This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\addtokomafont

labelinglabel

Soficity for group actions on sets and applications

David Gao Department of Mathematical Sciences, UCSD, 9500 Gilman Dr, La Jolla, CA 92092, USA weg002@ucsd.edu Srivatsav Kunnawalkam Elayavalli Department of Mathematical Sciences, UCSD, 9500 Gilman Dr, La Jolla, CA 92092, USA srivatsav.kunnawalkam.elayavalli@vanderbilt.edu https://sites.google.com/view/srivatsavke/home  and  Gregory Patchell Department of Mathematical Sciences, UCSD, 9500 Gilman Dr, La Jolla, CA 92092, USA gpatchel@ucsd.edu
Abstract.

In this article we develop a notion of soficity for actions of countable groups on sets. We show two equivalent perspectives, several natural properties and examples. Notable examples include arbitrary actions of both amenable groups and free groups, and actions of sofic groups with locally finite stabilizers. As applications we prove soficity for generalized wreath products (and amalgamated free generalized wreath products) of sofic groups where the underlying group action is sofic. This generalizes the result of Hayes and Sale [HS18], and proves soficity for many new families of groups.

1. Introduction

The study of finitary approximations of countable groups is a modern area of interest in group theory. One such finitary approximation property is the notion of soficity due to Gromov (for a detailed survey with references see [Pes08]). Groups that satisfy this property are known to additionally verify other important group theoretic questions such as Gottschalk surjunctivity conjecture [KL11], Kervaire conjecture (see Corollary 10.4 of [Pes08]), Kaplansky direct finiteness conjecture [ES04], determinant conjecture [Lüc02], Connes embedding problem [Con76] etc (see also [Bow10, Tho08]). There is therefore an interest to identify more examples of soficity in countable groups, irrespective of the notorious open problem of the existence of a non sofic group. There are currently many known examples of sofic groups (two elementary and important sources of examples include amenable groups, residually finite groups), and examples of group operations that preserve soficity. These include taking direct products of sofic groups, amalgamated free products and HNN extensions of sofic groups over amenable amalgams [P1̆1, ES11, DKP14, Pop14], wreath products of sofic groups [HS18].

In this paper we strictly generalize [HS18] and identify new examples of soficity in the setting of generalized wreath product groups. In order to identify the correct level of generality, we first had to understand and develop a natural notion of soficity for a group action on a set. To our knowledge, a satisfactory definition of this is currently unavailable in the literature (note that there are very satisfying definitions of soficity in the different more analytic setting of probability measure preserving actions, see [EL10a, Pau14]). Aside from our applications in this paper, we would like to place an emphasis on the fact that we fill this void, and begin developing a fruitful theory of soficity for group actions on sets, with potential other use in the future. We provide the definition below:

Let GG be a countable discrete group, XX be a countable discrete set, α:GX\alpha:G\curvearrowright X be an action, AA be a finite set, φ:GSym(A)\varphi:G\rightarrow\textrm{Sym}(A) be a map. For a finite subset FGF\subseteq G and ϵ>0\epsilon>0, φ\varphi is called (F,ϵ)(F,\epsilon)-multiplicative if d(φ(gh),φ(g)φ(h))<ϵd(\varphi(gh),\varphi(g)\varphi(h))<\epsilon for all g,hFg,h\in F, where dd denotes the normalized Hamming distance. For finite subsets FGF\subseteq G, EXE\subseteq X, and ϵ>0\epsilon>0, φ\varphi is called an (F,E,ϵ)(F,E,\epsilon)-orbit approximation of α\alpha if there exists a finite set BB and a subset SAS\subseteq A s.t. |S|>(1ϵ)|A||S|>(1-\epsilon)|A| and for each sSs\in S there is an injective map πs:EB\pi_{s}:E\hookrightarrow B s.t. πφ(g)s(x)=πs(α(g1)x)\pi_{\varphi(g)s}(x)=\pi_{s}(\alpha(g^{-1})x) for all sSs\in S, gFg\in F, xEx\in E, whenever φ(g)sS\varphi(g)s\in S and α(g1)xE\alpha(g^{-1})x\in E.

Definition (see Definition 2.1).

α\alpha is called sofic if for all finite subsets FGF\subseteq G, EXE\subseteq X, and ϵ>0\epsilon>0, there exists a finite set AA and a map φ:GSym(A)\varphi:G\rightarrow\textrm{Sym}(A) which is unital, (F,ϵ)(F,\epsilon)-multiplicative, and an (F,E,ϵ)(F,E,\epsilon)-orbit approximation of α\alpha.

If the reader is more comfortable working in the ultrapower framework, we redirect them to Proposition 2.9, where we present a natural equivalent definition involving the natural action of the universal sofic group on the Loeb measure space (see for instance [HE23]). This could even serve as an additional motivation for why we define soficity in this manner.

The first example is the left action of a sofic group GG on itself. We show (in Theorem 2.14 and Theorem 2.12) that this is sofic if and only if GG is sofic. More generally, we are able to show (in Theorem 2.14) that GG acting on the coset space G/HG/H is sofic where HH is any locally finite subgroup. On the other side, we have (see Theorem 2.12) that if the action of GG on a coset space G/HG/H is sofic, then there is a normal subgroup NN of GG such that NHN\leq H and G/NG/N is sofic. This by itself could give help to finding new examples of sofic groups. 111We would like to remind the readers that it remains a puzzling question whether G/NG/N is sofic for GG sofic and NN is a finite normal subgroup, say even |N|=2|N|=2.

The following are rather quick observations: If α:GX\alpha:G\curvearrowright X is the composition of a quotient map q:GHq:G\rightarrow H and a sofic action β:HX\beta:H\curvearrowright X, then α\alpha is sofic; If α:GX\alpha:G\curvearrowright X is sofic, then the restriction of α\alpha to each of its orbits is sofic; If α:GX\alpha:G\curvearrowright X is sofic and HH is a subgroup of GG, then α|H\alpha|_{H} is sofic; If G1G2G_{1}\subseteq G_{2}\subseteq\cdots is an increasing sequence of subgroups of GG whose union is GG, and α:GX\alpha:G\curvearrowright X restricted to each GiG_{i} is sofic, then α\alpha is sofic. Moreover we have that if the restriction of α:GX\alpha:G\curvearrowright X to each of its orbits is sofic, then α\alpha is sofic (see Proposition 2.16). This naturally reduces the study of sofic actions to transitive ones.

It is of course an open question whether all actions of sofic groups are sofic. Interestingly in Theorem 2.17 and Theorem 2.19 we find that there are two classes of groups whose arbitrary actions are sofic: amenable groups and free groups. We are ready to now state our main result (see Theorem 3.6) which recovers and generalizes the main result of Hayes and Sale [HS18].

Theorem A.

Let G,HG,H be sofic groups, α:HX\alpha:H\curvearrowright X be a sofic action. Then the generalized wreath product GαHG\wr_{\alpha}H is sofic.

The proof of the main result is very much inspired by and pushes the ideas of [HS18]. The above result also applies to the setting of amalgamated free generalized wreath product (see Theorem 3.7). The unfamiliar reader is directed to Definition 3.3. At this point we would like to mention that the natural variant of the above result can also be proved in the context of hyperlinearity, see Theorems 3.8, 3.9 and Corollary 3.10.

The following are some remarks and questions we wish to highlight out before concluding the introduction. Let (Ω,μ)(\Omega,\mu) be a standard probability space, GG be a sofic group, α:GX\alpha:G\curvearrowright X be a sofic action. Then the induced generalized Bernoulli shift G(ΩX,μ|X|)G\curvearrowright(\Omega^{X},\mu^{\otimes|X|}) is sofic in the sense of [P1̆1]. This places an aspect of our work in the context of Elek-Lippner’s work [EL10b] defining soficity for equivalence relations and of Paunescu [P1̆1]. The most important open question that arises from our work is whether the converse of our main result holds: Let G,HG,H be nontrivial countable groups, α:HX\alpha:H\curvearrowright X be an action. Then the generalized wreath product GαHG\wr_{\alpha}H is sofic if and only if GG and HH are sofic and the action α\alpha is sofic? As we highlight in Section 4 a positive answer to this question follows from a positive answer to the following stability question: Suppose we have actions αi:GiX\alpha_{i}:G_{i}\curvearrowright X which commute with each other and where ii ranges over a countable index set. The actions naturally give rise to an action α:iGiX\alpha:\oplus_{i}G_{i}\curvearrowright X. Then, is α\alpha sofic if and only if all αi\alpha_{i} are sofic?

2. Sofic actions

In the following, dd, when denoting a metric on a symmetric group of a finite set, shall always be understood to be the normalized Hamming distance, unless specified otherwise. The following is our finitary definition of sofic actions:

Definition 2.1.

Let GG be a countable discrete group, XX be a countable discrete set, α:GX\alpha:G\curvearrowright X be an action, AA be a finite set, φ:GSym(A)\varphi:G\rightarrow\textrm{Sym}(A) be a map (not necessarily a homomorphism):

  1. (1)

    φ\varphi is called unital if φ(1G)=1\varphi(1_{G})=1;

  2. (2)

    For a finite subset FGF\subseteq G and ϵ>0\epsilon>0, φ\varphi is called (F,ϵ)(F,\epsilon)-multiplicative if d(φ(gh),φ(g)φ(h))<ϵd(\varphi(gh),\varphi(g)\varphi(h))<\epsilon for all g,hFg,h\in F;

  3. (3)

    For finite subsets FGF\subseteq G, EXE\subseteq X, and ϵ>0\epsilon>0, φ\varphi is called an (F,E,ϵ)(F,E,\epsilon)-orbit approximation of α\alpha if there exists a finite set BB and a subset SAS\subseteq A s.t. |S|>(1ϵ)|A||S|>(1-\epsilon)|A| and for each sSs\in S there is an injective map πs:EB\pi_{s}:E\hookrightarrow B s.t. πφ(g)s(x)=πs(α(g1)x)\pi_{\varphi(g)s}(x)=\pi_{s}(\alpha(g^{-1})x) for all sSs\in S, gFg\in F, xEx\in E, whenever φ(g)sS\varphi(g)s\in S and α(g1)xE\alpha(g^{-1})x\in E;

  4. (4)

    Recall that GG is called sofic if for all finite subsets FGF\subseteq G and ϵ>0\epsilon>0 there exists a finite set AA and a map φ:GSym(A)\varphi:G\to\textrm{Sym}(A) which is unital, (F,ϵ)(F,\epsilon)-multiplicative, and d(1,φ(g))>1ϵd(1,\varphi(g))>1-\epsilon for all gF{e}.g\in F\setminus\{e\}.

  5. (5)

    α\alpha is called sofic if for all finite subsets FGF\subseteq G, EXE\subseteq X, and ϵ>0\epsilon>0, there exists a finite set AA and a map φ:GSym(A)\varphi:G\rightarrow\textrm{Sym}(A) which is unital, (F,ϵ)(F,\epsilon)-multiplicative, and an (F,E,ϵ)(F,E,\epsilon)-orbit approximation of α\alpha.

Now we will place the above definition in the non-finitary setting of ultraproducts. We need some preliminary definitions and notations.

Definition 2.2.

Let (Xn)(X_{n}) be a sequence of sets, 𝒰\mathcal{U} be a free ultrafilter on \mathbb{N}. Then 𝒰Xn\prod_{\mathcal{U}}X_{n}, the algebraic ultraproduct of (Xn)(X_{n}), is defined as 𝒰Xn=Xn/\prod_{\mathcal{U}}X_{n}=\prod X_{n}/\sim where fgf\sim g iff {n:f(n)=g(n)}𝒰\{n:f(n)=g(n)\}\in\mathcal{U}. We shall write (xn)𝒰(x_{n})_{\mathcal{U}} to mean the element of 𝒰Xn\prod_{\mathcal{U}}X_{n} represented by (xn)Xn(x_{n})\in\prod X_{n}. If (AnXn)(A_{n}\subseteq X_{n}) is a sequence of subsets, then we shall write (An)𝒰(A_{n})_{\mathcal{U}} to mean,

(An)𝒰={(xn)𝒰𝒰Xn:{n:xnAn}𝒰}(A_{n})_{\mathcal{U}}=\{(x_{n})_{\mathcal{U}}\in\prod_{\mathcal{U}}X_{n}:\{n:x_{n}\in A_{n}\}\in\mathcal{U}\}
Definition 2.3.

Let (Xn)(X_{n}) and (Yn)(Y_{n}) be two sequences of sets, 𝒰\mathcal{U} be a free ultrafilter on \mathbb{N}. A map φ:𝒰Xn𝒰Yn\varphi:\prod_{\mathcal{U}}X_{n}\rightarrow\prod_{\mathcal{U}}Y_{n} is called liftable if there exists a sequence of maps φn:XnYn\varphi_{n}:X_{n}\rightarrow Y_{n} such that φ((xn)𝒰)=(φn(xn))𝒰\varphi((x_{n})_{\mathcal{U}})=(\varphi_{n}(x_{n}))_{\mathcal{U}} for all (xn)𝒰𝒰Xn(x_{n})_{\mathcal{U}}\in\prod_{\mathcal{U}}X_{n}.

Definition 2.4 (see for instance [Pau14]).

Let [n]={1,,n}[n]=\{1,\cdots,n\}, μn\mu_{n} be the normalized counting measure on [n][n], 𝒰\mathcal{U} be a free ultrafilter on \mathbb{N}. Then the Loeb measure space, denoted by 𝒰([n],μn)\prod_{\mathcal{U}}([n],\mu_{n}), is defined to have the underlying set 𝒰[n]\prod_{\mathcal{U}}[n]. For a sequence (An[n])(A_{n}\subseteq[n]), we define μ𝒰((An)𝒰)=lim𝒰μn(An)\mu_{\mathcal{U}}((A_{n})_{\mathcal{U}})=\lim_{\mathcal{U}}\mu_{n}(A_{n}). This can be extended to a measure on the σ\sigma-algebra generated by all such (An)𝒰(A_{n})_{\mathcal{U}}.

Definition 2.5.

Let (Xn,dn)(X_{n},d_{n}) be a sequence of metric spaces, 𝒰\mathcal{U} be a free ultrafilter on \mathbb{N}. Then 𝒰(Xn,dn)\prod_{\mathcal{U}}(X_{n},d_{n}), the metric ultraproduct of (Xn,dn)(X_{n},d_{n}), is defined to have the underlying set Xn/\prod X_{n}/\sim where fgf\sim g iff lim𝒰dn(f(n),g(n))=0\lim_{\mathcal{U}}d_{n}(f(n),g(n))=0. We shall write (xn)𝒰(x_{n})_{\mathcal{U}} to mean the element of 𝒰Xn\prod_{\mathcal{U}}X_{n} represented by (xn)Xn(x_{n})\in\prod X_{n}. Then the metric on this space is defined by d𝒰((xn)𝒰,(yn)𝒰)=lim𝒰dn(xn,yn)d_{\mathcal{U}}((x_{n})_{\mathcal{U}},(y_{n})_{\mathcal{U}})=\lim_{\mathcal{U}}d_{n}(x_{n},y_{n}).

Definition 2.6.

Let 𝒰\mathcal{U} be a free ultrafilter on \mathbb{N}. Let XX consists of all liftable maps 𝒰([n],μn)𝒰\prod_{\mathcal{U}}([n],\mu_{n})\rightarrow\prod_{\mathcal{U}}\mathbb{N}. We observe that the set {x:f(x)=g(x)}\{x:f(x)=g(x)\} for any f,gXf,g\in X is always of the form (An)𝒰(A_{n})_{\mathcal{U}} and therefore measurable. Let 𝕏𝒰1\mathbb{X}^{1}_{\mathcal{U}} be X/X/\sim where fgf\sim g iff they coincide a.e. Let d𝒰d_{\mathcal{U}} be the metric on 𝕏𝒰1\mathbb{X}^{1}_{\mathcal{U}} given by d𝒰([f],[g])=μ𝒰{x:f(x)g(x)}d_{\mathcal{U}}([f],[g])=\mu_{\mathcal{U}}\{x:f(x)\neq g(x)\}. We observe that the universal sofic group 𝒰(Sn,d)\prod_{\mathcal{U}}(S_{n},d) naturally acts on the Loeb measure space via pmp automorphisms, and that the pre-composition of a liftable map with a sequence of permutations results in a liftable map. Therefore, 𝒰(Sn,d)\prod_{\mathcal{U}}(S_{n},d) naturally acts on 𝕏𝒰1\mathbb{X}^{1}_{\mathcal{U}} via pre-composition of inverses. This action shall be denoted by 𝕊𝒰1\mathbb{S}^{1}_{\mathcal{U}} and called the first universal sofic action. We observe that this action preserves the metric d𝒰d_{\mathcal{U}}.

Definition 2.7.

Let 𝒰\mathcal{U} be a free ultrafilter on \mathbb{N}. For each nn, let XnX_{n} be the collection of all maps [n][n]\rightarrow\mathbb{N} and dnd_{n} be the normalized Hamming distance on XnX_{n}. Let 𝕏𝒰2=𝒰(Xn,dn)\mathbb{X}^{2}_{\mathcal{U}}=\prod_{\mathcal{U}}(X_{n},d_{n}). We define the following action 𝕊𝒰2\mathbb{S}^{2}_{\mathcal{U}} of the universal sofic group 𝒰(Sn,d)\prod_{\mathcal{U}}(S_{n},d) on 𝕏𝒰2\mathbb{X}^{2}_{\mathcal{U}} by 𝕊𝒰2((gn)𝒰)((fn)𝒰)=(fngn1)𝒰\mathbb{S}^{2}_{\mathcal{U}}((g_{n})_{\mathcal{U}})((f_{n})_{\mathcal{U}})=(f_{n}\circ g_{n}^{-1})_{\mathcal{U}}. We easily observe that this is a well-defined action that preserves the metric d𝒰d_{\mathcal{U}}. We shall call 𝕊𝒰2\mathbb{S}^{2}_{\mathcal{U}} the second universal sofic action.

Lemma 2.8.

There is an isometric bijection ι:𝕏𝒰2𝕏𝒰1\iota:\mathbb{X}^{2}_{\mathcal{U}}\rightarrow\mathbb{X}^{1}_{\mathcal{U}} which is equivariant under the two universal sofic actions.

Proof.

For each (fn)𝒰𝕏𝒰2(f_{n})_{\mathcal{U}}\in\mathbb{X}^{2}_{\mathcal{U}}, defined ι((fn)𝒰)\iota((f_{n})_{\mathcal{U}}) to be the map 𝒰([n],μn)𝒰\prod_{\mathcal{U}}([n],\mu_{n})\rightarrow\prod_{\mathcal{U}}\mathbb{N} defined by ι((fn)𝒰)((xn)𝒰)=(fn(xn))𝒰\iota((f_{n})_{\mathcal{U}})((x_{n})_{\mathcal{U}})=(f_{n}(x_{n}))_{\mathcal{U}}. It is easy to verify that this indeed satisfies the conditions of the lemma. ∎

In light of the lemma above, we shall simply identify the two universal sofic actions and denote this action by 𝕊𝒰:𝒰(Sn,d)𝕏𝒰\mathbb{S}_{\mathcal{U}}:\prod_{\mathcal{U}}(S_{n},d)\curvearrowright\mathbb{X}_{\mathcal{U}}. Now we present the natural ultraproduct definition of sofic actions.

Proposition 2.9.

Let GG be a countable discrete group, XX be a countable discrete set, α:GX\alpha:G\curvearrowright X be an action. The following are equivalent,

  1. (1)

    α\alpha is sofic;

  2. (2)

    There exists a free ultrafilter 𝒰\mathcal{U} on \mathbb{N}, a group homomorphism φ:G𝒰(Sn,d)\varphi:G\rightarrow\prod_{\mathcal{U}}(S_{n},d), and a map π:X𝕏𝒰\pi:X\rightarrow\mathbb{X}_{\mathcal{U}} s.t. 𝕊𝒰(φ(g))(π(x))=π(α(g)x)\mathbb{S}_{\mathcal{U}}(\varphi(g))(\pi(x))=\pi(\alpha(g)x) for all xXx\in X, gGg\in G and s.t. d𝒰(π(x),π(y))=1d_{\mathcal{U}}(\pi(x),\pi(y))=1 for all xyXx\neq y\in X.

Proof.

()(\Rightarrow) Fix increasing sequences of finite subsets F1F2GF_{1}\subseteq F_{2}\subseteq\cdots\subseteq G and E1E2XE_{1}\subseteq E_{2}\subseteq\cdots\subseteq X s.t. iFi=G\cup_{i}F_{i}=G, iEi=X\cup_{i}E_{i}=X. Fix a decreasing sequence ϵi>0\epsilon_{i}>0 s.t. limiϵi=0\lim_{i}\epsilon_{i}=0. For each ii, let φi:GSym(Ai)\varphi_{i}:G\rightarrow\textrm{Sym}(A_{i}) be a unital, (Fi,ϵi)(F_{i},\epsilon_{i})-multiplicative, and an (Fi,Ei,ϵi)(F_{i},E_{i},\epsilon_{i})-orbit approximation of α\alpha. By taking the Cartesian products of AiA_{i} with auxiliary finite sets if necessary, we may assume |Ai||A_{i}| is strictly increasing. By definition of (Fi,Ei,ϵi)(F_{i},E_{i},\epsilon_{i})-orbit approximation, there exists a finite set BiB_{i} and a subset SiAiS_{i}\subseteq A_{i} s.t. |Si|>(1ϵi)|Ai||S_{i}|>(1-\epsilon_{i})|A_{i}| and for each sSis\in S_{i} there is an injective map πsi:EiBi\pi^{i}_{s}:E_{i}\hookrightarrow B_{i} s.t. πφi(g)si(x)=πsi(α(g1)x)\pi^{i}_{\varphi_{i}(g)s}(x)=\pi^{i}_{s}(\alpha(g^{-1})x) for all sSis\in S_{i}, gFig\in F_{i}, xEix\in E_{i}, whenever φi(g)sSi\varphi_{i}(g)s\in S_{i} and α(g1)xEi\alpha(g^{-1})x\in E_{i}. By embedding BiB_{i} into \mathbb{N} we may take the co-domain of πsi\pi^{i}_{s} to be \mathbb{N}. Let 𝒰\mathcal{U} be any free ultrafilter on \mathbb{N} containing the set {|Ai|:i1}\{|A_{i}|:i\geq 1\}. We may then define φ:G𝒰(Sn,d)\varphi:G\rightarrow\prod_{\mathcal{U}}(S_{n},d) by defining φ(g)=(gn)𝒰\varphi(g)=(g_{n})_{\mathcal{U}} with gn=φi(g)g_{n}=\varphi_{i}(g) whenever n=|Ai|n=|A_{i}| and gn=1g_{n}=1 otherwise. Since {|Ai|:i1}𝒰\{|A_{i}|:i\geq 1\}\in\mathcal{U} and φi\varphi_{i} is (Fi,ϵi)(F_{i},\epsilon_{i})-multiplicative, we see that φ\varphi is a group homomorphism.

We then define π:X𝕏𝒰\pi:X\rightarrow\mathbb{X}_{\mathcal{U}} as follows: For each xXx\in X, nn\in\mathbb{N}, define πx,n:[n]\pi_{x,n}:[n]\rightarrow\mathbb{N},

πx,n(s)={πsi(x),if n=|Ai| and xEi and sSi0,otherwise\pi_{x,n}(s)=\begin{cases}\pi^{i}_{s}(x),\textrm{if }n=|A_{i}|\textrm{ and }x\in E_{i}\textrm{ and }s\in S_{i}\\ 0,\textrm{otherwise}\end{cases}

For each xXx\in X, π(x)\pi(x) shall be represented by the sequence of maps πx,n\pi_{x,n}. We observe that as {|Ai|:i1}𝒰\{|A_{i}|:i\geq 1\}\in\mathcal{U}, it does not matter how πx,n\pi_{x,n} is defined when n{|Ai|:i1}n\notin\{|A_{i}|:i\geq 1\}. Now, let xyXx\neq y\in X, then for large enough ii we have x,yEix,y\in E_{i}. For sSis\in S_{i}, we then have πx,|Ai|(s)=πsi(x)πsi(y)=πy,|Ai|(s)\pi_{x,|A_{i}|}(s)=\pi^{i}_{s}(x)\neq\pi^{i}_{s}(y)=\pi_{y,|A_{i}|}(s) as πsi\pi^{i}_{s} is injective. As |Si||Ai|>1ϵi1\frac{|S_{i}|}{|A_{i}|}>1-\epsilon_{i}\rightarrow 1, d|Ai|(πx,|Ai|,πy,|Ai|)|Si||Ai|1d_{|A_{i}|}(\pi_{x,|A_{i}|},\pi_{y,|A_{i}|})\geq\frac{|S_{i}|}{|A_{i}|}\rightarrow 1, so d𝒰(π(x),π(y))=1d_{\mathcal{U}}(\pi(x),\pi(y))=1.

Finally, fix xXx\in X, gGg\in G. Then for large ii, xEix\in E_{i}, g1Fig^{-1}\in F_{i}, and α(g)xEi\alpha(g)x\in E_{i}. Now, for any sSiφi(g1)1Sis\in S_{i}\cap\varphi_{i}(g^{-1})^{-1}S_{i}, by definition of πx,n\pi_{x,n} and (Fi,Ei,ϵi)(F_{i},E_{i},\epsilon_{i})-orbit approximation we have,

πx,|Ai|(φi(g1)s)=πφi(g1)si(x)=πsi(α(g)x)=πα(g)x,|Ai|(s)\pi_{x,|A_{i}|}(\varphi_{i}(g^{-1})s)=\pi^{i}_{\varphi_{i}(g^{-1})s}(x)=\pi^{i}_{s}(\alpha(g)x)=\pi_{\alpha(g)x,|A_{i}|}(s)

Since |Siφi(g1)1Si||Ai|>12ϵi1\frac{|S_{i}\cap\varphi_{i}(g^{-1})^{-1}S_{i}|}{|A_{i}|}>1-2\epsilon_{i}\rightarrow 1, this means the maps given by 𝕊𝒰(φ(g))(π(x))\mathbb{S}_{\mathcal{U}}(\varphi(g))(\pi(x)) and by π(α(g)x)\pi(\alpha(g)x) coincide on a set of measure 1 in 𝒰([n],μn)\prod_{\mathcal{U}}([n],\mu_{n}), whence they are identified in 𝕏𝒰\mathbb{X}_{\mathcal{U}}. This proves the claim.

()(\Leftarrow) For each gGg\in G, we shall write φ(g)=(gn)𝒰\varphi(g)=(g_{n})_{\mathcal{U}}. Since φ(1G)=1\varphi(1_{G})=1, we shall in particular choose φ(1G)=(1)𝒰\varphi(1_{G})=(1)_{\mathcal{U}}. Let φn:GSn\varphi_{n}:G\rightarrow S_{n} be defined by φn(g)=gn\varphi_{n}(g)=g_{n}. Now, fix finite FGF\subseteq G, EXE\subseteq X, and ϵ>0\epsilon>0, we shall show that there exists nn s.t. φn\varphi_{n} is unital, (F,ϵ)(F,\epsilon)-multiplicative, and an (F,E,ϵ)(F,E,\epsilon)-orbit approximation of α\alpha. We observe that φn\varphi_{n} is unital for all nn. Since FF is finite and φ\varphi is a group homomorphism, there exists L1𝒰L_{1}\in\mathcal{U} s.t. for all nL1n\in L_{1}, φn\varphi_{n} is (F,ϵ)(F,\epsilon)-multiplicative.

Now, we observe that, in the definition of (F,E,ϵ)(F,E,\epsilon)-orbit approximation of α\alpha, it is not necessary that BB is a finite set, as, for an infinite BB, we may simply restrict BB to sSπs(E)\cup_{s\in S}\pi_{s}(E) and the latter set is finite. Thus, we may set B=B=\mathbb{N}. Choose ϵ>0\epsilon^{\prime}>0 s.t. 1|E|2ϵ|F||E|ϵ1ϵ1-|E|^{2}\epsilon^{\prime}-|F||E|\epsilon^{\prime}\geq 1-\epsilon. Now, for each xEx\in E, we represent π(x)\pi(x) as a sequence of maps (πx,n)𝒰(\pi_{x,n})_{\mathcal{U}} with πx,n:[n]=B\pi_{x,n}:[n]\rightarrow\mathbb{N}=B. We first observe that, for any xyEx\neq y\in E, as d𝒰(π(x),π(y))=1d_{\mathcal{U}}(\pi(x),\pi(y))=1, there exists L2,x,y𝒰L_{2,x,y}\in\mathcal{U} s.t. dn(πx,n,πy,n)>1ϵd_{n}(\pi_{x,n},\pi_{y,n})>1-\epsilon^{\prime} for all nL2,x,yn\in L_{2,x,y}. Let L2=xyEL2,x,yL_{2}=\cap_{x\neq y\in E}L_{2,x,y}. Since EE is finite, L2𝒰L_{2}\in\mathcal{U}. For each nL2n\in L_{2}, let,

S~n=xyE{s[n]:πx,n(s)πy,n(s)}\tilde{S}_{n}=\cap_{x\neq y\in E}\{s\in[n]:\pi_{x,n}(s)\neq\pi_{y,n}(s)\}

By assumption |S~n|n>1|E|2ϵ\frac{|\tilde{S}_{n}|}{n}>1-|E|^{2}\epsilon^{\prime}. For each sS~ns\in\tilde{S}_{n}, defined πsn:EB\pi^{n}_{s}:E\rightarrow B by πsn(x)=πx,n(s)\pi^{n}_{s}(x)=\pi_{x,n}(s). Then πsn\pi^{n}_{s} is injective for all sS~ns\in\tilde{S}_{n}.

Now, fix any gFg\in F, xEx\in E with α(g1)xE\alpha(g^{-1})x\in E. Recall that 𝕊𝒰(φ(g1))(π(x))=π(α(g1)x)\mathbb{S}_{\mathcal{U}}(\varphi(g^{-1}))(\pi(x))=\pi(\alpha(g^{-1})x). 𝕊𝒰(φ(g1))(π(x))\mathbb{S}_{\mathcal{U}}(\varphi(g^{-1}))(\pi(x)) is represented by the sequence of maps πx,nφn(g)\pi_{x,n}\circ\varphi_{n}(g) while π(α(g1)x)\pi(\alpha(g^{-1})x) is represented by the sequence of maps πα(g1)x,n\pi_{\alpha(g^{-1})x,n}. Thus, there exists L3,g,x𝒰L_{3,g,x}\in\mathcal{U} s.t. dn(πx,nφn(g),πα(g1)x,n)<ϵd_{n}(\pi_{x,n}\circ\varphi_{n}(g),\pi_{\alpha(g^{-1})x,n})<\epsilon^{\prime} for all nL3,g,xn\in L_{3,g,x}. Let L3=gF,xEα(g)EL3,g,xL_{3}=\cap_{g\in F,x\in E\cap\alpha(g)E}L_{3,g,x}. Again, as FF and EE are finite, L3𝒰L_{3}\in\mathcal{U}. For any nL2L3n\in L_{2}\cap L_{3}, define,

Sn=S~ngF,xEα(g)E{s[n]:[πx,nφn(g)](s)=πα(g1)x,n(s)}S_{n}=\tilde{S}_{n}\cap\bigcap_{g\in F,x\in E\cap\alpha(g)E}\{s\in[n]:[\pi_{x,n}\circ\varphi_{n}(g)](s)=\pi_{\alpha(g^{-1})x,n}(s)\}

Then |Sn|n>1|E|2ϵ|F||E|ϵ1ϵ\frac{|S_{n}|}{n}>1-|E|^{2}\epsilon^{\prime}-|F||E|\epsilon^{\prime}\geq 1-\epsilon. By definition, πφn(g)sn(x)=πsn(α(g1)x)\pi^{n}_{\varphi_{n}(g)s}(x)=\pi^{n}_{s}(\alpha(g^{-1})x) for all sSns\in S_{n}, gFg\in F, xEx\in E, whenever φn(g)sSn\varphi_{n}(g)s\in S_{n} and α(g1)xE\alpha(g^{-1})x\in E. This shows that for all nL1L2L3n\in L_{1}\cap L_{2}\cap L_{3}\neq\varnothing, φn\varphi_{n} is unital, (F,ϵ)(F,\epsilon)-multiplicative, and an (F,E,ϵ)(F,E,\epsilon)-orbit approximation of α\alpha. ∎

Remark 2.10.

If we regard the discrete set XX as equipped with the discrete metric, i.e., dX(x,y)=1d_{X}(x,y)=1 whenever xyx\neq y, then the distance condition in the second condition of the proposition above can be reduced to saying that π\pi is isometric. This inspires the following generalization of sofic actions to actions on separable metric spaces:

Definition 2.11.

Let GG be a countable discrete group, XX be a separable metric space with diameter bounded by 1, α:GX\alpha:G\curvearrowright X be an isometric action. Then α\alpha is called sofic if there exists a free ultrafilter 𝒰\mathcal{U} on \mathbb{N}, a group homomorphism φ:G𝒰(Sn,d)\varphi:G\rightarrow\prod_{\mathcal{U}}(S_{n},d), and an isometric embedding π:X𝕏𝒰\pi:X\rightarrow\mathbb{X}_{\mathcal{U}} s.t. 𝕊𝒰(φ(g))(π(x))=π(α(g)x)\mathbb{S}_{\mathcal{U}}(\varphi(g))(\pi(x))=\pi(\alpha(g)x) for all xXx\in X, gGg\in G.

In the case of countable sets equipped with the discrete metric, this simply reduces to our previous definition of sofic actions.

Theorem 2.12.

Let GG be a countable discrete group, HGH\leq G be a subgroup. If the left multiplication action α:GG/H\alpha:G\curvearrowright G/H is sofic, then there exists a normal subgroup NGN\trianglelefteq G s.t. NHN\leq H and G/NG/N is sofic. In particular, if the left multiplication action α:GG\alpha:G\curvearrowright G is sofic, then GG is sofic.

Proof.

By Proposition 2.9, there exists a free ultrafilter 𝒰\mathcal{U} on \mathbb{N}, a group homomorphism φ:G𝒰(Sn,d)\varphi:G\rightarrow\prod_{\mathcal{U}}(S_{n},d), and an injective map π:G/H𝕏𝒰\pi:G/H\rightarrow\mathbb{X}_{\mathcal{U}} s.t. 𝕊𝒰(φ(g))(π(x))=π(α(g)x)\mathbb{S}_{\mathcal{U}}(\varphi(g))(\pi(x))=\pi(\alpha(g)x) for all xG/Hx\in G/H, gGg\in G. Take N=ker(φ)N=\textrm{ker}(\varphi). Then G/NG/N embeds into 𝒰(Sn,d)\prod_{\mathcal{U}}(S_{n},d) and is thus sofic. Assume NHN\not\leq H. We may then let gNHg\in N\setminus H. φ(g)=1\varphi(g)=1, so π(H)=𝕊𝒰(φ(g))(π(H))=π(α(g)H)=π(gH)\pi(H)=\mathbb{S}_{\mathcal{U}}(\varphi(g))(\pi(H))=\pi(\alpha(g)H)=\pi(gH). As π\pi is injective, H=gHH=gH, a contradiction. Thus, we must have NHN\leq H. ∎

The converse of the “in particular” part of the above theorem is also true. In fact, we shall prove a stronger result in Theorem 2.14. First, though, we need a lemma:

Lemma 2.13.

Suppose GG is a sofic group. Let FGF\subseteq G be a finite subset and ϵ>0\epsilon>0. Then there exists a finite set AA and a unital, (F,ϵ)(F,\epsilon)-multiplicative map φ:GSym(A)\varphi:G\rightarrow\textrm{Sym}(A) such that |S|>(1ϵ)|A||S|>(1-\epsilon)|A|, where we define,

S1={sA:φ(g)sφ(h)s,g,hF,gh},S2={sA:φ(gh)s=φ(g)φ(h)s,g,hF},andS=S1S2.\begin{split}S_{1}&=\{s\in A:\varphi(g)s\neq\varphi(h)s,\forall g,h\in F,g\neq h\},\\ S_{2}&=\{s\in A:\varphi(gh)s=\varphi(g)\varphi(h)s,\forall g,h\in F\},and\\ S&=S_{1}\cap S_{2}.\end{split}
Proof.

Assume WLOG that FGF\subseteq G is a symmetric finite subset of GG containing the identity. Let F=FF={gh:g,hF}F^{\prime}=F\cdot F=\{gh:g,h\in F\} and ϵ=ϵ4|F|2\epsilon^{\prime}=\frac{\epsilon}{4|F|^{2}}.

Since GG is sofic, there exists a finite set AA and a map φ:GSym(A)\varphi:G\rightarrow\textrm{Sym}(A) which is unital and (F,ϵ)(F^{\prime},\epsilon^{\prime})-multiplicative and satisfies d(1,φ(g))>1ϵd(1,\varphi(g))>1-\epsilon^{\prime} for all non-identity gFg\in F^{\prime}.

Fix g,hFg,h\in F, ghg\neq h, then, since FF is symmetric and since φ\varphi is unital and (F,ϵ)(F^{\prime},\epsilon^{\prime})-multiplicative, we have,

d(φ(g)1,φ(g1))=d(φ(g)φ(g)1,φ(g)φ(g1))=d(1,φ(g)φ(g1))=d(φ(gg1),φ(g)φ(g1))<ϵ.\begin{split}d(\varphi(g)^{-1},\varphi(g^{-1}))&=d(\varphi(g)\varphi(g)^{-1},\varphi(g)\varphi(g^{-1}))\\ &=d(1,\varphi(g)\varphi(g^{-1}))\\ &=d(\varphi(gg^{-1}),\varphi(g)\varphi(g^{-1}))\\ &<\epsilon^{\prime}.\end{split}

Thus, since g1hFg^{-1}h\in F^{\prime} and is not the identity,

d(1,φ(g)1φ(h))d(1,φ(g1h))d(φ(g1h),φ(g1)φ(h))d(φ(g1)φ(h),φ(g)1φ(h))>1ϵϵd(φ(g1),φ(g)1)>13ϵ.\begin{split}d(1,\varphi(g)^{-1}\varphi(h))&\geq d(1,\varphi(g^{-1}h))-d(\varphi(g^{-1}h),\varphi(g^{-1})\varphi(h))-d(\varphi(g^{-1})\varphi(h),\varphi(g)^{-1}\varphi(h))\\ &>1-\epsilon^{\prime}-\epsilon^{\prime}-d(\varphi(g^{-1}),\varphi(g)^{-1})\\ &>1-3\epsilon^{\prime}.\end{split}

Hence, |{sA:φ(g)sφ(h)s}|=|{sA:sφ(g)1φ(h)s}|>(13ϵ)|A||\{s\in A:\varphi(g)s\neq\varphi(h)s\}|=|\{s\in A:s\neq\varphi(g)^{-1}\varphi(h)s\}|>(1-3\epsilon^{\prime})|A|. So,

|S1|=|g,hFgh{sA:φ(g)sφ(h)s}|>(13|F|2ϵ)|A||S_{1}|=|\bigcap_{\begin{smallmatrix}g,h\in F\\ g\neq h\end{smallmatrix}}\{s\in A:\varphi(g)s\neq\varphi(h)s\}|>(1-3|F|^{2}\epsilon^{\prime})|A|

We also observe that, for any fixed g,hFg,h\in F, |{sA:φ(gh)s=φ(g)φ(h)s}|>(1ϵ)|A||\{s\in A:\varphi(gh)s=\varphi(g)\varphi(h)s\}|>(1-\epsilon^{\prime})|A|. Therefore,

|S2|=|g,hF{sA:φ(gh)s=φ(g)φ(h)s}|>(1|F|2ϵ)|A||S_{2}|=|\bigcap_{g,h\in F}\{s\in A:\varphi(gh)s=\varphi(g)\varphi(h)s\}|>(1-|F|^{2}\epsilon^{\prime})|A|

As such, |S|=|S1S2|>(14|F|2ϵ)|A|=(1ϵ)|A||S|=|S_{1}\cap S_{2}|>(1-4|F|^{2}\epsilon^{\prime})|A|=(1-\epsilon)|A| as desired. ∎

Theorem 2.14.

Let GG be a sofic group, NGN\leq G be a locally finite subgroup of GG. Then the left multiplication action α:GG/N\alpha:G\curvearrowright G/N is sofic.

Proof.

Let FGF\subseteq G, EG/NE\subseteq G/N be finite subsets. Fix ϵ>0\epsilon>0. Let q:GG/Nq:G\rightarrow G/N be the natural quotient map. Fix σ:G/NG\sigma:G/N\rightarrow G an arbitrary section of qq. Let U={σ(x)1gσ(g1x):xE,gF}U=\{\sigma(x)^{-1}g\sigma(g^{-1}x):x\in E,g\in F\}. We observe that UU is finite and UNU\subseteq N. Let N=UN^{\prime}=\langle U\rangle. Since NN is locally finite, NN^{\prime} is a finite group. Let,

F=FN(Nσ(E)1)F^{\prime}=F\cup N^{\prime}\cup(N^{\prime}\cdot\sigma(E)^{-1})

We observe that since NN^{\prime} is a subgroup, 1GN1_{G}\in N^{\prime}, so σ(E)1Nσ(E)1F\sigma(E)^{-1}\subseteq N^{\prime}\cdot\sigma(E)^{-1}\subseteq F^{\prime}. Let ϵ=ϵ|E|+1\epsilon^{\prime}=\frac{\epsilon}{|E|+1}. Since FF^{\prime} is finite, by Lemma 2.13, there is a finite set AA and a unital, (F,ϵ)(F^{\prime},\epsilon^{\prime})-multiplicative map φ:GSym(A)\varphi:G\rightarrow\textrm{Sym}(A) such that |S|>(1ϵ)|A||S^{\prime}|>(1-\epsilon^{\prime})|A|, where we define,

S1={sA:φ(g)sφ(h)s,g,hF,gh},S2={sA:φ(gh)s=φ(g)φ(h)s,g,hF},andS=S1S2.\begin{split}S_{1}&=\{s\in A:\varphi(g)s\neq\varphi(h)s,\forall g,h\in F^{\prime},g\neq h\},\\ S_{2}&=\{s\in A:\varphi(gh)s=\varphi(g)\varphi(h)s,\forall g,h\in F^{\prime}\},and\\ S^{\prime}&=S_{1}\cap S_{2}.\end{split}

Now, on the set SS^{\prime}, we define a relation s1s2s_{1}\sim s_{2} if there exists nNn\in N^{\prime} s.t. s1=φ(n)s2s_{1}=\varphi(n)s_{2}. Since φ\varphi is unital, \sim is reflexive. As NN^{\prime} is a group, NFN^{\prime}\subseteq F^{\prime}, and by the definition of S2S_{2}, we see that \sim is symmetric and transitive. Hence, \sim is an equivalence relation. Let B=S/B=S^{\prime}/\sim and,

S=SxEφ(σ(x)1)1SS=S^{\prime}\cap\bigcap_{x\in E}\varphi(\sigma(x)^{-1})^{-1}S^{\prime}

We observe that as |S|>(1ϵ)|A||S^{\prime}|>(1-\epsilon^{\prime})|A| and ϵ=ϵ|E|+1\epsilon^{\prime}=\frac{\epsilon}{|E|+1}, we have |S|>(1ϵ)|A||S|>(1-\epsilon)|A|. Now, for sSs\in S, we define πs:EB\pi_{s}:E\hookrightarrow B to be πs(x)=[φ(σ(x)1)s]\pi_{s}(x)=[\varphi(\sigma(x)^{-1})s]_{\sim}. We observe that this is injective. Indeed, assume πs(x)=πs(y)\pi_{s}(x)=\pi_{s}(y), i.e., φ(σ(x)1)sφ(σ(y)1)s\varphi(\sigma(x)^{-1})s\sim\varphi(\sigma(y)^{-1})s. Then there exists nNn\in N^{\prime} s.t. φ(σ(x)1)s=φ(n)φ(σ(y)1)s\varphi(\sigma(x)^{-1})s=\varphi(n)\varphi(\sigma(y)^{-1})s. Since σ(E)1F\sigma(E)^{-1}\subseteq F^{\prime} and NFN^{\prime}\subseteq F^{\prime}, by the definition of S2S_{2} we have φ(σ(x)1)s=φ(n)φ(σ(y)1)s=φ(nσ(y)1)s\varphi(\sigma(x)^{-1})s=\varphi(n)\varphi(\sigma(y)^{-1})s=\varphi(n\sigma(y)^{-1})s. But then, as nσ(y)1Nσ(E)1Fn\sigma(y)^{-1}\in N^{\prime}\cdot\sigma(E)^{-1}\subseteq F^{\prime} and σ(x)1σ(E)1F\sigma(x)^{-1}\in\sigma(E)^{-1}\subseteq F^{\prime}, we have, by the definition of S1S_{1}, that σ(x)1=nσ(y)1\sigma(x)^{-1}=n\sigma(y)^{-1}. But then σ(x)=σ(y)n1σ(y)Nσ(y)N=y\sigma(x)=\sigma(y)n^{-1}\in\sigma(y)N^{\prime}\subseteq\sigma(y)N=y, i.e., x=yx=y. This shows that πs\pi_{s} must be injective.

Finally, for all sSs\in S, gFg\in F, xEx\in E, if φ(g)sS\varphi(g)s\in S and α(g1)x=g1xE\alpha(g^{-1})x=g^{-1}x\in E, then as σ(E)1F\sigma(E)^{-1}\subseteq F^{\prime} and FFF\subseteq F^{\prime}, and by the definition of S2S_{2},

πφ(g)s(x)=[φ(σ(x)1)φ(g)s]=[φ(σ(x)1g)s]\pi_{\varphi(g)s}(x)=[\varphi(\sigma(x)^{-1})\varphi(g)s]_{\sim}=[\varphi(\sigma(x)^{-1}g)s]_{\sim}

and,

πs(g1x)=[φ(σ(g1x)1)s]\pi_{s}(g^{-1}x)=[\varphi(\sigma(g^{-1}x)^{-1})s]_{\sim}

It now suffices to prove φ(σ(x)1g)sφ(σ(g1x)1)s\varphi(\sigma(x)^{-1}g)s\sim\varphi(\sigma(g^{-1}x)^{-1})s. Let n=σ(x)1gσ(g1x)n=\sigma(x)^{-1}g\sigma(g^{-1}x). By definition we have nUNn\in U\subseteq N^{\prime}. Since g1xEg^{-1}x\in E, as σ(E)1F\sigma(E)^{-1}\subseteq F^{\prime} and NFN^{\prime}\subseteq F^{\prime}, we have, by definition of S2S_{2},

φ(n)φ(σ(g1x)1)s=φ(nσ(g1x)1)s=φ(σ(x)1g)s\varphi(n)\varphi(\sigma(g^{-1}x)^{-1})s=\varphi(n\sigma(g^{-1}x)^{-1})s=\varphi(\sigma(x)^{-1}g)s

This concludes the proof. ∎

We record here some easy observations:

Proposition 2.15.
  1. (1)

    If α:GX\alpha:G\curvearrowright X is the composition of a quotient map q:GHq:G\rightarrow H and a sofic action β:HX\beta:H\curvearrowright X, then α\alpha is sofic;

  2. (2)

    If α:GX\alpha:G\curvearrowright X is sofic, then the restriction of α\alpha to each of its orbits is sofic;

  3. (3)

    If α:GX\alpha:G\curvearrowright X is sofic and HH is a subgroup of GG, then α|H\alpha|_{H} is sofic;

  4. (4)

    If G1G2G_{1}\subseteq G_{2}\subseteq\cdots is an increasing sequence of subgroups of GG whose union is GG, and α:GX\alpha:G\curvearrowright X restricted to each GiG_{i} is sofic, then α\alpha is sofic.

The converse of item 2 of the above proposition is also true, which naturally reduces the study of sofic actions to transitive ones:

Proposition 2.16.

If the restriction of α:GX\alpha:G\curvearrowright X to each of its orbits is sofic, then α\alpha is sofic.

Proof.

Fix finite subsets FGF\subseteq G, EXE\subseteq X, and ϵ>0\epsilon>0. We need to show there exists a unital, (F,ϵ)(F,\epsilon)-multiplicative, (F,E,ϵ)(F,E,\epsilon)-orbit approximation of α\alpha. Since EE is finite, it only intersects with finitely many orbits of α\alpha, which we shall denote by X1,,XnX_{1},\cdots,X_{n}. Let Ei=EXiE_{i}=E\cap X_{i}. Choose ϵ>0\epsilon^{\prime}>0 s.t. (1ϵ)n1ϵ(1-\epsilon^{\prime})^{n}\geq 1-\epsilon. Since α|Xi\alpha|_{X_{i}} is sofic, there exists φi:GSym(Ai)\varphi_{i}:G\rightarrow\textrm{Sym}(A_{i}) which is unital, (F,ϵ)(F,\epsilon^{\prime})-multiplicative, and an (F,Ei,ϵ)(F,E_{i},\epsilon^{\prime})-orbit approximation of α|Xi\alpha|_{X_{i}}. Define φ:GSym(A1××An)\varphi:G\rightarrow\textrm{Sym}(A_{1}\times\cdots\times A_{n}) by,

φ(g)(a1,,an)=(φ1(g)a1,φn(g)an)\varphi(g)(a_{1},\cdots,a_{n})=(\varphi_{1}(g)a_{1},\cdots\varphi_{n}(g)a_{n})

It is clear that φ\varphi is unital. We claim that φ\varphi is (F,ϵ)(F,\epsilon)-multiplicative. Indeed, for g,hFg,h\in F,

|A1××An|(1d(φ(gh),φ(g)φ(h)))=i=1n|{aAi:φi(gh)a=φi(g)φi(h)a}|=i=1n|Ai|(1d(φi(gh),φi(g)φi(h)))>|A1××An|(1ϵ)n|A1××An|(1ϵ)\begin{split}|A_{1}\times\cdots\times A_{n}|\cdot(1-d(\varphi(gh),\varphi(g)\varphi(h)))&=\prod_{i=1}^{n}|\{a\in A_{i}:\varphi_{i}(gh)a=\varphi_{i}(g)\varphi_{i}(h)a\}|\\ &=\prod_{i=1}^{n}|A_{i}|\cdot(1-d(\varphi_{i}(gh),\varphi_{i}(g)\varphi_{i}(h)))\\ &>|A_{1}\times\cdots\times A_{n}|\cdot(1-\epsilon^{\prime})^{n}\\ &\geq|A_{1}\times\cdots\times A_{n}|\cdot(1-\epsilon)\end{split}

so d(φ(gh),φ(g)φ(h))<ϵd(\varphi(gh),\varphi(g)\varphi(h))<\epsilon.

We now show φ\varphi is an (F,E,ϵ)(F,E,\epsilon)-orbit approximation of α\alpha to conclude the proof. We note that, as φi\varphi_{i} is an (F,Ei,ϵ)(F,E_{i},\epsilon^{\prime})-orbit approximation of α|Xi\alpha|_{X_{i}}, there exists a finite set BiB_{i}, SiAiS_{i}\subseteq A_{i} and, for each sSis\in S_{i}, πsi:EiBi\pi_{s}^{i}:E_{i}\hookrightarrow B_{i} satisfying the condition for an (F,Ei,ϵ)(F,E_{i},\epsilon^{\prime})-orbit approximation of α|Xi\alpha|_{X_{i}}. Let S=S1××SnS=S_{1}\times\cdots\times S_{n}. Then |S|>(1ϵ)n|A1××An|(1ϵ)|A1××An||S|>(1-\epsilon^{\prime})^{n}|A_{1}\times\cdots\times A_{n}|\geq(1-\epsilon)|A_{1}\times\cdots\times A_{n}|. Let B=i=1nBiB=\coprod_{i=1}^{n}B_{i}. For each s=(s1,,sn)Ss=(s_{1},\cdots,s_{n})\in S, define πs:E=i=1nEiB\pi_{s}:E=\coprod_{i=1}^{n}E_{i}\hookrightarrow B by,

πs(x)=πsii(x),when xEi\pi_{s}(x)=\pi_{s_{i}}^{i}(x),\textrm{when }x\in E_{i}

It is easy to see that the map is indeed injective. Finally, fix s=(s1,,sn)Ss=(s_{1},\cdots,s_{n})\in S, gFg\in F, xEiEx\in E_{i}\subseteq E. Assume φ(g)sS\varphi(g)s\in S and α(g1)xE\alpha(g^{-1})x\in E. Then α|Xi(g1)x=α(g1)xEi\alpha|_{X_{i}}(g^{-1})x=\alpha(g^{-1})x\in E_{i}. Also, φ(g)s=(φ1(g)s1,,φn(g)sn)S=S1××Sn\varphi(g)s=(\varphi_{1}(g)s_{1},\cdots,\varphi_{n}(g)s_{n})\in S=S_{1}\times\cdots\times S_{n} implies φi(g)siSi\varphi_{i}(g)s_{i}\in S_{i}, so,

πφ(g)s(x)=πφi(g)sii(x)=πsii(α|Xi(g1)x)=πsii(α(g1)x)=πs(α(g1)x)\pi_{\varphi(g)s}(x)=\pi_{\varphi_{i}(g)s_{i}}^{i}(x)=\pi_{s_{i}}^{i}(\alpha|_{X_{i}}(g^{-1})x)=\pi_{s_{i}}^{i}(\alpha(g^{-1})x)=\pi_{s}(\alpha(g^{-1})x)

This concludes the proof. ∎

Combining Theorem 2.14 and Proposition 2.16, we see that all actions by sofic groups with locally finite stabilizers are sofic. We are unable to settle the more general case of amenable stabilizers.

It is still open whether all actions by sofic groups are sofic. However, this does hold for amenable groups and free groups.

Theorem 2.17.

Any action α:GX\alpha:G\curvearrowright X where GG is an amenable group is sofic.

Proof.

Fix finite subsets FGF\subseteq G, EXE\subseteq X, and ϵ>0\epsilon>0. Assume WLOG that FF is symmetric and contains the identity. Let F=FFF^{\prime}=F\cdot F, ϵ=ϵ2|F|\epsilon^{\prime}=\frac{\epsilon}{2|F|}. As GG is amenable, we may then choose a Følner set AGA\subseteq G with |AgA|<ϵ|A||A\vartriangle gA|<\epsilon^{\prime}|A| for all gFg\in F^{\prime}. For gGg\in G, we choose φ(g)Sym(A)\varphi(g)\in\textrm{Sym}(A) to be any element of Sym(A)\textrm{Sym}(A) s.t. φ(g)a=ga\varphi(g)a=ga whenever gaAga\in A. It is clear that φ:GSym(A)\varphi:G\rightarrow\textrm{Sym}(A) is unital. We claim it is (F,ϵ)(F,\epsilon)-multiplicative. Indeed, let g,hFg,h\in F. For all aAa\in A with ha,ghaAha,gha\in A, by definition we have φ(g)φ(h)a=φ(gh)a\varphi(g)\varphi(h)a=\varphi(gh)a. So, as h1,h1g1FFh^{-1},h^{-1}g^{-1}\in F\cdot F,

|A|d(φ(g)φ(h),φ(gh))|{aA:haA}|+|{aA:ghaA}||Ah1A|+|Ah1g1A|<2ϵ|A|ϵ|A|\begin{split}|A|\cdot d(\varphi(g)\varphi(h),\varphi(gh))&\leq|\{a\in A:ha\notin A\}|+|\{a\in A:gha\notin A\}|\\ &\leq|A\vartriangle h^{-1}A|+|A\vartriangle h^{-1}g^{-1}A|\\ &<2\epsilon^{\prime}|A|\\ &\leq\epsilon|A|\end{split}

That is, d(φ(g)φ(h),φ(gh))<ϵd(\varphi(g)\varphi(h),\varphi(gh))<\epsilon.

We now show φ\varphi is an (F,E,ϵ)(F,E,\epsilon)-orbit approximation of α\alpha to conclude the proof. To this end, let S={sA:gsA,gF}=gF(AgA)S=\{s\in A:gs\in A,\forall g\in F\}=\cap_{g\in F}(A\cap gA), as FF is symmetric. Since |AgA|>(1ϵ)|A||A\cap gA|>(1-\epsilon^{\prime})|A|, we have |S|>(1|F|ϵ)|A|>(1ϵ)|A||S|>(1-|F|\epsilon^{\prime})|A|>(1-\epsilon)|A|. Now, let B=α(A1)EB=\alpha(A^{-1})\cdot E. Define πs:EB\pi_{s}:E\hookrightarrow B by πs(x)=α(s1)x\pi_{s}(x)=\alpha(s^{-1})x. It is clear these maps are injective. We also have, for all sSs\in S, gFg\in F, xEx\in E s.t. φ(g)sS\varphi(g)s\in S and α(g1)xE\alpha(g^{-1})x\in E,

πφ(g)s(x)=πgs(x)=α(s1g1)x=α(s1)α(g1)x=πs(α(g1)x)\pi_{\varphi(g)s}(x)=\pi_{gs}(x)=\alpha(s^{-1}g^{-1})x=\alpha(s^{-1})\alpha(g^{-1})x=\pi_{s}(\alpha(g^{-1})x)

where we have used the fact that, as gsAgs\in A, by the definition of φ(g)\varphi(g), φ(g)s=gs\varphi(g)s=gs. This concludes the proof. ∎

Remark 2.18.

Together with item 1 of Proposition 2.15, this implies the action of any group on a finite set is sofic.

Theorem 2.19.

Any action α:GX\alpha:G\curvearrowright X where GG is a free group is sofic.

Proof.

Fix finite subsets FGF\subseteq G, EXE\subseteq X, and ϵ>0\epsilon>0. Let the free generators of GG be {g1,g2,}\{g_{1},g_{2},\cdots\} (either a finite set or a sequence, depending on whether GG is finitely generated or not). Let F1={f1,,fn}F^{-1}=\{f_{1},\cdots,f_{n}\}. We write fk=gik,1ϵk,1gik,mkϵk,mkf_{k}=g_{i_{k,1}}^{\epsilon_{k,1}}\cdots g_{i_{k,m_{k}}}^{\epsilon_{k,m_{k}}}, where ϵ\epsilon’s are in {±1}\{\pm 1\}. Let BXB\subseteq X be a finite set containing EE and all elements of XX of the form α(gik,lϵk,lgik,mkϵk,mk)x\alpha(g_{i_{k,l}}^{\epsilon_{k,l}}\cdots g_{i_{k,m_{k}}}^{\epsilon_{k,m_{k}}})x for all xEx\in E, 1kn1\leq k\leq n, 1lmk1\leq l\leq m_{k}. We define a homomorphism ψ:GSym(B)\psi:G\rightarrow\textrm{Sym}(B) by defining, for a free generator gig_{i}, ψ(gi)\psi(g_{i}) to be any element of Sym(B)\textrm{Sym}(B) s.t. ψ(gi)b=α(gi)b\psi(g_{i})b=\alpha(g_{i})b whenever α(gi)bB\alpha(g_{i})b\in B, then extending it to a homomorphism from GG. Let A=Sym(B)A=\textrm{Sym}(B) regarded as a finite set. Then, the left multiplication action of Sym(B)\textrm{Sym}(B) on itself gives rise to an inclusion Sym(B)Sym(A)\textrm{Sym}(B)\hookrightarrow\textrm{Sym}(A). Let φ:GSym(A)\varphi:G\rightarrow\textrm{Sym}(A) be ψ\psi composed with this inclusion.

Since φ\varphi is a homomorphism, it is unital and (F,ϵ)(F,\epsilon)-multiplicative. So it suffices to show it is an (F,E,ϵ)(F,E,\epsilon)-orbit approximation of α\alpha. We let S=AS=A and for each sS=Sym(B)s\in S=\textrm{Sym}(B), we define πs(x)=s1x\pi_{s}(x)=s^{-1}x. It is clear this is an injective map from EE to BB. Now, for sSs\in S, gFg\in F, xEx\in E, we have πφ(g)s(x)=πψ(g)s(x)=s1ψ(g1)x\pi_{\varphi(g)s}(x)=\pi_{\psi(g)s}(x)=s^{-1}\psi(g^{-1})x. Here, we note that by the definition of ψ\psi and BB, it is easy to see that ψ(g1)x=α(g1)x\psi(g^{-1})x=\alpha(g^{-1})x. Thus, whenever α(g1)xE\alpha(g^{-1})x\in E, we have πφ(g)s(x)=s1ψ(g1)x=s1α(g1)x=πs(α(g1)x)\pi_{\varphi(g)s}(x)=s^{-1}\psi(g^{-1})x=s^{-1}\alpha(g^{-1})x=\pi_{s}(\alpha(g^{-1})x). This concludes the proof. ∎

3. Generalized wreath products

Definition 3.1.

Let G,HG,H be groups and α:HX\alpha:H\curvearrowright X an action on a set. The generalized wreath product GαHG\wr_{\alpha}H is the semidirect product GXβHG^{\oplus X}\rtimes_{\beta}H where β(h)((gx)xX)=(gα(h)1(x))xX\beta(h)((g_{x})_{x\in X})=(g_{\alpha(h)^{-1}(x)})_{x\in X}. We observe that the same construction can also applied to a tracial von Neumann algebra MM in place of GG and with direct sums replaced by tensor products.

Definition 3.2.

Let G,HG,H be groups, α:HX\alpha:H\curvearrowright X be an action on a set, and AGA\leq G be a subgroup. Then HH acts on the amalgamated free product AxXGx\ast^{x\in X}_{A}G_{x} where GxG_{x} are copies of GG by permuting GxG_{x} according to α\alpha. Denote this action by β\beta. Then the amalgamated free generalized wreath product, denoted by GαAHG\wr^{\ast_{A}}_{\alpha}H is given by GαAH=(AxXGx)βHG\wr^{\ast_{A}}_{\alpha}H=(\ast^{x\in X}_{A}G_{x})\rtimes_{\beta}H. In case A={1G}A=\{1_{G}\}, we shall call this group the free generalized wreath product and denote it by GαHG\wr^{\ast}_{\alpha}H. Note that the same construction can also be applied to an inclusion of tracial von Neumann algebras NMN\subseteq M in place of the inclusion of groups AGA\leq G.

The main applications in our present work in defining sofic actions are Theorem 3.6 and Theorem 3.7, which are heavily inspired by the work of [HS18]. First though, we need a few definitions and lemmas.

Definition 3.3.

Let GG be a countable discrete group, AA be a finite set. Let αA:Sym(A)A\alpha_{A}:\textrm{Sym}(A)\curvearrowright A be the natural action. We may then consider the generalized wreath product GαASym(A)G\wr_{\alpha_{A}}\textrm{Sym}(A), which is of the form GAβASym(A)G^{\oplus A}\rtimes_{\beta_{A}}\textrm{Sym}(A) where βA\beta_{A} is the action βA:Sym(A)GA\beta_{A}:\textrm{Sym}(A)\curvearrowright G^{\oplus A} induced by αA\alpha_{A}. We define a metric dG,Ad_{G,A} on this group given by

dG,A(g1σ1,g2σ2)=1|A||{aA:σ1(a)σ2(a) or g1(σ1(a))g2(σ2(a))}|d_{G,A}(g_{1}\sigma_{1},g_{2}\sigma_{2})=\frac{1}{|A|}|\{a\in A:\sigma_{1}(a)\neq\sigma_{2}(a)\textrm{ or }g_{1}(\sigma_{1}(a))\neq g_{2}(\sigma_{2}(a))\}|

where giGAg_{i}\in G^{\oplus A} and σiSym(A)\sigma_{i}\in\textrm{Sym}(A), with the former regarded as functions from AA to GG. It is easy to verify that this is indeed a metric and is invariant under both left and right multiplication.

Lemma 3.4.

Let GG be a sofic group, AA be a finite set. Then there exists a free ultrafilter 𝒰\mathcal{U} on \mathbb{N} and a sequence of finite sets FiF_{i}, s.t. (GAβASym(A),dG,A)(G^{\oplus A}\rtimes_{\beta_{A}}\textrm{Sym}(A),d_{G,A}) embeds into 𝒰(Sym(Fi),d)\prod_{\mathcal{U}}(\textrm{Sym}(F_{i}),d) isometrically.

Proof.

As GG is sofic, there exists a free ultrafilter 𝒰\mathcal{U} on \mathbb{N}, a sequence of finite sets EiE_{i}, and a group homomorphism ϕ:G𝒰(Sym(Ei),d)\phi:G\rightarrow\prod_{\mathcal{U}}(\textrm{Sym}(E_{i}),d) s.t. d𝒰(ϕ(g),ϕ(h))=1d_{\mathcal{U}}(\phi(g),\phi(h))=1 whenever ghg\neq h. We lift it to a sequence of maps ϕi:GSym(Ei)\phi_{i}:G\rightarrow\textrm{Sym}(E_{i}). Now, let Fi=Ei×AF_{i}=E_{i}\times A. We define πi:GAβASym(A)Sym(Fi)\pi_{i}:G^{\oplus A}\rtimes_{\beta_{A}}\textrm{Sym}(A)\rightarrow\textrm{Sym}(F_{i}) by,

πi(gσ)(x,a)=(ϕi(g[σ(a)])x,σ(a))\pi_{i}(g\sigma)(x,a)=(\phi_{i}(g[\sigma(a)])x,\sigma(a))

where gGAg\in G^{\oplus A}, σSym(A)\sigma\in\textrm{Sym}(A), xEix\in E_{i}, aAa\in A. Let π:GAβASym(A)𝒰(Sym(Fi),d)\pi:G^{\oplus A}\rtimes_{\beta_{A}}\textrm{Sym}(A)\rightarrow\prod_{\mathcal{U}}(\textrm{Sym}(F_{i}),d) be given by π(gσ)=(πi(gσ))𝒰\pi(g\sigma)=(\pi_{i}(g\sigma))_{\mathcal{U}}. We first verify that π\pi is a group homomorphism. Indeed, let g1,g2GAg_{1},g_{2}\in G^{\oplus A}, σ1,σ2Sym(A)\sigma_{1},\sigma_{2}\in\textrm{Sym}(A), then,

πi(g1σ1g2σ2)(x,a)=πi(g1βσ1(g2)σ1σ2)(x,a)=(ϕi(g1[σ1σ2(a)]βσ1(g2)[σ1σ2(a)])x,σ1σ2(a))=(ϕi(g1[σ1σ2(a)]g2[σ2(a)])x,σ1σ2(a))\begin{split}\pi_{i}(g_{1}\sigma_{1}g_{2}\sigma_{2})(x,a)&=\pi_{i}(g_{1}\beta_{\sigma_{1}}(g_{2})\sigma_{1}\sigma_{2})(x,a)\\ &=(\phi_{i}(g_{1}[\sigma_{1}\sigma_{2}(a)]\beta_{\sigma_{1}}(g_{2})[\sigma_{1}\sigma_{2}(a)])x,\sigma_{1}\sigma_{2}(a))\\ &=(\phi_{i}(g_{1}[\sigma_{1}\sigma_{2}(a)]g_{2}[\sigma_{2}(a)])x,\sigma_{1}\sigma_{2}(a))\end{split}

while,

πi(g1σ1)πi(g2σ2)(x,a)=πi(g1σ1)(ϕi(g2[σ2(a)])x,σ2(a))=(ϕi(g1[σ1σ2(a)])ϕi(g2[σ2(a)])x,σ1σ2(a))\begin{split}\pi_{i}(g_{1}\sigma_{1})\pi_{i}(g_{2}\sigma_{2})(x,a)&=\pi_{i}(g_{1}\sigma_{1})(\phi_{i}(g_{2}[\sigma_{2}(a)])x,\sigma_{2}(a))\\ &=(\phi_{i}(g_{1}[\sigma_{1}\sigma_{2}(a)])\phi_{i}(g_{2}[\sigma_{2}(a)])x,\sigma_{1}\sigma_{2}(a))\end{split}

so we have that,

d(πi(g1σ1g2σ2),πi(g1σ1)πi(g2σ2))=1|A|aAd(ϕi(g1[σ1σ2(a)]g2[σ2(a)]),ϕi(g1[σ1σ2(a)])ϕi(g2[σ2(a)]))d(\pi_{i}(g_{1}\sigma_{1}g_{2}\sigma_{2}),\pi_{i}(g_{1}\sigma_{1})\pi_{i}(g_{2}\sigma_{2}))=\frac{1}{|A|}\sum_{a\in A}d(\phi_{i}(g_{1}[\sigma_{1}\sigma_{2}(a)]g_{2}[\sigma_{2}(a)]),\phi_{i}(g_{1}[\sigma_{1}\sigma_{2}(a)])\phi_{i}(g_{2}[\sigma_{2}(a)]))

For any fixed aAa\in A, g1[σ1σ2(a)]g_{1}[\sigma_{1}\sigma_{2}(a)] is a fixed element of GG, and so is g2[σ2(a)]g_{2}[\sigma_{2}(a)], whence,

d(ϕi(g1[σ1σ2(a)]g2[σ2(a)]),ϕi(g1[σ1σ2(a)])ϕi(g2[σ2(a)]))0d(\phi_{i}(g_{1}[\sigma_{1}\sigma_{2}(a)]g_{2}[\sigma_{2}(a)]),\phi_{i}(g_{1}[\sigma_{1}\sigma_{2}(a)])\phi_{i}(g_{2}[\sigma_{2}(a)]))\rightarrow 0

as i𝒰i\rightarrow\mathcal{U}. Hence, d(πi(g1σ1g2σ2),πi(g1σ1)πi(g2σ2))0d(\pi_{i}(g_{1}\sigma_{1}g_{2}\sigma_{2}),\pi_{i}(g_{1}\sigma_{1})\pi_{i}(g_{2}\sigma_{2}))\rightarrow 0 as i𝒰i\rightarrow\mathcal{U}, i.e., π\pi is a group homomorphism.

We now show π\pi is isometric to conclude the proof. By definition of πi\pi_{i}, we see that,

d(πi(g1σ1),πi(g2σ2))=d(σ1,σ2)+1|A|aAσ1(a)=σ2(a)d(ϕi(g1[σ1(a)]),ϕi(g2[σ2(a)]))d(\pi_{i}(g_{1}\sigma_{1}),\pi_{i}(g_{2}\sigma_{2}))=d(\sigma_{1},\sigma_{2})+\frac{1}{|A|}\sum_{\begin{smallmatrix}a\in A\\ \sigma_{1}(a)=\sigma_{2}(a)\end{smallmatrix}}d(\phi_{i}(g_{1}[\sigma_{1}(a)]),\phi_{i}(g_{2}[\sigma_{2}(a)]))

For any fixed aAa\in A with σ1(a)=σ2(a)\sigma_{1}(a)=\sigma_{2}(a), g1[σ1(a)]g_{1}[\sigma_{1}(a)] is a fixed element of GG and so is g2[σ2(a)]g_{2}[\sigma_{2}(a)], hence d(ϕi(g1[σ1(a)]),ϕi(g2[σ2(a)]))=0d(\phi_{i}(g_{1}[\sigma_{1}(a)]),\phi_{i}(g_{2}[\sigma_{2}(a)]))=0 if g1[σ1(a)]=g2[σ2(a)]g_{1}[\sigma_{1}(a)]=g_{2}[\sigma_{2}(a)] and d(ϕi(g1[σ1(a)]),ϕi(g2[σ2(a)]))1d(\phi_{i}(g_{1}[\sigma_{1}(a)]),\phi_{i}(g_{2}[\sigma_{2}(a)]))\rightarrow 1 as i𝒰i\rightarrow\mathcal{U} otherwise. Hence, as i𝒰i\rightarrow\mathcal{U}, we have,

d(πi(g1σ1),πi(g2σ2))d(σ1,σ2)+1|A||{aA:σ1(a)=σ2(a) and g1[σ1(a)]g2[σ2(a)]}=dG,A(g1σ1,g2σ2)\begin{split}d(\pi_{i}(g_{1}\sigma_{1}),\pi_{i}(g_{2}\sigma_{2}))&\rightarrow d(\sigma_{1},\sigma_{2})+\frac{1}{|A|}|\{a\in A:\sigma_{1}(a)=\sigma_{2}(a)\textrm{ and }g_{1}[\sigma_{1}(a)]\neq g_{2}[\sigma_{2}(a)]\}\\ &=d_{G,A}(g_{1}\sigma_{1},g_{2}\sigma_{2})\end{split}

This proves the claim. ∎

As an immediate corollary, we get,

Corollary 3.5.

Let GG be a countable discrete group. Suppose there exists a free ultrafilter 𝒰\mathcal{U} on \mathbb{N}, a sequence of sofic groups GiG_{i}, and a sequence of finite sets AiA_{i} s.t. GG embeds into 𝒰(GiAiβAiSym(Ai),dGi,Ai)\prod_{\mathcal{U}}(G_{i}^{\oplus A_{i}}\rtimes_{\beta_{A_{i}}}\textrm{Sym}(A_{i}),d_{G_{i},A_{i}}), then GG is sofic.

Theorem 3.6.

Let G,HG,H be sofic groups, α:HX\alpha:H\curvearrowright X be a sofic action. Then the generalized wreath product GαHG\wr_{\alpha}H is sofic.

Proof.

Fix increasing sequences of finite subsets F1F2HF_{1}\subseteq F_{2}\subseteq\cdots\subseteq H and E1E2XE_{1}\subseteq E_{2}\subseteq\cdots\subseteq X s.t. iFi=H\cup_{i}F_{i}=H, iEi=X\cup_{i}E_{i}=X. Fix a decreasing sequence ϵi>0\epsilon_{i}>0 s.t. limiϵi=0\lim_{i}\epsilon_{i}=0. For each ii, let φi:HSym(Ai)\varphi_{i}:H\rightarrow\textrm{Sym}(A_{i}) be a unital, (Fi,ϵi)(F_{i},\epsilon_{i})-multiplicative, and an (Fi,Ei,ϵi)(F_{i},E_{i},\epsilon_{i})-orbit approximation of α\alpha. By definition of (Fi,Ei,ϵi)(F_{i},E_{i},\epsilon_{i})-orbit approximation, there exists a finite set BiB_{i} and a subset SiAiS_{i}\subseteq A_{i} s.t. |Si|>(1ϵi)|Ai||S_{i}|>(1-\epsilon_{i})|A_{i}| and for each sSis\in S_{i} there is an injective map πsi:EiBi\pi^{i}_{s}:E_{i}\hookrightarrow B_{i} s.t. πφi(g)si(x)=πsi(α(g1)x)\pi^{i}_{\varphi_{i}(g)s}(x)=\pi^{i}_{s}(\alpha(g^{-1})x) for all sSis\in S_{i}, gFig\in F_{i}, xEix\in E_{i}, whenever φi(g)sSi\varphi_{i}(g)s\in S_{i} and α(g1)xEi\alpha(g^{-1})x\in E_{i}. Let Gi=GBiG_{i}=G^{\oplus B_{i}}, which is sofic as GG is. Let pi:GXGEip_{i}:G^{\oplus X}\rightarrow G^{\oplus E_{i}} be the canonical projection map. Let qsi:GEiGBi=Giq^{i}_{s}:G^{\oplus E_{i}}\hookrightarrow G^{\oplus B_{i}}=G_{i} be the inclusion map induced by πsi:EiBi\pi^{i}_{s}:E_{i}\hookrightarrow B_{i}. Let Psi=qsipiP^{i}_{s}=q^{i}_{s}\circ p_{i}.

Let the action of HH on GXG^{\oplus X} via α\alpha be denoted by β\beta, i.e., GαH=GXβHG\wr_{\alpha}H=G^{\oplus X}\rtimes_{\beta}H. We define ρi:GαHGiAiβAiSym(Ai)\rho_{i}:G\wr_{\alpha}H\rightarrow G_{i}^{\oplus A_{i}}\rtimes_{\beta_{A_{i}}}\textrm{Sym}(A_{i}) as follows,

ρi(gh)=(sSiPsi(g)aAiSi1Gi)φi(h)\rho_{i}(gh)=(\oplus_{s\in S_{i}}P^{i}_{s}(g)\bigoplus\oplus_{a\in A_{i}\setminus S_{i}}1_{G_{i}})\cdot\varphi_{i}(h)

where gGXg\in G^{\oplus X} and hHh\in H. Let 𝒰\mathcal{U} be an arbitrary free ultrafilter on \mathbb{N}. Let ρ:GαH𝒰(GiAiβAiSym(Ai),dGi,Ai)\rho:G\wr_{\alpha}H\rightarrow\prod_{\mathcal{U}}(G_{i}^{\oplus A_{i}}\rtimes_{\beta_{A_{i}}}\textrm{Sym}(A_{i}),d_{G_{i},A_{i}}) be given by ρ(gh)=(ρi(gh))𝒰\rho(gh)=(\rho_{i}(gh))_{\mathcal{U}}.

We shall now prove that ρ\rho is a group homomorphism. Indeed, let g1,g2GXg_{1},g_{2}\in G^{\oplus X}, h1,h2Hh_{1},h_{2}\in H. The supports of g1g_{1} and g2g_{2} are finite, so for large enough ii, supp(g2)Ei\textrm{supp}(g_{2})\subseteq E_{i}, α(h1)supp(g2)Ei\alpha(h_{1})\cdot\textrm{supp}(g_{2})\subseteq E_{i}, and h1,h11,h2Fih_{1},h_{1}^{-1},h_{2}\in F_{i}. Furthermore, whenever h1,h11Fih_{1},h_{1}^{-1}\in F_{i}, as φi\varphi_{i} is (Fi,ϵi)(F_{i},\epsilon_{i})-multiplicative, we have,

d(φi(h11)1,φi(h1))=d(φi(h11)φi(h11)1,φi(h11)φi(h1))=d(1,φi(h11)φi(h1))=d(φi(h1h11),φi(h11)φi(h1))<ϵi\begin{split}d(\varphi_{i}(h_{1}^{-1})^{-1},\varphi_{i}(h_{1}))&=d(\varphi_{i}(h_{1}^{-1})\varphi_{i}(h_{1}^{-1})^{-1},\varphi_{i}(h_{1}^{-1})\varphi_{i}(h_{1}))\\ &=d(1,\varphi_{i}(h_{1}^{-1})\varphi_{i}(h_{1}))\\ &=d(\varphi_{i}(h_{1}h_{1}^{-1}),\varphi_{i}(h_{1}^{-1})\varphi_{i}(h_{1}))\\ &<\epsilon_{i}\end{split}

Let S~i={sSi:φi(h11)1s=φi(h1)s}\tilde{S}_{i}=\{s\in S_{i}:\varphi_{i}(h_{1}^{-1})^{-1}s=\varphi_{i}(h_{1})s\}. Then, for large enough ii, |S~i|>(12ϵi)|Ai||\tilde{S}_{i}|>(1-2\epsilon_{i})|A_{i}|. Now, for large enough ii,

ρi(g1h1g2h2)=ρi(g1βh1(g2)h1h2)=(sSiPsi(g1βh1(g2))aAiSi1Gi)φi(h1h2)=(sSiPsi(g1)Psi(βh1(g2))aAiSi1Gi)φi(h1h2)\begin{split}\rho_{i}(g_{1}h_{1}g_{2}h_{2})&=\rho_{i}(g_{1}\beta_{h_{1}}(g_{2})h_{1}h_{2})\\ &=(\oplus_{s\in S_{i}}P^{i}_{s}(g_{1}\beta_{h_{1}}(g_{2}))\bigoplus\oplus_{a\in A_{i}\setminus S_{i}}1_{G_{i}})\cdot\varphi_{i}(h_{1}h_{2})\\ &=(\oplus_{s\in S_{i}}P^{i}_{s}(g_{1})P^{i}_{s}(\beta_{h_{1}}(g_{2}))\bigoplus\oplus_{a\in A_{i}\setminus S_{i}}1_{G_{i}})\cdot\varphi_{i}(h_{1}h_{2})\end{split}

We observe that, when sS~iφi(h1)S~i=S~iφi(h11)1S~is\in\tilde{S}_{i}\cap\varphi_{i}(h_{1})\tilde{S}_{i}=\tilde{S}_{i}\cap\varphi_{i}(h_{1}^{-1})^{-1}\tilde{S}_{i}, Psi(βh1(g2))Gi=GBiP^{i}_{s}(\beta_{h_{1}}(g_{2}))\in G_{i}=G^{\oplus B_{i}}, regarded as a function from BiB_{i} to GG, is given by,

Psi(βh1(g2))(b)={g2(α(h11)x),if xEi and πsi(x)=b1G,otherwiseP^{i}_{s}(\beta_{h_{1}}(g_{2}))(b)=\begin{cases}g_{2}(\alpha(h_{1}^{-1})x),\textrm{if }x\in E_{i}\textrm{ and }\pi^{i}_{s}(x)=b\\ 1_{G},\textrm{otherwise}\end{cases}

We note that α(h11)xsupp(g2)\alpha(h_{1}^{-1})x\in\textrm{supp}(g_{2}) iff xα(h1)supp(g2)x\in\alpha(h_{1})\cdot\textrm{supp}(g_{2}). Since for large enough ii, α(h1)supp(g2)Ei\alpha(h_{1})\cdot\textrm{supp}(g_{2})\subseteq E_{i}, supp(g2)Ei\textrm{supp}(g_{2})\subseteq E_{i}, h11Fih_{1}^{-1}\in F_{i}, and assuming sS~iφi(h1)S~is\in\tilde{S}_{i}\cap\varphi_{i}(h_{1})\tilde{S}_{i}, we have,

Psi(βh1(g2))(b)={g2(α(h11)x),if xα(h1)supp(g2) and πsi(x)=b1G,otherwise={g2(x),if xsupp(g2) and πsi(α(h1)x)=b1G,otherwise={g2(x),if xsupp(g2) and πφi(h11)si(x)=b1G,otherwise={g2(x),if xEi and πφi(h11)si(x)=b1G,otherwise=Pφi(h11)si(g2)(b)\begin{split}P^{i}_{s}(\beta_{h_{1}}(g_{2}))(b)&=\begin{cases}g_{2}(\alpha(h_{1}^{-1})x),\textrm{if }x\in\alpha(h_{1})\cdot\textrm{supp}(g_{2})\textrm{ and }\pi^{i}_{s}(x)=b\\ 1_{G},\textrm{otherwise}\end{cases}\\ &=\begin{cases}g_{2}(x),\textrm{if }x\in\textrm{supp}(g_{2})\textrm{ and }\pi^{i}_{s}(\alpha(h_{1})x)=b\\ 1_{G},\textrm{otherwise}\end{cases}\\ &=\begin{cases}g_{2}(x),\textrm{if }x\in\textrm{supp}(g_{2})\textrm{ and }\pi^{i}_{\varphi_{i}(h_{1}^{-1})s}(x)=b\\ 1_{G},\textrm{otherwise}\end{cases}\\ &=\begin{cases}g_{2}(x),\textrm{if }x\in E_{i}\textrm{ and }\pi^{i}_{\varphi_{i}(h_{1}^{-1})s}(x)=b\\ 1_{G},\textrm{otherwise}\end{cases}\\ &=P^{i}_{\varphi_{i}(h_{1}^{-1})s}(g_{2})(b)\end{split}

So,

ρi(g1h1g2h2)=(sS~iφi(h1)S~iPsi(g1)Pφi(h11)si(g2)aAi(S~iφi(h1)S~i)κa)φi(h1h2)\rho_{i}(g_{1}h_{1}g_{2}h_{2})=(\oplus_{s\in\tilde{S}_{i}\cap\varphi_{i}(h_{1})\tilde{S}_{i}}P^{i}_{s}(g_{1})P^{i}_{\varphi_{i}(h_{1}^{-1})s}(g_{2})\bigoplus\oplus_{a\in A_{i}\setminus(\tilde{S}_{i}\cap\varphi_{i}(h_{1})\tilde{S}_{i})}\kappa_{a})\cdot\varphi_{i}(h_{1}h_{2})

for some κaGi\kappa_{a}\in G_{i}. On the other hand,

ρi(g1h1)ρi(g2h2)=(sSiPsi(g1)aAiSi1Gi)φi(h1)(sSiPsi(g2)aAiSi1Gi)φi(h2)=(sS~iφi(h1)S~iPsi(g1)Pφi(h1)1si(g2)aAi(S~iφi(h1)S~i)λa)φi(h1)φi(h2)\begin{split}\rho_{i}(g_{1}h_{1})\rho_{i}(g_{2}h_{2})&=(\oplus_{s\in S_{i}}P^{i}_{s}(g_{1})\bigoplus\oplus_{a\in A_{i}\setminus S_{i}}1_{G_{i}})\cdot\varphi_{i}(h_{1})\cdot(\oplus_{s\in S_{i}}P^{i}_{s}(g_{2})\bigoplus\oplus_{a\in A_{i}\setminus S_{i}}1_{G_{i}})\cdot\varphi_{i}(h_{2})\\ &=(\oplus_{s\in\tilde{S}_{i}\cap\varphi_{i}(h_{1})\tilde{S}_{i}}P^{i}_{s}(g_{1})P^{i}_{\varphi_{i}(h_{1})^{-1}s}(g_{2})\bigoplus\oplus_{a\in A_{i}\setminus(\tilde{S}_{i}\cap\varphi_{i}(h_{1})\tilde{S}_{i})}\lambda_{a})\cdot\varphi_{i}(h_{1})\varphi_{i}(h_{2})\end{split}

for some λaGi\lambda_{a}\in G_{i}. Thus, for large enough ii,

dGi,Ai(ρi(g1h1g2h2),ρi(g1h1)ρi(g2h2))d(φi(h1h2),φi(h1)φi(h2))+Ai(S~iφi(h1)S~i)|Ai|<ϵi+4ϵi=5ϵi0\begin{split}d_{G_{i},A_{i}}(\rho_{i}(g_{1}h_{1}g_{2}h_{2}),\rho_{i}(g_{1}h_{1})\rho_{i}(g_{2}h_{2}))&\leq d(\varphi_{i}(h_{1}h_{2}),\varphi_{i}(h_{1})\varphi_{i}(h_{2}))+\frac{A_{i}\setminus(\tilde{S}_{i}\cap\varphi_{i}(h_{1})\tilde{S}_{i})}{|A_{i}|}\\ &<\epsilon_{i}+4\epsilon_{i}\\ &=5\epsilon_{i}\\ &\rightarrow 0\end{split}

This proves that ρ\rho is a group homomorphism. Let N=ker(ρ)N=\textrm{ker}(\rho). By Corollary 3.5, (GαH)/N(G\wr_{\alpha}H)/N is sofic. Let ι:GαH(GαH)/N×H\iota:G\wr_{\alpha}H\rightarrow(G\wr_{\alpha}H)/N\times H be defined by ι(gh)=(ghN,h)\iota(gh)=(ghN,h) where gGXg\in G^{\oplus X} and hHh\in H. Since both (GαH)/N(G\wr_{\alpha}H)/N and HH are sofic, it now suffices to prove ι\iota is injective.

Clearly, ker(ι)GX\textrm{ker}(\iota)\subseteq G^{\oplus X}. Assume to the contrary that ker(ι)\textrm{ker}(\iota) is not trivial and let gker(ι){1}g\in\textrm{ker}(\iota)\setminus\{1\}. Then ρi(g)=(sSiPsi(g)aAiSi1Gi)\rho_{i}(g)=(\oplus_{s\in S_{i}}P^{i}_{s}(g)\bigoplus\oplus_{a\in A_{i}\setminus S_{i}}1_{G_{i}}). Since g1g\neq 1, there exists xXx\in X s.t. g(x)1Gg(x)\neq 1_{G}. For large enough ii, xEix\in E_{i}, and in such cases Psi(g)1GiP^{i}_{s}(g)\neq 1_{G_{i}} for all sSis\in S_{i}. Then we have,

dGi,Ai(ρi(g),1)=|Si||Ai|=1ϵi1d_{G_{i},A_{i}}(\rho_{i}(g),1)=\frac{|S_{i}|}{|A_{i}|}=1-\epsilon_{i}\rightarrow 1

In particular, ρ(g)1\rho(g)\neq 1, so gNg\notin N. But then ι(g)1\iota(g)\neq 1, a contradiction. This proves the claim. ∎

The following theorem can be proved following similar arguments as in the proof of Theorem 3.6. We include parts of a proof here for the convenience of the readers,

Theorem 3.7.

Let HH be a sofic group, α:HX\alpha:H\curvearrowright X be a sofic action, AGA\leq G be an inclusion of countable discrete groups s.t. the amalgamated free product of any countably many copies of GG over AA is sofic. Then the amalgamated free generalized wreath product GαAHG\wr^{\ast_{A}}_{\alpha}H is sofic. In particular, if G,HG,H are sofic groups and α:HX\alpha:H\curvearrowright X is a sofic action, then the free generalized wreath product GαHG\wr^{\ast}_{\alpha}H is sofic, and under the same conditions, if we in addition have an amenable subgroup AA of GG, then the amalgamated free generalized wreath product GαAHG\wr^{\ast_{A}}_{\alpha}H is sofic.

Proof outline.

Again, fix increasing sequences of finite subsets F1F2HF_{1}\subseteq F_{2}\subseteq\cdots\subseteq H and E1E2XE_{1}\subseteq E_{2}\subseteq\cdots\subseteq X s.t. iFi=H\cup_{i}F_{i}=H, iEi=X\cup_{i}E_{i}=X. Fix a decreasing sequence ϵi>0\epsilon_{i}>0 s.t. limiϵi=0\lim_{i}\epsilon_{i}=0. For each ii, let φi:HSym(Ai)\varphi_{i}:H\rightarrow\textrm{Sym}(A_{i}) be a unital, (Fi,ϵi)(F_{i},\epsilon_{i})-multiplicative, and an (Fi,Ei,ϵi)(F_{i},E_{i},\epsilon_{i})-orbit approximation of α\alpha. By definition of (Fi,Ei,ϵi)(F_{i},E_{i},\epsilon_{i})-orbit approximation, there exists a finite set BiB_{i} and a subset SiAiS_{i}\subseteq A_{i} s.t. |Si|>(1ϵi)|Ai||S_{i}|>(1-\epsilon_{i})|A_{i}| and for each sSis\in S_{i} there is an injective map πsi:EiBi\pi^{i}_{s}:E_{i}\hookrightarrow B_{i} s.t. πφi(g)si(x)=πsi(α(g1)x)\pi^{i}_{\varphi_{i}(g)s}(x)=\pi^{i}_{s}(\alpha(g^{-1})x) for all sSis\in S_{i}, gFig\in F_{i}, xEix\in E_{i}, whenever φi(g)sSi\varphi_{i}(g)s\in S_{i} and α(g1)xEi\alpha(g^{-1})x\in E_{i}. Let Gi=AbBiGbG_{i}=\ast^{b\in B_{i}}_{A}G_{b}, where GbG_{b} are copies of GG. Under the assumption that the amalgamated free product of any countably many copies of GG over AA is sofic, we see that GiG_{i} is sofic. Let pi:AxXGxAxEiGxp_{i}:\ast^{x\in X}_{A}G_{x}\rightarrow\ast^{x\in E_{i}}_{A}G_{x}, where GxG_{x} are copies of GG, be the map that is the identity map on AxEiGx\ast^{x\in E_{i}}_{A}G_{x} and sends everything else to 11. Let qsi:AxEiGxAbBiGb=Giq^{i}_{s}:\ast^{x\in E_{i}}_{A}G_{x}\hookrightarrow\ast^{b\in B_{i}}_{A}G_{b}=G_{i} be the inclusion map induced by πsi:EiBi\pi^{i}_{s}:E_{i}\hookrightarrow B_{i}. Let Psi=qsipiP^{i}_{s}=q^{i}_{s}\circ p_{i}.

Let the action of HH on AxXGx\ast^{x\in X}_{A}G_{x} via α\alpha be denoted by β\beta, i.e., GαAH=AxXGxβHG\wr^{\ast_{A}}_{\alpha}H=\ast^{x\in X}_{A}G_{x}\rtimes_{\beta}H. We define ρi:GαAHGiAiβAiSym(Ai)\rho_{i}:G\wr^{\ast_{A}}_{\alpha}H\rightarrow G_{i}^{\oplus A_{i}}\rtimes_{\beta_{A_{i}}}\textrm{Sym}(A_{i}) as follows,

ρi(gh)=(sSiPsi(g)aAiSi1Gi)φi(h)\rho_{i}(gh)=(\oplus_{s\in S_{i}}P^{i}_{s}(g)\bigoplus\oplus_{a\in A_{i}\setminus S_{i}}1_{G_{i}})\cdot\varphi_{i}(h)

where gAxXGxg\in\ast^{x\in X}_{A}G_{x} and hHh\in H. Let 𝒰\mathcal{U} be an arbitrary free ultrafilter on \mathbb{N}. Let ρ:GαH𝒰(GiAiβAiSym(Ai),dGi,Ai)\rho:G\wr_{\alpha}H\rightarrow\prod_{\mathcal{U}}(G_{i}^{\oplus A_{i}}\rtimes_{\beta_{A_{i}}}\textrm{Sym}(A_{i}),d_{G_{i},A_{i}}) be given by ρ(gh)=(ρi(gh))𝒰\rho(gh)=(\rho_{i}(gh))_{\mathcal{U}}. The remainder of the proof follows the same outline as the proof of Theorem 3.6 - the only additional fact we note here is that, as the support of any element in AxXGx\ast^{x\in X}_{A}G_{x} is finite, it is eventually contained in EiE_{i} for large enough ii, and so pip_{i}, and therefore Psi=qsipiP^{i}_{s}=q^{i}_{s}\circ p_{i}, is eventually multiplicative on any fixed finitely many elements of AxXGx\ast^{x\in X}_{A}G_{x}. ∎

A natural setting where the first line of the above applies is arbitrary free products of free groups amalgamated over any fixed subgroup (see [GJ21, GEMss]).

For the following results in the Hilbert-Schmidt setting, they can be proved using the same line of reasoning, where the map pip_{i} is replaced by the conditional expectation from M¯XM^{\bar{\otimes}X} (or NxXMx\ast^{x\in X}_{N}M_{x}) to M¯EiM^{\bar{\otimes}E_{i}} (or NxEiMx\ast^{x\in E_{i}}_{N}M_{x}, resp.) GiG_{i} shall now be replaced by Mi=M¯BiM_{i}=M^{\bar{\otimes}B_{i}} (or Mi=NbBiMbM_{i}=\ast^{b\in B_{i}}_{N}M_{b}, resp.) qsiq^{i}_{s} shall still be the natural inclusion map induced by πsi:EiBi\pi^{i}_{s}:E_{i}\hookrightarrow B_{i} and Psi=qsipiP^{i}_{s}=q^{i}_{s}\circ p_{i}. The map ρi\rho_{i} shall now be a map from MαHM\wr_{\alpha}H (or MαNHM\wr^{\ast_{N}}_{\alpha}H, resp.) to Mi¯Ai¯𝕄|Ai|()¯L(H)M_{i}^{\bar{\otimes}A_{i}}\bar{\otimes}\mathbb{M}_{|A_{i}|}(\mathbb{C})\bar{\otimes}L(H) given by,

ρi(mh)=[(sSiPsi(m)aAiSi0)φi(h)]λh\rho_{i}(mh)=[(\oplus_{s\in S_{i}}P^{i}_{s}(m)\bigoplus\oplus_{a\in A_{i}\setminus S_{i}}0)\cdot\varphi_{i}(h)]\otimes\lambda_{h}

where mM¯Xm\in M^{\bar{\otimes}X} (or NxXMx\ast^{x\in X}_{N}M_{x}, resp.) and hHh\in H, and where sSiPsi(m)aAiSi0\oplus_{s\in S_{i}}P^{i}_{s}(m)\bigoplus\oplus_{a\in A_{i}\setminus S_{i}}0 is understood as a diagonal matrix in Mi¯Ai¯𝕄|Ai|()M_{i}^{\bar{\otimes}A_{i}}\bar{\otimes}\mathbb{M}_{|A_{i}|}(\mathbb{C}), φi(h)\varphi_{i}(h) is now interpreted as a permutation matrix in 𝕄|Ai|()\mathbb{M}_{|A_{i}|}(\mathbb{C}), and λh\lambda_{h} is the unitary associated with hh in L(H)L(H). The remainder of the proof follows essentially the same outline, replacing the arguments verifying the conditions on the dGi,Aid_{G_{i},A_{i}} metric with the conditions on the preservation of the trace. (The final part of the argument, dealing with the kernel of ρ\rho, is no longer needed, as we included the λh\otimes\lambda_{h} term in the definition of ρi\rho_{i}.)

Theorem 3.8.

Let (M,τ)(M,\tau) be a Connes-embeddable tracial von Neumann algebra, HH be a hyperlinear group, α:HX\alpha:H\curvearrowright X be a sofic action. Then the generalized wreath product MαHM\wr_{\alpha}H is Connes-embeddable. In particular, if G,HG,H are hyperlinear groups and α:HX\alpha:H\curvearrowright X is a sofic action, then the generalized wreath product GαHG\wr_{\alpha}H is hyperlinear.

Theorem 3.9.

Let HH be a hyperlinear group, α:HX\alpha:H\curvearrowright X be a sofic action, NMN\subseteq M be an inclusion of tracial von Neumann algebras s.t. the amalgamated free product of any countably many copies of MM over NN is Connes-embeddable. Then the amalgamated free generalized wreath product MαNHM\wr^{\ast_{N}}_{\alpha}H is Connes-embeddable. In particular, if HH is a hyperlinear group, α:HX\alpha:H\curvearrowright X is a sofic action, and MM is a Connes-embeddable tracial von Neumann algebra, then the free generalized wreath product MαHM\wr^{\ast}_{\alpha}H is Connes-embeddable, and under the same conditions, if we in addition have an amenable subalgebra NN of MM, then the amalgamated free generalized wreath product MαNHM\wr^{\ast_{N}}_{\alpha}H is Connes-embeddable.

Corollary 3.10.

Let HH be a hyperlinear group, α:HX\alpha:H\curvearrowright X be a sofic action, AGA\leq G be an inclusion of countable discrete groups s.t. the amalgamated free product of any countably many copies of GG over AA is hyperlinear. Then the amalgamated free generalized wreath product GαAHG\wr^{\ast_{A}}_{\alpha}H is hyperlinear. In particular, if G,HG,H are hyperlinear groups and α:HX\alpha:H\curvearrowright X is a sofic action, then the free generalized wreath product GαHG\wr^{\ast}_{\alpha}H is hyperlinear, and under the same conditions, if we in addition have an amenable subgroup AA of GG, then the amalgamated free generalized wreath product GαAHG\wr^{\ast_{A}}_{\alpha}H is hyperlinear.

4. Concluding remarks and open questions

We document the following Proposition which places an aspect of our work in the context of Elek-Lippner’s work [EL10b] defining soficity for equivalence relations and of Paunescu [P1̆1]. We omit the proof because it is substantially similar to the arguments in the previous section. It is open whether the converse of the statement below holds.

Proposition 4.1.

Let (Ω,μ)(\Omega,\mu) be a standard probability space, GG be a sofic group, α:GX\alpha:G\curvearrowright X be a sofic action. Then the induced generalized Bernoulli shift G(ΩX,μ|X|)G\curvearrowright(\Omega^{X},\mu^{\otimes|X|}) is sofic in the sense of [P1̆1].

We ask the following two natural questions on the permanence of sofic actions which we are currently unable to answer:

Question 4.2.

Suppose we have actions αi:GiX\alpha_{i}:G_{i}\curvearrowright X which commute with each other and where ii ranges over a countable index set. Then the actions naturally give rise to an action α:iGiX\alpha:\oplus_{i}G_{i}\curvearrowright X. α\alpha is sofic iff all αi\alpha_{i} are sofic?

Question 4.3.

Suppose we have actions αi:GiX\alpha_{i}:G_{i}\curvearrowright X where ii ranges over a countable index set. Then the actions naturally give rise to an action α:iGiX\alpha:\ast_{i}G_{i}\curvearrowright X. α\alpha is sofic iff all αi\alpha_{i} are sofic?

The forward directions of both conjectures follow from item 3 of Proposition 2.15. By item 4 of Proposition 2.15, it suffices to consider the case where there are only two groups G1G_{1} and G2G_{2}. We document below one more natural question which is nothing but the converse of Theorem 3.6:

Question 4.4.

Let G,HG,H be nontrivial countable groups, α:HX\alpha:H\curvearrowright X be an action. Then the generalized wreath product GαHG\wr_{\alpha}H is sofic iff GG and HH are sofic and the action α\alpha is sofic?

A positive answer to Question 4.2 implies a positive answer to the above. Indeed, let Γ\Gamma be a sofic group. Then the left multiplication α1:ΓΓ\alpha_{1}:\Gamma\curvearrowright\Gamma is sofic by Theorem 2.14. The right multiplication action α2:ΓΓ\alpha_{2}:\Gamma\curvearrowright\Gamma, α2(γ)η=ηγ1\alpha_{2}(\gamma)\eta=\eta\gamma^{-1} is isomorphic to the left multiplication action, so as such is also sofic. α1\alpha_{1} and α2\alpha_{2} clearly commute, so the combined action α:ΓΓΓ\alpha:\Gamma\oplus\Gamma\curvearrowright\Gamma is sofic assuming a positive answer to Question 4.2. Let Δ:ΓΓΓ\Delta:\Gamma\rightarrow\Gamma\oplus\Gamma be the diagonal embedding, then by item 3 of Proposition 2.15, αΔ\alpha\circ\Delta is sofic. One can easily verify that αΔ\alpha\circ\Delta is the conjugation action of Γ\Gamma on itself. That is, the conjugation action of any sofic group on itself is sofic, assuming Question 4.2 has a positive answer.

Now, if GαHG\wr_{\alpha}H is sofic, then the above applies to it, so its conjugation action on itself is sofic. By item 3 of Proposition 2.15, we have the conjugation action β:HGαH\beta:H\curvearrowright G\wr_{\alpha}H is sofic. Now, as GG is nontrivial, we may fix gG{1G}g\in G\setminus\{1_{G}\}. For any xXx\in X, let gxGXGαHg_{x}\in G^{\oplus X}\subseteq G\wr_{\alpha}H be the element that, as a function from XX to GG, takes value 1G1_{G} at all elements of XX except xx and takes value gg at xx. Let the orbit of gxg_{x} under β\beta be 𝒪(gx)\mathcal{O}(g_{x}) and the orbit of xx under α\alpha be 𝒪(x)\mathcal{O}(x). Since β(h)gx=hgxh1=gα(h)x\beta(h)g_{x}=hg_{x}h^{-1}=g_{\alpha(h)x} and furthermore that gxgyg_{x}\neq g_{y} whenever xyx\neq y as g1Gg\neq 1_{G}, we easily see that the map 𝒪(x)ygy𝒪(gx)\mathcal{O}(x)\ni y\mapsto g_{y}\in\mathcal{O}(g_{x}) is a bijection that identifies β\beta restricted to 𝒪(gx)\mathcal{O}(g_{x}) and α\alpha restricted to 𝒪(x)\mathcal{O}(x). As β\beta is sofic, this implies, via item 2 of Proposition 2.15, that α\alpha restricted to 𝒪(x)\mathcal{O}(x) is sofic. Since xXx\in X is arbitrary, α\alpha restricted to each of its orbit is sofic, whence α\alpha is sofic by Proposition 2.16.

References

  • [Bow10] Lewis Phylip Bowen, A measure-conjugacy invariant for free group actions, Ann. of Math. (2) 171 (2010), no. 2, 1387–1400. MR 2630067
  • [Con76] A. Connes, Classification of injective factors. Cases II1,II_{1}, II,II_{\infty}, IIIλ,III_{\lambda}, λ1\lambda\not=1, Ann. of Math. (2) 104 (1976), no. 1, 73–115. MR 454659
  • [DKP14] Ken Dykema, David Kerr, and Mikaël Pichot, Sofic dimension for discrete measured groupoids, Trans. Amer. Math. Soc. 366 (2014), no. 2, 707–748. MR 3130315
  • [EL10a] G. Elek and G. Lippner, Sofic equivalence relations, J. Funct. Anal. 258 (2010), 1692–1708.
  • [EL10b] Gábor Elek and Gábor Lippner, Sofic equivalence relations, J. Funct. Anal. 258 (2010), no. 5, 1692–1708. MR 2566316
  • [ES04] Gábor Elek and Endre Szabó, Sofic groups and direct finiteness, J. Algebra 280 (2004), no. 2, 426–434. MR 2089244
  • [ES11] by same author, Sofic representations of amenable groups, Proc. Amer. Math. Soc. 139 (2011), no. 12, 4285–4291. MR 2823074
  • [GEMss] David Gao, Srivatsav Kunnawalkam Elayavalli, and Mahan Mj, Remarks on soficity of amalgamated free products, in progress.
  • [GJ21] David Gao and Marius Junge, Relative embeddability of von neumann algebras and amalgamated free products, 2021.
  • [HE23] Ben Hayes and Srivatsav Kunnawalkam Elayavalli, On sofic approximations of non amenable groups, 2023.
  • [HS18] Ben Hayes and Andrew W. Sale, Metric approximations of wreath products, Ann. Inst. Fourier (Grenoble) 68 (2018), no. 1, 423–455. MR 3795485
  • [KL11] David Kerr and Hanfeng Li, Entropy and the variational principle for actions of sofic groups, Invent. Math. 186 (2011), no. 3, 501–558. MR 2854085
  • [Lüc02] W. Lück, l2l^{2}-invariants: Theory and applications to geometry and kk-theory, Springer-Verlag, Berlin, 2002.
  • [Pau14] Liviu Paunescu, A convex structure on sofic embeddings, Ergodic Theory Dynam. System 34 (2014), no. 4, 1343–1352.
  • [Pes08] Vladimir G. Pestov, Hyperlinear and sofic groups: a brief guide, Bull. Symbolic Logic 14 (2008), no. 4, 449–480. MR 2460675
  • [Pop14] Sorin Popa, Independence properties in subalgebras of ultraproduct II1\rm II_{1} factors, J. Funct. Anal. 266 (2014), no. 9, 5818–5846. MR 3182961
  • [P1̆1] Liviu Păunescu, On sofic actions and equivalence relations, J. Funct. Anal. 261 (2011), no. 9, 2461–2485. MR 2826401
  • [Tho08] Andreas Thom, Sofic groups and Diophantine approximation, Comm. Pure Appl. Math. 61 (2008), no. 8, 1155–1171. MR 2417890