we obtain a filtered group (Γ, (Γk)k∈N) by k ⊂ setting Γk := Λ . This filtered group is called the filtered group associated with the pair (Λ, Γ). (3) If (Γ, (Γk)k∈N) is a filtered group, then its l-shift Sl(Γ, (Γk)k∈N) is defined as the filtered group (Γ, Γkl)k∈N.

DEFINITION 2.5. Let (Γ, (Γ ) N) and (Θ, (Θ ) N) be filtered groups and let k N. n n∈ n n∈ ∈ (i) A filtered morphism ρ : (Γ, (Γ ) N) (Θ, (Θ ) N) of filtered groups is a group homomorphism n n∈ → n n∈ ρ :Γ Θ such that ρ (Γ ) Θ for all n N. ∞ → ∞ n ⊂ n ∈ (ii) A k-local filtered morphism ρ : (Γ, (Γ ) N) (Θ, (Θ ) N) is a map ρ : Γ Θ such that n n∈ → n n∈ k k → k ρ (e)= e and ρ (gh)= ρ (g)ρ (h) for all g,h Γ which satisfy gh Γ . k k k k ∈ k ∈ k REMARK 2.6. (i) If ρ : (Γ, (Γ ) N) (Θ, (Θ ) N) is a filtered morphism of filtered groups, n n∈ → n n∈ then ρ restricts to a k-local filtered morphism ρ : Γ Θ for all k N. We refer to ρ as the ∞ k k → k ∈ k k-component of the filtered morphism ρ. (ii) Similarly, if ρ : (Γ, (Γ ) N) (Θ, (Θ ) N) is a k-local filtered morphism and l k, then we n n∈ → n n∈ ≤ define its l-component as the restriction ρ : Γ Θ of ρ to Γ , which is again an l-local filtered l l → l k l morphism. (iii) The composition of filtered morphisms is again a filtered morphism, and similarly the compo- sition of k-local filtered morphisms is a k-local filtered morphism. Filtered groups and filtered morphisms form a category, as do filtered groups and k-local filtered morphism for any fixed k N. ∈ In the category of filtered groups we can now define kernels and images in the obvious way.

DEFINITION 2.7. Let ρ : (Γ, (Γ ) N) (Θ, (Θ ) N) be a filtered morphism of filtered group. k k∈ → k k∈ (i) The filtered group (ρ∞(Γ), (ρk(Γk))k∈N) is called the image of ρ. (ii) The pair ker (ρ) := (ker(ρ ), (ker (ρ) := ker(ρ ) Γ ) N) is a filtered group, called the kernel ∞ ∞ k ∞ ∩ k k∈ of ρ. (iii) The sets ρk(Γk) and kerk(ρ) are called the partial images, respectively partial kernels of ρ. Partial images and partial kernels of local filtered morphisms can be defined similarly.

2. Approximate groups and related notions In this section we introduce approximate subgroups and various related notions. In particular, with every approximate subgroup we associate a filtered group and study the corresponding filtered morphisms.

DEFINITION 2.8 (T. Tao, [Tao08]). Let Γ be a group and K N. A subset Λ Γ is called a ∈ ⊂ K-approximate subgroup, if (AG1) Λ=Λ−1 := x−1 x Λ and e Λ. { | ∈ } ∈ (AG2) There exists a finite subset F Γ such that Λ2 ΛF and F = K. Λ ⊂ ⊂ Λ | Λ| We say that Λ is an approximate subgroup if it is a K-approximate subgroup for some K N. ∈ If Λ Γ is an approximate subgroup, then the group Λ∞ := Λk is called the enveloping group ⊂ k∈N of Λ, the pair (Λ, Λ∞) is called an approximate group, and the associated filtered group (Λ∞, (Λk) N) is S k∈ called a filtered approximate group. We say that an approximate group (Λ, Λ∞) is finite (countable) if Λ is finite (countable). To discuss some first examples we need the following notions.

DEFINITION 2.9. Let Γ be a group and let A, B Γ be subsets. ⊂ (i) If A B, then A is called left-syndetic in B if there exists a finite subset F Γ such that B AF , ⊂ ⊂ ⊂ and right-syndetic in B if there there exists a finite subset F Γ such that B F A. ⊂ ⊂ 2. APPROXIMATE GROUPS AND RELATED NOTIONS 17

(ii) A and B are called left-commensurable if there exist finite subsets F , F Γ such that A BF 1 2 ⊂ ⊂ 1 and B AF and right-commensurable if there exist finite subsets F , F Γ such that A F B ⊂ 2 1 2 ⊂ ⊂ 1 and B F A. ⊂ 2 If G is a topological group, then we define similar notions of left-/right-syndeticity and left-/right- commensurability for subsets of G by demanding that the sets F , F1, F2 in question be compact (rather than finite). By definition, if A is left-syndetic in B, then A and B are left-commensurable. Every non-amenable group contains a left-syndetic set which is not right-syndetic ([Pau10]), but for symmetric sets A and B the notions coincide.

EXAMPLE 2.10 (Trivial examples of approximate groups). We now give some trivial examples of approximate groups. All currently known examples arise from these trivial examples by the construc- tions outlined in the remainder of this section. (1) If Γ is a group, then Γ Γ is an approximate subgroup and hence the pair (Γ, Γ) is an approximate ⊂ group. We will occasionally identify a group Γ with the approximate group (Γ, Γ). (2) If Γ is a (discrete) group and Λ Γ is a syndetic symmetric unital subset, then (Λ, Λ∞) is an ⊂ approximate group. This gives plenty of examples of approximate groups, which are not groups. However, the corresponding approximate subgroups are still commensurable to the ambient gro- up. We refer to them as almost groups. (3) If Γ is a group and F is a finite symmetric subset of Γ containing the identity, then (F, F ∞) is an approximate group. More generally, if G is a locally compact group and W is a relatively compact (i.e. having compact closure) symmetric identity neighborhood in G, then (W, W ∞) is an approximate group. A slight variation of the second example is given as follows:

PROPOSITION 2.11. Let (Ξ, Ξ∞) be an approximate group. Assume that Λ Ξ is left-syndetic in Ξ, ⊂ symmetric (i.e. Λ=Λ−1) and contains the identity. Then Λ is an approximate subgroup of Ξ∞, and hence (Λ, Λ∞) is an approximate group.

PROOF. Let F , F be finite sets such that Ξ ΛF and Ξ2 ΞF , and let F := F F . Then F is 1 2 ⊂ 1 ⊂ 2 1 2 finite and since Λ Ξ we have Λ2 Ξ2 ΞF ΛF F =ΛF .  ⊂ ⊂ ⊂ 2 ⊂ 1 2 REMARK 2.12 (Approximate subgroups of approximate groups). If (Ξ, Ξ∞) and (Λ, Λ∞) are ap- proximate groups, then (Ξ, Ξ∞) is called an approximate subgroup of (Λ, Λ∞) provided Ξ Λ and ⊂ Ξ∞ Λ∞. Note that this is compatible with our previous terminology: If Γ is a group, then a sub- ⊂ set Λ Γ is an approximate subgroup of Γ in the sense of Definition 2.8 if and only if (Λ, Λ∞) is an ⊂ approximate subgroup of (Γ, Γ) in this sense. Our next goal is to establish the following basic facts concerning approximate subgroups:

PROPOSITION 2.13. Let Γ be a group and let Λ Γ be an approximate subgroup. ⊂ (i) There is a finite subset F Λ∞ such that for every l, k N with l < k, we have Λk Λk−lF l F l Λk−l. Λ ⊂ ∈ ⊂ Λ ∩ Λ (ii) For every finite subset F Λ∞ there exists a finite subset F ∗ Λ∞ such that ΛF F ∗Λ and F Λ ⊂ ⊂ ⊂ ⊂ ΛF ∗. (iii) Λk is an approximate subgroup of Γ for every k N. ∈ In particular, the filtration Λ Λ2 Λ3 ... ⊂ ⊂ ⊂ of Λ∞ consists of pairwise commensurable approximate subgroups. In some of our constructions below we will encounter certain asymmetric sets which otherwise behave like approximate subgroups; we will thus establish Proposition 2.13 in this wider context.

DEFINITION 2.14. Let Γ be a group. A subset Λ Γ is called a left-quasi-subgroup of Γ if it satisfies ⊂ the following conditions: 18 2. BASIC CONSTRUCTIONS OF APPROXIMATE GROUPS

(QG1) There exists a finite subset F Γ such that Λ−1 ΛF . 1 ⊂ ⊂ 1 (QG2) There exists a finite subset F Γ such that Λ2 ΛF . 2 ⊂ ⊂ 2 It is called a right-quasi-subgroup of Γ if it satisfies the dual conditions: (QG1op) There exists a finite subset F Γ such that Λ−1 F Λ. 1 ⊂ ⊂ 1 (QG2op) There exists a finite subset F Γ such that Λ2 F Λ. 2 ⊂ ⊂ 2 REMARK 2.15. (i) By definition, an approximate group is a unital symmetric left-quasi-subgroup. (ii) If Λ Γ is a left-quasi-subgroup, then Λ−1 is a right-quasi-subgroup and vice versa. In partic- ⊂ ular, an approximate subgroup could equivalently be defined as a unital symmetric right-quasi- subgroup. −1 2 (iii) In (QG1) we can choose F1 to be contained in (Λ ) Λ, and in (QG2) we can choose F2 to be contained in Λ−1Λ2. If Λ is an approximate subgroup (and hence both a left- and a right-quasi- subgroup), then we can find a finite set FΛ such that Λ2 F Λ ΛF and F Λ3. (2.1) ⊂ Λ ∩ Λ Λ ⊂ We will usually reserve the letter FΛ to denote such a set. We will see that the same set can be used in Proposition 2.13 (i). (iv) If d is any left-invariant metric on Γ and Λ Γ is a left-quasi-subgroup, then there exists a ⊂ constant C 0 such that for all x, y Λ we have ≥ ∈ d(x−1, Λ) C and d(xy, Λ) C. (2.2) ≤ ≤ If Λ is instead a right-quasi-subgroup, then (2.2) holds for any right-invariant metric d on Γ. From a metric point of view, this captures the notion of a subset of a group, in which multiplication and inversion are defined up to a bounded error. Asking for exact symmetry as in (AG1) seems less natural from the metric point of view, but it turns out to be convenient for many purposes. If Λ is a left- or right-quasi-subgroup of Γ, then Λ+ := Λ e is a unital left- or right-quasi- ∪{ } subgroup respectively. We will thus restrict attention to unital quasi-subgroups. If Λ is a unital left- or right-quasi-subgroup of Γ, then in analogy with the symmetric case we denote by ∞ Λ∞ := Λk k[=1 the enveloping semigroup of Λ.

LEMMA 2.16. Let l, k N with l < k, let Γ be a group, let Λ Γ be unital and let F Λ∞ be finite. ∈ 0 ⊂ ⊂ (i) If Λ Γ is a left-quasi-subgroup of Γ, then Λk Λk−lF l. ⊂ ⊂ 2 (ii) If Λ is a left-quasi-subgroup of Γ, then there exists a finite subset F ∗ Λ∞ such that F Λ ΛF ∗. ⊂ ⊂ (iii) If Λ Γ is a right-quasi-subgroup of Γ, then Λk F lΛk−l. ⊂ ⊂ 2 (iv) If Λ is a right-quasi-subgroup of Γ, then there exists a finite subset F ∗ Λ∞ such that ΛF F ∗Λ. ⊂ ⊂ PROOF. (i) For every k 2, (QG2) implies ≥ Λk = (Λk−2)Λ2 Λk−2ΛF =Λk−1F ⊂ 2 2 Thus the statement follows by induction. (ii) Since F is finite and each element of F is contained in some power of Λ, there exists k N such ∈ that F Λk. We thus deduce from (i) that ⊂ F Λ ΛkΛ=Λk+1 ΛF k. ⊂ ⊂ 2 ∗ k We may thus choose F := F2 . (iii) and (iv) follow by applying (i) and (ii) respectively to Λ−1. 

COROLLARY 2.17. If Γ is a group and Λ Γ is a unital left-quasi-subgroup of Γ such that Λ−1 Λ∞, ⊂ ⊂ then so is Λk for all k N. ∈ 3. MORPHISMS BETWEEN APPROXIMATE GROUPS 19

PROOF. Let Ξ := Λk. By (QG1) we have Ξ−1 = (Λ−1)k (ΛF )k =ΛF ΛF , ⊂ 1 1 ··· 1 and we may assume that F (Λ−1)2 Λ∞, hence we can apply Lemma 2.16.(ii) to move the finite 1 ⊂ ⊂ sets to the right and thereby to establish (QG1) for Ξ. As for (QG2), by Lemma 2.16.(i) we have Ξ2 =Λ2k ΛkF k = ΞF k. ⊂ 2 2 This finishes the proof. 

Proposition 2.13 now follows by specializing Lemma 2.16 and Corollary 2.17 to the symmetric case.

3. Morphisms between approximate groups We now turn to the slightly subtle problem of defining morphisms between approximate groups.

DEFINITION 2.18. A (global) morphism between approximate groups is a filtered morphism be- tween the associated filtered groups. Given k N,a k-local morphism between approximate groups is ∈ a k-local filtered morphism between the associated filtered groups More explicitly, if (Λ, Λ∞) and (Ξ, Ξ∞) are approximate groups, then a morphism ρ : (Λ, Λ∞) → (Ξ, Ξ∞) is given by a group homomorphism ρ :Λ∞ Ξ∞ which maps Λ to Ξ and thus induces ∞ → maps ρ :Λk Ξk, which are the k-components of ρ. Note that any k-local morphism ρ :Λk Ξk is k → k → symmetric, since for x Λ we have x−1,e Λ, and hence xx−1 = e implies ρ (x)ρ (x−1)= ρ (e)= ∈ { }⊂ k k k e. Since we identify each group Γ with the corresponding approximate group (Γ, Γ), we will simply write ρ : (Ξ, Ξ∞) Γ for a (global) morphism ρ : (Ξ, Ξ∞) (Γ, Γ). → → REMARK 2.19 (Local morphisms vs. Freiman homomorphisms). There are various (subtly dif- fering) notions of “Freiman homomorphism” in the additive combinatorics literature. According to [Bre14, Def. 1.4],a Freiman k-homomorphism between a subset A of a group G and a subset B of a group H is a map ρ : A B such that for all m 2k, x ,...,x A and ǫ ,...,ǫ 1 we have → ≤ 1 m ∈ 1 m ∈ {± } xǫ1 xǫm = e = ρ(x )ǫ1 ρ(x )ǫm = e. 1 ··· m ⇒ 1 ··· m Assume now that (Λ, Λ∞) and (Ξ, Ξ∞) are approximate groups and let ρ :Λ Ξ be a map. → If ρ is the restriction of a k-local morphism ρ :Λk Ξk, then for m 2k, x ,...,x A and k → ≤ 1 m ∈ ǫ ,...,ǫ 1 we have 1 m ∈ {± } xǫ1 xǫm = e xǫ1 xǫk = x−ǫk+1 x−ǫm ρ (xǫ1 xǫk )= ρ (x−ǫk+1 x−ǫm ) 1 ··· m ⇒ 1 ··· k k+1 ··· m ⇒ k 1 ··· k k k+1 ··· m ρ(xǫ1 ) ρ(xǫk )= ρ(x )−ǫk+1 ρ(x )−ǫm ρ(x )ǫ1 ρ(x )ǫm = e, ⇒ 1 ··· k k+1 ··· m ⇒ 1 ··· m hence ρ is a Freiman k-homomorphism. Conversely, let ρ be a Freiman k-homomorphism. Let x ,...,x ,y ,...,y Λ and set x := 1 k 1 k ∈ x x Λk and y := y y . If x = y, then 1 ··· k ∈ 1 ··· k x x y−1 y−1 = e ρ(x ) ρ(x )ρ(y )−1 ρ(y )−1 = e ρ(x ) ρ(x )= ρ(y ) ρ(y ). 1 ··· k k ··· 1 ⇒ 1 ··· k k ··· 1 ⇒ 1 ··· k 1 ··· k We can thus extend ρ to a map ρ :Λk Ξk, ρ (x x ) := ρ(x ) ρ(x ). k → k 1 ··· k 1 ··· k Now it is enough to assume that ρ is a Freiman m-homomorphism for m 3 k. Let x, y be as above ≥ 2 and assume that z := xy Λk. If we write z = z z with z ,...,z Λ, then, since 3k 2m, we ∈ 1 ··· k 1 k ∈ ≤ have x x y y z−1 z−1 = e ρ(x ) ρ(x )ρ(y ) ρ(y )ρ(z )−1 ρ(z )−1 = e, 1 ··· k 1 ··· k k ··· 1 ⇒ 1 ··· k 1 ··· k k ··· 1 20 2. BASIC CONSTRUCTIONS OF APPROXIMATE GROUPS and hence ρk(x)ρk(y)= ρk(z), which shows that ρ is a k-local morphism. To summarize, ρ = ρ , where ρ is a k-local morphism = ρ is a Freiman k-homomorphism k|Λ k ⇒ = ρ = ρ , ⇒ ⌊2k/3⌋|Λ where ρ is a 2k/3 -local morphism. ⌊2k/3⌋ ⌊ ⌋ In particular, all results about local morphisms are equivalent to results about Freiman homomor- phisms if one is willing to take a loss in the implied constants. In fact, most of the results concerning local morphisms established below actually hold for (extensions of) Freiman homomorphisms with the same constant, as is evident from their proofs. We will usually only establish the local versions in our proofs and not bother with optimizing constants.

REMARK 2.20 (Kernels and Images). Let ρ : (Ξ, Ξ∞) (Λ, Λ∞) be a global morphism of approx- → k imate groups. We define the kth partial kernel and the kth partial image as kerk(ρ) := ker(ρ∞) Ξ , k k ∩ respectively Imk(ρ) := ρ(Λ )= ρ(Λ) . By definition these are the partial kernels and images of the in- duced map between the associated filtered groups. Partial kernels and images can also be defined for local morphisms in the obvious way. We will see later that the partial images are pairwise commen- surable approximate subgroups for k 1, and that the partial kernels are pairwise commensurable ≥ approximate subgroups for k 2. ≥ The following notion will become important in our study of quasi-isometry types of finitely- generated approximate groups.

DEFINITION 2.21. Let ρ : (Ξ, Ξ∞) (Λ, Λ∞) be a global morphism of approximate groups. We → say that a ρ has finite kernel if for every k N the partial kernel ker (ρ) is finite. ∈ k We warn the reader that a global morphism ρ : (Ξ, Ξ∞) (Λ, Λ∞) may have finite kernel in the → sense of Definition 2.21 despite the fact that the map ρ : Ξ∞ Λ∞ has infinite kernel: ∞ → EXAMPLE 2.22. Let Γ1 be a group and let Γ2 be a finitely-generated infinite group with finite generating set S. Then the projection onto the first coordinate p : (Γ S, Γ Γ ) Γ 1 × 1 × 2 → 1 has finite kernel, since ker (p)= e Sk is finite, but the underlying group homomorphism Γ Γ k { }× 1 × 2 → Γ1 has infinite kernel. Since approximate groups are quasifications of groups, it is natural to also quasify the relevant morphisms. The definition of a “quasified group homomorphism” we will work with is that of a quasimorphism (with discrete target) in the sense of Definition B.42, which we recall here for the convenience of the reader:

DEFINITION 2.23. Let G and H be groups and let f : G H be a map. Then f is called a → quasimorphism if its left-defect set D(f) := f(y)−1f(x)−1f(xy) x, y G (3.1) { | ∈ } is finite. We refer the reader to Section 5 for a discussion of this definition and various related notions.

DEFINITION 2.24. Let (Ξ, Ξ∞) and (Λ, Λ∞) be approximate groups. (i) A map of pairs ρ : (Ξ, Ξ∞) (Λ, Λ∞) is called a global quasimorphism (or simply a quasimorphism) → if ρ : Ξ∞ Λ∞ is a quasimorphism in the sense of Definition B.42. → (ii) A map of pairs ρ : (Ξ, Ξk) (Λ, Λ∞) is called a k-local quasimorphism (or simply a k-quasimorph- k → ism) if the (left-)k-defect set D (ρ ) := ρ (y)−1ρ (x)−1ρ (xy) x, y such that xy Ξk (3.2) k k { k k k | ∈ } is finite. 4. INTERSECTIONS AND IMAGES UNDER LOCAL (QUASI-)MORPHISMS 21

REMARK 2.25 (Quasimorphisms and filtrations). Unlike morphisms, quasimorphisms between approximate groups do not preserve the respective filtrations. However, if ρ : (Ξ, Ξ∞) (Λ, Λ∞) is → a quasimorphism between approximate groups, then D(ρ) Λ∞ is finite, hence there exists N N ⊂ ∈ such that D(ρ) ΛN . We then have ρ(Ξ2) ρ(Ξ)2D(ρ) Λ2+N and thus ⊂ ⊂ ⊂ ρ(Ξk) Λk+(k−1)N (Λ2N )k, (3.3) ⊂ ⊂ which shows that ρ induces a morphism between the approximate groups (Ξ, Ξ∞) and (Λ2N , Λ∞) (rather than between (Ξ, Ξ∞) and (Λ, Λ∞)). This shift is the reason why we did not require in the definition of a k-local quasimorphism that ρ(Ξk) Λk, but only that ρ (Ξk) Λ∞. ⊂ k ⊂ 4. Intersections and images under local (quasi-)morphisms We now explain how to construct non-trivial examples of approximate groups starting from the trivial examples discussed above. We will see that many of the standard constructions for groups work with slight modifications. In the present section we consider intersections and direct images. The fact that the intersection of two subgroups is again a subgroup generalizes to the approximate setting as follows.

LEMMA 2.26 (Intersections). Let Γ be a group and let Λ, Ξ Γ be approximate subgroups and k,l 2. ⊂ ≥ Then Λk Ξl is an approximate subgroup of Γ. ∩ PROOF. By symmetry we may assume that l k. Let F Γ be a finite symmetric set containing ≤ ⊂ e such that Λ2 F Λ and Ξ2 F Ξ. Define a finite set T F 2k+2l−2 by ⊂ ⊂ ⊂ T := t F 2k+2l−2 tΛ Ξ = . { ∈ | ∩ 6 ∅} For every t T pick an element x tΛ Ξ. Given t T and x tΛ Ξ we have x−1x Λ2 Ξ2, ∈ t ∈ ∩ ∈ ∈ ∩ t ∈ ∩ hence x x (Λ2 Ξ2). This shows that ∈ t ∩ tΛ Ξ x (Λ2 Ξ2), ∩ ⊂ t ∩ and since Λk F k−1Λ and Ξl F l−1Ξ, with help of Remark 2.2 we deduce that ⊂ ⊂ (Λk Ξl)2 Λ2k Ξ2l F 2k−1Λ F 2l−1Ξ F 2l−1 (F 2k+2l−2Λ) Ξ ∩ ⊂ ∩ ⊂ ∩ ⊂ ∩ 2l−1  2l−1  2 2 2l−1  2 2 = F tΛ Ξ F xt(Λ Ξ ) = F xt (Λ Ξ ) ∩ ! ⊂ ∩ ! ! ∩ t[∈T t[∈T t[∈T 2l−1 k l F xt (Λ Ξ ). ⊂ ! ∩ t[∈T Since T and F are finite, this shows that Λk Ξl is an approximate subgroup of Γ.  ∩ Next we would like to establish the following two results: The image of an approximate subgroup under a group homomorphism is an approximate • subgroup. (Together with Proposition 2.13 this implies that the partial images of a global morphism of approximate groups are pairwise commensurable approximate subgroups.) The image of a group under a symmetric quasimorphism is an approximate subgroup. • In fact, we will establish both result simultaneously in the following general form:

PROPOSITION 2.27 (Direct images). Let G, H be groups and let f : G H be a (left-)quasimorphism. → If Ξ G is a left-quasi-subgroup, then Λ := f(Ξ) H is a left-quasi-subgroup. In particular, if Ξ is an ⊂ ⊂ approximate subgroup and f is symmetric, then Λ is an approximate subgroup.

PROOF. Let g Ξ. Using the inverse of (5.6), as well as (QG1) we have ∈ f(g)−1 f(g−1)D(f)f(e)−1 f(Ξ−1)D(f)f(e)−1 f(ΞF )D(f)f(e)−1. ∈ ⊂ ⊂ 1 By (5.4) we have f(ΞF ) f(Ξ)f(F )D(f)=Λf(F )D(f), 1 ⊂ 1 1 22 2. BASIC CONSTRUCTIONS OF APPROXIMATE GROUPS and thus Λ−1 Λf(F )D(f)2f(e)−1, which establishes (QG1) for Λ. As for (QG2) a similar computa- ⊂ 1 tion shows Λ2 = f(Ξ)2 f(Ξ2)(D(f))−1 f(ΞF )(D(f))−1 Λf(F )D(f)(D(f))−1. ⊂ ⊂ 2 ⊂ 2 This proves the first statement, and the second statement then follows from the fact that the image of a symmetric unital set under a symmetric map is symmetric and unital. 

REMARK 2.28. Proposition 2.27 implies in particular that if ρ : (Ξ, Ξ∞) (Λ, Λ∞) is a symmetric → global quasimorphism of approximate groups, then ρ(Ξ) is an approximate subgroup of Λ∞. In this case the condition that ρ be global can be weakened considerably. Indeed, since Ξ is symmetric, we have F = e , and F can be chosen to be contained in Ξ3. Then an inspection of the proof of 1 { } 2 Proposition 2.27 yields the following result.

COROLLARY 2.29 (Local direct images). Let ρ : (Ξ, Ξ∞) (Λ, Λ∞) be a symmetric 4-local quasimor- → phism of approximate groups. Then ρ(Ξ) is an approximate subgroup of Λ∞. 

5. Pre-images and (partial) kernels Not only intersections and direct images of subgroups, but also pre-images of subgroups under group homomorphisms are subgroups. We now investigate the situation for approximate subgroups and quasimorphisms. We will establish results like the following:

PROPOSITION 2.30 (Preimages). (i) If ρ : G H is a group homomorphism and Λ H is an → ⊂ approximate subgroup, then ρ−1(Λ2) is an approximate subgroup. (ii) If ρ : G A is a surjective quasimorphism with abelian target, Λ A is an approximate subgroup and → ⊂ S A is a symmetric set containing D(ρ), then ρ−1(Λ + S) is an approximate subgroup of G. ⊂ We will also establish a more technical version of (ii) for quasimorphisms with non-abelian target. All of these results follow from the following general lemma:

LEMMA 2.31. Let G and H be groups, f : G H be a surjective quasimorphism, let F H be a finite → ⊂ subset, and Λ H a subset. Assume the following: ⊂ (I) There exists a finite subset F H such that Λ2 ΛF . 2 ⊂ ⊂ 2 (II) Either F contains f(e)D(f)−1D(f) e or f is symmetric and F contains D(f) e . ∪{ } ∪{ } (III) There is a finite set F ∗ such that F Λ ΛF ∗ and F −1Λ ΛF ∗. ⊂ ⊂ −1 −1 −1 Then both Ξ1 := f (ΛF ) and Ξ2 := f (F ΛF ) satisfy (QG2).

PROOF. Let Ξ be either Ξ or Ξ . In both cases we have f(Ξ) F −1ΛF and thus by (5.4) 1 2 ⊂ f(Ξ2) f(Ξ)f(Ξ)D(f) F −1ΛF F −1ΛFD(f). ⊂ ⊂ ∗∗ ∗ ∗ 2 ∗∗ If we set F := F F2(F ) FD(f), then F is finite, and using (III), (I) and (III) again, we find f(Ξ2) F −1Λ2(F ∗)2FD(f) F −1ΛF (F ∗)2FD(f) ΛF ∗F (F ∗)2FD(f)=ΛF ∗∗. ⊂ ⊂ 2 ⊂ 2 Now we fix a section σ : H G of f, i.e., a map such that f σ = Id , and define a finite subset of G → ◦ H by F := σ(F ∗∗). As we just saw, if ξ , ξ Ξ, then there exist x Λ and y F ∗∗ such that f(ξ ξ )= xy. 1 2 ∈ ∈ ∈ 1 2 Then by (5.4), (5.6) and (II) we have b f(ξ ξ σ(y)−1) f(ξ ξ )f(σ(y)−1)D(f)= xyf(σ(y)−1)D(f)= xf(σ(y))f(σ(y)−1)D(f) ΛF. 1 2 ∈ 1 2 ⊂ We deduce that ξ ξ σ(y)−1 f −1(ΛF ) Ξ, and hence ξ ξ ΞF . This shows that Ξ2 ΞF and 1 2 ∈ ⊂ 1 2 ∈ ⊂ finishes the proof.  b b Now Part (i) of Proposition 2.30 follows directly: 5. PRE-IMAGES AND (PARTIAL) KERNELS 23

COROLLARY 2.32. Let G, H be groups, let Λ H be an approximate subgroup and let ρ : G H be a ⊂ → group homomorphism. (i) If ρ is surjective, then Ξ := ρ−1(Λ) is an approximate subgroup of G. If moreover Λ generates H, then Ξ generates G. (ii) If ρ is not necessarily surjective, then Ξ := ρ−1(Λ2) is an approximate subgroup of G.

PROOF. (i) Since Λ is symmetric and contains the identity and ρ is symmetric, Ξ satisfies (AG1). To show that Ξ satisfies (AG2), or equivalently (QG2), we apply Lemma 2.31 with F := e . Condition { } (I) holds, since Λ is an approximate subgroup, Condition (II) holds since D(ρ) = e , and Condition { } (III) holds since F = e . We deduce that ρ−1(ΛF )= ρ−1(Λ) = Ξ satisfies (QG2). This shows that Ξ is { } an approximate subgroup of G. Now assume that Λ generates H and let g G. Then ρ(g) Λk for some k N, say ρ(g)= λ λ ∈ ∈ ∈ 1 ··· k with λ Λ. If we choose pre-images ξ of λ , then ξ Ξ, and ξ := gξ−1 ξ−1 ker(ρ)= ρ−1 e i ∈ i i i ∈ 0 k ··· 1 ∈ { }⊂ Ξ. This shows that g = ξ ξ ξk Ξ∞ and finishes the proof. 0 2 ··· ∈ (ii) Let H′ := ρ(G). By Lemma 2.26 the subset Λ2 H′ is an approximate subgroup of H′, and thus ∩ we may apply (i) to this approximate subgroup.  In the case of quasi-morphisms, the situation is more complicated. To obtain an easy to state result, we will restrict to the case of symmetric quasimorphisms. Note that if ρ : G H is a symmetric → quasimorphism, then ρ(e) = e and hence e = ρ(e)−1ρ(e)−1ρ(ee) D(ρ). We thus have the following ∈ second corollary of Lemma 2.31.

COROLLARY 2.33. Let (Λ, Λ∞) be an approximate group, let G be a group and let ρ : G Λ∞ be a → surjective symmetric quasimorphism. Then Ξ := ρ−1(D(ρ)−1ΛD(ρ)) is an approximate subgroup of G.

PROOF. Firstly, since Λ is symmetric, we have Ξ−1 = ρ−1((D(ρ)−1ΛD(ρ))−1)= ρ−1(D(ρ)−1Λ−1D(ρ)) = Ξ, and hence Ξ is symmetric. It remains to check that the triple (ρ, F, Λ) with F := D(ρ) satisfies Condi- tions (I) – (III) of Lemma 2.31. Now (I) is satisfied by assumption and (II) is satisfied by definition of F . Finally, (III) follows from Proposition 2.13.(ii), applied to F and F −1. 

REMARK 2.34. There is a crucial difference between Corollary 2.32 and Corollary 2.33. In the latter we have to assume that Λ generates the range Λ∞, whereas in the former Λ was allowed to be an arbitrary approximate subgroup of H. The assumption that Λ generates the range in Corollary 2.33 is necessary since we appeal to Proposition 2.13.(ii) in the proof, in order to get Condition (III) of Lemma 2.31. However, if the target group happens to be abelian, then Condition (III) is automatically satisfied, and thus the appeal to Proposition 2.13 is not necessary. We thus obtain the following strengthening of Corollary 2.33 in the abelian case.

COROLLARY 2.35. Let G and A be groups with A abelian, let Λ A be an approximate subgroup and let ⊂ ρ : G A be a surjective symmetric quasimorphism. Then Ξ := ρ−1(Λ + D(ρ) D(ρ)) is an approximate → − subgroup of G. In fact, the same conclusion holds if D(ρ) D(ρ) is replaced by any symmetric set containing − D(ρ).  This concludes the proof of Proposition 2.30. We now specialize to the case of kernels, i.e. preim- ages of the trivial subgroup.

COROLLARY 2.36. Let ρ : (Ξ, Ξ∞) (Λ, Λ∞) be a global morphism of approximate groups. Then the → partial kernels ker (ρ) are approximate subgroups of Ξ∞, for all k 2. k ≥ PROOF. Since ker (ρ) = ker(ρ ) Ξk, this is a special case of Lemma 2.26.  k ∞ ∩ Note that the corollary does not apply to the case k = 1; we do not know under which assump- tions ker (ρ) = ker(ρ ) Ξ is an approximate subgroup. The corollary generalizes to the case of 1 ∞ ∩ quasimorphisms in the following way: given a quasimorphism G A of groups with abelian target, → 24 2. BASIC CONSTRUCTIONS OF APPROXIMATE GROUPS we denote by D (ρ) the smallest symmetric subset of A containing D(ρ). Since D(ρ) D(ρ) is finite, sym − symmetric and contains D(ρ), this set is always finite.

DEFINITION 2.37. Let G be a group, A an abelian group and let ρ : G A be a surjective symmet- → −1 ric quasimorphism. Then the quasi-kernel of ρ is defined as the preimage qker(ρ) := ρ (Dsym(ρ)). Applying Corollary 2.35 with Λ= e we may record: { } COROLLARY 2.38. The quasi-kernel of every surjective symmetric quasimorphism with abelian target is an approximate subgroup.  A different generalization of Corollary 2.36 concerns partial kernels of local morphisms:

LEMMA 2.39. Let (Ξ, Ξ∞), (Λ, Λ∞) be approximate groups. (i) If ρ : (Ξ, Ξ10) (Λ, Λ10) is a 10-local morphism, then the partial kernel ker (ρ ) is an approximate 10 → 2 10 subgroup of Ξ∞. 6k−2 6k−2 (ii) If k 2 and if ρ(6k−2) : (Ξ, Ξ ) (Λ, Λ ) is a (6k 2)-local morphism, then the partial kernels ≥ → − ∞ ker2(ρ(6k−2)),..., kerk(ρ(6k−2)) are approximate subgroups of Ξ .

PROOF. We first prove (i), in which we use the notation ρ := ρ10. Since ρ is symmetric and ρ(e)= e, we have that ker (ρ) = Ξ2 ρ−1(e) is also symmetric and contains e. Therefore it remains to check 2 ∩ property (AG2), or equivalently, to find a finite subset F ∗ of Ξ∞ such that (ker (ρ))2 F ∗ ker (ρ). We 2 ⊂ 2 will construct this set as in the proof of Lemma 2.26. By (2.1), there is a finite set F := F Ξ∞ such that Ξ2 F Ξ and F Ξ3. Define a finite set Ξ ⊂ ⊂ ⊂ T F 3 by ⊂ T := t F 3 tΞ ρ−1(e) = , { ∈ | ∩ 6 ∅} and note that tΞ F 3Ξ Ξ10 for every t T . For every t T , pick an element z tΞ ρ−1(e). ⊂ ⊂ ∈ ∈ t ∈ ∩ Given a t T and z tΞ ρ−1(e), we have z−1z Ξ2. Since ρ is a 10-local morphism and z−1, z, ∈ ∈ ∩ t ∈ t z−1z Ξ10, we have t ∈ ρ(z−1z)= ρ(z−1)ρ(z)= e z−1z ρ−1(e). t t ⇒ t ∈ Therefore z−1z Ξ2 ρ−1(e), hence z z (Ξ2 ρ−1(e)). This shows that t ∈ ∩ ∈ t ∩ tΞ ρ−1(e) z (Ξ2 ρ−1(e)). (5.1) ∩ ⊂ t ∩ Also note that (Ξ2 ρ−1(e))2 Ξ4 ρ−1(e). Indeed, if a,b (Ξ2 ρ−1(e))2, then ab Ξ4 and in ∩ ⊂ ∩ ∈ ∩ ∈ particular a,b,ab Ξ10. Since ρ is a 10-local morphism, this implies ρ(ab)= ρ(a)ρ(b)= e. ∈ Combining this observation with the fact that, by Proposition 2.13, we have Ξ4 F 3Ξ, we deduce ⊂ with (5.1) that (Ξ2 ρ−1(e))2 Ξ4 ρ−1(e) F 3Ξ ρ−1(e) = tΞ ρ−1(e) ∩ ⊂ ∩ ⊂ ∩ ∩ t∈T  [  2 −1 2 −1 zt(Ξ ρ (e)) = zt (Ξ ρ (e)). ⊂ ∩ ! ∩ t[∈T t[∈T Since T is finite, this shows that Ξ2 ρ−1(e) is an approximate subgroup of Ξ∞. ∩ The proof of part (ii) is analogous to (i). Indeed, set ρ := ρ and consider ker (ρ) = Ξk ρ−1(e). (6k−2) k ∩ Since Proposition 2.13 gives us Ξ2k F 2k−1Ξ, we define T := t F 2k−1 tΞ ρ−1(e) = , so that ⊂ { ∈ | ∩ 6 ∅} tΞ F 2k−1Ξ Ξ6k−2. By the same calculation as above we get ⊂ ⊂ (Ξk ρ−1(e))2 Ξ2k ρ−1(e) F 2k−1Ξ ρ−1(e) = tΞ ρ−1(e) ∩ ⊂ ∩ ⊂ ∩ ∩ t∈T  [  2 −1 2 −1 k −1 zt(Ξ ρ (e)) = zt (Ξ ρ (e)) zt (Ξ ρ (e)), ⊂ ∩ ! ∩ ⊂ ! ∩ t[∈T t[∈T t[∈T which proves that ker (ρ) is an approximate subgroup of Ξ∞. If one replaces the initial set (Ξk k ∩ ρ−1(e))2 by (Ξl ρ−1(e))2, for any l 2,...,k 1 , then the same calculation with the same ρ and ∩ ∈ { − } 5. PRE-IMAGES AND (PARTIAL) KERNELS 25 the appropriately adjusted set T yields that the partial kernels ker2(ρ),..., kerk−1(ρ) are approximate subgroups of Ξ∞. 

LEMMA 2.40 (Commensurability of partial kernels). Let (Ξ, Ξ∞) and (Λ, Λ∞) be approximate groups, with (Ξ, Ξ∞) countable. If ρ : (Ξ, Ξk+1) (Λ, Λk+1) is a (k + 1)-local morphism for some k 2, then k+1 → ≥ ker (ρ ) is left-syndetic in ker (ρ ), for all 2 j k. 2 k+1 j k+1 ≤ ≤ PROOF. We will assume without loss of generality that ρ : Ξ Λ is surjective. Since Ξ∞ is 1 → countable, it admits a proper left-invariant metric (see Lemma B.14), and we fix such a metric d once and for all. Now let ρ : (Ξ, Ξk+1) (Λ, Λk+1) be a (k + 1)-local morphism for some k 2. We k+1 → ≥ are going to show that there exists N N such that for all x ker (ρ ) and y ker (ρ ) we ∈ ∈ k k+1 ∈ 2 k+1 have d(x, y) < N. This will then imply d(y−1x, e) < N by left-invariance, and hence ker (ρ ) k k+1 ⊂ ker2(ρk+1)B(e,N), where B(e,N) denotes the ball of radius N around e. Since d is proper this will prove the lemma. Choose a finite set F Ξ∞ such that Ξk ΞF , which can be done by Proposition 2.13(i). We ⊂ ⊂ may assume without loss of generality that F Ξk+1. We then set C := max d(f,e). Given ⊂ 1 f∈F ξ ker (ρ ) = Ξk ρ−1 (e), we choose ξ Ξ and f F such that ξ = ξ f. Then d(ξ, ξ ) = ∈ k k+1 ∩ k+1 o ∈ ∈ o o d(ξ f, ξ ) = d(f,e) C . Moreover, since ρ is a (k + 1)-local morphism, ξ , f, ξ Ξk+1, and o o ≤ 1 k+1 0 ∈ ρk+1(ξ)= e, we have e = ρk+1(ξ)= ρk+1(ξo)ρk+1(f), and hence ρ (ξ )= ρ (ξ )= ρ (f)−1 (ρ (F ))−1 Λ. 1 o k+1 o k+1 ∈ k+1 ∩ Thus if we define F ′ := (ρ (F ))−1, then ξ ρ−1(F ′ Λ) and hence k+1 o ∈ 1 ∩ ker (ρ ) N (ρ−1(F ′ Λ)), (5.2) k k+1 ⊂ C1 1 ∩ where N denotes the C -neighborhood with respect to d. Since F is finite, so is F ′ Λ, say F ′ Λ= C1 1 ∩ ∩ λ ,...,λ . Since ρ is surjective, we find ξ ,...,ξ Ξ with ρ (ξ ) = λ for all i = 1,...,N. Now { 1 N } 1 1 N ∈ 1 i i let ξ′ ρ−1(λ ) and set ξ′′ := ξ′ξ−1. Since ξ′, ξ−1 Ξ, we have ξ′′ Ξ2, in particular, ξ, ξ′, ξ′′ Ξk+1. i ∈ 1 i i i i i i ∈ i ∈ ∈ We deduce that ρ (ξ′′)= ρ (ξ′)ρ (ξ−1)= λ λ−1 = e, and hence ξ′′ Ξ2 ρ−1 (e) = ker (ρ ). k+1 i k+1 i k+1 i i i i ∈ ∩ k+1 2 k+1 On the other hand, if we set C := max d(ξ ,e), then d(ξ′, ξ′′) = d(ξ′, ξ′ξ−1) = d(e, ξ−1) C . 2 i=1,...,N i i i i i i i ≤ 2 This shows that N ρ−1(F ′ Λ) = ρ−1(λ ) N (ker (ρ )), (5.3) 1 ∩ 1 i ⊂ C2 2 k+1 i=1 [ and combining (5.2) and (5.3) we obtain ker (ρ ) N (ker (ρ )), k k+1 ⊂ C1+C2 2 k+1 which finishes the proof. 

As a special case we see that if ρ : (Ξ, Ξ∞) (Λ, Λ∞) is a global morphism of countable approxi- → mate groups, then ker (ρ) is left-syndetic in ker (ρ) for all k 2. With Corollary 2.36, this implies that 2 k ≥ all of the partial kernels ker2(ρ), ker3(ρ), . . . are pairwise commensurable approximate subgroups of Ξ∞. As a special case we may record:

COROLLARY 2.41. A morphism ρ : (Ξ, Ξ∞) (Λ, Λ∞) of countable approximate groups has finite → kernels if and only if the second partial kernel ker2(ρ) is finite.  Also, Lemma 2.39 combined with Lemma 2.40 gives us the following:

COROLLARY 2.42. Let (Ξ, Ξ∞) and (Λ, Λ∞) be approximate groups, with (Ξ, Ξ∞) countable, and let ρ : (Ξ, Ξk) (Λ, Λk) be a k-local morphism for some k 10. Then all of the partial kernels k → ≥ ker2(ρk),..., ker k+2 (ρk) ⌊ 6 ⌋ are pairwise commensurable approximate subgroups of Ξ∞.  26 2. BASIC CONSTRUCTIONS OF APPROXIMATE GROUPS

6. Extensions of approximate groups Recall that in group theory, an extension of a group G by a group N is a short exact sequence of the p form 1 N ֒ G G 1, i.e. p : G G is a surjective group homomorphism and N = ker(p). By → → → → → ∼ abuse of language we will also refer to p : G G itself as a group extension of G. In analogy with the → group case we define:e e e DEFINITION 2.43. (i) A morphism p : (Ξ, Ξ∞) (Λ, Λ∞) between approximate groups is called → an extension of (Λ, Λ∞) if p : Ξ Λ is surjective. 1 → (ii) An extension p : (Ξ, Ξ∞) (Λ, Λ∞) is called a finite extension if p has finite kernels in the sense → of Definition 2.21. ∞ ∞ (iii) An extension p : (Ξ, Ξ ) (Λ, Λ ) is called minimal if p1 : Ξ Λ is bijective. ∞ → ∞ → ∞ (iv) If (Ξo, Ξo ) is an approximate subgroup of (Ξ, Ξ ) which surjects onto (Λ, Λ ), then the restric- tion p : (Ξ , Ξ∞) (Λ, Λ∞) is called a subextension of p. o o o → (v) Given an approximate group (Λ, Λ∞), a group Γ and a group extension p : Γ Λ∞ we denote → by Ext(p, (Λ, Λ∞)) the class of all extensions (Ξ, Ξ∞) (Λ, Λ∞) with Ξ∞ =Γ whose underlying → group extension is given by p. (vi) An extension p : (Ξ, Ξ∞) (Λ, Λ∞) is called maximal, if every (Ξ , Ξ∞) Ext(p , (Λ, Λ∞)) is a → o o ∈ ∞ subextension of p. The problem of classifying all extensions of (Λ, Λ∞) splits into the two sub-problems of classifying all group extension of p :Γ Λ∞; • → given a group extension p :Γ Λ∞, describing the class Ext(p, (Λ, Λ∞)). • → The first problem is purely group-theoretic. Concerning the second problem we observe: COROLLARY 2.44 (Maximal extension associated with a group extension). Let (Λ, Λ∞) be an ap- proximate group and p : Γ Λ∞ be a group extension. Define Ξ := p−1(Λ) Γ. Then (Ξ , Ξ∞ ) is → max ⊂ max max the unique maximal extension in Ext(p, (Λ, Λ∞)). ∞ ∞ PROOF. By Corollary 2.32 the pair (Ξmax, Ξmax) is an approximate group and Ξmax = Γ. Since p(Ξ )=Λ we have (Ξ , Ξ∞ ) Ext(p, (Λ, Λ∞)), and maximality follows from the fact that max max max ∈ p−1(Λ) is the largest subset of Γ which maps to Λ.  Minimal extensions of approximate groups are closely related to bounded group extensions as defined by Heuer [Heu20]. We briefly recall the relevant definition: p DEFINITION 2.45. A group extension 1 ker(p) Γ Λ∞ 1 is called bounded if there is a ∞ → → −→ → ∞ section σ :Λ Γ of p such that σ is a quasimorphism and the induced map ϕσ :Λ Aut(ker(p)) → −1 → given by ϕσ(g)(h)= σ(g)hσ(g) has finite image. In this case σ is called a good section. Bounded group extensions are classified by bounded group cohomology in a similar way in which group extensions are classified by usual group cohomology. We will need the following refinement: p DEFINITION 2.46. A bounded group extension 1 ker(p) Γ Λ∞ 1 is called symmetrically → → −→ → bounded if it admits a good section which is moreover symmetric. p Not every group extension 1 ker(p) Γ Λ∞ 1 admits a symmetric section, but it al- → → −→ → ways does when Λ∞ contains no 2-torsion [Sta09, Cor. 3.3]. In general, the classification of extensions admitting symmetric sections is related to symmetric group cohomology [Sta09]. Similarly, symmet- rically bounded extensions should be related to “symmetric bounded group cohomology”; we will, however, not pursue this direction here. Instead we confine ourselves with the following observation, which relates minimal extensions to (symmetric) bounded cohomology. COROLLARY 2.47. Let (Λ, Λ∞) be an approximate group, let p : Γ Λ∞ be a symmetrically bounded → group extension and let σ be a good section. Then Ξ := σ(Λ) is an approximate subgroup of Γ and p induces an extension p : (Ξ, Ξ∞) (Λ, Λ∞), which is a minimal extension. → PROOF. Ξ is an approximate subgroup by Proposition 2.27, and p1(Ξ) = Λ by construction.  7. MODEL SETS AND APPROXIMATE LATTICES 27

7. Model sets and approximate lattices relatively compact symmetric identity neighborhoods in locally compact groups are trivial exam- ples of approximate subgroups. We will see in this subsection how these can be used to construct interesting examples of approximate groups.

REMARK 2.48 (Countable approximate subgroups from identity neighborhoods). Recall from Ex- ample 2.10 that if H is a locally compact group and W is a relatively compact symmetric identity neighborhood in H, then (W, W ∞) is an approximate group. If H is discrete, then W is just a finite set, so assume that H is non-discrete. In this case, W will be an uncountable approximate group. There are several ways to obtain a countable approximate subgroup out of H. For example, if Γ H is a countable subgroup, then by Lemma 2.26 the intersection W 2 Γ is a countable approximate ⊂ ∩ subgroup of Γ (hence of H). Slightly more generally, if Γ is a countable group and ρ : Γ H is a → group homomorphism, then by Proposition 2.30 the pre-image ρ−1(W 2) is a countable approximate subgroup of Γ. The following lemma shows that one can actually get rid of the exponent 2.

LEMMA 2.49. Let W be a relatively compact symmetric identity neighborhood in a locally compact group H, and let ρ :Γ H be a group homomorphism. Then ρ−1(W ) is an approximate subgroup of Γ. → PROOF. We may replace H and W by the closure H′ of ρ(Γ) and the intersection W H′ respec- ∩ tively and thereby assume without loss of generality (since ρ(Γ) is dense in H) that ρ(Γ) Int W = H. By compactness we then find a finite subset F Γ such that W 2 ρ(F ) Int W ρ(F )W . Using the ⊂ ⊂ ⊂ fact that ρ is a group homomorphism, we conclude that ρ−1(W )2 ρ−1(W 2) ρ−1(ρ(F )W ) Fρ−1(W ), ⊂ ⊂ ⊂ This shows that ρ−1(W ) satisfies (AG2). Since it also satisfies (AG1) by construction, it is thus an approximate subgroup. 

The following special case is of particular importance.

EXAMPLE 2.50 (Cut-and-project construction). Let G, H be locally compact groups, and denote by proj : G H G and proj : G H H the canonical projections. Given a subgroup Γ < G H we G × → H × → × denote its projections by Γ := proj (Γ) and Γ := proj (Γ). Assume that the restriction p :Γ Γ G G H H → G of projG is injective, and define τ := proj p−1 :Γ H. H ◦ G → Then τ is a homomorphism, and thus for every relatively compact symmetric identity neighborhood W in H the subset Λ(Γ, W ) := τ −1(W ) is an approximate subgroup of Γ. Explicitly, we have Λ(Γ, W ) = proj (Γ (G W )), G ∩ × i.e. Λ(Γ, W ) arises from Γ by first cutting it with the “strip” (G W ) in G H, and then projecting × × down to G. We thus call Λ(Γ, W ) a cut-and-project set.

The following definition from [BH18a, BHP18] generalizes a classical definition of Meyer [Mey72].

DEFINITION 2.51. If Γ is a lattice in G H, then Λ(G, W ) is called a model set in G. If this lattice is × moreover cocompact, then it is called a uniform model set. A syndetic subset of a (uniform) model set is called (uniform) Meyer set. The following related notion was introduced in [BH18a]:

DEFINITION 2.52. An approximate subgroup Λ of an lcsc group G is called a uniform approximate lattice if it is discrete and syndetic. 28 2. BASIC CONSTRUCTIONS OF APPROXIMATE GROUPS

REMARK 2.53 (Alternative characterizations of uniform approximate lattice). As explained in [BH18a], an approximate subgroup Λ of a lcsc group G is a uniform approximate lattice if it is a Delone subset of G in the sense of Definition B.9 with respect to some left-admissible metric d on G. It is then in fact a Delone set with respect to any left-admissible metric d on G. We will see later (see Theorem 4.7 below) that uniform approximate lattices are closely related to to geometric isometric actions of approximate groups. The following was established in [BH18a].

PROPOSITION 2.54 (Approximate lattices from Meyer sets). Every uniform Meyer set is a uniform approximate lattice. It is linguistically convenient, though an epistemological accident, that the notion of a model set is also closely related to that of a good model of an approximate group as introduced by Machado [Mac20b], building on the pioneering work of Hrushovski [Hru12] and defined by analogy with the notion of a good model for an ultra approximate group as introduced by Breuillard, Green and Tao in their study of finite approximate groups [BGT12, Def. 3.5].

DEFINITION 2.55. Let (Λ, Λ∞) be an approximate group. A good model for (Λ, Λ∞) is a global morphism (Λ, Λ∞) H into a locally compact group H such that → (M1) ρ(Λ) is relatively compact in H (and hence all partial images are relatively compact); (M2) there exists an identity neighborhood U such that ρ−1(U) Λ. ⊂ The following was established in [Mac20b, Proposition 1 and Theorem 1]:

THEOREM 2.56 (Machado). For a uniform approximate lattice Λ in an lcsc group G the following are equivalent: (i) Λ is a uniform Meyer set. (ii) Λ is a commensurable to an approximate subgroup Λ′ < G such that (Λ′, (Λ′)∞) admits a good model. (iii) There exists a sequence of approximate subgroups (Λn)n∈N in G which are commensurable to Λ such that Λ2 Λ for all n N. n+1 ⊂ n ∈ Both Proposition 2.54 and Theorem 2.56 have versions for non-uniform Meyer sets, but we will not need those in our present context. In the case where G is abelian, every uniform approximate lattice is a Meyer set by a classical theorem of Meyer. This theorem has recently been generalized to include all amenable [Mac20b] and also certain non-amenable groups [Mac20a, Hru20]. CHAPTER 3

Large-scale geometry of approximate groups

As recalled in Appendix B, every countable group Γ gives rise to a canonical coarse equivalence class [Γ]c of metric spaces, and if Γ is finitely-generated, then this coarse equivalence class can be refined into a canonical QI type [Γ] [Γ] of metric spaces. ⊂ c In this chapter we generalize these notions to countable approximate groups. The definition of a canonical coarse class of a countable approximate group is in complete analogy with the group case. However, there are two competing definitions of a canonical QI type for countable approximate groups, and as we will see, the interplay between these two definitions is at the heart of geometric approximate group theory. Since every countable approximate group gives rise to a canonical coarse class, we can use coarse invariants of metric spaces to study countable approximate groups. Similarly, we can use QI invariants to study countable approximate groups which are finitely-generated in a suitable sense. We discuss some examples of such invariants, most of which will reappear in later chapters. We will freely use the language of (metric) coarse geometry; our terminology concerning various classes of coarse maps is summarized in Appendix A. We will also assume some basic facts concerning geometric group theory, not just for finitely-generated groups, but more generally for countable and locally compact second countable (lcsc) groups. See Appendix B for background and notation.

1. The coarse space associated with a countable approximate group Our first goal in this chapter is to associate with each countable approximate group (Λ, Λ∞) a canonical coarse equivalence class [Λ]c of metric spaces, generalizing the construction in the group case summarized in Section 2 of Appendix B. Our starting point is the following easy observation; see Definition B.3 for the notion of a left-admissible metric on a countable group.

LEMMA 3.1. Let Γ be a countable group, Λ Γ be a subset and d and d′ be left-admissible metrics on Γ. ⊂ (i) The identity (Λ, d ) (Λ, d′ ) is a coarse equivalence. |Λ×Λ → |Λ×Λ (ii) If d and d′ are moreover large-scale geodesic (e.g. word metrics with respect to finite symmetric generating sets), then the identity (Λ, d ) (Λ, d′ ) is a quasi-isometry. |Λ×Λ → |Λ×Λ PROOF. The identical map (Γ, d) (Γ, d′) is a coarse equivalence by Corollary B.11, and if d and → d′ are large-scale geodesic, then it is a quasi-isometry by Corollary A.30. Restricting to (Λ, d) we obtain a coarse embedding, respectively a quasi-isometric embedding. Since the image of Λ is given by Λ, the result follows. 

By Lemma 3.1.(i) the following notion is well-defined, i.e. independent of the choice of left-admis- sible metric d:

DEFINITION 3.2. Let (Λ, Λ∞) be a countable approximate group and let d be a left-admissible metric on Λ∞. Then [Λ] := [((Λ, d )] c |Λ×Λ c is called the canonical coarse class of Λ.

REMARK 3.3 (Independence of the ambient group). Assume that (Λ, Λ∞) is an approximate sub- group of a countable group Γ. If d is a left-admissible metric on Γ, then d ∞ ∞ is a left-admissible |Λ ×Λ 29 30 3. LARGE-SCALE GEOMETRY OF APPROXIMATE GROUPS metric on Λ∞, and hence [Λ] := [(Λ, d )] . c |Λ×Λ c Thus the canonical coarse class of Λ is intrinsic in the sense that it is independent of the ambient group used to define it. Note that, since the restriction of a left-admissible metric is proper, the class [Λ]c admits a repre- sentative which is a proper metric space. Also note that for the definition of the canonical coarse class we do not use that Λ is an approximate subgroup, so one could as well associate a canonical coarse class of (pseudo-)metric spaces with an arbitrary subset of a countable group. The following four lemmas record basic properties of the coarse equivalence classes of approxi- mate groups.

LEMMA 3.4. If (Λ, Λ∞) is a countable approximate group, then for every k 1 the inclusion Λ ֒ Λk ≥ → induces a coarse equivalence [Λ] [Λk] . c → c PROOF. If d is any left-admissible metric on Λ∞, then the isometric embedding Λ ֒ Λk is coarsely → onto. Indeed, if Λ2 ΛF , then Λk ΛF k−1 and thus for every g Λk there exists h Λ and f F k−1 ⊂ ⊂ ∈ ∈ ∈ such that g = hf and hence d(g,h)= d(hf,h)= d(f,e). k Thus if C := max k−1 d(f,e), then Λ N (Λ).  f∈F ⊂ C LEMMA 3.5. Let (Ξ, Ξ∞) be an approximate subgroup of a countable approximate group (Λ, Λ∞). Then . [the inclusion Ξ ֒ Λ induces a coarse embedding [Ξ] [Λ → c → c PROOF. Let d be a left-admissible metric on Λ∞. Then the restriction of d to Ξ∞ is a left-admissible metric, and hence the restrictions of d to Λ and Ξ respectively are left-admissible metrics. It follows that [Λ] and [Ξ] are represented by (Λ, d ) and (Ξ, d ) respectively, and since the inclusion (Ξ, d ) c c |Λ |Ξ |Ξ → (Λ, d ) is isometric, the lemma follows.  |Λ By Lemma A.5, a map h : X Y between pseudo-metric spaces is coarsely Lipschitz if and only → if for every t 0 there exists s 0 such that if x, x′ X satisfy d (x, x′) t, then d (f(x),f(x′)) s. ≥ ≥ ∈ X ≤ Y ≤ In the language of coarse geometry (see Remark A.9), this means that h is bornologous with respect to the pseudo-metric coarse structures on X and Y . In particular, this notion depends only on the coarse equivalence classes of X and Y , respectively.

LEMMA 3.6. Let (Ξ, Ξ∞), (Λ, Λ∞) be countable approximate groups and let f : (Ξ, Ξ2) (Λ, Λ2) be 2 → a 2-local quasimorphism. Then the restriction f : Ξ Λ is coarsely Lipschitz with respect to the canonical 1 → coarse classes of Ξ and Λ.

PROOF. Fix left-admissible metrics d on Ξ∞ and d′ on Λ∞ and denote by and ′ the corre- k·k k·k sponding norms. Given t,s > 0 we define B¯(t) := ξ Ξ2 ξ t and B¯′(s) := λ Λ2 λ ′ s . { ∈ |k k≤ } { ∈ |k k ≤ } Since f2 is a 2-local quasimorphism, it follows from (5.5) that there exists a finite set D = D2(f2) such that for all g,h Ξ we have f (g−1)f (h) f (g−1h)D−1, and consequently ∈ 1 1 ∈ 2 ′ −1 −1 ′ −1 −1 −1 ′ d (f1(g )f1(h),f2(g h)) = d (f2(g h) f1(g )f1(h),e) max d (x, e) =: C < . ≤ x∈D−1 ∞

Since d is proper, the set B¯(t) and thus also f2(B¯(t)) is finite, for every t> 0. We deduce that for every t> 0 there exists s> 0 such that f (B¯(t)) B¯′(s) ξ Ξ2 : ξ t f (ξ) ′ s 2 ⊂ ⇒ ∀ ∈ k k≤ ⇒ k 2 k ≤ g,h Ξ: g−1h t f (g−1h) ′ s ⇒ ∀ ∈ k k≤ ⇒ k 2 k ≤ g,h Ξ: g−1h t f (g−1)f (h) ′ s + C ⇒ ∀ ∈ k k≤ ⇒ k 1 1 k ≤ g,h Ξ: d(g,h) t d′(f (g),f (h)) s + C. ⇒ ∀ ∈ ≤ ⇒ 1 1 ≤ In view of Lemma A.5 this finishes the proof.  2. COARSE INVARIANTS OF COUNTABLE APPROXIMATE GROUPS 31

Combining this lemma with Lemma A.7 we deduce:

COROLLARY 3.7. Let (Ξ, Ξ∞), (Λ, Λ∞) be countable approximate groups and let f : (Ξ, Ξ2) (Λ, Λ2) 2 → be a 2-local (quasi-)isomorphism. Then [Ξ]c = [Λ]c.  We express this corollary by saying that the canonical coarse class of a countable approximate group depends only on its 2-local isomorphism type. Note that, by Remark 2.19, every 3-Freiman isomorphism is a 2-local isomorphism, and hence the coarse class only depends on the 3-Freiman isomorphism type in the same sense.

LEMMA 3.8. Let Λ, Λ′ be approximate subgroups of a countable group Γ. If Λ and Λ′ are commensurable, ′ then [Λ]c = [Λ ]c.

PROOF. Let d be a left-admissible metric on Γ and let F , F Γ be finite such that Λ Λ′F 1 2 ⊂ ⊂ 1 and Λ′ ΛF . Then every λ Λ can be written as λ = λ′f for some λ′ Λ′ and f F and ⊂ 2 ∈ ∈ ∈ 1 hence d(λ, λ′) = d(f,e), which is uniformly bounded. This shows that Λ is contained in a bounded d-neighborhood of Λ′, and by a symmetric argument Λ′ is contained in a bounded d-neighborhood of ′ Λ. This implies that (Λ, d ) and (Λ , d ′ ′ ) are coarsely equivalent.  |Λ×Λ |Λ ×Λ EXAMPLE 3.9 (Coarse kernel of a (local) morphism). Let (Ξ, Ξ∞), (Λ, Λ∞) be approximate groups, with (Ξ, Ξ∞) countable, and let ρ : (Ξ, Ξk) (Λ, Λk) be a k-local morphism, for some k 10. Then k → ≥ by Corollary 2.42 and Lemma 3.8 we have

[ker2(ρk)]c = [ker3(ρk)]c = = [ker⌊ k+2 ⌋(ρk)]c. ··· 6

This common coarse equivalence class will be called the coarse kernel of ρk and denoted [ker(ρk)]c. The coarse kernel [ker(ρ)]c of a global morphism ρ is defined similarly.

2. Coarse invariants of countable approximate groups By a coarse invariant of (proper) pseudo-metric spaces we mean an assignment which assigns to I each (proper) pseudo-metric space (X, d) some object (X, d) such that (X, d) = (Y, d) whenever I I I (X, d) and (Y, d) are coarsely equivalent. For example, every functor on the category of (metrizable) coarse spaces defines a coarse invariant, but there are also interesting numerical coarse invariants.

REMARK 3.10 (Coarse invariants of approximate groups). If is a coarse invariant of proper I pseudo-metric spaces and (Λ, Λ∞) is a countable approximate group, then we define (Λ) := (X, d), I I where (X, d) is any proper representative of [Λ] . By Corollary 3.7 we then have (Λ) = (Ξ) whenever c I I (Λ, Λ∞) and (Ξ, Ξ∞) are 2-locally isomorphic (for example, 3-Freiman isomorphic). In particular, is I an invariant in the sense that (Λ) = (Ξ) whenever (Λ, Λ∞) and (Ξ, Ξ∞) are (globally) isomorphic. I I We refer to such an invariant as a coarse invariant.

An example of a numerical coarse invariant is asymptotic dimension:

DEFINITION 3.11. Let n N . A metric space (X, d) has asymptotic dimension asdim X n if for ∈ 0 ≤ every R> 0 there exist collections (0),..., (n) of subsets of X with the following properties: U U (ASD1) n (i) is a cover of X; i=0 U (ASD2) For every i =0,...,n the collection (i) is uniformly bounded, i.e. there exists D > 0 such that S U for all U (i) we have diam U D; ∈U ≤ (ASD3) For every i = 0,...,n the collection (i) is R-disjoint, i.e. for all U, V (i) we have either U ∈ U U = V or d(U, V ) R. ≥ We say that asdim X = n if asdim X n, but not asdim X n 1. If there is no such n N so that ≤ ≤ − ∈ 0 the properties are satisfied for every R> 0, we say that asdim X = . ∞ 32 3. LARGE-SCALE GEOMETRY OF APPROXIMATE GROUPS

It evidently suffices for (X, d) to have the above properties for all sufficiently large R. Our def- inition of asymptotic dimension is one of many equivalent definitions, see [BD08]. It is sometimes referred to as the coloring definition of asdim, because the union of collections (0),..., (n) can be re- U U garded as one cover of X, which is uniformly bounded, and can be separated into (n+1) subfamilies U (i), each of which has its own color i, and is R-disjoint. Coarse invariance of asymptotic dimension U is established in the following lemma:

LEMMA 3.12. (i) If X is a subspace of Y , then asdim X asdim Y . ≤ (ii) If X and Y are coarsely equivalent, then asdim X = asdim Y . (iii) If X embeds coarsely into Y , then asdim X asdim Y . ≤ (0) (n) PROOF. (i) Let asdim Y = n. If R > 0, and ,..., satisfy (ASD1)–(ASD3) above, then the UY UY collections (i) := U X U (i) ,i =0, ,n UX { ∩ | ∈UY } ··· satisfy (ASD1)–(ASD3) for X, so asdim X n. ≤ (ii) Assume that asdim X = n, and that f : X Y is a coarse equivalence whose image is K- → relatively dense in Y . Given (0),..., (n) which satisfy (ASD1)–(ASD3) above for constants D and UX UX R, one can check that (i) := N (f(U)) U (i) ,i =0, ,n UY { K | ∈UX } ··· satisfiy (ASD1)–(ASD3) with constants D′ =Φ (D)+2K and R′ =Φ (D) 2K, where Φ and Φ are + − − − + lower and upper controls for f. This implies that asdim Y n = asdim X, and the opposite inequality ≤ follows by a symmetric argument. (iii) follows from (i) and (ii).  We thus obtain a first coarse numerical invariant of countable approximate groups:

DEFINITION 3.13. The asymptotic dimension of a countable approximate group (Λ, Λ∞) is

asdim Λ := asdim([Λ]c). By Remark 3.10 the asymptotic dimension of a countable approximate groups is invariant under 2-local (quasi-)isomorphism (in particular, Freiman 3-isomorphisms). We will discuss asymptotic di- mension of countable approximate groups in detail in Chapter 7 below. For later reference we record the following consequences of Lemmas 3.4, 3.5 and 3.12:

COROLLARY 3.14 (Monotonicity of asymptotic dimension for countable approximate groups). Let (Λ, Λ∞) be a countable approximate group, and let (Ξ, Ξ∞) (Λ, Λ∞) be an approximate subgroup. Then the ⊂ following hold: (i) asdim Λ asdim Λ∞. ≤ (ii) asdim Λk = asdim Λ for all k 1. ≥ (iii) asdim Ξ asdim Λ.  ≤ The numerical coarse invariant asdim is closely related to a more geometric coarse invariant which takes values in homeomorphism types of compact spaces:

EXAMPLE 3.15 (Higson corona). Let (X, d) be a proper metric space; a bounded continuous func- tion f : X C is called a Higson function if for all x X and r> 0 we have → ∈ lim sup f(x) f(y) d(x, y) r =0. x→∞ {| − | | ≤ } ∗ The Higson functions form a unital C -subalgebra Ch(X) of Cb(X), hence its Gelfand spectrum hX := spec(Ch(X)) defines a compactification of X, called the Higson compactification and the boundary νX := hX X is called the Higson corona of X. By[Roe03, Cor. 2.42] the assignment which assigns to \ I a proper metric space X the homeomorphism type [νX] of its Higson corona is a coarse invariant. We thus define the Higson corona of a countable approximate group (Λ, Λ∞) as νΛ := [νX], where (X, d) 2. COARSE INVARIANTS OF COUNTABLE APPROXIMATE GROUPS 33 is any proper representative of [Λ] . As a special case of [Dra00, Thm. 6.2] we see that if asdim Λ < , c ∞ then asdim Λ = dim νΛ, where dim denotes topological (or covering) dimension. Thus for countable approximate groups of finite asymptotic dimension, the asymptotic dimension can be recovered from the Higson corona. Here is a related coarse invariant:

EXAMPLE 3.16 (Space of coarse ends). With every metric space (X, d) one can associate via Con- struction A.38 a topological space (X), whose elements are called the coarse ends of X. By Theorem E∞ A.39 the homeomorphism type of (X) is a coarse invariant of X. In particular, we may define E∞ (Λ) as the homeomorphism type of (X) where X is some (hence any) representative of [Λ] . E∞ E∞ c It turns out that the possible homeomorphism types of spaces of coarse ends of approximate groups are severely restricted. To give a flavor of the results to come, we mention the following result; see Definition 3.28 below for the notion of a geometrically finitely-generated approximate group.

COROLLARY 3.17. If (Λ, Λ∞) is a geometrically finitely-generated approximate groups and (Λ) con- E∞ tains at least 3 points, then it is uncountable and homeomorphic to a Cantor space. In particular, a geometrically finitely-generated approximate group has either 0, 1, 2 or uncountably many coarse ends.

PROOF. We will later see that if (Λ, Λ∞) is geometrically finitely-generated in the sense of Defi- nition 3.28, then [Λ]c admits a representative which is quasi-cobounded, large-scale geodesic and of coarse bounded geometry (see Proposition 3.36 and Corollary 3.49). The corollary then follows from Theorem A.44.  Other coarse invariants can be defined using the Vietoris–Rips complex. The following presenta- tion follows [HW21].

CONSTRUCTION 3.18 (Vietoris–Rips complex). Let (X, d) be a metric space. For r 0, the Vietoris– ≥ (0) Rips complex VRr(X) is the following abstract simplicial complex: The vertex set is VRr(X) := X and if x ,...,x VR (X)(0) are vertices, then (x ,...,x ) is an n-simplex of VR (X) if d(x , x ) r 0 n ∈ r 0 n r i j ≤ for all i, j 0,...,n . We denote by VR (X) the geometric realization of VR (X). ∈{ } | r | r Given s>r> 0 we denote by ιs : VR (X) ֒ VR (X) the natural inclusion map, which is the r | r | → | s | identity on the set of vertices.

LEMMA 3.19. Let (X, d), (Y, d) be metric spaces and let ϕ, ψ : X Y be coarse Lipschitz maps which → are at bounded distance. Then for every p> 0 there exists q = q(p) > 0 with the following properties: (i) The maps ϕ, ψ extend to simplicial maps ϕ , ψ : VR (X) VR (Y ) . ∗ ∗ | p | → | q | (ii) The maps ϕ∗ and ψ∗ are homotopic.

PROOF. Let Φ be an upper control function for both ϕ and ψ and let C 0 be such that ϕ and ψ ≥ are C-close. We then define q(p) := Φ(p)+ C. (i) Let (x ,...,x ) be a simplex in VR (X). The for all i, j 0,...,n we have d(x , x ) p, and 0 n p ∈{ } i j ≤ hence d(ϕ(x ), ϕ(x )) Φ(p) q(p). i j ≤ ≤ This shows that ϕ∗(x0,...,xn) := (ϕ(x0),...,ϕ(xn)) is a simplex in VRq(p)(Y ), and hence we obtain the desired extension ϕ∗. The same argument also applies to ψ. (ii) If (x ,...,x ) is a simplex in VR (X), then for all i, j 0,...,n we have 0 n p ∈{ } d(ϕ(x ), ψ(x )) d(ϕ(x ), ϕ(x )) + d(ϕ(x ), ψ(x )) Φ(p)+ C = q(p), i j ≤ i j j j ≤ hence (ϕ(x0),...,ϕ(xn), ψ(x0),...,ψ(xn)) is a simplex in VRq(p)(Y ). This shows that if σ is the re- alization of the simplex (x0,...,xn), then ϕ∗(σ) and ψ∗(σ) are contained in a common simplex. The lemma follows.  34 3. LARGE-SCALE GEOMETRY OF APPROXIMATE GROUPS

REMARK 3.20 (K-acyclicity). In the sequel, by a homotopy functor K we mean a functor from the homotopy category of triangulizable topological spaces (i.e. spaces which arise as geometric realiza- tions of abstract simplicial complexes) into some category . We fix such a homotopy functor K and C assume that admits a final object 0. C As in Remark A.6 we define a category Coa0 of coarse metric spaces as follows: Objects in Coa0 are non-empty pseudo-metric spaces and morphisms are closeness classes of coarse Lipschitz maps. By Lemma A.7 two objects in Coa0 are isomorphic if and only if they are coarsely equivalent. We say that an object X in Coa0 is K-acyclic if the following holds: For every r > 0 there exists s> 0 such that the morphism

Ks := K(ιs): K( VR (X) ) K( VR (X) ). r r | r | → | s | .induced by the inclusion map ιs : VR (X) ֒ VR (X) factors through 0 r | r | → | s | It follows from Lemma 3.19 that if X and Y are isomorphic in Coa0, then X is K-acyclic if and only if Y is. In fact, we have the following more general statement.

LEMMA 3.21. Let [f]: X Y be a morphism in Coa which is represented by a coarse Lipschitz map f. → 0 If [f] admits a left-inverse [g] and Y is K-acyclic, then so is X.

PROOF. Let p > 0 and using Lemma 3.19.(i) choose r > 0 so large that f induces a map f∗ : ( VR (X) VR (Y ) . By K-acyclicity of Y we then find s> 0 such that ιs : VR (Y ) ֒ VR (Y | p | → | r | r | r | → | s | induces a morphism Ks : K( VR (Y ) ) K( VR (Y ) ) factoring through 0. Finally, choose t > 0 r | r | → | s | such that g induces a morphism VR (Y ) VR (X) . Then g f induces the trivial morphism | s | → | t | ◦ K((f g) )= K(g ) Ks K(f ): K( VR (X) ) K( VR (X) ). ◦ ∗ ∗ ◦ r ◦ ∗ | p | → | t |

Since g f IdX , it follows from Lemma 3.19.(ii) (after enlarging t if necessary), that (f g)∗ ◦ t ∼ ◦ ≃ (IdX )∗ = ιp. The lemma follows. 

We may thus define:

DEFINITION 3.22. A countable approximate group (Λ, Λ∞) is called K-acyclic with respect to a homotopy functor K if some (hence any) representative of [Λ]c is K-acyclic.

REMARK 3.23 (Basepoints). The following modification of K-acyclicity is sometimes useful. As- sume that K is a functor from the homotopy category of pointed topological spaces. We then say that an object X in Coa0 is K-acyclic, if the morphism

Ks := K(ιs): K( VR (X) , ) K( VR (X) , ). r r | r | ∗ → | s | ∗ -induced by the inclusion map ιs : VR (X) ֒ VR (X) factors through 0 for every choice of base r | r | → | s | point X = V X (0) = V X (0). K-acyclicity of approximate groups is then defined accordingly. ∗ ∈ | r | | s | The following invariants are studied in [HW21]:

EXAMPLE 3.24 (Finiteness properties of approximate groups). We say that a countable approxi- mate group (Λ, Λ∞) has the finiteness property (F ) if it is π -acyclic for all k 0,...,n 1 , where n k ∈{ − } π denotes the kth homotopy group (note that k goes from 0 to n 1, and not to n). It has finiteness k − property (FP ) if it is H -acyclic for all k 0,...,n 1 , where H denotes the kth simplicial (or n k ∈ { − } k singular) homology group. It is established in [HW21] that in the case where Λ=Λ∞ is a group, this definition is equivalent to the usual definition of finiteness properties of countable groups. By the Hurewicz theorem, πk-acyclicity implies Hk-acyclicity for every k, hence Property (Fn) implies Property (FPn) for every n. Similarly, Properties (F2) and (FPn) together imply Property (Fn). 3. EXTERNAL AND INTERNAL QI TYPE 35

3. External and internal QI type

For finitely generated groups Γ, the canonical coarse class [Γ]c contains a canonical QI type [Γ] which is represented e.g. by any Cayley graph of Γ. The fact that a countable group Γ is finitely generated can be characterized in many different ways, see Corollary B.21. These different charac- terizations give rise to inequivalent definitions of finite generation for countable approximate groups and, consequently, different QI types. The simplest and most algebraic definition of being finitely generated for a countable approximate group (Λ, Λ∞) is the one introduced in [BH18a]; while it has many deficits from the geometric point of view, it will still turn out to be useful for us.

DEFINITION 3.25. An approximate group (Λ, Λ∞) is called algebraically finitely-generated if Λ∞ is finitely generated as a group.

If (Λ, Λ∞) is an algebraically finitely-generated approximate group, then Λ∞ admits a large-scale geodesic left-admissible metric d; for example, every word metric with respect to a finite generating set is large-scale geodesic and left-admissible. We then refer to the metric d as an external metric |Λ×Λ on Λ. If d′ is any other large-scale geodesic left-admissible metric on Λ∞, then by Corollary B.21 the identity map (Λ∞, d) (Λ∞, d′) is a quasi-isometry. It thus restricts to a quasi-isometry (Λ, d ) → |Λ×Λ → (Λ, d′ ). To summarize: |Λ×Λ PROPOSITION 3.26. If (Λ, Λ∞) is an algebraically finitely-generated approximate group, then any two external metrics on Λ define the same QI type. 

DEFINITION 3.27. If (Λ, Λ∞) is algebraically finitely generated and d is an external metric on Λ, ∞ then the QI type of (Λ, d) is denoted [Λ]ext and referred to as the external QI type of (Λ, Λ ).

Note that, by definition, [Λ] [Λ] . A different, more geometric refinement of [Λ] is based on ext ⊂ c c finiteness properties (see Example 3.24).

DEFINITION 3.28. A countable approximate group (Λ, Λ∞) is called geometrically finitely generated if it has finiteness property (F1). In order to associate a QI type with every geometrically finitely generated approximate group (Λ, Λ∞), we need the following observation:

THEOREM 3.29 (Large-scale geodesic representatives). Let (Λ, Λ∞) be a countable approximate group, let d be a left-admissible metric on Λ∞ and let d = d . Then the following are equivalent: |Λ×Λ ∞ (i) (Λ, Λ ) has finiteness property (F1). (ii)b There exists a proper metric d′ on Λ which isb coarsely equivalent to d and such that (Λ, d′) is large-scale geodesic. (iii) There exists a representative X [Λ] which is large-scale geodesic. ∈ c (iv) Every representative X [Λ] is coarsely connected. ∈ c (v) (Λ, d) is coarsely connected.

PROOF. (i) (v): By definition, (Λ, Λ∞) has finiteness property (F ) iff for every r > 0 there ⇐⇒ 1 exists s> 0 such that the map (π )s : π ( VR ) π ( VR ) 0 r 0 | r | → 0 | s | is trivial. This property only depends on the 1-skeleton of the respective complexes, and it means that any two points x, y Λ= VR (0) can be connected by a pathin VR (1). The latter means that there ∈ | r | | s | exist x ,...,x with x = x and x = y such that (x , x ) is an edge in VR , for i 0,...,n 1 0 n 0 n i i+1 | s | ∈{ − } or equivalently d(x , x ) s. This, however, means precisely that (Λ, d) is s-coarsely connected. i i+1 ≤ (ii) = (iii) = (iv) = (v): These follow from Lemma A.26 and the fact that every space which ⇒ ⇒ ⇒ is quasi-isometric to (Λ, d) is contained in [Λ]c. 36 3. LARGE-SCALE GEOMETRY OF APPROXIMATE GROUPS

(v) = (ii) We fix C > 0 such that (Λ, d) is coarsely C -connected. For every C C we then ⇒ 0 0 ≥ 0 obtain a large-scale geodesic metric dC on Λ by setting d (x, y) := min n N x ,...,x Λ: x = x ,y = x , d(x , x ) C for all i 1,...,n . C { ∈ 0 | ∃ 0 n ∈ 0 n i−1 i ≤ ∈{ }} We claim that this metric is coarsely equivalent to d for all sufficiently large C. Since d(x, y) 1 ≤ C dC (x, y) it suffices to show that r> 0 t> 0 : d(x, y) r = d (x, y) t. ∀ ∃ ≤ ⇒ C ≤ For this we choose a finite symmetric set F with Λ2 ΛF and given r> 0 we abbreviate ⊂ s := max d(e,f) f F and L(r) := max d (e, z) z Λ B(e, r + s) . { | ∈ } { C0 | ∈ ∩ } We choose C so large that C C +2s. Now let r 0 and x, y Λ with d(x, y) r. Since x−1y Λ2, ≥ 0 ≥ ∈ ≤ ∈ there exist z Λ and f F such that x−1y = zf. By definition, b b ∈ ∈ −1 d(e, z) d(e, x y)+ d(zf, z) r + s = dC0 (e, z) L(r). b ≤ b ≤ ⇒ ≤ There thus exists an n L(r) and elements x0,..., xn Λ such that ≤ b b b b b ∈ b x = e, x = z and d(x , x ) C , for all i 1,...,n . 0 n i−1 i ≤ 0 ∈{ } 2 b b Since Λ is s-relatively dense in Λ , we can find elements x1,...,xn Λ such that b b b b b ∈ x = x and d(x , xx ) s, for all i 1,...,n . 0 i i ≤ ∈{ } We then have b b d(x , x ) d(x , xx )+ d((xx , xx )+ d(xx , x ) s + C + s C, for all i 1,...,n . i−1 i ≤ i−1 i−1 i−1 i i i ≤ 0 ≤ ∈{ } If we set xn+1 := y, then we also have b b b b b b b d(x , x )= d(x ,y) d(x , xx )+ d(xx ,y) s + d(yf,y) 2s C. n n+1 n ≤ n n n ≤ ≤ ≤ This shows that dC (x, y) n +1 L(r)+1. Also note that d proper implies d is proper, and therefore ≤ ≤ b c b c b so is dC , which finishes the proof.  b A basic result in geometric group theory, known as “Gromov’s trivial lemma” (see Corollary A.30), ensures that two large-scale geodesic metric spaces are quasi-isometric if and only if they are coarsely equivalent. In particular, if (Λ, Λ∞) is geometrically finitely-generated, then by Theorem 3.29 the set [Λ] := X [Λ] X is large-scale geodesic (3.1) int { ∈ c | } defines a QI-type which refines the coarse class [Λ]c.

DEFINITION 3.30. If (Λ, Λ∞) is a geometrically finitely-generated approximate group, then the QI ∞ type [Λ]int given by (3.1) is called the internal QI type of (Λ, Λ ).

This QI type is indeed “internal” in the sense that if depends only on [Λ]c:

LEMMA 3.31 (Properties of the internal QI type). Let (Λ, Λ∞), (Ξ, Ξ∞) be countable approximate groups with [Λ]c = [Ξ]c. Then Λ is geometrically finitely-generated if and only if Ξ is, and in this case [Λ]int = [Ξ]int. Moreover, the following hold:

(i) [Λ]int is invariant under 2-local quasi-isomorphisms (in particular, under 2-local isomorphisms and Frei- man 3-isomorphisms). (ii) If (Λ, Λ∞) is geometrically finitely-generated, then for all k N the approximate group (Λk, Λ∞) is k ∈ geometrically finitely generated and [Λ]int = [Λ ]int. (iii) If ρ : (Ξ, Ξk) (Λ, Λk) is a k-local morphism for some k 10 and ker (ρ ) is geometrically finitely- k → ≥ 2 k generated, then all partial kernels up to ker k+2 (ρk) are geometrically finitely-generated and ⌊ 6 ⌋

[ker2(ρk)]int = [ker3(ρk)]int = = [ker⌊ k+2 ⌋(ρk)]int. ··· 6 PROOF. The first statement is immediate from (3.1). Then (i) follows from Lemma A.7, (ii) follows from Lemma 3.4, (iii) follows from Example 3.9.  3. EXTERNAL AND INTERNAL QI TYPE 37

In the situation of Lemma 3.31.(iii), we denote [ker(ρk)]int := [ker2(ρk)]int and refer to it as the internal QI kernel of ρk; we apply the same terminology to global morphisms.

REMARK 3.32 (Representing external and internal QI types). We have seen that the external QI ∞ type [Λ]ext of an algebraically finitely-generated approximate group (Λ, Λ ) can be represented by an external metric on Λ. Similarly, it follows from Theorem 3.29 that the internal QI type [Λ]int of a geometrically finitely-generated approximate group (Λ, Λ∞) can always be represented by a proper metric d on Λ. We refer to any such metric as an internal metric on Λ. The proof of Theorem 3.29 provides an explicit construction of such an internal metric on Λ as a “large-scale path metric”. If d is an internal metric on Λ, then (Λ, d) is proper and large-scale geodesic, and by Lemma A.26 every proper large-scale geodesic metric space is quasi-isometric to a locally finite graph. In particular, ∞ if (Λ, Λ ) is a geometrically finitely generated approximate group, then [Λ]int can be represented by a ∞ locally finite graph XΛ. We refer to any such graph as a generalized Cayley graph for (Λ, Λ ). In the study of geometrically finitely-generated approximate groups, it will often be useful to choose a convenient representative of its internal QI type:

DEFINITION 3.33. If (Λ, Λ∞) is a geometrically finitely generated approximate group, then a proper geodesic metric space (X, d) [Λ] is called an apogee for (Λ, Λ∞). ∈ int Every geometrically finitely-generated approximate group admits an apogee; for example we can always choose a generalized Cayley graph. It is an interesting questions which metric spaces can represent geometrically finitely-generated approximate groups. To discuss this, we introduce the fol- lowing notion:

DEFINITION 3.34. Let (X, d) be a pseudo-metric space, let K 1, C 0 and r> 0. We say that X ≥ ≥ is (K,C,r)-quasi cobounded if for all x, y X there exists a (K,C,C)-quasi-isometry f : X X such ∈ → that d(f(x),y) < r. It is called quasi-cobounded if it is (K,C,r)-quasi cobounded for some (K,C,r) and cobounded if it is (1, 0, 0)-quasi-cobounded.

REMARK 3.35. There are various equivalent formulations of quasi-coboundedness: (i) To see that a given space (X, d) is quasi-cobounded for some constants (K′, C′, r′), it clearly suffices to show that there exist K 1, C 0, r > 0 and o X with the following property: for ≥ ≥ ∈ all x X there exists a (K,C,C)-quasi-isometry f : X X such that d(f(x), o) < r. Moreover, ∈ → in this case the constants (K′, C′, r′) depend only on the constants (K,C,r). (ii) Assume that (X, d) is (K,C,r)-quasi-cobounded and let x, y X. Pick a quasi-isometry f : X ∈ → X such that d(f(x),y) < r and define f(z), if z = x f : X X, f(z) := 6 → y, if z = x.  This f is at distance at moste r from f (hencee a (K, C +3r, 3r)-quasi-isometry) with f(x) = y. It follows that a space is quasi-cobounded if and only if there exist constants K′ 1 and C′ 0 ≥ ≥ such thate for all x, y X there exists a (K′, C′, C′)-quasi-isometry with f(x)= y. e ∈ It follows from (ii) that a quasi-cobounded space in our sense is what is called a coarsely quasi-symmetric space in [ALC18]. Note that, by definition, quasi-coboundedness is a QI-invariant. We thus say that a QI class is quasi-cobounded if one (hence any) of its representatives is quasi-cobouned.

PROPOSITION 3.36 (Quasi-coboundedness of the internal QI class). If (Λ, Λ∞) is a geometrically ∞ finitely-generated approximate group, then [Λ]int is quasi-cobounded. In particular, every apogee of (Λ, Λ ) is a proper, geodesic, quasi-cobounded space.

∞ PROOF. Let d∞ be a left-admissible metric on Λ . Since Λ is an approximate group, it is of finite index in Λ2, and hence both Λ and Λ2 are geometrically finitely-generated. Let d and d′ be internal 38 3. LARGE-SCALE GEOMETRY OF APPROXIMATE GROUPS metrics on Λ and Λ2 respectively. The inclusion ι :Λ Λ2 is a coarse equivalence with respect to → ′ the restrictions of d∞, and hence also with respect to d and d . By Corollary A.30 it is thus a quasi- isometry and we denote by p : (Λ2, d′) (Λ, d) a quasi-inverse of ι. Now for every λ Λ the map → ∈ L(λ) : (Λ, d) (Λ, d) given by L(λ)(x) := p(λx) is a quasi-isometry with L(λ)(e) = λ, and these → quasi-isometries have uniform QI constants. The proposition follows.  Most of the results that we will establish about apogees of geometrically finitely-generated ap- proximate groups will actually be true for general proper geodesic quasi-cobounded spaces.

4. Distortion In the previous section we have defined two notions of being finitely generated for countable approximate groups, which are equivalent when restricted to countable groups by Corollary B.21. The following example shows that they are not equivalent in general:

EXAMPLE 3.37 (Algebraically finitely-generated does not imply geometrically finitely-generated). Let Γ := F be the free group on two generators a and b and let ρ := ϕ : Γ Z be the ab- 2 ab → counting quasimorphism (see Example B.47). This is a surjective symmetric quasimorphism with D(ρ)= 1, 0, 1 (see Remark B.51). Thus its quasi-kernel {− } Λ := qker(ϕ )= w F ρ(ab) 1 ab { ∈ 2 | | |≤ } is an approximate subgroup of F , and since a,b Λ we have Λ∞ = F , hence (Λ, Λ∞) is algebraically 2 ∈ 2 finitely-generated. It is easy to see that Λ considered as a subset of the Cayley tree of F2 is not coarsely connected. In view of Theorem 3.29 this shows that (Λ, Λ∞) is not geometrically finitely-generated. On the other hand we will see in Corollary 4.39 below that every geometrically finitely-generated approximate group is algebraically finitely-generated. However, even if (Λ, Λ∞) is both geometrically and (hence) algebraically finitely generated, then [Λ]int need not coincide.

DEFINITION 3.38. An approximate group (Λ, Λ∞) is called undistorted if it is both geometrically and algebraically finitely-generated and if [Λ]int = [Λ]ext. It is called distorted if it is both geometrically and algebraically finitely-generated, but [Λ] = [Λ] . int 6 ext By definition, a geometrically and algebraically finitely-generated approximate group (Λ, Λ∞) is undistorted if and only if every external metric is also an internal metric. Note that this condition (which amounts to equality [Λ]int = [Λ]ext of subsets of [Λ]c) is a priori stronger than just the existence of a quasi-isometry between a representative of [Λ]int and a representative of [Λ]ext.

EXAMPLE 3.39 (A distorted approximate group). The following example of a distorted approxi- mate group goes back to [BH18a, Example 3.26], due to Emily Stark. Consider the Baumslag-Solitar group Λ∞ := BS(1, 2) := a,b bab−1 = a2 of type (1, 2) and let Λ := a b,b−1 . By definition, h | i h i∪{ } Λ is symmetric, contains the identity and generates Λ∞. A short calculation involving the defining relation (and using that (b−1ab)2 = a) shows that Λ2 Λ e,b,b−1,b−1a , ⊂ { } hence (Λ, Λ∞) is an approximate group. We choose the finite generating set S := a±1,b±1 and ∞ 2n n{ −n } denote by dS the associated word metric on Λ . Now let wn := a . Since wn = b ab , we have w 2n+1. In general, the distance between ak and al in the external metric d is logarithmic k nkS ≤ S|Λ×Λ in k l . On the other hand, let d be the metric on Λ such that d(ak,al) := k l and such that | − | | − | d(ak,b±1) := k +1 and d(b,b−1) := 2. Then d is large-scale geodesic and coarsely equivalent to | | d , hence an internal metric on Λ. Since d is not quasi-isometric to d , the approximate S|Λ×Λ S|Λ×Λ group (Λ, Λ∞) is distorted. We will later see that distortion is an obstruction to the existence of geometric isometric actions of approximate groups. If an approximate group (Λ, Λ∞) is distorted, then its internal QI type is often the more useful model in geometric applications. For example, we have seen in Lemma 3.31 that [Λ]int 5. GROWTH OF APPROXIMATE GROUPS 39 is invariant under 2-local (quasi-)isomorphisms (and hence Freiman 3-isomorphisms) and depends only on [Λ]c. The latter is not the case for the external QI type: ∞ Z Z REMARK 3.40. If (Λ, Λ ) is as in Example 3.39, then [Λ]c ∼= [ ]c, and hence [Λ]int ∼= [ ]int by Lemma 3.31. On the other hand, [Λ] = [Z] . This shows that Lemma 3.31 has no counterpart for ext 6∼ ext external QI classes: The external QI type can even change when one passes to a finite index (approxi- mate) subgroup. We now give a general criterion to ensure that a given approximate group is undistorted; this criterion reformulates a result from [BH18a].

DEFINITION 3.41. Let (X, d) be a proper metric space. A subset A X is called weakly (R,K,C)- ⊂ quasi-convex in X provided that for all x, y A there is a (K, C)-quasi-geodesic segment between x ∈ and y contained in NR(A). By definition, weak quasi-convexity is invariant under quasi-isometries of pairs, hence we can define:

DEFINITION 3.42. An approximate group (Λ, Λ∞) is called weakly quasi-convex if it is algebraically finitely-generated and Λ is a weakly quasi-convex subset of Λ∞ with respect to some (hence any) word metric on Λ∞. Note that (Λ, Λ∞) is weakly quasi-convex in Λ∞ if and only if there exists k = k(S) N such that ∈ for all x, y Λ there exists a geodesic from x to y in the Cayley graph of Λ∞ with respect to S whose ∈ vertices are contained in Λk. We then call k a convexity parameter of (Λ, Λ∞) with respect to S. We can now state the following criterion, which is slightly adapted from [BH18a].

PROPOSITION 3.43. If (Λ, Λ∞) is a weakly quasi-convex approximate group, then it is geometrically finitely-generated and undistorted.

PROOF. It suffices to show that (Λ, dS Λ×Λ) is quasi-isometric to a connected graph. For this let ∞ | k be a convexity parameter for (Λ, Λ ) with respect to S and denote by Γk the full subgraph of the Cayley graph of Λ∞ with respect to S on the vertex set Λk, and by Γ Γ the connected component ⊂ k of Λ. Then the assumption implies that the inclusion (Λ, d ) Γ is a quasi-isometry, and Γ is a S|Λ×Λ → connected graph, hence geodesic. 

Note that it was established in [BH18a, Prop. 3.27] that every uniform model set is weakly-quasi convex. Thus uniform model sets provide examples of undistorted approximate groups. We will later extend this result to arbitrary uniform approximate lattices.

REMARK 3.44 (Concerning terminology). In the paper [BH18a], which introduced geometric ap- proximate group theory, the authors referred to algebraically finitely-generated approximate groups simply as finitely-generated approximate groups, and to the external QI type [Λ]ext as the canonical QI type. They then wrote [Λ] instead of [Λ]ext. However, the focus of the relevant parts of [BH18a] was mostly on uniform approximate lattices, and we will see that uniform approximate lattices are always undistorted. Since the writing of [BH18a] it has become increasingly clear that there are many interesting ex- amples of distorted approximate groups. To deal with these examples, it is necessary to distinguish carefully between internal and external QI types.

5. Growth of approximate groups The purpose of this section is to define growth of approximate groups in analogy with growth of groups. We first recall a very general notion of growth for metric spaces of coarse bounded geometry in the sense of Block and Weinberger [BW92]. The following definition is taken from [ALC18]; for the terminology see Definition A.32 and Remark A.34. 40 3. LARGE-SCALE GEOMETRY OF APPROXIMATE GROUPS

DEFINITION 3.45. Let (X, d) be a pseudo-metric space of coarse bounded geometry with basepoint o X and let Λ X be a quasi-lattice. ∈ ⊂ (i) The growth function of Λ is defined as γ (r) := Λ B(o, r) . Λ,o | ∩ | (ii) The growth of X, denoted gr(X), is the growth equivalence class of the function γΛ,o.

EXAMPLE 3.46. We say that a pseudo-metric space (X, d) of coarse bounded geometry has exponential growth if 0 < lim inf log γ(r) < for some (hence any) γ gr(X); • r ∞ ∈ subexponential growth if sup log γ(r) 0 for some (hence any) γ gr(X); • r ≤ ∈ polynomial growth if sup log γ(r) < for some (hence any) γ gr(X). • log r ∞ ∈ Note in particular that if X has polynomial growth, then there exists d N such that for every γ ∈ ∈ gr(X) there exists C > 0 such that for all r R, ∈ γ(r) Crd. ≤ By Lemma A.35 the growth of X is independent of the choice of o and Λ. Moreover we have:

PROPOSITION 3.47 (QI invariance of growth). If X is of coarse bounded geometry and Y is quasi- isometric to X, then Y is of coarse bounded geometry and gr(X) = gr(Y ).

PROOF. Let f : X Y be a (K,C,C′)-quasi-isometry and let Λ X be a quasi-lattice so that → 1 ⊂ Λ := f(Λ ) is a quasi-lattice in Y by Lemma A.33. Let x X be a basepoint and y := f(x ) Y . 2 1 0 ∈ 0 0 ∈ Assume that y Λ B(y , r) and let x Λ such that y = f(x). We then have ∈ 2 ∩ 0 ∈ 1 d(f(x),f(x )) r = d(x, x ) Kr + KC = Λ B(y , r) f(Λ B(x , Kr + Kc)). 0 ≤ ⇒ 0 ≤ ⇒ 2 ∩ 0 ⊂ 1 ∩ 0 This shows that γ γ , and hence the proposition follows by symmetry.  Λ2,y0  Λ1,x0 In view of this proposition we can thus talk unambiguously about a QI class of coarse bounded geometry (meaning that one, hence any of its representatives has coarse bounded geometry), and of its growth. In order to apply this to approximate groups we need the following observation:

LEMMA 3.48. Let (Λ, Λ∞) be a countable approximate group and let d be the restriction of a left-admissible metric d on Λ∞ to Λ. Then (Λ, d) is of coarse bounded geometry and Λ is a quasi-lattice in (Λ, d).

∞ PROOFb . Given z Λ we denote by B(z, r) the closed ball of radius r around z with respect to d ∞ ∈ 3 in Λ . Fix y Λ and set Φ+(r) := Λ B(y, r) . Then for every x Λ we have ∈ | ∩ | ∈ b Λ B(x, r) = yx−1(Λ B(x, r)) Λ3 B(y, r) =Φ (r). | ∩ | | ∩ | ≤ | ∩ | + We deduce that Λ is a quasi-lattice in itself with respect to d, and hence (Λ, d) is of coarse bounded geometry.  Since quasi-lattices are preserved under coarse equivalences (see Lemma A.33) we deduce:

COROLLARY 3.49. Let (Λ, Λ∞) be a countable approximate group. (i) If (Λ, Λ∞) is algebraically finitely-generated, then every X [Λ] is of coarse bounded geometry and if ∈ ext d is any external metric on Λ, then gr(X) is represented by the function γ(r) := x Λ d(x, e) r . |{ ∈ | ≤ }| (ii) If (Λ, Λ∞) is geometrically finitely-generated, then every X [Λ] is of coarse bounded geometry and if ∈ int d is any internal metric on Λ, then gr(X) is represented by the function γ(r) := x Λ d(x, e) r .  |{ ∈ | ≤ }| DEFINITION 3.50. Let (Λ, Λ∞) be a countable approximate group. If (Λ, Λ∞) is algebraically (re- spectively geometrically) finitely generated and X [Λ]ext (respectively X [Λ]int), then gr(X) is ∈ ∞ ∈ called the external growth (respectively internal growth) of (Λ, Λ ) and denoted grext(Λ) (respectively grint(Λ)). 5. GROWTH OF APPROXIMATE GROUPS 41

∞ If (Λ, Λ ) is undistorted, then we simply write gr(Λ) := grint(Λ) = grext(Λ) and refer to gr(Λ) as the growth of Λ. If Λ is a countable discrete group then we recover the usual notion of growth of a group.

PROPOSITION 3.51 (Growth inequalities). For every algebraically and geometrically finitely-generated approximate group (Λ, Λ∞) we have gr (Λ) gr (Λ) gr(Λ∞). int  ext  PROOF. Let d be an external metric on Λ. As in the proof of Theorem 3.29 we can then find an internal metric of the form d (x, y) := inf n x ,...,x Λ: d(x , x ) < C for all i 1,...,n , and x = x, x = y . C { | ∃ 0 n ∈ i−1 i ∈{ } 0 n } For all x, y Λ we then have d (x, y) C d(x, y), and hence ∈ C ≥ · x Λ d (x, e) r x Λ d(x, e) Cr . |{ ∈ | C ≤ }| ≤ |{ ∈ | ≤ }| This implies the first inequality and the second inequality is obvious. 

COROLLARY 3.52. In the situation of Proposition 3.51 we have the following implications: (i) Λ∞ has pol. growth = Λ has external pol. growth = Λ has internal pol. growth. ⇒ ⇒ (ii) Λ has internal exp. growth = Λ has external exp. growth = Λ∞ has exp. growth.  ⇒ ⇒ Both inequalities in Proposition 3.51 can be strict. For the first inequality this follows from the following example:

EXAMPLE 3.53. If (Λ, Λ∞) is as in Example 3.39, then (Λ, Λ∞) has linear internal growth in the sense that grint(Λ) = grint(Z) is represented by γint(r) := r. On the contrary, grext(Λ) is represented by r ∞ γext =2 and hence (Λ, Λ ) has exponential external growth. To see that the second inequality can also be strict, we consider the following example:

EXAMPLE 3.54. Let Λ∞ := Z2 and Λ := (a,b) Z2 b 1 . Then gr (Λ) is linear, whereas { ∈ | | | ≤ } ext gr(Λ∞) is quadratic.

CHAPTER 4

Geometric quasi-actions of approximate groups

The main idea of geometric group theory is to study finitely-generated groups via their isomet- ric actions on metric spaces. A particular role is played by actions which are geometric, since these provide models for the canonical QI class of the group by the Milnor–Schwarz lemma. While every algebraically finitely-generated approximate group admits an isometric action on a proper geodesic metric space, this action can usually not be chosen to be geometric – distortion turns out to be an important obstruction. This issue can be resolved by relaxing the conditions of an isometric action and considering quasi-isometric quasi-actions (qiqacs) instead. One can then show that every (alge- braically or geometrically) finitely-generated approximate group admits a geometric qiqac on a proper geodesic metric space, and thus these quasi-actions should be seen as the main protagonists of geomet- ric approximate group theory. Correspondingly, we will formulate our version of the Milnor–Schwarz lemma in the context of these quasi-actions.

1. Isometric actions of approximate groups Throughout this section, (Λ, Λ∞) denotes an approximate group. We have the obvious notion of a (global) isometric action of (Λ, Λ∞):

DEFINITION 4.1. Let (X, d) be a pseudo-metric space and o X. ∈ (i) A morphism ρ : (Λ, Λ∞) Is(X) is called an isometric action of (Λ, Λ∞) on X. → (ii) If ρ : (Λ, Λ∞) Is(X) is an isometric action, then the subset → ρ(Λ)(o) := ρ(λ)(o) λ Λ { | ∈ } is called the Λ-orbit of o under ρ. Despite the name, Λ-orbits under an isometric action ρ need not be invariant under ρ(Λ).

DEFINITION 4.2. Let X be a metric space, be a set of self-maps of X and Y X be a subset. A ⊂ Then Y is called quasi-invariant under if there exists D 0 such that for every f in we have A ≥ A d (Y,f(Y )) D, where d denotes Hausdorff distance. Haus ≤ Haus PROPOSITION 4.3. Let ρ : (Λ, Λ∞) Is(X) be an isometric action on a pseudo-metric space (X, d). → (i) Every Λ-orbit is quasi-invariant under ρ(Λ). (ii) Any two Λ-orbits are at bounded distance.

PROOF. (i)Let o X and let := ρ(Λ)(o). Choose a finite set F Λ∞ such that Λ2 F Λ and set ∈ O ⊂ ⊂ R := sup d(o, f.o) f F ; then for every λ Λ we have d ( ,ρ(λ)( )) R. { | ∈ } ∈ Haus O O ≤ (ii) This follows from the fact that d(ρ(λ)(x),ρ(λ)(y)) = d(x, y) for all λ Λ and x, y X.  ∈ ∈ Every infinite countable approximate group admits an isometric action with infinite orbits, which is, however, usually not cobounded:

EXAMPLE 4.4. If (Λ, Λ∞) is an arbitrary countable approximate group and d is a left-admissible metric on Λ∞, then (Λ, Λ∞) acts isometrically on (Λ∞, d) via ρ(λ)(x) := λx. Thus every countable approximate group admits an isometric action. If (Λ, Λ∞) is algebraically finitely-generated, then it even admits an isometric actions on a locally finite graph (e.g. any Cayley graph of Λ∞). However, these actions are not very much of interest to us since the Λ-orbits are not cobounded (in a sense made precise below) unless Λ is of finite index in Λ∞ and thus essentially a group.

43 44 4. GEOMETRIC QUASI-ACTIONS OF APPROXIMATE GROUPS

In order to define a more interesting class of isometric actions, we mimic the group case. We recall that an action of a discrete group on a proper metric space is geometric if and only if it is cobounded and proper (see the discussion in Remark B.23 concerning various equivalent definitions). The notions of coboundedness and properness can be defined in very large generality:

DEFINITION 4.5. Let (X, d) be a proper pseudo-metric space, be a set and ρ : Map(X,X) A A → be a map. (i) ρ is called cobounded if for some x X the set ρ( ).x := ρ(a)(x) a is relatively dense in ∈ A { | ∈ A} X. (ii) ρ is called proper if the map X X X, (a, x) (x, ρ(a)(x)) is proper, where carries the A× → × 7→ A discrete topology. In particular we define:

DEFINITION 4.6. Let (X, d) be a proper pseudo-metric space. An isometric action ρ : (Λ, Λ∞) → Is(X) is called (1) cobounded if ρ :Λ Is(X) Map(X,X) is cobounded in the sense of Definition 4.5; 1 → ⊂ (2) proper if ρ :Λ Is(X) Map(X,X) is proper in the sense of Definition 4.5; 1 → ⊂ (3) geometric if it is cobounded and proper. The goal of the remainder of this section is to characterize those approximate groups which admit geometric isometric actions on proper metric spaces. The final result will be as follows (see Definition 2.52 and the discussion thereafter for the notion of a uniform approximate lattice):

THEOREM 4.7 (Approximate groups with geometric isometric actions). The following are equivalent: (i) (Λ, Λ∞) admits a geometric isometric action on a large-scale geodesic proper metric space. (ii) (Λ, Λ∞) is a finite extension of a uniform approximate lattice in a compactly-generated lcsc group. In this case, (Λ, Λ∞) is algebraically and geometrically finitely-generated and undistorted. We now turn to the proof of the equivalence (i) (ii) of Theorem 4.7; the final statement will ⇐⇒ be proved in Section 5. The implication (ii) (i) in Theorem 4.7 comes from the following simple ⇒ observation:

PROPOSITION 4.8. Let Λ be a uniform approximate lattice in a lcsc group G, let Λ∞ < G be its enveloping group and let d be a left-admissible metric on G. Denote by λ : G Is(G, d) the isometric action of G on G → itself by left-multiplication. Then the isometric action of (Λ, Λ∞) on (G, d) given by λ (Λ, Λ∞) ֒ G G Is(G, d) → −−→ is geometric.

PROOF. Since Λ=Λ.e is relatively dense in G, the action is cobounded. Properness follows from properness of the left-regular action of G on itself together with discreteness of Λ. 

PROOF OF THEOREM 4.7. (ii) (i): Assume that π : (Λ, Λ∞) (Ξ, Ξ∞) is a finite extension ⇒ → and that (Ξ, Ξ∞) is a uniform approximate lattice in a compactly-generated lcsc group G. Since G is compactly generated, we can find a left-admissible metric d on G such that (G, d) is (proper and) large- scale geodesic (Proposition B.16). We fix such a metric once and for all. We consider the composition λ .(ρ : (Λ, Λ∞) π (Ξ, Ξ∞) ֒ G G Is(G, d −→ → −−→ Since π is onto, we have ρ(Λ).e = Ξ, hence ρ is cobounded. As for properness, consider A := λ Λ ρ(λ)B (e) B (e) = . R { ∈ | R ∩ R 6 ∅} From the properness part of Proposition 4.8 we know that F := π(A ) Λ is finite. Now assume R R ⊂ that λ , λ A such that π(λ ) = π(λ ). Then π(λ−1λ ) = e and hence λ−1λ ker(π :Λ2 Ξ2), 1 2 ∈ R 1 2 1 2 1 2 ∈ 2 → which is finite by assumption. This shows that AR fibers over the finite set FR with finite fibers, hence it is finite as well, showing properness of the action ρ. This finishes the proof of Thm. 4.7 (ii) (i).  ⇒ 1. ISOMETRIC ACTIONS OF APPROXIMATE GROUPS 45

We now turn to the converse implication. Note that if an approximate group (Λ, Λ∞) acts geomet- rically by isometries on a proper metric space X, then in particular Λ∞ and hence Is(X) acts cobound- edly on X. For this reason we now restrict attention to metric spaces which admit a cobounded action of the isometry group. We will call such metric spaces cobounded for short. We recall from Proposition B.20 that the isometry group of a large-scale geodesic cobounded proper metric space is a compactly- generated lcsc group.

PROPOSITION 4.9. Let X be a large-scale geodesic cobounded proper metric space and let Λ be an approx- imate subgroup of G := Is(X). Then the following are equivalent: (i) The isometric action of (Λ, Λ∞) on X is geometric. (ii) Λ is relatively dense in G and the map p :Λ G G G, (λ, g) (g,λg) is proper. × → × 7→ PROOF. Since G is a compactly-generated lcsc group it admits a large-scale geodesic left-admis- sible metric d (Proposition B.16). Fix a basepoint o X and let ι = ι : (G, d ) (X, d) be the G ∈ o G → corresponding orbit map. By Proposition B.20 the map ι is a quasi-isometry. Denote by µ :Λ G G, (λ, g) λg and ρ :Λ X X, (λ, x) λ(x), × → 7→ × → 7→ the natural actions of Λ on G and X respectively. Then we have a commuting diagram µ Λ G / G × id×ι ι   Λ X / X. × ρ Since ι is a quasi-isometry, it follows that Λ is relatively dense in G if and only if Λ.o = ι(Λ) is relatively dense in X, which means that the action is cobounded. Similarly, the map p :Λ G G G, × → × (λ, g) (x, λg) is proper if and only if the map p :Λ X X X, (λ, x) (x, λ.x) is proper. This 7→ X × → × 7→ establishes the desired equivalence.  We can reformulate condition (ii) of Proposition 4.9 using the following proposition:

PROPOSITION 4.10. Let G be a compactly-generated lcsc group and Λ G be an approximate subgroup. ⊂ Then the following are equivalent: (i) Λ is relatively dense in G and the map p :Λ G G G, (λ, g) (g,λg) is proper. × → × 7→ (ii) Λ is an approximate lattice in G.

PROOF. (ii) (i): If Λ is an approximate lattice in G, then Λ is relatively dense in G. Moreover, ⇒ it is uniformly discrete and hence locally finite with respect to any left-admissible metric on G. In particular, if we fix such a metric, then for every R> 0 the intersection B (e) Λ is finite. Since 2R ∩ λ Λ B (e) λB (e) B (e) Λ, { ∈ | R ∩ R }⊂ 2R ∩ this implies that the map p :Λ G G G is proper. × → × (i) (ii) Since Λ is relatively dense by assumption, we only need to show that Λ is uniformly ⇒ discrete. Assume for contradiction that this is not the case. Then there exist elements g ,h G with n n ∈ g = h such that x := g−1h converges to e. Note that x Λ2 e and choose F Λ3 finite such n 6 n n n n n ∈ \{ } Λ ⊂ that Λ2 =ΛF . Passing to a subsequence we may assume that x Λ f for some fixed f F . This Λ n ∈ { } ∈ Λ means that there exist λ Λ f −1 such that λ f −1, i.e., g := f −1 G is an accumulation point n ∈ \{ } n → ∈ of Λ. Fix a left-admissible metric d on G and let R := d(g,g2) and R := 2R . Now for every ǫ (0, R ] 0 0 ∈ 0 there exist infinitely many n N with λ B (g). For such n we then have d(gλ ,g2)= d(λ ,g) <ǫ. ∈ n ∈ ǫ n n In particular, gλ B (g)λ and gλ B (g2) B (g) B (g), n ∈ R n n ∈ ǫ ⊂ R0+ǫ ⊂ R hence B (g)λ B (g) = . This shows that the map (λ, g) (g,gλ) is not proper. Applying R n ∩ R 6 ∅ 7→ inverses to both sides and using that G and Λ are symmetric it follows that p is not proper, which is a contradiction.  46 4. GEOMETRIC QUASI-ACTIONS OF APPROXIMATE GROUPS

PROOF OF THEOREM 4.7. (i) = (ii): Assume that (Λ, Λ∞) admits a geometric isometric action ⇒ on a large-scale geodesic proper metric space (X, d), say ρ : (Λ, Λ∞) Is(X). Recall from Proposition → B.16 that G := Is(X) is a compactly-generated lcsc group and denote by (Ξ, Ξ∞) the image of ρ. Then by the implication (i) (ii) of Proposition 4.9 and the implication (i) (ii) of Proposition 4.10 we ⇒ ⇒ deduce that Ξ is an approximate lattice in G, and thus it remains to show only that the extension (Λ, Λ∞) (Ξ, Ξ∞) is finite. We first observe that since ρ : (Λ, Λ∞) Is(X) is a proper isometric → → action, then by Lemma 4.25 the induced isometric action ρ : (Λk, Λ∞) Is(X) is also proper. Now if k → k N and λ ker (ρ), then for any basepoint o X and every R> 0 we have ∈ ∈ k ∈ λB (o) B (o) = . R ∩ R 6 ∅

From properness of ρk we thus deduce that kerk(ρ) is finite, hence ρ has finite kernels. This finishes the proof of the implication (i) (ii).  ⇒ At this point we have established the equivalence (i) (ii) in Theorem 4.7. The proof of the ⇐⇒ final statement is based on the Milnor–Schwarz lemma (Theorem 4.37) and will be given in Section 5 below. Note that an approximate lattice in a discrete group is just an almost group in the sense of Example 2.10; we may thus conclude:

COROLLARY 4.11. Let (Λ, Λ∞) be an approximate group which is not a finite extension of an almost group. Then (Λ, Λ∞) does not admit any geometric isometric action on a large-scale geodesic proper metric space with discrete isometry group. 

2. Quasi-isometric quasi-actions (qiqacs) Since not all approximate groups are finite extensions of uniform approximate lattices in a com- pactly generated lcsc group (for example, distorted approximate groups are not), it follows from The- orem 4.7 that not every approximate group admits a geometric isometric action. In order to develop a far-reaching geometric theory of approximate groups one thus has to relax the definition of an iso- metric action.

DEFINITION 4.12. Let (Λ, Λ∞) be an approximate group, (X, d) be a pseudo-metric space and o X. Given k N, let K 1 and C , C′ 0 be constants. ∈ ∈ k ≥ k k ≥ (i) A map of pairs ρ : (Λ, Λ∞) (QI(X), QI(X)) is called a (K , C , C′ )-quasi-isometric quasi- → k k k action (or (K , C , C′ )-qiqac for short) of (Λ, Λ∞) on X if for every k 1 we have ρ(Λk) k k k ≥ ⊂ QI ′ (X) and f f Kk ,Ck,Ck

′ k f d(ρ(λ1)ρ(λ2).x, ρ(λ1λ2).x) < C (x X, λ1, λ2 Λ ). (2.1) k ∈ ∈ ′ (ii) If there exists K > 0 such that Kk K for all k N, then a (Kk, Ck, Ck)-qiqac is called a ′ ′≤ ∈ ′ (K, Ck, Ck)-uniform qiqac.A (1, Ck, Ck)-uniform qiqac is called a (Ck, Ck)-rough quasi-action (or roqac for short). (iii) A map of pairs ρ : (Λ, Λ∞) (QI(X), QI(X)) is called a weak qiqac if it factor through a morphism → ρ : (Λ, Λ∞) (QI(X), QI(X)) (called the associated morphism). → (iv) If ρ : (Λ, Λ∞) (QI(X), QI(X))fis a (Kf , C , C′ )-qiqac, then the set → k k k ρ(Λ).o := ρ(λ).o λ Λ f f { | ∈ } is called the Λ-quasi-orbit of o under ρ.

In the sequel, when given a group G and an approximate group (Λ, Λ∞), we will often denote a map of pairs (Λ, Λ∞) (G, G) simply by (Λ, Λ∞) G. This is in accordance with our previous abuse → → of notation for morphisms (see comments after Def. 2.18) and applies in particular to qiqacs. 2. QUASI-ISOMETRIC QUASI-ACTIONS (QIQACS) 47

REMARK 4.13. (i) By definition, every isometric action is a uniform qiqac and every uniform qiqac is a qiqac. Moreover, a map ρ : (Λ, Λ∞) QI(X) is a weak qiqac if and only if for all → λ , λ Λ, 1 2 ∈ Cλ1,λ2 := sup d(ρ(λ1)ρ(λ2).x, ρf(λ1λ2).x) < . x∈X ∞ In particular, every qiqac is a weak qiqac, but the definition is more restrictive, in that it demands k the Cλ1,λ2 to be bounded uniformly as λ1 and λ2 vary inside some fixed Λ , and requires that the QI-constants of ρ(λ) are bounded uniformly as λ varies over some Λk. (ii) Λ-quasi-orbits are quasi-invariant under ρ(Λ) and Λ-quasi-orbits through different base-points for a given qiqac are at mutually bounded distance (by a quasified version of the proof of Pro- postion 4.3). (iii) It is immediate from the above definition that every qiqac ρ : (Λ, Λ∞) (QI(X), QI(X)) restricts → to a qiqac ρ : (Λk, Λ∞) (QI(X), QI(X)), called the induced qiqac . → (iv) The conditions in the definition of a qiqac can be weakened. It sufficesf to assumef that ρ(Λ) is uniform and that (2.1) holdsf for allfλ Λ and λ Λk. Using the latter assumption one can 1 ∈ 2 ∈ then prove by induction that (2.1) holds for all λ , λ Λk (for a larger uniform constant), and 1 2 ∈ conclude that ρ(Λk) is uniformly close to (ρ(Λ))k, hence uniform if ρ(Λ) is. We chose the current formulation in order to make the existence of the induced qiqac more obvious. (v) Our definition of a qiqac is of course not the only possible relaxation of the definition of an isometric action. As far as the involved QI constants are concerned, one might as well ask for stronger or weaker uniformity conditions and it may seem at this point that our definition is rather arbitrary. However, in reality the wiggle room is quite limited: If one asks for more unifor- mity then the left-regular quasi-action defined in Section 4 below will no longer be a quasi-action, and if one relaxes the uniformity too much, then one can no longer establish basic results such as the Milnor–Schwarz lemma. (vi) The definition of a qiqac does not require any relation between ρ(g) and ρ(g)−1 a priori. However, such a relation does in fact always hold by the following lemma.

LEMMA 4.14. If ρ : (Λ, Λ∞) QI(X) is a qiqac, then there exist constants D such that for all x X → k ∈ and g Λk, ∈ −1 −1 d(ρ(g) x,f ρ(g )x) d(ρ(e2)x, ρ(e)ρ(e)x)= d(ρ(e)x, ρ(e)ρ(e)x) K−1d(x, ρ(e)x) C , 1 ≥ 1 − 1 and hence d(ρ(e)x, x) < 2K C independently of x. We deduce that for g Λk we have 1 1 ∈ d(ρ(g)−1x, ρ(g−1)x) K d(x, ρ(g)ρ(g−1)x)+ C ≤ k k K (d(x, ρ(gg−1)x)+ d(ρ(gg−1)x, ρ(g)ρ(g−1)x))) + C ≤ k k K (d(x, ρ(e)x)+ C )+ C ≤ k k k K (2K C + C )+ C , ≤ k 1 1 k k which is again independent of x.  Important examples of qiqacs arise from isometric actions by a process called quasi-conjugation:

CONSTRUCTION 4.15 (Quasi-conjugation). Let ρ : (Λ, Λ∞) QI(X) be a qiqac, let ϕ : X Y be → → a quasi-isometry and let ϕ : Y X be a quasi-inverse of ϕ. Then we obtain a qiqac of (Λ, Λ∞) on Y → by f (ϕ, ϕ) ρ : (Λ, Λ∞) QI(Y ), λ ϕ ρ(λ) ϕ. ∗ → 7→ ◦ ◦ We refer to (ϕ, ϕ)∗ρ as the quasi-conjugation of ρ by (ϕ, ϕ). This construction applies in particular to the case where ρ is an isometric action to begin with. In this case, (ϕ, ϕ)∗ρ is actually a uniform qiqac. More generally, the class of uniform qiqacs is invariant under quasi-conjugation, whereas the class of roqacs is not. 48 4. GEOMETRIC QUASI-ACTIONS OF APPROXIMATE GROUPS

DEFINITION 4.16. Let (Λ, Λ∞) be an approximate group, (X, d), (Y, d′) be pseudo-metric spaces, and let ρ ,ρ : (Λ, Λ∞) QI(X) and ρ : (Λ, Λ∞) QI(Y ) be qiqacs. For every k N let K 1 and 0 1 → 2 → ∈ k ≥ C , C′ 0 be constants. k k ≥ f f k (i) We say that ρ0 and ρ1 are (Ck)-equivalent if d(ρ0(γ).x, ρ1(γ).x) < Ck for all γ Λ and x X. ′ ∈′ ∈ (ii) We say that ρ1 and ρ2 are (Kk, Ck, Ck)-quasi-conjugate if there exist (Kk, Ck, Ck)-quasi-isometries ϕ : X Y and ϕ : Y X such that ϕ ϕ and ϕ ϕ are C′ -close to the respective identities and → → ◦ ◦ k such that (ϕ, ϕ)∗ρ1 and ρ2 are (Ck)-equivalent.

REMARK 4.17. Note that if ρ0 and ρ1 are equivalent qiqacs, then the associated morphisms into QI(X) coincide. More generally, if they are quasi-conjugate qiqacs on the same space X, then the associated morphisms are conjugate by an element of QI(X), hence the name. However, being quasi-conjugate is stronger than just being conjugate inside QI(X): The defini- tions require an additional uniformity, which is needed, for example, to ensure that quasi-orbits of equivalent quasi-actions are at bounded Hausdorff distance from each other.

CONSTRUCTION 4.18 (Quasi-restrictions of qiqacs, I). Assume that ρ is a qiqac of an approximate group (Λ, Λ∞) on a pseudo-metric space (X, d) and let Y X be a subset which is quasi-invariant ⊂ under ρ(Λ∞). By definition, this means that the inclusion ( ι : Y ֒ Y := Y ρ(λ)(Y → ∪ ∞ λ∈[Λ is coarsely onto. Thus there exists a quasi-isometrye p : X X with the following properties: → p is of finite distance from the identity and p(x)= x for all x X Y . • ∈ \ p e is a quasi-inverse to ι. • |Y Then ρ := p ρ defines a qiqac which is equivalent to ρ and such that Y is invariante under ρ(Λ∞). We ◦ may thus turn every ρ(Λ∞)-quasi-invariant subset into a ρ(Λ∞)-invariant subset at the cost of replacing ρ by ane equivalent qiqac. We then obtain a qiqac e ρ : (Λ, Λ∞) QI(Y ), λ ρ(λ) , |Y → 7→ |Y whose equivalence class is uniquely determined by ρ. We will refer to any qiqac in this equivalence f class as a quasi-restriction of ρ toeY . e In Construction 4.18 we have explained how to quasi-restrict a qiqac of an approximate group (Λ, Λ∞) to a subset which is quasi-invariant under ρ(Λ∞). A variant of this construction still applies if the respective subset is merely invariant under ρ(Λ), albeit the details become more technical:

CONSTRUCTION 4.19 (Quasi-restrictions of qiqacs, II). Assume that ρ is a qiqac of an approximate group (Λ, Λ∞) on a pseudo-metric space (X, d) and let Y X be a subset which is quasi-invariant ⊂ under ρ(Λ). Set Y := Y and for n N define 1 ∈ Yn+1 := Y1 Y2 Yn ρ(λ)(Y ). ∪ ∪···∪ ∪ n λ[∈Λ If F Λ∞ is finite with Λ2 F Λ, then Λn F n−1Λ, and hence ρ(Λn)(Y ) is at bounded distance from ⊂ ⊂ ⊂ ρ(Λn−1)(Y ). If ρ is an isometric action, then this bound can in fact be chosen independently of n, but in general it will depend on n. In any case, we may deduce that the isometric inclusions ιn+1 : Y Y n n → n+1 are quasi-isometries. We choose quasi-inverses pn+1 : Y Y such that n n+1 → n pn+1ιn+1(y)= y for all y Y and d(ιn+1(pn+1(x)), x) δ for all x Y , n n ∈ n n n ≤ n ∈ n+1 where δn is a constant depending on the QI constants of ρ. Again, if ρ was actually an isometric action, then the sequence (δn) can be chosen to be bounded. In the sequel, for all k

n Then for all k

′ e ′ If ρ is a (Kk, Ck, Ck)-qiqac, then ρY (g) will be a (Kk, Ck + σk, Ck + σk)-quasi-isometry of Y = Y1. Moreover, if h Λl and y Y , then we have ∈ ∈ e k+l+1 k+1 l+1 d (ρY (gh)(y), ρY (g)ρY (h)(y)) = d p1 (ρ(gh)(y)),p1 (ρ(g)(p1 (ρ(h)(y)))) d pk+l+1(ρ(gh)(y)),ρ(gh)(y) + d ρ(gh)(y),ρ(g)(pl+1(ρ(h)(y))) ≤ 1 1  e l+1e e k+1 l+1 +d ρ(g)(p1 (ρ(h)(y)),p1 (ρ(g)(p1 (ρ(h)(y)))) ,  which is bounded by some constant depending on k and l. This shows that ρY is a qiqac. Since this construction generalizes Construction 4.19 we still refer to ρY as a quasi-restriction of ρ to Y . e Note that Construction 4.19 applies in particular to quasi-orbits. e 3. Geometric quasi-actions Imitating the definition in the case of isometric actions we make the following definition:

DEFINITION 4.20. Let (Λ, Λ∞) be an approximate group and let X be a proper metric space. A qiqac ρ : (Λ, Λ∞) QI(X) is called → (1) cobounded if ρ1 :Λ QI(X) Map(X,X) is cobounded in the sense of Definition 4.5; f → ⊂ (2) proper if ρ :Λ QI(X) Map(X,X) is proper in the sense of Definition 4.5; 1 → ⊂ (3) geometric if it is coboundedf and proper. f REMARK 4.21. (1) By definition, a qiqac ρ is cobounded if some of its quasi-orbits are rela- tively dense. It then follows from Remark 4.13 that all quasi-orbits are relatively dense. (2) Properness of ρ means that for every R> 0 and some x X the set ∈ λ Λ B (x) ρ(λ).B (x) = { ∈ | R ∩ R 6 ∅} is finite. This then holds for every x X. In fact, from Lemma 4.25 we can draw the stronger ∈ conclusion that for every k 1, R> 0 and x X the set ≥ ∈ µ Λk B (x) ρ(µ).B (x) = { ∈ | R ∩ R 6 ∅} is finite. (3) Every isometric action of an approximate group is in particular a qiqac, and hence we obtain notions of a proper, cobounded or geometric isometric action. For isometric actions of groups we recover the usual definitions.

The property of being geometric is invariant under quasi-conjugation:

PROPOSITION 4.22. If two qiqacs are quasi-conjugate, then one is proper, cobounded or geometric if and only if the other one has the corresponding property.

The proof of Proposition 4.22 has nothing to do with approximate groups. In fact, the proposition is just a special case of the following two lemmas, which are formulated in the generality of Definition 4.5. 50 4. GEOMETRIC QUASI-ACTIONS OF APPROXIMATE GROUPS

LEMMA 4.23. Let X, Y be proper metric spaces, let ϕ : X Y be a quasi-isometry and ϕ¯ : Y X be a → → quasi-inverse to ϕ. Let A be a set, ρ : A QI(X) be a map with uniform image and define ρ¯ : A QI(Y ) by → → ρ¯(a) := ϕ ρ(a) ϕ¯. Then the following hold: ◦ ◦ (i) ρ¯ is cobounded if and only if ρ is cobounded.f f (ii) ρ¯ is proper if and only if ρ is proper.

PROOF. We choose K > 1 and C > 0 such that the following hold: ϕ and all ρ(a) are (K,C,C)- quasi-isometries for all a A. Moreover, there exists a (K,C,C)-quasi-isometry ϕ¯ such that ϕϕ¯ and ∈ ϕϕ¯ are at distance at most C from the respective identities. Let x X and y := ϕ(x ). o ∈ o o (i) Assume that ρ(A).x is R-relatively dense in X. Then for every y Y we find a = a(y) A o ∈ ∈ such that d(ρ(a)(x ), ϕ¯(y)) R. Then o ≤ d(¯ρ(a)y ,y) = d(ϕρ(a)¯ϕϕ(x ),y) d(ϕρ(a)¯ϕϕ(x ), ϕρ(a)(x )) + d(ϕρ(a)(x ),y) o o ≤ o o o Kd(ρ(a)¯ϕϕ(x ),ρ(a)(x )) + C + Kd(¯ϕϕρ(a)(x ), ϕ¯(y)) + C ≤ o o o K2d(¯ϕϕ(x ), x )+ KC + C + Kd(ρ(a)(x ), ϕ¯(y)) + KC + C ≤ 0 0 o K2C + KC + C + KR + KC + C = KR + K2C +2KC +2C. ≤ Since y Y was arbitrary, the set ρ¯(A).y is (KR + K2C +2KC +2C)-relatively dense in Y , and ∈ o hence ρ¯ is cobounded. The converse follows by reversing the roles of X and Y . (ii) Assume that ρ is proper, i.e. for every R> 0 and every x X the set ∈ Aρ := a A B(x, R) ρ(a).B(x, R) = x,R { ∈ | ∩ 6 ∅} is finite. Now let y Y and assume that ∈ B(y, R) ρ¯(a).B(y, R) = (3.1) ∩ 6 ∅ for some R 0 and a A. We then find y ,y B(y, R) such that ρ¯(a)(y ) = y , and we define ≫ ∈ 1 2 ∈ 1 2 x :=ϕ ¯(y), x :=ϕ ¯(y ) and x :=ϕ ¯(y ). Then x , x B (x) and 1 1 2 2 1 2 ∈ KR+C x2 =ϕ ¯(y2)=ϕ ¯ρ¯(a)(y1)=ϕϕρ ¯ (a)¯ϕ(y1)=ϕϕρ ¯ (a)x1, hence d(ρ(a)x1, x2) < C and thus d(ρ(a)x , x) d(ρ(a)x , x )+ d(x , x) < C + KR + C = KR +2C =: R′. 1 ≤ 1 2 2 ρ We deduce that x ,ρ(a)x B ′ (x) and hence a A ′ . This shows that for every R 0 the set { 1 1} ⊂ R ∈ x,R ≫ of a A satisfying (3.1) is finite, whence ρ¯ is proper. The converse follows again by reversing the roles ∈ of X and Y . 

Similarly we have:

LEMMA 4.24. Let (X, d) be a proper pseudo-metric space, C > 0, let A be a set and let ρ ,ρ : A 0 1 → Map(X,X) be such that d(ρ (a)(x),ρ (a)(x)) < C for all a A and x X. Then ρ is cobounded, 0 1 ∈ ∈ 0 respectively proper, if and only if ρ1 has the corresponding property.

PROOF. The statement about coboundedness follows immediately from the fact that ρ0(A).x and ρ (A).x are at bounded Hausdorff distance for each x X. As for properness, assume that ρ is 1 ∈ 0 proper and assume that for some γ A, o X and R > 0 we have B(o, R) ρ (γ).B(o, R) = , say ∈ ∈ ∩ 1 6 ∅ x′ = ρ (γ)(x) is in this intersection, for some x, x′ B(o, R). Then x′′ := ρ (γ)x B(o, R + C) and 1 ∈ 0 ∈ hence B(o, R + C) ρ (γ).B(o, R + C) = . This shows that γ was contained in a finite set, hence ρ is ∩ 0 6 ∅ 1 proper. 

At this point we have established Proposition 4.22. We also mention the following closure prop- erty of geometric qiqacs.

LEMMA 4.25. If a qiqac ρ : (Λ, Λ∞) QI(X) is proper, cobounded or geometric, then the induced → quasi-isometric quasi-action ρ : (Λk, Λ∞) QI(X) has the same property. → f f 4. LEFT-REGULAR QUASI-ACTIONS 51

PROOF. The statement about coboundedness is immediate since ρ(Λk).x ρ(Λ).x for every x ∞ k ⊃ ∈ X. As for properness, let Fk Λ be finite such that Λ ΛFk. Since Fk is finite, we can choose ′ ⊂ ⊂ k′ > k such that F Λk . Choose constants K > 1 and C > 0 such that ρ is a (K , C , C )-qiqac. k ⊂ k k k k k From now on we assume that B(x, R) ρ(g).B(x, R) = for some g Λk, x X and R> 0. This ∩ 6 ∅ ∈ ∈ means that there exist x′, x′′ B(x, R) such that ∈ x′′ = ρ(g)(x′). (3.2) We need to show that g is confined to a finite subset of Λk. We first claim that M(x, R) := max sup d(ρ(f)x′, x′) < . (3.3) f∈Fk x′∈B(x,R) ∞ Indeed, if f F and x′ B(x, R), then ∈ k ∈ ′ ′ ′ ′ d(ρ(f)x , x ) d(ρ(f)x ,ρ(f)x)+ d(ρ(f)x, x)+ d(x, x ) Kk′ R + Ck′ + d(ρ(f)x, x)+ R, ≤ ≤ ′ which is bounded uniformly in x′. Secondly, we can write g = λf for some λ Λ and f F Λk . ∈ ∈ k ⊂ For all x′ B(x, R) we then have ∈ ′ ′ ′ ′ ′ ′ d(ρ(g)x ,ρ(λ)x ) = d(ρ(λf)x ,ρ(λ)x ) d(ρ(λ)ρ(f)x ,ρ(λ)x )+ C ′ ≤ k ′ ′ K d(ρ(f)x , x )+ C + C ′ K M(x, R)+ C + C ′ . ≤ k k k ≤ k k k ′′′ ′ Let us abbreviate C(x) := KkM(x, R)+ Ck + Ck′ an define x := ρ(λ)x . Then d(x′′, x′′′)= d(ρ(g)x′,ρ(λ)x′) C(x), ≤ hence x′, x′′′ B(x, R + C(x)). We thus deduce that B(x, R + C(x)) ρ(λ).B(x, R + C(x)) = . By ∈ ∩ 6 ∅ properness of the original qiqac we conclude that λ is confined to a finite subset of Λ. Since g = λf for some f F , this implies that g is confined to a finite subset of Λk. This finishes the proof.  ∈ k We have seen in Proposition 3.36 that every apogee of a geometrically finitely-generated approxi- mate group is necessarily quasi-cobounded. The same conclusion also holds for pseudo-metric spaces which admit a cobounded qiqac of an approximate group. We can actually make a more precise state- ment using the following terminology:

DEFINITION 4.26. Let X be a pseudo-metric space. We say that X is semi-uniformly quasi-cobounded if there exist constants K 1 and C, R 0 and a set of (K, C)-quasi-isometries of X with the ≥ ≥ A following properties: For all x, y X there exists g A such that d(g(x),y) < R and there exist ∈ ∈ constants C 0 such that k consists of (K, C )-quasi-isometries for all k N. We say that X is k ≥ A k ∈ uniformly quasi-cobounded if moreover the sequence Ck can be chosen to be bounded. We record the following immediate consequences of this definition for later reference:

PROPOSITION 4.27. If a metric space (X, d) admits a geometric qiqac of an approximate group (Λ, Λ∞), then it is quasi-cobounded. If this qiqac can be chosen to be uniform, then it is even semi-uniformly quasi- cobounded (but not necessarily uniformly quasi-cobounded). 

4. Left-regular quasi-actions The purpose of this section is to show that every algebraically or geometrically finitely-generated approximate group admits a geometric qiqac on a metric space (X, d), which in the case of a geo- metrically finitely-generated group can be chosen to be a proper geodesic metric space. The case of an algebraically finitely-generated algebraic group was already treated in [BH18a], but we recall it here, since the argument in the geometrically finitely-generated case is based on similar ideas. Our inspiration comes again from the group case:

EXAMPLE 4.28 (The group case). If Γ is a finitely-generated group with finite generating set S and associated word metric dS, then left-regular action of Γ on itself defines a geometric isometric action on (Γ, dS). If X is any representative of the canonical QI type of [Γ], then we can quasi-conjugate the 52 4. GEOMETRIC QUASI-ACTIONS OF APPROXIMATE GROUPS left-regular action to obtain a geometric qiqac of Γ on X. Note that the resulting quasi-conjugacy class of qiqacs does not depend on the generating set S. We will give similar constructions both for the internal and external QI type. For the moment let (Λ, Λ∞) be an arbitrary approximate group (without any finiteness assumptions). We fix once and for all a left-admissible metric d on Λ∞. Following [BH18a] we are going to construct a geometric quasi-action of (Λ, Λ∞) on (Λ, d ). |Λ×Λ Let us denote by L : (Λ, Λ∞) Is(Λ∞, d) the left-regular action of Λ∞ and note that Λ= L(Λ)(e) → is just the quasi-orbit of e under L; in particular, Λ is quasi-invariant under L(Λ). We may thus form the quasi-restriction λ := L : (Λ, Λ∞) QI(Λ, d ) Λ → |Λ×Λ as in Construction 4.19. Let us describe λ more explicitly to get an estimate of the QI-constants of this qiqac: e f Let F Λ3 be such that Λ2 ΛF and set δ := max d(f,e). Then, for all n N, the inclusions ⊂ ⊂ f∈F ∈ ιn+1 :Λn Λn+1 are (1, 0,δ)-quasi-isometries, so we can find quasi-inverses pn+1 :Λn+1 Λn such n → n → that pn+1ιn+1(x)= x for all x Λn and d(ιn+1(pn+1(x)), x) δ for all x Λn+1, n n ∈ n n ≤ ∈ and such that each pn+1 is a (1, 2δ, 0)-quasi-isometry. Thus, if for all k

2 3 + k+1 λ(g):Λ i Λ j Λ j ... j Λ . 2 3 4 k+1 p1 p2 p3 pk

We deduce that λ(g):Λ Λ is a (1, 2kδ)-quasi-isometric embedding. Moreover, if g Λk, h Λl and → ∈ ∈ x Λ, then we have ∈ d(λ(gh)(x), λ(g)λ(h)(x)) 4(k + l)δ. ≤ Therefore, if k = l we get d(λ(gh)(x), λ(g)λ(h)(x)) 8kδ. Since λ(e)=Id , note that for all x X and ≤ Λ ∈ g Λk we have ∈ d(λ(g)λ(g−1)(x), x) 8kδ and d(λ(g−1)λ(g)(x), x) 8kδ. ≤ ≤ In particular, λ(g)(Λ) is 8kδ-relatively dense in Λ, hence λ(g) is a (1, 2kδ, 8kδ)-quasi-isometry with quasi-inverse λ(g−1). We have thus established:

PROPOSITION 4.29. For every left-admissible metric d on Λ∞, the map λ : (Λ, Λ∞) QI(Λ, d ), → |Λ×Λ g λ(g) is a cobounded quasi-isometric quasi-action, in fact a cobounded (2kδ, 8kδ)-rough quasi-action.  7→ f At this point the arguments for the external and internal QI type start to diverge. Let us first deal with the external case. Thus assume that Λ∞ is finitely-generated and fix a finite-generating set ∞ S of Λ and associated word metric dS. From Proposition 4.29 we then obtain a cobounded rough quasi-action λ : (Λ, Λ∞) QI(Λ, d ), g λ(g). (4.1) → S|Λ×Λ 7→ k+1 We recall that this quasi-action depends on a choice of projections (p )k≥1. f k k+1 LEMMA 4.30. (i) For any choice of finite generating set S and projections (pk )k≥1, the quasi-action (4.1) is proper, hence geometric. (ii) For any two choices the resulting quasi-actions are quasi-conjugate. 4. LEFT-REGULAR QUASI-ACTIONS 53

PROOF. (i) Denote by B(o, R) the ball of radius R around o in Λ∞ and set B (o, R) := B(o, R) Λ. Λ ∩ Assume that λ(g).B (o, R) B (o, R) = for some g Λ. Then there exist x, x′ B (o, R) such Λ ∩ Λ 6 ∅ ∈ ∈ Λ that λ(g)x = x′. Set x′′ := gx. Then d(x′′, λ(g)x) 2δ, hence x, x′′ B(o, R +2δ). We deduce that ≤ ∈ g.B(o, R) B(o, R) = , which by properness of the action of Λ∞ on (Λ∞, d ) confines g to a finite ∩ 6 ∅ S subset of Λ∞. (ii) We first fix a generating set S and define δ as above. By construction we then have d(λ(g)(h),gh) < 2kδ for all g Λ∞ and h Λ, i.e., we have d(λ(g)(h),gh) < 2kδ independently of the choice of pro- k+1 ∈ ∈ jections pk . We deduce that any two left-regular quasi-actions with respect to the same generating set are equivalent. Now let S and S′ be two different finite generating sets for Λ∞. We can then find a quasi-isometry ∞ ∞ ′ ϕ : (Λ , d ) (Λ , d ′ ) with quasi-inverse ϕ¯. Then quasi-conjugation by (ϕ, ϕ ) turns left-regular S → S quasi-actions with respect to S into left-regular quasi-actions with respect to S′, where the correspond- ing projections are also quasi-conjugate. This finishes the proof. 

REMARK 4.31 (Uniqueness of external left-regular quasi-action). Note that if (X, d) is an arbitrary representative of [Λ]ext, then it is quasi-isometric to (Λ, d) for some (hence any) external metric d on Λ, hence we can quasi-conjugate the geometric quasi-action on (Λ, d) defined above to a geometric quasi- action on X. We refer to any quasi-action obtained in this way as an external left-regular quasi-action. By Lemma 4.30 the external left-regular quasi-action is unique up to quasi-conjugacy. The previous construction shows that if (Λ, Λ∞) is algebraically finitely-generated, then it admits a geometric quasi-action on any chosen representative of [Λ]ext. We now want to establish a similar result for representatives of the internal QI type. This is more subtle and requires us to appeal to the uniform version of Gromov’s trivial lemma (Proposition A.31). Thus let (Λ, Λ∞) be a geometrically finitely-generated approximate group and let (X, d) be a rep- ∞ resentative of [Λ]int. Also, let d be the restriction of a left-admissible metric on Λ to Λ. By Proposition 4.29 we have a rough quasi-action b λ : (Λ, Λ∞) QI(Λ, d), g λ(g), (4.2) → 7→ and by the choice of (X, d) we have mutually coarsely inverse coarse equivalences b f b ψ : (Λ, d) (X, d) and ψ : (X, d) (Λ, d). → → Now if f QI(Λ, d), then ψ f ψ is a coarse self-equivalence of (X, d). Since (X, d) is large-scale ∈ ◦ ◦ b b geodesic, it is thus a quasi-isometry by the symmetric version of Gromov’s trivial lemma (Corollary A.30). We thusf obtainb a well-defined map ψ : QI(Λ, d) QI(X, d), f ψ f ψ. ∗ → 7→ ◦ ◦ Moreover, by Proposition A.31 this map preserves uniform subsets of quasi-isometries, i.e. for all f b f K 1, C, C′ 0 there exist k 1, c,c′ 0 such that ≥ ≥ ≥ ≥ ψ QI ′ (Λ, d) QI ′ (X, d). ∗ K,C,C ⊂ k,c,c We deduce:   f b f PROPOSITION 4.32. The map λ := ψ λ : (Λ, Λ∞) QI(X, d) is a uniform quasi-action. ∗ ◦ → PROOF. Choose C , C′ > 0 for all n N such that λ is a (1, C , C′ )-qiqac. Then for every n N n n b f n n n ∈ ∈ we have ρ(Λ ) QI ′ (Λ, d Λ×Λ) and hence by the previous remark we find constants k 1 ⊂ Kn,Cn,Cn | ≥ ′ n b n and cn,c 0 such that λ(Λ ) QI ′ (Λ, d). On the other hand if λ1, λ2 Λ , then for all x Λ n ≥ ⊂ k,cn,cn ∈ ∈ we have f b ′ fd(λ(λ1)λ(λ2).x, λ(λ1λ2).x) < Cn, and thus b b b b ′ d(λ(λ1)λ(λ2).x, λ(λ1λ2).x) < Φ+(Cn), where Φ+ is an upper control for ψ. This finishes the proof.  54 4. GEOMETRIC QUASI-ACTIONS OF APPROXIMATE GROUPS

In the sequel we refer to a geometric quasi-action of (Λ, Λ∞) which arises in this way as an internal left-regular quasi-action of (Λ, Λ∞) on (X, d). Appealing to Proposition A.31 again, one sees that any two internal left-regular quasi-actions are quasi-conjugate. Finally one deduces:

PROPOSITION 4.33. If (Λ, Λ∞) is a geometrically finitely-generated approximate group, then every left- regular quasi-action of (Λ, Λ∞) is geometric.

PROOF. We consider a left-regular quasi-action λ : (Λ, Λ∞) QI(X, d) for some (X, d) [Λ] with → ∈ notation as above. Coboundedness of λ is immediate from the fact that ψ is coarsely onto. As for properness, one first observes as in the external case that λ is proper.f Since coarse equivalences are proper maps, one deduces that λ is also proper.  b To summarize:

PROPOSITION 4.34. If (Λ, Λ∞) is a geometrically finitely-generated approximate group and X [Λ] , ∈ int then (Λ, Λ∞) admits a uniform qiqac on X.  We record a number of useful consequences of Proposition 4.34. With Proposition 4.27 and Theo- rem 3.29 we obtain:

COROLLARY 4.35. If (Λ, Λ∞) is a geometrically finitely-generated approximate group, then every X ∈ [Λ]int is a semi-uniformly quasi-cobounded coarsely connected space. Moreover, we can choose X to be a generalized Cayley graph for (Λ, Λ∞); this then shows:

COROLLARY 4.36. If (Λ, Λ∞) is a geometrically finitely-generated approximate group, then it admits a uniform geometric qiqac on a quasi-cobounded proper metric space, and in fact on a quasi-cobounded locally finite graph. 

5. The Milnor-Schwarz lemma and its variants The classical Milnor–Schwarz lemma says that orbit maps of geometric actions of finitely-genera- ted groups on proper large-scale geodesic spaces are quasi-isometries. In fact, finite generation of the group in question does not have to be assumed but follows automatically from the existence of such a geometric action. There are several directions in which this can be extended: Firstly, one can replace actions by quasi-action without a change in the conclusion. Secondly, one can replace the assumption of large- scale geodesicity by coarse connectedness; the conclusion then becomes that orbit maps are coarse equivalences. All of these versions still hold for geometrically finitely-generated approximate groups, and with weaker conclusions also for arbitrary approximate groups.

THEOREM 4.37 (General Milnor–Schwarz lemma for approximate groups). Let (Λ, Λ∞) be an ap- proximate group, (X, d) be a coarsely-connected proper pseudo-metric space and let ρ : (Λ, Λ∞) QI(X) be a → (K , C , C′ )-qiqac of (Λ, Λ∞) on X. Let x X be a basepoint and denote by ι :Λ X the associated orbit k k k 0 ∈ → map given by λ ρ(λ).x . f 7→ 0 (i) If ρ is cobounded, then there exists R > 0 such that F := µ Λ2 B (x ) ρ(µ).B (x ) = { ∈ | R 0 ∩ R 0 6 ∅} generates Λ∞. If moreover ρ is geometric, then the following hold. (ii) (Λ, Λ∞) is algebraically finitely-generated. (iii) The orbit map ι : (Λ, d ) (X, d), λ ρ(λ).x is a coarse equivalence for some (hence any) finite F |Λ×Λ → 7→ 0 generating set F of Λ∞. ′ (iv) If all of the sequences (Kk), (Ck) and (Ck) are bounded, then the upper control of ι can be chosen to be affine linear, i.e. there exists constants K 1 and C 0 such that d(ι(µ), ι(λ)) Kd (µ, λ)+ C. In ≥ ≥ ≤ F particular, this happens if ρ is an isometric action or Λ=Λ∞ is a group. If moreover X is large-scale geodesic, then the following hold. 5. THE MILNOR-SCHWARZ LEMMA AND ITS VARIANTS 55

(v) The lower control of ι can be chosen to be affine linear, i.e. there exists constants K 1 and C 0 such −1 ≥ ≥ that d(ι(µ), ι(λ)) K dF (µ, λ) C. ≥ − ′ (vi) If all of the sequences (Kk), (Ck) and (Ck) are bounded, then ι : (Λ, dF ) (X, d) is a quasi-isometry → ∞ and hence (X, d) represents [Λ]ext. In particular, this happens if ρ is an isometric action or Λ=Λ is a group. If moreover (Λ, Λ∞) is geometrically finitely-generated, then

(vii) (X, d) represents [Λ]int. In particular, if a geometrically finitely-generated approximate group quasi-acts geometrically by quasi-isometri- es on a large-scale geodesic proper pseudo-metric space, then this space represents its internal and external QI type.

PROOF. Throughout the proof we will assume without loss of generality that (Kk) and (Ck) are ′ increasing sequences and that Ck < Ck. We also choose (Dk) as in Lemma 4.14. (i) Fix a basepoint x X and a constant C 0 such that X is C-coarsely connected and the 0 ∈ ≥ quasi-orbit := ρ(Λ).x is C-relatively dense in X. Then for every λ Λ there exist a C-path from O 0 ∈ x to x = ρ(λ).x , i.e., there exist points x , x ,...,x X such that d(x , x ) < C. Since is 0 n 0 0 1 n ∈ i i+1 O C-relatively dense, there exist λ ,...,λ Λ such that λ = e, λ = λ and d(ρ(λ ).x , x ) < 2C. We 0 n ∈ 0 n i 0 i then have

d(ρ(λi).x0,ρ(λi+1).x0) < 3C, hence there exists R> 0 (depending on C and the implied quasi-isometry constants) such that −1 d(ρ(λi λi+1).x0, x0) < R and thus λ−1λ ,...,λ−1 λ F . On the other hand, { 0 1 n−1 n}⊂ λ = λ−1λ = (λ−1λ )(λ−1λ ) (λ−1 λ ), 0 n 0 1 1 2 ··· n−1 n and since λ Λ was arbitrary we conclude that F = Λ =Λ∞. ∈ h i h i (ii) This follows from (i) and Remark 4.21. (iii), (iv) Since the statement does not depend on the choice of generating system, we may assume that F := µ Λ2 B (x ) ρ(µ).B (x ) = where R is chosen as in (i). (Note again that F is finite { ∈ | R 0 ∩ R 0 6 ∅} by Remark 4.21.) Since ρ is cobounded, the orbit map ι has relatively dense image. It thus remains to find upper and lower controls for ι, and to show that the upper control can be chosen affine linear if ′ ′ (Kn) and (Cn) (and hence (Cn) by our convention Cn < Cn) are bounded. Since F is finite, the maximum

CF := max d(x0,ρ(f)x0) f∈F exists, and we define an upper control Φ+ by Φ+(n) := K1(K2nCF +2C2n)n + (K1 + 1)C1 + D1 and note that this upper control is affine linear if and only if (Kn) and (Cn) are bounded. Now denote by ∞ dρ the pseudo-metric on Λ given by

dρ(µ, λ) := d(ρ(µ).x0,ρ(λ).x0)= d(ι(µ), ι(λ)), and observe that for µ, λ Λ we have ∈ d (µ, λ) = d(ρ(µ).x ,ρ(λ).x ) K d(x ,ρ(µ)−1ρ(λ).x )+ C K d(x ,ρ(µ−1)ρ(λ).x )+ C + D ρ 0 0 ≤ 1 0 0 1 ≤ 1 0 0 1 1 K d(x ,ρ(µ−1λ).x )+ K C + C + D . ≤ 1 0 0 1 1 1 1 Now if µ, λ Λ with d (µ, λ) = n, then there exist f ,...,f F such that µ−1λ = f f , and ∈ F 1 n ∈ 1 ··· n hence d (µ, λ) K d(x ,ρ(µ−1λ).x ) + (K + 1)C + D = K d(x ,ρ(f f ).x ) + (K + 1)C + D ρ ≤ 1 0 0 1 1 1 1 0 1 ··· n 0 1 1 1 n−1 K1 d(ρ(f1 fk).x0,ρ(f1 fk+1).x0) + (K1 + 1)C1 + D1, ≤ ··· ··· ! Xk=0 56 4. GEOMETRIC QUASI-ACTIONS OF APPROXIMATE GROUPS where we will take that, for k =0, the expression ρ(f ...f ).x stands for (just) x . So let 0 k n 1. 1 k 0 0 ≤ ≤ − Since F Λ2, we have f f Λ2k, leading to the estimate ⊂ 1 ··· k ∈ d(ρ(f f ).x ,ρ(f f ).x ) 1 ··· k 0 1 ··· k+1 0 d(ρ(f f ).x ,ρ(f f )ρ(f ).x )+ d(ρ(f f )ρ(f ).x ,ρ(f f ).x ) ≤ 1 ··· k 0 1 ··· k k+1 0 1 ··· k k+1 0 1 ··· k+1 0 (K d(x ,ρ(f ).x )+ C )+ C K C +2C K C +2C . ≤ 2k 0 k+1 0 2k 2k ≤ 2k F 2k ≤ 2n F 2n Plugging this into our previous estimate for dρ(µ, λ) we deduce that n−1 dρ(µ, λ) K1 K2nCF +2C2n + (K1 + 1)C1 + D1 ≤ ! kX=0 = K1(K2nCF +2C2n)n + (K1 + 1)C1 + D1 = Φ+(n).

Using that n = dF (µ, λ) and unraveling the definition of dρ we thus obtain d(ι(µ), ι(λ)) Φ (d (µ, λ)). (5.1) ≤ + F To obtain a lower control we introduce the following notation: Given n N we define ∈ B := λ Λ d (e, λ)

REMARK 4.38. Assume that (Λ, Λ∞) is distorted and let d be an external metric on Λ. Then (Λ, Λ∞) quasi-acts geometrically by quasi-isometries on (Λ, d), but by assumption (Λ, d) does not represent [Λ]int. This shows in particular that (vii) does not hold in general without the assumption that (X, d) be large-scale geodesic. We now derive a number of consequences of Theorem 4.37. Firstly, we obtain the following impli- cation:

COROLLARY 4.39 (Geometric finite generation implies algebraic finite generation). If (Λ, Λ∞) is a geometrically finitely generated group, then it is also algebraically finitely generated.

PROOF. By Corollary 4.36, (Λ, Λ∞) admits a geometric qiqac on a proper geodesic metric space. The corollary then follows from Theorem 4.37.(ii).  Secondly, we obtain a new geometric criterion for undistortedness:

COROLLARY 4.40 (Undistortedness from isometric actions). Let (Λ, Λ∞) be a geometrically finitely- generated approximate group which acts geometrically by isometries on a large-scale geodesic proper pseudo- metric space (X, d). Then (Λ, Λ∞) is algebraically finitely-generated and undistorted. 5. THE MILNOR-SCHWARZ LEMMA AND ITS VARIANTS 57

PROOF. Finite generation follows from Theorem 4.37.(ii). On the other hand, by Theorem 4.37.(vi) and (vii) we have [Λ]ext = [(X, d)] = [Λ]int, which shows that (Λ, Λ∞) is undistorted.  Thirdly, we can now finish the proof of Theorem 4.7:

PROOF OF THEOREM 4.7, CONCLUDED. If (Λ, Λ∞) is as in Theorem 4.7, then by Theorem 4.37.(ii) the group Λ∞ is finitely-generated. Moreover, if S is any finite generating set of Λ∞, then by Theorem 4.37.(vi) there is a quasi-isometry between (X, d) and (Λ, d ), and consequently (Λ, d ) is S|Λ×Λ S|Λ×Λ large-scale geodesic. This shows that (Λ, Λ∞) is geometrically finitely-generated. Moreover, some (hence any) external metric on Λ is an internal metric, and thus (Λ, Λ∞) is undistorted. Since the equivalence (i) (ii) was already established in Section 1, this finishes the proof.  ⇐⇒ Finally, it turns out – somewhat surprisingly – that under some mild hypothesis every geometric qiqac can be replaced by a uniform geometric qiqac on the same space. This is a consequence of the implication (iii) (ii) of the following corollary. ⇒ COROLLARY 4.41 (Making geometric quasi-actions uniform). Let (Λ, Λ∞) be a geometrically finitely- generated approximate group and let (X, d) be a large-scale geodesic proper pseudo-metric space. Then the following are equivalent: ∞ (i) (X, d) represents the internal QI type of (Λ, Λ ), i.e. [(X, d)] = [Λ]int. (ii) There exists a geometric uniform qiqac of (Λ, Λ∞) on X. (iii) There exists a geometric qiqac of (Λ, Λ∞) on X.

PROOF. The implication (i) (ii) was established in Proposition 4.34, the implication (ii) (iii) is ⇒ ⇒ obvious and (iii) (i) is part of the Milnor-Schwarz lemma (Theorem 4.37).  ⇒ A closer inspection of the proof actually shows that every geometric qiqac of (Λ, Λ∞) on X is quasi-conjugate to a uniform one.

CHAPTER 5

Hyperbolic approximate groups and their boundaries

In this chapter we develop the foundations of a theory of hyperbolic approximate groups. Hy- perbolic groups have a number of special properties not shared by general hyperbolic spaces: For example, they are always visible and have bounded growth at some scale; the latter ensures that they can be realized by convex subsets of real-hyperbolic space. Moreover, their Gromov boundaries are doubling and locally self-similar. We will show in the current chapter that all of these results can also be established for hyperbolic approximate groups, except that local self-similarity has to be replaced by a quasified version. We also show that, as in the group case, the Gromov boundary of a hyperbolic approximate group has either 0, 2 or infinitely many points and that it is minimal provided the approximate group is non-elementary. However, some care has to be taken concerning the definition of a non-elementary hyperbolic approximate group for this to hold, since some of the equivalent characterizations of non- elementary hyperbolic groups no longer hold in the approximate setting.

1. Hyperbolic approximate groups Hyperbolic spaces can be defined as metric spaces with some large-scale version of negative cur- vature. In fact, as we discuss in Appendix C, there are (at least) three subtly differing definitions of a hyperbolic space in the literature, which we refer to as Gromov hyperbolic, Rips hyperbolic and Morse hyperbolic spaces respectively. For proper geodesic metric spaces, all three definitions coincide, hence we can unambiguously talk about hyperbolic proper geodesic metric spaces in this case. We will occasionally have to work with spaces which are not geodesic (in particular, in the context of Morse boundaries), and in this case we prefer the notion of Morse hyperbolicity, since unlike the other two notions it is QI-invariant. Let us briefly recall the definition; see Definition C.2 for the notion of a δ-slim triangle.

DEFINITION 5.1. A metric space (X, d) is called Morse hyperbolic if it is large-scale geodesic and for all K 1 and C 0 there exists δ 0 so that every (K, C)-quasi-geodesic triangle is δ-slim. ≥ ≥ ≥ REMARK 5.2 (Morse hyperbolic vs. Gromov hyperbolic). (i) Since quasi-isometries preserve slim quasi-geodesic triangles, it is immediate from the definition that being Morse hyperbolic is a QI-invariant among metric spaces (cf. Theorem C.7). (ii) By Theorem C.7, Morse hyperbolic spaces can also be characterized as those quasi-geodesic spaces in which every quasi-geodesic is Morse (with a Morse gauge which is uniform in the parameters of the quasi-geodesic), hence the name. (iii) Not every Gromov hyperbolic space is Morse hyperbolic, since Gromov hyperbolic spaces need not be quasi-geodesic. Conversely, not every Morse hyperbolic space is Gromov hyperbolic: The graph of the function f : R R, x x with the induced metric as a subspace of R2 is quasi- → 7→ | | isometric (in fact, bi-Lipschitz) to the real line, hence it is Morse hyperbolic. However, it is not Gromov hyperbolic (see [BS07a, Rem. 4.1.3]). (iv) A Morse hyperbolic space is Gromov hyperbolic if and only if it is quasi-ruled [BHM11]. In particular, every proper geodesic Gromov hyperbolic space is Morse hyperbolic. (v) One can characterize Morse hyperbolic spaces extrinsically as those metric spaces which are quasi-isometric to proper geodesic Gromov hyperbolic spaces (see Theorem C.7). Thus the class

59 60 5. HYPERBOLIC APPROXIMATE GROUPS AND THEIR BOUNDARIES

of Morse hyperbolic spaces is the smallest QI-invariant class of metric spaces which contains all proper geodesic Gromov hyperbolic spaces. Recall that a finitely-generated group Γ is called hyperbolic if some (hence any) of its Cayley graphs is Gromov hyperbolic in the sense of Definition C.1. Equivalently, every representative of the canonical QI class [Γ] is Morse hyperbolic. (On the contrary, every Cayley graph of a non-elementary hyperbolic group is quasi-isometric to a metric space which is not Gromov hyperbolic, see [BHM11].) We extend this definition to the case of approximate groups:

DEFINITION 5.3. An approximate group (Λ, Λ∞) is called hyperbolic if it is geometrically finitely- generated and some (hence any) X [Λ] is Morse hyperbolic. ∈ int By Remark 5.2, a geometrically finitely-generated approximate group is hyperbolic if and only if some (hence any) apogee, or equivalently some (hence any) generalized Cayley graph, is Gromov hyperbolic. In conjunction with Corollary 4.35 and Remark 5.2 this leads to the following observation:

PROPOSITION 5.4. If (Λ, Λ∞) is a hyperbolic approximate group, then every apogee X of (Λ, Λ∞) is a semi-uniformly quasi-cobounded proper geodesic Gromov hyperbolic space. In particular, every generalized Cayley graph of (Λ, Λ∞) is a semi-uniformly quasi-cobounded Gromov hyperbolic graph. 

2. Hyperbolic approximate groups are visual The first important property of hyperbolic groups which we extend to the broader context of hyperbolic approximate groups is visuality:

DEFINITION 5.5. A proper geodesic Gromov hyperbolic space X is called visual if for some (hence for any) base point o X, there is a positive constant D such that every point in X has distance at ∈ most D from some geodesic ray emanating from o.

REMARK 5.6. Note that in [BS00], a metric space X is said to be visual if for some (hence for any) base point o X, there is a positive constant D such that each point in X lies on a D-roughly geodesic ∈ ray emanating from o. In[BS07a], a Gromov hyperbolic space is said to be visual if for some (hence for any) basepoint o X, there is a positive constant D such that for every x X there is a ξ ∂ X such ∈ ∈ ∈ s that d(o, x) (x ξ) + D. These two definitions are equivalent to each other and also to the definition ≤ | o above for proper geodesic hyperbolic spaces. One should think of visuality as a coarse version of the geodesic extension property (the ability to extend a geodesic segment into a geodesic ray). Infinite hyperbolic approximate groups are visual in the following sense:

PROPOSITION 5.7 (Hyperbolic approximate groups are visual). Every apogee of an infinite hyperbolic approximate group (Λ, Λ∞) is visual. In view of Proposition 5.4 this is a special case of the following general result, whose proof follows closely the well-known argument in the group case (see e.g. [Gen19]).

PROPOSITION 5.8. Every unbounded proper geodesic hyperbolic quasi-cobounded space X is visual.

PROOF. Since X is quasi-cobounded, there exist constants K 1 and C, C′, R 0, and a collection ≥ ≥ QI(X) of (K,C,C′)-quasi-isometries such that for all x, y X, there is a g with g(y) A ⊆ ∈ ∈ A ∈ B(x, R). We choose such a collection once and for all. Letfo X be a basepoint and x X an arbitrary point. We first prove that there is a bi-infinite ∈ ∈ geodesic γ at bounded distance from x. Since X is unbounded, we can pick two sequences of points (x ) and (y ) in X with the property that d(x ,y ) . For each pair (x ,y ) we then choose a n n n n → ∞ n n geodesic segment [x ,y ] connecting them. For every n N we denote by a the midpoint of [x ,y ]. n n ∈ n n n We choose a quasi-isometry g such that a′ := g (a ) B(x, R) and define x′ := g (x ) and n ∈ A n n n ∈ n n n y′ := g (y ). Then d(x′ ,a′ ) and d(y′ ,a′ ) . Moreover, g ([x ,y ]) is a quasi-geodesic n n n n n → ∞ n n → ∞ n n n segment from x′ to y′ which contains a′ , hence is of bounded distance from x. Since is uniform, n n n A 3. BOUNDARIES OF HYPERBOLIC APPROXIMATE GROUPS 61 both the implied QI-constants and the distance from x of gn([xn,yn]) can be bounded independently of n. ′ ′ ′ ′ Now let [xn,yn] be a geodesic segment from xn to yn. By Lemma C.5 the quasi-geodesic segments ′ ′ [xn,yn] and gn([xn,yn]) are at uniformly bounded distance, hence x is at uniformly bounded distance ′ ′ from all of the geodesic segments [xn,yn]. We can thus apply Lemma A.16 to obtain a bi-infinite geodesic line γx which is at uniformly bounded distance from x. For every n Z we now choose a geodesic segment [o, γx(n)] from o to γx(n) and consider the ∈ ′ geodesic triangle formed by [o, γx( n)], γx([ n,n]) and [o, γx(n)]. By Lemma C.3 this triangle is δ - ′ − − slim for some uniform δ . In particular, γx(0) (and hence x) is at uniformly bounded distance from either [o, γ ( n)] or [o, γ (n)]. x − x Another application of Lemma A.16 shows that for some sequence (nk) the geodesic segments [o, γ ( n )] and [o, γ (n )] converge to geodesic rays emanating from o, and at least one of them is at x − k x k uniformly bounded distance from x. 

3. Boundaries of hyperbolic approximate groups In this section, (Λ, Λ∞) will always denote a hyperbolic approximate group. Our goal is to asso- ciate with (Λ, Λ∞) a boundary ∂Λ, defined up to homeomorphism. We first recall the definition of the Gromov boundary of a Gromov hyperbolic spaces.

REMARK 5.9 (Sequential vs. ray boundary). It turns out that there are several definitions of bound- ary for hyperbolic spaces, which however coincide in favorable circumstances (cf. Section 2). For a general Gromov hyperbolic space X (not necessarily geodesic), one can define the sequential bound- ary, ∂sX, of X that consists of equivalence classes of sequences that converge to infinity. When, in addition, X is proper and geodesic, then one can define a boundary consisting of equivalence classes of rays, ∂rX, where two rays are considered equivalent if they have bounded Hausdorff distance. As explained in Section 2, both ∂sX and ∂rX carry natural . Roughly speaking, two equiva- lence classes of rays are close in ∂rX if they fellow travel for a long time. When X is a proper geodesic Gromov hyperbolic metric space, then there the map [γ] [(γ(n)) N] defines a homeomorphism 7→ n∈ ∂ X ∂ X, and hence we will identify both boundaries and simply denote them by ∂X. r → s REMARK 5.10 (Boundaries of Morse hyperbolic spaces). Now let (X, d) be a general Morse hy- perbolic space. In this situation one could define the sequential boundary ∂sX, but this is not a QI-invariant and hence of limited usefulness to us. Instead, if (X′, d′) is a proper geodesic Gromov hyperbolic space which is quasi-isometric to (X, d), then we call ∂X′ a Gromov boundary of X (see Definition C.19). This definition has the disadvantage that a Morse hyperbolic space has many differ- ent Gromov boundaries, but all of these are homeomorphic and their common homeomorphism type ∂[X] := [∂X1] is a QI-invariant.

DEFINITION 5.11. The Gromov boundary of a hyperbolic approximate group (Λ, Λ∞) is defined as the homeomorphism type ∂Λ := ∂[X], where X is an arbitrary representative of [Λ]int.

Note that by construction, ∂Λ depends only on [Λ]int, and hence on [Λ]c (by Lemma 3.31); in particular it is invariant under 2-local isomorphisms (or Freiman 3-isomorphisms. Moreover, if is a I homeomorphism invariant of compact, metrizable topological spaces, then we may define (∂Λ) := (∂X), I I where X is an apogee of (Λ, Λ∞). Following Example C.21 we then refer to (∂Λ) asa topological bound- I ary invariant of Λ. The following topological boundary invariant will play a mayor role in Chapter 7 below:

EXAMPLE 5.12 (Topological dimension of the boundary). If (Λ, Λ∞) is a hyperbolic approximate group, then we define topological dimension of its boundary by dim ∂Λ := dim ∂X, 62 5. HYPERBOLIC APPROXIMATE GROUPS AND THEIR BOUNDARIES where X is an apogee of (Λ, Λ∞) and dim denotes topological (or covering) dimension (see Definition 7.3 below). Boundaries of hyperbolic groups have a number of additional properties compared to boundaries of general Gromov hyperbolic spaces, We now investigate how these properties generalize to hyper- bolic approximate groups.

REMARK 5.13 (Boundaries of hyperbolic groups are locally self-similar). Recall that a metric space (X, d) is called locally self-similar with parameters λ 1, R > 1 if for every R > R and every subset ≥ 0 0 A X with diameter diam A 1 there exists a map f : A X such that ⊂ ≤ R → 1 R d(x , x ) d(f(x ),f(x )) λR d(x , x ), x , x A. λ · 1 2 ≤ 1 2 ≤ · 1 2 ∀ 1 2 ∈ It is an important fact in the theory of hyperbolic groups that the Gromov boundary of a non-element- ary hyperbolic group is locally self-similar with respect to any visual metric. This is no longer true for hyperbolic approximate groups, but a certain quasified version of local self-similarity still holds.

DEFINITION 5.14. We say a metric space X is locally quasi-similar to (subsets of) a metric space Y if there are constants λ 1, K 1, and R > 1 such that for every R > R and every subset C X ≥ ≥ 0 0 ⊂ with diam C 1 , there exists a map f : C Y such that ≤ R → 1 K RK (d (x , x ))K d (f(x ),f(x )) λ √R K d (x , x ), x , x C. (3.1) λ X 1 2 ≤ Y 1 2 ≤ X 1 2 ∀ 1 2 ∈ We say that X is locally quasi self-similar if is locally quasi-similarp to itself.

PROPOSITION 5.15 (Local quasi self-similarity of boundaries). If (Λ, Λ∞) is a hyperbolic approximate group, then any representative of ∂Λ is locally quasi self-similar with respect to an arbitrary visual metric. We will actually establish the following proposition, which in view of Proposition 5.4 implies Proposition 5.15:

PROPOSITION 5.16. The boundary ∂rX of a proper geodesic hyperbolic quasi-cobounded space X is locally quasi-self similar with respect to an arbitrary visual metric. For the proof of Proposition 5.16 we adopt the following notation.

NOTATION 5.17. If (X, d) is a quasi-cobounded hyperbolic space and o X is a basepoint, then ∈ there exist constants K 1, C 0, r> 0 such that for all x, y X there exists a (K, C)-quasi-isometry ≥ ≥ ∈ f : X X, which is also a (K,C,o)-quasi-isometry in the sense of (3.4) from Appendix C, such that → d(f(x),y) < r. We then say that X is (K,C,r,o)-quasi cobounded. Moreover, we will refer to a visual metric on ∂rX with basepoint o and parameters a and c as an (a,c,o)-visual metric for short.

PROOF OF PROPOSITION 5.16. Assume that (X, d) is (K1, C1, r, o)-quasi cobounded and that d0 is an (a0,c0, o)-visual metric on ∂rX. We are going to show that (∂rX, d0) is locally quasi-self similar with parameters

− 1−2K 1 R := c +1,K := K and λ := max c1+2K1 a4K1δ+C1+2δ+r,c 1 aC1+r . 0 0 1 { 0 0 0 0 } Thus let R > R0 and let B ∂rX with diamd (B) 1/R. Since log (R/c0) > 0, one may argue as in ⊂ 0 ≤ a0 the proof of [BS07a, Thm. 2.3.2] to construct a point g X with d(o, g) = log (R/c ) such that for all ∈ a0 0 ξ′, ξ′′ B one has ∈ (ξ′ ξ′′) d(o, g) (ξ′ ξ′′) (ξ′ ξ′′) d(o, g)+4δ. (3.2) | o − ≤ | g ≤ | o − On the other hand, since X is (K1, C1, r, o)-quasi cobounded, we can find a (K1, C1, o)-quasi isometry h such that h(g) B(o, r), and we define ∈ f := ∂ h : B ∂ X. r |B → r 3. BOUNDARIES OF HYPERBOLIC APPROXIMATE GROUPS 63

By Lemma C.24 we then have K−1(ξ′ ξ′′) C (f(ξ′) f(ξ′′)) K (ξ′ ξ′′) + C +2δ. 1 | g − 1 ≤ | h(g) ≤ 1 | g 1 Since d(o, h(g)) < r we deduce that

(f(ξ′) f(ξ′′)) (f(ξ′) f(ξ′′)) + d(o, h(g)) < K (ξ′ ξ′′) + C +2δ + r, (3.3) | o ≤ | h(g) 1 | g 1 and similarly

(f(ξ′) f(ξ′′)) (f(ξ′) f(ξ′′)) d(o, h(g)) > K−1(ξ′ ξ′′) C r. (3.4) | o ≥ | h(g) − 1 | g − 1 − Combining (3.2) and (3.3) we obtain (f(ξ′) f(ξ′′)) K (ξ′ ξ′′) + C +2δ + r | o ≤ 1 | g 1 K ((ξ′ ξ′′) d(o, g)+4δ)+ C +2δ + r ≤ 1 | o − 1 = K (ξ′ ξ′′) K d(o, g)+4K δ + C +2δ + r. 1 | o − 1 1 1 d(o,g) Combining this with the visual inequalities for d0 (see (3.2) in Appendix C) and the fact that a0 = R/c0 we obtain

′ ′′ ′ ′′ −1 −(f(ξ )|f(ξ ))o d0(f(ξ ),f(ξ )) c0 a0 ≥ · ′ ′′ −1 −K1(ξ |ξ )o+K1d(o,g)−4K1δ−C1−2δ−r c0 a0 ≥ · ′ ′′ −1 −4K1δ−C1−2δ−r K1 −(ξ |ξ )o K1 = c0 a0 (R/c0) (a0 ) · · · ′ ′′ −(ξ |ξ ) = c−1−2K1 a−4K1δ−C1−2δ−r RK1 (c a o )K1 0 · 0 · · 0 0 λ−1 RK d (ξ′, ξ′′)K . ≥ · · 0 If instead of starting from (3.2) and (3.3) we start from (3.2) and (3.4), then a similar computation yields

− ′ ′′ − −K 1(ξ |ξ ) +K 1d(o,g)+C +r d (f(ξ′),f(ξ′′)) c a 1 o 1 1 0 ≤ 0 · 0 K ′ ′′ = c aC1+r K1 R/c 1 a−(ξ |ξ )o 0 · 0 · 0 · 0 −1 ′ ′′ 1−2K K qK = c 1 aC1p+r √1 R 1 c a−(ξ |ξ )o 0 · 0 · · 0 0 λ K√R K d (ξ′, ξ′′). q ≤ · · 0 This shows that (X, d0) is locally quasi-self similar withp parameters (R0,K,λ). 

Another fundamental property of Gromov boundaries of hyperbolic groups – also in connection with asymptotic dimension – is the following doubling property with respect to an arbitrary visual metric.

DEFINITION 5.18. Let (Z, d) be a metric space. We say that Z is doubling with doubling constant N N if for all t> 0 and all ξ Z there exist ξ ,...,ξ Z such that ∈ ∈ 1 N ∈ N B(ξ, 2t) B(ξ ,t). ⊂ i i=1 [ We say that Z is doubling at small scales with local doubling constant N if this holds for all sufficiently small t> 0.

Note that if Z is compact, then doubling at small scales implies doubling. We are going to show:

THEOREM 5.19 (Boundaries of hyperbolic approximate groups are doubling). If X is an apogee of a hyperbolic approximate group (Λ, Λ∞), then ∂X is doubling with respect to an arbitrary visual metric. 64 5. HYPERBOLIC APPROXIMATE GROUPS AND THEIR BOUNDARIES

The proof of the corresponding result in the group case is based on local self-similarity. In fact, every metric space which is locally similar to a compact metric space is doubling at small scales (see [BS07a, Lemma 2.3.4]), and hence doubling if it is compact. Unfortunately, we do not know if every space which is locally quasi similar to a compact metric space is doubling at small scales. Nevertheless we can make the argument work for visual metrics on boundaries of approximate groups. Here we use the fact that every snowflake version of a visual metric is again a visual metric, hence locally quasi-self-similar. In view of Proposition 5.16 this argument applies in the following generality:

THEOREM 5.20. The boundary ∂rX of a proper geodesic hyperbolic quasi-cobounded space X is doubling.

PROOF. Since ∂rX is compact it suffices to establish doubling at small scales. For this we assume that X is (K1, C1, r, o)-quasi cobounded and that d+ is an (a+,c+, o)-visual metric on X. We define the K1 K1 associated snowflake metric by d0 := d+; this metric is then (a0,c0, o)-visual, where a0 = a+ and cK1 = c (∂ X, d ) 0 +. By Proposition 5.16 the spacep r 0 is locally quasi-self similar with certain parameters (R0,K,λ). The proof of Proposition 5.16 actually shows that we may assume K = K1. Next we fix ρ (0, (4c cK1 λ)−1). This choice of ρ ensures that ∈ + 0 ρ< 1/(4λ) and K1 4c λρ c < 1. (3.5) + · 0 Since ∂rX is compact we can choose N N such thatp every subset B ∂rX can be covered by at ∈ ⊂ most N balls of radius ρ with respect to the d0-metric, with centers in B. We now define a function R : (0, ) (0, ) by R(t) := K1 λρ/t and choose t > 0 such that for all t (0,t ) we have ∞ → ∞ 0 ∈ 0 R(t) > R . We now fix t (0,t ) and set R := R(t) so that 0 ∈ 0 p RK1 = λρ/t, and hence t = λρ/RK1 . (3.6)

d+ Fix ξ ∂rX and denote by B (ξ, 2t) the ball of radius 2t around ξ with respect to the d+-metric. We ∈ d+ claim that the diameter of B (ξ, 2t) with respect to the metric d0 is less than 1/R. To prove the claim, ′ ′′ d+ let ξ , ξ B (ξ, 2t). Using the visual inequalities of d+ and d0 we obtain ∈ ′ ′′ ′ ′′ −(ξ |ξ ) −(ξ |ξ ) 4t d (ξ′, ξ′′) c−1a o = c−1c−K1 (c a o )K1 c−1c−K1 d (ξ′, ξ′′)K1 , ≥ + ≥ + + + 0 0 0 ≥ + 0 0 hence passing to the supremum and using (3.6) and (3.5) we obtain

d K1 K K K1 K diam (B + (ξ, 2t)) 4tc c 1 = 1 4c c λρ/RK1 = 1 4c λρ c 1/R < 1/R, d0 ≤ + 0 + · 0 · + · 0 · finishing the proof of the claim.q p q p d+ Since diamd0 (B (ξ, 2t)) < 1/R and since (∂rX, d0) is locally quasi-self similar with parameters (R ,K , λ), we can now find a function f : Bd+ (ξ, 2t) ∂ X such that for all ξ , ξ Bd+ (ξ, 2t) we 0 1 → r 1 2 ∈ have 1 K RK1 (d (ξ , ξ ))K1 d (f(ξ ),f(ξ )) λ √1 R K1 d (ξ , ξ ). (3.7) λ 0 1 2 ≤ 0 1 2 ≤ 0 1 2 On the other hand, by our choice of N we can find ξ ,...,ξ Bd+ (ξ,p2t) such that 1 N ∈ N N f(Bd+ (ξ, 2t)) Bd0 (f(ξ ),ρ), and hence Bd+ (ξ, 2t) f −1(Bd0 (f(ξ ),ρ)). ⊂ i ⊂ i i=1 i=1 [ [ It thus suffices to show that for i =1,...,N we have f −1(Bd0 (f(ξ ),ρ)) Bd+ (ξ ,t). (3.8) i ⊂ i For this, let ξ′ f −1(Bd0 (f(ξ ),ρ)). Using (3.7) we obtain i ∈ i ρ > d (f(ξ ),f(ξ′)) λ−1RK1 (d (ξ , ξ′))K1 = λ−1RK1 d (ξ , ξ′), 0 i i ≥ 0 i i · + i i and combining this with (3.6) we obtain

′ K1 d+(ξi, ξi) < λρ/R = t. Since ξ′ f −1(Bd0 (f(ξ ),ρ)) was chosen arbitrarily this establishes (3.8) and finishes the proof.  i ∈ i Theorem 5.19 now follows from Theorem 5.20 in view of Proposition 5.4. 4. EMBEDDINGS OF HYPERBOLIC APPROXIMATE GROUPS 65

4. Embeddings of hyperbolic approximate groups In this section we consider the problem of finding a particularly nice apogee X for a hyperbolic approximate group (Λ, Λ∞). The most classical family of examples of proper geodesic hyperbolic n spaces is given by the real hyperbolic spaces (H )n≥2. More generally, every closed convex subset of Hn (with the induced metric) is a proper CAT(-1) space, and hence a proper geodesic hyperbolic space. We are going to show:

THEOREM 5.21 (Good apogees for hyperbolic approximate groups). Every hyperbolic approximate group (Λ, Λ∞) admits an apogee X which is a closed convex subset of Hn.

If Λ is finite, then we can just choose X to be a point; we will thus assume from now on that Λ is infinite. Our main tool for the proof of Theorem 5.21 is the Bonk–Schramm embedding theorem; to state this theorem we briefly recall the relevant terminology. The following is a very weak version of the doubling property:

DEFINITION 5.22. A metric space (X, d) has bounded growth at some scale if there exist constants R>r> 0 and N N such that every open ball of radius R in X can be covered by N open balls of ∈ radius r.

With this terminology understood, the embedding theorem can be formulated as follows ([BS00, Thm. 1.1]):

THEOREM 5.23 (Bonk-Schramm embedding theorem). If X is a Gromov hyperbolic proper geodesic metric space with bounded growth at some scale, then there exists n N such that X is roughly similar to a ∈ convex subset of hyperbolic n-space Hn.

Here, two metric spaces X and Y are called roughly similar if there exists f : X Y and constants → k,λ > 0 such that λd(x, y) k d(f(x),f(y)) λd(x, y)+ k − ≤ ≤ and f(X) is k-relatively dense in Y . In other words, up to re-scaling the metric, X is quasi-isometric to Y . If X is a Gromov hyperbolic proper geodesic metric space with bounded growth at some scale, then by [BS00, Thm. 9.2] the Gromov boundary ∂rX of X is doubling. In particular, the following theorem generalizes Theorem 5.20.

THEOREM 5.24. Every proper quasi-cobounded metric space X, in particular every apogee of a hyperbolic approximate group, has bounded growth at some scale.

PROOF. Assume that X is (K , C , r, o)-quasi-cobounded. Given x X we choose a (K , C , C )- 1 1 ∈ 1 1 1 quasi-isometry f : X X with d(o, f(x)) r. We also choose a quasi-inverse g of f and set h := g f. → ≤ ◦ Then g is a (K2, C2, C2)-quasi-isometry and h is a (K3, C3, C3)-quasi-isometry which is C3-close to the identity, where K2, C2,K3, C3 are constants depending only on K1 and C1. Given positive real numbers R3 and r3 we define real numbers r0, r1, r2, R0, R1, R2 by r := r := r C , r := K−1(r C ) and R := R +C , R := K R +2K C , R := K R +C +r. 2 1 3− 3 0 2 1− 2 2 3 3 1 3 2 3 3 0 1 1 1 We now choose R and r in such a way that R > r > 0 for all i 0,..., 3 . By properness of X 3 3 i i ∈ { } there exists n N and o ,...,o X such that ∈ 1 n ∈ N B(o, R ) B(o , r ). 0 ⊂ i 0 i=1 [ Moreover, N f(B(x, R )) B(f(x),K R + C ) B(o, K R + C + r)= B(o, R ) B(o , r ). 1 ⊂ 1 1 1 ⊂ 1 1 1 0 ⊂ i 0 i=1 [ 66 5. HYPERBOLIC APPROXIMATE GROUPS AND THEIR BOUNDARIES

If we set xi := g(oi), then applying g to the previous expression yields N N N h(B(x, R )) g(B(o , r )) B(g(o ),K r + C ) = B(x , r ). 1 ⊂ i 0 ⊂ i 2 0 2 i 1 i=1 i=1 i=1 [ [ [ Now let y Im(h) B(x, R ), say y = h(z). Then ∈ ∩ 2 d(x, z) K d(h(x),h(z)) + K C K (d(x, y)+ C )+ K C ≤ 3 3 3 ≤ 3 3 3 3 < K3R2 +2K3C3 = R1, and hence z B(x, R ). This shows that ∈ 1 N Im(h) B(x, R ) h(B(x, R )) B(x , r ). ∩ 2 ⊂ 1 ⊂ i 1 i=1 [ Since Im(h) is C3 relatively dense in X, we deduce that N N B(x, R )= B(x, R C ) N (Im(h) B(x, R )) B(x , r + C )= B(x , r ). 3 2 − 3 ⊂ C3 ∩ 2 ⊂ i 1 3 i 3 i=1 i=1 [ [ Since R3, r3 and N are independent of x, this finishes the proof. 

COROLLARY 5.25. For every quasi-cobounded proper geodesic Gromov hyperbolic space, in particular for every apogee of a hyperbolic approximate group, there exists n N such that X is roughly similar to a ∈ quasi-cobounded convex subset of Hn for some n N.  ∈ Theorem 5.21 now follows immediately from Corollary 5.25. We now discuss some consequences of Theorem 5.21 concerning boundaries of hyperbolic approximate groups. This will lead us to a refinement of Theorem 5.21. From now on X will always denote a closed convex subset of Hn which is an apogee for a hyper- bolic approximate group (Λ, Λ∞) (and in particular quasi-cobounded).

REMARK 5.26 (Boundaries of good apogees as hyperbolic limit sets). In the sequel we denote by n n H the Gromov compactification of hyperbolic n-space and by ∂Hn := H Hn its Gromov boundary, \ which is an (n 1)-sphere. − If Y is an arbitrary subset of Hn then we define its limit set (Y ) as the subset L (Y ) := Y ∂Hn ∂Hn, n L ∩ ⊂ where the closure is taken in H ; this is a special case of the notion of a Gromov limit set as defined in Definition 6.2 below. Since X is a closed convex subset of Hn, it is a CAT(-1) space and in particular uniquely geodesic. Its Gromov boundary ∂X can thus be identified with geodesic rays emanating from some basepoint o X. Each such geodesic then defines a point in ∂Hn and thus we obtain a continuous embedding ∈ ι : ∂X ∂Hn; it follows from Arzelà–Ascoli Lemma A.16 that the image is given by the limit set ∂X. → More precisely, we can fix a basepoint o X and given ξ (X) we may choose x X with x ξ. ∈ ∈ L n ∈ n → Since X is closed and convex, the geodesic segments [o, xn] subsequentially converge to a geodesic ray in X with endpoint ξ. This shows that (X) ι(∂X), and the converse inclusion is obvious. In L ⊂ particular we have ∂X = (X), ∼ L where, by construction, (X) is a compact subset of the sphere ∂Hn. L We now consider the problem of recovering X from its limit set in Hn.

REMARK 5.27 (Hulls in real hyperbolic spaces). There are various different ways to associate with a compact subset L ∂Hn a set in Hn which has L as limit set: ⊂ In Definition D.27 we define (in large generality) the notion ofa weak hull H(L) Hn of L. In the ⊂ present case, the definition can be made more explicit: If we set L(2) := (ξ, η) L2 ξ = η , then for { ∈ | 6 } 4. EMBEDDINGS OF HYPERBOLIC APPROXIMATE GROUPS 67 all (ξ, η) L(2) there is a unique (up to reparametrization) geodesic γ bi-asymptotic to ξ and η, and ∈ ξ,η by Remark D.28 we then have

H(L)= γξ,η(R), (2) (ξ,η[)∈L i.e. the weak hull is the union of all geodesics connecting points in L. Since L is compact, the weak hull is quasi-convex (Proposition D.30) and if L is non-empty, then L = (H(L)) (Corollary D.33). L The slight disadvantage of the weak hull is that it is only quasi-convex, but typically not convex. We thus define (L) := conv(H(L)); H this is the smallest closed convex subset of Hn which contains all geodesics between points in L and is thus called the convex hull of L. We establish in Proposition D.31 below that (L) is at bounded H Hausdorff distance from H(L); in particular this implies that ( (L)) = (H(L)) = L. (4.1) L H L LEMMA 5.28. Let Y Hn be an unbounded quasi-cobounded closed convex subset. Then ( (Y )) Y ⊂ H L ⊂ and Y is at bounded Hausdorff distance from ( (Y )). H L PROOF. Assume first that x H( (Y )) is contained in the weak hull. Then there exist ξ, η (Y ) ∈ L ∈ L with ξ = η such that x [ξ, η], and hence x ,y Y with x ξ and y η. Since ξ = η 6 ∈ n n ∈ n → n → 6 we have (x y ) D for some D > 0, and since Y is convex it follows from Lemma A.16 that n| n 0 ≤ suitable reparametrizations of the geodesic segments [xn,yn] converge uniformly on compacta to a parametrization of [ξ, η]. Since x [ξ, η] and Y is closed we deduce that x Y . This proves that ∈ ∈ H( (Y )) Y , and since Y is closed and convex we have L ⊂ ( (Y )) Y. H L ⊂ Conversely, let us fix a basepoint o Y . Since Y is convex, it is geodesic, hence visual by Proposition ∈ 5.8. There thus exists D> 0 such that for every x Y there is a geodesic ray γ emanating from o with ∈ dist(x, γ) < D. Since ξ := γ( ) (Y ) we have γ = [o, ξ) ( (Y )) and thus Y N ( ( (Y ))). ∞ ∈ L ∈ H L ⊂ D H L The lemma follows.  In particular, we may apply this to our apogee X and obtain a new apogee ( (X)) for Λ. This H L apogee is also a closed convex subset of Hn, but it has a number of advantages:

REMARK 5.29 (A characterization of the convex hull). If X Hn is an apogee for an infinite ap- ⊂ proximate group (Λ, Λ∞), then so is every closed convex set X′ Hn which is at bounded Hausdorff ⊂ distance from X. All of these apogees have the same limit set L := (X) = (X′) and thus contain L L (L) by Lemma 5.28. In particular, (L) is the unique smallest closed convex subset of Hn with limit H H set L which is an apogee for Λ. To summarize, we have the following refinement of Theorem 5.21:

COROLLARY 5.30 (Bonk–Schramm realization). For every infinite hyperbolic approximate group (Λ, Λ∞) there exist n N and a compact subset L ∂Hn with the following properties: ∈ ⊂ (i) (L) is an apogee for Λ. H (ii) L = ( (L)) = ∂ (L) is a representative of ∂Λ (with the topology inherited from ∂Hn). L H ∼ H (iii) The Λ-quasi-orbits under the left-regular qiqac on (L) are quasi-lattices and the corresponding quasi- H orbit maps are quasi-isometries.  In the sequel we refer to L as in Corollary 5.30 as a Bonk-Schramm realization of ∂Λ. As a first application of Corollary 5.30 we study hyperbolic approximate groups with finite Gromov boundaries:

REMARK 5.31 (Approximate groups with finite Gromov boundaries). Let (Λ, Λ∞) be a hyperbolic approximate group. Assume that some representative L of ∂Λ is finite; then all representatives of ∂Λ are finite with the same cardinality ∂Λ := L . | | | | 68 5. HYPERBOLIC APPROXIMATE GROUPS AND THEIR BOUNDARIES

The case ∂Λ =0 occurs if and only if [Λ] is represented by a bounded space which means that | | int Λ is finite. Assume now that Λ is infinite so that k := ∂Λ > 0. | | Let L ∂Hn be a Bonk–Schramm realization of ∂Λ. Then L = k and [Λ] is represented by ⊂ | | int (L). If k =1, then (L)= , which is impossible, hence we always have k 2. H H ∅ ≥ Since ideal hyperbolic polyhedra in hyperbolic n-space are thin, the hull (L) intersects the com- H plement of any sufficiently large ball in Hn in k components, corresponding to the k elements of L. This shows that the number of ends of L is given by (Λ) = (L) = k. |E∞ | |E | With Corollary 3.17 we deduce that k 2 and hence k = 2. This however means that [Λ] is rep- ≤ int resented by the convex hull of two points in Hn, which is a geodesic line or, in other words, that [Λ]int = [Z]. We have thus established the following proposition.

PROPOSITION 5.32 (Gromov boundaries of hyperbolic approximate groups). For a hyperbolic ap- proximate group (Λ, Λ∞) exactly one of the following three possibilities occurs: (Type 1) Λ is finite and hence ∂Λ = (Λ) =0. | | |E∞ | (Type 2) Λ is quasi-isometric to Z and ∂Λ = (Λ) =2. | | |E∞ | (Type 3) Λ is infinite with ∂Λ = (and hence (Λ) 1, 2, ).  | | ∞ |E∞ |∈{ ∞} 5. Non-elementary hyperbolic approximate groups The purpose of this section is to generalize the notion of a non-elementary hyperbolic group to the setting of approximate groups. The following well-known theorem provides several equivalent characterizations of elementary hyperbolic groups:

THEOREM 5.33 (Characterizations of elementary hyperbolic groups). Let Γ be a hyperbolic group. Then the following conditions are equivalent: (i) Γ is finite or there exists a subgroup Γ′ < Γ of index 2 which fixes a point in ∂Γ. ≤ (ii) Γ is finite or there exists a finite index subgroup Γ′ < Γ which fixes a point in ∂Γ. (iii) Γ is finite or there exists a finite index subgroup Γ′′ < Γ which is infinite cyclic. (iv) ∂Γ 2. | |≤ (v) ∂Γ is finite.

PROOF. The implications (i) = (ii) and (iii) = (iv) = (v) are obvious, and (v) = (iv) is ⇒ ⇒ ⇒ ⇒ a special case of Proposition 5.32. Also, if (iv) holds, then either ∂Γ = and hence Γ is finite, or ∅ ∂Γ = ξ , ξ , in which case (i) holds with Γ′ := Stab (ξ ). Thus the only non-trivial implication is { 1 2} Γ 1 (ii) = (iii), which follows from [GdlH90, Theorem 8.3.30]. ⇒ More precisely, if Γ is finite, then there is nothing to show, otherwise the group Γ′ from (ii) has an infinite point stabilizer in ∂Γ′ = ∂Γ, hence by [GdlH90, Theorem 8.3.30] contains a further finite index subgroup Γ′′ which is infinite cyclic.  Note that passing to an index 2 subgroup may be necessary in condition (i), as the example of an infinite dihedral group shows.

DEFINITION 5.34. A hyperbolic group Γ is called elementary if it satisfies the equivalent conditions of Theorem 5.33, otherwise it is called non-elementary.

NOTATION 5.35. For the remainder of this section, (Λ, Λ∞) denotes a hyperbolic approximate group. If Λ is infinite, then we will always fix a Bonk-Schramm realization L ∂Hn of the Gro- ⊂ mov boundary of ∂Λ so that, by Corollary 5.30, the space X := (L) Hn is an apogee for Λ H ⊂ and L represents ∂Λ. We then denote by ρ : (Λ, Λ∞) QI(X) a left-regular quasi-action and by → ∂ρ : (Λ, Λ∞) Homeo(L) the associated boundary action. → f Our next goal is to define a notion of a non-elementary hyperbolic approximate group. All of the properties in Theorem 5.33 have counterparts in the theory of hyperbolic approximate groups, but the 5. NON-ELEMENTARY HYPERBOLIC APPROXIMATE GROUPS 69 non-trivial implication (ii) = (iii) breaks down. Not only is the iteration argument behind the proof ⇒ of [GdlH90, Theorem 8.3.30] not available in the context of approximate groups, but it turns out that the condition (iii) is actually strictly stronger than the other conditions:

EXAMPLE 5.36. If Λ is a uniform approximate lattice in R, then Λ is quasi-isometric to R and hence ∂Γ = 2. In particular, Λ is infinite, but it need not contain any infinite cyclic subgroup (let alone of | | finite index). For example, the set of vertices of the Fibonacci tiling does not. Other than that, the proof of Theorem 5.33 carries over almost verbatim to give the following corollary:

COROLLARY 5.37. Given a hyperbolic approximate group (Λ, Λ∞) we consider the following conditions: (i) Λ is finite or there exists a finite index approximate subgroup Λ′′ Λ which is an infinite cyclic group. ⊂ (ii) ∂Λ 2. | |≤ (iii) ∂Λ is finite. (iv) Λ is finite or there exists a finite index approximate subgroup Λ′ Λ2 such that (Λ′)∞ fixes a point in ⊂ ∂Λ. (v) Λ is finite or there exists an approximate subgroup Λ′ Λ∞ which is commensurable to Λ and such that ⊂ (Λ′)∞ fixes a point in ∂Λ. Then the implications (i) = (ii) (iii) = (iv) = (v) hold. ⇒ ⇐⇒ ⇒ ⇒ PROOF. (i) = (ii) (iii) are immediate from Proposition 5.32 and (iv) = (v) is obvious, ⇒ ⇐⇒ ⇒ hence it remains to prove (ii) = (iv). ⇒ If ∂Λ= then Λ is finite, hence (iv) holds. Otherwise we have ∂Λ= ξ , ξ and the stabilizer of ∅ { 1 2} ξ defines a finite index subgroup Γ of Λ∞. We may thus choose Λ′ := Λ2 Γ; this stabilizes ξ and is 1 ∩ 1 an approximate subgroup of Λ2 by Lemma 2.26.  We define the strongest possible definition of non-elementary hyperbolic group by negating the weakest of the conditions in Corollary 5.37:

DEFINITION 5.38. Let (Λ, Λ∞) be a hyperbolic approximate subgroup. (i) We say that Λ has fixpoints at infinity if there exists some ξ L such that ∂ρ(λ)(ξ)= ξ for all λ Λ ∈ ∈ (and hence all λ Λ∞). ∈ (ii) We say that Λ is non-elementary if no Λ′ Λ∞ which is commensurable to Λ has fixpoints at ⊂ infinity. By Corollary 5.37 every non-elementary hyperbolic approximate group has an infinite boundary. We do not know whether the converse is true. It is a classical fact in the theory of hyperbolic groups (see e.g. [GdlH90, Cor. 8.2.26]) that the action of a non-elementary hyperbolic group on its boundary is minimal; here an action of a group on a metric space (X, d) is called minimal if every orbit is dense. Similarly we define:

DEFINITION 5.39. An action ρ : (Λ, Λ∞) Homeo(X) of an approximate group (Λ, Λ∞) on a → topological space X is minimal if every ρ(Λ)-orbit in X is dense. Note that minimality of an action (Λ, Λ∞) y X is a priori stronger than minimality of Λ∞ y X. Fixpoints are an obvious obstruction to minimality (if the space has at least two points), and in the case of the boundary action of a hyperbolic approximate groups they turn out to be the only obstruction:

PROPOSITION 5.40 (Boundary minimality). For a hyperbolic approximate group (Λ, Λ∞), the following are equivalent: (i) (Λ, Λ∞) has no fixpoints at infinity. (ii) The action of (Λ, Λ∞) on some (hence any) representative of ∂Λ is minimal. (iii) The action of Λ∞ on some (hence any) representative of ∂Λ is minimal. In particular, these hold if (Λ, Λ∞) is non-elementary. 70 5. HYPERBOLIC APPROXIMATE GROUPS AND THEIR BOUNDARIES

PROOF. Assume that (Λ, Λ∞) has no fixpoints at infinity and let ξ L; then the orbit closure ∈ L′ := ρ(Λ).ξ is a compact subset of ∂Hn which contains at least one other point than ξ, say η. Now the weak hull H(L′) contains the geodesic [η, ξ] and thus H(L′) = . ′ 6 ′ ∅ It follows from Lemma 6.27 that dHaus(ρ(Λ).H(L ), H(L )) < D for some D > 0. On the other hand, since ρ is cobounded, we know that there exists an R 0 such that for any z H(L′) we have ≥ ′ ∈ dHaus(ρ(Λ).z, H(L)) < R. Setting T := R + D, we conclude dHaus(H(L ), H(L)) < T . Passing to the respective limit sets we deduce with Proposition D.32 that L′ = (H(L′)) = (H(L)) = L. L L This shows that (Λ, Λ∞) acts minimally on L (hence any representative of ∂Λ) and proves the impli- cation (i) = (ii). The implication (ii) = (iii) = (i) are obvious.  ⇒ ⇒ ⇒ Boundary minimality is a useful tool in the investigation of the topology of ∂Λ, as illustrated by the following result.

PROPOSITION 5.41. Assume that (Λ, Λ∞) is an infinite hyperbolic approximate group with Bonk-Schramm realization L ∂Hn which has no fixpoints at infinity. Then precisely one of the following two options holds: ⊂ (i) L =2. | | (ii) L is a perfect compact subset of ∂Hn. In particular, the boundary of a non-elementary hyperbolic approximate group is a perfect compact subset of a sphere.

PROOF. Note first that L is compact as a closed subset of ∂Hn. Now assume that L is not perfect and pick an isolated point ξ L. Since ∂ρ(Λ∞) acts by homeomorphisms on L, it follows that every ∈ point in ∂ρ(Λ∞).ξ is an isolated point of L. This implies that ρ(Λ).ξ is a discrete subset of the compact space L, hence finite. We deduce with Proposition 5.40 that L itself is finite. By Proposition 5.32 we thus have either L =2 or L =0, but since Λ is infinite, the latter is impossible. This shows that L is | | | | either perfect and compact or of cardinality 2, and the proposition follows.  We now collect everything that we have learned so far about the topology of boundaries of non- elementary hyperbolic approximate groups in the following corollary:

COROLLARY 5.42 (Topology of boundaries). If (Λ, Λ∞) is a non-elementary hyperbolic group and L is a representative for ∂Λ, then the following hold: (i) L is a perfect compact space whose homeomorphism group acts minimally. (ii) L embeds homeomorphically into a finite-dimensional sphere, in particular L is of finite topological dimen- sion. (iii) There exists a metric d on L which generates the topology and such that (L, d) is locally quasi-self-similar and doubling.  CHAPTER 6

Limit sets of quasi-isometric quasi-actions

In this chapter we consider the behavior of quasi-isometric quasi actions (qiqacs) of approximate groups at infinity. To this end we first explain how to extend a qiqac of a given approximate group on a proper geodesic metric space to an action by homeomorphisms on a suitably defined boundary of the space. Inside the boundary we then find an invariant subset, called the limit set of the qiqac, which provides important geometric information about the qiqac at hand.

1. Morse and Gromov limit sets of approximate groups In this section we define the boundary action of a qiqac and the associated limit set. Throughout we will work in the following setting: NOTATION 6.1. The following notations and conventions will be used throughout this chapter: (Λ, Λ∞) denotes an approximate group; • X is a proper geodesic metric space with basepoint o X; • ∈ ρ: (Λ, Λ∞) QI(X) is a (K , C , C′ )-qiqac. • → k k k Our basic constructions apply to a wide variety of boundaries. For example, if X is Gromov hyperbolic, then we canf define a limit set in the Gromov boundary of X. If X is an arbitrary proper geodesic metric space, then we can define a limit set in the Morse boundary of X (see Appendix D for the definition of Morse boundaries). In fact, one could as well define limit sets in other type of boundaries - rather than trying to axiomatize the situation, we will discuss limit sets in Gromov and Morse boundaries and leave it to the reader to adapt the constructions presented here to their context. For the first part of this section, up to Proposition 6.8, we will assume that X is Gromov hyperbolic. We then denote by ∂X := ∂ X the sequential model of the Gromov boundary of X and by X = X ∂X s ∪ the Gromov compactification. Note that ∂X is always compact, and it is non-empty as soon as X is unbounded. DEFINITION 6.2. Given a subset Y X we define its Gromov limit set as (Y ) := Y ∂X, where ⊂ L ∩ the closure is taken inside X. Note that if two subsets Y,Z X are of bounded Hausdorff distance, then (Y ) = (Z), which ⊂ L L implies: LEMMA 6.3. For all o, o′ X and k N we have ∈ ∈ (ρ(Λ).o)= (ρ(Λ).o′)= (ρ(Λk).o) ∂X. L L L ⊂ PROOF. The first equality follows from the fact that, by Remark 4.13, any two Λ-quasi-orbits are at bounded Hausdorff distance. Similarly, the quasi-orbits of Λ and Λk are of bounded Hausdorff distance since Λ Λk is of finite index by Proposition 2.13.  ⊂ DEFINITION 6.4. The set (Λ) := (ρ(Λ).o) is called the Gromov limit set of the qiqac ρ. Lρ L By Lemma 6.3 the Gromov limit set of ρ is independent of the basepoint o. CONSTRUCTION 6.5 (Boundary action of a qiqac). Recall that every quasi-isometry f : X X → induces a homeomorphism ∂f : ∂X ∂X. Explicitly, if (x ) is a sequence in X which represents → n ξ ∂X, then the sequence (f(x )) represents f(ξ). In particular, we may define a map ∈ n ∂ρ:Λ∞ Homeo(∂X), ∂ρ(λ) := ∂(ρ(λ)). → 71 72 6. LIMIT SETS OF QUASI-ISOMETRIC QUASI-ACTIONS

LEMMA 6.6 (Boundary action). The map ∂ρ is a group homomorphism and hence defines an action of (Λ, Λ∞) on ∂X by homeomorphisms.

PROOF. Let ξ = [(x )] ∂X and µ, λ Λ∞. Since ρ is a qiqac, there exists a constant R (depending n ∈ ∈ on λ and µ) such that for all x X we have d(ρ(λ)(ρ(µ)(x)),ρ(λµ)(x)) < R. We deduce that ∈ ∂ρ(λ)(∂ρ(µ)(ξ)) = [(ρ(λ)(ρ(µ)(xn))] = [(ρ(λµ)(xn))] = ∂ρ(λµ)(ξ), and hence ∂ρ(λ) ∂ρ(µ)= ∂ρ(λµ).  ◦ DEFINITION 6.7. ∂ρ : (Λ, Λ∞) Homeo(∂X) is called the boundary action of the qiqac ρ. → The following crucial observation follows essentially from Lemma 6.3:

PROPOSITION 6.8 (Invariance of the limit set). The limit set (Λ) is invariant under the boundary Lρ action ∂ρ, i.e., ∂ρ(Λ∞)( (Λ)) (Λ). Lρ ⊂ Lρ PROOF. We fix a basepoint o X so that (Λ) = (ρ(Λ).o). Given µ Λ∞ we can choose k N ∈ Lρ L ∈ ∈ such that µ Λk. Since ρ is a qiqac, a similar argument as in the proof of Lemma 6.6 shows that ∈ ρ(µ)(ρ(Λ).o) is at bounded Hausdorff distance from ρ(Λk+1).o. Since the quasi-orbits of Λ and Λk+1 are at bounded Hausdorff distance, the result follows.  For every qiqac of (Λ, Λ∞) on a Gromov hyperbolic space X we have thus managed to construct an action of (Λ, Λ∞) by homeomorphism on the corresponding Gromov limit set. This construction is functorial in the following sense: Assume that ρ′ is another qiqac of (Λ, Λ∞) on a proper geodesic hyperbolic space Y and that ϕ : X Y is a quasi-isometry with quasi-inverse ψ and that ρ′ is → equivalent to ϕ ρ ψ. In this case, ϕ(ρ(Λ).o) is at bounded Hausdorff distance from ρ′(Λ).ϕ(o), ◦ ◦ and hence ∂ϕ maps (Λ) homeomorphically onto ′ (Λ). We then obtain the following commuting Lρ Lρ diagram: ∂ϕ ∂X ∂Y

∂ϕ|Lρ(Λ) (Λ) ′ (Λ) Lρ Lρ We now want to establish similar results for qiqacs on arbitrary proper geodesic metric spaces, which are not necessarily Gromov hyperbolic. A convenient choice of boundary for this purpose is the so-called Morse boundary; we refer the reader to Appendix D for background concerning this boundary. From now on, (X, d) is allowed to be an arbitrary proper geodesic metric space and o X denotes M ∈ a fixed choice of basepoint. We denote by ∂s Xo the sequential model of the Morse boundary of X with basepoint o (see Remark D.5 for the definition). Since a change of basepoint induces a homeo- M morphism on the boundary, we will usually suppress the basepoint in the notation and write ∂s X M for ∂s Xo. M REMARK 6.9 (On the scope of our setting). If X is Gromov hyperbolic, then ∂s X = ∂sX is just (the sequential model of) the Gromov boundary of X, and hence the results about Morse limit sets below reduce to the results concerning Gromov limit sets above. Assume now that X is not Gromov hyperbolic; then one of two cases will happen: If there are M no Morse geodesics in X, then ∂s Xo is empty. This happens for example if every geodesic in X bounds a half-flat, in particular in higher rank buildings and symmetric spaces. If there are Morse M geodesics, then ∂s Xo is actually non-compact [Liu19, Theorem 5.7]. By [Sis16] this is the case if X is quasi-isometric to an acylindrically hyperbolic group (for example a relatively hyperbolic group, a mapping class group, Out(Fn) or a non-directly decomposable right angled Artin group, cf.[Osi16]), but there are examples outside this class, including finitely-generated groups such as Tarski monsters and infinite torsion groups [OOS09, Fin17]. 1.MORSEANDGROMOVLIMITSETSOFAPPROXIMATEGROUPS 73

In our more general setting, there is no canonical topology on X ∂M X, and hence we cannot ∪ s define the limit set of a subset Y X as in the hyperbolic case. Instead, following Definition D.20, we ⊂M define the Morse limit set of Y in ∂s X as the closed subset given by (Y )= ξ ∂M X Morse gauge N and (y ) X(N) Y such that lim y = ξ . L ∈ s ∃ n ⊂ o ∩ n n o However, if X is hyperbolic, then ∂M X = ∂ X is simply (the sequential model of) the Gromov bound- s s ary of X, and in this case the Morse limit set coincides with the Gromov limit set. In this sense, the Morse limit set generalizes the Gromov limit set. As in the hyperbolic case, we observe that if two subsets Y,Z X are of bounded Hausdorff ⊂ distance, then (Y )= (Z). The proof of Lemma 6.3 thus yields: L L LEMMA 6.10. For all o, o′ X and k N we have ∈ ∈ (ρ(Λ).o)= (ρ(Λ).o′)= (ρ(Λk).o) ∂M X.  L L L ⊂ s DEFINITION 6.11. The set (Λ) := (ρ(Λ).o) is called the (Morse) limit set of the qiqac ρ. Lρ L Note that by Lemma 6.10, the limit set of ρ is independent of the basepoint o. As in the hyperbolic case, the qiqac ρ also gives rise to a boundary action of Λ∞ on the Morse boundary, which preserves the limit set, but the proof is slightly more involved: Firstly, by Corollary D.10 every quasi-isometry f : X X induces a homeomorphism ∂M f : ∂M X ∂M X, and hence → s s → s we may define a map: ∂ ρ:Λ∞ Homeo(∂M X), ∂ ρ(λ) := ∂ (ρ(λ)). s → s s s The following two lemmas show that this defines an action of Λ∞ which preserves the limit set; the proofs are based on the same ideas as the proofs of Lemma 6.6 and Lemma 6.8; the only additional ingredient is the fact that quasi-isometries preserve Morse geodesics.

LEMMA 6.12 (Boundary action). The map ∂sρ is a group homomorphism and hence defines an action of ∞ M (Λ, Λ ) on ∂s X by homeomorphisms.

M ∞ (N) PROOF. Let ξ = [(x )] ∂ X and λ, µ Λ . We choose a Morse gauge N such that (x ) Xo . n ∈ s ∈ n ⊂ Since ρ is a qiqac we can argue as in the proof of Lemma 6.6 to find a constant R (depending on λ and µ) such that for all x X we have ∈ d(ρ(λ)(ρ(µ)(x)),ρ(λµ)(x)) < R. (1.1)

To finish the proof we need to check that [(ρ(λ)(ρ(µ)(xn)))] is equivalent to [(ρ(λµ)(xn))] in the Morse boundary. In view of (1.1) and the definition of the Morse boundary as a direct limit, this amounts ′ to showing that there is a Morse gauge N such that (ρ(λ)(ρ(µ)(xn))) and (ρ(λµ)(xn)) are subsets of ′ (N ) Xo . However, this follows from the fact that Morse geodesics are robust under quasi-isometries and change of basepoint. 

PROPOSITION 6.13 (Invariance of the limit set). The limit set (Λ) is invariant under the boundary Lρ action ∂sρ, i.e., ∂ ρ(Λ∞)( (Λ)) (Λ). s Lρ ⊂ Lρ ∞ PROOF. Let µ Λ and ξ = [(xn)] ρ(Λ). We can then choose k N and a Morse gauge k∈ ∈ L (N) ∈ N such that µ Λ and (xn) is contained in Xo ρ(Λ).o. Since Morse geodesics are robust under ∈ ∩ ′ quasi-isometry and change of basepoint, there exists a Morse gauge N so that the sequence (ρ(µ)(xn)) ′ (N ) is contained in Xo and represents ∂sρ(µ)(ξ). Since ρ is a qiqac, the same argument as in the proof of Lemma 6.12 now shows that this sequence is at bounded Hausdorff distance from ρ(Λk+1).o, hence the result follows. 

DEFINITION 6.14. ∂ ρ : (Λ, Λ∞) Homeo(∂ (X)) is called the boundary action of the qiqac ρ. s → s 74 6. LIMIT SETS OF QUASI-ISOMETRIC QUASI-ACTIONS

As in the hyperbolic case, we have the following functoriality result: If Y is another proper geo- desic metric space and ϕ : X Y is a quasi-isometry with quasi-inverse ψ, then the qiqac ρ induces → a qiqac ρ′ on Y by ρ′(λ) := ϕ ρ(λ) ψ. As in the hyperbolic case we get that ϕ(ρ(Λ).o) is at bounded ◦ ◦ Hausdorff distance from ρ′(Λ).ϕ(o) and we obtain:

LEMMA 6.15 (Functoriality). In the situation above, the following diagram Λ∞-equivariantly commutes:

M ∂sϕ M ∂s X ∂s Y

∂sϕ|Lρ(Λ) (Λ) ′ (Λ) Lρ Lρ In particular, ∂ ϕ maps (Λ) homeomorphically onto ′ (Λ).  s Lρ Lρ As a special case we observe that the homeomorphism type of the limit set is invariant under quasi-conjugacies. We conclude this section by discussing qiqacs ρ which have full limit set in the sense that (Λ) = ∂M X. An obvious class of examples is given by cobounded qiqacs: Lρ s EXAMPLE 6.16 (Limit sets of cobounded qiqacs). Assume that the qiqac ρ is cobounded. In this case the quasi-orbit ρ(Λ).o is at bounded Hausdorff distance from X and hence (Λ) = (ρ(Λ).o)= (X)= ∂M X. Lρ L L s In particular, every cobounded action has full limit set. However, the limit set can also be full without the action being cobounded (even in the Gromov hyperbolic case).

EXAMPLE 6.17 (A small action with full limit set). Let Γ be a non-elementary hyperbolic group with finite generating set S, and the word metric dS. It is known that in this case the action of Γ on ∞ the Gromov boundary ∂sΓ of (Γ, dS ) is minimal ([Gro87]), i.e. every orbit is dense. Now let (Λ, Λ ) be an infinite approximate group with Λ∞ =Γ and consider the natural isometric action ρ of (Λ, Λ∞) ∞ on (Λ , dS ) given by left-multiplication. Since Λ is infinite, the Gromov limit set ρ(Λ) ∂sΓ is non- ∞ L ⊂ empty. Since it is also closed and Λ -invariant, and since ∂sΓ is minimal, we deduce that ρ(Λ) = ∞ L ∂sΓ, i.e. ρ has full limit set. On the other hand, if Λ does not have finite index in Λ , then ρ is not cobounded.

2. Conical limit points We keep the notation of the last section (see Notation 6.1). We will be concerned with a special kind of limit points, defined as follows:

DEFINITION 6.18. For Y X, a limit point ξ (Y ) is called a conical limit point of Y if for all ⊂ ∈ L geodesic rays γ representing ξ there exists K > 0 and a strictly increasing sequence k N such that n ∈ B(γ(k ),K) Y = . n ∩ 6 ∅ The set of all conical limit points of Y is denoted con(Y ) and called the conical limit set of Y . L In particular, given a qiqac ρ of (Λ, Λ∞) on X, a point ξ (Λ) is called a conical limit point of ρ if ∈ Lρ ξ is a conical limit point of Λ.o for some o X. The conical limit set of ρ is ∈ con(Λ) := ξ (Λ) ξ is a conical limit point . Lρ { ∈ Lρ | } It is immediate from the definition that if Y,Z X are two subspaces at bounded Hausdorff ⊂ distance, then (Y )= (Z). This implies that con(Λ) is independent of the choice of basepoint o and L L Lρ that con(Λ) = con(Λn) for all n N. From the latter property we deduce, by the same argument as Lρ Lρ ∈ in the case of , the following invariance property of the conical limit set: Lρ PROPOSITION 6.19 (Invariance of the conical limit set). The conical limit set con(Λ) is a ∂ρ(Λ∞)- Lρ invariant subset of (Λ).  Lρ 2. CONICAL LIMIT POINTS 75

Although there are many situations of interest in which every limit point is conical (see in partic- ular Proposition 6.34 below), in general the conical limit set can be strictly smaller than the full limit set; the following proposition provides a large class of examples for this phenomenon. Concerning the statement of the proposition we observe that by Corollary 2.38 the quasikernel of a symmetric quasi- morphism is an approximate subgroup; this approximate subgroup then acts on the ambient group, and if the ambient group is finitely-generated then the action is isometric with respect to any choice of word metric.

PROPOSITION 6.20 (Conical limit sets of quasikernels). Let Γ be a non-elementary hyperbolic group with Gromov boundary ∂Γ and let f :Γ Z be a surjective symmetric quasimorphism. Assume that qker(f) → generates Γ. Then (qker(f)) = ∂Γ, where the conical limit set of qker(f) in ∂Γ is given by L con(qker(f)) = ξ ∂Γ f is bounded along every geodesic γ representing ξ . L { ∈ | } PROOF. The first statement is just a special case of Example 6.17. For the proof of the seconds statement we fix a finite generating set S for Γ and realize ∂Γ as the boundary of the Cayley graph of Γ with respect to S. Let ξ ∂Γ and let γ be a geodesic representing ξ. Assume first that ξ con(qker(f)). Then there ∈ ∈ L exist an integer K > 0 and a sequence n and elements x B(γ(n ),K) qker(f), where k → ∞ k ∈ k ∩ the balls are taken with respect to the word metric d = d . We then have f(x ) d(f), where d(f) S | k | ≤ denotes the defect of f, and since γ(n )= x s s for some s S S−1 e we obtain k k 1 ··· K j ∈ ∪ ∪{ }

lim f(γ(t)) lim f(γ(nk)) lim f(xk) + K d(f)+ max f(s) < . | |≤ | |≤ | | s∈S∪S−1∪{e} | | ∞ t→∞ k→∞ k→∞    Conversely, assume that lim f(γ(t)) = . We claim that ξ con(qker(f)). Assume otherwise t→∞ | | ∞ 6∈ L for contradiction; then there exist a natural number K > 0 and a sequence n and elements k ր ∞ x qker(f) such that x qker(f) and d(γ(n ), x ) K. Since lim f(γ(t)) = we have k ∈ k ∈ k k ≤ t→∞ | | ∞ f(γ(n )) . Then as above one computes | k | → ∞

f(xk) f(γnk ) K d(f)+ max f(s) , | | ≥ | |− s∈S∪S−1∪{e} | |   which is unbounded. This contradiction finishes the proof of the second statement.  Note that in the situation of Proposition 6.20 the conical limit set is dense in the full limit set, since the boundary action of a non-elementary hyperbolic group is minimal. Nevertheless, in many cases of interest it is a proper subset of the full limit set, and this additional information can be exploited. In this regard it is instructive to consider the case of homogeneous counting quasimorphisms in free groups (see Example B.47).

EXAMPLE 6.21 (A class of homogeneous counting quasimorphisms). We consider the situation of Notation B.50: F is a non-abelian free group with free generating set S := a ,...,a and u W is r { 1 r} ∈ r a cyclically reduced non-self-overlapping word of length u 2. Moreover, we abbreviate by k kS ≥ Λ := qker(ϕcyc)= g F ϕcyc(g) d(ϕcyc) u u { ∈ r | | u |≤ u } the quasi-kernel of the associated cyclic counting morphism ϕcyc : F Z. By Remark B.51, ϕcyc u r → u is the homogenization of the counting quasimorphism ϕu, in particular it is conjugation-invariant, symmetric, and surjective, hence Λu is an approximate subgroup of Fr by Corollary 2.38. Since u is non-self-overlapping and of length 2, its initial and final letter are distinct, and hence ≥ ϕ (s)=0 for all s S. This shows, firstly, that ϕ is not a homomorphism (since ϕ = 0 and u ∈ u u|S ϕ (u)=1) and hence d(ϕ ) 1 and, secondly, that S Λ , and hence Λ∞ = F . u u ≥ ⊂ u u r The action of Fr on its Cayley tree T2r with respect to S now induces an isometric action of (Λ , Λ∞) on T . Since T is hyperbolic, we can consider the limit set (Λ ) ∂T as well as the u u 2r 2r L u ⊂ 2r corresponding conical limit set con(Λ ) ∂T . L u ⊂ 2r We are going to show: 76 6. LIMIT SETS OF QUASI-ISOMETRIC QUASI-ACTIONS

THEOREM 6.22 (Recovering quasikernels from their conical limit sets). Let Fr be a free group of rank r 3 with free generating set S and let u, v be two cyclically reduced, non-self-overlapping words in F of ≥ r lengths u , v 2. Then the following hold: | |S | |S ≥ (i) (Λ )= (Λ )= ∂T . L u L v 2r (ii) con(Λ )= con(Λ ) if and only if Λ =Λ if and only if u = v±1. L u L v u v In particular, cyclically reduced non-self-overlapping words of length 2 in F are uniquely determined by the ≥ r conical limit sets of the associated cyclic counting quasimorphisms, whereas the full limit set is independent of the chosen word.

REMARK 6.23 (Concerning the proof of Theorem 6.22). Part (i) of Theorem 6.22 is just a special case of Example 6.17. Concerning (ii) we have the obvious implications

u = v±1 = Λ =Λ = con(Λ )= con(Λ ). ⇒ u v ⇒ L u L v The non-trivial implication con(Λ ) = con(Λ ) = u = v±1 is based on three ingredients: our L u L v ⇒ starting point is Proposition 6.20, which allows us to identify the conical limit sets in question. In addition, we will need a number of rather specific combinatorial arguments concerning free groups (summarized in Lemma B.39 and Lemma B.40 below) and a reconstruction theorem for quasimor- phisms (Corollary B.59) due to Gabi Ben Simon and the second author.

PROOF OF THEOREM 6.22. Let u, v be as in the theorem and assume that con(Λ ) = con(Λ ). L u L v We then have to show that v = u or v = u−1. Throughout the proof we are going to abbreviate := con(Λ ) = con(Λ ). We also set N := w F ϕcyc(w) = 0 and define an equivalence L L u L v u { ∈ r | u 6 } relation σ on N by wσ w′ : ϕcyc(w)ϕcyc(w′) > 0. The set N and equivalence relation σ will u u u ⇐⇒ u u v v be defined accordingly. We make the following two claims: Claim 1: Nu = Nv. Claim 2: σu = σv. Assuming the two claims for the moment, we complete the proof of the theorem: By the first claim we have Nu = Nv and we denote this set by N. Moreover, the equivalence classes of σu are given by cyc cyc cyc Pos(ϕu ) = w Fr ϕu (w) > 0 and Pos( ϕu )), and similarly the equivalence classes of σv are { cyc∈ | cyc } − cyc cyc given by Pos(ϕv ) and Pos( ϕv ). We thus deduce from Claim 2 that either Pos(ϕv ) = Pos(ϕu ) or cyc cyc − cyc Pos(ϕv ) = Pos( ϕu ). It then follows from the reconstruction theorem (Corollary B.59) that ϕu and cyc − −1 ϕv are linearly dependent. However, according to [HT18] this is only possible if v = u or v = u . It thus remains to establish the two claims. We first prove Claim 1. Define N ′ := w F ϕcyc(w) = 0 and w is cyclically reduced and u { ∈ r | u 6 } define N ′ accordingly. By conjugacy-invariance of cyclic counting quasimorphisms we have w N v ∈ u (respectively w N ) if and only if some conjugate of w is contained in N ′ (respectively, N ′ ), hence ∈ v u v it suffices to show that N ′ = N ′ . For this let w F be cyclically reduced and non-trivial. Then there u v ∈ r exists a unique geodesic ray w∞ in T which emanates from e and contains wn for all n N, and 2r ∈ we denote by ξ ∂Γ its endpoint. If w N ′ , then ϕcyc(wn) = n ϕcyc(w) and thus ξ by ∈ ∈ u | u | | u | → ∞ 6∈ L Proposition 6.20. If w N ′ , then f(wn)=0 for all n N and hence f is bounded along w∞, which by 6∈ u ∈ Proposition 6.20 implies ξ . Repeating the same argument for v instead of u then yields ∈ L w N ′ ξ w N ′ , ∈ u ⇐⇒ 6∈ L ⇐⇒ ∈ v ′ ′ and thus Nu = Nv, which proves the claim. The proof of Claim 2 is similar, but combinatorially more involved. Since we already know that N = N , we will denote this set simply by N. We also set C := max C , C , where C and C are u v 0 { u v} u v defined as in Lemma B.40; we then define

3 C1 := max sup f(x) + d(f)+1 . f∈{ϕcyc,ϕcyc} u v x∈B(e,C0) | | 2 ! 2. CONICAL LIMIT POINTS 77

We first observe that xσ y if and only if x′σ y′ for some (hence all) x′,y′ F which are conjugate to u u ∈ r some positive power of x and y respectively; the same holds for σv instead of σu. We now define N ′′ := w N w cyclically reduced, w > C , min ϕcyc(w) , ϕcyc(w) > C . { ∈ | k kS 0 {| u | | v |} 1} Since w N if and only if some conjugate of a positive power of w lies in N ′′, it suffices to show that ∈ ′′ ′′ cyc cyc σu and σv agree on N . Thus let x, y N and assume that x✚σuy; thus either ϕu (x) > 0 > ϕu (y) or cyc cyc ∈ ϕu (y) > 0 > ϕu (x). cyc cyc cyc We first consider the case in which ϕu (x) > 0 > ϕu (y). Since ϕu is homogeneous and takes integer values we can then find a,b N such that ∈ ϕcyc(xa)= ϕcyc(yb) > 0. (2.1) u − u cyc a cyc b a b This implies in particular that #u (x ) > 0 and #u−1 (y ) > 0. Moreover, x and y are cyclically reduced and xa > u and yb > u−1 . It then follows from Lemma B.39 (applied once to k kS k kS k kS k kS the pair (xa,u) and once to the pair (yb,u−1) that there exist x′,y′ F which are cyclically equivalent ∈ r to xa, respectively yb such that

cyc ′ ′ cyc ′ ′ ϕu (x )= ϕu(x ) and ϕu (y )= ϕu(y ). (2.2) In particular, x′ is conjugate to xa and y′ is conjugate to yb and min x′ , y′ u . By Lemma {k kS k kS}≥k k B.40 we now find w,z F with the following properties: ∈ r w,z B(e, C ) B(e, C ); • ∈ u ⊂ 0 the product x′wy′z is reduced and cyclically reduced; • every copy of u or u−1 in (x′wy′z)n is contained in one of the n copies of x′ or in one of the n • copies of y′. Combining these properties with (2.1) and (2.2) we deduce that for all n N we have ∈ ′ ′ n ′ ′ cyc ′ cyc ′ cyc a cyc b ϕu((x wy z) )= nϕu(x )+ nϕu(y )= nϕu (x )+ nϕu (y )= nϕu (x )+ nϕu (y )=0. cyc ′ ′ This implies that ϕu (x wy z)=0 and as in the proof of Claim 1 one sees that if ξ is the endpoint of the geodesic (x′wy′z)∞, then ξ . We have thus proved that if x✚σ y and ϕcyc(x) > 0 > ϕcyc(y), then ∈ L u u u ( ) there exist powers x′ and y′ of conjugates of x and y respectively and words w,z of length C such ∗ ≤ 0 that x′wy′z is reduced and cyclically reduced and its powers converge to an element ξ . ∈ L If instead we have ϕcyc(y) > 0 > ϕcyc(x), then ( ) holds after reversing the roles of x and y. We u u ∗ now claim that, conversely, if Condition ( ) holds for either the pair (x, y) or the pair (y, x), then x✚σ y. ∗ u Since the relation σ is symmetric, we may assume that ( ) holds for (x, y) (N ′′)2. Also assume for u ∗ ∈ contradiction that xσuy. cyc cyc cyc ′ cyc ′ Since ϕu (x) and ϕu (y) have the same sign, the same is true for ϕu (x ) and ϕu (y ), and since ϕcyc(x) and ϕcyc(y) are larger than C , the same is true for ϕcyc(x′) and ϕcyc(y′). With the notation | u | | u | 1 | u | u as in ( ) we then have ∗ ϕcyc(x′wy′z) ϕcyc(x′)+ ϕcyc(y′) ϕcyc(w) ϕcyc(z) 3d(ϕcyc) | u | ≥ | u u | − | u | − | u |− u = ϕcyc(x′) + ϕcyc(y′) ϕcyc(w) ϕcyc(z) 3d(ϕcyc) | u | | u | − | u | − | u |− u cyc cyc 2C1 2 sup ϕu (x) 3d(ϕu ) > 0 ≥ − x∈B(e,C0) | |− by the very definition of C . As in the proof of Claim 1 this implies that ξ , which is a contra- 1 6∈ L diction. The same arguments can also applied to v instead of u. We thus obtain for all x, y N ′′ the ∈ equivalences xσ y The Condition ( ) holds neither for (x, y) nor for (y, x) u ⇐⇒ ∗ xσ y. ⇐⇒ v This finishes the proof.  78 6. LIMIT SETS OF QUASI-ISOMETRIC QUASI-ACTIONS

3. Qiqacs with stable quasi-orbits We keep the notation of the previous sections (see Notation 6.1). We observe that if the quasi-orbit ρ(Λ).o is quasi-convex in X, then every quasi-orbit of X is quasi-convex in X; we then simply say that ρ has quasi-convex quasi-orbits. For proper qiqacs on hyperbolic spaces having quasi-convex quasi-orbits is a rather strong condition with strong consequences for the limit set:

PROPOSITION 6.24 (Consequences of quasi-convex quasi-orbits in the hyperbolic case). If X is Gromov hyperbolic, (Λ, Λ∞) is geometrically finitely-generated and ρ is proper and has quasi-convex quasi- orbits, then the following hold: (i) := ρ(Λ).o [Λ] . O ∈ int (ii) (Λ, Λ∞) is a hyperbolic approximate group. (iii) If A is an apogee for Λ, then there is a quasi-isometric embedding ι : A X whose image is given by . → O .(iv) Any map ι as in (iii) induces a continuous embedding ∂ι : ∂A ֒ ∂X whose image is given by (Λ) → Lρ In particular (Λ) is a representative of ∂Λ. Lρ PROOF. (i) We first note that since X is quasi-convex, it is quasi-geodesic by Remark A.21, O ⊂ and hence large-scale geodesic by Lemma A.28. We also note that is quasi-invariant under ρ(Λ) and O denote by ρ the quasi-restriction of ρ to (cf. Construction 4.19). O Since ρ is proper, this quasi-restriction is proper as well, and by definition it is cobounded, hence geometric.e Since Λ is geometrically finitely-generated and is large-scale geodesic we may thus apply O the Milnor–Schwarz (Theorem 4.37) to conclude that [Λ] . O ∈ int (ii) follows from (i) and Corollary C.9. (iii) follows from (i). (iv) follows from (i) and Corollary D.25.  If we remove the assumption that X be Gromov hyperbolic, then having quasi-convex orbits by itself will no longer allow one to draw any similar conclusions. However, we can still establish a version of Proposition 6.24 if we replace quasi-convexity by the stronger condition of stability: Recall from Definition D.22 that a quasi-convex subset Y X is called N-stable (with respect to some Morse ⊂ gauge N) provided every pair of points in Y can be connected by an N-Morse geodesic in X. We say that the qiqac ρ has stable quasi-orbits if some quasi-orbit is N-stable for some Morse gauge N (and hence every quasi-orbit is stable with a Morse gauge which may depend on the given quasi-orbit).

PROPOSITION 6.25 (Consequences of stable quasi-orbits). If (Λ, Λ∞) is geometrically finitely-generated and ρ is proper and has stable quasi-orbits, then the following hold: (i) := ρ(Λ).o [Λ] . O ∈ int (ii) (Λ, Λ∞) is a hyperbolic approximate group. (iii) If A is an apogee for Λ, then there is a quasi-isometric embedding ι : A X whose image is given by . → O .(iv) Any map ι as in (iii) induces a continuous embedding ∂ι : ∂A ֒ ∂X whose image is given by (Λ) → Lρ In particular, (Λ) is a representative of ∂Λ, and hence compact. Lρ PROOF. (i) Literally as in Proposition 6.24. (ii) follows from (i) and the fact that a stable subspace is Morse hyperbolic (see Proposition D.24). (iii) follows from (i). (iv) follows from (i) and Proposition D.24. 

4. Convex cocompact qiqacs We keep the notation of the previous sections (see Notation 6.1). In addition we are going to assume that Λ is infinite and that ρ is proper. These assumptions imply in particular that all quasi- orbits of Λ are unbounded. The main goal of this section is to characterize proper qiqacs with stable quasi-orbits geometrically. It will turn out they are precisely the “convex-cocompact” qiqacs and that all of their limit points are conical. 4. CONVEX COCOMPACT QIQACS 79

As the name suggests, a “convex-cocompact” qiqac is a qiqac which quasi-restricts to a cobounded qiqac on a quasi-convex subset of X. To construct such a subset we need to recall from Definition D.27 the notion of the weak hull H(Z) X of a subset Z ∂M X: By definition, H(Z) is the union of ⊂ ∈⊂ s all bi-infinite geodesic lines which are bi-asymptotic to a pair of distinct points in Z. The following observation shows that - in the generality of our present setting (cf. Notation 6.1) - the weak hull of the limit set of ρ provides a candidate for a Λ-cobounded subspace of X, provided the limit set is compact:

PROPOSITION 6.26 (Weak hull of the limit set). If (Λ) is compact, then H( (Λ)) is Λ-quasi- Lρ Lρ invariant. Since the limit set is Λ∞-invariant by Proposition 6.13, the proposition is actually a special case of the following general observation:

LEMMA 6.27. Assume a set Z ∂M X is compact and Λ∞-invariant. Then H(Z) is Λ-quasi-invariant. ⊂ s PROOF. We need to show that there exists a constant D 0 so that for any λ Λ we have ≥ ∈ d (H(Z),ρ(λ).H(Z)) < D. Thus let λ Λ and x H(Z); there then exist ξ−, ξ+ Z with ξ− = ξ+ Haus ∈ ∈ ∈ 6 and a geodesic line γ bi-asymptotic to (ξ−, ξ+) such that x γ(R). ∈ We first observe that since Z is compact we have Z ∂N X for some fixed Morse gauge N. By ⊂ s Theorem D.15.(ii) we deduce that γ is N ′-Morse for some Morse-gauge N ′ depending only on N. Let γ′ := ρ(λ).γ; this is an N ′′-Morse (K, C)-quasi-geodesic, where N ′′, K and C are independent of γ and λ and depend only on N ′ (hence N) and the QI constants of ρ (see [CS15, Lemma 2.5 (2)]). We set D := 2N ′′(K, C). Let γ′ be a geodesic segment joining γ′( n) and γ′(n). By[CS15, Lemma 2.5 (3)] the geodesics γ′ n − n are bounded Hausdorff distance from γ′([ n,n]), where the bound depends only on N,K,C and thus − by [CS15, Lemma 2.5 (1)] are N ′′′-Morse, where N ′′′ depends only on N,K,C. By the Arzelà–Ascoli Lemma A.16 and [Cor17, Lemma 2.10], this converges to an N ′′′-Morse bi-infinite geodesic γ′′ that is of Hausdorff distance at most D from γ′. In particular, ρ(λ)(x) is of distance at most D from γ′′. Since γ′′ is bi-asymptotic to (∂ ρ(λ)(ξ−), ∂ ρ(λ)(ξ+)) Z2, hence contained in H(Z), the lemma follows.  s s ∈ REMARK 6.28 (Quasi-restriction to the weak hull). Assume that (Λ) is compact and hence quasi- Lρ invariant under ρ(Λ). We can then quasi-restrict the qiqac ρ to a qiqac ρH on H( (Λ)) (see Con- (Λ) Lρ struction 4.19). Then, by construction, ρH(Λ) is a qiqac on H( ρ(Λ)) and it is at bounded distance from L k ρ in the following sense: for any k N there exists a constant Rk so that fore any λ Λ and x X we ∈ ∈ ∈ have e d(ρH(Λ)(λ).x, ρ(λ).x) < Rk. (4.1)

DEFINITION 6.29. Let (Λ, Λ∞) be an infinite geometrically finitely-generated approximate group e and let ρ: (Λ, Λ∞) QI(X) be a qiqac on a proper geodesic space X. We say that ρ is convex cocompact → if: (CC1) Λ quasi-acts properlyf on X via ρ; (CC2) (Λ) ∂M X is non-empty and compact; Lρ ⊂ s (CC3) Λ quasi-acts coboundedly on H( (Λ)) via ρ := ρH . Lρ (Lρ(Λ)) We can now state the main theorem of this section: e e THEOREM 6.30 (Convex cocompactness vs. stable quasi-orbits). Suppose that X is a proper geodesic metric space, (Λ, Λ∞) is a geometrically finitely-generated approximate group, and ρ: (Λ, Λ∞) QI(X) is a → qiqac. Then the following are equivalent: (i) ρ is convex cocompact. f (ii) ρ is proper and has stable quasi-orbits. In this case, (Λ, Λ∞) is a hyperbolic approximate group and (Λ) is compact and represents ∂Λ. Lρ In the group case, a similarly formulated equivalence is proved in [CD19]. It will be convenient to use the following reformulation of Condition (CC3), which is immediate from (4.1): 80 6. LIMIT SETS OF QUASI-ISOMETRIC QUASI-ACTIONS

LEMMA 6.31. Condition (CC3) is equivalent to the condition (CC3’) There exists D> 0 such that for every z H( (Λ)) there is some λ Λ with d(ρ(λ).o, z)

It seems worthwhile to record the following special case of Theorem 6.30; in the group case, this result is due to Swenson [Swe01].

COROLLARY 6.33 (Convex cocompactness vs. quasi-convex quasi-orbits). Suppose that X is a proper geodesic hyperbolic space, (Λ, Λ∞) is a geometrically finitely-generated approximate group, and ρ: (Λ, Λ∞) → QI(X) is a qiqac. Then the following are equivalent: (i) ρ is convex cocompact. f(ii) ρ is proper and has quasi-convex quasi-orbits. In this case, (Λ, Λ∞) is a hyperbolic approximate group and (Λ) represents ∂Λ. Lρ It turns out that convex cocompactness of a qiqac is enough to ensure that all limit points are conical:

PROPOSITION 6.34 (Limit points are conical). If ρ: (Λ, Λ∞) QI(X) is a convex cocompact qiqac on → a proper geodesic metric space X, then every limit point of ρ is conical. f PROOF. We may assume that Λ is infinite, since otherwise the proposition holds vacuously. By Proposition 5.32 this implies that ∂Λ 2. By Theorem 6.30 we may now fix a Morse gauge N such | | ≥ that := ρ(Λ).o is N-stable. Denote L := ( ) and note that by Theorem 6.30 we have L 2. O L O (N) | |≥ Let ξ L. Since is N-stable, there exists a a sequence (x ) in ρ(Λ).o Xo such that ξ = ∈ O n ∩ [(xn)]. Let γn be a geodesic joining o to xn. By Arzelà–Ascoli (Lemma A.16) we know that the γn subsequentially converge to an N-Morse geodesic γ representing ξ. Since L 2 we can choose | | ≥ ξ′ L ξ . By Theorem D.15 there then exists a bi-infinite geodesic line α which is bi-asymptotic to ∈ \{ } (ξ′, ξ), and thus contained in H(L). 4. CONVEX COCOMPACT QIQACS 81

Again, by Theorem D.15, there is a constant δ depending only on N and constants T,T ′ (depend- ing on γ and α) such that d (γ([T, )), α([T ′, ))) < δ. Since α([T ′, ]) H(L), the proposition Haus ∞ ∞ ∞ ⊂ now follows from (CC3’).  It follows from the proposition that the isometric actions studied in Theorem 6.22 are not convex cocompact. In general, convex cocompact isometric actions are rare:

REMARK 6.35 (Convex cococompact isometric actions). There is a major difference between con- vex cocompact qiqacs and convex cocompact isometric actions: Indeed, if ρ: (Λ, Λ∞) Is(X) is a → convex cocompact isometric action on a proper geodesic metric space X, then the weak hull H( (X)) Lρ is not only quasi-invariant, but actually invariant under ρ(Λ), and hence under ρ(Λ∞). This is be- cause isometries map geodesics with endpoints in (X) to geodesics in (X). As a consequence, if Lρ Lρ (Λ, Λ∞) acts convex cocompactly on X, then it acts geometrically on H( (X)). Note that the latter is Lρ quasi-convex (by Proposition D.30), and hence large-scale geodesic. If we denote by G the setwise stabilizer of (X) in Is(X, d), we may thus deduce that ρ(Λ) is a Lρ uniform approximate lattice in G by Proposition 4.10. In particular, if G is discrete or Λ∞ acts properly, then Λ is an almost group (by the same argument as in the proof Corollary 4.11). To summarize:

COROLLARY 6.36 (Obstructions to convex cocompact isometric actions). Let (Λ, Λ∞) be a geomet- rically finitely-generated approximate group which is not an almost group. Assume that (Λ, Λ∞) acts iso- metrically and convex cocompactly on a proper geodesic metric space X. Then the image of Λ∞ in Is(X) is non-discrete. In particular, the limit set has a non-discrete stabilizer Is(X) and Λ∞ does not act properly on X.  Note that there exist plenty of Fuchsian groups which act convex cocompactly on the hyperbolic plane and whose limit set has a discrete stabilizer in PSL2(R); these are mainly responsible for the rich zoo of convex cocompact Fuchsian groups. In the world of approximate groups which are not almost groups, such examples cannot exist. However, to the best of our knowledge no similar obstructions exist for convex cococompact qiqacs. This shows once again that qiqacs, rather than isometric actions, should be the focus of geometric approximate group theory.

CHAPTER 7

Asymptotic dimension

1. Summary of results The goal of this chapter is to establish a number of fundamental results concerning asymptotic dimension of countable approximate groups, as defined in Definition 3.13 above. The main result of this chapter concerns hyperbolic approximate groups. Namely, for these approximate groups, their as- ymptotic dimension can be related to topological dimension of their boundary, as defined in Example 5.12 above:

THEOREM 7.1 (Asymptotic dimension of hyperbolic approximate groups). If (Λ, Λ∞) is a hyperbolic approximate group with Gromov boundary ∂Λ, then asdim Λ = dim ∂Λ+1.

Equivalently, Theorem 7.1 says that for some (hence any) apogee X of a hyperbolic approximate group (Λ, Λ∞) we have asdim X = dim ∂X +1, where ∂X is the Gromov boundary of X. For any such apogee and any visual metric d on ∂X we will actually establish the stronger result, namely that asdim X = dim ∂X +1= ℓ- dim(∂X,d)+1, where ℓ- dim denotes linearly controlled metric dimension as defined in Definition 7.9 below. This will prove in particular that ℓ- dim ∂Λ := ℓ- dim(∂X,d) depends neither on the choice of apogee X nor on the choice of visual metric d on ∂X, but only on [Λ]int. We refer the reader to Section 2 below for precise definitions of dim and ℓ-dim. ′ For hyperbolic groups, Theorem 7.1 was conjectured by Gromov ([Gro93], 1.E1) and established by Sergei Buyalo and Nina Lebedeva [BL07], based on earlier work by many authors (like [BM91], [Sa95],[BS07b]). Our proof follows the steps of the proof in the group case, as presented in [BS07a]. We would like to point out that one of the proofs ([BL07, Thm. 1.1], which is [BS07a, Thm. 12.2.1]) contains a gap that we were unable to fix. Nina Lebedeva kindly provided us with a list of corrections for this proof, and our arguments will be parallel to her corrected version (see Section 7). The proof of Theorem 7.1 will occupy most of the current chapter; we will actually establish a more general theorem which applies to all quasi-cobounded proper geodesic Gromov hyperbolic spaces (see Theorem 7.18 below). Besides Theorem 7.1, we also establish two smaller results, which are useful when dealing with asymptotic dimension of approximate groups. To describe the first of these results, we recall from Example 3.9 the definition of the coarse kernel of a morphism between approximate groups. With this notion understood, we have the following result:

THEOREM 7.2 (Asymptotic Hurewicz mapping theorem for countable approximate groups). Let (Ξ, Ξ∞), (Λ, Λ∞) be countable approximate groups and let f : (Ξ, Ξ∞) (Λ, Λ∞) be a global morphism of → approximate groups. Then asdim Ξ asdim Λ + asdim([ker(f)] ).  ≤ c In fact, we establish a local version of this theorem in Theorem 7.15. The proof is relying on an as- ymptotic version of the classical (dimension lowering) Hurewicz mapping theorem from [BDLM08].

83 84 7. ASYMPTOTIC DIMENSION

Our second minor result concerns asymptotic dimension of countable approximate groups which are not algebraically finitely-generated. If (Λ, Λ∞) is such an approximate group, then we provide in Theorem 7.16 a formula for asdim Λ in terms of the finitely-generated subgroups of Λ∞. Before we turn to the proofs of these theorems, in the following section we collect some back- ground results concerning dimension theory.

2. Background on dimension theory For the reader’s convenience, we summarize in this section some basic properties of topological dimension dim and linearly controlled metric dimension ℓ-dim. We start with some general notation concerning covers of metric spaces. If X is a set, then a collection of subsets of X is called a cover of X if X = . If X is a U U topological space, then a cover is called an open cover if every U is an open subset of X. While in U ∈U S classical dimension theory one is mostly concerned with open covers of topological or metric spaces, let us emphasize that the covers considered in the definition of asymptotic dimension need not be open. The multiplicity (or order) mult of a cover is defined as the maximal number n such that U U there exist U ,...,U with U U = . With this notation understood we have the following 1 n ∈U 1 ∩···∩ n 6 ∅ definition of topological dimension (see e.g. [Mun75]):

DEFINITION 7.3. For a nonempty topological space X, its topological dimension (or covering dimen- sion) dim X is the minimal integer n N such that for every open cover of X, there is an open cover ∈ 0 U of X which refines , and which has the multiplicity mult n +1. If no such n N exists, we V U V ≤ ∈ 0 say that dim X = . We also set dim := 1. ∞ ∅ − REMARK 7.4. Note that the definition of topological dimension dim is often stated for finite open covers (see, for example, [Eng95, Def. 1.6.7]or [Coo15, Def. 1.1.8]). But as long as a topological space U X is Hausdorff and paracompact, this definition is equivalent to ours (see [Eng95, Prop. 3.2.2]). We will often be interested in covers of metric spaces. If is a cover of a metric space (X, d), then U we can define its mesh as mesh := sup diam U U . U { | ∈U} The following definition is taken from [Eng95, Ch. 1.6].

DEFINITION 7.5. Let (X, d) be a nonempty metric space. The metric dimension µ dim X is the small- est integer n N such that for every ε > 0 there is an open cover of X with mult n +1 and ∈ 0 U U ≤ mesh ε. U ≤ REMARK 7.6. In [BS07a, p. 107], the dimension dim X of a metric space (X, d) is defined as µ dim X. This is somewhat unfortunate since [Eng95, Ex. 1.10.23] shows that µ dim X and dim X do not coincide for general metric spaces. The reason that this will not cause any problems for us in the sequel, when applying results from [BS07a], is that we will only consider topological dimension of non-empty compact metric spaces, and for these we have the following result.

THEOREM 7.7. If (X, d) is a non-empty compact metric space, then µ dim X = dim X.

PROOF. See[Eng95], Definition 1.6.7 and Theorem 1.6.12.  Thus, when we know we are dealing with a compact metric space X, we may use Definition 7.5 as our definition for dim X. Our next goal is to introduce the notion of linearly controlled metric dimension of a metric space. For this we need the definition of the Lebesgue number of an open cover.

DEFINITION 7.8. For an open cover of a metric space X, the Lebesgue number of is the number U U L( ) := inf L( , x), where L( , x) := sup dist(x, X U). U x∈X U U U∈U \ Note that L( ) may be strictly larger than mesh (and even infinite, as will be the case for L( ) U U U−1 in the proof of Theorem 7.21 below). The following definition of linearly controlled metric dimension is given in Subsection 9.1.4 of [BS07a]. b 3. AN ASYMPTOTIC HUREWICZ MAPPING THEOREM FOR COUNTABLE APPROXIMATE GROUPS 85

DEFINITION 7.9. If (X, d) is a non-empty metric space, then its linearly controlled metric dimension ℓ-dim X is the minimal integer n N with the following property: there exists a δ (0, 1) such that ∈ 0 ∈ for every sufficiently small r > 0 there is an open cover of X with mult n +1, mesh r U U ≤ U ≤ and L( ) δr. If no such n N exists, we say that ℓ-dim X = . Since ℓ-dim of a metric space U ≥ ∈ 0 ∞ depends on the choice of metric, we will often emphasize this choice by writing ℓ-dim(X, d) instead of just ℓ-dim X. In [BL07], linearly controlled metric dimension is referred to as capacity dimension; the term micro- scopic Assouad-Nagata dimension is also sometimes used. For general metric spaces (X, d) we always have the inequality ℓ-dim X dim X (see [LS05, Proposition 2.2]), but the converse is not true in ≥ general. See [BS07a, Chapter 11] for some examples of calculation of ℓ-dim and dim.

REMARK 7.10 (Coloring definitions of dim and ℓ-dim). Our definition of asymptotic dimension (Definition 3.11 above) was in terms of colored coverings. There are similar coloring versions of defini- tions of dim and ℓ-dim, which are often easier to work with than the original definitions, and which we briefly recall here. As explained on page 109 of [BS07a], a family of subsets of a space X is disjoint U if it is pairwise disjoint, i.e., if its multiplicity mult =1. U For m N, a family is called m-colored if it is the union of m families m i, and each family i ∈ U ∪i=1U U (of the same color i) is disjoint. Clearly, the multiplicity of an m-colored family is at most m, i.e., each point contained in the union of this family can be contained in at most m different elements of the family. As it turns out (see in [BS07a, Section 9.1.6]), for a metric space X, in the definitions for µ dim and ℓ-dim above one can replace the requirement that multiplicity of the relevant family be n +1 ≤ by the requirement that this family be (n + 1)-colored. Thus, for a nonempty compact metric space X, and n N , we have dim X n if and only if for every ε > 0, there is an (n + 1)-colored open cover ∈ 0 ≤ = n i of X with mesh ε. U ∪i=0U U ≤ Similarly, for a nonempty metric space X, and n N , we have ℓ-dim X n if and only if there ∈ 0 ≤ exists a δ (0, 1) such that for every sufficiently small r > 0 there is an (n + 1)-colored open cover ∈ = n i of X with mesh r and L( ) δr. U ∪i=0U U ≤ U ≥ 3. An asymptotic Hurewicz mapping theorem for countable approximate groups In this section we establish (a local version of) Theorem 7.2. To motivate our result, we first recall the classical Hurewicz dimension-lowering theorem for maps.

THEOREM 7.11 (Hurewicz dimension-lowering theorem). Let X and Y be separable metric spaces and let f : X Y be a closed map. Then → dim X dim Y + dim(f), where dim(f) := sup dim(f −1(y)) y Y .  ≤ { | ∈ } In order to state a version of this theorem for asymptotic dimension, we need the following defi- nition from [BD01]:

DEFINITION 7.12. Let Y := Yα α∈A be a collection of metric spaces and let n N0. We say that { } u ∈ the asymptotic dimension of Y is uniformly bounded by n, and write asdim(Y) n, if for all R> 0 there (0) (n) ≤ exists D > 0 such that for every α A there exist collections α ,..., α of subsets of Y satisfying ∈ U U α properties (ASD1)–(ASD3) from Definition 3.11 with respect to R and D. The following asymptotic Hurewicz mapping theorem is due to G. Bell and A. Dranishnikov (see [BD08]):

THEOREM 7.13 (Asymptotic Hurewicz mapping theorem, first version). Let f : X Y be a Lip- → schitz map from a geodesic metric space X to a metric space Y . If for every ρ > 0 the collection Yρ := u f −1(B(y,ρ)) satisfies asdim(Y ) n, then asdim X asdim Y + n.  { }y∈Y ρ ≤ ≤ A generalization of this result, due to N. Brodskiy, J. Dydak, M. Levin and A. Mitra ([BDLM08]), states: 86 7. ASYMPTOTIC DIMENSION

THEOREM 7.14 (Asymptotic Hurewicz mapping theorem, second version). Let h : X Y be a −1 → coarsely Lipschitz map between metric spaces. If for every ρ > 0 the collection Yρ := h (B(y,ρ)) y∈Y u { } satisfies asdim(Y ) n, then asdim X asdim Y + n.  ρ ≤ ≤ Note that in the original statement of this theorem, [BDLM08, Theorem 1.2], what they call a “large scale uniform” function is precisely what is called a coarsely Lipschitz function in this book. Also, instead of asdim X asdim Y + n, they write asdim X asdim Y + asdim h, where asdim h is ≤ ≤ u defined as sup asdim A A X and asdim(h(A)) = 0 . However, the property asdim(Y ) n can { | ⊆ } ρ ≤ be shown to be equivalent to the map h having an n-dimensional control function (terminology of [BDLM08]), and by Corollary 4.10 of [BDLM08], this is equivalent to asdim h n. ≤ For countable approximate groups we obtain the following consequence:

THEOREM 7.15 (Local asymptotic Hurewicz mapping theorem for approximate groups). Let (Ξ, Ξ∞), (Λ, Λ∞) be countable approximate groups and let f : (Ξ, Ξ6) (Λ, Λ6) be a 6-local morphism of 6 → approximate groups. Then asdim Ξ asdim Λ + asdim([ker (f )] ). ≤ 6 6 c 6 PROOF. Replacing (Λ, Λ ) by the image of f6 if necessary, we may assume that f6 is surjective as a map of pairs. We fix left-admissible metrics d on Ξ∞ and d′ on Λ∞. Given ξ Ξ∞ and λ Λ∞, we ∈ ∈ denote by B(ξ,ρ) and B′(λ, ρ) the balls of radius ρ around ξ, respectively λ, in Ξ∞, respectively Λ∞. By Lemma 3.6 the restriction f : Ξ Λ is coarsely Lipschitz. In view of Theorem 7.14 it thus 1 → suffices to show that, for every ρ> 0, the collection Y := f −1(B′(λ, ρ) Λ) ρ { 1 ∩ }λ∈Λ satisfies the inequality u asdim(Y ) asdim(f −1(e)). (3.1) ρ ≤ 6 Thus fix ρ > 0, and for every λ Λ pick ξ = ξ(λ) f −1(λ). Also note that (λB′(e,ρ)) Λ ∈ ∈ 1 ∩ ⊂ λ B′(e,ρ) Λ2 , since if z = λb = λ, for some b B′(e,ρ), λ Λ, then b = λ−1λ Λ2. Now we have ∩ ∈ ∈ ∈ −1 ′ −1 ′ −1 ′ 2 f (B (λ, ρ) Λ) = f ((f1(ξ)B (e,ρ)) Λ) f f1(ξ) B (e,ρ) Λ 1 ∩ 1e ∩ e ⊂ 3 e ∩ = z Ξ3 f (z) f (ξ)(B′(e,ρ) Λ2) { ∈ | 3 ∈ 1 ∩ }  z Ξ3 f (ξ−1z) B′(e,ρ) Λ2 ⊂ { ∈ | 4 ∈ ∩ } ξf −1(B′(e,ρ) Λ2). ⊂ 4 ∩ Since left-multiplication by ξ yields an isometry of Ξ∞, we have reduced (3.1) to proving that asdim(f −1(B′(e,ρ) Λ2)) asdim(f −1(e)) (3.2) 4 ∩ ≤ 6 for every ρ > 0. Now by properness of d′, the ball B′(e,ρ) and hence also its intersection with Λ2 are finite, say B′(e,ρ) Λ2 = λ ,...,λ . Since f surjects onto Λ2, we can find ξ ,...,ξ Ξ2 such that ∩ { 1 N } 2 1 N ∈ f2(ξi)= λi. Define R := max d(e, ξi). 1≤i≤N Then for i =1,...,N and η f −1(λ ), we have η ξ−1 Ξ4Ξ2 = Ξ6, and f (η ξ−1)= λ λ−1 = e. Since i ∈ 4 i i i ∈ 6 i i i i d(η ,f −1(e)) d(η , η ξ−1)= d(e, ξ−1) R, we deduce that i 6 ≤ i i i i ≤ f −1(λ ) N (f −1(e)), 4 i ⊂ R 6 and consequently N f −1(B′(e,ρ) Λ2)= f −1(λ ) N (f −1(e)). 4 ∩ 4 i ⊂ R 6 i=1 [ We deduce that asdim(f −1(B′(e,ρ) Λ2)) asdim(N (f −1(e))) = asdim(f −1(e)), 4 ∩ ≤ R 6 6 which establishes (3.2) and finishes the proof.  4. ASYMPTOTIC DIMENSION OF NON-FINITELY GENERATED COUNTABLE APPROXIMATE GROUPS 87

Since the coarse kernel of a global morphism f : (Ξ, Ξ∞) (Λ, Λ∞) satisfies → [ker(f)]c := [ker2(f)]c = [ker3(f)]c = [ker4(f)]c = [ker5(f)]c = [ker6(f)]c = ..., Theorem 7.2 it now an immediate consequence of Theorem 7.15.

4. Asymptotic dimension of non-finitely generated countable approximate groups Throughout this section, (Λ, Λ∞) denotes a countable approximate group. If (Λ, Λ∞) is algebraically finitely-generated, then we can choose a finite symmetric generating set S of Λ∞, and asdim Λ can be computed as the asymptotic dimension of Λ with respect to the restric- ∞ tion of the word metric dS. If (Λ, Λ ) is not algebraically finitely-generated, then the computation of asdim Λ is more complicated. One could choose a weighted word metric d on Λ∞ and then compute the asymptotic dimension of Λ with respect to the restriction of d, but working with such weighted word metrics is rather inconvenient. In the case of non-finitely-generated countable groups, an alternative approach was given by A. Dranishnikov and J. Smith in [DS06, Thm. 2.1]: they show that the asymptotic dimension of a count- able group Γ is the supremum over the asymptotic dimensions of all finitely-generated subgroups of Γ. Their techniques yield the following result for approximate groups, which is unfortunately slightly more technical to state than the corresponding theorem in the group case:

THEOREM 7.16. Let (Λ, Λ∞) be a countable approximate group. Then

asdim Λ = sup  asdim Ξ Ξ Λ2, Ξ is an approx. subgroup of Γ  , (4.1) { | ⊂ } Γ<Λ∞  [  Γ finitely generated    i.e., the asymptotic dimension of Λ is the supremum of the asymptotic dimensions of approximate subgroups of finitely-generated subgroups of Λ∞, with these approximate subgroups contained in Λ2.

PROOF. If (Λ, Λ∞) is an approximate group, then so is (Λ2, Λ∞), and by Corollary 3.14 (ii) we have asdim Λ = asdim Λ2. If (Ξ, Ξ∞) is any approximate subgroup of (Λ2, Λ∞), then (iii) of Corollary 3.14 yields asdim Λ = asdim Λ2 asdim Ξ. ≥ Passing to the supremum over all approximate subgroups (Ξ, Ξ∞) of finitely-generated subgroups of Λ∞ yields the inequality for the expression (4.1). ≥ Concerning the converse inequality, let us fix a weight function w :Λ∞ [0, ) and denote by d → ∞ the associated weighted word metric on Λ∞. If the supremum on the right side of (4.1) equals , then ∞ by the first part of the proof we have asdim Λ = and the proof is finished. So let us assume that ∞ the supremum on the right in (4.1) equals m, for some m N . We want to show that asdim Λ m, ∈ 0 ≤ i.e., that for any R > 0 (large enough), we can find a uniformly bounded, (m + 1)-colored cover = m (i) of Λ, where each (i) is R-disjoint. V ∪i=0V V Let R> 0 be given. We define T := g Λ∞ w(g) R and Γ := T . R { ∈ | ≤ } R h Ri Since w is proper, note that the set TR is finite, and thus the group ΓR is finitely-generated. We also define Ξ := Λ2 Γ . R ∩ R Since both Λ and ΓR are symmetric and contain e, the same is true of ΞR. In addition, we claim that for all sufficiently large R, the subset Ξ Γ is an approximate subgroup of the finitely-generated R ⊂ R group Γ . Indeed, let F Λ∞ be a finite set such that Λ4 F Λ2. Then w is bounded on F and thus R ⊂ ⊂ F T for all sufficiently large R. For all such R and ξ , ξ Ξ we then find λ Λ2 and f F such ⊂ R 1 2 ∈ R ∈ ∈ that ξ ξ = fλ. Since F T , we have 1 2 ⊂ R λ = f −1ξ ξ T Γ Γ =Γ , 1 2 ∈ R R R R 88 7. ASYMPTOTIC DIMENSION and thus λ Ξ . This shows that Ξ2 F Ξ , and the claim follows. So let us agree that the constant ∈ R R ⊂ R R> 0 given above was indeed large enough so that ΞR defined above is an approximate subgroup of ΓR. By our assumption, we have asdim ΞR m. ≤ ∞ ∞ Now we shall focus on (left) cosets of ΓR in Λ . First, observe that if g,h Λ are in two distinct ∞ ∈ ΓR-cosets of Λ , then d(g,h) > R. (4.2) −1 Namely, in this case the element g h is not contained in ΓR, hence cannot be written as a product of −1 elements of weight R. Therefore d(g,h)= g h w > R, as claimed. ≤ k k ∞ Furthermore, let us now choose a set ZR of representatives of ΓR-cosets in Λ so that ∞ Λ = zΓR. z∈GZR Define Z0 := z Z Λ (zΓ ) = so that R { ∈ R | ∩ R 6 ∅} Λ= Λ (zΓ ). ∩ R z∈Z0 GR We may assume without loss of generality that Z0 Λ, since every coset which intersects Λ non- R ⊂ trivially admits a coset representative in Λ. Under this assumption we then have Λ= Λ (zΓ )= z(z−1Λ Γ ) z(Λ2 Γ )= zΞ =: Λ . ∩ R ∩ R ⊂ ∩ R R R z∈Z0 z∈Z0 z∈Z0 z∈Z0 GR GR [R [R Also note that e Λ Λ zΛ2 Λ3. ⊂ R ⊂ ⊂ z∈Z0 [R Since Z0 Z , the elements of Z0 represente distinct cosets of Γ . Consequently, zΓ z′Γ = for R ⊂ R R R R ∩ R ∅ all z = z′ in Z0 , and thus the union defining Λ is disjoint, i.e., 6 R R Λ = zΞ . Re R z∈Z0 GR e Moreover, if z = z′ are distinct elements of Z0 and g zΞ and h z′Ξ , then d(g,h) > R by (4.2). 6 R ∈ R ∈ R Since we know that asdim Ξ m, for the same R given above we can choose the families R ≤ (0),..., (m) of subsets of Ξ so that each family (i) is R-disjoint, D-bounded (for some D > 0), U U R U and so that = m (i) is a cover for Ξ . Using this , we define U ∪i=0U R U (i) := zU z Z0 ,U (i) , i =0, . . . , m, VR { | ∈ R ∈U } (i) that is, we transport the (m + 1)-colored cover of ΞR to each zΞR, and R represents the family of U (i) V sets of the same color in z∈Z0 zΞR. Then the sets in R are still D-bounded, by left-invariance of the R V metric. Moreover, let V = zU and V ′ = z′U ′ be two distinct elements of (i). If z = z′, then V and V ′ F VR 6 are of distance at least R by (4.2). If z = z′, then U = U ′, and hence V and V ′ are of distance at least R 6 by R-disjointedness of (i). So := m (i) is a uniformly bounded, (m + 1)-colored cover of Λ , U VR ∪i=0VR R with each (i) R-disjoint. Thus =Λ = m (Λ (i)) is a uniformly bounded, (m + 1)-colored VR V ∩VR ∪i=0 ∩VR cover of Λ, with each (i) =Λ (i) R-disjoint. Therefore asdim Λ m, which finishes the proof. e  V ∩VR ≤ Note that the theorem holds trivially if (Λ, Λ∞) is algebraically finitely-generated, so the interest is in the case where Λ∞ is not finitely-generated.

REMARK 7.17. If (Ξ, Ξ∞) is an approximate subgroup of (Λ, Λ∞) such that Ξ (and hence Ξ∞) is contained in a finitely-generated subgroup Γ of Λ∞, then asdim Ξ can be computed as follows: Let S be a finite generating set of Γ and let dS be the associated word metric. Then dS restricts to a left- admissible word metric on Ξ∞ and hence to an external metric on Ξ, and we thus have asdim Ξ = asdim(Ξ, d ). S |Ξ×Ξ 5.ASYMPTOTICDIMENSIONVS.DIMENSIONOFTHEBOUNDARY 89

Thus Theorem 7.16 allows us to compute asdim Λ using only word metrics on finitely-generated sub- groups of Λ∞ (rather than a weighted word metric on Λ∞).

5. Asymptotic dimension vs. dimension of the boundary The remainder of this chapter is devoted to the proof of Theorem 7.1. We recall from Proposition 5.4 that if X is an apogee of a hyperbolic approximate group, then it is proper, geodesic, Gromov hyperbolic and quasi-cobounded. Thus Theorem 7.1 is implied by the following theorem:

THEOREM 7.18. If X is a quasi-cobounded proper geodesic Gromov hyperbolic space with Gromov bound- ary ∂X and d denotes a visual metric on ∂X, then asdim X = dim ∂X +1= ℓ- dim(∂X,d)+1. Let us emphasize again that X being proper, geodesic and Gromov hyperbolic means that ∂X = ∂rX = ∂sX is compact, so it is safe to use (the coloring version of) Definition 7.5 of dim ∂X (see Remark 7.10). For cobounded (rather than just quasi-cobounded) spaces, this is [BS07a, Thm. 12.3.3]. Our con- tribution here lies in weakening the assumption from coboundedness to quasi-coboundedness. One of the two inequalities in the theorem actually holds in the complete generality of proper geodesic hyperbolic spaces without any (quasi-)coboundedness assumptions, see [BS07a, Thm. 10.1.2]:

THEOREM 7.19. For every proper geodesic hyperbolic space X, we have asdim X dim ∂X +1.  ≥ We will establish Theorem 7.18 by showing that for every quasi-cobounded proper geodesic Gro- mov hyperbolic space X, with Gromov boundary ∂X with a visual metric d, we have asdim X ℓ- dim(∂X,d)+1 dim(∂X)+1. (5.1) ≤ ≤ The full strength of the quasi-coboundedness assumption is probably not necessary to establish these inequalities. On the other hand, (5.1) does not hold for arbitrary proper geodesic Gromov hy- perbolic spaces, as the following example shows:

EXAMPLE 7.20 (Hyperbolic Shashlik). Let γ : [0, ) Hn be a geodesic ray in hyperbolic n-space ∞ → and let x , x ,... be points of γ([0, )) such that d(x , x ) 2n+2, for all n N. We consider the 1 2 ∞ n n+1 ≥ ∈ hyperbolic shashlik ∞ X := γ([0, )) B(x , 2n) Hn, ∞ ∪ n ⊂ n=1 [ which is a proper geodesic Gromov hyperbolic space. Since X contains arbitrarily large balls in Hn, we have asdim X = asdim Hn = n. On the other hand, X contains a single geodesic ray, and thus ∂X = and in particular dim ∂X =0. {∗} The main step in the proof of (5.1) (and hence Theorem 7.18) is the following theorem, which generalizes [BS07a, Thm. 12.2.1]

THEOREM 7.21. If a metric space X is locally quasi-similar to a compact metric space Y and its linearly controlled metric dimension ℓ-dim X is finite, then ℓ-dim X dim Y . ≤ We will prove Theorem 7.21 in Section 7 below. We now explain how Theorem 7.21 is used to prove Theorem 7.18. Note that Theorem 7.18 holds trivially if X is bounded: in this case, asdim X =0 and ∂X = , hence dim = 1 = asdim X 1. We may thus focus on the unbounded case. ∅ ∅ − − From now on let X be an unbounded quasi-cobounded proper geodesic Gromov hyperbolic space, and choose any visual metric d on its Gromov boundary ∂X.

REMARK 7.22. In the proof of Theorem 7.18, we will use three properties of X: (i) The space X is visual (see Proposition 5.8). 90 7. ASYMPTOTIC DIMENSION

(ii) ∂X is locally quasi-self-similar, i.e., locally quasi-similar to itself, with respect to any visual metric (see Proposition 5.16). (iii) ∂X is doubling, with respect to any visual metric (see Theorem 5.20). Equipped with these three properties we can now complete the proof of Theorem 7.18 using the same embedding argument as in [BS07a, Chapter 12]. Here is how the argument works. First, the fact that (∂X,d) is doubling, so in particular, doubling at small scales, implies by [BS07a, Lemma 12.2.2] that ℓ- dim(∂X,d) < . ∞ Then, we will use the following theorem of Buyalo [Buy05] (reproduced as [BS07a, Theorem 12.1.1] and based on earlier work of Buyalo and Schroeder [BS05]): THEOREM 7.23 (Buyalo–Schroeder embedding theorem). Let Z be a visual Gromov hyperbolic space such that n := ℓ- dim ∂ Z < . Then there exists a quasi-isometric embedding Z T ... T , where s ∞ → 1 × × n+1 T1,...,Tn+1 are simplicial trees.  REMARK 7.24 (Terminology concerning trees). Following [BS07a, Ex. 1.2.5], we call a geodesic metric space ( , d) a metric tree if every triangle in T is a (possibly degenerate) tripod. If T is a tree T in the sense of , i.e. a connected graph without loops, then we obtain a metric tee ( , d) T as follows: = T is the geometric realization of T , and the metric on is the unique path-metric T | | T such that every edge has length 1. We then call ( , d) a simplicial tree. No assumption is made here T concerning the cardinality of the vertex set of T nor on the valences of the vertices. If all vertices of T have finite valences (and hence the vertex set if countable), then the metric d is proper and ( , d) is T called a locally finite simplicial tree. The conclusion of [BS07a, Theorem 12.1.1] is that a certain metric space admits a quasi-isometric embedding into a product of metric trees. However, as remarked on [BS07a, p.147], the proof actually constructs an embedding into a product of (typically non-locally finite) simplicial trees. So, considering the fact that our space X is Gromov hyperbolic and visual, and assuming that n = ℓ- dim(∂X,d) (since we know it is finite), let us apply Theorem 7.23 to X and fix a quasi-isometric embedding F : X T ... T of our space X into a product of (n + 1) simplicial trees. By → 1 × × n+1 [BS07a, Prop. 10.2.1] we have asdim T 1 for every metric tree T , and hence by the product theorem ≤ for asymptotic dimension (see [BD08, Thm. 32]) we have asdim(T T ) n +1. 1 ×···× n+1 ≤ Now recall from Lemma 3.12.(iii) that asdim is monotone under coarse embeddings. We thus obtain (compare [BS07a, Cor. 12.1.11]) that asdim X asdim T ... T n +1= ℓ- dim(∂X,d)+1. ≤ 1 × × n+1 ≤ Finally, since (∂X,d) is locally quasi self-similar and has finite ℓ-dim, we apply Theorem 7.21 to get ℓ- dim(∂X,d) dim ∂X, so ≤ asdim X ℓ- dim(∂X,d)+1 dim ∂X +1, ≤ ≤ which is the desired inequality. Together with the inequality of Theorem 7.19 this gives us the state- ment of Theorem 7.18. We have thus reduced the proof of Theorem 7.18 to the proof of Theorem 7.21. REMARK 7.25. We may as well note here that, since ℓ-dim Z dim Z is true for any metric space ≥ Z, then for our space X and any choice of visual metric d on ∂X we have both ℓ- dim(∂X,d) ≥ dim ∂X (true in general) and ℓ- dim(∂X,d) dim ∂X (from Theorem 7.21), so we get the equality ≤ ℓ- dim(∂X,d) = dim ∂X without referring to inequalities involving asdim X.

6. Preliminaries for the proof of Theorem 7.21 In the previous section we have reduced the proof of our main theorem (Theorem 7.18) to the proof of Theorem 7.21. We will give this proof in Section 7 below. Before we do so, we collect in this section various preliminary notions and results, which will be used in the proof. 6. PRELIMINARIES FOR THE PROOF OF THEOREM 7.21 91

6.1. Restrictions of coverings. In the proof of Theorem 7.21 we will have to restrict various open coverings to subspaces. We thus introduce some terminology to deal with such restrictions. From now on let (X, d) be a metric space and let A X be a subset. If is a family of subsets of ⊂ U X, then we define the restriction of to A by U = U A U . U|A { ∩ | ∈ U} We say that a family of (open) subsets of X is an (open) covering family of A in X if A. This U U ⊇ is equivalent to being an (open) cover of A. We introduce the following notion of the Lebesgue U|A S number for open covering families:

DEFINITION 7.26. If A is a subset of a metric space X and is an open covering family of A in X, U then the Lebesgue number of ( ,A,X) is defined as U LX ( , A) := inf LX ( , x), where LX ( , x) := sup dist(x, X U). (6.1) U x∈A U U U∈U \ Note that for A = X we have

LX ( ,X) = inf LX ( , x)= L( ), U x∈X U U hence in the case A = X we recover the usual definition of the Lebesgue number. The following inequalities are immediate from the definitions:

LEMMA 7.27. If is an open cover of a metric space X and A B X, then U ⊆ ⊆ L( )= L ( ,X) L ( , A) L ( , A).  U X U ≤ X U ≤ B U|B We emphasize that, in the situation of Lemma 7.27, the covering family of A in B may have U|B larger Lebesgue number than the original family. In the case A = B the last inequality of Lemma 7.27 specializes to L ( , A) L ( , A)= L( ), X U ≤ A U|A U|A and it is important to note that even in this special case the inequality can be strict. Similarly, if is an U open cover of a metric space X and A X is a subset, then ⊂ mesh mesh , (6.2) U|A ≤ U but equality does not hold in general. Thus, when restricting an open cover of X to a subspace, the Lebesgue number may increase and the mesh may decrease. The gap in the original proof of [BL07, Thm. 1.1] ([BS07a, Thm. 12.2.1]) is caused by the fact that the authors do not take these effects into account carefully enough. This illustrates the importance of keeping the ambient space in the Lebesgue number notation and mesh notation, which we may omit only when the ambient space is the obvious one. 6.2. Saturation of covering families. Let (X, d) be a metric space. Given a subset W X and ⊂ r > 0, we will denote by Nr(W ) the open r-neighborhood of the set W , and we set N−r(W ) := X N (X W ). \ r \ DEFINITION 7.28. Let U X and let , be families of subsets of X. ⊂ U V (i) The saturation U of the set U by the family is the union of U with all members V with ∗V V ∈ V U V = . ∩ 6 ∅ (ii) The saturation of the family by the family is the family = U U . U V U∗V { ∗ V | ∈U} The following result is [BL07, Lemma 3.1] (see also[BS07a, Proposition 9.6.1]):

PROPOSITION 7.29. Suppose that X is a metric space and let A, B X. Let = c and = ⊂ U c∈C U V c be covering families of A and B, respectively, which are both open in X and m-colored with m = c∈C V S C 1. If mesh LX (U,A) , then the family = c given by |S |≥ V ≤ 2 W c∈C W c c c c c − U − U := N LX ( ,A) ( ) V SN LX ( ,A) (U) V = for all U W 2 U ∗V ∪ ∈V | 2 ∩ ∅ ∈U   n o 92 7. ASYMPTOTIC DIMENSION is an open m-colored covering family of A B in X, with L ( , A B) min LX (U,A) ,L ( ,B) and ∪ X W ∪ ≥ { 2 X V } mesh max mesh , mesh .  W ≤ { U V} In the sequel we will denote the family constructed in Proposition 7.29 by = ⊛ . W W U V 6.3. The mesh and the Lebesgue number under quasi-homotheties. An important role in the original theorem of Buyalo and Lebedeva is played by [BL07, Lemma 3.4] (see [BS07a, Lemma 12.2.3]) which describes the behavior of the mesh and the Lebesgue number under λ-homotheties. This lemma generalizes to quasi-homotheties as follow:

LEMMA 7.30. Let h : A B be a map between metric spaces such that, for some λ 1, K 1 and for → ≥ ≥ some R> 0, we have

1 K RK (d (a ,a ))K d (h(a ),h(a )) λ √R K d (a ,a ), a ,a A. (6.3) λ · A 1 2 ≤ B 1 2 ≤ · A 1 2 ∀ 1 2 ∈ p Let C A, let be an open covering family of h(C) in h(A), and let = h−1( )= h−1(U) U . Then ⊆ U U U { | ∈ U} is an open covering family of C in A, h : A h(A) is bijective, and the following is true: U → (i) 1 RK e (mesh )K mesh λ K√R K√mesh , and e e e e λ · U ≤ U ≤ · U 1 K K K√ K (ii) λ R (LA( , C)) Lh(A)( ,h(C)) λ R LA( , C). · U ≤ e U ≤ · U PROOF OF LEMMA 7.30. First note that h is continuous,p so is a family of open sets in A, and e U there is a bijective correspondence between elements of and : each U from corresponds to U = U U U h−1(U) from , and h(U)= U, that is, h( )= . U U U For (i): e e e e e e mesh = sup diam U = sup sup dB(b1,b2) = sup sup dB(b1,b2) U e e e e e U∈U U∈U b1,b2∈U h(U)∈h(U) b1,b2∈h(U) (6.3) e e K K = sup sup dB(h(a1),h(a2)) sup sup λ √R dA(a1,a2) U∈U a1,a2∈U ≤ U∈U a1,a2∈U ·

K K K p = λ √R K sup diam U = λ √R √mesh . · U∈U · U r The other inequality for mesh is proven analogously. For (ii):

Lh(A)( ,h(C)) = inf Lh(A)( ,b) = inf sup distB(b,h(A) U) b∈h(C) b∈h(C) e e U U U∈U \

e = inf supe distB(h(a),h(A) h(U)) e h(a)∈h(C) h(U)∈h(U) \

= inf sup distB(h(a),h(A U)) = inf sup inf dB(h(a),h(x)) a∈C U∈U \ a∈C U∈U x∈A\U (6.3) K K K K inf sup inf λ √R dA(a, x)= λ √R inf sup distA(a, A U) ≤ a∈C U∈U x∈A\U · · a∈C U∈U \

K p K K p = λ √R K inf LA( ,a)= λ √R LA( , C). · a∈C U · U q p The other inequality for the Lebesgue number is proven analogously. 

7. Proof of Theorem 7.21 We now turn to the proof of Theorem 7.21. This will also conclude the proof of Theorem 7.18 and thereby conclude this chapter. Our proof is based on the corrected proof of [BL07, Thm. 1.1] (which is [BS07a, Thm. 12.2.1]), using the list of corrections provided to us by Nina Lebedeva in private correspondence.The authors wish to wholeheartedly thank Nina Lebedeva for sharing these corrections with them. 7. PROOF OF THEOREM 7.21 93

PROOF OF THEOREM 7.21. Let N := ℓ-dim X < . If dim Y = we are done, so let us assume ∞ ∞ that n := dim Y < . We have to show that N n. ∞ ≤ From ℓ-dim X = N, we know that there is a constant δ (0, 1) such that for every sufficiently small ∈ τ > 0 there exists an (N + 1)-colored open cover = N j of X with mesh τ and L( ) δτ. V ∪j=0V V ≤ V ≥ Since X is locally quasi-similar to Y , there exist constants λ 1 and K 1 such that for every ≥ ≥ sufficiently large R> 1 and every V X with diam V 1 , there is a map h : V Y such that ⊂ ≤ R V → 1 K RK(d (v , v ))K d (h (v ),h (v )) λ √R K d (v , v ), v , v V. (7.1) λ X 1 2 ≤ Y V 1 V 2 ≤ X 1 2 ∀ 1 2 ∈ Since mesh τ, and we would like to build a map with propertyp (7.1) for each V so we need V ≤ ∈ V diam V 1 , let us require that τ = 1 (thus for τ sufficiently small, R is sufficiently large). Now for ≤ R R every V , we fix a choice of such a map h : V Y once and for all. ∈V V → Using that Y is compact and dim Y = n, we can choose, for each j 0,...,N , a finite (n + 1)- n ∈ { } colored open cover of Y , say = c, such that the following hold: Uj Uj Uj c=0 [ e δK e e (i) mesh , and U0 ≤ 2K λ · 1 K2 K2 (ii) mesh ej+1 2 min L( j ) , mesh j , for every j =0,...,N 1. U ≤ 2K λ1+K · U U − ·      Notice that (i) ande (ii) imply: e e δK mesh , j =0,...,N. (7.2) Uj ≤ 2K λ ∀ · Also, using notation l := min L( ), mesh , notice that (7.2), (i) and (ii) imply j { Uj e Uj } δK lje e, j =0,...,N, and (7.3) ≤ 2K λ ∀ · 1 2 l (l )K , j =0,...,N 1, (7.4) j+1 ≤ 2K λ1+K2 · j ∀ − from which, considering that l 0,· 1 , it follows that j ∈ 2  1 lj+1 lj , j =0,...,N 1. (7.5) ≤ K√2 · ∀ − 1 K Also define the constant l := min L( ) j =0,...,N > 0. 2N−j Uj |     ′ We proceed by shrinking a little: for each Ve , consider the smaller subset V := N −δτ (V ). V ∈ V 2 Since L( ) δτ, we still have that V ′ V is an open cover of X. Put V := h (V ′). Now for V ≥ { | ∈ V} V every j 0,...,N , and every V j, we define the family ∈{ } ∈V e := U V U = . Uj,V { ∈ Uj | ∩ 6 ∅} This is a finite, open, at most (n + 1)-colored covering family of V in Y . Note that need not Uj,V e e e e e Uj,V be contained in h (V ) in general, so it is important to distinguish between and . V Uj,V Uj,V |hV (V ) Continuinge with the proof, one considers e e e e := h−1( )= h−1(U) U = h−1(U h (V )) U = h−1( ). Uj,V V Uj,V { V | ∈ Uj,V } { V ∩ V | ∈ Uj,V } V Uj,V |hV (V ) ′ Each j,V defines an at most (n + 1)-colored open covering family of V in V . U e e e e e e e e One can now apply Lemma 7.30 to the map h : V Y , with C = V ′, = and V → U Uj,V |hV (V ) = j,V , to obtain U U e e 1 K RK (mesh )K mesh λ √R K mesh , and (7.6) λ · Uj,V ≤ Uj,V |hV (V ) ≤ · Uj,V p 1 e K K RK (L ( , V ′))K L ( , V ) λ √R L ( , V ′). (7.7) λ · V Uj,V ≤ hV (V ) Uj,V |hV (V ) ≤ · V Uj,V q e e 94 7. ASYMPTOTIC DIMENSION

From the left side of (7.6) and from (6.2) it follows that 1 RK (mesh )K mesh , (7.8) λ · Uj,V ≤ Uj,V while from the right side of (7.7) and from Lemma 7.27, it folleows that L ( , V ) λ K√R K L ( , V ′). (7.9) Y Uj,V ≤ · V Uj,V q e e ′ Given a j 0,...,N , define X := j V . Then X j = 0,...,N form an open cover of ∈ { } j ∪V ∈V { j | } X. The family of sets

j := j,V U j U V[∈V is an open covering family of the set Xj in X. The main step of the proof is to establish the following properties of this family: (1) for each j 0,...,N , the family is at most (n + 1)-colored, ∈{ } Uj (2) mesh τ K√λ K mesh , j 0,...,N , Uj ≤ Uj ∀ ∈{ } q τ K δτ (3) L ( ,X ) min K (L( )) , , j 0,...,N , and X Uj j ≥ λ · e Uj 2 ∀ ∈{ } 1 (4) mesh j+1 LXn( j ,Xj ), j 0,...,No 1 . U ≤ 2 · U e∀ ∈{ − } The only thing to check for (1) is that within each of the colors c 0,...,n , the sets are disjoint. ∈{ c } c Indeed, if U1, U2 are elements of j of the same color c, i.e., U1,U2 , this means U1 , U ∈ Uj ∈ Uj,V1 U c V = V U = h−1(U ) U = h−1(U ) U , U c 2 j,V2 . If 1 2, then 1 V1 1 , 2 V1 2 , for some 1 2 j,V1 , but these are ∈ U c c ∈ U disjoint since j , as well as j,V1 , consists of disjoint elements, c = 0,...,n. If V1 = V2, then these U U je e ∀ e e e 6 two are disjoint as different elements of , so U1 and U2 are disjoint, too. e e V The inequality in (2) follows from(7.8) and the fact that mesh mesh : Uj,V ≤ Uj (7.8) K 1 1 (mesh j,V ) λ mesh j,V λe meshe j U ≤ · RK · U ≤ · RK · U

1 K K mesh √λ meshe , V j e ⇒ Uj,V ≤ R · Uj ∀ ∈V q K K mesh j = sup mesh j,V e τ √λ mesh j . ⇒ U j U ≤ · U V ∈V q e ′ For the statement in (3), note that, since elements V of Xj are disjoint, we get

LX ( j ,Xj ) = inf sup dist(x, X U) = inf inf sup dist(x, X U) j ′ U x∈Xj U∈Uj \ V ∈V x∈V U∈Uj,V \ = inf inf sup min dist(x, X V ), dist(x, V U) . j ′ V ∈V x∈V U∈Uj,V { \ \ } ′ δτ Since x V = N− δτ (V ) implies dist(x, X V ) , we have ∈ 2 \ ≥ 2 δτ LX ( j ,Xj ) inf inf sup min , dist(x, V U) U ≥ V ∈Vj x∈V ′ 2 \ U∈Uj,V   δτ = min , inf inf sup dist(x, V U) V ∈Vj x∈V ′ ( 2 U∈Uj,V \ )

δτ ′ = min , inf LV ( j,V , V ) . (7.10) 2 V ∈Vj U   On the other hand, Lemma 7.27 L( )= L ( , Y ) L ( , V )= L ( , V ), V j , (7.11) Uj Y Uj ≤ Y Uj Y Uj,V ∀ ∈V e e e e e e 7. PROOF OF THEOREM 7.21 95 where the last equality is true because the only sets U from that matter for y V are those that are Uj ∈ in . Uj,V Using (7.9) and (7.11), we may deduce that e e e e (7.11) (7.9) L( ) L ( , V ) λ K√R K L ( , V ′), V j Uj ≤ Y Uj,V ≤ · V Uj,V ∀ ∈V q K K ′ K ′ L( ej) λ √Re infe LV ( j,V , V )= λ √R K inf LV ( j,V , V ) ⇒ U ≤ · V ∈Vj U · V ∈Vj U q q e 1 1 K ′ (L( j )) inf LV ( j,V , V ). (7.12) ⇒ λK · R · U ≤ V ∈Vj U Therefore, (7.10) and (7.12) imply e (7.10) (7.12) δτ ′ δτ τ K LX ( j ,Xj) min , inf LV ( j,V , V ) min , (L( j )) , j =0,...,N. U ≥ 2 V ∈Vj U ≥ 2 λK · U ∀     Considering the proof of (4), we shall use (ii), (7.2), (2) and (3): e

(2) (ii) K K K 1 1 1 K K mesh j+1 τ √λ mesh j+1 τ √λ min (L( j )) , (mesh j ) U ≤ U ≤ · 2 · K√λ · λK · U U q 1 τ τ n o e = min (L( ))K , (meshe )K . e (7.13) 2 λK Uj λK Uj τ δτn o Using (7.2) it can be shown that (mesh )K , which,e when combinede with (7.13) yields λK Uj ≤ 2 1 τ δτ (3) 1 mesh min (L( ))Ke, L ( ,X ), j =0,...,N 1. Uj+1 ≤ 2 λK Uj 2 ≤ 2 X Uj j ∀ −   Note that in the calculations provinge (3) and (4), we have also proven: τ L ( ,X ) lK, for j =0,...,N, (7.14) X Uj j ≥ λK · j 1 τ mesh lK, for j =0,...,N 1, (7.15) Uj+1 ≤ 2 · λK · j − δτ and mesh , for j =0,...,N, (7.16) Uj ≤ 2 where (7.16) for j =0 follows from (i) and (2).

The final goal of the construction is to build an open (n + 1)-colored cover of X that will have U properties which guarantee that ℓ-dim X n. The way to build it is by induction in finitely many ≤ steps. First, put −1 := X , and note that L( −1) = inf sup dist(x, X U) = inf dist(x, ) = . U { } U x∈X b \ x∈X ∅ ∞ U∈U−1 Also put :=b , and assume that, for some bj 0,...,N 1 , we have already constructed families U0 U0 ∈{ − } ,..., so that the following properties are true, for indexes m 0,...,j (except (II) , which does U0 Uj ∈{ } 0 not makeb sense): b (I) b is an (n + 1)-colored open covering family of X ... X in X, m Um 0 ∪ ∪ m (II) mesh 1 L ( , m−2X ), m Um ≤ 4 · X Um−2 ∪i=0 i (III) meshb 1 L ( , m−1X ), m Um ≤ 2 · X Um−1 ∪i=0 i (IV) mesh δτ , b m Um ≤ 2 m b 1 m−1 (V)m LX ( m, i=0Xi) min LX ( m−1, i=0 Xi),LX ( m,Xm) , and Ub ∪ ≥ 2 U ∪ U m τ K (VI)m LX ( m, Xi) K nl . o Ub ∪i=0 ≥ λ · m b b 96 7. ASYMPTOTIC DIMENSION

Before proceeding, we should check steps m =0 and m =1, for the basis of induction. We leave to the reader to check (I)0 and (III)0 through (VI)0 (skipping (II)0, which is void). For m = 1: (II) is true because mesh 1 L ( ) = . Property (III) follows from (4): 1 U1 ≤ 4 X U−1 ∞ 1 mesh 1 L( ,X )= 1 L( ,X ). So now we apply Proposition 7.29 to produce := ⊛ . U1 ≤ 2 · U0 0 2 · U0 0 U1 U0 U1 Clearly (I)1 is true, while properties (IV)1 and (V)1 followb directly from conclusions of Proposition 7.29: L ( ,X X ) min 1 Lb ( ,X ),L ( ,X ) is exactly the statement of (V)b , whileb from X U1 0 ∪ 1 ≥ 2 X U0 0 X U1 1 1 mesh 1 max mesh 0, meshn 1 , (IV)0 and (7.16) we geto the statement of (IV)1. It remains to show U ≤b { U U } b (VI)1, which follows from (V)1, (VI)0, (7.5) and (7.14). First note: b b K 1 (VI)0 1 τ τ 1 (7.5) τ L ( ,X ) lK = l lK , 2 X U0 0 ≥ 2 · λK · 0 λK · K√ · 0 ≥ λK · 1  2  (7.14) τ bK and L ( ,X ) K l , so these two facts combined with (V) yield (VI) . X U1 1 ≥ λ · 1 1 1 For the step of induction, assuming, as above, that properties (I) – (VI) are true m j, let us m m ∀ ≤ prove these properties for index m = j +1. Start by checking (II)j+1: K (7.15) 1 τ (7.5) 1 τ 1 1 τ (VI)j−1 1 mesh lK l = lK L ( , j−1X ). Uj+1 ≤ 2 · λK · j ≤ 2 · λK · K√ · j−1 4 · λK · j−1 ≤ 4 · X Uj−1 ∪i=0 i  2  Now we can show (III)j+1, because from (II)j+1 and (4) we get: b

1 1 (V)j 1 mesh min L ( , j−1X ), L ( ,X ) L ( , j X ). Uj+1 ≤ 4 · X Uj−1 ∪i=0 i 2 X Uj j ≤ 2 · X Uj ∪i=0 i   Now, according to Proposition 7.29, web may produce := ⊛ . Clearlyb (I) is true, property Uj+1 Uj Uj+1 j+1 (V)j+1 is delivered directly by the conclusion about the Lebesgue number of Proposition 7.29, while from mesh max mesh , mesh , and propertiesb b (IV) and (7.16) for j + 1, we get the Uj+1 ≤ { Uj Uj+1} j statement of (IV)j+1. It remains to show (VI)j+1. First note: b b K (VI)j (7.5) 1 j 1 τ K τ 1 τ K LX ( j , Xi) l = lj l , 2 U ∪i=0 ≥ 2 · λK · j λK · K√2 · ≥ λK · j+1   (7.14) b τ K and L ( ,X ) K l . Therefore X Uj+1 j+1 ≥ λ · j+1 (V)j+1 1 τ L ( , j+1X ) min L ( , j X ),L ( ,X ) lK . X Uj+1 ∪i=0 i ≥ 2 X Uj ∪i=0 i X Uj+1 j+1 ≥ λK · j+1   b b After finitely many steps of induction, we obtain an open (n + 1)-colored cover := of X = U UN N X , which satisfies (IV) , that is, mesh = mesh δτ . Also, after applying property (V) N ∪i=0 i N U UN ≤ 2 times, as well as (3) and the fact that min min ai,bi = min min ai, min bi , we get b i∈I { } b { i∈I i∈I } (V ) 1 L( )= L( ) min L ( ,X ) j 0,...,N U UN ≥ 2N−j X Uj j | ∈{ }   (3) τ 1 1 δτ b min min (L( ))K , j 0,...,N ≥ λK · 2N−j Uj 2N−j · 2 | ∈{ }     τ 1 1 δτ = min min (L( e ))K j 0,...,N , λK · 2N−j Uj | ∈{ } 2N · 2     τ 1 δτ δτ 2l 1 = min l, = e min , . λK · 2N · 2 2 · δλK 2N     Since we can choose τ > 0 to be arbitrarily small, and the constants δ, λ, l, K and N are independent ′ 2l 1 ′ δ of τ, this shows that N = ℓ-dim X n: just put δ := min K , N and τ := τ, and now we can say ≤ δλ 2 2  7. PROOF OF THEOREM 7.21 97 that there exists δ′ in (0, 1) such that for all τ ′ > 0 small enough, we can find an open, (n + 1)-colored cover of X with mesh τ ′ and L( ) δ′τ ′.  U U ≤ U ≥

CHAPTER 8

Growth in non-elementary hyperbolic approximate groups

In this chapter we establish that non-elementary hyperbolic approximate groups have exponential internal growth. Note that the arguments usually employed in the group case do not work in this situation; in particular, non-elementary hyperbolic approximate groups do not have to contain non- abelian free subsemigroups. In our proof we distinguish two cases. For non-elementary hyperbolic approximate groups of asymptotic dimension 2 we can establish exponential growth via classical arguments comparing ≥ asymptotic dimension and critical exponent in quasi-cobounded hyperbolic spaces. These arguments, however, break down if the approximate group in question has asymptotic dimension 1. However, by assembling the results of previous sections, we are able to show that every non- elementary hyperbolic approximate group of asymptotic dimension 1 is quasi-isometric to a finitely- generated free group. This then allows us to reduce the case of asymptotic dimension 1 to the group case.

1. The case of asymptotic dimension 1 The goal of this section is to establish the following theorem; here, following [MSW00], a locally simplicial tree is called bushy, if every vertex is at uniformly bounded distance from a vertex which has at least 3 unbounded complementary components. THEOREM 8.1 (Non-elementary hyperbolic approximate groups of asymptotic dimension 1). Gi- ven a non-elementary hyperbolic approximate group (Λ, Λ∞), the following are equivalent: (i) asdim Λ = 1. (ii) ℓ-dim ∂Λ=0. (iii) dim ∂Λ=0. (iv) Some (hence any) representative of ∂Λ is totally disconnected. (v) Some (hence any) representative of ∂Λ is homeomorphic to a Cantor space. (vi) Some (hence any) representative of [Λ]int admits a quasi-isometric embedding into a simplicial tree. (vii) [Λ]int can be represented by a simplicial tree. (viii) [Λ]int can be represented by a locally finite bushy simplicial tree with uniformly bounded valences. (ix) Λ is quasi-isometric to a finitely-generated non-abelian free group. EXAMPLE 8.2. Examples of non-elementary hyperbolic approximate groups of asymptotic dimen- sion 1 can be constructed as uniform model sets in SL2(Qp), using an irreducible uniform S-arithmetic lattice in SL (Q ) SL (R). Indeed, any such model set acts geometrically on the Bruhat-Tits tree of 2 p × 2 SL2(Qp) and to see that there are no fixpoints at infinity one can e.g. use the fact that these model sets are Zariski-dense in SL2(Qp) by [BHS19]. Concerning the proof of Theorem 8.1, we note that the equivalences (i) (ii) (iii) are ⇐⇒ ⇐⇒ immediate from Theorems 7.1 and 7.18. In view of Corollary 5.42 the equivalence (iii) (iv) is a ⇐⇒ special case of the following classical result in dimension theory (see e.g. [Coo15, Thm. 2.7.1]): THEOREM 8.3 (Zero-dimensional compact metrizable spaces). If Z is a compact metrizable space, then dim Z =0 if and only if Z is totally-disconnected. The implication (iv) = (v) is immediate from Corollary 5.42 and the following classical theorem ⇒ of Brouwer [Bro10]:

99 100 8. GROWTH IN NON-ELEMENTARY HYPERBOLIC APPROXIMATE GROUPS

THEOREM 8.4 (Brouwer). Every non-empty totally-disconnected perfect compact metrizable space is homeomorphic to a Cantor space.  Since the Cantor space has topological dimension 0 (see e.g. [Coo15, Thm. 2.7.1]) we have thus established the equivalence of the Conditions (i)–(v). As for the remaining implications, the implication (ii) = (vi) is a special case of Theorem 7.23 ⇒ above. The implication (vi) = (vii) = (viii) follows from the following construction: ⇒ ⇒ CONSTRUCTION 8.5 (Growing bushy trees). Let (Λ, Λ∞) be a geometrically finitely-generated in- finite approximate group, let d be an internal metric on Λ and let ( , d ) be a simplicial tree. We T T assume that we are given a quasi-isometric embedding ϕ : (Λ, d) ( , d ). 0 → T T Since all edges in have length 1, the isometric embedding V ֒ of the vertex set of has T → T T relatively dense image, so it is a quasi-isometry. Composing ϕ0 with a quasi-inverse of this quasi- isometry we obtain a quasi-isometric embedding ϕ : (Λ, d) (V, d ). We denote by V the → T |V ×V 0 image of ϕ; since (Λ, Λ∞) is geometrically finitely-generated, (Λ, d) is coarsely connected, and hence V is C-coarsely connected for some C 0. 0 ≥ Now let denote the convex hull of V . Since can be obtained from V by connecting each T0 0 T0 0 vertex in V to its nearest neighbors in V by a geodesic segment of length C, it follows that is at 0 ≤ T0 Hausdorff distance at most C from V0. Consequently, the map ϕ induces a quasi-isometry (denoted by the same letter) ϕ : (Λ, d) ( , d ) and ( , d ) represents [Λ] . It remains to show that the tree → T0 T0 T0 T0 int is locally finite with uniformly bounded valences and bushy. T0 By Corollary 3.49, the space (Λ, d) has coarse bounded geometry, and hence ( , d ) has coarse T0 T0 bounded geometry by Proposition 3.47. In particular, ( , d ) is uniformly locally finite. We have thus T0 T0 established that (Λ, Λ∞) is quasi-isometric to a locally finite simplicial tree with uniformly bounded valences. Moreover, since (Λ, Λ∞) is non-elementary, ∂ is infinite. This implies that there exists at least one T0 vertex with at least 3 unbounded complementary components. Moreover, ∂ is quasi-cobounded, T0 and hence the set of all such vertices is relatively dense. This shows that is bushy and finishes the T0 proof. As pointed out in [MSW00, Sec. 2.1], any two locally finite bushy simplicial trees with uniformly bounded valences are quasi-isometric to each other, hence quasi-isometric to (the Cayley tree of) a finitely-generated non-abelian free group. This shows that (viii) = (ix). Since the boundary of a ⇒ finitely-generated non-abelian free group is homeomorphic to a Cantor space, we also deduce that (ix) = (v), and hence we have established Theorem 8.1. We record the immediate consequences for ⇒ later use: COROLLARY 8.6. Let (Λ, Λ∞) be a non-elementary hyperbolic approximate group of asymptotic dimension 1 without fixpoints at infinity. Then any apogee of Λ has infinitely many ends.  COROLLARY 8.7. Let (Λ, Λ∞) be a non-elementary hyperbolic approximate group of asymptotic dimension 1 without fixpoints at infinity. Then Λ has exponential internal growth.  REMARK 8.8. If Γ is a non-elementary hyperbolic group of asymptotic dimension 1, then it follows from Theorem 8.1 that Γ quasi-acts geometrically on a locally finite bushy simplicial tree with uni- formly bounded valences. By the celebrated QI rigidity theorem of Mosher–Sageev–Whyte [MSW00], this qiqac is in fact quasi-conjugate to an isometric action on a (possibly different) locally finite bushy simplicial tree with uniformly bounded valences. This shows that every non-elementary hyperbolic group has asymptotic dimension 1 if and only if it admits a geometric isometric action on a locally finite bushy simplicial tree with uniformly bounded valences. In the approximate group case, we only obtain a geometric qiqac, and it is unclear to us whether it is quasi-conjugate to an actual isometric action on a tree.

2. The case of higher asymptotic dimension In this section we are finally going to establish the following result: 2. THE CASE OF HIGHER ASYMPTOTIC DIMENSION 101

THEOREM 8.9 (Exponential growth). Every non-elementary hyperbolic approximate group (Λ, Λ∞) has exponential internal growth. For hyperbolic approximate groups of asymptotic dimension 1 this was established in Corollary 8.7. Thus for the rest of this section we are going to assume that (Λ, Λ∞) is a non-elementary hyperbolic approximate group with asdim Λ 2. We will work in the following setting: ≥ NOTATION 8.10 (General Setting). Throughout this section we fix a Bonk–Schramm realization L ∂Hn of ∂Λ so that in particular, ⊂ dim L = dim ∂Λ = asdim Λ 1 1. (2.1) − ≥ Then X := (L) Hn is a closed convex apogee for Λ with ∂X = (X) = L. We fix a left-regular H ⊂ L quasi-action ρ : (Λ, Λ∞) QI(X) with ρ(e)=Id. → We also denote by pr : Hn X the nearest point projection and fix a basepoint o X. We X → ∈ then denote by ι :Λ X, λf ρ(λ)(o) the associated orbit map. Since ρ is geometric, the map ι is a → 7→ quasi-isometric embedding with respect to some (hence any) internal metric on Λ. We denote by := ρ(Λ).o the image of ι (i.e. the quasi-orbit of o), and given λ Λ we set O ∈ λ := d(o, ι(λ)). k k Finally we abbreviate B := λ Λ λ R and A := B B so that Λ= A . R { ∈ |k k≤ } R R+1 \ R R REMARK 8.11 (Exponential growth rate). Since d is an internal metric on Λ, internalF exponential growth of Λ amounts to showing that 0 <δ(Λ) < , where ∞ 1 δ(Λ) := lim inf log B < R · | R| ∞ is the exponential growth rate of Λ. Since (Λ, d) is large-scale geodesic, it is quasi-isometric to a locally finite graph, and since it has bounded local geometry any such graph has bounded valencies. This implies that δ(Λ) < , and thus to establish exponential growth we only need to show that δ(Λ) > 0. ∞ DEFINITION 8.12. The growth series of Λ is defined as −αkλk Sα := e , λX∈Λ and the critical exponent δΛ is defined as δ := inf α S converges . Λ { | α } With this terminology at hand, we are going to establish the following more precise version of Theorem 8.9 (in the case of higher asymptotic dimension):

THEOREM 8.13 (Asymptotic dimension and exponential growth rate). Under our standing assump- tion that asdim Λ 2 we have ≥ δ(Λ) δ asdim Λ 1 1. ≥ Λ ≥ − ≥ In particular, δ(Λ) > 0, i.e. Λ has internal exponential growth. The proof of the first inequality is standard:

LEMMA 8.14 (Exponential growth rate vs. critical exponent). We have δ(Λ) δ . ≥ Λ PROOF. Let ǫ> 0 and α := δ(Λ) + ǫ. Since 1 1 α> lim inf log B + ǫ = lim inf log B + ǫ, R +1 · | R+1| R · | R+1| we can find R such that for all R > R we have αR log B + ǫR/2. For all R > R and all 0 0 ≥ | R+1| 0 λ A = B B we then have ∈ R R+1 \ R e−αkλk e−αR B −1 e−ǫR/2 A −1 e−ǫR/2. ≤ ≤ | R+1| · ≤ | R| · 102 8. GROWTH IN NON-ELEMENTARY HYPERBOLIC APPROXIMATE GROUPS

This implies

∞ R1 ∞ S = e−αkλk e−αkλk + A −1 e−ǫR/2 α ≤ | R| · RX=0 λX∈AR RX=0 λX∈AR R=XR1+1 λX∈AR ∞ B + e−ǫR/2 < . ≤ | R+1| ∞ R=XR+1 This shows that S converges for every α>δ(Λ), and thus δ(Λ) δ .  α ≥ Λ In order to prove the final inequality δ asdim Λ 1, (2.2) Λ ≥ − we will interpret the right-hand side as a lower bound on the Hausdorff dimension of the conical limit set of Λ.

REMARK 8.15 (Conical limit set). Since ρ is geometric, it is convex cocompact and thus, by Propo- sition 6.34 every limit point of Λ (or ) in ∂X is conical. On the other hand, since is relatively dense O O in X and the latter is a closed convex subset of Hn, we have (Λ) = (X)= L. This shows that if we L L consider Λ as a subset of Hn, then con(Λ) = (Λ) = L. (2.3) L L REMARK 8.16 (Hausdorff dimension). Let E be a subset of a metric space (Z, d). Given α> 0 and δ [0, ] we define ∈ ∞ ∞ ∞ Hδ (E, d) := inf rα x E, r (0,δ): E B(x , r ) . α  j | ∃ j ∈ j ∈ ⊂ j j  j=1 j=1 X [  δ ∞ Then Hα(E, d) := limδ→0 Hα(E, d) is called the α-dimensional Hausdorff measure ofE and Hα (E, d) is called the α-dimensional Hausdorff content of E. With this notation understood, the Hausdorff dimension of E (with respect to d) can be characterized as dim (E, d) := inf α H (E, d)=0 = inf α H∞(E, d)=0 = inf α H (E, d) < . Haus { | α } { | α } { | α ∞} Each Hausdorff measure is monotone and σ-additive on Borel sets [Rog70, Theorem 27], and hence if a Borel set E can be written as a countable union of Borel sets EM , then we have

dimHaus(E, d) = sup dimHaus(EM , d). M

PROPOSITION 8.17 (Hausdorff dimension of the conical limit set). If d is any metric on ∂Hn which induces the usual topology, then dim ( con(Λ), d) asdim Λ 1. Haus L ≥ − PROOF. By Theorem 7.1 we have dim L = dim ∂Λ = asdim Λ 1. − Now L is a separable metrizable space, and thus for L the topological dimension dim L coincides with the small inductive dimension ind(L) by [Eng95, Thm. 1.6.4 and Thm. 4.1.3]. Finally, as explained in [HW41, Section VII.4], it follows from [HW41, Thm. VII.2] that ind(L) (which unfortunately is denoted dim L in loc. cit.) is bounded above by the Hausdorff dimension dimHaus(L, d). In view of Remark 8.15 this establishes the proposition. 

We have thus reduced the proof of Theorem 8.13 to the following statement:

THEOREM 8.18 (Critical exponent and Hausdorff dimension). We have δ dim ( con(Λ)). Λ ≥ Haus L 2. THE CASE OF HIGHER ASYMPTOTIC DIMENSION 103

The remainder of this section is devoted to the proof of Theorem 8.18, which is essentially the same as the proof of [BJ97, Thm. 2.1]. Firstly, we need a lemma which allows us to provide upper bounds on Hausdorff dimension:

LEMMA 8.19 (Upper dimension bound). Let (X, d) be a metric space and let α> 0. Assume that there exists a sequence (x ) in X and a sequence (r ) of positive real numbers such that each x X is contained in n n ∈ infinitely many of the balls B(x , r ) and such that rα < . Then dim (X) α. n n n ∞ Haus ≤ PROOF. Since every x X is contained inP infinitely many balls, the collection of balls N := ∈ U (B(x , r )) covers X for every N N. By definition of the α-dimensional Hausdorff content we n n n≥N ∈ thus have ∞ H∞(X) C := rα. α ≤ N n nX=N Since C is finite we have C 0 and thus H∞(X)=0, which implies the lemma.  1 n → α As a second ingredient we need some estimates from hyperbolic trigonometry.

NOTATION 8.20. We are going to work in the Poincaré ball model of Hn; thus from now on Hn denotes the unit ball in Rn with the hyperbolic metric and we assume that the basepoint is given by o = 0. Given x Hn we denote by x the hyperbolic distance of x from 0 and by x [0, 1) the ∈ k k k kE ∈ Euclidean distance of x from 0 so that 1+ x x = 2arctanh( x ) = log k kE . k k k kE 1 x  −k kE  We then identify ∂Hn with the unit sphere in Rn and denote by d∠ the angular metric on ∂Hn. By definition, if η, ξ ∂Hn, then d∠(ξ, η) is the arclength of the shortest segment of a great circle which ∈ contains ξ and η.

LEMMA 8.21. For every C > 0 there exists M > 0 such that the following holds: If x Hn with C ∈ x > 1 and γ : [0, ) X is a geodesic ray in Hn emanating from 0 with d(γ, x) C, then k kE 2 ∞ → ≤ x x π d∠ ξ, M (1 x ) or d∠ ξ, . x ≤ C −k kE x ≥ 2  k kE   k kE  PROOF. Since the statement is invariant under isometries, we may assume without loss of gener- ality that n =2. Now let x θ := d∠ ξ, . x E π  k k  We will assume that θ < 2 . Note that, by definition, θ is just the hyperbolic angle between the rays through ξ and x. If C′ is the hyperbolic distance between x and its nearest point projection onto γ, then C′ C, and by basic hyperbolic trigonometry we have ≤ sinh(C′) sinh(C) sinh(C)(1 x 2 ) sin(θ)= = −k kE . sinh( x ) ≤ sinh(2arctanh( x )) 2 x k k k kE k kE 2 π 1 1 1−kxkE Since θ < we have sin(θ) θ and since x E we have 4(1 x E). We deduce that 2 ≥ 2 k k ≥ 2 kxkE ≤ −k k θ 2 sin(θ) < 4 sinh(C) (1 x ). ≤ · −k kE We may thus choose MC := 4 sinh(C). 

PROOF OF THEOREM 8.18. It will be convenient for us to work in the ball model of Hn; we will use the notation of Notation 8.20. Let ǫ> 0 and α := δ + ǫ; by definition of δ this means that S converges. Since d(0, x)= x = Λ Λ α k k log 1+kxkE , convergence of S is equivalent to the statement that 1−kxkE α   (1 ρ(λ)(0) )α < , (2.4) −k kE ∞ λX∈Λ 104 8. GROWTH IN NON-ELEMENTARY HYPERBOLIC APPROXIMATE GROUPS see [Nic89, Sec. 1.6]. For every x Hn 0 we denote by γ the unique geodesic ray from 0 through ∈ \{ } x x. We then define x π : Hn 0 ∂Hn, x γ ( )= . ∞ \{ } → 7→ x ∞ x k kE We extend this map to all of Hn by setting π (0) := (1, 0 ..., 0)⊤ and enumerate Λ= λ , λ ,... . For ∞ { 1 2 } every M N we then define ∈ ξ := π (ρ(λ )(0)) and r(M) := M(1 ρ(λ )(0) ). n ∞ n n −k n kE (M) n We then set := B(ξ , rn ) n N and denote by E the set of all ξ ∂H which are contained UM { n | ∈ } M ∈ in infinitely many elements of . We also define E := E . By (2.4) and Lemma 8.19 we then have UM M dim (E ) α for all M N and thus dim (E) α by Remark 8.16; letting ǫ 0 we therefore Haus M ≤ ∈ Haus ≤S → obtain dim (E) δ . Haus ≤ Λ To conclude the proof, it suffices to show that every conical limit point is contained in the set E. Thus let ξ be a conical limit point and let γ be a geodesic ray emanating from 0 which represents ξ. Then there exist C > 0 and infinitely many elements x1, x2,..., in the quasi-orbit ρ(Λ).o which are at distance < C from γ. Since the quasi-orbit is locally finite, we have

xn π 1 d∠ ξ, < and x for almost all n. x 2 k nkE ≥ 2  k nkE  If we now choose M := M as in Lemma 8.21, then the lemma shows that ξ E . This finishes the C ∈ M proof.  APPENDIX A

Background from large-scale geometry

1. Notations concerning (pseudo-)metric spaces In this section we fix our notation concerning metric spaces. In fact, it will be convenient for us to allow for a slight generalization of metric spaces called pseudo-metric spaces:

DEFINITION A.1. Let X be a set. A function d : X X [0, ) is called a pseudo-metric if for all × → ∞ x,y,z X the following hold: ∈ (i) d(x, x)=0, (ii) d(x, y)= d(y, x), (iii) d(x, z) d(x, y)+ d(y,z). ≤ In this case the pair (X, d) is called a pseudo-metric space. If (X, d) is a pseudo-metric space, then we obtain an equivalence relation on X by setting x ∼ ∼ y : d(x, y)=0. The pseudo-metric d then descends to a metric d¯ on the quotient space X¯ := X/ , ⇔ ∼ and we refer to (X,¯ d¯) as the metric quotient of (X, d). Our reason to consider pseudo-metric spaces is that they arise naturally in the context of isometric group actions:

EXAMPLE A.2. Assume that a group Γ acts by isometries on a metric space (X, d ) and let o X X ∈ be a basepoint. Then we obtain a left-invariant pseudo-metric d on Γ by setting d(g,h) := dX (g.o, h.o). In general, this pseudo-metric is not a metric. We will use the following notations concerning pseudo-metric spaces: If (X, d) is a pseudo-metric space, a point x X and R> 0, we denote by ∈ B(x, R) := x′ X d(x, x′) < R and B¯(x, R) := x′ X d(x, x′) R { ∈ | } { ∈ | ≤ } the open, respectively closed ball of radius R around x. Given A X we denote by N (A) the (open) ⊂ R R-neighborhood of A in X, i.e.

NR(A) := B(x, R). x[∈A The open balls B(x, R) form the basis for a topology τd on X, which is Hausdorff if and only if d is a metric. By abuse of language, if (X, τd) satisfies a topological property, we will also say that (X, d) has this property. For example, we say that (X, d) is connected if (X, τd) is. We say that (X, d) is proper if for some (hence any) o X the map x d(o, x) is proper. Equivalently, the balls B(o, R) are relatively ∈ 7→ compact for all R> 0 and some (hence any) o X. ∈ DEFINITION A.3. Let (X, d) be a pseudo-metric space, let A X and let R>r> 0. ⊂ (i) A is called r-uniformly discrete in X if d(x, y) > r for all x, y A with x = y. ∈ 6 (ii) A is called relatively dense in X if NR(A)= X. (iii) A is called (r, R)-Delone if it is r-uniformly discrete and R-relatively dense. In this case (r, R) are called Delone parameters for A. It is well-known that every pseudo-metric space contains a Delone set (cf. [CdlH16, Prop. 3.C.3]). There are other terms for the properties from the previous definition, especially for (ii): such a set A is often called quasi-dense, or coarsely dense.

105 106 A. BACKGROUND FROM LARGE-SCALE GEOMETRY

2. Coarse maps and coarse equivalences The following terminology follows [CdlH16, Chapter 3] (except for coarse equivalence, which is called metric coarse equivalence in [CdlH16], even among pseudo-metric spaces).

DEFINITION A.4. Let (X, d ), (Y, d ) be pseudo-metric spaces and f,f ′ : X Y be maps. X Y → (i) A non-decreasing function Φ : [0, ) [0, ) is called an upper control function. A non- + ∞ → ∞ decreasing function Φ : [0, ) [0, ] with lim Φ (t)= is called a lower control function. − ∞ → ∞ t→∞ − ∞ (ii) An upper control function Φ+ (respectively a lower control function Φ−) is an upper control for f (respectively a lower control for f) if d (f(x),f(x′)) Φ (d (x, x′)) (respectively d (f(x),f(x′)) Y ≤ + X Y Φ (d (x, x′))) for all x, x′ X. ≥ − X ∈ (iii) f is called coarsely Lipschitz (respectively coarsely expansive, a coarse embedding) if it admits an upper control (respectively a lower control, both an upper and a lower control). (iv) f is called coarsely proper if pre-images of bounded sets are bounded. (v) f is called essentially surjective or coarsely surjective if its image is relatively dense in (Y, dY ). It is called a coarse equivalence if it is a coarse embedding and essentially surjective. (vi) f and f ′ are close, denoted f f ′, if they are at uniformly bounded distance in the sense that ∼ ′ sup dY (f(x),f (x)) < . x∈X ∞ We denote by [f] the equivalence class of f under closeness. (vii) A map g : Y X is called a coarse inverse to f if [g f] = [Id ] and [f g] = [Id ]. → ◦ X ◦ Y We refer the reader to [CdlH16, Chapter 3] for basic properties of coarsely expansive and coarsely Lipschitz functions. The terminology varies a lot throughout the literature (cf. [CdlH16, Remark 3.A.4]). In [Roe03], coarsely Lipschitz maps between metric spaces are referred to as (uniformly) bornologous, while in [BDLM08] they are called large-scale uniform; in [BD08], coarsely expansive maps are called uniformly expansive maps. Coarsely Lipschitz maps and coarsely expansive maps can also be characterized without explicit mentioning of control functions (see [CdlH16, Prop. 3.A.5]):

LEMMA A.5. A map f : X Y between pseudo-metric spaces is coarsely Lipschitz if and only if for all → r 0, there exists s 0 such that, if x, x′ X satisfy d (x, x′) r, then d (f(x),f(x′)) s. ≥ ≥ ∈ X ≤ Y ≤ A map f : X Y between pseudo-metric spaces is coarsely expansive if and only if for all s 0, there → ≥ exists r 0 such that, if x, x′ X satisfy d (x, x′) r, then d (f(x),f(x′)) s.  ≥ ∈ X ≥ Y ≥ REMARK A.6 (Metric coarse category). Note that coarsely expansive (respectively coarsely Lips- chitz) functions are closed under composition, and that composition descends to closeness classes. We may thus define a category Coa0 as follows: Objects are non-empty pseudo-metric spaces, and mor- phisms from (X, d) to (Y, d) are closeness classes of coarsely Lipschitz maps. This category is called the metric coarse category in [CdlH16, Def. 3.A.7], and in this category the following hold ([CdlH16, Prop. 3.A.16]).

LEMMA A.7 (Morphisms in the metric coarse category). Let f : X Y be a coarsely Lipschitz map → between non-empty pseudo-metric spaces and [f] the corresponding morphism in Coa0.

(i) [f] is a monomorphism in Coa0 if and only if f is coarsely expansive. (ii) [f] is an epimorphism in Coa0 if and only if f is coarsely surjective. (iii) [f] is an isomorphism in Coa0 if and only if f is a coarse equivalence if and only if [f] is a monomorphism and an epimorphism.

Explicitly, (iii) states that a coarsely Lipschitz map f : X Y (with X = ) is a coarse equivalence → 6 ∅ if and only if there exists a coarsely Lipschitz map g : Y X which is a coarse inverse for f. →

DEFINITION A.8. Two pseudo-metric spaces (X, dX ) and (Y, dY ) are called coarsely equivalent if there exists a coarse equivalence f : X Y . → 3. QUASI-ISOMETRIES, BI-LIPSCHITZ MAPS AND ROUGH ISOMETRIES 107

By Lemma A.7, for non-empty pseudo-metric spaces, coarse equivalence is the same as isomor- phism in the category Coa0. In particular, coarse equivalence defines an equivalence relation on pseudo-metric spaces. An invariant of metric spaces will be called a coarse invariant if it is invari- ant under this equivalence relation. We denote by [(X, d)]c (or simply [X]c) the coarse equivalence class of the pseudo-metric space (X, d).

REMARK A.9. The above notions can be defined in the wider context of coarse spaces in the sense of Roe [Roe03] (cf. [CdlH16, 3.E]). Recall that if X is a set, then a non-empty collection of subsets E of X X is called a coarse structure (or a uniform bornology) on X and the pair (X, ) is called a coarse × E space if the following properties hold. (E1) ∆(X)= (x, x) x X , { | ∈ }∈E (E2) If E , and F E, then F , ∈E ⊂ ∈E (E3) If E , E , then E E , 1 2 ∈E 1 ∪ 2 ∈E (E4) If E , then E−1 , ∈E ∈E (E5) If E, F , then E F , ∈E ◦ ∈E where E−1 = (x′, x) (x, x′) E , E F = (x′, x′′) x X with (x′, x) E, (x, x′′) F . { | ∈ } ◦ { | ∃ ∈ ∈ ∈ } The elements of are then called controlled sets or entourages. A subset B X is called -bounded if E ⊂ E B B is controlled. × If (X, ) and (Y, ) are coarse spaces, then a map f : X Y is called proper if pre-images of EX EY → bounded sets under f are bounded, and bornologous if (f f)( ) . A proper bornologous map × EX ⊂EY is called a coarse map. Two coarse maps f,f ′ : X Y are considered equivalent, denoted f f ′, if → ∼ (f(x),f ′(x)) x X is controlled. Coarse spaces and equivalence classes of coarse maps form a { | ∈ } category Coa. (Let us note here that what we call a bornologous map, in [CdlH16, Def. 3.E.1] it is called a coarse map, i.e., their coarse maps do not have to satisfy the requirement of properness.) If (X, d) is a pseudo-metric space, then

d := E X X sup d(x, y) < E { ⊂ × | (x,y)∈E ∞} defines a coarse structure on X, called the coarse structure generated by d. The -bounded subsets Ed of X are just the d-bounded subsets. If (X, dX ) and (Y, dY ) are pseudo-metric spaces, then a map f : (X, ) (Y, ) is bornologous if and only if it is coarsely Lipschitz. On the other hand, while EdX → EdY f : (X, ) (Y, ) being coarsely expansive implies that f is proper, a proper map f need not be EdX → EdY coarsely expansive (see Remark 3.A.13 (2), [CdlH16]). Still, it is true that if two pseudo-metric spaces are coarsely equvialent, then their associated coarse spaces are isomorphic in the category Coa.

3. Quasi-isometries, bi-Lipschitz maps and rough isometries We will be particularly interested in coarse equivalences with (at most) affine control functions.

DEFINITION A.10. Let (X, d ), (Y, d ) be pseudo-metric spaces and f : X Y be a map. Given X Y → C, C′ 0 and K 1 we say that f is ≥ ≥ (i) a (K, C)-quasi-isometric embedding if for all x , x X, 1 2 ∈ 1 d (x , x ) C d (f(x ),f(x )) K d (x , x )+ C; K · X 1 2 − ≤ Y 1 2 ≤ · X 1 2 (ii) a (K,C,C′)-quasi-isometry if it is a (K, C)-quasi-isometric embedding whose image is, moreover, C′-relatively dense in Y ; (iii) a K-bi-Lipschitz embedding if it is a (K, 0)-quasi-isometric embedding; (iv) a K-bi-Lipschitz equivalence if it is a surjective bi-Lipschitz embedding; (v) a C-coarse isometric embedding if it is a (1, C)-quasi-isometric embedding; (vi) a (C, C′)-coarse isometry if it is a (1,C,C′) quasi-isometry; (vii) an isometric embedding if it is a (1, 0)-quasi-isometric embedding; (viii) an isometry if it is a surjective isometric embedding. 108 A. BACKGROUND FROM LARGE-SCALE GEOMETRY

REMARK A.11. If the precise constants do not matter, we simply refer to f as a quasi-isometric embedding, quasi-isometry etc. An equivalence class of quasi-isometries will be called a (K,C,C′)- quasi-isometry class if it admits a representative which is a (K,C,C′)-quasi-isometry, and we will use similar terminology for the other types of maps. A collection of quasi-isometric embeddings, quasi- isometries etc. (or equivalence classes thereof) is called uniform if the implied constants are uniformly bounded. If f : X Y is a (K , C )-quasi-isometric embedding and g : Y Z is a (K , C )-quasi- → 1 1 → 2 2 isometric embedding, then g f : X Z is a (K1K2,K2C1 + C2) quasi-isometric embedding. If ′ ◦ → ′ f is a (K1, C1, C1)-quasi-isometry and g is a (K2, C2, C2)-quasi-isometry, then f g is a (K1K2,K2C1 + ′ ′ ◦ C2,K2C1 + C2 + C2)-quasi-isometry. In particular, if A is a uniform collection of quasi-isometries from X Y and B is a uniform collection of quasi-isometries from Y to Z, then the set B A of all → ◦ compositions is a uniform collection of quasi-isometries. We say that (X, dX ) and (Y, dY ) are quasi-isometric (respectively bi-Lipschitz, coarsely-isometric, iso- metric) if there exists a quasi-isometry (bi-Lipschitz equivalence, coarse isometry, isometry) f : X Y . → If f is a (K,C,C′)-quasi-isometry, then f admits a coarse inverse f¯ which is a (K,KC +2KC′, KC + KC′)-quasi-isometry. It follows that quasi-isometry (and similarly bi-Lipschitzness, coarse isometry, isometry) is an equivalence relation on pseudo-metric spaces. Note that a coarse inverse for a quasi- isometry is usually referred to as quasi-inverse. In the sequel we will denote the equivalence class of (X, dX ) under quasi-isometry by [X, dX ] (or simply [X]) and refer to it as the QI type of X. Note that [X, dX ] = [X,¯ d¯X ], hence every QI type can be represented by a metric (as opposed to pseudo-metric) space.

REMARK A.12. As mentioned earlier, every pseudo-metric space (X, d) contains a Delone set A ⊂ X. Since A is relatively dense in X, the inclusion A ֒ X is a quasi-isometry. It follows that every → pseudo-metric space is quasi-isometric to a uniformly discrete metric space. In particular, every proper pseudo-metric space is quasi-isometric to a locally finite metric space.

REMARK A.13. Given a pseudo-metric space (X, dX ), we denote by QIK,C,C′ (X) the set of all (K,C,C′)-quasi-isometries from X to itself. We also set f QI(X) := QIK,C,C′ (X). K≥1 [′ f C,C ≥0f By the previous remark, QI(X) is a semigroup under composition, which contains the isometry group

Is(X) = QI1,0,0(X) as a subgroup. Moreover, the quotient QI(X) := QI(X)/∼ is a group with re- spect to composition of representatives,f and we refer to this group as the quasi-isometry group of X. Composingf the inclusion Is(X) QI(X) with the quotient map QI(X) f QI(X) we obtain a canon- → → ical homomorphism Is(X) QI(X). In general, the canonical homomorphism is neither injective nor → surjective. f f

4. Geodesics and quasi-geodesics in pseudo-metric spaces

DEFINITION A.14. Let (X, d) be a pseudo-metric space, K 1, C 0 and I R be connected. A ≥ ≥ ⊂ (K, C)-quasi-isometric embedding γ : I X is called a: → (i) (K, C)-quasi-geodesic segment if I is bounded; (ii) (K, C)-quasi-geodesic ray if I is bounded below; (iii) (K, C)-quasi-geodesic line if I = R. We say that γ is a quasi-geodesic segment, ray or line (or simply a quasi-geodesic) if it is a (K, C)-quasi- geodesic segment, ray or line for some K and C as above. If K = 1 we refer to γ as a C-rough geodesic segment, a C-rough geodesic ray or a C-rough geodesic, instead, and If K = 1 and C = 0, i.e., γ is an isometric embedding, then γ is referred to as a geodesic segment,a geodesic ray,ora geodesic line. 4.GEODESICSANDQUASI-GEODESICSINPSEUDO-METRICSPACES 109

Unless otherwise stated, rays will have domain [0, ). Note that by this definition, all geodesics ∞ are parametrized by arc length. We will sometimes need to reparametrize geodesics:

DEFINITION A.15. Let I R be connected. A map γ : I X is said to be a linearly reparametrized ⊂ → geodesic, if there exists a constant λ such that d(γ(t),γ(t′)) = λ t t′ for all t,t′ I. | − | ∈ We now state two corollaries of the Arzelà–Ascoli Theorem that will be useful in proofs about the Gromov and Morse boundaries.

LEMMA A.16 (Arzelà–Ascoli). Let X be a proper metric space and o X. Let a 0 b and ∈ n ≤ ≤ n let (β :[a ,b ] X) N be a sequence of geodesics in X. Assume that there exists C > 0 such that n n n → n∈ d(β (0), o) < C for all n N. n ∈ (i) If a and b , then (β ) has a subsequence which converges uniformly on compact sets to n → −∞ n → ∞ n n a bi-infinite geodesic line β : R X. → (ii) If a = 0 and b , then (β ) has a subsequence which converges uniformly on compact sets to a n n → ∞ n n geodesic ray β : [0, ) X. ∞ → It was observed by Morse [Mor24] that geodesics in negatively curved spaces have additional properties; one of these properties is nowadays called the Morse property and can be defined as follows:

DEFINITION A.17. Let X be a proper geodesic metric space, and let N : [1, ) [0, ) [0, ) ∞ × ∞ → ∞ be an arbitrary function. A quasi-geodesic γ : I X is called N-Morse if any (K, C)-quasi-geodesic → segment q with endpoints on γ is contained in the N(K, C)-neighborhood (γ) of γ. If γ is a NN(K,C) quasi-geodesic ray we obtain the notion of a N-Morse quasi-geodesic ray, and if γ is a quasi-geodesic line then we call γ an N-Morse quasi-geodesic line. If γ is an N-Morse quasi-geodesic, then we call N a Morse gauge for γ. Two important classes of metric spaces for us will be quasi-geodesic and geodesic metric spaces.

DEFINITION A.18. A pseudo-metric space (X, d) is called a (K, C)-quasi-geodesic space if for all x, y X there exists a (K, C)-quasi-geodesic segment γ :[a,b] X such that γ(a)= x and γ(b)= y. ∈ → If K = 1 it is called a C-roughly geodesic space, and if moreover C = 0 then it is called a geodesic space. We say that (X, d) is quasi-geodesic if it is (K, C)-quasi-geodesic, for some K and C as above.

REMARK A.19 (QI-invariance). Since the image of a geodesic under a quasi-isometry is a quasi- geodesic, the image of a geodesic metric space under a quasi-isometric embedding is a quasi-geodesic metric space. Moreover, if X is a relatively dense subset of a metric space Y , then X is quasi-geodesic if and only if Y is quasi-geodesic. Combining these two facts we see that being quasi-geodesic is QI-invariant among pseudo-metric spaces. In general, a subspace of a geodesic metric space need not be geodesic, or even quasi-geodesic.

DEFINITION A.20. Let (X, d) be a geodesic metric space. A subspace Y is called a quasi-convex subspace of X if there exists C > 0 such that every geodesic in X with endpoints in Y is contained in NC(Y ).

REMARK A.21. We note that if Y X is a quasi-convex subset of a geodesic metric space, then Y ⊂ is quasi-geodesic with respect to the induced metric. Indeed, assume that Y is R-quasi-convex for some constant Rl; we may then join any two points in Y by a geodesic in X in the R-neighborhood of Y . It follows that one can construct a (1, 2R + 1)- quasi-geodesic in Y with bounded Hausdorff distance 2R from the geodesic. Thus Y is quasi-geodesic where the quasi-constants depend only on R.

EXAMPLE A.22. If Γ is a connected graph with edge set and w : [0, ) is an arbitrary E E → ∞ function, then there exists a unique path metric dw on (the geometric realization of) Γ such that (the geometric realization of) any given edge e is isometric to an interval of length w(e). If w is the ∈ E 110 A. BACKGROUND FROM LARGE-SCALE GEOMETRY constant function 1, then we refer to d as the canonical metric on Γ. Each of the metric spaces ( Γ , d ) | | w is geodesic; in particular, every graph is a geodesic metric space with respect to its canonical metric. The canonical metric of a graph is proper if and only if the graph is locally finite; by contrast, every countable graph admits a proper metric of the form dw for some weight function w. We will see in the next section that, up to quasi-isometry, locally finite graphs (with their canonical metrics) are the only examples of proper geodesic metric spaces.

5. Coarsely connected and large-scale geodesic spaces

DEFINITION A.23. Let (X, d) be a pseudo-metric space. (i) Let n N and c > 0. A sequence (x ,...,x ) of points in X is called a c-path of length n from x ∈ 0 n 0 to x if d(x , x ) 0 such that for all x, x′ X there exists a c-path from ∈ x to x′. The terminology is justified by the following observation, see [CdlH16, Prop. 3.B.7]:

LEMMA A.24. A pseudo-metric space is coarsely connected if and only if it is coarsely equivalent to a connected metric space. In particular, coarse connectedness is a coarse invariant, hence a QI-invariant.  We compare the notion of being quasi-geodesic (Definition A.18) to the following terminology from [CdlH16]:

DEFINITION A.25. Let (X, d) be a pseudo-metric space. Moreover, let Φ+ be an upper control and let a,c > 0, b 0. ≥ (i) (X, d) is (Φ ,c)-coarsely geodesic if for all x, x′ X there exists a c-path from x to x′ of length + ∈ n Φ (d(x, x′)). ≤ + (ii) (X, d) is (a,b,c)-large-scale geodesic if it is (Φ+,c)-coarsely geodesic with respect to the affine con- trol Φ+(x)= ax + b.

We say that (X, d) is coarsely geodesic (respectively large-scale geodesic) if it is (Φ+,c)-coarsely geo- desic ((a,b,c)-large-scale geodesic), for some Φ+ and c as above (a,b,c as above). For the following lemma see again [CdlH16, Prop. 3.B.7 and Lemma 3.B.6(6)]:

LEMMA A.26 (Coarse connectedness properties). (i) Every large-scale geodesic space is coarsely geodesic, and every coarsely geodesic space is coarsely con- nected. (ii) A pseudo-metric space is coarsely geodesic if and only if it is coarsely equivalent to a geodesic metric space. (iii) A pseudo-metric space is large-scale geodesic if and only if it is quasi-isometric to a geodesic metric space. (iv) A proper pseudo-metric space is large-scale geodesic if and only if it is quasi-isometric to a locally finite graph (with its canonical metric). In particular, coarse geodesicity is a coarse invariant and large-scale geodesicity is a QI-invariant. 

REMARK A.27. The proof of (iv) is actually constructive: Indeed, let (X, d) be an (a,b,c)-large- scale geodesic pseudo-metric space. We construct a metric graph Γ0 as follows. The vertex set of Γ0 is simply X, and two vertices v , v X are connected by an edge if and only if they are distinct and 1 2 ∈ d(v , v ) c. Finally let w be the constant function on edges of Γ taking value c. Then by [CdlH16, 1 2 ≤ 0 Lemma 3.B.6(6)] the metric graph ( Γ , d ) is quasi-isometric to (X, d), albeit in general not locally | 0| w finite. Now assume that (X, d) is moreover proper. By Remark A.12 we can replace (X, d) by a quasi- isometric space (X′, d′) which is locally finite. Applying the above construction then yields a locally finite graph ( Γ , d ) which is quasi-isometric to (X′, d′) (and hence to (X, d)). In general, the metric | 0| w d will not be the canonical metric on Γ , but will differ from the latter by a factor c. However, this w | 0| can be corrected by subdividing each edge of Γ into c edges; then (X, d) is quasi-isometric to the 0 ⌊ ⌋ resulting graph Γ with its canonical metric. Note that Γ is locally finite, since Γ0 is. 6. SPACES OF COARSE BOUNDED GEOMETRY 111

LEMMA A.28. A metric space (X, d) is large-scale geodesic if an only if it is quasi-geodesic.

PROOF. The forward implication follows from Remark A.19 and Lemma A.26. For the reverse implication, assume that X is (K, C)-quasi-geodesic. Given x, x′ X we pick a ∈ (K, C)-quasi-geodesic segment γ :[a,b] X joining them. The points γ(a),γ(a + 1),...,γ(a + b → { ⌊ − a 1),γ(b) form a (K + C)-path of length n b a +1 b a +1. To show that X is large-scale ⌋− } ≤⌊ − ⌋ ≤ − geodesic, it suffices to find constants a > 0 and b 0, which are independent of the choice of x, x′, 0 0 ≥ such that n a d(x, x′)+ b . Since γ is a (K, C)-quasi-geodesic, we know that b a Kd(x, x′)+ KC. ≤ 0 0 − ≤ Thus we get n Kd(x, x′)+ KC +1.  ≤ The usefulness of large-scale geodesic metric spaces in coarse geometry is based on the fact that in large-scale geodesic spaces, coarse equivalences are quasi-isometries. This was first observed by Gromov [Gro93, 0.2.D], who refers to this fact as a “trivial lemma”.

LEMMA A.29. Let (X, dX ) be a large-scale geodesic metric space, (Y, dY ) an arbitrary metric space and f : X Y a map. Assume that there exists an upper control Φ : [0, ) [0, ) such that for all x , x X → + ∞ → ∞ 1 2 ∈ d (f(x ),f(x )) Φ (d (x , x )). Y 1 2 ≤ + X 1 2 Then there exist constants a 1, b 0 such that for all x , x X we have ≥ ≥ 1 2 ∈ d (f(x ),f(x )) a d (x , x )+ b. Y 1 2 ≤ · X 1 2 PROOF. Assume that (X, d ) is (a ,b ,c )-large-scale geodesic, and, given x, x′ X, let x = X 0 0 0 ∈ x ,...,x = x′ in X be such that d(x , x ) c , and let n a d (x, x′)+ b . Then 0 n i i−1 ≤ 0 ≤ 0 X 0 n d (f(x),f(x′)) d(f(x ),f(x )) n Φ (c ) Φ (c ) a d (x, x′)+Φ (c ) b . Y ≤ i i−1 ≤ · + 0 ≤ + 0 · 0 X + 0 · 0 i=1 X We may thus choose a := Φ+(c0)a0 and b := Φ+(c0)b0. This finishes the proof. 

COROLLARY A.30. Every coarse equivalence between large-scale geodesic spaces is a quasi-isometry, hence two large-scale geodesic metric spaces are coarsely equivalent if and only if they are quasi-isometric.  While this symmetric version of Gromov’s trivial lemma is often sufficient in applications, we will also need the asymmetric version of Lemma A.29 in the proof of our generalized Milnor–Schwarz lemma. Moreover, it will be important to us to have also a uniform version of Gromov’s lemma at our disposal. Given metric spaces (X, dX ), (Y, dY ) and a collection A of coarse equivalences from X to Y , we say that collection A is uniform if there exist an upper control function Φ+ and a lower control function Φ , which are an upper, respectively lower control for every f A. Similarly, given − ∈ a collection B of quasi-isometries from X to Y , we say that collection B is uniform if the elements of B have uniformly bounded QI-constants. With this terminology understood we have:

PROPOSITION A.31. Let (X, dX ), (Y, dY ) be large-scale geodesic metric spaces. Then every uniform collection of coarse equivalences from X to Y is a uniform collection of quasi-isometries from X to Y .

PROOF. For the upper bound we argue as in the proof of Lemma A.29 and observe that the QI constants a := Φ+(c0)a0 and b := Φ+(c0)b0 depend only on the upper control function and the con- stants a0,b0,c0, which depend only on X. For the lower bounds we apply the same argument to a quasi-inverse of the coarse equivalence in question. 

6. Spaces of coarse bounded geometry The following terminology was introduced by Block and Weinberger [BW92]:

DEFINITION A.32. Let (X, d) be a pseudo-metric space, let R 0 and let Φ be an upper control ≥ + function. A subset Λ X is Φ -uniformly locally finite if Λ B(x, r) Φ (r) holds for all x X and ⊂ + | ∩ |≤ + ∈ r > 0. This Λ is called a (R, Φ+)-quasi-lattice if it is moreover R-relatively dense. The space (X, d) is called of coarse bounded geometry if it contains a quasi-lattice. 112 A. BACKGROUND FROM LARGE-SCALE GEOMETRY

It turns out that having coarse bounded geometry is a coarse invariant:

LEMMA A.33. Let f : X Y be a coarse equivalence with control functions (Ψ , Ψ ) and C-relatively → − + dense image and let Λ X be a (R, Φ )-quasi-lattice. Then f(Λ) is a (Ψ (R)+ C, Φ )-quasi-lattice with ⊂ + + ++ upper control function Φ++ given by Φ (r) := Φ (Ψ (r + C)), where Ψ (s) := sup t [0, ) Ψ (t) s . ++ + − − { ∈ ∞ | − ≤ } PROOF. Since Λ is R-relatively dense in X, the set f(Λ) is Ψ (R)-relatively dense in f(X) and b b + thus Ψ+(R)+ C relatively dense in Y . Now let y f(Λ) B(y , r), for some y Y and r > 0. There then exists x X such that ∈ ∩ 0 0 ∈ ∈ d(f(x),y ) < C and thus y B(f(x), r + C). Choose λ Λ such that y = f(λ). Then 0 ∈ ∈ Ψ (d(λ, x)) d(f(λ),f(x)) r + C = d(λ, x) Ψ (r + C). − ≤ ≤ ⇒ ≤ − We deduce that b f(Λ) B(y , r)= f(Λ B(x, Ψ (r + C))) = f(Λ) B(y , r) Φ (Ψ (r + C)), ∩ 0 ∩ − ⇒ | ∩ 0 |≤ + − and the lemma follows.  b b REMARK A.34 (Growth equivalence of functions). Let f,g : [0, ) [0, ) be two functions. We ∞ → ∞ say that g dominates f and write f g if there are a,b 1 and c, d 0 such that for all r c we  ≥ ≥ ≥ have f(r) ag(br + d). In fact, it is always possible to choose d = 0 (at the cost of increasing b and ≤ c). We write f g if f g and g f. This defines an equivalence relation, and the corresponding ≈   equivalence class [f] of f is called the growth equivalence class of f.

LEMMA A.35 (Growth of quasi-lattices). Let (X, d) be a metric space of coarse bounded geometry and let Λ , Λ X be two quasi-lattices. Then for any x , x X, the functions 1 2 ⊂ 1 2 ∈ γ (r) := B(x , r) Λ and γ (r) := B(x , r) Λ 1 | 1 ∩ 1| 2 | 2 ∩ 2| are growth-equivalent.

PROOF. Choose R and Φ such that Λ and Λ are (R, Φ )-quasi-lattices. Fix a choice of x , x + 1 2 + 1 2 ∈ X, then choose δ > d(x , x ) and let λ B(x , r) Λ , for some r > 0. Choose λ Λ such that 1 2 1 ∈ 1 ∩ 1 2 ∈ 2 d(λ , λ ) R. Then 1 2 ≤ d(λ , x ) d(λ , λ )+ d(λ , x )+ d(x , x ) < R + r + δ. 2 2 ≤ 2 1 1 1 1 2 Since λ B(λ , R), we obtain 1 ∈ 2 B(x , r) Λ B(λ , R), 1 ∩ 1 ⊂ 2 λ2∈B(x2,r[+R+δ)∩Λ2 and hence γ (r) γ (r + R + δ) Φ (R). This shows that γ γ , and similarly γ γ .  1 ≤ 2 · + 1  2 2  1 7. Coarse ends of metric spaces The purpose of this section is to discuss a coarse version (due to Àlvarez and Candel) of the notion of ends of a topological space. We start by recalling the classical notation; our notation is set up in such a way that the correspondence to the coarse version becomes particularly apparent. We briefly recall the definition of the space of ends of a topological space X.

CONSTRUCTION A.36 (Ends). Let X be a topological space. We denote by (X) the poset of K compact subsets of X, ordered by inclusion. Given K (X) we denote by the set of unbounded ∈ K UK connected components of X K, considered as a discrete topological space. ′ \ If K K are compact subsets of X, then we define ηK,K′ : K′ K by mapping every ′ ⊂ ′ U → U U ′ to the unique U with U U. With respect to the maps η ′ the family ( ) ∈ UK ∈ UK ⊂ K,K UK K∈K(X) is an inverse system and the space of ends of X can be defined as the inverse limit

(X) := lim K . E ← U 7. COARSE ENDS OF METRIC SPACES 113

We now introduce a coarse version of the space of ends for metric spaces, following Àlvarez and Candel [ALC18, Chapter 6]. We need the following notion:

DEFINITION A.37 (c-connected components). If (X, d) is a metric space and c> 0, then we can de- fine an equivalence relation on X by setting x y if there is a c-path from x to y. The equivalence ∼c ∼c classes of this equivalence relation are precisely the maximal c-connected subsets of X, and they are called the c-connected components of X.

CONSTRUCTION A.38 (Coarse ends). Let (X, d) be a metric space. We denote by (X) the poset B of bounded subsets of X, ordered by inclusion. Given c> 0 and B (X), we denote by the set ∈B Uc,B of unbounded c-connected components of X B, considered as a discrete topological space. ′ ′ \ Now let c >c> 0 and B B be bounded subsets of X. We define ηc,B,B′ : c,B′ c,B by ′ ⊂ ′ U → U mapping every U c,B′ to the unique U c,B with U U. Similarly, we define θc,c′,B : c,B ∈ U ∈ U ′ ⊂ ′ U → ′ by mapping every U to the unique U ′ such that U U . Then we obtain a Uc ,B ∈ Uc,B ∈ Uc ,B ⊂ commuting square

ηc,B,B′ o ′ (7.1) Uc,B Uc,B

θc,c′,B θc,c′,B′    ηc′,B,B′  ′ o ′ ′ . Uc ,B Uc ,B With respect to the maps η ′ , the family ( ) is an inverse system and we define c,B,B Uc,B B∈B(X) c(X) := lim c,B. E ← U The elements of (X) are called the c-ends of X. By (7.1) the maps θ ′ ′ induce continuous maps Ec c,c ,B θ ′ : (X) ′ (X), and with respect to these maps the family ( (X)) is a direct system. We c,c Ec → Ec Ec c>0 may thus define ∞(X) := lim c(X); E → E the elements of (X) are called the coarse ends of X. E∞ The following is established in [ALC18, Prop. 6.35]:

THEOREM A.39 (Coarse invariance of coarse ends). Let X and Y be metric spaces and let f : X Y → be a coarsely proper coarsely Lipschitz map. Then f induces a continuous map f : (X) (Y ), which ∞ E∞ →E∞ depends only on the closeness class of f. In particular, the homeomorphims type of (X) is a coarse invariant E∞ of metric spaces.  While the space of coarse ends can be defined for an arbitrary metric space X and is a coarse invariant in this context, some assumptions on X are needed to make this space well-behaved. The following is [ALC18, Prop. 6.29]:

PROPOSITION A.40 (Compactness condition). If X is a coarsely connected metric space of coarse bounded geometry, then (X) is compact.  E∞ Our next goal is to compare the space (X) of coarse ends to the space (X) for proper metric E∞ E spaces X. We first observe that there is always a natural continuous map (X) (X): E →E∞ CONSTRUCTION A.41 (Comparison maps). Assume that (X, d) is a proper metric space and let o X be a basepoint. We abbreviate B := B(o, r); then the sequence B is cofinal both in (X) ∈ r n K and (X). Moreover, since for every c > 0 every connected component is contained in a unique B c-connected component, we have maps cn,c : . Passing to the inverse limit over n we UBn → Uc,Bn obtain a comparison map c : (X) (X), and then taking a direct limit over all c> 0 we obtain ∞,c E →Ec a comparison map c : (X) (X). ∞,∞ E →E∞ In [ALC18, Sec. 6.3] it is shown that the criterion applies, for example, to connected graphs and complete Riemannian manifolds. We extend this list as follows: 114 A. BACKGROUND FROM LARGE-SCALE GEOMETRY

PROPOSITION A.42 (Ends vs. coarse ends). If (X, d) is any proper geodesic metric space, then the comparison map yields a homeomorphism (X) = (X). E ∼ E∞ The proof is based on the following observation, which is similar to [ALC18, Prop. 6.31]:

LEMMA A.43. Let (X, d) be a proper metric space. Assume that there exists c 0 such that for all 0 ≥ c>c , there exists n (c) N such that for all n n there exists some m = m(c,n) N such that every 0 0 ∈ ≥ 0 ∈ coarse c-connected component of X B is contained in a unique connected component of X B and such \ n \ m that lim m(c,n)= for every n N. Then c : (X) (X) is a homeomorphism. n→∞ ∞ ∈ ∞,∞ E →E∞ PROOF. The condition ensures (for sufficiently large n and c) the existence of maps µn,c : Uc,Bn → , which induce maps µ∞,c : c(X) (X) and hence a map µ∞,∞ : ∞(X) (X), which UBm(c,n) E → E E → E can be shown to be inverse to the comparison map, as in the proofof[ALC18, Prop. 6.31]. 

PROOF OF PROPOSITION A.42. We show that Lemma A.43 applies with c0 := 0, n0(c) := 5c and m(c,n) := n 2c. Indeed, let c and n be chosen accordingly and let U be a coarse c-connected compo- − nent of X B . Let x, y U be two points of distance c and let γ be a geodesic segment in X which \ n ∈ ≤ connects x and y. Since the endpoints of γ lie outside B and γ has length c we have γ X B , where n ⊂ \ m m := n 2c. We deduce that x and y lie in the same connected component of X B . Consequently, − \ m U is contained in a single connected component of X B , and the proposition follows.  \ m It is well-known that if X is a Cayley graph of a finitely-generated group, then the set of ends of X (which coincides with the set of coarse ends by Proposition A.42) is homeomorphic to a Cantor space as soon as it contains at least three points. It is an interesting question in which generality this result holds. The following general criterion was established by Àlvarez and Candel [ALC18, Thm. 6.39]:

THEOREM A.44 (Álvarez-Candel criterion). If X is a quasi-cobounded large-scale geodesic metric space of coarse bounded geometry, then either (X) 2 or (X) is homeomorphic to a Cantor space.  |E∞ |≤ E∞ Combining Proposition A.42 and Theorem A.44 we obtain:

COROLLARY A.45. If X is a quasi-cobounded proper geodesic metric space of coarse bounded geometry, then either (X) 2 or (X) is a Cantor space.  |E |≤ E APPENDIX B

Some notions from geometric group theory

1. Left-invariant (pseudo-)metrics on groups A pseudo-metric d on a group G is called left-invariant if d(gh,gk) = d(h, k) for all g,h,k G. ∈ Left-invariant pseudo-metrics on groups correspond to pseudo-norms in the sense of the following definition (see e.g. [DS06]).

DEFINITION B.1. A pseudo-norm on a group G is a map : G [0, ) with the following k·k → ∞ properties (N1) e =0, k k (N2) g = g−1 for all g G, k k k k ∈ (N3) gh g + h for all g,h G. k k≤k k k k ∈ It is called a norm if instead of (N1) we have (N1’) g =0 if and only if g = e. k k Given a pseudo-norm we define d : G G [0, ) by d (g,h) := g−1h , and given a k·k k·k × → ∞ k·k k k pseudo-metric d : G G [0, ) we define : G [0, ) by g := d(e,g). With these notations × → ∞ k·kd → ∞ k kd understood, the following lemma follows straight from the definitions:

LEMMA B.2. The maps d and d define mutually inverse bijections between the set 7→ k · kd k · k 7→ k·k of left-invariant pseudo-metrics on G and the set of pseudo-norms on G. Under these bijections, left-invariant metrics correspond to norms.  If G carries a topology, we are interested in left-invariant metrics on G which to some extent are compatible with the given topology.

DEFINITION B.3. Let G be a topological group. A pseudo-metric d on G is called (i) locally bounded if every point in G has a d-bounded open neighborhood; (ii) left-adapted if it is left-invariant, proper (i.e. closed balls are compact) and locally bounded (i.e. every point has a neighborhood of finite diameter); (iii) left-admissible if it is left-adapted and induces the given topology on G. Note that every continuous pseudo-metric is automatically locally bounded. We emphasize that, following [CdlH16], we do not require left-adapted pseudo-metrics to be continuous. Any topological group which admits a left-admissible metric is necessarily locally compact, and among locally com- pact groups, the existence of a left-admissible pseudo-metric is characterized by Struble’s theorem (see [CdlH16, Thm. 2.B.4]).

PROPOSITION B.4. A locally compact group G admits a left-admissible metric if and only if it is second- countable.  Throughout this book we refer to a locally compact second-countable group G as an lcsc group. To check that a given left-adapted metric on a lcsc group is left-admissible, it suffices to check its continuity (see [CdlH16, Prop. 2.9]):

LEMMA B.5. Every proper continuous metric on a locally compact space induces the given topology. In particular, every continuous left-adapted metric on an lcsc group is left-admissible. 

115 116 B. SOME NOTIONS FROM GEOMETRIC GROUP THEORY

Lcsc groups arise naturally as isometry groups of pseudo-metric spaces:

EXAMPLE B.6. Let (X, d) be a proper pseudo-metric space; then on the isometry Is(X, d) the topologies of pointwise convergence, of uniform convergence on compacta and the compact-open topology all coincide and define a lcsc topology on Is(X, d) [CdlH16, 5.B.5]. In this book we will al- ways consider Is(X, d) as a lcsc group with respect to this topology. It can be shown that the action of Is(X, d) on X is jointly continuous and proper [CdlH16, Lemma 5.B.4], In fact, by a theorem of Malicki and Solecki [CdlH16, Thm 5.B.14], every lcsc group can be realized as Is(X, d) for a proper metric space (X, d), hence we can think of lcsc groups simply as isometry groups of proper metric spaces. A particularly important class of examples of lcsc groups for our purposes are countable discrete groups. Specializing Proposition B.4 and Lemma B.5 to this context we obtain:

COROLLARY B.7. A discrete group Γ admits a left-admissible metric if and only if Γ is countable. More- over, for a metric d on a countable discrete group Γ the following are equivalent: (i) d is left-invariant and proper; (ii) d is left-adapted; (iii) d is left-admissible.  While left-admissible metrics on lcsc groups are far from unique, different left-admissible metrics share a number of important properties. For example, we will see in the next section that they all lie in the same coarse equivalence class. However, there are also a number of finer metric properties which are shared by all left-admissible metrics, and we now give an example of such a property. Recall from Definition A.3 the definition of a Delone set in a metric space. The following proposi- tion guarantees that Delone sets in lcsc groups with respect to a left-admissible metric admit a purely group theoretic characterization:

PROPOSITION B.8 (Characterization of Delone sets in lcsc groups, [BH18a]). Let G be a lcsc group, Λ G a subset and d a left-admissible metric. ⊂ (i) Λ is uniformly discrete in (G, d) if and only if the identity e G is not an accumulation point of Λ−1Λ. ∈ (ii) Λ is relatively dense in (G, d) if and only if it is left-syndetic. In particular, the property of being a Delone set in (G, d) is independent of the choice of d. 

DEFINITION B.9. A subset Λ of a lcsc group G is called a Delone set in G if it is a Delone set with respect to some (hence any) left-admissible metric on G.

2. The coarse class of a countable group With every countable discrete group and, more generally with every lcsc group we can associate a canonical coarse equivalence class of metric spaces. This construction is based on the following coarse uniqueness property of left-admissible metrics (see [CdlH16, Cor. 4.A.6]):

PROPOSITION B.10. Let d be a left-adapted pseudo-metric on a lcsc group G and let d′ be a left-adapted pseudo-metric on a closed subgroup H < G. Then the inclusion map (H, d′) ֒ (G, d) is a coarse embedding. In → particular, if d and d′ are left-adapted pseudo metrics on G, then (G, d) and (G, d′) are coarsely equivalent.  Let us spell out the special case of countable discrete groups, since it lies at the heart of our ap- proach:

COROLLARY B.11. Let Γ1 be a countable group and Γ2 < Γ1 be a subgroup. If d1 and d2 are left-admissible ,metrics on Γ and Γ respectively, then the inclusion (Γ , d ) ֒ (Γ , d ) is a coarse embedding. In particular 1 2 2 2 → 1 1 if d, d′ are two left-admissible pseudo-metrics on a countable group Γ, then the identity (Γ, d) (Γ, d′) is a → coarse equivalence. 

DEFINITION B.12. If G is a lcsc group, then the coarse class [G]c of G is the coarse equivalence class of the metric space (G, d), where d is some (hence any) left-admissible pseudo-metric on G. 3. THE CANONICAL QI CLASS OF A FINITELY-GENERATED GROUP 117

The definition applies in particular to the case of a countable discrete group Γ. In order to give an explicit representative of [Γ]c we need to construct an explicit left-admissible metric on Γ. This can be done by means of the following construction:

DEFINITION B.13. Let Γ be a countable group and S Γ be a symmetric subset. A function ⊂ w : S e [0, ) is called a weight function on S if it is proper and satisfies w−1(0) = e and ∪{ } → ∞ { } w(s)= w(s−1) for all s S. ∈ LEMMA B.14. Let S be a symmetric generating set of a countable group Γ and let w : S e [0, ) ∪{ } → ∞ be a weight function. Then

n g := inf w(s ) g = s s , s S k kS,w i | 1 ··· n i ∈ (i=1 ) X defines a norm on Γ, and the associated metric dS,w is left-admissible. In particular, (Γ, dS,w) represents [Γ]c.

PROOF. (N1) follows from w−1(0) = e , (N2) follows from symmetry of S and w and (N3) holds { } by construction. Properness of w implies that the associated metric is proper. 

If S is finite, then we may choose w := 1, that is, w(s)=1, s S. In this case, d := d is called ∀ ∈ S S,1 a word metric on Γ.

3. The canonical QI class of a finitely-generated group

With any generating set S of a lcsc group G we can associate a word metric dS = dS,1 by the same formula as in Lemma B.14. This metric is proper if and only if S is compact. In this case it is even left-adapted and large-scale geodesic [CdlH16, Prop. 4.B.4].

LEMMA B.15. If G is a compactly-generated lcsc group and S is a compact symmetric generating set of G, then the associated word metric dS is left-adapted and large-scale geodesic. 

Recall that a lcsc group G is called compactly generated if it admits a compact generating set S. By Lemma B.15, every compactly-generated lcsc group admits a left-adapted large-scale geodesic metric. While the word metric dS from Lemma B.15 is in general not continuous, it is always possible to find a continuous left-adapted large-scale geodesic metric ([CdlH16, Prop. 4.B.4]). For example, if G is a connected Lie group, then any left-invariant Riemannian metric on G is a large-scale geodesic left-admissible metric. This shows:

PROPOSITION B.16. Let G be a compactly-generated lcsc group. Then there exists a large-scale geodesic left-admissible metric d on G. 

In fact, Proposition B.16 can be turned into a characterization of compactly-generated lcsc groups:

PROPOSITION B.17 (Geometric characterizations of compactly-generated groups). Let G be a lcsc group. Then the following are equivalent: (i) G is compactly-generated. (ii) Every left-admissible metric on G is coarsely connected. (iii) G admits a coarsely connected left-admissible metric. (iv) G admits a large-scale geodesic left-admissible metric. (v) Every representative of [G]c is coarsely connected. (vi) Every representative of [G]c is coarsely geodesic. (vii) [G]c admits a large-scale geodesic representative.

If any one of these equivalent conditions holds, then any two large-scale geodesic representatives of [G]c are quasi-isometric. 118 B. SOME NOTIONS FROM GEOMETRIC GROUP THEORY

PROOF. The implication (i) (iv) follows from Proposition B.16. Using Lemma A.26 we also have ⇒ implications (iv) (iii) (v) (ii) and (iv) (vii) (vi) (v). To see that (ii) (i), observe that if d is ⇒ ⇒ ⇒ ⇒ ⇒ ⇒ ⇒ a left-admissible metric on G such that (G, d) is C-coarsely connected, then G is generated, then the compact set B(e, C) generates G, hence G is compactly-generated. This process yields the equivalence of conditions (i)-(vii), and the final statement is immediate from Gromov’s trivial lemma (Corollary A.30). 

DEFINITION B.18. The canonical QI-type of a compactly-generated lcsc group is [G] := X [G] Xis large-scale geodesic . { ∈ c | } Note that the canonical QI type of a compactly-generated lcsc group is indeed a QI type by Propo- sition B.17 and Lemma A.26. It follows from Lemma B.15 that if S is a compact generating set of a lcsc group G, then (G, dS ) is a representative of [G]. By Proposition B.16 the canonical QI type can always be represented by a continuous metric on G. We emphasize once more that the canonical QI type of a lcsc group is only defined if the group is compactly generated.

EXAMPLE B.19 (Examples of compactly-generated lcsc groups). The following are examples of compactly-generated lcsc groups. (i) Every connected lcsc group G is compactly-generated - any compact identity neighborhoodisa generating set. (ii) More generally, every lcsc group G with finitely many connected components is compactly- generated - any compact identity neighborhood which intersects all of the components is a gen- erating set. (iii) Isometry groups of large-scale geodesic proper metric spaces are compactly-generated if they act coboundedly; this is a consequence of the following proposition.

PROPOSITION B.20. Let (X, d) be a large-scale geodesic proper metric space and assume that G := Is(X, d) acts coboundedly on X. Then the following hold: (i) G is compactly generated (hence the canonical QI type [G] is well-defined). (ii) If S is any compact symmetric generating set of G, then every orbit map defines a quasi-isometry (G, d ) S → (X, d). In particular, (X, d) represents the canonical QI type [G] of G.

PROOF. (i) This is contained in [CdlH16, Prop. 5.B.10]. (ii) By [CdlH16, Prop. 5.B.10], the action of G on X is geometric in the sense of [CdlH16, Def. 4.C.1]. It thus follows from [CdlH16, Theorem 4.C.5] that if o X is an arbitrary basepoint, then the ∈ map d : G G [0, ) given by d (g,h) := d(g.o, h.o) defines a left-adapted pseudo-metric on G o × → ∞ o such that the orbit map (G, d ) (X, d) with respect to o is a quasi-isometry. Since d is left-adapted o → o we have (G, d ) [G] , and hence (X, d) [G] . Since (X, d) is large-scale geodesic, so is (G, d ) and o ∈ c ∈ c o hence we even have (G, d ) [G] and (X, d) [G]. Finally, if S is a compact generating set of G, then o ∈ ∈ the identical map (G, d ) (G, d ) is a coarse equivalence by Proposition B.10, and since both spaces o → S are large-scale geodesic (see Lemma B.15) it is even a quasi-isometry. The proposition follows.  Again, we are particularly interested in the case of countable discrete groups. For such groups, being finitely generated is also equivalent to the classical finiteness property (F1) (see e.g. [Geo08, Prop. 7.2.1]). We thus obtain the following version of Proposition B.17:

PROPOSITION B.21 (Geometric characterizations of finitely-generated groups). Let Γ be a countable group. Then the following are equivalent. (i) Γ is finitely-generated. (ii) Γ is of type F1, i.e. admits a classifying space with finite 1-skeleton. (iii) Every left-admissible metric on Γ is coarsely connected. (iv) Γ admits a coarsely connected left-admissible metric. (v) Γ admits a large-scale geodesic left-admissible metric. (vi) Every representative of [Γ]c is coarsely connected. 4. FREE GROUPS AND COMBINATORICS OF WORDS 119

(vii) Every representative of [Γ]c is coarsely geodesic. (viii) [Γ]c admits a large-scale geodesic representative.

If any one of these equivalent conditions holds, then any two large-scale geodesic representatives of [Γ]c are quasi-isometric. 

Recall from Lemma B.15 that if S is any finite symmetric generating set of a finitely-generated group Γ with associated word metric dS , then (Γ, dS ) represents the canonical QI class [Γ]. Since this QI class is large-scale geodesic , we know from Proposition A.26 that it can be represented by a locally finite graph. In fact, it is easy to construct such a graph explicitly:

EXAMPLE B.22 (Cayley graphs of finitely-generated groups). If Γ is finitely generated with finite symmetric generating set S, then the Cayley graph of Γ with respect to S is the locally finite graph Cay(Γ,S) with vertex set Γ and edges given by pairs (g,gs) with g Γ and s S. If we restrict the ∈ ∈ canonical metric of Cay(Γ,S) to the vertices then we recover the word metric dS . Since the vertex set of any locally finite graph is relatively dense with respect to the canonical metric, the isometric inclusion (Γ, d ) ֒ Cay(Γ,S) is a quasi-isometry. This shows that Cay(Γ,S) is a locally finite graph S → which represents [Γ].

We conclude this section by recalling the Milnor-Schwarz lemma for countable discrete groups; for this we need the notion of a geometric action.

REMARK B.23 (On the notion of geometric actions of groups). According to [CdlH16, Def. 4.C.1], an isometric action of a topological group G on a pseudo-metric space (X, d) is called (i) cobounded, if some (hence any) G-orbit is relatively dense in X; (ii) locally bounded, if for every g G and B X bounded there exists an open neighborhood U of g ∈ ⊂ such that UB X is bounded; ⊂ (iii) metrically proper, if g G d(x, g.x) R is relatively compact for all x X and R 0; { ∈ | ≤ } ∈ ≥ (iv) geometric, if it is cobounded, locally bounded and metrically proper. However, in many situations of interest, geometric actions admit a much simpler characterization. Assume for example that (X, d) is a non-empty proper metric space, and that G is a locally compact group acting continuously on X by isometries. (This includes the case where G is discrete.) Then the G action on X is automatically locally bounded, and it is cobounded if and only if it is cocompact (i.e. there exists a compact subset K X such that G.K = X), and metrically proper if and only if it is ⊂ proper (i.e. the map G X X X given by (g, x) (x, g.x) is proper). In particular, such an × → × 7→ action is geometric if and only if it is proper and cocompact. In the case of actions of discrete groups on proper metric spaces, this is the definition given in most books on geometric group theory.

LEMMA B.24 (Milnor–Schwarz lemma). Let (X, d) be a large-scale geodesic proper metric space and let Γ be a countable discrete group acting geometrically on X. Then Γ is finitely generated and (X, d) is a representative of the canonical QI class [Γ]. 

4. Free groups and combinatorics of words Throughout this section let r 2 be a natural number and S = a ,...,a be a set of cardinality ≥ { 1 r} r. We denote by F the free group over S. Given g F we denote by g := g the word length of r ∈ r k k k kS g with respect to S.

r NOTATION B.25 (Reduced products). If u,u ,...,u F e then we write u = u u provided 1 l ∈ r \{ } 1 ··· l u = u1 ul and u = u1 + + ul ; we then say that u is the reduced product of u1,...,ul. ··· k k k k ··· k k r Set S∗ = S S−1; then every g F e can be written uniquely as g = g g with g ,...,g S∗ ∪ ∈ r \{ } 1 ··· l 1 l ∈ and g = l; we then call (g ,...,g ) the reduced word representing g. Similarly, we say that the empty k k 1 l word is the reduced word representing e. 120 B. SOME NOTIONS FROM GEOMETRIC GROUP THEORY

r r DEFINITION B.26. Let u = u u and v = v v with l,n N, u ,...,u , v ,...,v S∗. If for 1 ··· l 1 ··· n ∈ 1 l 1 n ∈ some k 1,...,n l +1 we have ∈{ − } vk = u1,...,vk+l = ul, then we say that u occurs in v at position k. We then denote by #u(v) the number of such occurrences of u in v.

It will sometimes be convenient to also define #u(e) := 0 and #e(v)= v for u, v Fr. Note that 5 k k ∈ in the definition of # (v) occurrences of u in v may overlap, e.g. # 2 (a )=4. u a1 1

DEFINITION B.27. Let u,x,y F e with u 2. We say that w is a u-blocker for (x, y) if xwy ∈ r \{ } k k≥ is a reduced product and

#u(xwy) = #u(x) + #u(y) and #u−1 (xwy) = #u−1 (x) + #u−1 (y).

EXAMPLE B.28 (An unblockable word). Let u = a1a2 F2. We claim that there is no u-blocker −1 ∈ −1 for the pair (a1,a1 ) or in fact for any pair (x, y) where x ends in a1 and y starts with a1 . Assume for contradiction that w was such a blocker. Since xw is reduced and does not contain any copies of −1 −1 −1 a1a2 or a2 a1 overlapping w, we see that w has to start from either a1 or a2 . Moreover, every a1 in −1 −1 −1 w can only be followed by a1 or a2 , and every a2 can only be followed by a1 or a2 . But this means −1 that w has to end with either a1 or a2 . In the former case, wy is not reduced, and in the latter case we generate a copy of u−1. This is a contradiction. We are going to show that this issue does not exist in free groups of higher rank:

THEOREM B.29 (Existence of blockers). Assume that r 3 and let u F with u 2. Then there ≥ ∈ r k k ≥ exists a constant C > 0 with the following property: For all x, y F e there exists a u-blocker w for the u ∈ r \{ } pair (x, y) which satisfies u w < C . k k≤k k u The following proof of Theorem B.29 arose from discussions with Alexey Talambutsa. We are grateful to him for allowing us to include the proof in this book. Until the end of the proof of Lemma r B.35 let r 3 and fix u F e with l := u 2; we write u = u u with u ,...,u S∗ and ≥ ∈ r \{ } k k ≥ 1 ··· l 1 l ∈ l 2. ≥ LEMMA B.30 (Reduction to the equal length case). Theorem B.29 holds if for every (x, y) F 2 with ∈ r x = y = l 1 there exists a u-blocker of length u . k k k k − ≥k k PROOF. On the one hand, the property of w being a u-blocker for (x, y) depends only on the last (l 1) letters of x and the first (l 1) letters of y, hence we may assume that x , y l 1. Since − − k k k k ≤ − there are only finitely many words of length at most (l 1) this implies in particular that the sizes of − blockers are uniformly bounded (by some constant Cu depending on u), provided such blockers exist. On the other hand, if x or y have length less than (l 1), then we can prolong them to words x′ − and y′ of length (l 1) by adding letters to the left of x and the right of y; every u-blocker for (x′,y′) is − then automatically a blocker for (x, y). 

From now on we will only consider elements x, y F with x = y = l 1 < u . Note that ∈ r k k k k − k k this assumption implies that

#u(x) = #u−1 (x) = #u(y) = #u−1 (y)=0, and thus w is a u-blocker for (x, y) if and only if

#u(xwy) = #u−1 (xwy)=0. To establish the existence of such a blocker we consider the following class of graphs which is a variant of the classical De Bruijn graphs [dB46] and was studied by Reiner Martin in his PhD thesis [Mar95]: 4. FREE GROUPS AND COMBINATORICS OF WORDS 121

DEFINITION B.31. Let l 2. The (oriented, labelled) l-th De Bruijn–Martin graph Γ of F is the ≥ l,r r graph with vertex set V := w Fr w = l 1 and set of oriented edges E~ := u Fr u = l . r { ∈ |k k − } { ∈ |k k } If u = u u with u ,...,u S∗, then the edge u is considered as an edge from the vertex u u 1 ··· l 1 l ∈ 1 ··· l−1 to the vertex u u and labelled by ℓ(u) := u . Moreover, the pair of edges (u,u−1) E~ 2 is called an 2 ··· l l ∈ inverse edge pair. To relate these graphs to the problem at hand, we use the following terminology:

DEFINITION B.32. Let y be a vertex of Γl,r. We say that a vertex v is a u-acceptance state for y if the product vy is reduced with #u(vy) = #u−1 (vy)=0. We then have the following reformulation of Theorem B.29

PROPOSITION B.33 (Graph theoretic formulation). Theorem B.29 holds provided for all vertices x, y of Γ there exists an oriented edge path of length l in Γ from x to a u-acceptance state v for y which does not l,r ≥ l,r use either of the edges u or u−1.

PROOF. Assume that there is a path as in the proposition with labels given by w1,...,wN with N l. We define w := w w . By definition of Γ we then have w = N and xw is reduced and ≥ 1 ··· N l,r k k ends with v. We claim that w is a u-blocker for (x, y). For this we first note that since w ends in v and v is an acceptance state for y, the word xwy is reduced. It thus remains to show that neither u nor u−1 appears in xwy. Since w l, every occurrence of u or u−1 in xwy occurs either in xw or in vy; the latter is k k ≥ excluded since v is an acceptance state for y. On the other hand, an occurrence of u or u−1 in xw means precisely that at some point the considered path from x to v crosses one of the two forbidden edges. This finishes the proof.  Proposition B.33 reduces the proof of Theorem B.29 to showing that acceptance states always exist and that they can be reached from any vertex, avoiding any given forbidden edge pair. This is established by the following two lemmas.

LEMMA B.34 (Acceptance states exist). For every vertex y of Γl,r there exists a u-acceptance state. l−2 PROOF. The total number of vertices of Γl,r is given by 2r(2r 1) . Now let v be a vertex; we r r − write v = v v and y = y y with v ,...,v ,y ,...,y S∗. Then v is not a u-acceptance 1 ··· l−1 1 ··· l−1 1 l−1 1 l−1 ∈ state if one of the following cases occurs: v = y−1 ; this is the case for (2r 1)l−2 vertices; • 1 l−1 − v v y equals u or u−1; this is the case for at most 2 vertices, namely v = u u and • 1 ··· l−1 1 1 ··· l−1 v = u−1 u−1; l ··· 2 v v y y equals u or u−1; in this case there are at most two choices for v v and thus at • 2 ··· l 1 2 2 ··· l most 2(2r 1) choices for v; − ... • v y y equals u or u−1; here there are 2(2r 1)l−2 choices for v. • l−1 1 ··· l−1 − The number α of u-acceptance states thus satisfies α 2r(2r 1)l−2 (2r 1)l−2 2 2(2r 1) 2(2r 1)l−2 ≥ − − − − − − −···− − (2r 1)l−1 1 = 2r(2r 1)l−2 (2r 1)l−2 2 − − − − − − (2r 1) 1 − − (2r 1)l−1 1 = (2r 1)l−2 2r 1 − − − · − − (2r 1)l−2(r 1)  − −  2r 1 1 > (2r 1)l−2 2r 1 − = (2r 1)l−2 2r 3 0. − · − − r 1 − · − − r 1 ≥  −   −  Thus α> 0 and the lemma follows.  122 B. SOME NOTIONS FROM GEOMETRIC GROUP THEORY

Note that the previous lemma even applies in the case r = 2. This is also the case for Part (i), but not for Part (ii) of the following lemma.

LEMMA B.35 (Strong connectedness of de Bruijn-Martin graphs). Let x, y be vertices of Γl,r. (i) There exists an oriented edge path of length l from x to y in Γ . ≥ l,r (ii) There exists an oriented edge path of length l from x to y in Γ which does not use either of the edges u ≥ l,r or u−1.

PROOF. (i) Connectedness of the unoriented De Bruijn–Martin graph was proved in [HT18, Lem- ma 2.10], and the argument actually shows that if x = y there always exists an oriented edge path from 6 x to y (and hence from y to x) of length 1. Concatenating such paths we can construct arbitrarily ≥ long edge paths from x to y for any two vertices x and y. r (ii) Since u = u u with u ,...,u S∗, the forbidden edges and their labels are given by 1 ··· l 1 l ∈ −1 u u u u l u u and u−1 u−1 1 u−1 u−1. 1 ··· l−1 −−→ 2 ··· l l ··· 2 −−−→ l−1 ··· 1 In view of (i) it suffices to show that there exist oriented edge paths from u u to u u and 1 ··· l−1 2 ··· l from u−1 u−1 to u−1 u−1 which bypass the forbidden edges u and u−1. l ··· 2 l−1 ··· 1 Since r 3 (and thus S∗ 6), we can find t,t′ S∗ such that ≥ | |≥ ∈ t u ,u−1,u−1 ,u ,u−1 and t′ u ,u−1,u−1,t−1 . 6∈ { 1 1 l−1 l l } 6∈ { 1 2 l } Since t = u−1 , there is an oriented edge path in Γ such that 6 l−1 l,r u u t u u t t ... t tl−1. 1 ··· l−1 −−→ 2 ··· l−1 −−→ −−→ −1 All of the edges in this path are labelled by t, which is different from ul and u1 , and hence avoid the two forbidden edges. Since t′ = t−1, we may also define an oriented edge path in Γ by 6 l,r ′ ′ ′ tl−1 t tl−2t′ t ... t (t′)l−1. −−→ −−→ −−→ All of these edges start from a vertex with initial letter t, and since t u ,u−1 , none of them can be 6∈ { 1 l } forbidden. Finally, since t′ = u−1 we may define an oriented edge path in Γ by 6 2 l,r u u u (t′)l−1 2 (t′)l−2u 3 ... l u u . −−→ 2 −−→ −−→ 2 ··· l All of these edges start from a vertex with initial letter t′, and since t′ u ,u−1 , none of them can 6∈ { 1 l } be forbidden. Concatenating the three paths above we obtain an edge path in Γ from u u l,r 1 ··· l−1 to u u which avoids the two forbidden edges. The argument for a bypass from u−1 u−1 to 2 ··· l l ··· 2 u−1 u−1 is similar.  l−1 ··· 1 At this point we have established Theorem B.29. r r REMARK B.36 (Cyclic words). Let u = u u and v = v v be elements of F with l,n N 1 ··· l 1 ··· k r ∈ and u ,...,u , v ,...,v S∗. 1 l 1 k ∈ We say that u is cyclically reduced if u−1 = u . This implies that un = n u , i.e. the powers of u lie l 6 1 k k k k on a (unique) geodesic of the Cayley tree of Fr with respect to S. If u and w are cyclically reduced, then we say that they are cyclically equivalent if there exists a cyclic permutation σ of 1,...,l such { } that w = u u . We refer to the equivalence class [u] of u as a cyclic word. σ(1) ··· σ(l) If u and v are cyclically reduced then we say that u cyclically occurs in w at position s if vs = u1, ..., v[s+1] = u2,..., v[k+l−1] = ul, where [m] 1,...,k denotes the residue of m modulo k. The number ∈{ } cyc of such cyclic occurrences depends only on [v], and hence will be denote by #u ([v]). If w Fr e is a word which is not necessarily cyclically reduced then we can write w uniquely r ∈ \{ } as w = tw t−1 with t, w F and v cyclically reduced. We then refer to w as the cyclic reduction of o o ∈ r o o w and to [w] := [wo] as the associated cyclic word. In particular, this allows us to define #u([w]) for a cyclically reduced word w and an arbitrary w F e . Finally, we also set #cyc([e]) := 0 for every ∈ r \{ } u cyclically reduced word u. 4. FREE GROUPS AND COMBINATORICS OF WORDS 123

LEMMA B.37 (Homogeneity). If u, w F e are cyclically reduced words, then we have #cyc([w]n)= ∈ r \{ } u n #cyc([w]) for all n N. · u ∈ r PROOF. Since W n = w w is a reduced product, we deduce that u cyclically occurs in w at posi- ··· tion s if and only if it occurs cyclically in wn at positions s,s + w ,...,s + (n 1) w . Conversely, k k − k k every cyclic occurrence of u in wn is of this form, and the lemma follows.  r REMARK B.38 (Normal forms). If u = u u with u ,...,u S∗, then we denote by 1 ··· l 1 l ∈ P (u) := u u k 1,...,l 1 and S(u) := u u k 2,...,l { 1 ··· k | ∈{ − }} { k ··· l} | ∈{ } the collections of proper prefixes, respectively proper suffixes of u. If v F e is another non-trivial ∈ r \{ } word, then we say that u and v do not overlap if P (u) S(v) = = P (v) S(u) and none of the two ∩ ∅ ∩ words is a proper subword of the other. We say that u is non-self-overlapping if u ⋔ u. We also say that (u, v) is an independent pair if u and v are non-self-overlapping with u = v and u ⋔ v. It was 6 established in [HS16, Lemma 2.5 and Cor. 2.6] that if u is cyclically reduced and non-self-overlapping, then (u,u−1) is an independent pair. Assume from now on that this is the case. Given w Fr e and g1,...,gm Fr e , we call (g1,...,gm) a decomposition of w if r ∈ \{r } ∈ \{ } w = g g . If g = a b for some j 1,...,m , then (g ,...,g ,a ,b ,...,g ) is another de- 1 ··· m j j j ∈ { } 1 j−1 j j m composition of w, called a simple refinement of (g1,...,gm). We call a decomposition of w a refinement of (g1,...,gm) if it can be obtained by a sequence of simple refinements. We say that a decomposition (g , ,g ) is u-maximal if the number of g with g u,u−1 is maximal. By [HS16, Lemma 5.13] 1 ··· m i i ∈ { } every w F e admits a unique u-maximal decomposition (g ,...,g ) which is not a refinement ∈ r \{ } 1 m of a u-maximal decomposition. We call (g1,...,gm) the u-decomposition of w. Note that in a u-maximal −1 −1 decomposition, every gj is either equal to u or equal to u or does not contain any copy of u or u ; moreover, every other element of the decomposition is either a copy of u or u−1. For example, if −1 −1 −1 −1 −1 −1 −1 −1 w = a1a2a1 a2 a1 a2a1 , then its a1a2-decomposition is (a1a2,a1 ,a2 a1 ,a2a1 ). The following observation will be needed in the proof of Theorem 6.22:

LEMMA B.39. Let u F be cyclically reduced and non-self-overlapping, and let x F be cyclically ∈ r ∈ r reduced with #cyc(x) > 0 and x u 2. Then there exists x′ F which is cyclically equivalent to x u k k≥k k ≥ ∈ r such that cyc cyc ′ ′ ′ − ′ #u (x ) = #u(x ) and #u−1 (x ) = #u 1 (x ). (4.1)

PROOF. Since #u(x) > 0, the word x admits a u-decomposition (g1,...,gn) such that gj = u for some j 1,...,n . Then x′ := g g g is cyclically equivalent to x and has u as a prefix, ∈ { } j j+1 ··· j−1 respectively suffix. To see that (4.1) holds for x′, assume that u (or u−1) occurs cyclically in x′ at position s in x′. Then s x′ u +1, since otherwise the occurrence of u (or u−1) would overlap ≤k k−k k −1 −1 with gj, contradicting the fact that (u,u ) is an independent pair. But then u (or u ) actually occurs in x′ at position s, and hence (4.1) holds.  We will also need the following consequence of Theorem B.29:

LEMMA B.40. Let r 3 and let u F with l := u 2. Then there exists a constant C with the ≥ ∈ r k k ≥ u following properties: if x′,y′ F with min x′ , y′ u , then there exist w,z F with ∈ r {k k k k}≥k k ∈ r u w C and u z C k k≤k k≤ u k k≤k k≤ u such that x′wy′z is reduced and cyclically reduced, and such that every copy of u or u−1 in (x′wy′z)n is contained in one of the n copies of x′ or in one of the n copies of y′.

PROOF. Let Cu be the constant from Theorem B.29. By the theorem we can find a blocker w for the pair (x′,y′) and a blocker z for the pair (y′, x′) of the desired lengths. Then x′wy′ and wy′z are reduced, and hence x′wy′z is reduced and cyclically reduced. Moreover, since all four factors of x′wy′z have length u , no copy of u or u−1 can overlap more than two of the factors, and then the blocker ≥ k k properties of w and z imply that every copy of u or u−1 in (x′wy′z)n is contained in one of the n copies of x′ or in one of the n copies of y′.  124 B. SOME NOTIONS FROM GEOMETRIC GROUP THEORY

5. Quasimorphisms and quasi-homomorphisms Quasimorphisms are generalizations of group homomorphisms which allow multiplicativity to be relaxed by a finite error. The following metric definition goes back to Ulam [Ula60], see also [FK16, BOT13].

DEFINITION B.41. Let G, H be groups, let d be a metric on H and C 0. Amap f : G H is ≥ → called a (d, C)-quasi-homomorphism if for all x, y G, ∈ d(f(xy),f(x)f(y)) C. (5.1) ≤ It is called a d-quasi-homomorphism if it is a (d, C)-quasi-homomorphism for some C 0. ≥ A metric-free way to express that multiplication fails by a small amount can be defined as follows.

DEFINITION B.42. Let G be a group, H a locally compact group and let f : G H be a map. Then → f is called a quasimorphism if its left-defect set D(f) := f(y)−1f(x)−1f(xy) x, y G (5.2) { | ∈ } is relatively compact in H.

REMARK B.43. A priori what we call a quasimorphism in Definition B.42 should be called a left- quasimorphism, and one should define a right-quasimorphism by demanding that the right-defect set D∗(f) := f(x)f(y)f(xy)−1 x, y G (5.3) { | ∈ } be relatively compact. However, we will mostly be interested in situations where either H is abelian (hence there is no point in distinguishing left and right) or where H is discrete. In the latter situation it was proved by Heuer in [Heu20, Prop. 2.3] the two notions are actually equivalent. The following remark clarifies the precise relation between metrically defined quasi-homomor- phisms and quasimorphisms.

REMARK B.44 (Quasimorphisms vs. quasi-homomorphisms). The notionofa d-quasi-homomor- phism f : G H clearly depends on the choice of metric d on H. Assume now that H is a countable → group, considered as a discrete topological space. Then H admits a left-invariant proper metric dl (see Proposition B.14) and similarly a right-invariant proper metric d . Then for a map f : G H we have r → the equivalences f is a d -quasi-homomorphism f is a (left-)quasimorphism l ⇔ f is a (right-)quasimorphism f is a d -quasi-homomorphism ⇔ ⇔ r

In particular, for countable targets, the notion of a quasimorphism is a special case of the notion of a quasi-homomorphism. Similar statements can be made for quasimorphisms with lcsc targets. How- ever, the definition of a quasimorphism makes sense without the assumption that H be countable (or lcsc). We warn the reader that some authors, including [FK16], reserve the term quasimorphism for quasimorphisms (in our sense) with abelian target. We do not follow this convention here. For later reference we record that if f : G H is a quasimorphism, then by definition for all x, y G, we have → ∈ f(xy)= f(x)f(y)(f(y)−1f(x)−1f(xy)) f(x)f(y)D(f) (5.4) ∈ and similarly f(x)f(y)= f(xy)(f(y)−1f(x)−1f(xy))−1 f(xy)D(f)−1. (5.5) ∈ In particular, we have f(x)f(x−1),f(x−1)f(x) f(e)D(f)−1. (5.6) { }⊂ 5. QUASIMORPHISMS AND QUASI-HOMOMORPHISMS 125

REMARK B.45 (Real-valued quasimorphisms). Let Γ be a group. We consider maps f : Γ R, → where R is considered as an additive group with metric given by the Euclidean metric d. (i) Every quasimorphism f : Γ R can be written as f = f + b, where f : Γ R, g → sym sym → 7→ f(g) f(g−1) is a symmetric quasimorphism and b is bounded. In this sense, every real-valued − quasimorphism is at bounded distance from a symmetric one. (ii) If Γ is a group and f is a real-valued function on Γ, then it is a quasimorphism if and only if d(f) := sup f(gh) f(g) f(h) < . g,h∈Γ | − − | ∞ In this case we have D(f) [ d(f), d(f)], and hence d(f) is called the defect of f. ⊂ − (iii) A real-valued quasimorphism f : G R is called homogeneous if f(gn)= n f(g) for every g Γ → · ∈ and n Z. Every real-valued quasimorphism f : Γ R is at bounded distance from a unique ∈ → homogeneous quasimorphism f :Γ R. Explicitly, we have → f(gn) e f(g) = lim , n→∞ n where the limit exists by Fekete’s lemma.e Every homogeneous quasimorphism is automatically symmetric and conjugation-invariant (see e.g. [Cal09, Sec. 2.2]).

EXAMPLE B.46 (Perturbed homomorphisms). If Γ is an infinite group, then there are plenty of real- valued quasimorphisms on Γ. Indeed, if f :Γ R is a homomorphism (possibly trivial), then f + b o → o is a quasimorphism for every function b : Γ R which takes finitely many values. Quasimorphisms → of this form are called perturbed homomorphisms. If Γ is amenable (as a discrete group), then every real-valued quasimorphism on Γ is a perturbed homomorphism, but for many non-amenable groups there are also plenty of exotic examples of real- valued quasimorphisms which are not perturbed homomorphisms. The most basic examples are counting quasimorphisms on free groups:

EXAMPLE B.47 (Counting quasimorphisms). Let Fr be a free group of rank r with free generating set S := a ,...,a ; we identify elements of F with reduced words over S S−1. If u F e is a { 1 r} r ∪ ∈ r \{ } non-trivial reduced word, then the function

ϕ : F Z, v # (v) # −1 (v) u r → 7→ u − u (with notation as in Definition B.26) defines a symmetric quasimorphism on Fr (see e.g. [Cal09, p. 22]). It is called the (big) u-counting quasimorphism. We will need two variants of this construction: firstly, the homogenization f : F R of ϕ is called the homogeneous u-counting quasimorphism.. Secondly, u r → u if u is cyclically reduced, then the function cyc cyc cyc ϕ : F Z, v # ([v]) # − ([v]) u r → 7→ u − u 1 is called the cyclic u-counting quasimorphism. Counting quasimorphisms were introduced by Brooks [Bro81], whereas cyclic counting quasi- morphisms were introduced by Fa˘ıziev [Fai94˘ ] and popularized by the paper [Gri95].

LEMMA B.48. Assume that u is cyclically reduced. Then cyc fu = ϕu . (5.7)

In particular, fu is a Z-valued homogeneous quasimorphism and fu(w) depends only on [w].

PROOF. We have to show that for every w F we have ∈ r cyc cyc f (w) = # ([w]) # −1 ([w]). (5.8) u u − u For w = e this holds by definition, thus assume that w = e. We first observe that both sides of (5.8) 6 remain unchanged if we replace w by its cyclic reduction, and that both sides multiply by n if we replace w by wn. We may thus assume that w is cyclically reduced and that w u . k kS ≥k kS 126 B. SOME NOTIONS FROM GEOMETRIC GROUP THEORY

In this situation, every occurrence of u in wn yields a cyclic occurrence of u in [wn], and every such cyclic occurrence of u yields an occurrence of u in wn+1. Applying the same argument to u−1 yields n n cyc n cyc n n+1 # (w ) # −1 (w ) # ([w ]) # −1 ([w ]) n +1 # (w ) # −1 n+1 u − u u − u u − u w . n ≤ n ≤ n · n +1 cyc cyc Now the expression in the middle equals # ([w]) # −1 ([w]), and the two outer expressions both u − u converge to fu(w). 

REMARK B.49 (Groups with many non-trivial quasimorphisms into R). It turns out that for ev- ery non-abelian free group, the space of finite real linear combinations of counting quasimorphisms modulo perturbed homomorphisms is infinite-dimensional; in fact, the only linear relations between equivalence classes of counting quasimorphisms are “the obvious ones” (see [HT18]). In general, if H is a subgroup of a group G, then not every quasimorphism f : H R can be → extended to a quasimorphism on G; in fact, not every group containing a free subgroup admits a non- trivial real-valued quasimorphism. However, if H is “hyperbolically embedded” in G, then every homogeneous quasimorphism on H extends [HO13]. It follows that if G is a group which contains a hyperbolically embedded non-abelian free group, then it has plenty of non-trivial quasimorphism; such groups are known as acylindrically hyperbolic groups [Osi16]. One can actually show that, conversely, every homogeneous quasimorphism on an acylindrically hyperbolic group arises as an extension of a quasimorphism on an hyperbolically embedded free subgroup (of rank 2), see [HS19]. We will use the following class of counting quasimorpisms to illustrate properties of conical limit sets.

NOTATION B.50 (Special counting quasimorphisms). Let S := a ,...,a be a set of cardinality { 1 r} r 2 and let F denote the free group over S. We denote by W the set of all cyclically reduced and ≥ r r non-self-overlapping words of length 2 in F and set ≥ r (F ) := ϕ : F Z u W and cyc(F ) := ϕcyc : F Z u W . Qo r { u r → | ∈ r} Qo r { u r → | ∈ r} REMARK B.51 (Properties of special counting quasimorphisms). Let u W and w F e with ∈ r ∈ r \{ } cyclic reduction wo. cyc (i) ϕu is the homogenization of ϕu (by Lemma B.48), in particular it is symmetric and conjugation- invariant. (ii) If w has u-decomposition (g1,...,gn) in the sense of Remark B.38, then ϕ (w)= j 1,...,n g = u j 1,...,n g = u−1 . u { ∈{ } | j }−{ ∈{ } | j } (iii) It follows from (ii) that ϕ (un)= n for every n Z; consequently also ϕcyc(un)= n for all n Z. u ∈ u ∈ In particular, for u W the quasimorphisms ϕ : F Z and ϕcyc : F Z are surjective. ∈ r u r → u r → REMARK B.52 (Quasimorphisms with non-commutative targets). It was pointed out by Thurston long ago that “constructing non-trivial quasimorphisms with non-commutative targets is difficult”. Recently, a far reaching structure theory for general quasimorphisms with non-commutative target was developed by Fujiwara and Kapovich [FK16], which confirms this statement. While homomorphisms are always symmetric, this is not the case for general quasimorphisms. We have seen that every R-valued quasimorphism is at bounded distance from a symmetric quasi- morphism. In general we can say the following:

LEMMA B.53 (Quasi-homomorphisms are almost symmetric). If f : G H is a (d, C)-quasimor- → phism with respect to some left-invariant metric d, then d(f(e),e) C and for all g G we have ≤ ∈ d(f(g−1),f(g)−1) 2C. ≤ PROOF. Applying (5.1) with x = y = e we obtain d(e,f(e)) = d(f(e),f(e)f(e)) = d(f(ee),f(e)f(e)) C, ≤ 6. RECONSTRUCTION THEOREMS FOR HOMOGENEOUS QUASIMORPHISMS 127 and thus applying (5.1) again we obtain for all g G, ∈ d(f(g−1),f(g)−1) = d(f(g)f(g−1),e) d(f(g)f(g−1),f(gg−1)) + d(e,f(e)) 2C, ≤ ≤ which finishes the proof. 

REMARK B.54 (Quasi-homomorphisms vs. symmetric quasimorphisms). It is an easy consequence of Lemma B.53 that for a left-invariant metric d, every d-quasi-homomorphism f : G H is at → bounded distance from a symmetric function f : G H. Indeed, we can choose a subset G+ G → ⊂ such that G = e G+ (G+)−1 and define f : G G { } ∪ ∪ e → f(g), g G+, ∈ f(g) := f(ge−1)−1, g (G+)−1,  ∈  e, g = e. e However, this function need not be a d-quasi-homomorphism again, so it is not clear to us whether every d-quasi-homomorphism is at bounded distance from a symmetric d-quasi-homomorphism. If the metric d happens to be bi-invariant, then the answer to this question is positive. In fact, in this case any function f : G H which is at bounded distance from f is a d-quasi-homomorphism, → since for all x, y G we have ∈ e d(f(xy), f(x)f(y)) d(f(xy),f(xy)) + d(f(xy),f(x)f(y)) + d(f(x)f(y), f(x)f(y)) ≤ 2C + C + d(f(x)f(y),f(x)f(y)) + d(f(x)f(y), f(x)f(y)) e e e ≤ e e e 3C + d(f(y), f(y)) + d(f(x), f(x)) 7C. ≤ e ≤ e e e However, admitting a bi-invariant metric is a rather restrictive condition on the group H. e e 6. Reconstruction theorems for homogeneous quasimorphisms In our study of conical limit points we will use a reconstruction theorem for homogeneous quasi- morphisms, which is based on work by Ben Simon and the second named author [BSH12]; this re- quires the following terminology.

NOTATION B.55 (Sandwiched order). Let f :Γ R be a non-zero homogeneous quasimorphism → on a group Γ. We say that a partial order on Γ is bi-invariant if g h implies gx hx and xg xh ≤ ≤ ≤ ≤ for all g,h,x Γ; such an order is then determined by the associated order semigroup S := g Γ ∈ ≤ ≤ { ∈ | g e . We say that is sandwiched by f if there exists constants C > C > 0 such that ≥ } ≤ 2 1 g Γ f(g) > C S g Γ f(g) > C ; { ∈ | 1}⊂ ≤ ⊂{ ∈ | 2} in fact, the lower bound implies the upper bound with C1 := 0 [BSH12, Lemma 3.2]. We denote by IPO(f) the set of all invariant partial orders on Γ which are sandwiched by f. If is sandwiched by f, then there exists g Γ such that for all h Γ there exists n N with ≤ ∈ ∈ ∈ f(gn) h. Any such element g is called dominant, and following [EP00] we define the growth function ≥ associated with the pair (g, ) by ≤ 1 p n γ≤,g(h) := lim min p 0 g h . n→∞ n · { ≥ | ≥ } The following is [BSH12, Prop. 1.3]

LEMMA B.56 (Sandwich lemma). Let f : Γ R be a non-zero homogeneous quasimorphism and let → IPO(f). Then for every dominant g we have f(g) > 0 and ≤∈ f(h) γ (h)= .  g f(g) The sandwich lemma has the following immediate consequence:

COROLLARY B.57 (First reconstruction theorem). Let f ,f :Γ R be non-zero homogeneous quasi- 1 2 → morphisms. If IPO(f ) IPO(f ) = , then f = C f for some C > 0.  1 ∩ 2 6 ∅ 1 · 2 128 B. SOME NOTIONS FROM GEOMETRIC GROUP THEORY

We can give a more convenient reformulation using the following notion:

DEFINITION B.58. If f :Γ R is a homogeneous quasimorphism, then the positivity set Pos(f) is → given by Pos(f) := g Γ f(g) 0 . { ∈ | ≥ } COROLLARY B.59 (Second reconstruction theorem). Let f ,f : Γ R be non-zero homogeneous 1 2 → quasimorphisms. If Pos(f ) Pos(f ), then f = C f for some C > 0.  1 ⊂ 2 1 · 2 PROOF. Let g Pos(f ) (and hence g Pos(f )) and set P := Pos(f ). Denote by d(f) the 0 ∈ 1 ∈ 2 2 supremum of the defect set of f and choose n N with n> 2 d(f2) . Then for every g PgnP we have ∈ f2(g0) ∈ 0 d(f2) f2(g) n f2(g0) 2d(f2) > 2 f2(g0) 2d(f2) 0, ≥ · − f2(g0) − ≥ and hence PgnP P . This shows that gnP is a semigroup, and hence there exists a bi-invariant 0 ⊂ 0 partial order on Γ with ≤ S = e g(gnP )g−1. ≤ { } ∪ 0 g\∈Γ We claim that is sandwiched by both f and f ; this will imply the corollary by the first reconstruc- ≤ 1 2 tion theorem. For the proof of the claim let j 1, 2 and set C := n f (g )+ d(f ). Now if x Γ ∈ { } 1 · j 0 j j ∈ with f(x ) > C , then for all g Γ we have j 1 ∈ f (g−ng−1x g) f (g−n)+ f(g−1x g) d(f)= n f (g )+ f(x ) d(f) > 0, j 0 j ≥ j 0 j − − · j 0 j − and hence g−ng−1x g gnPos(f ) P = x g(gnP )g−1 S , 0 j ∈ 0 j ⊂ ⇒ j ∈ 0 ⊂ ≤ g\∈Γ which shows that f sandwiches and finishes the proof.  j ≤ APPENDIX C

Morse hyperbolic spaces and their Gromov boundaries

1. Gromov, Rips and Morse hyperbolic spaces In this section we collect some basic notions concerning hyperbolic spaces and their boundaries. We use three definitions of hyperbolicity. The first definition, called the 4-point or Gromov definition of hyperbolicity, is defined for any metric space.

DEFINITION C.1 (Gromov hyperbolic spaces). Let (X, d) be a metric space. Given three points x, y, w X, the Gromov product of x and y with respect to the basepoint w is defined as ∈ 1 (x y) = (d(w, x)+ d(w,y) d(x, y)) . | w 2 − Given a δ 0, a metric space X is said to be δ-Gromov hyperbolic if for all x,y,z,w X: ≥ ∈ (x y) min (x z) , (y z) δ. | w ≥ { | w | w}− A metric space X is called Gromov hyperbolic if it is δ-Gromov hyperbolic for some δ 0. ≥ The second definition, credited to Rips, requires that the metric space be geodesic because it char- acterizes hyperbolicity by asking that all geodesic triangles be “slim”.

DEFINITION C.2 (Rips hyperbolic spaces). Let (X, d) be a geodesic metric space. Given a δ 0,a ≥ geodesic triangle in X is said to be δ-slim if each of its sides is contained in the δ-neighborhood of the union of the other two sides. The space (X, d) is called δ-Rips hyperbolic if every geodesic triangle in X is δ-slim. It is called Rips hyperbolic if it is δ-Rips hyperbolic for some δ 0. ≥ Rips’ definition is often easier to handle; in the case of geodesic metric spaces, it coincides with Gromov’s definition:

PROPOSITION C.3 (Rips vs. Gromov hyperbolicity, [BH99, Prop. III.H.1.22]). For every δ 0 there ≥ exists δ′ 0 with the following property: If a geodesic metric space (X, d) is δ-Gromov hyperbolic (respectively ≥ δ-Rips hyperbolic), then it is δ′-Rips hyperbolic (respectively δ′-Gromov hyperbolic). In particular, a geodesic metric space is Gromov hyperbolic if and only if it is Rips hyperbolic.

Among geodesic metric spaces, Gromov hyperbolicity (or, equivalently, Rips hyperbolicity) is actually a QI-invariant, see e.g. [BH99, Thm. III.H.1.9]:

THEOREM C.4 (QI-invariance of hyperbolicity among geodesic metric spaces). If X, X′ are quasi- isometric geodesic metric spaces, then X is hyperbolic if and only if X′ is hyperbolic. 

The proof of this theorem uses a particularly nice property of hyperbolic spaces, that says, roughly, that (the image of) a quasi-geodesic segment is close to any geodesic segment joining its endpoints. The following lemma details this property.

LEMMA C.5 (Morse lemma). There exists a function Θ(K,L,δ) so that the following holds: If X is a δ- Gromov hyperbolic geodesic space, then for every (K,L)-quasi-geodesic segment f :[a,b] X, the Hausdorff → distance between the image of f and any geodesic segment between f(a) and f(b) is at most Θ(K,L,δ). 

129 130 C.MORSEHYPERBOLICSPACESANDTHEIRGROMOVBOUNDARIES

Occasionally, in particular in the context of Morse boundaries, we will have to deal with non- geodesic hyperbolic spaces. In this wider context, Gromov hyperbolicity is no longer a quasi-isometry invariant.

EXAMPLE C.6. Let X R2 be the graph of the function x x with the metric induced from the ⊂ 7→ | | Euclidean metric of R2. Then X is not hyperbolic, but π : X R, (x, y) x is a quasi-isometry (even 1 → 7→ bi-Lipschitz), and R is 0-Gromov hyperbolic. See [BS07a, Remark 4.1.3] for a more refined example. The next theorem, though, gives four characterizations of a quasi-isometrically invariant notion.

THEOREM C.7 (Conditions for Morse hyperbolicity). Let (X, d) be a proper metric space. Then the following are equivalent: (1) X is quasi-isometric to a proper geodesic hyperbolic metric space. (2) There exists a K 1, C 0 so that X is (K, C)-quasi-geodesic and for all K′ K, C′ C ≥ ≥ ′ ′ ′ ′ ≥ ≥ there exists a Morse gauge N(K′,C′), depending on K , C , so that every (K , C )-quasi-geodesic is N(K′,C′)-Morse. (3) There exists a K 1, C 0 so that X is (K, C)-quasi-geodesic and for all K′ K, C′ C there ≥ ≥ ′ ′ ′ ′ ≥ ≥ exists δ ′ ′ 0, depending on K , C , so that all (K , C )-quasi-geodesic triangles are δ ′ ′ - (K ,C ) ≥ (K ,C ) slim as in Definition C.2. (4) X is large scale geodesic and for all K 1 and C 0 there exists a δ 0 so that every (K, C)-quasi- ≥ ≥ ≥ geodesic triangle is δ-slim.

DEFINITION C.8. A proper metric space (X, d) is Morse hyperbolic if it satisfies one (hence any) of the four equivalent properties in Theorem C.7. We record the following consequence for later reference:

COROLLARY C.9 (Quasi-convex subsets). Every quasi-convex subset of a geodesic Morse hyperbolic space is Morse hyperbolic.

PROOF. It is convenient to work with Definition (3) of a Morse hyperbolic space. Thus let X be geodesic and Morse hyperbolic and let Y X be C-quasi-convex. It then follows from Remark A.21 ⊂ that Y is quasi-geodesic. In fact, by a similar argument as in the remark one can show that every quasi-geodesic in Y is uniformly close to a quasi-geodesic in X, and hence that every quasi-geodesic triangle in Y is at uniformly bounded Hausdorff distance from a quasi-geodesic triangle in X, The corollary follows.  The remainder of this section is devoted to the proof of Theorem C.7. Firstly, recall from Lemma A.28 that being a quasi-geodesic metric space is equivalent to being a large scale geodesic metric space. Since for K′ K and C′ C every (K, C)-quasi-geodesic is also ≥ ≥ a (K′, C′)-quasi-geodesic, we deduce that (3) (4). We prefer to work with Condition (3) in the ⇐⇒ sequel, since it is closer in spirit to Conditions (1) and (2). We are going to show that (1) = (2) = ⇒ ⇒ (3) = (1). The implication (1) = (2) follows from the following lemma: ⇒ ⇒ LEMMA C.10 (Stability of quasi-geodesics). Assume that X is quasi-isometric to a proper geodesic hyperbolic metric space. Then there exist K 1, C 0 such that X is (K, C)-quasi-geodesic and for all ′ ′ ≥ ≥ ′ ′ ′ ′ K K, C C there exists a Morse gauge N ′ ′ , depending on K , C , so that every (K , C )-quasi- ≥ ≥ (K ,C ) geodesic is N(K′,C′)-Morse.

PROOF. Since X is quasi-isometric to a geodesic metric space, it is large-scale geodesic by Lemma A.26, hence by Lemma A.28 there exist K 1, C 0 such that X is (K, C)-quasi-geodesic. ≥ ≥ Now we fix a (K , C )-quasi-isometry f : X Y into a proper geodesic δ-hyperbolic metric space. 1 1 → Let x, y X and let γ be a (K′, C′)-quasi-geodesic joining x and y, where K K′, C C′. We ∈ ≤ ≤ know that f(γ) is a (K′′, C′′)-quasi-geodesic with endpoints f(x) and f(y), where K′′, C′′ depend on ′ ′ K1,K , C1, C . Since Y is geodesic and δ-hyperbolic, we know that there exists a geodesic segment [f(x),f(y)], which is moreover N-Morse, for some Morse gauge N which depends only on δ. So we 2.BOUNDARIESOFGROMOVANDMORSEHYPERBOLICSPACES 131 know that f(γ) is in the N ′(K′′, C′′)-neighborhood of [f(x),f(y)]. Indeed, since Y is geodesic, f(γ) and [f(x),f(y)] have Hausdorff distance bounded by M =2N ′(K′′, 2(K′′ + C′′)) + K′′ + C′′ [Cor17, Lemma 2.1]. Let β bea (k,c)-quasi-geodesic with endpoints on γ. We know that f(β) is a (k′,c′)-quasi-geodesic ′ ′ ′ ′ with endpoints on f(γ), where k ,c depend on K1,k,C1,c. We also note that f(β) is a (k ,c + 2N(K′′, C′′)-quasi-geodesic with endpoints on [f(x),f(y)], so it is in the M ′ = N ′(k′,c′+2N(K′′, C′′))- neighborhood of [f(x),f(y)]. Thus we can conclude that f(β) is in the M + M ′ neighborhood of f(γ). ′ It follows that β lies in the K1(M + M + C1)-neighborhood of γ. Thus γ is N(K′,C′)-Morse, where ′ N(K′,C′) = K1(M + M + C1).  The implication (2) = (3) is provided by the following lemma: ⇒ LEMMA C.11 (Slim quasi-triangles). Assume X is (K, C)-quasi-geodesic and there exists a Morse gauge N , depending on K, C, so that every (K, C)-quasi-geodesic is N -Morse. Then there exists δ (K,C) (K,C) (K,C) ≥ 0, depending on K, C, so that all (K, C)-quasi-geodesic triangles are δ(K,C)-slim as in Definition C.2.

PROOF. We follow closely the proof Lemma 6.2 in [MM99]. By abuse of notation for any two points x, y X we will use [x, y] to denote a (K, C)-quasi-geodesic joining x and y. Let x,y,z X. ∈ ∈ We will show that [x, y] lies in a δ-neighborhood of [x, z] [y,z], for δ depending on K,C,N. ∪ Let z′ be a point on [x, y] closest to z. We will show that the concatenation [x, z′] [z′,z] is a ∪ (3K, 2C)-quasi-geodesic. Let u [x, z′] and v [z′,z]. We note that d(z′, v) d(u, v), because z′ is ∈ ∈ ≤ a closest point to v. Let the difference in parameters along [x, z′] between u and z′ be b a and the | − | difference in parameters along [z′,z] between z′ and v be c d . | − | Putting everything together, we get the following two inequalities:

b a + d c Kd(u,z′)+ KC + Kd(z′, v)+ KC d(u, v) d(u,z′)+ d(z′, v) | − | | − |≤ ≤ K(d(u, v)+ d(z′, v)) + KC + Kd(z′, v)+ KC K b a + C + K d c + C ≤ ≤ | − | | − | K(2d(u, v)) + KC + Kd(u, v)+ KC = K( b a + d c )+2C ≤ | − | | − | 3Kd(u, v)+2KC ≤ ′ ′ Now since all (K, C)-quasi-geodesics are N(K,C)-Morse, we have that [x, z ] [z ,z] is in the ′ ∪ δ(K,C) = N(K,C)(3K, 2C)-neighborhood of [x, z] and thus, in particular, [x, z ] is. We can repeat this ar- gument substituting y for x and we see that all of [x, y] is in the δ -neighborhood of [x, z] [y,z].  (K,C) ∪ Finally, the following lemma establishes the implication (3) = (1) and finishes the proof: ⇒ LEMMA C.12. Let X be a proper (K, C)-quasi-geodesic metric space with the property that for all K′ ′ ′ ′ ′ ′ ≥ K, C C there exists δ ′ ′ 0, depending on K , C , so that all (K , C )-quasi-geodesic triangles are ≥ (K ,C ) ≥ δ(K′,C′)-slim as in Definition C.2. Then X is quasi-isometric to a hyperbolic proper geodesic metric space.

PROOF. Since X is a (K, C)-quasi-geodesic metric space, it is large scale geodesic (Lemma A.28) and thus there exists a proper geodesic metric space Y and a (K′, C′)-quasi-isometry f : Y X → (Lemma A.26). Let x,y,z Y and let T be a geodesic triangle in Y with vertices x,y,z. We note that T ′ = f(T ), is ′ ′ ∈ ′ a (K , C )-quasi-geodesic triangle. By assumption on X there exists a δ(K′,C′) > 0 so that T is δ(K′,C′)- slim. It follows from the quasi-isometric inequality that T is (Kδ(K′,C′) + C)-slim, hence Y is Rips (equivalently, Gromov) hyperbolic. 

2. Boundaries of Gromov and Morse hyperbolic spaces From now on let (X, d) be a Gromov hyperbolic space (not necessarily geodesic) and let o X be ∈ a basepoint. A sequence of points x X converges to infinity if { i}⊂ lim (xi xj )o = . i,j→∞ | ∞ 132 C.MORSEHYPERBOLICSPACESANDTHEIRGROMOVBOUNDARIES

We say that two sequences (xi) and (yi) that converge to infinity are equivalent if

lim (xi yi)o = . i→∞ | ∞ One can check that neither of these notions depend on the choice of basepoint.

DEFINITION C.13. If (X, d) is a Gromov hyperbolic metric space, then its sequential Gromov bound- ary ∂sX is defined as the set of equivalence classes of sequences converging to infinity. To topologize the set X := X ∂ X, one observes first that the Gromov product can be extended ∪ s to X as follows: If ξ, η ∂ X, then we set ∈ s (ξ η)o := inf lim inf (xi yj)o , | i,j→∞ { | } where the infimum is taken over all sequences (xi) and (yj ) in X representing ξ and η respectively. If ξ ∂ X and y X, then ∈ s ∈ (y ξ)o = (ξ y)o = inf lim inf(xi y)o, | | i→∞ | where the infimum is again taken over all sequences (xi) representing ξ. Then the Gromov topology on X is the unique topology which restricts to the metric topology on X and such that a neighborhood basis of ξ ∂ X is given by the sets ∈ s (ξ)= x X (ξ x) > R , UR ∈ | | o as R ranges through (0, ). It turns out that this topology is independent of the choice of basepoint o. ∞ From now on we will always consider X as a topological space with respect to the Gromov topology.

LEMMA C.14 (Gromov compactification of a proper hyperbolic space). The subset X X is open ⊂ and hence ∂sX is closed in X. If (X, d) is proper, then X and ∂sX are compact.  Thus if X is a proper hyperbolic space (not necessarily geodesic), then X is a compactification of X, i.e. a compact space which contains X as a dense open subset. Now let X be a proper geodesic hyperbolic space. In this case, the Gromov compactification X ad- mits the following more explicit description (see [BH99, Section III.H.3]: Recall that (quasi-)isometric embedding c : [0, ) X is called a (quasi-)geodesic ray in X and that two quasi-geodesic rays c,c′ in ∞ → X are asymptotic provided sup d(c(t),c′(t)) t 0 < . { | ≥ } ∞ This property is actually equivalent to finite Hausdorff distance of their images, hence independent of the chosen parametrization. Denote by ∂rX the collection of all asymptoticity classes of quasi-geodesic rays in X. Given a quasi-geodesic ray c, we denote by c( ) the associated class in ∂ X. ∞ r By [BH99, Lemma III.H.3.1], every asymptoticity class admits a representative c0 which is a geo- desic ray emanating from o. Thus, ∂ X = c( ) c : [0, ) X geodesic ray,c(0) = o . r { ∞ | ∞ → } This is sometimes taken as the definition of ∂rX, but we prefer our original definition, which has better functoriality properties.

LEMMA C.15 ([BH99, Lemma III.H.3.13]). If c is a quasi-geodesic ray in a proper geodesic hyperbolic space X, then the sequence (c(n)) N converges to infinity and there is a well-defined bijection ∂ X ∂ X n∈ r → s given by c( ) [(c(n)) N].  ∞ 7→ n∈ In view of this lemma we refer to ∂rX as the ray model of the Gromov boundary of X. We will often identify ∂rX and ∂sX via the bijection above, and denote either of these spaces simply by ∂X. In (r) particular, we will topologize ∂rX via this identification. Similarly, we will identify X := X ∂rX (r) ∪ and X, and use this identification to topologize X . The resulting topology admits the following more explicit description: 3. VISUAL METRICS 133

LEMMA C.16 ([BH99, Lemma III.H.3.6]). Assume that X is proper, geodesic and δ-Rips hyperbolic and let k> 2δ. Let c : [0, ) X be a geodesic ray with c (0) = o and for every n N let 0 ∞ → 0 ∈ V (c ) := c( ) c : [0, ) X geodesic ray,c(0) = o, d(c(n),c (n)) < k . (2.1) n 0 { ∞ | ∞ → 0 } (r) Then V (c ) n N is a fundamental system of (not necessarily open) neighborhoods of c( ) in X .  { n 0 | ∈ } ∞ Now let X, X′ be proper geodesic hyperbolic metric spaces and let f : X X′ be a quasi- → isometry. If c is a quasi-geodesic ray in X, then f c is a quasi-geodesic ray in X′, and asymptoticity ◦ is preserved. Thus f induces a map ∂ f : ∂ X ∂ X′, c( ) (f c)( ), r r → r ∞ 7→ ◦ ∞ (r) (r) ′(r) and hence a map f : X X . It follows from Lemma C.16 that ∂rf is continuous (see [BH99, → (r) Thm. III.H.3.9]), even if f itself is not continuous (and hence f is not continuous). Using the identi- fication from Lemma C.15 we also obtain maps f : X X′ and ∂ f : ∂ X ∂ X′, the latter of which → s s → s is continuous.

COROLLARY C.17 (Boundary extensions of quasi-isometries). If X, X′ are proper geodesic hyperbolic ′ ′ metric spaces and f : X X is a quasi-isometry, then ∂rf : ∂rX ∂rX is a homeomorphism. In particular, ′ → ′ → ′ if X and X are quasi-isometric, then ∂rX and ∂rX (and hence ∂sX and ∂sX ) are homeomorphic. 

REMARK C.18. While the extension f of a quasi-isometric embedding f : X X′ does not need to → be continuous on X, it is still continuous at infinity in the following sense: If xn converges to infinity in ′ X, then (f(xn)) converges to infinity in X , and lim f(xn)= ∂f(lim xn). In particular, if X is discrete, then f is continuous. We now use quasi-isometry invarianve of Gromov boundaries of proper geodesic Gromov hy- perbolic spaces to define Gromov boundaries for Morse hyperbolic spaces. Thus for the remainder of this section, X denotes a Morse hyperbolic space. We denote by [X] the quasi-isometry class of X and by [X] [X] the subclass of proper geodesic representatives of [X]. By Theorem C.7 this geod ⊂ subclass is non-empty and by Remark 5.2 every X [X] is a proper geodesic Gromov- (hence 1 ∈ geod Rips-)hyperbolic metric space.

DEFINITION C.19. If X [X] then the Gromov boundary ∂X of X is called a Gromov bound- 1 ∈ geod 1 1 ary of the Morse hyperbolic space X.

REMARK C.20 (Uniqueness properties of Gromov boundaries). By definition, a Morse hyperbolic space has many different Gromov boundaries. However, if X ,X [X] , then there exists a quasi- 1 2 ∈ geod isometry f : X X and by Corollary C.17 this map extends to a boundary map ∂f : ∂X ∂X . 1 → 2 1 → 2 Any property preserved by such boundary maps is thus an invariant of X.

EXAMPLE C.21 (Topological boundary invariants). By Corollary C.17, all Gromov boundaries of X have the same (compact, metrizable) homeomorphism type, which we denote by ∂[X]. If is any I homeomorphism-invariant of compact metrizable spaces, then we may define (∂[X]) := (∂X ), where X [X]. I I 1 1 ∈ We then refer to (X) as a topological boundary invariant of X. A typical example of such an invariant is I topological dimension (cf. Definition 7.3): we may thus define the topological dimension of the boundary, dim(∂[X]), for every Morse hyperbolic space X. We will sometimes abuse notation and simply denote this invariant by dim ∂X.

3. Visual metrics In this section, (X, d) denotes a proper geodesic hyperbolic space. Given a basepoint o X and a ∈ constant a> 1, we define a map ′ ρ : ∂ X ∂ X [0, ), ρ (ξ, ξ′) := a−(ξ|ξ )o . o,a s × s → ∞ o,a 134 C.MORSEHYPERBOLICSPACESANDTHEIRGROMOVBOUNDARIES

This map is then a K-quasi-metric for some K 1 in the sense that it is symmetric, with ρ (ξ, ξ′)=0 ≥ o,a iff ξ = ξ′, and for all ξ, ξ′, ξ′′ ∂ X we have ∈ s ρ (ξ, ξ′′) K max ρ (ξ, ξ′), ρ (ξ′, ξ′′) . o,a ≤ { o,a o,a } The dependence on the parameters is as follows [BS07a, Rem. 2.2.4]: For all a,a′ > 1 and o, o′ X we ∈ have ′ ′ ′ ρo,a(ξ, ξ ) ′ log a −d(o,o ) d(o,o ) ′ ′ ′ log a a ′ a and ρo,a (ξ, ξ )= ρo,a(ξ, ξ ) . (3.1) ≤ ρo′,a(ξ, ξ ) ≤

DEFINITION C.22. A metric d on ∂sX is called visual with basepoint o X and parameters a> 1 ′ ∈ and c 1 if it is bi-Lipschitz equivalent to ρo,a, i.e., for all ξ, ξ ∂sX we have ≥ ′ ∈ ′ c−1a−(ξ|ξ )o = c−1ρ (ξ, ξ′) d(ξ, ξ′) cρ (ξ, ξ′)= ca−(ξ|ξ )o . (3.2) o,a ≤ ≤ o,a These inequalities are then called the visual inequalities for d.

In the geodesic case, visual metrics on ∂rX are defined analogously, using the identification ∂rX ∼= ∂sX. By[BS07a, Thm. 2.2.7], every hyperbolic space (geodesic or not) admits a visual metric on its Gromov boundary. By (3.1) any two visual metrics with the same parameter a are bi-Lipschitz. More generally, if d and d′ are visual metrics with respective parameters a and a′, then d′ is bi-Lipschitz to ′ dα, where α := log a , i.e. there exists c 1 such that log a ≥ c−1d(ξ, ξ′)α d′(ξ, ξ′) c d(ξ, ξ′)α. (3.3) ≤ ≤ One can show that any visual metric on ∂sX induces the given topology [BH99, Proposition III.H.3.21]. We call a > 1 a visuality parameter for X if there exists a visual metric on ∂sX with parameters (a,c) for some c 1. ≥ LEMMA C.23. The set of visuality parameters is a non-empty connected set, hence either (1, ) or of the ∞ form (1,amax) or of the form (1,amax], for some amax > 1.

PROOF. Non-emptiness follows from the existence of a visual metric. If d is a visual metric with parameters (a,c) and ǫ (0, 1), then dǫ is a visual metric with parameters (aǫ,cǫ).  ∈ We recall from [BS07a, Cor. 4.4.2 and Prop. 4.3.2] that if X is a geodesic hyperbolic metric space with basepoint o X and f : X X is a (K , C )-quasi-isometry, then there exist constants K 1 ∈ → 0 0 ≥ and C 0 depending only on K , C , X and o such that f is a (K, C)-quasi-isometry and moreover ≥ 0 0 K−1(x x′) C (f(x) f(x′)) K(x x′) + C, (3.4) | o − ≤ | f(o) ≤ | o for all x, x′ X. We then say that f is a (K,C,o)-quasi-isometry. By[BS07a, Lemma 2.2.2] this carries ∈ over to the boundary in the following form:

LEMMA C.24. If X is a geodesic δ-Gromov hyperbolic metric space, then for every (K,C,o)-quasi-isometry f : X X the boundary extension ∂ f : ∂ X ∂ X satisfies → r r → r K−1(ξ′ ξ′′) C (∂ f(ξ′) ∂ f(ξ′′)) K(ξ′ ξ′′) + C +2δ, | o − ≤ r | r f(o) ≤ | o for all o X and ξ′, ξ′′ ∂ X.  ∈ ∈ r Note that by taking exponentials, from Lemma C.24 we can deduce the quantitative estimates concerning the behavior of ∂rf with respect to any visual metric, which involve the hyperbolicity constant of X, quasi-isometry constants of f and the parameter of the visual metric in question. APPENDIX D

Morse boundaries of proper geodesic metric spaces

In this appendix we discuss a vast generalization of the notion of Gromov boundary known as Morse boundary, which can be defined for arbitrary proper geodesic spaces and is a quasi-isometry invariant for such spaces.

1. Definition and basic properties Since the definition of the Morse boundary is based on the notion of a Morse geodesic, we briefly recall the relevant definitions from Definition A.17:

REMARK D.1 (Morse geodesic rays and asymptoticity). Let X be a proper geodesic metric space and let N : [1, ) [0, ) [0, ) be an arbitrary function. Recall that a quasi-geodesic ray γ : I X ∞ × ∞ → ∞ → is called N-Morse if any (K, C)-quasi-geodesic segment q with endpoints on γ is contained in the N(K, C)-neighborhood (γ) of γ; we then also call N a Morse gauge for γ. We say that two NN(K,C) Morse geodesic rays c and d in X are asymptotic if they have bounded Hausdorff distance. This defines an equivalence relation on the space of Morse geodesic rays, and the equivalence class of such a ray c is denoted by c( ). The following observation (see [Cor17, Prop. 2.4]) will be crucial for us: There exists ∞ ′ a constant CN (depending only on N) such that two N-Morse geodesic rays c and c are equivalent if and only if d(c(t),c′(t)) C for all t. ≤ N DEFINITION D.2. Let X be a proper geodesic metric space with basepoint o X. The Morse ∈ boundary of X based at o is the subset ∂M X := c( ) c : [0, ) X Morse geodesic ray,c(0) = o ∂ X r o { ∞ | ∞ → }⊂ r REMARK D.3. In general, it may happen that ∂M X = . For example, this happens if every r o ∅ geodesic in X bounds a half-flat, as is the case in a symmetric space (or Bruhat-Tits building) of higher M rank. If X is hyperbolic, then ∂r Xo = ∂rX. We will describe here two different, but equivalent, approaches to define a quasi-isometry invari- ant topology on the Morse boundary; for more details see [Cor17].

REMARK D.4 (Direct limit topology on the Morse boundary). Let X be a (not necessarily hyper- bolic) proper geodesic metric space and let o X be a basepoint. Consider all of the geodesic rays ∈ emanating from o up to asymptoticity. For each Morse gauge N we choose this subset ∂N X = c( ) ∂ X c : [0, ) X is an N-Morse geodesic ray with c(0) = o ∂M X r o { ∞ ∈ r | ∞ → }⊂ r o N and topologize ∂r Xo as in Lemma C.16, by replacing 2δ by the CN from Remark D.1. We then define a partial order on the set of all Morse gauges as follows: Given N,N ′ we set N N ′ if and only ′ ′ M ′ ∈M ≤N N if N(λ, ǫ) N (λ, ǫ) for all λ, ǫ N. One can show that if N N , then the inclusion ∂r Xo ֒ ∂r Xo ≤ ∈ ≤M → is continuous. We can thus topologize the Morse boundary ∂r Xo as the direct limit M N ∂r Xo = lim ∂r Xo. N∈−→(M,≤) M From now on we always consider the Morse boundary ∂r Xo as a topological space with respect to the above direct limit topology. If X is hyperbolic, then we simply recover the (ray model of the) Gromov boundary ∂rX by Lemma C.5. In general, however, the direct limit topology on the Morse

135 136 D.MORSEBOUNDARIESOFPROPERGEODESICMETRICSPACES

N boundary may be non-compact. We emphasize that ∂r Xo is defined using N-Morse geodesic rays rather than N-Morse quasi-geodesic rays. It is possible to define the direct limit topology using Morse quasi-geodesic rays, but for this one has to carefully keep track of the quasi-isometry constants in- volved. We now provide an alternative model for the Morse boundary, which is more in line with the sequential model of the Gromov boundary:

REMARK D.5 (Sequential description of the Morse boundary). If X is a proper geodesic metric space with basepoint o X, then an alternative description of the Morse boundary ∂M X and its ∈ r o direct limit topology can be given as a direct limit of Gromov boundaries of certain hyperbolic sub- (N) spaces of X. More precisely, given a Morse gauge N we denote by Xo X the set of all y X ⊂ ∈ such that there exists a N–Morse geodesic segment from o to y in X, which we denote [o, y]. Note that there is no reason why this should be connected of even geodesic. However, by [CH17, Proposition 3.2] it is always Gromov hyperbolic with respect to the restricted metric, and hence we can consider (N) its sequential Gromov boundary ∂sXo which is a (possibly empty) compact topological space with respect to the topology induced by some (hence any) visual metric. For N N ′ the isometric inclu- ′ ′ (N) (N ) (N) (N ) ≤ sions Xo ֒ Xo induce continuous inclusions ∂ Xo ֒ ∂ Xo , and hence we can define the → s → s sequential Morse boundary as M (N) ∂s Xo := lim ∂sXo . N∈−→(M,≤) M (N) We note that every (pre-) compact subset of ∂s Xo is contained in one of the strata ∂sXo (see [CD19, (N) N Lemma 4.1]). One can show that there are natural homeomorphisms ∂ Xo ∂ X which are s → r o compatible with inclusions, see [CH17]. This implies in particular, that M M ∂s Xo ∼= ∂r Xo, M hence we refer to ∂s Xo as the sequential model of the Morse boundary. While this model is very much in line with the sequential approach to the Gromov boundary, we still need to assume that X be (N) geodesic in order to obtain a sensible definition of the spaces Xo . M Note that every element ξ ∂s Xo can be represented by a sequence (xn) of points which are ∈ (N) contained in some fixed hyperbolic approximant Xo . We then write ξ = lim xn and say that ξ is an N-boundary point.

NOTATION D.6 (Representation by geodesic rays and lines). In the sequel we will mostly work M with the sequential model ∂s Xo of the Morse boundary. Nevertheless, we sometimes want to repre- sent such points by geodesic rays: If γ is a geodesic ray in X (not necessarily emanating from o), then we say that γ is asymptotic to ξ ∂M X if it is asymptotic (in the sense of Remark D.1) to a geodesic ray ∈ s o γ′ emerging from o such that γ( ) corresponds to ξ via the homeomorphism ∂M X = ∂M X . Note ∞ s o ∼ r o that any geodesic which is asymptotic to a point ξ in the Morse boundary is automatically Morse. In fact, if ξ is an N-boundary point, then γ is N ′-Morse for a gauge N ′ depending only on N and d(γ(0), o). If γ : R X is a bi-infinite geodesic line in X, then we can define two geodesic rays γ± : [0, ) → ∞ → X by γ+(t) := γ(t) and γ−(t) := γ( t). We then say that γ is bi-asymptotic to (ξ−, ξ+) (∂M X )2 if − ∈ s o γ− is asymptotic to ξ− and γ+ is asymptotic to ξ+. As in the ray case one observes that if ξ± are both N-Morse and γ is a geodesic line which is bi-asymptotic to (ξ−, ξ+), then γ is N ′-Morse for a Morse gauge N ′ depending only on N. It turns out that all geodesic lines which are bi-asymptotic to the same pair of points are at uni- formly bounded Hausdorff distance ([CD19]):

PROPOSITION D.7 (Bi-asymptoticity classes of geodesic lines). For any Morse gauge N, there exists ′ ′ − + (N) (N) a constant K > 0 such that if γ,γ are geodesic lines bi-asymptotic to (ξ , ξ ) ∂ Xo ∂ Xo , then ∈ s × s ′ ′ dHaus(γ,γ )

Note that if X and X′ are two proper geodesic metric spaces and f : X X′ is a quasi-isometric → embedding which maps a given basepoint o X to o′ X′, then it induces a map ∈ ∈ ′ ∂ f : ∂ X ∂ X ′ , c( ) (f c)( ). r r o → r o ∞ 7→ ◦ ∞ M M ′ The following example from [Cor17] shows that this map does not in general map ∂r Xo to ∂r Xo′ :

EXAMPLE D.8 (Failure of functoriality of the Morse boundary). Let X be the hyperbolic plane and let c : R X be a bi-infinite geodesic. Then the restriction c : [0, ) X of c is a Morse quasi- → ∞ → geodesic ray. Now let X′ be the space obtained from X by gluing a Euclidean half-plane along c. Then thee inclusion f : X X′ is a (quasi-)isometric embedding, but f c is no longer Morse.e → ◦ ′ M M ′ e Even if f : X X is a quasi-isometric embedding such that ∂rf(∂r Xo) ∂r Xo′ , then the map M M→ ′ ⊂ ∂ f : ∂ X ∂ X ′ need not be continuous with respect to the direct limit topology. The following r r o → r o condition is immediate from the definitions (cf. [Cor17, Proposition 4.2]):

PROPOSITION D.9. Let X,X′ be proper geodesic metric spaces with basepoints o, o′ and let f : X X′ → be a quasi-isometric embedding with f(o)= o′. If for every N there exists N ′ such that ∈M ∈M ′ (N) ′(N ) ∂ X ∂ X ′ , r o ⊂ r o then ∂rf restricts to a map M M M ′ ∂ f : ∂ X ∂ X ′ , r r o → r o which is continuous and in fact a homeomorphism onto its image. 

A quasi-isometric embedding satisfying the condition of the proposition will be called Morse- preserving. Clearly every quasi-isometry is Morse-preserving, and hence one concludes that the Morse boundary with its direct limit topology is a quasi-isometry invariant among proper geodesic metric spaces:

COROLLARY D.10 (Quasi-isometry invariance). Let X,X′ be proper geodesic metric spaces with base- ′ ′ ′ M M M ′ points o, o and f : X X be a quasi-isometry with f(o) = o . Then ∂ f : ∂ X ∂ X ′ is a homeo- → r r o → r o morphism. 

Also note that if o and o′ are two basepoints in X, then there exists a self-quasi-isometry of X (at bounded distance from the identity) which maps o to o′. We may thus record:

M COROLLARY D.11 (Basepoint independence). Up to homeomorphism the Morse boundary ∂r Xo of a proper geodesic metric space X with respect to a basepoint o is independent of the basepoint o. 

M In view of the corollary we will sometimes be sloppy about basepoints and simply write ∂r X M instead of ∂r Xo if we only care about its homeomorphism type.

REMARK D.12 (Morse boundary for large-scale geodesic spaces). If X is a proper large-scale ge- odesic metric space, then X is quasi-isometric to a proper geodesic metric space X′. If X′′ is any other proper-geodesic metric space in the same QI type, then X′ and X′′ have homeomorphic Morse boundaries by Corollary D.10. We may thus define the Morse boundary ∂M X as their common home- omorphism type.

REMARK D.13 (Boundary maps in the sequential model). A Morse-preserving quasi-isometric ′ ′ M M ′ embedding f : (X, o) (X , o ) also induces a map ∂ f : ∂ X ∂ X ′ via the identifications of → s s o → s o the ray model and the sequential model. Explicitly, this map can be described as follows: Every point M ξ ∂s Xo can be represented by a sequence (xn) of points, which are contained in a fixed hyperbolic ∈ (N) approximant Xo . One can show that if f is Morse-preserving, then (f(xn)) will be contained in ′ ′ N some fixed hyperbolic approximant (X )o′ and ∂sf(ξ) is represented by this sequence. 138 D.MORSEBOUNDARIESOFPROPERGEODESICMETRICSPACES

2. Ideal Morse triangles Throughout this section, X denotes a proper geodesic metric space and o X is a basepoint. ∈ − + (N) − + ± DEFINITION D.14. Let ξ , ξ ∂ Xo with ξ = ξ . If γ are geodesic rays emanating from ∈ s 6 o which are asymptotic to ξ± respectively and γ is a geodesic line which is bi-asymptotic to (ξ−, ξ+), then (γ−,γ,γ+) is called an ideal N-Morse triangle with vertices (o, ξ−, ξ+). In this section we are going to show that, even if X is not hyperbolic, ideal N-Morse triangles always behave like ideal triangles in hyperbolic spaces:

− + (N) − + THEOREM D.15 (Ideal Morse triangles are slim). If ξ , ξ ∂ Xo with ξ = ξ , then there exists ∈ s 6 an ideal N-Morse triangle (γ−,γ,γ+) with vertices (o, ξ−, ξ+). Moreover, the following hold: (i) The Hausdorff distance between any two ideal N-Morse triangles with vertices (o, ξ−, ξ+) is bounded by a constant depending only on N. (ii) If (γ−,γ,γ+) is such a triangle, then the geodesics γ−,γ,γ+ are N ′-Morse for some N ′ depending only on N. (iii) There exists a constant δ depending only on N such that every N-Morse triangle with vertices (o, ξ−, ξ+) is δ-slim.

Note that (i) is immediate from Remark D.1 and Proposition D.7. It is thus enough to construct a single ideal N-Morse triangle with vertices (o, ξ−, ξ+) and to show that this specific triangle satisfies (ii) and (iii). For this we will use the following construction:

(N) CONSTRUCTION D.16 (Limit geodesics and triangles). Let (xn), (yn) Xo be sequences asymp- − + (N) − + ∈ totic to ξ , ξ ∂ Xo respectively, with ξ = ξ . Let γ = [o, x ], γ = [o, y ] be N-Morse ∈ s 6 x,n n y,n n geodesic segments. Since X is proper, the Arzelà–Ascoli Lemma A.16 implies that there exist geodesic rays γx,γy such that γx,n,γy,n subsequentially converge uniformly on compact sets to γx,γy. Let γn be a geodesic joining γ (n) and γ (n) for each n N. By Proposition 3.11 in [Cor17], (γ ) has a Morse x y ∈ n subsequential limit γ. It is then obvious that (γx,γ,γy) is an ideal N-Morse triangle with vertices (o, ξ−, ξ+) and satisfies (ii).

+ DEFINITION D.17. In the situation of Construction D.16, γx and γy are limit legs based at o for ξ and ξ− respectively, and γ is called a limit geodesic line from ξ− to ξ+. The ideal N-Morse triangle (γx,γ,γy) is called a limit triangle based at o. Working with concrete limit triangles is often more convenient for practical computation than working with abstract ideal Morse triangles; for example, the following is established in [CD19]:

(N) PROPOSITION D.18 (Limit triangles are slim). For any Morse gauge N, if (xn), (yn) Xo are − + (N) ⊂ asymptotic to ξ = ξ ∂ Xo , then any limit triangle is 4N(3, 0)-slim.  6 ∈ s This shows in particular that limit triangles satisfy Property (iii) of Theorem D.15 and thereby fin- ishes the proof of the theorem. For later reference we record the following consequence from Property (i) of Theorem D.15:

COROLLARY D.19. If γ is a geodesic ray which is asymptotic to some ξ ∂M X , then it is of bounded ∈ s o Hausdorff distance to a limit leg for ξ. Similarly, if γ is a geodesic line which is bi-asymptotic to some (ξ−, ξ+) ∈ ∂M X ∂M X , then it is at bounded Hausdorff distance from a limit geodesic. Moreover, the bounds depend s o × s o only on the respective Morse gauges. 

3. Limit sets and stable subspaces Throughout this section, X denotes a proper geodesic metric space and o X denotes a basepoint. ∈ M We can associate with every subset Y of X a corresponding subset of the Morse boundary ∂s Xo: 3. LIMIT SETS AND STABLE SUBSPACES 139

DEFINITION D.20. Let X be a proper geodesic metric space and let Y X be a subset. Then the M ⊂ Morse limit set of Y in ∂s X is (Y )= ξ ∂M X N and (y ) X(N) Y such that lim y = ξ . L ∈ s ∃ n ⊂ o ∩ n n o Note that if X happens to be Gromov hyperbolic, then (Y ) ∂X is just the Gromov limit set as L ⊂ defined in Definition 6.2. A number of basic properties of Gromov limit sets carry over to the current, more general setting. In particular, if two subspaces Y, Y ′ X are at bounded Hausdorff distance, ⊂ then (Y )= (Y ′). L L EXAMPLE D.21 (Some limit sets). If Y is bounded, then the limit set (Y ) is empty; the converse L is not true, since for example the Morse limit set of every flat of dimension > 2 is empty. The limit set of a geodesic ray is either empty (if it is not Morse) or a singleton (if it is Morse); the limit set of a geodesic line γ has at most 2 points, and if γ is bi-asymptotic to (ξ−, ξ+) then its limit set is ξ−, ξ+ . { } To obtain more interesting examples, we are going to need the following definition, which is a variant of a definition of Durham and Taylor [DT15].

DEFINITION D.22. Let X be a proper geodesic metric space, let Y X be a subset and let N be a ⊂ Morse gauge. The subset Y is called N-stable if it is quasi-convex and if every pair of points in Y can be connected by a geodesic which is N-Morse in X.

REMARK D.23 (Properties of stable subsets). Let X be a proper geodesic metric space and Y X. ⊂ (1) A quasi-convex subset Y X is stable if and only if every pair of points in Y can be con- ⊂ nected by a quasi-geodesic which is N-Morse for some Morse gauge N, since every such quasi-geodesic is then uniformly close to an actual geodesic, which is N ′-Morse for a uni- form N ′. (2) By definition, if Y X is stable and x, y Y , then there exists a Morse geodesic connecting ⊂ ∈ x and y inside X. Since Y is quasi-convex, this geodesic is contained in NC(Y ), and hence we can find a quasi-geodesic in Y connecting x and y of distance at most C. Thus any two points in Y can be connected inside Y by a Morse quasi-geodesic with uniform QI constants and uniform Morse gauge. (3) From (2) one deduces that every stable subspace is large-scale geodesic and quasi-geodesic. (4) Note that a quasi-geodesic proper metric space in which any two points can be joined by a uniform Morse quasi-geodesic is Morse hyperbolic by Theorem C.7. It thus follows from (2) and (3) that every stable subset of a proper geodesic metric space is necessarily Morse hyperbolic. (5) If X is a proper geodesic metric space and Y is a proper metric space, then a quasi-isometric embedding f : Y X is called a stable embedding if its image is a stable subset in the sense → of Definition D.22. This definition is equivalent to the original definition of Durham and Taylor ([DT15]). Note that if Y is a stable subset of X, then its inclusion is a quasi-isometric embedding (since Y is quasi-convex), hence a stable embedding. (6) If X is hyperbolic, then every quasi-convex subset of X is stable by definition. It is immediate from the definitions that if X, Y are proper geodesic metric spaces, then every stable embedding f : Y X is Morse-preserving. Let o Y be a basepoint and set o′ := f(o). Since → M ∈ Y is hyperbolic by Remark D.23, we have ∂s Y = ∂sY , and thus by Proposition D.9 we obtain an injection .[(( f : ∂ Y ֒ ∂M X, [(x )] [(f(x ∂ s s → s n → n which is a homeomorphism onto its image. Note that the image of ∂sf is compact, since ∂sY is. This argument does not apply directly to the inclusion of an arbitrary stable subspace ι : Y X, since Y → might not be geodesic. Nevertheless, we note that ι(Y ) is Morse hyperbolic and thus there exists a hyperbolic geodesic metric space Y ′ [Y ] and, thus a quasi-isometry g : Y ′ Y . So the map g ι ∈ geod → ◦ 140 D.MORSEBOUNDARIESOFPROPERGEODESICMETRICSPACES

′ M is a stable embedding and we get that ∂sY = ∂Y (Definition C.19) topologically embeds in ∂s X and its image is (Y ) [CH17, Theorem 3.16, Proposition 3.18]. This proves the following proposition: L PROPOSITION D.24. Let X be a proper geodesic metric space and let Y X be a stable subspace. Then ⊂ there exists a proper geodesic hyperbolic metric space Y ′ [Y ] and a quasi-isometry g : Y ′ Y , and the ∈ geod → following hold: (i) The inclusion ι : Y X induces a map ∂ (g ι): ∂ Y ′ ∂M X. → s ◦ s → s (ii) The image of ∂ (g ι) is precisely the limit set (Y ). s ◦ L In particular, Y is Morse-hyperbolic and the limit set (Y ) is a compact subset of the Morse boundary ∂M X L s and homeomorphic to the Gromov boundary ∂Y of Y in the sense of Definition C.19. 

For ease of reference we spell out the hyperbolic case:

COROLLARY D.25. If Y is a quasi-convex subset of a proper geodesic Gromov hyperbolic space X, then Y is Morse hyperbolic and its Gromov limit set (Y ) ∂X is homeomorphic to the Gromov boundary ∂Y of Y L ⊂ in the sense of Definition C.19. 

4. Weak hulls Throughout this section, X denotes a proper geodesic metric space and o X denotes a basepoint. ∈ We have seen in the previous section that limit sets of stable subsets are compact. In this section we are going to establish the converse:

PROPOSITION D.26. Let X be a proper geodesic metric space and let Z ∂M X be a subset. Then the ⊂ s o following are equivalent: (i) Z is compact. (ii) There exists a stable subset Y X such that (Y )= Z. ⊂ L As mentioned above, the implication (i) = (ii) was already established in Proposition D.24. For ⇒ M the converse implication we need a way to construct a subset of X from a given subset of ∂s Xo. In the case of a CAT(-1) space we could simply take the convex hull; in the general case, some more care has to be taken.

DEFINITION D.27. Let Z ∂M X and set Z(2) := (ξ, η) Z2 ξ = η . The weak hull H(Z) of Z is ⊂ s o { ∈ | 6 } then defined as

H(Z) := γ(R) γ : R X is bi-asymptotic to (ξ, η) X. { | → }⊂ (2) (ξ,η[)∈Z By definition, if Z < 2 then the weak hull is empty. | | REMARK D.28 (Variants of the weak hull). If X is a CAT(-1) space and Z ∂X, then for all ⊂ (ξ, η) Z(2) there is a unique bi-infinite geodesic line γ which is bi-asymptotic to (ξ, η). In the ∈ ξ,η general case there may be more than one geodesic line which is bi-asymptotic to a given (ξ, η) Z(2). ∈ However, by Theorem D.15 for each pair (ξ, η) Z(2) there exists always at least one such geodesic, ∈ and we can pick a preferred geodesic γξ,η (for example, a limit geodesic) for each pair. We can then define

Ho(Z) := γξ,η(R). (4.1) (2) (ξ,η[)∈Z If X is CAT(-1), then we simply have H(Z) = H0(Z), but in general the inclusion H0(Z) H(Z) is (N) ⊂ strict. However, if Z ∂ Xo for a fixed Morse gauge N, then it follows from Theorem D.15 that ⊂ s H0(Z) and H(Z) are at bounded Hausdorff distance. In particular this holds if X is Gromov hyperbolic or Z is relatively compact (see Remark D.5). 4. WEAK HULLS 141

EXAMPLE D.29 (Ideal polygons). If Z is a finite subset in the (Morse) boundary of the hyperbolic plane H2, then H(Z) = H (Z) H2 consists of the boundary of the ideal polygon with vertex set Z o ⊂ together with the “diagonal” geodesic lines which connect the vertices. If Z > 2, then the weak hull | | is not convex; however, since ideal polygons in hyperbolic space are slim, it is of bounded Hausdorff distance from its convex hull, and thus at least quasi-convex. In general we cannot expect the weak hull of an arbitrary subset of the Morse boundary to be quasi-convex. However, the situation is better for compact subsets ([CD19, Prop. 4.2]):

PROPOSITION D.30. Let Z ∂M X be compact. Then H(Z) is a stable subspace of X and in particular ⊂ s o Morse hyperbolic and quasi-convex.  In the case of real hyperbolic space we can do even better: We can find a closed convex set at bounded Hausdorff distance from the weak hull:

PROPOSITION D.31 (Weak hull vs. convex hull in hyperbolic spaces). Let L ∂Hn and set ⊂ (L) := conv(H(L)). H Then H(L) is at bounded Hausdorff distance from (L). H PROOF. It suffices to construct a closed convex set containing X := H(L) which is at bounded H Hausdorff distance from X. For this, using quasi-convexity of X, we fix a constant K such that every geodesic segment containing two points of X is contained in a K-neighbourhood of X. Let o H(L). Given a geodesic ray γ : [0, ) Hn emanating from o we define ∈ ∞ → t := sup t 0 γ(t) X [0, ]. γ { ≥ | ∈ } ∈ ∞ We then denote by the set of all such geodesic rays γ : [0, ) Hn which leave X in the sense Gl ∞ → that tγ < . If γ l then pγ := γ(tγ + 1) N2(X) o . There thus exists a unique hyperplane ∞ n ∈ G ∈ \{ } n H(γ,pγ) H through pγ which is perpendicular to γ at pγ . This hyperplane separates H into two ⊂ + halfspaces, and we denote by H (γ,pγ) the closure of the halfspace containing o. We now claim that H(L) is at bounded Hausdorff distance from the closed convex set := H+(γ,p ). H γ γ\∈Gl It is immediate that X , for if γ is a geodesic ray from o through some x X then the segment of ⊂ H ∈ γ from o to x is contained in . Conversely let x and let γ be a geodesic ray emanating from o H ∈⊂ H through x. If γ does not leave X or x = γ(t) with t

PROPOSITION D.32. Let Z ∂M X be compact. Then (H(Z)) = Z. ⊂ s o L PROOF. We first show that Z (H(Z)). Let η, ξ Z. By definition of the Morse limit set we have (N) ⊂ L ∈ η, ξ Xo for some Morse gauge N. Let γ ,γ be N-Morse geodesic rays based at o representing η ∈⊂ η ξ and ξ respectively. Let α be a bi-infinite geodesic line joining η and ξ. Since ideal Morse triangles are δ-slim (Theorem D.15), we know that, for some constant δ depending on N, eventually γη, respectively γ will be δ-close to α. It follows that ξ, η (H(Z)). ξ ∈ L We now show that (H(Z)) Z. Let [(x )] (H(Z)). By definition we know that (x ) H(Z), L ⊂ n ∈ L n ⊂ so xn αn, where αn is a bi-infinite geodesic line. By Proposition D.30, since Z is compact, we know ∈ ′ that the αn are all N-Morse for some Morse gauge N. Let γn and γn be two geodesic rays based at o representing the endpoints of αn. Since triangles are δ-thin (Theorem D.15), we know that xn is either ′ ′ δ away from some point on γn or some point on γn. Up to renaming γn and γn we may assume that 142 D.MORSEBOUNDARIESOFPROPERGEODESICMETRICSPACES xn is always δ-close to some point on γn. Note that the γn represent points in Z. Up to replacing by a subsequence, we know by Arzelà–Ascoli Lemma A.16 that γ converges uniformly on compact sets { n} to a geodesic ray γ based at o. Since Z is closed, γ represents a point in Z. And since the sequence of γ converges uniformly, we know that x is δ-close to γ and the result follows.  n { n} As a special case we record:

COROLLARY D.33. Let X be a proper, geodesic δ-Rips hyperbolic metric space. Let Z be a non-empty closed subset of ∂X. Then (H(Z)) = Z.  L Index

K-acyclic, 34 morphism of, 19 R-disjoint, 31 undistorted, 38 #u(v), 120 approximate group action asdim, 31 orbit, 43 N -filtered group, 15 approximate subgroup, 16 dim, 84 asymptotic dimension, 31 ℓ-dim, 85 coloring definition of, 32, 85 µ dim, 84 uniformly bounded, 85 kth filtration step, 15 u-acceptance state, 121 bi-asymptotic, 136 u-blocker, 120 bi-invariant partial order, 127 bi-Lipschitz embedding, 107 action bi-Lipschitz equivalence, 107 approximate group action Bonk-Schramm realization, 67 minimal, 69 bornologous, 107 group action map, 107 cobounded, 44, 119 bounded cocompact, 119 -bounded, 107 E geometric, 44, 119 bushy simplicial tree, 99 locally bounded, 119 metrically proper, 119 canonical coarse class, 29 minimal, 69 canonical metric (on a graph), 110 proper, 44 canonical QI-type, 118 algebraically finitely-generated approx. group, Cayley graph, 119 35 close functions, 106 almost group, 17 coarse bounded geometry ambient group, 15 space of, 111 apogee, 37 coarse class approximate group, 16 of a lcsc group, 116 K-acyclic, 34 coarse embedding, 106 k-local morphism of, 19 coarse ends, 33, 113 approximate subgroup of, 17 coarse equivalence, 106 asdim of, 32 coarse invariant, 31, 107 coarse invariant of, 31 coarse inverse, 106 countable, 16 coarse isometric embedding, 107 distorted, 38 coarse isometry, 107 filtered, 16 coarse map, 107 finite, 16 equivalence, 107 good model, 28 coarse space, 107 Higson corona of, 32 coarse structure, 107 isometric action of, 43 generated by d, 107

143 144 INDEX coarsely connected, 110 bi-Lipschitz, 107 coarsely expansive, 106 coarse isometric, 107 coarsely geodesic space, 110 isometric, 107 coarsely Lipschitz, 106 quasi-isometric, 107 coarsely proper, 106 ends coarsely quasi-symmetric, 37 c-ends, 113 coarsely surjective, 106 coarse ends, 113 cobounded, 45 space of, 112 cobounded action, 44 entourages, 107 cobounded map, 44 enveloping group, 16 cobounded space, 37 enveloping semigroup, 18 compactly generated group, 117 essentially surjective, 106 conical limit point, 74 exponential growth rate, 101 conical limit set, 74 external QI type, 35 connected components c-connected components, 113 filtered group, 15 controlled sets, 107 l-shift, 16 convergence, 131 filtered morphism, 16 convex hull, 67 k-component, 16 convexity parameter, 39 k-local, 16 cover, 84 image, 16 colored, 85 kernel, 16 covering dimension, 84 partial image, 16 covering family, 91 partial kernel, 16 cut-and-project set, 27 fixpoints at infinity, 69 cyclic reduction, 122 Freiman k-homomorphism, 19 cyclic word, 122 associated, 122 generalized Cayley graph, 37 cyclical occurrence, 122 geodesic cyclically equivalent words, 122 line, 108 cyclically reduced word, 122 linearly reparametrized, 109 ray, 108 De Bruijn–Martin graph, 121 segment, 108 defect of a function, 125 geodesic space, 109 Delone geometric action, 44 (r, R)-Delone, 105 geometrically finitely generated approx. parameters, 105 group, 35 set, 105 global morphism, 19 set in a lcsc group, 116 global quasimorphism, 20 dimension good model, 28 capacity, 85 Gromov boundary covering, 84 of a hyperbolic approximate group, 61 linearly controlled metric, 83, 85 ray model of, 132 metric, 84 sequential, 71, 132 microscopic Assouad-Nagata, 85 Gromov compactification, 71 topological, 84 Gromov hyperbolic, 129 dominating function, 112 metric space, 129 doubling at small scales, 63 Gromov limit set doubling property, 63 of a subset, 71 of qiqac, 71 embedding Gromov product, 129 INDEX 145

Gromov topology, 132 left-quasi-subgroup, 17 group left-quasimorphism, 124 compactly generated, 117 left-syndetic, 16 hyperbolic, 60 limit geodesic line, 138 lcsc, 29, 115 limit legs, 138 group extension, 26 limit point bounded, 26 conical, 74 symmetrically bounded, 26 limit set growth, 40 conical, 74 bounded growth, 65 limit triangle, 138 equivalence class, 112 linearly controlled metric dimension, 83, 85 exponential, 40 locally bounded external, 40 pseudo-metric, 115 internal, 40 locally quasi self-similar, 62 polynomial, 40 locally quasi-similar, 62 subexponential, 40 locally self-similar, 62 growth function, 40 lower control function, 106

Higson compactification, 32 map Higson corona, 32 cobounded, 44 Higson function, 32 proper, 44 homotopy functor, 34 mesh, 84 hyperbolic metric approximate group, 60 continuous, 117 Gromov, 129 external, 35 group, 60 internal, 37 elementary, 68 visual, 134 non-elementary, 68 metric coarse category, 106 Morse, 130 metric dimension, 84 Rips, 129 metric quotient, 105 shashlik, 89 metric space cobounded, 45 independent pair of words, 123 metric tree, 90 internal QI type, 36 Meyer set, 27 inverse edge pair, 121 uniform, 27 isometric action, 43 model set, 27 isometric embedding, 107 uniform, 27 isometry, 107 morphism, 19 minimal extension, 26 large-scale geodesic space, 110 extension of, 26 lcsc group, 29, 115 finite extension, 26 Lebesgue number, 84 maximal extension, 26 of covering family, 91 partial image, 20 left-adapted partial kernel, 20 pseudo-metric, 115 Morse left-admissible boundary, 135 pseudo-metric, 115 sequential, 136 left-commensurable, 17 sequential model, 136 left-defect set, 20, 124 gauge, 109 left-invariant preserving, 137 pseudo-metric, 115 quasi-geodesic, 109 146 INDEX

quasi-geodesic line, 109 rough quasi-action, 46 quasi-geodesic ray, 109 uniform, 46 Morse hyperbolic, 59, 130 weak, 46 space, 59, 130 quasi-action Morse limit set, 73 external left-regular, 53 multiplicity internal left-regular, 54 of cover, 84 qiqac, 46 quasi-isometric, 46 non-elementary, 69 quasi-cobounded non-overlapping words, 123 semi-uniformly, 51 non-self-overlapping word, 123 space, 37 norm, 115 uniformly, 51 quasi-conjugate, 48 open cover, 84 quasi-conjugation, 47 order quasi-convex subspace, 109 bi-invariant partial, 127 quasi-geodesic of cover, 84 line, 108 sandwiched, 127 ray, 108 order semigroup, 127 segment, 108 path quasi-geodesic space, 109 c-path, 110 quasi-homomorphism, 124 perturbed homomorphism, 125 quasi-invariant, 43 positivity set Pos(f), 128 quasi-isometric embedding, 107 proper quasi-isometry, 107, 134 action, 44 quasi-isometry group, 108 map, 44, 107 quasi-kernel, 24 metric, 124 quasi-lattice, 111 space, 105 quasi-metric, 134 pseudo-metric, 105 quasi-orbit, 46 left-adapted, 115 quasi-restriction, 48, 49 left-admissible, 115 quasimorphism, 20, 124 left-invariant, 115 k-local, 20 locally bounded, 115 counting, 125 space, 105 cyclic, 125 pseudo-norm, 115 defect, 125 homogeneous, 125 QI kernel, 37 homogeneous counting, 125 QI-type, 108 canonical, 118 reduced product, 119 qiqac, 46 refinement (word decomposition), 123 associated morphism, 46 simple refinement, 123 boundary action of, 72, 73 relatively compact, 17 cobounded, 49 relatively dense convex cocompact, 79 R-relatively dense, 105 geometric, 49 right-commensurable, 17 Gromov limit set of, 71 right-defect set, 124 induced, 47 right-quasi-subgroup, 18 limit set, 73 right-quasimorphism, 124 proper, 49 right-syndetic, 16 quasi-isometric-quasi-action, 46 Rips-hyperbolic, 129 roqac, 46 rough geodesic line, 108 INDEX 147 rough geodesic ray, 108 of boundary of approximate group, 61 rough geodesic segment, 108 triangle roughly geodesic space, 109 slim, 129 roughly similar spaces, 65 uniform approximate lattice, 27 sandwiched order, 127 uniform bornology, 107 saturation uniform collection of a family by a family, 91 of coarse equivalences, 111 of a set by a family, 91 of maps, 108 section of quasi-isometries, 111 good, 26 uniformly bounded family, 31 self-similar uniformly discrete, 105 locally, 62 r-uniformly discrete, 105 semi-uniformly quasi-cobounded, 51 uniformly locally finite subset, 111 sequential Gromov boundary, 132 uniformly quasi-cobounded, 51 simplicial tree, 90 unital subset, 15 bushy, 99 upper control function, 106 locally finite, 90 slim triangle, 129 Vietoris–Rips complex, 33 space of ends, 112 visual inequalities, 134 stable embedding, 139 visual metric, 62, 134 stable subset, 78, 139 visual space, 60 subextension, 26 visuality parameter, 134 symmetric map, 15 symmetric subset, 15 weak hull, 66, 140 weakly quasi-convex, 39 topological dimension, 84 weight function, 117 coloring definition of, 85 word metric, 117

Bibliography

[ALC18] Jesús A. Álvarez López and Alberto Candel. Generic coarse geometry of leaves, volume 2223 of Lecture Notes in Mathe- matics. Springer, Cham, 2018. [BD01] G. Bell and A. Dranishnikov. On asymptotic dimension of groups. Algebr. Geom. Topol., 1:57–71, 2001. [BD08] G. Bell and A. Dranishnikov. Asymptotic dimension. Topology Appl., 155(12):1265–1296, 2008. [BDLM08] N. Brodskiy, J. Dydak, M. Levin, and A. Mitra. A Hurewicz theorem for the Assouad-Nagata dimension. J. Lond. Math. Soc. (2), 77(3):741–756, 2008. [BG08] Jean Bourgain and Alex Gamburd. Uniform expansion bounds for Cayley graphs of SL2(Fp). Ann. of Math. (2), 167(2):625–642, 2008. [BG13] Michael Baake and Uwe Grimm. Aperiodic order. Vol. 1, volume 149 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 2013. A mathematical invitation, With a foreword by Roger Penrose. [BGT12] Emmanuel Breuillard, Ben Green, and Terence Tao. The structure of approximate groups. Publ. Math. Inst. Hautes Études Sci., 116:115–221, 2012. [BH99] Martin R. Bridson and André Haefliger. Metric spaces of non-positive curvature, volume 319 of Grundlehren der Mathe- matischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999. [BH18a] Michael Björklund and Tobias Hartnick. Approximate lattices. Duke Math. J., 167(15):2903–2964, 2018. [BH18b] Michael Björklund and Tobias Hartnick. Spectral theory of approximate lattices in nilpotent lie groups. Preprint, 2018. on arXiv: https://arxiv.org/abs/1811.06563. [BH21] Michael Björklund and Tobias Hartnick. Analytic properties of approximate lattices. to appear in Annales de l’Institut Fourier, 2021. [BHM11] Sébastien Blachère, Peter Haïssinsky, and Pierre Mathieu. Harmonic measures versus quasiconformal measures for hyperbolic groups. Ann. Sci. Éc. Norm. Supér. (4), 44(4):683–721, 2011. [BHP18] Michael Björklund, Tobias Hartnick, and Felix Pogorzelski. Aperiodic order and spherical diffraction, I: auto- correlation of regular model sets. Proc. Lond. Math. Soc. (3), 116(4):957–996, 2018. [BHS19] Michael Björklund, Tobias Hartnick, and Thierry Stulemeijer. Borel density for approximate lattices. Forum of Math- ematics, Sigma, 7:e40, 2019. [Bir34] Garrett Birkhoff. The topology of transformation-sets. Ann. of Math. (2), 35(4):861–875, 1934. [BJ97] Christopher J. Bishop and Peter W. Jones. Hausdorff dimension and Kleinian groups. Acta Math., 179(1):1–39, 1997. [BL07] S. V. Buyalo and N. D. Lebedeva. Dimensions of locally and asymptotically self-similar spaces. Algebra i Analiz, 19(1):60–92, 2007. [BM91] Mladen Bestvina and Geoffrey Mess. The boundary of negatively curved groups. J. Amer. Math. Soc., 4(3):469–481, 1991. [BOT13] M. Burger, N. Ozawa, and A. Thom. On Ulam stability. Israel J. Math., 193(1):109–129, 2013. [Bre14] Emmanuel Breuillard. A brief introduction to approximate groups. In Thin groups and superstrong approximation, volume 61 of Math. Sci. Res. Inst. Publ., pages 23–50. Cambridge Univ. Press, Cambridge, 2014. [Bro10] L.E.J. Brouwer. On the structure of perfect sets of points. Proc. Akad. Amsterdam, 12:785–794, 1910. [Bro81] Robert Brooks. Some remarks on bounded cohomology. In Riemann surfaces and related topics: Proceedings of the 1978 Stony Brook Conference (State Univ. New York, Stony Brook, N.Y., 1978), volume 97 of Ann. of Math. Stud., pages 53–63. Princeton Univ. Press, Princeton, N.J., 1981. [BS00] M. Bonk and O. Schramm. Embeddingsof Gromov hyperbolic spaces. Geom. Funct. Anal., 10(2):266–306, 2000. [BS05] Sergei Buyalo and Viktor Schroeder. Embedding of hyperbolic spaces in the product of trees. Geom. Dedicata, 113:75– 93, 2005. [BS07a] Sergei Buyalo and Viktor Schroeder. Elements of asymptotic geometry. EMS Monographs in Mathematics. European Mathematical Society (EMS), Zürich, 2007. [BS07b] Sergei Buyalo and Viktor Shroeder. The hyperbolic dimension of metric spaces. Algebra i Analiz, 19(1):93–108, 2007. [BSH12] Gabi Ben Simon and Tobias Hartnick. Reconstructing quasimorphisms from associated partial orders and a question of Polterovich. Comment. Math. Helv., 87(3):705–725, 2012. [Buy05] S. V. Buyalo. Capacity dimension and embedding of hyperbolic spaces into the product of trees. Algebra i Analiz, 17(4):42–58, 2005. [BW92] Jonathan Block and Shmuel Weinberger. Aperiodic tilings, positive scalar curvature and amenability of spaces. J. Amer. Math. Soc., 5(4):907–918, 1992.

149 150 BIBLIOGRAPHY

[Cal09] Danny Calegari. scl, volume 20 of MSJ Memoirs. Mathematical Society of Japan, Tokyo, 2009. [Car15] Pietro Kreitlon Carolino. The Structure of Locally Compact Approximate Groups. ProQuest LLC, Ann Arbor, MI, 2015. Thesis (Ph.D.)–University of California, Los Angeles. [CD19] Matthew Cordes and Matthew Gentry Durham. Boundary convex cocompactness and stability of subgroups of finitely generated groups. Int. Math. Res. Not. IMRN, (6):1699–1724, 2019. [CdlH16] Y. Cornulier and P. de la Harpe. Metric geometry of locally compact groups, volume 25 of EMS Tracts in Mathematics. European Mathematical Society (EMS), Zürich, 2016. [CH17] Matthew Cordes and David Hume. Stability and the Morse boundary. J. Lond. Math. Soc. (2), 95(3):963–988, 2017. [CI11] Ionut Chifan and Adrian Ioana. On relative property (T) and Haagerup’s property. Trans. Amer. Math. Soc., 363(12):6407–6420, 2011. [Coo15] Michel Coornaert. Topological dimension and dynamical systems. Universitext. Springer, Cham, 2015. Translated and revised from the 2005 French original. [Cor17] Matthew Cordes. Morse boundaries of proper geodesic metric spaces. Groups Geom. Dyn., 11(4):1281–1306, 2017. [CS15] Ruth Charney and Harold Sultan. Contracting boundaries of CAT(0) spaces. J. Topol., 8(1):93–117, 2015. [dB46] N. G. de Bruijn. A combinatorial problem. Nederl. Akad. Wetensch., Proc., 49:758–764 = Indagationes Math. 8, 461–467 (1946), 1946. [Dra00] A. N. Dranishnikov. Asymptotic topology. Uspekhi Mat. Nauk, 55(6(336)):71–116, 2000. [DS06] A. Dranishnikov and J. Smith. Asymptotic dimension of discrete groups. Fund. Math., 189(1):27–34, 2006. [DT15] Matthew Gentry Durham and Samuel J. Taylor. Convex cocompactness and stability in mapping class groups. Al- gebr. Geom. Topol., 15(5):2839–2859, 2015. [Eng95] Ryszard Engelking. Theory of dimensions finite and infinite, volume 10 of Sigma Series in Pure Mathematics. Heldermann Verlag, Lemgo, 1995. [EP00] Y. Eliashberg and L. Polterovich. Partially ordered groups and geometry of contact transformations. Geom. Funct. Anal., 10(6):1448–1476, 2000. [ES83] P. Erd˝os and E. Szemerédi. On sums and products of integers. In Studies in pure mathematics, pages 213–218. Birkhäuser, Basel, 1983. [Fa˘i94] V. A. Fa˘iziev. Pseudocharacters on free groups. Izv. Ross. Akad. Nauk Ser. Mat., 58(1):121–143, 1994. [Fin17] Elisabeth Fink. Morse geodesics in torsion groups. arXiv, 2017. [FK16] Koji Fujiwara and Michael Kapovich. On quasihomomorphisms with noncommutative targets. Geom. Funct. Anal., 26(2):478–519, 2016. [Fre73] G. A. Fre˘ıman. Foundations of a structural theory of set addition. American Mathematical Society, Providence, R. I., 1973. Translated from the Russian, Translations of Mathematical Monographs, Vol 37. [GdlH90] É. Ghys and P. de la Harpe, editors. Sur les groupes hyperboliques d’après Mikhael Gromov, volume 83 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1990. Papers from the Swiss Seminar on Hyperbolic Groups held in Bern, 1988. [Gen19] A. Genevois. Criterion for visuality of hyperbolic spaces. MathOverflow, 2019. URL: https://mathoverflow.net/q/324832 (version: 2019-03-07). [Geo08] Ross Geoghegan. Topological methods in group theory, volume 243 of Graduate Texts in Mathematics. Springer, New York, 2008. [Gle52] Andrew M. Gleason. Groups without small subgroups. Ann. of Math. (2), 56:193–212, 1952. [Gri95] R. I. Grigorchuk. Some results on bounded cohomology. In Combinatorial and geometric group theory (Edinburgh, 1993), volume 204 of London Math. Soc. Lecture Note Ser., pages 111–163. Cambridge Univ. Press, Cambridge, 1995. [Gro87] Mikhael Gromov. Hyperbolic groups. In Essays in group theory, volume 8 of Math. Sci. Res. Inst. Publ., pages 75–263. Springer, New York, 1987. [Gro93] M. Gromov. Asymptotic invariants of infinite groups. In Geometric group theory, Vol. 2 (Sussex, 1991), volume 182 of London Math. Soc. Lecture Note Ser., pages 1–295. Cambridge Univ. Press, Cambridge, 1993. [Hel08] H. A. Helfgott. Growth and generation in SL2(Z/pZ). Ann. of Math. (2), 167(2):601–623, 2008. [Heu20] Nicolaus Heuer. Low-dimensional bounded cohomology and extensions of groups. Math. Scand., 126(1):5–31, 2020. [HO13] Michael Hull and Denis Osin. Induced quasicocycles on groups with hyperbolically embedded subgroups. Algebr. Geom. Topol., 13(5):2635–2665, 2013. [Hru12] Ehud Hrushovski. Stable group theory and approximate subgroups. J. Amer. Math. Soc., 25(1):189–243, 2012. [Hru20] Ehud Hrushovski. Beyond the lascar group. Preprint, 2020. on arXiv: https://arxiv.org/abs/2011.12009. [HS16] Tobias Hartnick and Pascal Schweitzer. On quasioutomorphism groups of free groups and their transitivity proper- ties. J. Algebra, 450:242–281, 2016. [HS19] Tobias Hartnick and Alessandro Sisto. Bounded cohomology and virtually free hyperbolically embedded sub- groups. Groups Geom. Dyn., 13(2):677–694, 2019. [HT18] Tobias Hartnick and Alexey Talambutsa. Relations between counting functions on free groups and free monoids. Groups Geom. Dyn., 12(4):1485–1521, 2018. [HW41] Witold Hurewicz and Henry Wallman. Dimension Theory. Princeton Mathematical Series, vol. 4. Princeton University Press, Princeton, N. J., 1941. BIBLIOGRAPHY 151

[HW21] Tobias Hartnick and Stefan Witzel. Higher finiteness properties of arithmetic approximate lattices: the rank theorem for number fields. Preprint in preparation, 2021. [Liu19] Qing Liu. Dynamics on the Morse boundary. arXiv, 2019. [LS05] Urs Lang and Thilo Schlichenmaier. Nagata dimension, quasisymmetric embeddings, and Lipschitz extensions. Int. Math. Res. Not., (58):3625–3655, 2005. [Mac20a] Simon Machado. Approximate lattices in higher-rank semi-simple groups. Preprint, 2020. on arXiv: https://arxiv.org/abs/2011.01835. [Mac20b] Simon Machado. Good models, infinite approximate subgroups and approximate lattices. Preprint, 2020. on arXiv: https://arxiv.org/abs/2011.01829. [Mar95] Reiner Martin. Non-uniquely ergodic foliations of thin-type, measured currents and automorphisms of free groups. ProQuest LLC, Ann Arbor, MI, 1995. Thesis (Ph.D.)–University of California, Los Angeles. [Mey72] YvesMeyer. Algebraic numbers and harmonic analysis. North-Holland Publishing Co., Amsterdam-London; American Elsevier Publishing Co., Inc., New York, 1972. North-Holland Mathematical Library, Vol. 2. [MM99] Howard A. Masur and Yair N. Minsky. Geometry of the complex of curves. I. Hyperbolicity. Invent. Math., 138(1):103–149, 1999. [Mor24] Harold Marston Morse. A fundamental class of geodesics on any closed surface of genus greater than one. Trans. Amer. Math. Soc., 26(1):25–60, 1924. [MSW00] Lee Mosher, Michah Sageev, and Kevin Whyte. Quasi-actions on trees i. bounded valence. Annals of Mathematics, 158, 11 2000. [Mun75] James R. Munkres. Topology: a first course. Prentice-Hall, Inc., Englewood Cliffs, N.J., 1975. [MZ52] Deane Montgomery and Leo Zippin. Small subgroups of finite-dimensional groups. Ann. of Math. (2), 56:213–241, 1952. [Nic89] Peter J. Nicholls. The ergodic theory of discrete groups, volume 143 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1989. [OOS09] Alexander Yu. Olshanskii, Denis V. Osin, and Mark V. Sapir. Lacunary hyperbolic groups. Geom. Topol., 13(4):2051– 2140, 2009. [Osi16] D. Osin. Acylindrically hyperbolic groups. Trans. Amer. Math. Soc., 368(2):851–888, 2016. [Pau10] Vern Paulsen. Syndetic sets and amenability. Proceedings of the American Mathematical Society, 140, 02 2010. [Plü69] Helmut Plünnecke. Eigenschaften und Abschätzungen von Wirkungsfunktionen. BMwF-GMD-22. Gesellschaft für Math- ematik und Datenverarbeitung, Bonn, 1969. [Roe03] John Roe. Lectures on coarse geometry, volume 31 of University Lecture Series. American Mathematical Society, Provi- dence, RI, 2003. [Rog70] C. A. Rogers. Hausdorff measures. Cambridge University Press, London-New York, 1970. [Ruz99] Imre Z. Ruzsa. An analog of Freiman’s theorem in groups. Number 258, pages xv, 323–326. 1999. Structure theory of set addition. [Sa95] Jacek Swi´ ˛atkowski. On the asymptotic homological dimension of hyperbolic groups. Bull. London Math. Soc., 27(3):209–221, 1995. [Sis16] Alessandro Sisto. Quasi-convexity of hyperbolically embedded subgroups. Mathematische Zeitschrift, pages 1–10, 2016. [Sta09] Mihai D. Staic. Symmetric cohomology of groups in low dimension. Arch. Math. (Basel), 93(3):205–211, 2009. [Swe01] Eric L. Swenson. Quasi-convex groups of isometries of negatively curved spaces. Topology Appl., 110(1):119–129, 2001. Geometric topology and geometric group theory (Milwaukee, WI, 1997). [Tao08] Terence Tao. Product set estimates for non-commutative groups. Combinatorica, 28(5):547–594, 2008. [Tao14] Terence Tao. Hilbert’s fifth problem and related topics, volume 153 of Graduate Studies in Mathematics. American Mathe- matical Society, Providence, RI, 2014. [Ula60] S. M. Ulam. A collection of mathematical problems. Interscience Tracts in Pure and Applied Mathematics, no. 8. Inter- science Publishers, New York-London, 1960.