Hostname: page-component-cd9895bd7-lnqnp Total loading time: 0 Render date: 2024-12-26T00:30:17.709Z Has data issue: false hasContentIssue false

A dichotomy for topological full groups

Published online by Cambridge University Press:  15 September 2022

Eduardo Scarparo*
Affiliation:
School of Mathematics and Statistics, University of Glasgow, Glasgow, United Kingdom
Rights & Permissions [Opens in a new window]

Abstract

Given a minimal action $\alpha $ of a countable group on the Cantor set, we show that the alternating full group $\mathsf {A}(\alpha )$ is non-amenable if and only if the topological full group $\mathsf {F}(\alpha )$ is $C^*$ -simple. This implies, for instance, that the Elek–Monod example of non-amenable topological full group coming from a Cantor minimal $\mathbb {Z}^2$ -system is $C^*$ -simple.

Type
Article
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of The Canadian Mathematical Society

1 Introduction

Given an action $\alpha $ of a group on the Cantor set X, the topological full group of $\alpha $ , denoted by $\mathsf {F}(\alpha )$ , is the group of homeomorphisms on X which are locally given by $\alpha $ .

In [Reference Juschenko and MonodJM13], Juschenko and Monod showed that topological full groups of Cantor minimal $\mathbb {Z}$ -systems are amenable. Together with results of Matui [Reference Matte BonMat06], this gave rise to the first examples of infinite, simple, finitely generated, amenable groups. On the other hand, in [Reference Elek and MonodEM13], Elek and Monod constructed an example of a free minimal $\mathbb {Z}^2$ -subshift whose topological full group contains a free group.

A group $\Gamma $ is said to have the unique trace property if its reduced $C^*$ -algebra $C^*_r(\Gamma )$ has a unique tracial state and to be $C^*$ -simple if $C^*_r(\Gamma )$ is simple. In [Reference Breuillard, Kalantar, Kennedy and OzawaBKKO17], Breuillard et al. showed that $\Gamma $ has the unique trace property if and only if it does not contain any non-trivial amenable normal subgroup, and in [Reference KennedyKen20], Kennedy showed that $\Gamma $ is $C^*$ -simple if and only if it does not contain any nontrivial amenable uniformly recurrent subgroup (URS).

By using this new characterization of $C^*$ -simplicity, Le Boudec and Matte Bon showed in [Reference Le Boudec and Matte BonLBMB18] that the topological full group of a free minimal action of a countable non-amenable group on the Cantor set is $C^*$ -simple, and asked whether the same conclusion holds if one does not assume freeness. In [Reference Brix and ScarparoBS19], Brix and the author showed that it suffices to assume that the action is topologically free. In [Reference Kerr and Tucker-DrobKTD19], Kerr and Tucker-Drob obtained examples of $C^*$ -simple topological full groups coming from actions of amenable groups.

Given an action $\alpha $ of a group on the Cantor set, Nekrashevych introduced in [Reference NekrashevychNek19] the alternating full group of the action, which we denote by $\mathsf {A}(\alpha )$ . This is a normal subgroup of $\mathsf {F}(\alpha )$ generated by certain copies of finite alternating groups. It was shown in [Reference NekrashevychNek19] that if $\alpha $ is minimal, then $\mathsf {A}(\alpha )$ is simple and is contained in every nontrivial normal subgroup of $\mathsf {F}(\alpha )$ .

In [Reference MatuiMB18], Matte Bon obtained a classification of URSs of topological full groups. By using this result, we show the following theorem.

Theorem (Theorem 3.5)

Let $\alpha $ be a minimal action of a countable group on the Cantor set. The following conditions are equivalent:

  1. (i) $\mathsf {A}(\alpha )$ is non-amenable.

  2. (ii) Any group H such that $\mathsf {A}(\alpha )\leq H\leq \mathsf {F}(\alpha )$ is $C^*$ -simple.

  3. (iii) There exists a $C^*$ -simple group H such that $\mathsf {A}(\alpha )\leq H\leq \mathsf {F}(\alpha )$ .

As a consequence, we obtain the following corollary.

Corollary (Corollary 3.6)

Let $\alpha $ be a minimal action of a countable group on the Cantor set. Then $\mathsf {F}(\alpha )$ has the unique trace property if and only if it is $C^*$ -simple. If $\mathsf {A}(\alpha )=\mathsf {F}(\alpha )'$ , then $\mathsf {F}(\alpha )$ is non-amenable if and only if it is $C^*$ -simple.

It is still an open problem whether $\mathsf {A}(\alpha )$ always coincides with $\mathsf {F}(\alpha )'$ , but in many cases, this is known to be true. For example, it follows from results of Matui [Reference MatuiMat15] that this is the case for free Cantor minimal $\mathbb {Z}^n$ -systems. This implies that the example of non-amenable topological full group coming from an action of $\mathbb {Z}^2$ in [Reference Elek and MonodEM13] is $C^*$ -simple.

2 Preliminaries

2.1 Topological dynamics

Given a locally compact Hausdorff space X, we denote by $\mathcal {B}(X)$ the Borel $\sigma $ -algebra of X, and by $\mathcal {P}(X)$ the space of regular probability measures on X.

If $\Gamma $ is a group acting by homeomorphisms on X, we say that X is a locally compact $\Gamma $ -space. If X admits no nontrivial $\Gamma $ -invariant closed subspaces, then we say that X (or the action) is minimal. Given $U\subset X$ , let $\operatorname {St}_{\Gamma} (U)$ consist of the elements of $\Gamma $ which fix pointwise U, and $\operatorname {St}_{\Gamma} (U)^0$ consist of the elements of $\Gamma $ which fix pointwise a neighborhood of U. To ease the notation, given $x\in X$ , we let $\Gamma _x:=\operatorname {St}_{\Gamma} (\{x\})$ and $\Gamma _x^0:=\operatorname {St}_{\Gamma} (\{x\})^0$ .

Denote by $\operatorname {Sub}(\Gamma )$ the set of subgroups of $\Gamma $ , endowed with the Chabauty topology; this is the restriction to $\operatorname {Sub}(\Gamma )$ of the product topology on $\{0, 1\}^{\Gamma} $ , where every subgroup $\Lambda \in \operatorname {Sub}(G)$ is identified with its characteristic function $\boldsymbol {1}_\Lambda \in \{0, 1\}^{\Gamma} $ . Notice that the space of amenable subgroups $\operatorname {Sub}_{am}(\Gamma )$ is closed in $\operatorname {Sub}(\Gamma )$ . We consider $\operatorname {Sub}(\Gamma )$ as a compact $\Gamma $ -space under the action by conjugation. A subgroup $\Lambda \leq \Gamma $ is said to be confined if $\{e\}$ is not in the closure of the $\Gamma $ -orbit of $\Lambda $ .

An invariant random subgroup (IRS) is a $\Gamma $ -invariant regular probability measure on $\operatorname {Sub}(\Gamma )$ . We say an IRS is amenable if its support is contained in $\operatorname {Sub}_{am}(\Gamma )$ . By [Reference Breuillard, Kalantar, Kennedy and OzawaBKKO17, Corollary 4.3] and [Reference Bader, Duchesne and LécureuxBDL16, Corollary 1.5], $\Gamma $ has the unique trace property if and only if its unique amenable normal subgroup is $\{e\}$ , if and only if its unique amenable IRS is $\delta _{\{e\}}$ .

A URS is a $\Gamma $ -invariant closed minimal subspace $\mathcal {U}\subset \operatorname {Sub}(\Gamma )$ . We say $\mathcal {U}$ is amenable if every element of $\mathcal {U}$ is amenable. By [Reference KennedyKen20, Theorem 4.1], $\Gamma $ is $C^*$ -simple if and only if its only amenable URS is $\{\{e\}\}$ . Alternatively, $\Gamma $ is $C^*$ -simple if and only if it does not contain any confined amenable subgroup.

Suppose $\Gamma $ is countable and X is a minimal compact $\Gamma $ -space. Let $\operatorname {Stab}_{\Gamma} \colon X\to \operatorname {Sub}(\Gamma )$ be the map given by $\operatorname {Stab}_{\Gamma} (x):=\Gamma _x$ and $\operatorname {Stab}_{\Gamma} ^0\colon X\to \operatorname {Sub}(\Gamma )$ be the map given by $\operatorname {Stab}_{\Gamma} ^0(x):=\Gamma _x^0$ , for $x\in X$ . Notice that $\operatorname {Stab}_{\Gamma} $ and $\operatorname {Stab}_{\Gamma} ^0$ are Borel measurable and $\Gamma $ -equivariant. Moreover, the set $Y:=\{x\in X:\Gamma _x=\Gamma _x^0\}$ is dense in X and $\operatorname {Stab}_{\Gamma} (Y)$ is a URS, the so-called stabilizer URS of the action $\Gamma \!\curvearrowright \! X$ (for a proof of these last claims, see [Reference Le Boudec and Matte BonLBMB18, Section 2]).

2.2 Topological full groups

Fix an action $\alpha $ of a group $\Gamma $ on the Cantor set X. We say that a homeomorphism $h\colon U\to V$ between clopen subsets $U,V\subset X$ is locally given by $\alpha $ if there exist $g_1,\dots ,g_n\in \Gamma $ and clopen sets $A_1,\dots ,A_n\subset U$ such that $U=\bigsqcup _{i=1}^n A_i$ and $h|_{A_i}=g_i|_{A_i}$ , for $1\leq i \leq n$ . The topological full group of $\alpha $ , denoted by $\mathsf {F}(\alpha )$ , is the group of homeomorphisms $h\colon X\to X$ which are locally given by $\alpha $ .

Given $d\in \mathbb {N}$ , a d-multisection is a collection of d disjoint clopen sets $(A_i)_{i=1}^d\subset X$ and $d^2$ homeomorphisms $(h_{i,j}\colon A_i\to A_j)_{i,j=1}^d$ which are locally given by $\alpha $ and such that, for $1\leq i,j,k\leq d$ , it holds that $h_{j,k}h_{i,j}=h_{i,k}$ and $h_{i,i}=\mathrm {Id}_{A_i}$ .

Given $d\in \mathbb {N}$ , let $S_d$ and $A_d$ be the symmetric and alternating groups, respectively. Given a d-multisection $\mathcal {F}=((A_i)_{i=1}^d,(h_{i,j})_{i,j=1}^d)$ and $\sigma \in S_d$ , let $\mathcal {F}(\sigma )\in \mathsf {F}(\alpha )$ be given by $\mathcal {F}(\sigma )|_{A_i}:=h_{\sigma (i),i}$ , for $1\leq i \leq n$ and $\mathcal {F}(\sigma )(x)=x$ for $x\notin \bigsqcup _{i=1}^n A_i$ . The alternating full group $\mathsf {A}(\alpha )$ is the subgroup of $\mathsf {F}(\alpha )$ generated by

$$ \begin{align*}\{\mathcal{F}(\sigma):d\in\mathbb{N}, \mathcal{F} \text{ is a }d\text{-multisection, } \sigma\in A_d\}.\end{align*} $$

Notice that $\mathsf {A}(\alpha )$ is normal in $\mathsf {F}(\alpha )$ and that $\mathsf {A}(\alpha )$ is contained in the derived subgroup $\mathsf {F}(\alpha )'$ . If $\alpha $ is a minimal action of a countable group on the Cantor set, then $\mathsf {A}(\alpha )$ is simple [Reference NekrashevychNek19, Theorem 4.1].

Remark 2.1 Alternatively, $\mathsf {F}(\alpha )$ and $\mathsf {A}(\alpha )$ can be described as groups of bisections of the groupoid of germs of $\alpha $ . This is the point of view adopted in [Reference MatuiMB18, Reference NekrashevychNek19]. Conversely, given an effective groupoid G with unit space $G^{(0)}$ homeomorphic to the Cantor set, denote by $\alpha $ the natural action of the topological full group of G on $G^{(0)}$ . Then the topological and alternating full groups of G coincide with $\mathsf {F}(\alpha )$ and $\mathsf {A}(\alpha )$ , respectively [Reference Nyland and OrtegaNO19, Corollary 4.7].

3 $\boldsymbol{C}^{\boldsymbol{*}}$ -simplicity of full groups

Given a locally compact $\Gamma $ -space X and $U\subset X$ open not necessarily invariant, let

$$ \begin{align*}\mathcal{P}_{\Gamma}(U):=\{\mu\in\mathcal{P}(U):\forall g\in\Gamma\ \forall A\in\mathcal{B}(U),\ \ \mu(A\cap g^{-1}U)=\mu(gA\cap U)\}. \end{align*} $$

Alternatively, one can characterize $\mathcal {P}_{\Gamma} (U)$ as the measures $\mu \in \mathcal {P}(U)$ such that $\mu (gA)=\mu (A)$ for every $g\in \Gamma $ and $A\in \mathcal {B}(U)$ such that $gA\subset U$ .

Proposition 3.1 Let X be a compact $\Gamma $ -space and $U\subset X$ open such that $X=\Gamma. U$ . Then the map $j\colon \mathcal {P}_{\Gamma} (X)\to \mathcal {P}_{\Gamma} (U)$ given by $j(\nu ):=\frac {\nu |_{\mathcal {B}(U)} }{\nu (U)}$ is a well-defined bijection.

Proof Take $g_1,\dots ,g_n\in \Gamma $ such that $X=\bigcup _{i=1}^n g_iU$ . For $1\leq i \leq n$ , let

$$ \begin{align*}A_i:=U\setminus\bigcup_{j=1}^{i-1}g_i^{-1}g_j U.\end{align*} $$

Then $X=\bigsqcup _{i=1}^n g_iA_i$ . Given $\nu \in \mathcal {P}_{\Gamma} (X)$ , obviously $\nu (U)\geq 1/n$ , so that j is a well-defined map. Moreover, given $A\in \mathcal {B}(X)$ , we have

(1) $$ \begin{align} \nu(A)=\sum_{i=1}^n\nu(g_i A_i\cap A)=\sum_{i=1}^n\nu(A_i\cap g_i^{-1} A). \end{align} $$

Since each $A_i$ is contained in U, this implies that $\nu $ is determined by its restriction to $\mathcal {B}(U)$ .

If $\nu _1,\nu _2\in \mathcal {P}_{\Gamma} (X)$ are such that $j(\nu _1)=j(\nu _2)$ , then $\nu _1|_U=\frac {\nu _1(U)}{\nu _2(U)}\nu _2|_U$ . Furthermore, by (1), we have

$$ \begin{align*}1=\nu_1(X)=\sum_{i=1}^n\nu_1(A_i)=\sum_{i=1}^n\frac{\nu_1(U)}{\nu_2(U)}\nu_2(A_i)=\frac{\nu_1(U)}{\nu_2(U)}\nu_2(X)=\frac{\nu_1(U)}{\nu_2(U)},\end{align*} $$

and hence $\nu _1(U)=\nu _2(U)$ . Consequently, $\nu _1=\nu _2$ and j is injective.

Let us now show that j is surjective. Given $\mu \in \mathcal {P}_{\Gamma} (U)$ and $A\in \mathcal {B}(X)$ , let

$$ \begin{align*}\nu(A):=\sum_{i=1}^n\mu(A_i\cap g_i^{-1}A).\end{align*} $$

Given $B\in \mathcal {B}(U)$ , we have

$$ \begin{align*}\nu(B)=\sum_{i=1}^n\mu(A_i\cap g_i^{-1}B)=\sum_{i=1}^n\mu(g_iA_i\cap B)=\mu(B),\end{align*} $$

so that $\nu |_{\mathcal {B}(U)}=\mu $ .

We claim that $\nu $ is $\Gamma $ -invariant. Fix $A\in \mathcal {B}(X)$ and $g\in \Gamma $ , and we will show that $\nu (A)=\nu (gA)$ .

For $1\leq i\leq n$ , let $h_i:=g^{-1}g_i$ , $B_i:=A_i\cap g_i^{-1}A$ , and $C_i:=A_i\cap g_i^{-1}gA=A_i\cap h_i^{-1} A$ . By definition of $\nu $ , we have $\nu (A)=\sum _{i=1}^n\mu (B_i)$ and $\nu (gA)=\sum _{i=1}^n\mu (C_i)$ .

Moreover, one can readily check that $A=\bigsqcup _{i=1}^n g_i B_i=\bigsqcup _{i=1}^n h_iC_i$ .

For $1\leq i,j\leq n$ , let $B_{i,j}:=B_i\cap g_i^{-1}h_jC_j$ and $C_{i,j}:=h_j^{-1}g_iB_i\cap C_j$ .

Notice that, for $1\leq i\leq n$ ,

$$ \begin{align*}\bigsqcup_{j=1}^n B_{i,j}=B_i\cap g_i^{-1}A=B_i\end{align*} $$

and, for $1\leq j \leq n$ ,

$$ \begin{align*}\bigsqcup_{i=1}^n C_{i,j}=h_j^{-1}A\cap C_j=C_j.\end{align*} $$

Furthermore, $g_i B_{i,j}=h_jC_{i,j}$ , and hence $\mu (B_{i,j})=\mu (C_{i,j})$ , since $B_{i,j}$ and $C_{i,j}$ are contained in U for every $i,j$ . Therefore,

$$ \begin{align*} \nu(A)=\sum_{i=1}^n\mu(B_i)=\sum_{i,j=1}^n \mu(B_{i,j})=\sum_{i,j=1}^n\mu(C_{i,j})=\sum_{j=1}^n\mu(C_j)=\nu(gA). \end{align*} $$

Finally, we have that $j(\frac {\nu }{\nu (X)})=\frac {\nu |_{\mathcal {B}(U)}/\nu (X)}{\nu (U)/\nu (X)}=\nu |_{\mathcal {B}(U)}=\mu .$

Remark 3.2 Let $\Gamma \!\curvearrowright \! X$ and $\Lambda \!\curvearrowright \! Y$ be actions on compact spaces. The actions are said to be Kakutani equivalent [Reference LiLi18, Definition 2.14] if there exist clopen sets $A\subset X$ and $B\subset Y$ such that $X=\Gamma .A$ , $Y=\Lambda. B$ , and the partial transformation groupoids obtained by restriction to A and B are isomorphic. Proposition 3.1 implies that Kakutani equivalence induces a bijection between $\mathcal {P}_{\Gamma} (X)$ and $\mathcal {P}_\Lambda (X)$ .

The proof of the following result is analogous to [Reference Nyland and OrtegaNO19, Lemma 4.9(2)].

Lemma 3.3 Let $\alpha $ be a minimal action of a group $\Gamma $ on the Cantor set X. Given $U\subset X$ clopen, $x\in U$ , and $g\in \Gamma $ such that $g(x)\in U$ , there exists a neighborhood V of x and $h\in \operatorname {St}_{\mathsf {A}(\alpha )}(U^{\mathsf {c}})$ such that $g|_V=h|_V$ .

Proof Case 1: $g(x)\neq x$ . Take $k\in \Gamma $ such that $k(g(x))\in U\setminus \{x,g(x)\}$ . Let V be a clopen neighborhood of x such that V, $g(V)$ , and $kg(V)$ are disjoint subsets of U. Then the homeomorphisms $h_{2,1}:=g|_V$ and $h_{3,1}:=kg|_V$ give rise to a $3$ -multisection $\mathcal {F}$ such that $\mathcal {F}((123))|_V=g|_V$ and $\mathcal {F}((123))\in \operatorname {St}_{\mathsf {A}(\alpha )}(U^{\mathsf {c}})$ .

Case 2: $g(x)=x$ . Take $k\in \Gamma $ such that $k(x)\in U\setminus \{ x\}$ . By Case 1, there are $h_1,h_2\in \operatorname {St}_{\mathsf {A}(\alpha )}(U^{\mathsf {c}})$ and $V_1,V_2$ neighborhoods of x and $k(x)$ , respectively, such that $k|_{V_1}=h_1|_{V_1}$ and $gk^{-1}|_{V_2}=h_2|_{V_2}$ . Then $V:=V_1\cap k^{-1}(V_2)$ is a neighborhood of x such that $h_2h_1|_V=g|_V$ .

The next lemma uses the same idea of [Reference MatuiMB18, Corollary 6.5].

Lemma 3.4 Let $\alpha $ be a minimal action of a countable group $\Gamma $ on the Cantor set X and H a group such that $\mathsf {A}(\alpha )\leq H\leq \mathsf {F}(\alpha )$ . Then H is not $C^*$ -simple if and only if $\mathsf {A}(\alpha )_x^0$ is amenable for all $x\in X$ .

Proof Suppose H is not $C^*$ -simple. Then H contains a confined amenable subgroup. By [Reference MatuiMB18, Theorem 6.1], there exists $Q\subset X$ finite such that $\operatorname {St}_{\mathsf {A}(\alpha )}^0(Q)$ is amenable. Given $x\in X$ , take a net $(g_i)\subset \mathsf {F}(\alpha )$ such that $g_iq\to x$ for any $q\in Q$ (existence of such a net $(g_i)$ follows from minimality and proximality of $\mathsf {F}(\alpha )\!\curvearrowright \! X$ ; see [Reference MatuiMB18, Lemma 5.12]). Take K a limit point of $g_i\operatorname {St}_{\mathsf {A}(\alpha )}^0(Q)g_i^{-1}$ . One can readily check that $\mathsf {A}(\alpha )_x^0\leq K$ , and hence $\mathsf {A}(\alpha )_x^0$ is amenable.

Conversely, if $\mathsf {A}(\alpha )_x^0$ is amenable for all $x\in X$ , then since $\mathsf {A}(\alpha )_x^0$ is nontrivial for every x, it follows that the stabilizer URS $\mathcal {U}$ of $\mathsf {A}(\alpha )\!\curvearrowright \! X$ is a nontrivial amenable URS of $\mathsf {A}(\alpha )$ . By [Reference MatuiMB18, Theorem 6.1], any element of $\mathcal {U}$ is a confined subgroup of $\mathsf {F}(\alpha )$ (hence of H as well). Therefore, H is not $C^*$ -simple.

Theorem 3.5 Let $\alpha $ be a minimal action of a countable group $\Gamma $ on the Cantor set X. The following conditions are equivalent:

  1. (i) $\mathsf {A}(\alpha )$ is non-amenable.

  2. (ii) Any group H such that $\mathsf {A}(\alpha )\leq H\leq \mathsf {F}(\alpha )$ is $C^*$ -simple.

  3. (iii) There exists a $C^*$ -simple group H such that $\mathsf {A}(\alpha )\leq H\leq \mathsf {F}(\alpha )$ .

Proof The implications (ii) $\implies $ (iii) $\implies $ (i) are immediate.

(i) $\implies $ (ii): Suppose that there exists H non- $C^*$ -simple such that $\mathsf {A}(\alpha )\leq H\leq \mathsf {F}(\alpha )$ . By Lemma 3.4, $\mathsf {A}(\alpha )_x^0$ is amenable for every $x\in X$ .

Fix a clopen nonempty set U properly contained in X. Since, for any $x\in U^{\mathrm {c}}$ , we have $\Lambda :=\operatorname {St}_{\mathsf {A}(\alpha )}(U^{\mathsf {c}})\leq \mathsf {A}(\alpha )_x^0$ , it follows that $\Lambda $ is amenable.

Let $\mu \in \mathcal {P}_\Lambda (U)$ , and we claim that $\mu \in \mathcal {P}_{\Gamma} (U)$ . By regularity, it suffices to show that, for any $K\subset U$ compact and $g\in \Gamma $ such that $g(K)\subset U$ , it holds that $\mu (gK)=gK$ . By Lemma 3.3, there are $h_1,\dots ,h_n\in \Lambda $ and a partition $K=\bigsqcup _{i=1}^n K_i$ into compact sets such that $g|_{K_i}=h_i|_{K_i}$ for $1\leq i\leq n$ . Therefore,

$$ \begin{align*}\mu(gK)=\sum_{i=1}^n\mu(gK_i)=\sum_{i=1}^n\mu(h_iK_i)=\sum_{i=1}^n\mu(K_i)=\mu(K)\end{align*} $$

and $\mu \in \mathcal {P}_{\Gamma} (U)$ .

We conclude from Proposition 3.1 that there is $\nu \in \mathcal {P}_{\Gamma} (X)=\mathcal {P}_{\mathsf {F}(\alpha )}(X)$ . Furthermore, by minimality of the action, $\nu $ has full support. Let $\rho :=(\operatorname {Stab}_{\mathsf {A}(\alpha )}^0)_*\nu $ .

Given $g\in \Lambda \setminus \{e\}$ , we have $\rho (\{K\in \operatorname {Sub}(\mathsf {A}(\alpha )):g\in K\})\geq \nu (U^{\mathsf {c}})>0$ . Hence, $\rho $ is a nontrivial amenable IRS on $\mathsf {A}(\alpha )$ . Since $\mathsf {A}(\alpha )$ is simple, this implies that $\mathsf {A}(\alpha )$ is amenable.

The following is an immediate consequence of Theorem 3.5.

Corollary 3.6 Let $\alpha $ be a minimal action of a countable group on the Cantor set. Then $\mathsf {F}(\alpha )$ has the unique trace property if and only if it is $C^*$ -simple. If $\mathsf {A}(\alpha )=\mathsf {F}(\alpha )'$ , then $\mathsf {F}(\alpha )$ is non-amenable if and only if it is $C^*$ -simple.

Remark 3.7 If $\alpha $ is an action of a group on the noncompact Cantor set X, then the topological full group $\mathsf {F}(\alpha )$ is the group of homeomorphisms on X which are locally given by $\alpha $ and have compact support. Moreover, $\mathsf {A}(\alpha )$ is defined by requiring that the domains of the partial homeomorphisms of the multisections to be compact-open (as in [Reference MatuiMB18, Definition 5.1]). By arguing as in [Reference MatuiMB18, Corollary 6.5], the same conclusion of Theorem 3.5 and Corollary 3.6 holds in the noncompact case.

Example 3.8 It follows from [Reference MatuiMat12, Lemma 6.3] and [Reference MatuiMat15, Theorem 4.7] that, given a free minimal action $\alpha $ of $\mathbb {Z}^n$ on the Cantor set, it holds that $\mathsf {F}(\alpha )'=\mathsf {A}(\alpha )$ . Hence, the example of non-amenable topological full group coming from a Cantor minimal $\mathbb {Z}^2$ -system in [Reference Elek and MonodEM13] is $C^*$ -simple.

Acknowledgment

I am grateful to Eduard Ortega and Nicolás Matte Bon for helpful comments. I also thank the anonymous referee for suggestions which helped to improve the presentation of this work and for pointing out an incorrection in the original formulation of Proposition 3.1.

Footnotes

This project has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (Grant Agreement No. 817597).

References

Bader, U., Duchesne, B., and Lécureux, J., Amenable invariant random subgroups . Israel J. Math. 213(2016), no. 1, 399422, with an appendix by Phillip Wesolek.CrossRefGoogle Scholar
Breuillard, E., Kalantar, M., Kennedy, M., and Ozawa, N., ${C}^{\ast }$ -simplicity and the unique trace property for discrete groups . Publ. Math. Inst. Hautes Études Sci. 126(2017), 3571.CrossRefGoogle Scholar
Brix, K. A. and Scarparo, E., ${{C}}^{\ast }$ -simplicity and representations of topological full groups of groupoids . J. Funct. Anal. 277(2019), no. 9, 29812996.CrossRefGoogle Scholar
Elek, G. and Monod, N., On the topological full group of a minimal Cantor ${\mathbf{Z}}^2$ -system . Proc. Amer. Math. Soc. 141(2013), no. 10, 35493552.CrossRefGoogle Scholar
Juschenko, K. and Monod, N., Cantor systems, piecewise translations and simple amenable groups . Ann. of Math. (2) 178 (2013), no. 2, 775787.CrossRefGoogle Scholar
Kennedy, M., An intrinsic characterization of ${C}^{\ast }$ -simplicity . Ann. Sci. Éc. Norm. Supér. (4) 53(2020), no. 5, 11051119.CrossRefGoogle Scholar
Kerr, D. and Tucker-Drob, R., Dynamical alternating groups, stability, property Gamma, and inner amenability. Preprint, 2019. arXiv:1902.04131Google Scholar
Le Boudec, A. and Matte Bon, N., Subgroup dynamics and ${C}^{\ast }$ -simplicity of groups of homeomorphisms . Ann. Sci. Éc. Norm. Supér. (4) 51(2018), no. 3, 557602.CrossRefGoogle Scholar
Li, X., Dynamic characterizations of quasi-isometry and applications to cohomology . Algebr. Geom. Topol. 18(2018), no. 6, 34773535.CrossRefGoogle Scholar
Matte Bon, N., Rigidity properties of full groups of pseudogroups over the Cantor set. Preprint, 2018. arXiv:1801.10133 Google Scholar
Matui, H., Some remarks on topological full groups of Cantor minimal systems . Internat. J. Math. 17(2006), no. 2, 231251.CrossRefGoogle Scholar
Matui, H., Homology and topological full groups of étale groupoids on totally disconnected spaces . Proc. Lond. Math. Soc. (3) 104(2012), no. 1, 2756.CrossRefGoogle Scholar
Matui, H., Topological full groups of one-sided shifts of finite type . J. Reine Angew. Math. 705(2015), 3584.CrossRefGoogle Scholar
Nekrashevych, V., Simple groups of dynamical origin . Ergodic Theory Dynam. Systems 39(2019), no. 3, 707732.CrossRefGoogle Scholar
Nyland, P. and Ortega, E., Topological full groups of ample groupoids with applications to graph algebras . Internat. J. Math. 30(2019), no. 4, 66, Article ID 1950018.CrossRefGoogle Scholar