Uniform simplicity of groups with proximal action

By Światosław R. Gal and Jakub Gismatullin, with an appendix by Nir Lazarovich

Abstract

We prove that groups acting boundedly and order-primitively on linear orders or acting extremely proximally on a Cantor set (the class including various Higman-Thomson groups; Neretin groups of almost automorphisms of regular trees, also called groups of spheromorphisms; the groups of quasi-isometries and almost-isometries of regular trees) are uniformly simple. Explicit bounds are provided.

1. Introduction

Let be a group. It is called -uniformly simple if for every nontrivial and nontrivial conjugacy class the element is the product of at most elements from . A group is uniformly simple if it is -uniformly simple for some natural number . Uniformly simple groups are called sometimes, by other authors, groups with finite covering number or boundedly simple groups (see, e.g., Reference 15Reference 19Reference 22). We call boundedly simple if is allowed to depend on . The purpose of this paper is to prove results on uniform simplicity, in particular Theorems 1.1, 1.2, and 1.3 below, for a number of naturally occurring infinite permutation groups.

Every uniformly simple group is simple. It is known that many groups with geometric or combinatorial origin are simple. In this paper we prove that, in fact, many of them are uniformly simple.

Below are our main results.

Let be a linearly ordered set. Let denote the group of order-preserving bijections of . We say that is boundedly supported if there are such that only if . The subgroup of boundedly supported elements of will be denoted by .

Theorem 1.1 (Theorem 3.1 below).

Assume that is proximal on a linearly ordered set (i.e., for every and from there exists such that ). Then its commutator group is six-uniformly simple and the commutator width of this group is at most two.

For the definition of a commutator width of a group see the beginning of Section 2. Observe that every doubly-transitive (i.e., transitive on ordered pairs) action is proximal.

This theorem immediately applies e.g. to and the class of Higman-Thomson groups for , where the latter are defined as follows. We fix natural numbers . The Higman-Thompson group is defined as piecewise affine, order-preserving transformations of whose breaking points (i.e., singularities) belong to and the slopes are for (see Reference 5, Proposition 4.4). The Thompson group is the group in the above series. Moreover, is independent of (up to isomorphism) Reference 5, 4.1. The Higman-Thompson groups satisfy the assumptions of Theorem 1.1 due to Lemmata 3.2 and 3.3 from Section 3.

Our result implies that is six-uniformly simple. Whereas Droste and Shortt proved in Reference 13, Theorem 1.1(c) that is two-uniformly simple. In fact, they proved that if is proximal (they use the term “feebly 2-transitive” for proximal action) and additionally closed under -patching of conjugate elements, then is two-uniformly simple. Thus, our Theorem 1.1 covers a larger class of examples than that from Reference 13 (as we assume only proximality), but with slightly worse bound for uniform simplicity.

The uniform simplicity of the Thompson group was proven implicitly by Bardakov, Tolstykh, and Vershinin Reference 2, Corollary 2.3 and Burago and Ivanov Reference 6. Although their proofs generalize to the general result given above, we write it down for several reasons. Namely, in the above cited papers some special properties of the linear structure of the real line is used, while the result is true for a general class of proximal actions on linearly ordered sets. The Droste and Shortt argument uses -patching, which is not suitable for our case. Furthermore, although in the examples mentioned above the action is doubly-transitive, the right assumption is proximality, which is strictly weaker than double-transitivity. In Theorem 4.2 we construct a bounded and proximal transitive action which is not doubly-transitive. This is discussed in detail in Section 4. The second reason for proving Theorem 1.1 is that a topological analogue of proximality, namely extremal proximality (see the beginning of Section 5), plays a crucial role in the proofs of the subsequent results. Extremal proximality was defined by S. Glasner in Reference 21, p. 96 and Reference 20, p. 328 for a general minimal action of a group on a compact Hausdorff space.

In Section 5 we go away from order-preserving actions, and consider groups acting on a Cantor set, and also groups almost acting on trees. The following theorem is Corollary 6.6(2).

Theorem 1.2.

The commutator subgroup of the Neretin group of spheromorphisms and the commutator subgroup of the Higman-Thomson group are nine-uniformly simple. The commutator width of each of those groups is at most three.

The group was introduced by Neretin in Reference 27, 4.5, 3.4 as the group of spheromorphisms (also called almost automorphisms) of a -regular tree . We will recall the construction in Section 6.

The Higman-Thompson group is defined as the group of automorphisms of the Jónsson-Tarski algebra Reference 5, 4A. It can also be described as a certain group of homeomorphisms of a Cantor set Reference 5, p. 57. Moreover, one can view as a subgroup of and the latter as a group acting spheromorphically on the -regular tree Reference 25, Section 2.2, that is, they are subgroups of . If is even, then . For odd , Reference 5, Theorem 4.16, Reference 17, Theorem 5.1. The group is also known as the Thompson group . It is known that .

Given a group acting on a tree , in the beginning of Section 5 we will define, following Neretin, the group of partial actions on the boundary of . Theorem 1.2 is a corollary of a more general theorem about uniform simplicity of partial actions.

Theorem 1.3.

Assume that a group acts on a leafless tree , whose boundary is a Cantor set, such that does not fix any proper subtree (e.g., is finite) nor a point in the boundary of . Then the commutator subgroup of is nine-uniformly simple.

This is an immediate corollary of Theorems 5.1 and 6.4. The latter theorem concerns several characterizations of extremal proximality of the group action on the boundary of a tree.

Section 7 is devoted to the proof that the groups of quasi-isometries and almost-isometries of regular trees are five-uniformly simple.

The uniform simplicity of homeomorphism groups of certain spaces has been considered since the beginning of the 1960s, e.g., by Anderson Reference 1. He proved that the group of all homeomorphisms of with compact support and the group of all homeomorphisms of a Cantor set are two-uniformly simple (and have commutator width one). His arguments used an infinite iteration arbitrary close to every point, which is not suitable for the study of spheromorphism groups and the Higman-Thompson groups.

-uniform simplicity is a first-order logic property (for a fixed natural number ). That is, it can be written as a formula in a first-order logic. Therefore, -uniform simplicity is preserved under elementary equivalence: if is -uniform simple, then all other groups elementary equivalent with are also -uniformly simple. In particular, all ultraproducts of Neretin groups, and Higman-Thompson groups mentioned above, are nine-uniformly simple.

Another feature of uniformly simple groups comes from Reference 19, Theorem 4.3, where the second author proves the following classification fact about actions of uniformly simple groups (called boundedly simple in Reference 19) on trees: if a uniformly simple group acts faithfully on a tree without invariant proper subtree or invariant end, then essentially is a bi-regular tree (see Section 6 for definitions).

In Section 4 we discuss the primitivity of actions on linearly ordered sets (i.e., lack of proper convex congruences). In fact, we prove that primitivity and proximality are equivalent notions for bounded actions (Theorem 4.1). Our primitivity appears in the literature as o-primitivity; see, e.g., Reference 26, Section 7.

Calegari Reference 9 proved that subgroups of the pl1 homeomorphism of the interval, in particular the Thompson group , have trivial stable commutator length. Essentially by Reference 7, Lemma 2.4 by Burago, Polterovich, and Ivanov (that we will explain for the completeness of the presentation) we reprove in Lemma 2.2 the commutator width of the commutator subgroup (and other groups covered by Theorems 3.1 and 1.1) of the Thompson group .

Let us discuss examples and nonexamples of bounded and uniform simplicity. It is known that a simple Chevalley group (that is, the group of points over an arbitrary infinite field of a quasi-simple quasi-split connected reductive group) is uniformly simple Reference 15Reference 22. In fact, there exists a constant (which is conjecturally , at least in the algebraically closed case Reference 23) such that, any such Chevalley group is -uniformly simple, where is the relative rank of Reference 15.

Full automorphism groups of right-angled buildings are simple, but never boundedly simple, because of the existence of nontrivial quasi-morphisms Reference 10, Reference 11, Theorem 1.1 (except if the building is a bi-regular tree Reference 19, Theorem 3.2).

Compact groups are never uniformly simple. More generally, a topological group is called a sin group if every neighborhood of the identity contains a neighborhood of which is invariant under all inner automorphisms. Every compact group is sin (as if is such a neighborhood, then has nonempty interior). Clearly every infinite Hausdorff sin-group is not uniformly simple. Moreover, many compact linear groups (e.g., ) are boundedly simple, because of the presence of the dimension with good properties. (See also the discussion at the end of this section.)

Let us conclude the introduction by relating simplicity and the notion of central norms on groups. Let be a group. A function is called a seminorm if

for all , and

for all .

A seminorm is called

trivial if for all ,

central if ,

a norm if for ,

discrete if , and

bounded if .

A seminorm is discrete if and only if the topology it induces is discrete. A discrete seminorm is a norm. Every central seminorm is conjugacy invariant: .

A typical example of a nontrivial central and discrete norm is a word norm attached to a subset (cf. Reference 16, 2.1):

For a nontrivial central norm we define an invariant . Of course, if is either nondiscrete or unbounded, then . We define to be the supremum of for all nontrivial central norms on .

Proposition 1.4.

Let be a group. Then:

(1)

is simple if and only if any nontrivial central seminorm on is a norm;

(2)

is boundedly simple if and only if every central seminorm on is a bounded norm;

(3)

if is uniformly simple, then every central seminorm on is a bounded and discrete norm;

(4)

is -uniformly simple if and only if .

Proof.

It is obvious that the kernel of a central seminorm is closed under multiplication and conjugacy invariant. Thus, it is a normal subgroup. On the other hand, if is a proper normal subgroup of , then

is a nontrivial central seminorm. It is a norm only if .

Suppose that is a central seminorm and assume that is boundedly simple. Choose . There exists such that every element is a product of at most conjugates of and . Thus, by the triangle inequality and conjugacy invariance, . The number is finite and independent on . For the converse take and consider the word norm attached to . It is obvious that is a central seminorm on . Thus is boundedly simple, as is bounded.

Suppose is -uniformly simple, i.e., is independent on and takes nontrivial central norm . By the triangle inequality we conclude that , which proves the necessity of the condition in . For the converse, take , and consider the word norm above. We have that is -uniformly simple. This completes the proof of (4), which implies (3).

In particular, the infinite alternating group is simple but not boundedly simple. To see the latter, observe that the cardinality of the support is an unbounded central norm. This norm is maximal up to scaling. Indeed, essentially by Reference 14, Lemma 2.5 every element is a product of at most conjugates of .

Also, it is easy to see that is boundedly simple, but not uniformly simple. The angle of rotation is a full invariant of an element of that group, and this function is a central norm. Clearly it is not discrete. As before, one can observe that if is a rotation by an angle , then every other rotation can be obtained by at most conjugates of .

Every universal sofic group Reference 14, Section 2 is boundedly simple, but not uniformly simple. Namely, let be the full symmetric group on letters with the normalized Hamming norm: for . Take

as the metric ultraproduct of over a nonprincipal ultrafilter . Then the proof of Reference 14, Proposition 2.3(5) shows that is boundedly simple. Furthermore, is not uniformly simple, as is a nondiscrete central norm on (see Proposition 1.4). As before, this norm is maximal up to scaling due to Reference 14, Lemma 2.5.

2. Burago-Ivanov-Polterovich method

The symbol will always denote a group. For we use the following notation: and . By we mean the conjugacy class of .

Let be a nontrivial conjugacy class in . By -commutator we mean an element of . If we will use the name -commutator as a synonym of -commutator, for short. Of course , thus the set of -commutators is closed under inverses and conjugation.

The commutator length of an element is the minimal number of commutators sufficient to express as their product. The commutator width of is the maximum of the commutator lengths of elements of its derived subgroup .

We say that and commute up to conjugation if there exist such that and commute.

Lemma 2.1.

Assume that and commute. Then is a product of two -commutators. More precisely can be written as a product of two conjugates of and two conjugates of by elements from the group generated by and .

Proof.

We have . Also, , since commute with .

Following Burago, Ivanov, and Polterovich Reference 7, Sec. 2.1 assume that is a subgroup, , and . We say that an element -displaces if

(hence also for ).

We will say that displaces if it 1-displaces . We say that is -displaceable in if there exists such that -displaces (this property is called strongly -displaceable in Reference 7, Sec. 2.1). In particular, elements of a displaceable subgroup commute up to conjugation.

Lemma 2.2 (Reference 7, Lemma 2.5).

Assume that -displaces . Let be a product of at most commutators . Then there exist , , and such that .

Burago, Polterovich, and Ivanov Reference 7, Theorem 2.2(i) proved that if for every some conjugate of -displaces , then every element of is a product of seven -commutators. We get a better result under a stronger assumption.

Proposition 2.3.

Assume that is such that for every finitely generated subgroup and , there exists a conjugate of which -displaces . Then every element of is a product of two commutators in and three -commutators in . Moreover,

Proof.

Every element can be expressed as a product of commutators of elements of for some . Call the group they generate . Since some conjugate of , say , -displaces , by Lemma 2.2, there exist , , and such that .

Since some conjugate of displaces the group generated by and , by Lemma 2.1, is a product of two -commutators. Thus is a product of three -commutators.

The “moreover” part follows from the fact that and that and are equal. The last claim follows from the fact that if commutes up to commutation with , then

Note that the assumption of the above corollary implies that neither nor is finitely generated. However, we will use this approach to prove uniform simplicity of the Higman-Thompson groups which are known to be finitely generated.

Lemma 2.4.

Assume that every two elements in commute up to conjugation. Then every commutator in can be expressed as a commutator in . In particular, is perfect.

Proof.

Let and belong to . Choose and such that and commute and also and commute. Then

Proposition 2.5.

Let displace . Assume that, for every , every finitely generated subgroup is -displaceable in . Then every element of is a product of four -commutators from . In particular .

Proof.

By Lemma 2.2 every element of is a product of two commutators of . By Lemma 2.4 they can be chosen to be commutators of elements of . By Lemma 2.1 each of them is a product of two -commutators over .

3. Bounded actions on ordered sets

The purpose of this section is to prove that numerous simple Higman-Thompson groups acting as order-preserving piecewise-linear transformations are, in fact, uniformly simple.

We always assume that a group acts faithfully on the left by order-preserving transformations on a linearly ordered set . Given a map , we define the support of to be . Given and we define . By we will denote the set . The group of all bounded automorphisms of is denoted by .

We call such an action

proximal, if for every such that and there is satisfying ;

bounded, if for every there are such that .

Note that being proximal implies that is dense without endpoints.

Theorem 3.1.

Assume that acts faithfully, order-preserving, boundedly, and proximally on a linearly ordered set . Then its commutator group is six-uniformly simple and the commutator width of is at most two.

Proof.

We apply Proposition 2.3. Let be an arbitrary nontrivial element of . Let be such that . Replacing by we may assume that . Choose such that . Then . Let be an arbitrary finitely generated subgroup of . Then there exists an interval, say , containing supports of all generators of , hence also containing supports of all elements of . By the proximality of the action, we may assume (possibly conjugating ) that . It is clear that such a conjugate of -displaces . Thus Proposition 2.3 applies.

Let us apply Theorem 3.1 to the Higman-Thompson groups of order-preserving piecewise-linear maps. We first recall the definitions. Let be integers. Recall that (, respectively) is defined as piecewise affine (we allow only finitely many pieces), order-preserving bijections of (, respectively) whose breaking points of the derivatives belong to and the slopes are for (see the bottom of page 53 and the top of page 56 in Reference 5).

Define (, respectively) to be the subgroup of (, respectively) consisting of all such transformations that are boundedly supported, that is, for some (, respectively).

We use the following lemma. The first part of it is a known result Reference 3.

Lemma 3.2.
(1)

The groups and are isomorphic Reference 3, Proposition C10.1.

(2)

The commutator subgroups of and are equal.

Proof.

(2) It is obvious that . Let us prove . Note that (because for , the element acts as the identity in some small neighborhoods of and ). Thus, if , then for some . Therefore . A slight modification of above gives which

sends piecewise affinely onto ,

is the identity on ,

sends piecewise affinely onto .

Then is another isomorphism between and such that (we regard as a subgroup of ). Write for . Then .

We consider the action of on and its orbits. Let be the ideal of generated by .

Lemma 3.3 (Reference 3, Theorem A4.1, Corollary A5.1).
(1)

is -invariant.

(2)

acts in a doubly-transitive way on . In particular, the action is proximal.

As a corollary of the above lemmata we get that groups satisfy the assumptions of Theorem 1.1.

Corollary 3.4.

is six-uniformly simple and the commutator width of it is at most two.

Remark 3.5.

Theorem 3.1 applies to the following groups.

Bieri and Strebel Reference 3 define a more general class of groups acting boundedly on . They take a subgroup in the multiplicative group and a -submodule and define to be a group of boundedly supported automorphisms of consisting of piecewise affine maps with slopes in and singularities in . They define an augmentation ideal of and prove that acts highly transitively on . Thus is six-uniformly simple.

Another example of doubly-transitive and bounded action on a linear order (thus satisfying the assumptions of Theorem 3.1) was considered by Chehata in Reference 12, who studied partially affine transformations of an ordered field and proved that this group is simple. Theorem 3.1 implies that the Chehata group is six-uniformly simple.

4. Proximality, primitivity, and double-transitivity

In this section we prove (Theorem 4.1) that proximality (from the previous section) and order-primitivity are equivalent properties for bounded group actions. In general, these properties are inequivalent. The action of the group of integers on itself is primitive but neither proximal nor bounded. We also give an example of bounded, transitive, and proximal action, which is not doubly-transitive (Theorem 4.2).

An action of a group on a linearly ordered set is called primitive (or order-primitive by some authors), if for any other linearly ordered set and homomorphism and order-preserving equivariant map (that is, ), the map is injective or is a singleton.

Theorem 4.1.

Every proximal action is primitive. Any bounded and primitive action is proximal.

Proof.

Assume the action is not primitive. Choose , , and such that and . Reversing the order if necessary, we may assume . Set . This choice contradicts proximality, as if , then

Assume that action is bounded, but not proximal. Let , , , and witness the latter. For , , consider the relation on defined as

if and there is no such that .

By the assumption . Let be the transitive closure of . The symmetric closure of is transitively closed, thus is an equivalence relation, which has convex classes. Moreover, is -invariant, that is, implies for all . It is enough to prove that is not total, that is, for some , because then the quotient map

proves nonprimitivity of the action ( has a natural -action).

First, we claim that there is such that . Indeed, if there is no such group element, define a map by the formula

This map would contradict primitivity.

Choose and from such that . Then is a countable family of intervals in , which are pairwise disjoint. We claim that , as otherwise there are , , such that and contains for some , which is impossible.

Clearly, if acts proximally on , then it acts in such a way on any orbit. Thus, we will restrict to transitive actions.

Examples of actions we discuss above are doubly-transitive (cf. Lemma 3.3(2) and Remark 3.5). Thus they are proximal. This property seems to be easier to check than double-transitivity. We construct below an example of bounded, transitive, and proximal action, which is not doubly-transitive. The reader may compare this result with a result of Holland Reference 24, Theorem 4, which says that every bounded, transitive, primitive, and closed under , action must be doubly-transitive. Moreover, any group acting boundedly and transitively cannot be finitely generated. Indeed, a finite number of elements have supports in a common bounded interval, thus the whole group is supported in that interval, so does not act transitively.

Theorem 4.2.

There exists a subgroup acting transitively and proximally but not doubly-transitively.

Proof.

For each we will define a countable linear order , a group acting on it, and a function such that:

(1)

;

(2)

is a -equivariant linear bounded suborder of ;

(3)

for , acts transitively and proximally on by order-preserving transformations (but not doubly-transitively);

(4)

is -invariant: for , , and .

Then we take , which acts boundedly, transitively, and proximally, but not doubly-transitively on , because of , which is a -invariant map .

Since is a countable and, by proximality, dense linear order without ends, it is isomorphic to .

In the following inductive construction we will define three auxiliary points from .

We put and , where acts on by translations. Let and , , .

Assume we have constructed , , and . Let

and for all . In plain words, consists of all functions from to which differ from a constant function (denoted by ) taking the value , only at finitely many places. Define a linear order on by putting if , with the convention that for all . Note that embeds into :

Consider , with the following action of :

Define

The interval contains the embedded copy of .

Extend the action of to the whole of by the identity on the complement . Thus the action of on is bounded. Define yet another automorphism of by . Let be the group generated by and . The action of on is clearly transitive.

For every pair from , define .

For and let be such that (such exists by proximality of the action of on ). Then

which proves the proximality of the action of on .

Finally, define . Clearly, is -invariant, hence the action of on is not doubly-transitive.

The element stabilizes and has unbounded orbits on . Thus the stabilizer of has unbounded orbits on . This is enough to conclude that the action is proximal.

Question 4.3.

Is there any transitive, proximal bounded action without the property that point stabilizers have unbounded orbits?

5. Extremely proximal actions on a Cantor set and uniform simplicity

The main goal of the present section is prove Theorem 5.1, which gives a criterion for a group acting on a Cantor set to be nine-uniformly simple.

Let be a Cantor set. Assume that a discrete group acts on by homeomorphisms. By the topological full group of we define (see, e.g., Reference 18)

Throughout this section we assume that:

the group acts faithfully by homeomorphisms on a Cantor set ;

is a topological full group, i.e., ;

the action is extremely proximal, i.e., for any nonempty and proper clopen sets there exists such that .

The second assumption is not hard to satisfy as .

Theorem 5.1.

Assume that satisfies the above assumptions. Then , the commutator subgroup of , is nine-uniformly simple. The commutator width of is at most three. Therefore, if is perfect (i.e., ), then is nine-uniformly simple.

Before proving Theorem 5.1, we need a couple of auxiliary lemmata.

Suppose and . By the Hausdorff property of , if , then there exists a clopen subset containing such that . In such a situation we define an element exchanging and :

Such an element belongs to , since is a topological full group. Observe that and for .

Lemma 5.2.

Assume acts extremely proximally on a Cantor set .

(1)

acts extremely proximally on .

(2)

For any nontrivial and a proper clopen there is such that .

(3)

Let be nontrivial. Then there exists such that is supported outside a clopen subset.

Proof.

(1) Let and be nonempty and proper clopen subsets of . Shrinking , if necessary, we may assume that (that is, we may always take and , such that ; then implies ). By extremal proximality, find elements , , , and in such that , , , and . Define and .

It is straightforward to check that, since , , and are pairwise disjoint, we have which is equivalent to

and similarly for . In particular, and belong to . Furthermore, .

(2) Choose to be a nonempty clopen such that . Choose, by (1), such that . Then .

(3) We may choose clopens and such that . If satisfies , then (such an exists by (2)).

If is the identity on the proof is finished. Otherwise define . We may find such that and . Notice that , , and are pairwise disjoint.

Choose such that . Put , , and . As in (1), we have that and if , then and .

Hence is the identity on . Indeed, let . Then . Thus , i.e., . Therefore .

For any clopen , let be the subgroup of consisting of elements of supported on .

Lemma 5.3.

Let be a proper clopen set. Then there exists a proper clopen such that .

Proof.

Let be such that . Let . Define :

Then is a homeomorphism, which induces an isomorphism given by

for any and . Since is the identity on , for any . Therefore, if , then .

Lemma 5.4.

Assume that are clopens. There exists such that for all , the sets are pairwise disjoint.

Proof.

Choose clopen such that . By extremal proximality, choose and such that and . Define by

Then the sets are pairwise disjoint. Indeed, it is sufficient to prove that for all . Since , we have . As , for , which is disjoint from .

Since conjugates to , the element satisfies the claim.

Proof of Theorem 5.1.

Let be an element of and let be a nontrivial conjugacy class of . By Lemmata 5.2(3) and 5.3 we have that for some and for some proper clopen .

We claim that is a product of four -commutators in . Choose and . We apply Proposition 2.5. Namely, let denote the union of groups such that is a clopen contained in . Clearly, is a proper subgroup of . By Lemma 5.2(2), we may choose such that . Thus, displaces . Let be a finitely generated subgroup of . The union of supports of its generators is a clopen , properly contained in , since . Hence . Choose such that . Let be as in Lemma 5.4. Then -displaces . Thus Proposition 2.5 applies and is a product of four -commutators.

By Lemma 2.2, the commutator width of is at most two. By Lemma 5.2(3), every element decomposes as a product of a conjugate of a given nontrivial element from , say a commutator, and an element conjugate into . Thus every element of is a product of three commutators.

6. Groups almost acting on trees

In this section we apply Theorem 5.1 to groups almost acting on trees.

By a graph (whose elements are called vertices) we mean a set, equipped with a symmetric relation called adjacency. A path is a sequence of vertices indexed either by a set or (in such a case we call the path a ray) such that consecutive vertices are adjacent, and no vertices whose indices differ by two coincide (i.e., there are no backtracks). A graph is called a tree if it is connected (nonempty) and has no cycles, i.e., paths of positive length starting and ending at the same vertex (in particular, the adjacency relation is irreflexive).

Ends of are classes of infinite rays in . Two rays are equivalent if they coincide except for some finite (not necessarily of the same cardinality) subsets. The set of all ends of is denoted by and is called the boundary of .

Given a pair of adjacent vertices (called an oriented edge) , we call the set of terminal vertices of paths starting at a halftree of and we will denote it by . The classes of rays starting at will be called the end of a halftree and will be denoted by . By we denote the pair .

We endow with a topology, where the basis of open sets consists of ends of all halftrees.

A valency of a vertex is the cardinality of the set of vertices adjacent to . A vertex of valency one is called a leaf. If every vertex has valency at least three but finite, then the boundary is easily seen to be compact, totally disconnected, without isolated points, and metrizable. Thus, is a Cantor set. In such a case, every end of a halftree is a clopen (open and closed) subset of .

A spheromorphism is a class of permutations of which preserve all but finitely many adjacency (and nonadjacency) relations. Two such maps are equivalent if they differ on a finite set of vertices (see, e.g., Reference 17, Section 3). We denote the group of all spheromorphisms of by . If is infinite, then the natural map is an embedding. Every spheromorphism induces a homeomorphism of its boundary .

For an integer , by we denote the regular tree whose vertices have degree . The group was introduced by Neretin in Reference 27, 4.5, 3.4 as the group of spheromorphisms of the -regular tree . It is abstractly simple Reference 25.

In what follows, we will be interested in subgroups acting extremely proximally on the boundary (see Theorem 6.4 and Corollary 6.7 below). The whole group of automorphisms of is such an example. Another example (cf. Example 6.8) is the automorphism group of a bi-regular tree , (i.e., every vertex of is black or white, every black vertex is adjacent with white vertices, every white — with black vertices). We prove that the group of partial -actions on is then nine-uniformly simple.

The group itself is virtually 8-uniformly simple Reference 19, Theorem 3.2. (Bounded simplicity in Reference 19 means uniform simplicity in our context.)

There is a connection between the notion of a spheromorphism and a topological full group acting on a boundary of a tree.

Example 6.1.
(1)

Any subdivision of into clopens can be refined to , a subdivision into ends of halftrees (since any clopen in is a finite union of boundaries of halftrees). Therefore the Neretin group can be characterized as .

(2)

Another, well studied, example comes from considering

One may induce cyclic orders by planar representation of . The group is the Higman-Thompson group Reference 17, Section 5, Reference 25, 2.2.

(3)

The previous two examples can be generalized in the following manner (see Reference 8, Section 3.2). Let be a function from the set of (undirected) edges of the -regular tree such that for every vertex , the restriction of to the set of edges starting at gives a bijection with . We say that such is a proper coloring of . Let be a subgroup of permutations of . Using proper coloring and we define the universal group to be

In fact is independent (up to conjugation in ) of the choice of proper coloring . We prove (see Corollary 6.6) that is nine-uniformly simple, provided that is transitive on . If is generated by a -cycle, then from (2). If , then .

We call an action for a group on a tree minimal if there is no proper -invariant subtree of . Given a subset of a tree, we define its convex hull to be the set of all vertices which lie on paths with both ends in the set . It is a subtree. The action is minimal if and only if the convex hull of any orbit is the whole tree.

Example 6.2.

Every action on a leafless tree with a finite quotient is minimal. The converse is not true (see Example 6.8).

Indeed, the distance from a -orbit is a bounded function. Hence the complement of an orbit cannot contain an infinite ray. Thus every vertex lies on a path with endpoints in a given orbit.

Lemma 6.3 (Reference 28, Lemma 4.1).

Assume that a group acts minimally on a leafless tree . Then for every vertex and an edge the orbit intersects the halftree .

Proof.

If is all contained in , so is its convex hull. Thus the claim.

We call an action for a group on a tree parabolic if has a fixed point in .

An action of a group by homeomorphisms on a topological space is called minimal if there is no proper nonempty closed invariant set (equivalently, if every orbit is dense). This notion should not cause confusion with the notion of minimal actions on trees. (A tree is a set equipped with a relation as opposed to its geometric realization which is a topological space.)

Theorem 6.4.

Assume that is a leafless tree such that is a Cantor set. Let act on . The following are equivalent:

(1)

The action of on is extremely proximal (see the beginning of Section 5 for the definitions).

(2)

The action of on is extremely proximal.

(3)

The action of on is minimal and does not support any -invariant probability measure.

(4)

The action of on is minimal and not parabolic, that is, there is no proper -invariant subtree of and has no fixed point in .

Proof.

() This is straightforward.

() Let be a closed, nonempty, proper, and -invariant subset of . Choose and a proper clopen containing . Define . Then there is no such that , since for some ; thus a contradiction.

Similarly, let be a -invariant measure on . Decompose , where the ’s are disjoint nonempty clopens. We may assume that . Then there is no such that . Indeed, for any we may decompose (by compactness) as a finite disjoint union of clopens such that for some and then

is a contradiction. Hence, the action is not extremely proximal.

() If there is an infinite -invariant subtree of or a fixed point , then either or is a -invariant closed subset of .

Suppose that there exists a finite -invariant subtree of . We use the following definition. Given a vertex of , we define the visual measure associated to to be the unique measure on with the following property: if is any injective path starting at , then

where is the valence of . The visual metric is obviously invariant under the action of the stabilizer of in .

We can consider the average of the visual measures associated to the vertices of this subtree . It will be a -invariant measure on .

() By Lemma 6.3 we may assume that, for every pair of edges and , there is such that either or is strictly contained in . It is enough to show that one can find such that the latter holds, i.e., (indeed, since ends of halftrees constitute a basis, we can find edges and such that and for nonempty proper clopens and in ; if there is such that , then ).

It is enough to prove this claim for . Indeed, if there exists such that and , then .

Assume that there exists such that . Let be a path such that and . Then , defined as , is a bi-infinite path. Let be its end as . Choose such that . Consider the bi-infinite path from to . It coincides with and for some . Therefore for big enough. Hence, . Thus the claim.

Remark 6.5.

Only clause (3) from Theorem 6.4 concerns an action of a group on a tree. The other parts of Theorem 6.4 are about actions on a Cantor set. We do not know if there is a straight argument for proving equivalence of and from Theorem 6.4, without referring to actions on trees.

Below is an application of Theorems 5.1 and 6.4 to the Neretin groups and the Higman-Thompson groups.

Corollary 6.6.
(1)

Suppose is a transitive permutation subgroup and let be a proper coloring of (see Example 6.1(3)). Then acts transitively on the directed edges of , and thus is nine-uniformly simple.

(2)

Fix natural numbers . The commutator subgroup of the Neretin group and the Higman-Thompson group , are nine-uniformly simple and have commutator width bounded by three.

Proof.

Let . Then the action of on is not parabolic as there is no -fixed edge adjacent to , hence no -fixed ray. It is minimal since the action is transitive.

Therefore, in the case of the Neretin group and the Higman-Thompson group , Theorem 5.1 applies immediately due to Theorem 6.4.

Suppose is a family of pairwise disjoint ends of halftrees for . If is a pointwise stabilizer of in (see Example 6.1(2)), then is isomorphic to Reference 17, Section 5. Moreover, is its own topological full group acting extremely proximally on . Hence we get the conclusion for .

Corollary 6.7.

Suppose is a free group of rank . Then acts on its Cayley graph, which is . This action is transitive and clearly not parabolic. Thus the induced action on the boundary is extremely proximal. Therefore is nine-uniformly simple by Theorem 5.1.

Example 6.8 (Reference 28, Section 5, Reference 19, p. 232).

We apply our results to trees constructed by Tits. Any connected graph of finite valence, with at least one edge, can appear as a quotient of a (finite valence) tree.

Assume that is a function from oriented edges of into the set of positive integers. By a result of Tits, there is a tree and a group acting on such that and, for any and such that is an edge in , there are exactly vertices in adjacent to (or none if it is not an edge of ).

If is such that the sum over edges starting at a given vertex is at least three (but finite), then the boundary of is a Cantor set.

If values of are at least two, the group action of on is minimal and not parabolic Reference 28, 5.7, i.e., the action of on is extremely proximal due to Theorem 6.4, and is nine-uniformly simple due to Theorem 5.1.

Corollary 6.9.

The groups of quasi-isometries and almost-isometries of a regular tree are five-uniformly simple.

Proof.

This follows from Lazarovich’s results from the appendix. Let be one of those groups. By Theorem 7.4, . Since is a subgroup of , it acts extremely proximally on (see Lemma 7.1) as a topological full group (see Lemma 7.2). This already proves nine-uniform simplicity.

Let and be two elements of . By Lemma 5.2 there exists , a conjugate of , such that fixes a clopen in . By Lemma 7.3, is a commutator of two elements fixing an open set in . Thus, by Lemma 2.1, is a product of two -commutators.

7. Appendix by Nir Lazarovich: Simplicity of and

We begin by recalling the following definitions.

For and , a -quasi-isometry between two metric spaces and is a map such that for all ,

and for all there exists such that .

A -almost-isometry is a -quasi-isometry.

A map is a quasi-isometry (resp., almost-isometry) if there exist and (resp., ) for which it is a -quasi-isometry (resp., -almost-isometry).

Two quasi-isometries are equivalent if they are at bounded distance (with respect to the supremum metric).

The group of all quasi-isometries (resp., almost-isometries) from a metric space to itself, up to equivalence, is denoted by (resp., ). Thus, for , we have the following containments:

where the last containment follows from the following lemma.

Lemma 7.1.

The group acts faithfully on .

Proof.

Let be a quasi-isometry. Let , and let , , be three distinct points such that is the median of , , , that is, is the unique intersection of all three (bi-infinite) geodesics , , . Then, by the stability of quasi-geodesics in Gromov hyperbolic spaces Reference 4, Theorem 1.7, is at bounded distance (which does not depend on the vertex ) from the midpoint of , , . This implies that if induces the identity map at the boundary, then .

In fact, the proof above is valid whenever the space is a proper geodesic Gromov hyperbolic space which has a Gromov boundary of cardinality at least three whose convex hull is at bounded distance from (e.g., any nonelementary hyperbolic group).

For what follows, let be the group or for .

Lemma 7.2.

The group is a topological full group.

Proof.

Fix , and let be a disjoint cover of such that for some . For each let be such that is a disjoint cover of . We may assume, by changing each on a bounded set, that .

Let us define

It is clear that if is in , then it induces the element on the boundary.

Let be the maximal quasi-isometry constants of , and let be the diameter of the bounded set .

We claim the following: for all , . Indeed, if are both in some or in , then the inequality is obvious. If and for some , then and therefore

Similarly, one shows this inequality for and .

Furthermore, the element , defined as

satisfies that for all , for the appropriate , , and . Moreover, it is easy to see that , from which we deduce that is a quasi-isometry.

Lemma 7.3.

Every element in that fixes an open set at the boundary is a commutator of two elements fixing a common set at the boundary.

Proof.

Let . Let be a bi-infinite line geodesic contained in and such that is the starting point of .

Let be a translation along , and let be the function defined by

The function is in since all the functions have the same quasi-isometry constants and .

Let be a 1-almost-isometry defined as

Then we still have as commutes with . However both and fix . Thus the claim.

Theorem 7.4.

The group is perfect and has commutator width at most .

Proof.

It suffices to show that each element of can be written as a product of two elements of which fix an open set at the boundary, as both of them are single commutators by Lemma 7.3.

Let ; there exists such that . Let be a halftree whose boundary contains and for which and are disjoint, and do not cover the whole of . Let be the map defined by

We see that fixes , and thus the claim.

Remark 7.5.

Since, for all , the trees and are quasi-isometric, the groups and are isomorphic.

Acknowledgments

The first author would like to thank Mati Rubin for a fruitful discussion and the Technion — Israel Institute of Technology for hospitality when working on the preliminary version of this paper. The second author would like to thank Hebrew University of Jerusalem for hospitality during the preparation of the paper. The authors gratefully acknowledge the support from the Erwin Schrödinger Institute in Vienna at the final stage of the work, during the meeting “Measured group theory 2016”.

Mathematical Fragments

Theorem 1.1 (Theorem 3.1 below).

Assume that is proximal on a linearly ordered set (i.e., for every and from there exists such that ). Then its commutator group is six-uniformly simple and the commutator width of this group is at most two.

Theorem 1.2.

The commutator subgroup of the Neretin group of spheromorphisms and the commutator subgroup of the Higman-Thomson group are nine-uniformly simple. The commutator width of each of those groups is at most three.

Theorem 1.3.

Assume that a group acts on a leafless tree , whose boundary is a Cantor set, such that does not fix any proper subtree (e.g., is finite) nor a point in the boundary of . Then the commutator subgroup of is nine-uniformly simple.

Proposition 1.4.

Let be a group. Then:

(1)

is simple if and only if any nontrivial central seminorm on is a norm;

(2)

is boundedly simple if and only if every central seminorm on is a bounded norm;

(3)

if is uniformly simple, then every central seminorm on is a bounded and discrete norm;

(4)

is -uniformly simple if and only if .

Lemma 2.1.

Assume that and commute. Then is a product of two -commutators. More precisely can be written as a product of two conjugates of and two conjugates of by elements from the group generated by and .

Lemma 2.2 (Reference 7, Lemma 2.5).

Assume that -displaces . Let be a product of at most commutators . Then there exist , , and such that .

Proposition 2.3.

Assume that is such that for every finitely generated subgroup and , there exists a conjugate of which -displaces . Then every element of is a product of two commutators in and three -commutators in . Moreover,

Lemma 2.4.

Assume that every two elements in commute up to conjugation. Then every commutator in can be expressed as a commutator in . In particular, is perfect.

Proposition 2.5.

Let displace . Assume that, for every , every finitely generated subgroup is -displaceable in . Then every element of is a product of four -commutators from . In particular .

Theorem 3.1.

Assume that acts faithfully, order-preserving, boundedly, and proximally on a linearly ordered set . Then its commutator group is six-uniformly simple and the commutator width of is at most two.

Lemma 3.2.
(1)

The groups and are isomorphic Reference 3, Proposition C10.1.

(2)

The commutator subgroups of and are equal.

Lemma 3.3 (Reference 3, Theorem A4.1, Corollary A5.1).
(1)

is -invariant.

(2)

acts in a doubly-transitive way on . In particular, the action is proximal.

Remark 3.5.

Theorem 3.1 applies to the following groups.

Bieri and Strebel Reference 3 define a more general class of groups acting boundedly on . They take a subgroup in the multiplicative group and a -submodule and define to be a group of boundedly supported automorphisms of consisting of piecewise affine maps with slopes in and singularities in . They define an augmentation ideal of and prove that acts highly transitively on . Thus is six-uniformly simple.

Another example of doubly-transitive and bounded action on a linear order (thus satisfying the assumptions of Theorem 3.1) was considered by Chehata in Reference 12, who studied partially affine transformations of an ordered field and proved that this group is simple. Theorem 3.1 implies that the Chehata group is six-uniformly simple.

Theorem 4.1.

Every proximal action is primitive. Any bounded and primitive action is proximal.

Theorem 4.2.

There exists a subgroup acting transitively and proximally but not doubly-transitively.

Theorem 5.1.

Assume that satisfies the above assumptions. Then , the commutator subgroup of , is nine-uniformly simple. The commutator width of is at most three. Therefore, if is perfect (i.e., ), then is nine-uniformly simple.

Lemma 5.2.

Assume acts extremely proximally on a Cantor set .

(1)

acts extremely proximally on .

(2)

For any nontrivial and a proper clopen there is such that .

(3)

Let be nontrivial. Then there exists such that is supported outside a clopen subset.

Lemma 5.3.

Let be a proper clopen set. Then there exists a proper clopen such that .

Lemma 5.4.

Assume that are clopens. There exists such that for all , the sets are pairwise disjoint.

Example 6.1.
(1)

Any subdivision of into clopens can be refined to , a subdivision into ends of halftrees (since any clopen in is a finite union of boundaries of halftrees). Therefore the Neretin group can be characterized as .

(2)

Another, well studied, example comes from considering

One may induce cyclic orders by planar representation of . The group is the Higman-Thompson group Reference 17, Section 5, Reference 25, 2.2.

(3)

The previous two examples can be generalized in the following manner (see Reference 8, Section 3.2). Let be a function from the set of (undirected) edges of the -regular tree such that for every vertex , the restriction of to the set of edges starting at gives a bijection with . We say that such is a proper coloring of . Let be a subgroup of permutations of . Using proper coloring and we define the universal group to be

In fact is independent (up to conjugation in ) of the choice of proper coloring . We prove (see Corollary 6.6) that is nine-uniformly simple, provided that is transitive on . If is generated by a -cycle, then from (2). If , then .

Lemma 6.3 (Reference 28, Lemma 4.1).

Assume that a group acts minimally on a leafless tree . Then for every vertex and an edge the orbit intersects the halftree .

Theorem 6.4.

Assume that is a leafless tree such that is a Cantor set. Let act on . The following are equivalent:

(1)

The action of on is extremely proximal (see the beginning of Section 5 for the definitions).

(2)

The action of on is extremely proximal.

(3)

The action of on is minimal and does not support any -invariant probability measure.

(4)

The action of on is minimal and not parabolic, that is, there is no proper -invariant subtree of and has no fixed point in .

Corollary 6.6.
(1)

Suppose is a transitive permutation subgroup and let be a proper coloring of (see Example 6.1(3)). Then acts transitively on the directed edges of , and thus is nine-uniformly simple.

(2)

Fix natural numbers . The commutator subgroup of the Neretin group and the Higman-Thompson group , are nine-uniformly simple and have commutator width bounded by three.

Corollary 6.7.

Suppose is a free group of rank . Then acts on its Cayley graph, which is . This action is transitive and clearly not parabolic. Thus the induced action on the boundary is extremely proximal. Therefore is nine-uniformly simple by Theorem 5.1.

Example 6.8 (Reference 28, Section 5, Reference 19, p. 232).

We apply our results to trees constructed by Tits. Any connected graph of finite valence, with at least one edge, can appear as a quotient of a (finite valence) tree.

Assume that is a function from oriented edges of into the set of positive integers. By a result of Tits, there is a tree and a group acting on such that and, for any and such that is an edge in , there are exactly vertices in adjacent to (or none if it is not an edge of ).

If is such that the sum over edges starting at a given vertex is at least three (but finite), then the boundary of is a Cantor set.

If values of are at least two, the group action of on is minimal and not parabolic Reference 28, 5.7, i.e., the action of on is extremely proximal due to Theorem 6.4, and is nine-uniformly simple due to Theorem 5.1.

Lemma 7.1.

The group acts faithfully on .

Lemma 7.2.

The group is a topological full group.

Lemma 7.3.

Every element in that fixes an open set at the boundary is a commutator of two elements fixing a common set at the boundary.

Theorem 7.4.

The group is perfect and has commutator width at most .

References

Reference [1]
R. D. Anderson, On homeomorphisms as products of conjugates of a given homeomorphism and its inverse, Topology of 3-manifolds and related topics (Proc. The Univ. of Georgia Institute, 1961), Prentice-Hall, Englewood Cliffs, N.J., 1962, pp. 231–234. MR0139684,
Show rawAMSref \bib{MR0139684}{article}{ author={Anderson, R. D.}, title={On homeomorphisms as products of conjugates of a given homeomorphism and its inverse}, conference={ title={Topology of 3-manifolds and related topics}, address={Proc. The Univ. of Georgia Institute}, date={1961}, }, book={ publisher={Prentice-Hall, Englewood Cliffs, N.J.}, }, date={1962}, pages={231--234}, review={\MR {0139684}}, }
Reference [2]
Valery Bardakov, Vladimir Tolstykh, and Vladimir Vershinin, Generating groups by conjugation-invariant sets, J. Algebra Appl. 11 (2012), no. 4, 1250071, 16, DOI 10.1142/S0219498812500715. MR2959420,
Show rawAMSref \bib{MR2959420}{article}{ author={Bardakov, Valery}, author={Tolstykh, Vladimir}, author={Vershinin, Vladimir}, title={Generating groups by conjugation-invariant sets}, journal={J. Algebra Appl.}, volume={11}, date={2012}, number={4}, pages={1250071, 16}, issn={0219-4988}, review={\MR {2959420}}, doi={10.1142/S0219498812500715}, }
Reference [3]
Robert Bieri and Ralph Strebel, On groups of PL-homeomorphisms of the real line, Mathematical Surveys and Monographs, vol. 215, American Mathematical Society, Providence, RI, 2016. MR3560537,
Show rawAMSref \bib{MR3560537}{book}{ author={Bieri, Robert}, author={Strebel, Ralph}, title={On groups of PL-homeomorphisms of the real line}, series={Mathematical Surveys and Monographs}, volume={215}, publisher={American Mathematical Society, Providence, RI}, date={2016}, pages={xvii+174}, isbn={978-1-4704-2901-0}, review={\MR {3560537}}, }
Reference [4]
Martin R. Bridson and André Haefliger, Metric spaces of non-positive curvature, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 319, Springer-Verlag, Berlin, 1999, DOI 10.1007/978-3-662-12494-9. MR1744486,
Show rawAMSref \bib{MR1744486}{book}{ author={Bridson, Martin R.}, author={Haefliger, Andr\'e}, title={Metric spaces of non-positive curvature}, series={Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]}, volume={319}, publisher={Springer-Verlag, Berlin}, date={1999}, pages={xxii+643}, isbn={3-540-64324-9}, review={\MR {1744486}}, doi={10.1007/978-3-662-12494-9}, }
Reference [5]
Kenneth S. Brown, Finiteness properties of groups, Proceedings of the Northwestern conference on cohomology of groups (Evanston, Ill., 1985), J. Pure Appl. Algebra 44 (1987), no. 1-3, 45–75, DOI 10.1016/0022-4049(87)90015-6. MR885095,
Show rawAMSref \bib{MR885095}{article}{ author={Brown, Kenneth S.}, title={Finiteness properties of groups}, booktitle={Proceedings of the Northwestern conference on cohomology of groups (Evanston, Ill., 1985)}, journal={J. Pure Appl. Algebra}, volume={44}, date={1987}, number={1-3}, pages={45--75}, issn={0022-4049}, review={\MR {885095}}, doi={10.1016/0022-4049(87)90015-6}, }
Reference [6]
Dmitri Burago and Sergei Ivanov, A remark on the group of PL-homeomorphisms in dimension one, Geometric and probabilistic structures in dynamics, Contemp. Math., vol. 469, Amer. Math. Soc., Providence, RI, 2008, pp. 141–148, DOI 10.1090/conm/469/09164. MR2478468,
Show rawAMSref \bib{MR2478468}{article}{ author={Burago, Dmitri}, author={Ivanov, Sergei}, title={A remark on the group of PL-homeomorphisms in dimension one}, conference={ title={Geometric and probabilistic structures in dynamics}, }, book={ series={Contemp. Math.}, volume={469}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2008}, pages={141--148}, review={\MR {2478468}}, doi={10.1090/conm/469/09164}, }
Reference [7]
Dmitri Burago, Sergei Ivanov, and Leonid Polterovich, Conjugation-invariant norms on groups of geometric origin, Groups of diffeomorphisms, Adv. Stud. Pure Math., vol. 52, Math. Soc. Japan, Tokyo, 2008, pp. 221–250. MR2509711,
Show rawAMSref \bib{MR2509711}{article}{ author={Burago, Dmitri}, author={Ivanov, Sergei}, author={Polterovich, Leonid}, title={Conjugation-invariant norms on groups of geometric origin}, conference={ title={Groups of diffeomorphisms}, }, book={ series={Adv. Stud. Pure Math.}, volume={52}, publisher={Math. Soc. Japan, Tokyo}, }, date={2008}, pages={221--250}, review={\MR {2509711}}, }
Reference [8]
Marc Burger and Shahar Mozes, Groups acting on trees: from local to global structure, Inst. Hautes Études Sci. Publ. Math. 92 (2000), 113–150 (2001). MR1839488,
Show rawAMSref \bib{MR1839488}{article}{ author={Burger, Marc}, author={Mozes, Shahar}, title={Groups acting on trees: from local to global structure}, journal={Inst. Hautes \'Etudes Sci. Publ. Math.}, number={92}, date={2000}, pages={113--150 (2001)}, issn={0073-8301}, review={\MR {1839488}}, }
Reference [9]
Danny Calegari, Stable commutator length in subgroups of , Pacific J. Math. 232 (2007), no. 2, 257–262, DOI 10.2140/pjm.2007.232.257. MR2366352,
Show rawAMSref \bib{MR2366352}{article}{ author={Calegari, Danny}, title={Stable commutator length in subgroups of ${\rm PL}^+(I)$}, journal={Pacific J. Math.}, volume={232}, date={2007}, number={2}, pages={257--262}, issn={0030-8730}, review={\MR {2366352}}, doi={10.2140/pjm.2007.232.257}, }
Reference [10]
Pierre-Emmanuel Caprace, Automorphism groups of right-angled buildings: simplicity and local splittings, Fund. Math. 224 (2014), no. 1, 17–51, DOI 10.4064/fm224-1-2. MR3164745,
Show rawAMSref \bib{MR3164745}{article}{ author={Caprace, Pierre-Emmanuel}, title={Automorphism groups of right-angled buildings: simplicity and local splittings}, journal={Fund. Math.}, volume={224}, date={2014}, number={1}, pages={17--51}, issn={0016-2736}, review={\MR {3164745}}, doi={10.4064/fm224-1-2}, }
Reference [11]
Pierre-Emmanuel Caprace and Koji Fujiwara, Rank-one isometries of buildings and quasi-morphisms of Kac-Moody groups, Geom. Funct. Anal. 19 (2010), no. 5, 1296–1319, DOI 10.1007/s00039-009-0042-2. MR2585575,
Show rawAMSref \bib{MR2585575}{article}{ author={Caprace, Pierre-Emmanuel}, author={Fujiwara, Koji}, title={Rank-one isometries of buildings and quasi-morphisms of Kac-Moody groups}, journal={Geom. Funct. Anal.}, volume={19}, date={2010}, number={5}, pages={1296--1319}, issn={1016-443X}, review={\MR {2585575}}, doi={10.1007/s00039-009-0042-2}, }
Reference [12]
C. G. Chehata, An algebraically simple ordered group, Proc. London Math. Soc. (3) 2 (1952), 183–197, DOI 10.1112/plms/s3-2.1.183. MR0047031,
Show rawAMSref \bib{MR0047031}{article}{ author={Chehata, C. G.}, title={An algebraically simple ordered group}, journal={Proc. London Math. Soc. (3)}, volume={2}, date={1952}, pages={183--197}, issn={0024-6115}, review={\MR {0047031}}, doi={10.1112/plms/s3-2.1.183}, }
Reference [13]
Manfred Droste and R. M. Shortt, Commutators in groups of order-preserving permutations, Glasgow Math. J. 33 (1991), no. 1, 55–59, DOI 10.1017/S001708950000803X. MR1089954,
Show rawAMSref \bib{MR1089954}{article}{ author={Droste, Manfred}, author={Shortt, R. M.}, title={Commutators in groups of order-preserving permutations}, journal={Glasgow Math. J.}, volume={33}, date={1991}, number={1}, pages={55--59}, issn={0017-0895}, review={\MR {1089954}}, doi={10.1017/S001708950000803X}, }
Reference [14]
Gábor Elek and Endre Szabó, Hyperlinearity, essentially free actions and -invariants. The sofic property, Math. Ann. 332 (2005), no. 2, 421–441, DOI 10.1007/s00208-005-0640-8. MR2178069,
Show rawAMSref \bib{MR2178069}{article}{ author={Elek, G\'abor}, author={Szab\'o, Endre}, title={Hyperlinearity, essentially free actions and $L^2$-invariants. The sofic property}, journal={Math. Ann.}, volume={332}, date={2005}, number={2}, pages={421--441}, issn={0025-5831}, review={\MR {2178069}}, doi={10.1007/s00208-005-0640-8}, }
Reference [15]
Erich W. Ellers, Nikolai Gordeev, and Marcel Herzog, Covering numbers for Chevalley groups, Israel J. Math. 111 (1999), 339–372, DOI 10.1007/BF02810691. MR1710745,
Show rawAMSref \bib{MR1710745}{article}{ author={Ellers, Erich W.}, author={Gordeev, Nikolai}, author={Herzog, Marcel}, title={Covering numbers for Chevalley groups}, journal={Israel J. Math.}, volume={111}, date={1999}, pages={339--372}, issn={0021-2172}, review={\MR {1710745}}, doi={10.1007/BF02810691}, }
Reference [16]
Światosław R. Gal and Jarek Kedra, On bi-invariant word metrics, J. Topol. Anal. 3 (2011), no. 2, 161–175, DOI 10.1142/S1793525311000556. MR2819193,
Show rawAMSref \bib{MR2819193}{article}{ author={Gal, \'Swiatos\l aw R.}, author={Kedra, Jarek}, title={On bi-invariant word metrics}, journal={J. Topol. Anal.}, volume={3}, date={2011}, number={2}, pages={161--175}, issn={1793-5253}, review={\MR {2819193}}, doi={10.1142/S1793525311000556}, }
Reference [17]
L. Garncarek and N. Lazarovich, The Neretin groups, arXiv:1502.00991v2, 2015.
Reference [18]
Thierry Giordano, Ian F. Putnam, and Christian F. Skau, Full groups of Cantor minimal systems, Israel J. Math. 111 (1999), 285–320, DOI 10.1007/BF02810689. MR1710743,
Show rawAMSref \bib{MR1710743}{article}{ author={Giordano, Thierry}, author={Putnam, Ian F.}, author={Skau, Christian F.}, title={Full groups of Cantor minimal systems}, journal={Israel J. Math.}, volume={111}, date={1999}, pages={285--320}, issn={0021-2172}, review={\MR {1710743}}, doi={10.1007/BF02810689}, }
Reference [19]
Jakub Gismatullin, Boundedly simple groups of automorphisms of trees, J. Algebra 392 (2013), 226–243, DOI 10.1016/j.jalgebra.2013.06.023. MR3085032,
Show rawAMSref \bib{MR3085032}{article}{ author={Gismatullin, Jakub}, title={Boundedly simple groups of automorphisms of trees}, journal={J. Algebra}, volume={392}, date={2013}, pages={226--243}, issn={0021-8693}, review={\MR {3085032}}, doi={10.1016/j.jalgebra.2013.06.023}, }
Reference [20]
Shmuel Glasner, Topological dynamics and group theory, Trans. Amer. Math. Soc. 187 (1974), 327–334, DOI 10.2307/1997056. MR0336723,
Show rawAMSref \bib{MR0336723}{article}{ author={Glasner, Shmuel}, title={Topological dynamics and group theory}, journal={Trans. Amer. Math. Soc.}, volume={187}, date={1974}, pages={327--334}, issn={0002-9947}, review={\MR {0336723}}, doi={10.2307/1997056}, }
Reference [21]
Shmuel Glasner, Proximal flows, Lecture Notes in Mathematics, Vol. 517, Springer-Verlag, Berlin-New York, 1976. MR0474243,
Show rawAMSref \bib{MR0474243}{book}{ author={Glasner, Shmuel}, title={Proximal flows}, series={Lecture Notes in Mathematics, Vol. 517}, publisher={Springer-Verlag, Berlin-New York}, date={1976}, pages={viii+153}, review={\MR {0474243}}, }
Reference [22]
Nikolai Gordeev and Jan Saxl, Products of conjugacy classes in Chevalley groups. I. Extended covering numbers, Israel J. Math. 130 (2002), 207–248, DOI 10.1007/BF02764077. MR1919378,
Show rawAMSref \bib{MR1919378}{article}{ author={Gordeev, Nikolai}, author={Saxl, Jan}, title={Products of conjugacy classes in Chevalley groups. I. Extended covering numbers}, journal={Israel J. Math.}, volume={130}, date={2002}, pages={207--248}, issn={0021-2172}, review={\MR {1919378}}, doi={10.1007/BF02764077}, }
Reference [23]
N. L. Gordeev, Products of conjugacy classes in algebraic groups. I, II, J. Algebra 173 (1995), no. 3, 715–744, 745–779, DOI 10.1006/jabr.1995.1111. MR1327877,
Show rawAMSref \bib{MR1327877}{article}{ author={Gordeev, N. L.}, title={Products of conjugacy classes in algebraic groups. I, II}, journal={J. Algebra}, volume={173}, date={1995}, number={3}, pages={715--744, 745--779}, issn={0021-8693}, review={\MR {1327877}}, doi={10.1006/jabr.1995.1111}, }
Reference [24]
Charles Holland, Transitive lattice-ordered permutation groups, Math. Z. 87 (1965), 420–433, DOI 10.1007/BF01111722. MR0178052,
Show rawAMSref \bib{MR0178052}{article}{ author={Holland, Charles}, title={Transitive lattice-ordered permutation groups}, journal={Math. Z.}, volume={87}, date={1965}, pages={420--433}, issn={0025-5874}, review={\MR {0178052}}, doi={10.1007/BF01111722}, }
Reference [25]
Christophe Kapoudjian, Simplicity of Neretin’s group of spheromorphisms, Ann. Inst. Fourier (Grenoble) 49 (1999), no. 4, 1225–1240. MR1703086,
Show rawAMSref \bib{MR1703086}{article}{ author={Kapoudjian, Christophe}, title={Simplicity of Neretin's group of spheromorphisms}, journal={Ann. Inst. Fourier (Grenoble)}, volume={49}, date={1999}, number={4}, pages={1225--1240}, issn={0373-0956}, review={\MR {1703086}}, }
Reference [26]
Stephen H. McCleary, Lattice-ordered permutation groups: the structure theory, Ordered groups and infinite permutation groups, Math. Appl., vol. 354, Kluwer Acad. Publ., Dordrecht, 1996, pp. 29–62. MR1486196,
Show rawAMSref \bib{MR1486196}{article}{ author={McCleary, Stephen H.}, title={Lattice-ordered permutation groups: the structure theory}, conference={ title={Ordered groups and infinite permutation groups}, }, book={ series={Math. Appl.}, volume={354}, publisher={Kluwer Acad. Publ., Dordrecht}, }, date={1996}, pages={29--62}, review={\MR {1486196}}, }
Reference [27]
Yu. A. Neretin, Combinatorial analogues of the group of diffeomorphisms of the circle, Izv. Ross. Akad. Nauk Ser. Mat. 56 (1992), no. 5, 1072–1085, DOI 10.1070/IM1993v041n02ABEH002264; English transl., Russian Acad. Sci. Izv. Math. 41 (1993), no. 2, 337–349. MR1209033,
Show rawAMSref \bib{MR1209033}{article}{ author={Neretin, Yu. A.}, title={Combinatorial analogues of the group of diffeomorphisms of the circle}, journal={Izv. Ross. Akad. Nauk Ser. Mat.}, volume={56}, date={1992}, number={5}, pages={1072--1085}, issn={1607-0046}, translation={ journal={Russian Acad. Sci. Izv. Math.}, volume={41}, date={1993}, number={2}, pages={337--349}, issn={1064-5632}, }, review={\MR {1209033}}, doi={10.1070/IM1993v041n02ABEH002264}, }
Reference [28]
Jacques Tits, Sur le groupe des automorphismes d’un arbre, Essays on topology and related topics (Mémoires dédiés à Georges de Rham), Springer, New York, 1970, pp. 188–211. MR0299534,
Show rawAMSref \bib{MR0299534}{article}{ author={Tits, Jacques}, title={Sur le groupe des automorphismes d'un arbre}, conference={ title={Essays on topology and related topics (M\'emoires d\'edi\'es \`a Georges de Rham)}, }, book={ publisher={Springer, New York}, }, date={1970}, pages={188--211}, review={\MR {0299534}}, }

Article Information

MSC 2010
Primary: 20E08 (Groups acting on trees), 20E32 (Simple groups)
Secondary: 20F65 (Geometric group theory), 22E40 (Discrete subgroups of Lie groups)
Keywords
  • Boundedly simple groups
  • trees
  • automorphism groups
  • spheromorphisms
  • almost automorphisms
  • Higman-Thomson groups
  • Neretin group
Author Information
Światosław R. Gal
Instytut Matematyczny, Uniwersytetu Wrocławskiego, pl. Grunwaldzki 2/4, 50-384 Wrocław, Poland – and – Weizmann Institute of Science, Rehovot 76100, Israel
sgal@math.uni.wroc.pl
Jakub Gismatullin
Instytut Matematyczny, Uniwersytetu Wrocławskiego, pl. Grunwaldzki 2/4, 50-384 Wrocław, Poland – and – Instytut Matematyczny, Polskiej Akademii Nauk, ul. Śniadeckich 8, 00-656 Warszawa, Poland
gismat@math.uni.wroc.pl
Homepage
MathSciNet
Contributor Information
Nir Lazarovich
Departement Mathematik, Eidgenössische Technische Hochschule Zürich, Rämistrasse 101, 8092 Zürich, Switzerland
nir.lazarovich@math.ethz.ch
MathSciNet
Additional Notes

The research leading to these results has received funding from the European Research Council under the European Unions Seventh Framework Programme (FP7/2007-2013)/ERC Grant Agreement No. 291111. The first author partially supported by Polish National Science Center (NCN) grant 2012/06/A/ST1/00259 and the European Research Council grant No. 306706.

The second author is partially supported by the NCN grants 2014/13/D/ST1/03491, 2012/07/B/ST1/03513.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 4, Issue 5, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2017 by the authors under Creative Commons Attribution-Noncommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/18
  • MathSciNet Review: 3693109
  • Show rawAMSref \bib{3693109}{article}{ author={Gal, \'Swiatos\l aw}, author={Gismatullin, Jakub}, title={Uniform simplicity of groups with proximal action}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={4}, number={5}, date={2017}, pages={110-130}, issn={2330-0000}, review={3693109}, doi={10.1090/btran/18}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

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

For more information please visit the AMS MathViewer documentation.