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

Extensibility and denseness of periodic semigroup actions

Raimundo Briceño Álvaro Bustos-Gajardo  and  Miguel Donoso-Echenique Facultad de Matemáticas, Pontificia Universidad Católica de Chile. Santiago, Chile {raimundo.briceno, abustog, miguel.donosoe}@uc.cl
Abstract.

We study periodic points and finitely supported invariant measures for continuous semigroup actions. Introducing suitable notions of periodicity in both topological and measure-theoretical contexts, we analyze the space of invariant Borel probability measures associated with these actions. For embeddable semigroups, we establish a direct relationship between the extensibility of invariant measures to the free group on the semigroup and the denseness of finitely supported invariant measures. Applying this framework to shift actions on the full shift, we prove that finitely supported invariant measures are dense for every left amenable semigroup that is residually a finite group and for every finite-rank free semigroup.

Key words and phrases:
Countable semigroup; embeddable semigroup; natural extension; free semigroup; periodic measure; Markov tree chain
2010 Mathematics Subject Classification:
Primary 22D40, 37A15, 20M05; Secondary 20M30, 37B10, 60J10.
R.B. was partially supported by ANID/FONDECYT Regular 1240508. A.B.-G. was supported by ANID/FONDECYT Postdoctorado 3230159 (year 2023). M.D. was partially supported by Beca de Magíster Nacional ANID 22230917 (year 2023).

Introduction

The study of periodic orbits is a fundamental aspect of the qualitative theory of dynamical systems. In particular, the question of whether periodic points are dense in a given system is relevant to topological and smooth dynamics [3, 11]. The measure-theoretical analogue of this problem is whether periodic measures are dense in the set of invariant measures, a topic that has been extensively addressed in classical ergodic theory [27, 29].

In the context of group actions, a concrete formulation of the first question is whether, for a countable group GG acting continuously on a compact metric space XX, the set of periodic points Per(X,G)\operatorname{Per}(X,G) is dense in XX. In this setting, the concept of periodicity must be adapted to the group framework, and the potential denseness of periodic points imposes algebraic constraints on GG, specifically requiring it to be residually finite. Moreover, denseness of periodic points characterizes residual finiteness in the following sense: a countable group GG is residually finite if and only if Per(X,G)\operatorname{Per}(X,G) is dense in 𝒜G\mathcal{A}^{G} for the shift action [8, Theorem 2.7.1].

More recently, interest has grown [7, 22, 26, 28] in the measure-theoretical analogue of this characterization: determining for which countable groups GG periodic measures are weak-* dense in the space of GG-invariant Borel probability measures G(𝒜G)\mathcal{M}_{G}(\mathcal{A}^{G}). This leads to a natural dichotomy among countable groups, distinguishing those for which periodic measures are weak-* dense from those for which they are not. The groups for which periodic measures are dense are said to have the pa property, and they are necessarily residually finite. However, not every residually finite group has the pa property.

We may further ask if this still holds true for the subset of ergodic periodic measures; the groups with this property will be said to have the epa property. Clearly, the epa property implies the pa property, but it is not known if the converse implication holds in general. It is known that all amenable, residually finite groups have the epa property (see [26], where denseness is established for more general systems with specification). On the other hand, Bowen showed that free groups on finitely many generators have the pa property [5]. However, the construction outlined in this work does not suffice to conclude that these groups have the epa property.

Notice that these properties are intrinsic to the group, as the action is prescribed once the group is set. This is not the only case where the structure of the space of invariant measures tell us something about the algebraic properties of the acting group (e.g., see [18]).

In this work, we aim to extend these questions and results to the more general setting of semigroup actions. To do so, we first need to examine the notion of periodicity, which is more subtle in the semigroup setting than in the group case, as pre-periodic behavior that is not periodic appears. With the appropriate definitions in place, our first result is as follows.

{restatable}

thmfirstproposition Let SS be a left reductive semigroup. Then:

  1. (i)

    SS is residually a finite semigroup if and only if the set of pre-periodic points of 𝒜S\mathcal{A}^{S} is dense in 𝒜S\mathcal{A}^{S} for every finite alphabet 𝒜\mathcal{A}.

  2. (ii)

    SS is residually a finite group if and only if the set of periodic points of 𝒜S\mathcal{A}^{S} is dense in 𝒜S\mathcal{A}^{S} for every finite alphabet 𝒜\mathcal{A}.

Next, we turn our attention to the measure-theoretical case and introduce the (e)pa property for semigroups, extending the notion developed for groups. To study this notion, we make use of the tool of natural extensions. Given a continuous action SXS\curvearrowright X of an embeddable semigroup SS and a receiving SS-group 𝐆\mathbf{G}, we consider its topological 𝐆\mathbf{G}-extension GX𝐆G\curvearrowright X_{\mathbf{G}}, as studied in [6]. Then, for the projection map π:X𝐆X\pi\colon X_{\mathbf{G}}\rightarrow X, we let π:G(X𝐆)S(X)\pi_{*}\colon\mathcal{M}_{G}(X_{\mathbf{G}})\rightarrow\mathcal{M}_{S}(X) be the push-forward map and define the set of 𝐆\mathbf{G}-extensible measures as Ext𝚪(X,S)=im(π)\operatorname{Ext}_{\bm{\Gamma}}(X,S)=\operatorname{im}(\pi_{*}). With this in place, we establish the following connection between the set of 𝐆\mathbf{G}-extensible measures and the weak-* closure of the set of SS-periodic measures (resp. ergodic SS-periodic measures).

{restatable}

thmsecondtheorem Let SS be an embeddable semigroup, and let 𝚪=(Γ,γ)\bm{\Gamma}=(\Gamma,\gamma) be a realization of the free SS-group. If SXS\curvearrowright X is a continuous action and 𝒫(X𝚪,Γ)\mathcal{P}(X_{\bm{\Gamma}},\Gamma) (resp. 𝒫erg(X𝚪,Γ)\mathcal{P}^{\mathrm{erg}}(X_{\bm{\Gamma}},\Gamma)) is weak-* dense in Γ(X𝚪)\mathcal{M}_{\Gamma}(X_{\bm{\Gamma}}), then

Ext𝚪(X,S)=𝒫(X,S)¯(resp. Ext𝚪(X,S)=𝒫erg(X,S)¯).\operatorname{Ext}_{\bm{\Gamma}}(X,S)=\overline{\mathcal{P}(X,S)}\quad\text{(resp. $\operatorname{Ext}_{\bm{\Gamma}}(X,S)=\overline{\mathcal{P}^{\mathrm{erg}}(X,S)}$)}.

In particular, if

  1. (i)

    Γ\Gamma has the (e)pa property, and

  2. (ii)

    Ext𝚪(𝒜S,S)=S(𝒜S)\textup{Ext}_{\bm{\Gamma}}(\mathcal{A}^{S},S)=\mathcal{M}_{S}(\mathcal{A}^{S}) for every finite alphabet 𝒜\mathcal{A},

then SS has the (e)pa property.

We then investigate the 𝚪\bm{\Gamma}-extensibility of measures for the free SS-group 𝚪\bm{\Gamma}, and use Theorem Introduction to establish the (e)pa property for some classes of semigroups. First, we consider the class of left amenable semigroups and prove the following result.

{restatable}

thmthirdtheorem Let SS be a left amenable semigroup that is residually a finite group. Then, SS has the epa property.

Next, we consider the free case, and obtain the following.

{restatable}

thmfourththeorem Let 𝔽d\mathbb{F}_{d} be the free group on generators {a1,,ad}\{a_{1},\dots,a_{d}\}. For Σ{a1±1,,ad±1}\Sigma\subseteq\{a_{1}^{\pm 1},\dots,a_{d}^{\pm 1}\}, consider the subsemigroup S=Σ+S=\left<\Sigma\right>^{+}, so that 𝐅d=(𝔽d,ι)\mathbf{F}_{d}=(\mathbb{F}_{d},\iota) is the free SS-group for the inclusion ι:S𝔽d\iota\colon S\to\mathbb{F}_{d}. Then, Ext𝐅d(𝒜S,S)=S(𝒜S)\textup{Ext}_{\mathbf{F}_{d}}(\mathcal{A}^{S},S)=\mathcal{M}_{S}(\mathcal{A}^{S}) and SS has the pa property.

In particular, Theorem Introduction allows us to conclude that the free semigroup 𝔽d+\mathbb{F}_{d}^{+} on dd generators has the pa property. Finally, we provide a characterization of the free SS-group for actions of certain subsemigroups SS of free groups of finite rank 𝔽d\mathbb{F}_{d}, which is reminiscent of the main result in [6], but in a measure-theoretical context.

{restatable}

thmfifththeorem Let dd be a positive integer and consider the free group 𝔽d\mathbb{F}_{d} on the generators {a1,,ad}\{a_{1},\dots,a_{d}\}. Given Σ{a1±1,,ad±1}\Sigma\subseteq\{a_{1}^{\pm 1},\dots,a_{d}^{\pm 1}\} such that Σ{a1,,ad}\Sigma\supseteq\{a_{1},\dots,a_{d}\}, consider the subsemigroup S=Σ+S=\left<\Sigma\right>^{+}, and let 𝐆\mathbf{G} be a receiving SS-group. The following are equivalent:

  1. (i)

    𝐆\mathbf{G} is a realization of the free SS-group, i.e., 𝐆𝐅d\mathbf{G}\simeq\mathbf{F}_{d}.

  2. (ii)

    Every SS-invariant measure in (𝒜S)\mathcal{M}(\mathcal{A}^{S}) is 𝐆\mathbf{G}-extensible.

  3. (iii)

    Every fully supported SS-invariant Markov measure in (𝒜S)\mathcal{M}(\mathcal{A}^{S}) is 𝐆\mathbf{G}-extensible.

The paper is organized as follows. In §1, we introduce the basic concepts of semigroup actions. In §2, we discuss the notion of periodicity for semigroup actions—in the topological and measure-theoretical settings—and its connection to residual finiteness, leading to the proof of Theorem Introduction. In §3, we consider natural extensions and explore the relationship between the (e)pa property for a semigroup SS and its free SS-group, establishing Theorem Introduction. In §4, we prove the (e)pa property for certain semigroups, as given in Theorem Introduction and Theorem Introduction, and conclude with a measure-theoretical characterization of the free SS-group for certain subsemigroups SS of free groups of finite rank 𝔽d\mathbb{F}_{d}, as given in Theorem Introduction.

1. Preliminaries

1.1. Semigroups and embeddings into groups

A semigroup is a set SS together with an associative binary operation (s,t)st(s,t)\mapsto st. A monoid is a semigroup which has a—necessarily unique—identity element 1S1_{S}. Throughout this work, we shall only deal with countable, discrete semigroups, with a special emphasis on monoids.

A subsemigroup of SS is a subset TST\subseteq S such that ttTtt^{\prime}\in T for all t,tTt,t^{\prime}\in T. In this scenario, we write TST\leq S. A common way to identify a subsemigroup TT is to take this set to be comprised of all possible products of elements of a subset KSK\subseteq S. More precisely, we define the subsemigroup generated by KK as follows:

K+={k0kn1:n,k0,,kn1K}S,\langle K\rangle^{+}=\{k_{0}\cdots k_{n-1}:n\in\mathbb{N},k_{0},\dots,k_{n-1}\in K\}\subseteq S,

where the empty product is the identity 1S1_{S} when SS is a monoid. We say that KK generates a subsemigroup TST\leq S if T=K+T=\langle K\rangle^{+}.

A semigroup homomorphism is a function θ:ST\theta\colon S\rightarrow T satisfying θ(st)=θ(s)θ(t)\theta(st)=\theta(s)\theta(t) for every s,tSs,t\in S. A semigroup homomorphism between monoids is called a monoid homomorphism if, additionally, θ(1S)=1T\theta(1_{S})=1_{T}. An embedding will be an injective homomorphism, and an isomorphism will be a bijective homomorphism. If there is an isomorphism STS\to T, we say SS and TT are isomorphic and denote this by STS\simeq T. In this case, SS and TT will be, in most cases, essentially the same for our purposes.

1.1.1. Free semigroups and presentations

We want to give a semigroup structure to the quotient of a semigroup by an equivalence relation. An equivalence relation \mathcal{R} on a semigroup SS is said to be a congruence if for every a,b,cSa,b,c\in S, aba\mathbin{\mathcal{R}}b implies both cacbca\mathbin{\mathcal{R}}cb and acbcac\mathbin{\mathcal{R}}bc. Given such a relation, we denote by π:SS/\pi_{\mathcal{R}}\colon S\rightarrow S/\mathcal{R} the quotient map π(s)=[s]\pi_{\mathcal{R}}(s)=[s]_{\mathcal{R}}.

If \mathcal{R} is a congruence on SS and aaa\mathbin{\mathcal{R}}a^{\prime}, bbb\mathbin{\mathcal{R}}b^{\prime}, transitivity implies that ababab\mathbin{\mathcal{R}}a^{\prime}b^{\prime}, and thus the binary operation S/×S/S/S/\mathcal{R}\times S/\mathcal{R}\longrightarrow S/\mathcal{R} given by [a][b]=[ab][a]_{\mathcal{R}}\cdot[b]_{\mathcal{R}}=[ab]_{\mathcal{R}} is well-defined and gives S/S/\mathcal{R} a semigroup structure, for which π\pi_{\mathcal{R}} is a semigroup homomorphism. For groups, the notions of quotient by a congruence and quotient by a normal subgroup are equivalent; this is easily seen by noting that there is a bijective correspondence associating to a congruence \mathcal{R} the normal subgroup N={gG:g1G}GN_{\mathcal{R}}=\{g\in G:g\mathrel{\mathcal{R}}1_{G}\}\trianglelefteq G and vice versa.

Given any set BB, the free semigroup generated by BB is the set 𝔽(B)+\mathbb{F}(B)^{+} of all finite words of elements from BB, with concatenation as its binary operation. There is a canonical inclusion ι:X𝔽(B)+\iota\colon X\hookrightarrow\mathbb{F}(B)^{+}, which comes with a universal property, analogous to that of free groups. If |B1|=|B2|\lvert B_{1}\rvert=\lvert B_{2}\rvert, then 𝔽(B1)+𝔽(B2)+\mathbb{F}(B_{1})^{+}\simeq\mathbb{F}(B_{2})^{+}. Thus, we may talk of the free semigroup generated by dd elements, 𝔽d+:=𝔽({a1,,ad})+\mathbb{F}_{d}^{+}:=\mathbb{F}(\{a_{1},\dotsc,a_{d}\})^{+} for an arbitrary, fixed set {a1,,ad}\{a_{1},\dotsc,a_{d}\}, and then 𝔽(B)+𝔽d+\mathbb{F}(B)^{+}\cong\mathbb{F}_{d}^{+} for any set BB of dd elements. The empty word ε\varepsilon will be the identity element of 𝔽(B)+\mathbb{F}(B)^{+}.

If S=B+S=\langle B\rangle^{+} for some BSB\subseteq S, then SS is isomorphic to a quotient of 𝔽(B)+\mathbb{F}(B)^{+}. This allows us to give a combinatorial description of a semigroup: given a set BB of generators and a collection R𝔽(B)+×𝔽(B)+R\subseteq\mathbb{F}(B)^{+}\times\mathbb{F}(B)^{+} of relations, the semigroup presentation BR+\langle B\mid R\rangle^{+} corresponds to the “largest” semigroup generated by BB for which all equations u=vu=v, for (u,v)R(u,v)\in R, hold true. This is formally the quotient 𝔽(B)+/\mathbb{F}(B)^{+}/\mathcal{R}, where \mathcal{R} is the smallest congruence in 𝔽(B)+\mathbb{F}(B)^{+} containing RR as a subset. Note that many different sets of relations may describe the same semigroup.

1.1.2. The free group on a semigroup

A pair 𝐆=(G,η)\mathbf{G}=(G,\eta) is an SS-group if GG is a group and η:SG\eta\colon S\rightarrow G is a semigroup morphism with η(S)=G\langle\eta(S)\rangle=G. If η\eta is an embedding, 𝐆\mathbf{G} is called a receiving SS-group, and whenever a semigroup admits an embedding into a group, we will say SS is embeddable. A morphism between two receiving SS-groups (G,η)(G,\eta) and (G,η)(G^{\prime},\eta^{\prime}) is a group morphism θ:GG\theta\colon G\rightarrow G^{\prime} such that θη=η\theta\circ\eta=\eta^{\prime}.

Definition 1.1.

Let SS be a semigroup. A free group on the semigroup SS, or a free SS-group, is an SS-group 𝚪=(Γ,γ)\bm{\Gamma}=(\Gamma,\gamma) such that for every SS-group (G,η)(G,\eta) there is a unique morphism θ:ΓG\theta\colon\Gamma\rightarrow G with θγ=η\theta\circ\gamma=\eta.

S{S}Γ{\Gamma}G{G}γ\scriptstyle{\gamma}η\scriptstyle{\eta}θ\scriptstyle{\theta}

Note that, if (Γ,γ)(\Gamma,\gamma) is the free SS-group and (G,η)(G,\eta) is any SS-group, the morphism θ:ΓG\theta\colon\Gamma\to G granted by the universal property of (Γ,γ)(\Gamma,\gamma) must be surjective, as its image contains a generating set for GG.

The free SS-group always exists and it is unique up to isomorphism of SS-groups (see [10, §12] for more details). Accordingly, we will speak of the free SS-group whenever we talk about properties that are stable under isomorphisms of SS-groups, and of a realization of the free SS-group when we want to refer to a specific pair (Γ,γ)(\Gamma,\gamma).

If we know a presentation BR+\langle B\mid R\rangle^{+} for SS, the group BR\langle B\mid R\rangle provides a concrete way of viewing the free SS-group. More precisely, if +\mathcal{R}^{+} is the congruence generated by RR in 𝔽(B)+\mathbb{F}(B)^{+} and \mathcal{R} is the congruence generated by RR in 𝔽(B)\mathbb{F}(B), let

S=𝔽(B)+/+andΓ=𝔽(B)/.S=\mathbb{F}(B)^{+}/\mathcal{R}^{+}\quad\text{and}\quad\Gamma=\mathbb{F}(B)/\mathcal{R}.

Then, if ι:SΓ\iota\colon S\rightarrow\Gamma is the semigroup morphism given by ι([b]+)=[b]\iota([b]_{\mathcal{R}^{+}})=[b]_{\mathcal{R}} for bBb\in B, extended homomorphically to all of SS, the pair (BR,ι)(\langle B\mid R\rangle,\iota) will be a realization of the free SS-group. In this case, the pair (BR,ι)(\langle B\mid R\rangle,\iota) will be referred to as the canonical realization of the free SS-group (see [6, Proposition 3.2]). For example, the following are the associated groups to canonical realizations of semigroups:

  1. (1)

    d=a1,adaiaj=ajai\mathbb{Z}^{d}=\langle a_{1},\dots a_{d}\mid a_{i}a_{j}=a_{j}a_{i}\rangle for d=a1,adaiaj=ajai+\mathbb{N}^{d}=\langle a_{1},\dots a_{d}\mid a_{i}a_{j}=a_{j}a_{i}\rangle^{+};

  2. (2)

    BS(m,n)=a,babm=bna\text{BS}(m,n)=\langle a,b\mid ab^{m}=b^{n}a\rangle for BS(m,n)+=a,babm=bna+\text{BS}(m,n)^{+}=\langle a,b\mid ab^{m}=b^{n}a\rangle^{+};

  3. (3)

    𝔽d=a1,,ad\mathbb{F}_{d}=\langle a_{1},\dotsc,a_{d}\mid\varnothing\rangle for 𝔽d+=a1,,ad+\mathbb{F}_{d}^{+}=\langle a_{1},\dotsc,a_{d}\mid\varnothing\rangle^{+}.

In these three instances, the semigroup morphism ι\iota associated to each canonical realization is the one that sends generators to generators in the natural way. Moreover, in every case, ι\iota is an embedding. In particular, all these cases correspond to embeddable semigroups. If BR+\langle B\mid R\rangle^{+} turns out to be embeddable into BR\langle B\mid R\rangle via ι\iota, we can always assume that BR+\langle B\mid R\rangle^{+} is a monoid by artifically adding an identity 1S1_{S} if necessary. In this case, defining ι(1S)=1Γ\iota(1_{S})=1_{\Gamma} yields an embedding as well. Thus, in any setting where embeddability plays a role, the semigroup BR+\langle B\mid R\rangle^{+} will be understood as a monoid. For instance, 𝔽d+\mathbb{F}_{d}^{+} will be understood as the free monoid on dd elements by adjoining the empty word ε\varepsilon to the free semigroup generated by dd elements, etc.

Embeddable semigroups are necessarily bicancellative, meaning that a=ba=b whenever ac=bcac=bc or ca=cbca=cb hold. Nonetheless, not every bicancellative semigroup can be embedded into a group [23]. In fact, given a semigroup presentation S=BR+S=\langle B\mid R\rangle^{+}, determining whether SS can be embedded into its free SS-group, or more generally into an arbitrary group, is not an easy matter. The relevant thing about the free SS-group is that it characterizes embeddability, meaning that a semigroup SS can be embedded into a group if and only if it can be embedded into its free SS-group [10, Theorem 12.4].

A useful result due to Adian [1] states that bicancellative one-relator finitely presented semigroups can be embedded into a group.

1.1.3. The left reversible case

A semigroup SS is said to be left reversible if for every s,tSs,t\in S there are x,ySx,y\in S such that sx=tysx=ty, or equivalently if for all s,tSs,t\in S, sStSsS\cap tS\neq\varnothing. Groups and Abelian semigroups are left reversible semigroups. In contrast, the free monoid 𝔽d+\mathbb{F}_{d}^{+} is not left reversible for d2d\geq 2, since a1𝔽d+ad𝔽d+=a_{1}\mathbb{F}_{d}^{+}\cap a_{d}\mathbb{F}_{d}^{+}=\varnothing.

In [25, Theorem 1], Ore showed that reversibility is a sufficient condition for a bicancellative semigroup to be embeddable into a group. In this scenario, the possible receiving SS-groups have a nice characterization, for which we need the following definition:

Definition 1.2.

Let SS be a semigroup. A receiving SS-group 𝐆=(G,η)\mathbf{G}=(G,\eta) is a group of right fractions of SS if for every gGg\in G there exist s,tSs,t\in S such that g=η(s)η(t)1g=\eta(s)\eta(t)^{-1}.

If 𝐆\mathbf{G} and 𝐆\mathbf{G}^{\prime} are groups of right fractions of SS, then they are isomorphic as SS-groups [9, Theorem 1.25], and the group of right fractions may thus be denoted by 𝚪S=(ΓS,γ)\bm{\Gamma}_{S}=(\Gamma_{S},\gamma).

Remark 1.3.

In [13], Dubreil proved that if SS is a bicancellative semigroup, then SS is left reversible if and only if 𝚪S\bm{\Gamma}_{S} exists. Whenever the group of right fractions of SS exists, it is the only receiving SS-group up to isomorphism of SS-groups, and it is hence isomorphic to the free SS-group [6, Corollary 2.20 and Remark 3.4].

Remark 1.4.

If SS is embeddable but not left reversible, then receiving SS-groups are not necessarily unique modulo isomorphism of SS-groups. This is shown in [6, Example 2.27], where both BS(1,2)\text{BS}(1,2) and 𝔽2\mathbb{F}_{2}—paired with adequate embeddings—play the role of receiving 𝔽2+\mathbb{F}_{2}^{+}-groups, but are non isomorphic even as groups.

1.2. Semigroup actions and topological natural extensions

We will understand an action α\alpha of SS over a set XX, denoted as S𝛼XS\overset{\alpha}{\curvearrowright}X, by a function α:S×XX\alpha\colon S\times X\rightarrow X such that α(s,α(t,x))=α(st,x)\alpha(s,\alpha(t,x))=\alpha(st,x) for all xXx\in X and s,tSs,t\in S. If SS is a monoid, we additionally request that α(1S,x)=x\alpha(1_{S},x)=x for all xXx\in X. Most of the time α(s,x)\alpha(s,x) will be written simply as sxs\cdot x, and the function α(s,):XX\alpha(s,\cdot)\colon X\rightarrow X as simply αs\alpha_{s} or ss. An action SXS\curvearrowright X is said to be surjective if αs\alpha_{s} is surjective for each sSs\in S. Given two monoids SS and TT, a monoid morphism θ:ST\theta\colon S\to T and two actions SXS\curvearrowright X and TYT\curvearrowright Y, a function φ:YX\varphi\colon Y\rightarrow X will be called θ\theta-equivariant if φ(θ(s)y)=sφ(x)\varphi(\theta(s)\cdot y)=s\cdot\varphi(x) for all yYy\in Y and sSs\in S. If S=TS=T and θ=idS\theta=\text{id}_{S}, we say that φ\varphi is equivariant. A subset AXA\subseteq X is SS-invariant for an action S𝛼XS\overset{\alpha}{\curvearrowright}X if sA:={sx:xA}AsA:=\{s\cdot x:x\in A\}\subseteq A for all sSs\in S, and completely SS-invariant if sA=AsA=A for every sSs\in S. In this case it makes sense to consider the action SAS\curvearrowright A given by restricting α|S×A:S×AA\alpha|_{S\times A}\colon S\times A\rightarrow A. Given xXx\in X, we define its SS-orbit as the SS-invariant set Sx={sx:sS}Sx=\{s\cdot x:s\in S\}.

1.2.1. Structure preserving actions

Suppose that XX is a compact metric space. We will denote by (X)\mathcal{B}(X) the associated Borel σ\sigma-algebra. Given a Borel probability measure μ\mu on (X,(X))(X,\mathcal{B}(X)), we will denote by (X,μ)(X,\mu) the corresponding Borel probability space.

A continuous SS-action (or simply continuous action, if SS is implicit) will be an action S𝛼XS\overset{\alpha}{\curvearrowright}X such that αs\alpha_{s} is continuous for each sSs\in S. Given two continuous actions S𝛼XS\overset{\alpha}{\curvearrowright}X and S𝛽YS\overset{\beta}{\curvearrowright}Y, a continuous equivariant function φ:YX\varphi\colon Y\to X is called a topological factor map if it is surjective, and a topological conjugacy if it is a homeomorphism. In the first case, we say that α\alpha is a topological factor of β\beta (and that β\beta is a topological extension of α\alpha), while in the second we say that both actions are topologically conjugate.

Given two Borel probability spaces (X,μ)(X,\mu) and (Y,ν)(Y,\nu), a measurable map φ:XY\varphi\colon X\rightarrow Y is measure preserving if μ(φ1A)=ν(A)\mu(\varphi^{-1}A)=\nu(A) for all A(Y)A\in\mathcal{B}(Y). Given a continuous action S𝛼XS\overset{\alpha}{\curvearrowright}X, we say that μ\mu is SS-invariant if each αs\alpha_{s} is measure-preserving, i.e., if μ(s1A)=μ(A)\mu(s^{-1}A)=\mu(A) for all A(X)A\in\mathcal{B}(X) and sSs\in S. In this case, we also say that α\alpha is a probability measure-preserving (p.m.p.) action, and denote it by S(X,μ)S\curvearrowright(X,\mu). Given two p.m.p. actions S𝛼(X,μ)S\overset{\alpha}{\curvearrowright}(X,\mu) and S𝛽(Y,ν)S\overset{\beta}{\curvearrowright}(Y,\nu), we say that α\alpha is factor of β\beta (and that β\beta is an extension of α\alpha) if there exist a full measure SS-invariant sets XXX^{\prime}\subseteq X and YYY^{\prime}\subseteq Y and a measure preserving equivariant map φ:YX\varphi\colon Y^{\prime}\rightarrow X^{\prime}. In this case, φ\varphi is called a factor map. If φ\varphi is moreover a bi-measurable map, we say that α\alpha and β\beta are (measure-theoretically) conjugate and that φ\varphi is a (measure-theoretical) conjugacy.

The space of Borel probability measures on (X,(X))(X,\mathcal{B}(X)) will be denoted by (X)\mathcal{M}(X). We can endow (X)\mathcal{M}(X) with the weak-* topology, which is metrizable and compact. Convergence is characterized by:

μnμXf𝑑μnXf𝑑μfor all f𝒞(X),\mu_{n}\to\mu\iff\int_{X}f\,d\mu_{n}\to\int_{X}f\,d\mu\quad\text{for all }f\in\mathcal{C}(X),

and a basis for the topology consists of the sets

V(μ,,ε)={μ(X):|f𝑑μf𝑑μ|<ε,f},V(\mu,\mathcal{F},\varepsilon)=\left\{\mu^{\prime}\in\mathcal{M}(X):\left|\int f\,d\mu^{\prime}-\int f\,d\mu\right|<\varepsilon,\forall f\in\mathcal{F}\right\},

where ε>0\varepsilon>0, 𝒞(X)\mathcal{F}\subseteq\mathcal{C}(X) is finite, and μ(X)\mu\in\mathcal{M}(X).

Given a continuous action SXS\curvearrowright X, the space of SS-invariant measures on (X,(X))(X,\mathcal{B}(X)), denoted by S(X)\mathcal{M}_{S}(X), is convex and weak-* closed in (X)\mathcal{M}(X). Depending on SXS\curvearrowright X, the space S(X)\mathcal{M}_{S}(X) may be empty or not. A measure μS(X)\mu\in\mathcal{M}_{S}(X) is ergodic if every set A(X)A\in\mathcal{B}(X) such that μ(s1AA)=0\mu(s^{-1}A\triangle A)=0 for all sSs\in S satisfies μ(A){0,1}\mu(A)\in\{0,1\}.

Given μ(X)\mu\in\mathcal{M}(X), we define the support of μ\mu as

supp(μ)={xX:μ(U)>0 for every open neighborhood U of x}.\operatorname{supp}(\mu)=\{x\in X:\mu(U)>0\text{ for every open neighborhood }U\text{ of }x\}.

The set supp(μ)\operatorname{supp}(\mu) is closed, and μ(A)=0\mu(A)=0 whenever Asupp(μ)=A\cap\operatorname{supp}(\mu)=\varnothing.

Remark 1.5.

If μS(X)\mu\in\mathcal{M}_{S}(X), then supp(μ)\operatorname{supp}(\mu) is an SS-invariant subset. Indeed, take xsupp(μ)x\in\operatorname{supp}(\mu) and sSs\in S. Given an open neighborhood UU of sxs\cdot x, we have that s1Us^{-1}U is an open neighborhood of xx, so μ(U)=μ(s1U)>0\mu(U)=\mu(s^{-1}U)>0. This proves that sxsupp(μ)s\cdot x\in\operatorname{supp}(\mu), so supp(μ)\operatorname{supp}(\mu) is SS-invariant.

1.2.2. Shift actions

Given a compact metric space 𝒜\mathcal{A}, we consider the product space 𝒜S\mathcal{A}^{S} of all functions x:S𝒜x\colon S\to\mathcal{A} endowed with the product topology. There is a natural continuous and surjective action S𝒜SS\curvearrowright\mathcal{A}^{S} given by

(sx)(t)=x(ts) for all s,tS,(s\cdot x)(t)=x(ts)\quad\text{ for all }s,t\in S,

which will be called the shift action and denoted by σ\sigma. An important special case is when 𝒜\mathcal{A} is a finite set—an alphabet—endowed with the discrete topology. If |𝒜|2|\mathcal{A}|\geq 2, the space of configurations 𝒜S\mathcal{A}^{S} is a Cantor space, and together with the action S𝒜SS\curvearrowright\mathcal{A}^{S} is called the full SS-shift. A closed SS-invariant subset of 𝒜S\mathcal{A}^{S} will be called an SS-subshift.

Given subsets QPSQ\subseteq P\subseteq S and p𝒜Pp\in\mathcal{A}^{P}, we define the cylinder set induced by pp at QQ as

[p;Q]={x𝒜S:x(t)=p(t) for all tQ}.[p;Q]=\{x\in\mathcal{A}^{S}:x(t)=p(t)\text{ for all }t\in Q\}.

The collection of all cylinder sets [x;Q][x;Q] with x𝒜Sx\in\mathcal{A}^{S} and QQ finite is a basis of clopen sets for the topology of 𝒜S\mathcal{A}^{S}.

1.3. Periodic configurations in groups

Let GG be a group and 𝒜G\mathcal{A}^{G} be the full GG-shift for some alphabet 𝒜\mathcal{A}. An element x𝒜Gx\in\mathcal{A}^{G} is periodic if it has a finite GG-orbit, that is, if |Gx|<|Gx|<\infty. Equivalently, x𝒜Gx\in\mathcal{A}^{G} is periodic if and only if its stabilizer

StabG(x)={gG:gx=x}\text{Stab}_{G}(x)=\{g\in G:g\cdot x=x\}

is a finite-index subgroup of GG, i.e., [G:StabG(x)]<[G:\text{Stab}_{G}(x)]<\infty. These notions extend to any group action GXG\curvearrowright X , and we denote by Per(X,G)\operatorname{Per}(X,G) the set of periodic points in XX.

A group GG is residually finite if for every pair of distinct elements g,hGg,h\in G there is a finite group FF and a group morphism θ:GF\theta\colon G\rightarrow F such that θ(g)θ(h)\theta(g)\neq\theta(h). More generally, if (p)(\textsc{p}) is a class of semigroups, we say that a semigroup SS is residually (p)(\textsc{p}) if for any distinct s,tSs,t\in S there exists a semigroup TT in the class (p)(\textsc{p}) and a semigroup morphism θ:ST\theta\colon S\to T such that θ(s)θ(t)\theta(s)\neq\theta(t). Residually finite groups have several equivalent characterizations, some of which are listed below:

Proposition 1.6.

Let GG be a group. The following are equivalent:

  1. (i)

    GG is residually finite.

  2. (ii)

    For any gGg\in G there exists a finite group FF and a group morphism θ:GF\theta\colon G\to F such that θ(g)1F\theta(g)\neq 1_{F},

  3. (iii)

    GG is isomorphic to a subgroup of a Cartesian product of (possibly infinitely many) finite groups.

  4. (iv)

    The set Per(𝒜G,G)\operatorname{Per}(\mathcal{A}^{G},G) is dense in 𝒜G\mathcal{A}^{G} for every finite alphabet 𝒜\mathcal{A}.

Evidently, every finite group is automatically residually finite. Some less trivial examples of residually finite groups include d\mathbb{Z}^{d} and all free groups, including 𝔽d\mathbb{F}_{d} for all dd\in\mathbb{N}; similarly, the linear groups GLd()\mathrm{GL}_{d}(\mathbb{Z}) for any d1d\geq 1 have this property. Furthermore, the class of residually finite groups is closed under taking subgroups, Cartesian products and inverse limits (but not under quotients). In contrast, divisible groups, that is, those for which the equation xn=gx^{n}=g has a solution for any n1n\geq 1 and gGg\in G, are never residually finite; these include ,\mathbb{Q},\mathbb{R} and \mathbb{C}. Similarly, non-Hopfian groups, that is, those for which a surjective, non-injective morphism GGG\to G exists, cannot be residually finite either; an example of such a group is the Baumslag–Solitar group BS(2,3)\mathrm{BS}(2,3) [21].

2. Periodicity for semigroup actions

In general, periodicity is a concept that tries to capture both finiteness and circularity. In the case of group actions, due to invertibility, the finite orbit condition is sufficient for precluding the existence of any transient portion in the orbit, thus assuring a circular behavior. However, even in the case of a single non-invertible surjective endomorphism, that is, an action of the semigroup \mathbb{N}, there could exist finite orbits with a transient initial section, yielding the distinction between periodic and pre-periodic points.

2.1. Periodic points and finitely supported invariant measures

The definition of periodicity in the context of semigroup actions is subtle and there may be various notions that capture it. However, one of our main purposes is to study finitely supported invariant measures. With this goal in mind, we introduce the following definition of periodicity.

Definition 2.1.

Let SXS\curvearrowright X be a semigroup action. An element xXx\in X will be called pre-periodic if |Sx|<|Sx|<\infty, and periodic if |Sx|<|Sx|<\infty and the set SxSx is completely SS-invariant, that is, tSx=SxtSx=Sx for all tSt\in S. The set of SS-periodic points of XX will be denoted as Per(X,S)\operatorname{Per}(X,S).

Example 2.2.

Complete SS-invariance does not necessarily imply |Sx|<|Sx|<\infty, although in the S=S=\mathbb{N} case it does. For instance, the element x{0,1}2x\in\{0,1\}^{\mathbb{N}^{2}} given by x(n,m)=1x(n,m)=1 if and only if n=mn=m, has infinite translates, and every element of 2\mathbb{N}^{2} acts bijectively upon its orbit.

A natural question is whether, as in the case of \mathbb{N} actions, the orbit of every pre-periodic point contains a periodic point or, equivalently, the support of an SS-invariant measure. However, this is not necessarily the case for general semigroup actions.

Example 2.3.

Consider 𝔽2+=a,b+\mathbb{F}_{2}^{+}=\langle a,b\rangle^{+} acting on the full 𝔽2+\mathbb{F}_{2}^{+}-shift {0,1}𝔽2+\{0,1\}^{\mathbb{F}_{2}^{+}}. Let x{0,1}𝔽2+x\in\{0,1\}^{\mathbb{F}_{2}^{+}} be the configuration defined by x(ε)=0x(\varepsilon)=0 and, for all s𝔽2+s\in\mathbb{F}_{2}^{+}, x(as)=1x(as)=1 and x(bs)=0x(bs)=0. The orbit of this configuration does not contain a periodic point, as can be seen in Figure 1, so there is no subset of this orbit which can support an invariant measure. Although these kind of orbits exhibit a valid form of periodicity, they will not be taken in consideration here.

x{x}y{y}b\scriptstyle{b}a\scriptstyle{a}b\scriptstyle{b}a\scriptstyle{a}
0{0}1{1}0{0}1{1}0{0}1{1}0{0}1{1}0{0}1{1}0{0}1{1}0{0}1{1}0{0}a\scriptstyle{a}b\scriptstyle{b}
Figure 1. A diagram of the 𝔽2+\mathbb{F}_{2}^{+}-orbit {x,y}\{x,y\} of xx and the representation of xx as a labeling of 𝔽2+\mathbb{F}_{2}^{+}.
Remark 2.4.

In the context of group actions, we saw that every periodic element induces a finite-index subgroup (namely, Stab(x)\text{Stab}(x)). For semigroup actions, if xXx\in X has finite SS-orbit, it induces a congruence on SS by

sxtsy=ty for all ySx{x}.s\mathrel{\mathcal{R}_{x}}t\iff s\cdot y=t\cdot y\text{ for all }y\in Sx\cup\{x\}.

Moreover, x\mathcal{R}_{x} has finite index, i.e., finitely many equivalence classes, so in order to admit periodic points, a semigroup must contain finite-index congruences. Note that, in the group case, the analogous to the congruence x\mathcal{R}_{x} is the largest normal subgroup contained in the stabilizer, i.e., the intersection of the stabilizers of all elements in GxGx.

Note that every element tSt\in S defines a function t:SxSxt\colon Sx\to Sx by ytyy\mapsto t\cdot y. The full transformation monoid of SxSx will be the set of all functions SxSxSx\to Sx together with function composition, denoted by 𝒯(Sx)\mathcal{T}(Sx). We have that sxts\mathrel{\mathcal{R}_{x}}t if and only if ss and tt define the same elements of 𝒯(Sx)\mathcal{T}(Sx), obtaining an embedding S/x𝒯(Sx)S/\mathcal{R}_{x}\to\mathcal{T}(Sx). By a subgroup of 𝒯(Sx)\mathcal{T}(Sx), we will understand a submonoid of 𝒯(Sx)\mathcal{T}(Sx) that is a group.

Proposition 2.5.

Let SXS\curvearrowright X be a semigroup action and xXx\in X. Then, xx is pre-periodic if and only if S/xS/\mathcal{R}_{x} is a finite semigroup, and xx is periodic if and only if S/xS/\mathcal{R}_{x} is a finite subgroup of 𝒯(Sx)\mathcal{T}(Sx).

Proof.

If xx is pre-periodic, there are finitely many functions SxSxSx\to Sx, hence finitely many equivalence classes for x\mathcal{R}_{x}. Conversely, if xx is not pre-periodic, there exists an infinite subset QSQ\subseteq S such that sxsxs\cdot x\neq s^{\prime}\cdot x for all s,sQs,s^{\prime}\in Q with sss\neq s^{\prime}. This immediately implies that |S/x||Q|=|S/\mathcal{R}_{x}|\geq|Q|=\infty.

If xx is periodic, every tSt\in S defines a bijection SxSxSx\to Sx, so there is an injective semigroup morphism S/xSym(Sx)S/\mathcal{R}_{x}\to\mathrm{Sym}(Sx). Since Sym(Sx)\mathrm{Sym}(Sx) is a finite group, S/xS/\mathcal{R}_{x} must be a finite group as well. Conversely, suppose that S/xS/\mathcal{R}_{x} is a subgroup of 𝒯(Sx)\mathcal{T}(Sx) and let tSt\in S. There exists some tSt^{\prime}\in S such that [tt]x[tt^{\prime}]_{\mathcal{R}_{x}} is the identity class in S/xS/\mathcal{R}_{x}, that is, ttt\circ t^{\prime} is the identity as a function SxSxSx\to Sx. Thus, for any sSs\in S we have t(tsx)=(tt)(sx)=sxt\cdot(t^{\prime}s\cdot x)=(tt^{\prime})\cdot(s\cdot x)=s\cdot x, i.e., tSx=SxtSx=Sx. ∎

Notice that the action of 𝔽2+\mathbb{F}_{2}^{+} upon the orbit of the point xx defined in Example 2.3 is transitive, i.e., for all s,tSs,t\in S, sxS(tx)s\cdot x\in S(t\cdot x). This holds true for every periodic point.

Proposition 2.6.

If xXx\in X is periodic, then the action SSxS\curvearrowright Sx is transitive. In particular, two SS-periodic orbits are either equal or disjoint.

Proof.

Suppose that SSxS\curvearrowright Sx is not transitive. Then, there exists tSt\in S such that xS(tx)x\not\in S(t\cdot x). Let

k=min{i1: there is a n>i with tnx=tix},k=\min\{i\geq 1:\text{ there is a }n>i\text{ with }t^{n}\cdot x=t^{i}\cdot x\},

which exists because SxSx is finite. By definition, there is an n>kn>k with tnx=tkxt^{n}\cdot x=t^{k}\cdot x, and we must have tn1xtk1xt^{n-1}\cdot x\neq t^{k-1}\cdot x. Note that

t(tn1x)=tnx=tkx=t(tk1x),t(t^{n-1}\cdot x)=t^{n}\cdot x=t^{k}\cdot x=t(t^{k-1}\cdot x),

contradicting the injectivity of tt upon SxSx. Therefore, SSxS\curvearrowright Sx is transitive.

For the final statement, let x1,x2Xx_{1},x_{2}\in X be two SS-periodic elements. If ySx1Sx2y\in Sx_{1}\cap Sx_{2}, there are elements s1,s2Ss_{1},s_{2}\in S such that y=s1x1=s2x2y=s_{1}\cdot x_{1}=s_{2}\cdot x_{2}. Since the action SSxS\curvearrowright Sx is transitive, we may choose an element tSt\in S with ts1x1=x1ts_{1}\cdot x_{1}=x_{1}, so ts2x2=x1ts_{2}\cdot x_{2}=x_{1}, which implies Sx1Sx2Sx_{1}\subseteq Sx_{2}. By symmetry, we get Sx2Sx1Sx_{2}\subseteq Sx_{1}, and so both orbits coincide. ∎

Remark 2.7.

Notice that if xx is pre-periodic for an \mathbb{N}-action, then the transitivity of Sx\mathbb{N}\curvearrowright Sx is equivalent to xx being periodic. However, Proposition 2.6 combined with Example 2.3 show that periodicity is a strictly stronger notion for general semigroup actions.

A periodic measure in S(X)\mathcal{M}_{S}(X) will be a measure μS(X)\mu\in\mathcal{M}_{S}(X) such that supp(μ)\operatorname{supp}(\mu) is finite. The set of periodic measures will be denoted by 𝒫(X,S)\mathcal{P}(X,S), and the set of periodic ergodic measures in S(X)\mathcal{M}_{S}(X) will be denoted as 𝒫erg(X,S)\mathcal{P}^{\mathrm{erg}}(X,S). The following result justifies our chosen definition of periodicity.

Lemma 2.8.

Let SXS\curvearrowright X be a continuous action, and xXx\in X a pre-periodic point. Then, there is an SS-invariant measure μS(X)\mu\in\mathcal{M}_{S}(X) with supp(μ)=Sx\textup{supp}(\mu)=Sx if and only if xx is periodic.

Proof.

Suppose xx is not periodic and choose tSt\in S such that t:SxSxt\colon Sx\rightarrow Sx is not bijective. Then, since SxSx is finite, there exists ySxy\in Sx such that t1(y)(Sx)=t^{-1}(y)\cap(Sx)=\varnothing. Take any SS-invariant measure μS(X)\mu\in\mathcal{M}_{S}(X) with supp(μ)Sx\textup{supp}(\mu)\subseteq Sx. Then, μ(XSx)=0\mu(X-Sx)=0, so

μ({y})=μ(t1(y))=μ(t1(y)Sx)=0.\mu(\{y\})=\mu(t^{-1}(y))=\mu(t^{-1}(y)\cap Sx)=0.

Thus, supp(μ)Sx\text{supp}(\mu)\subsetneq Sx, so no invariant measure can be supported upon the whole of SxSx. Conversely, assume each tSt\in S acts as a bijection of SxSx and consider the measure given by

μ(A)=|ASx||Sx| for every A(X).\mu(A)=\frac{|A\cap Sx|}{|Sx|}\quad\text{ for every }A\in\mathcal{B}(X).

Define φt:t1ASxASx\varphi_{t}\colon t^{-1}A\cap Sx\rightarrow A\cap Sx by ytyy\mapsto t\cdot y. For each tSt\in S, the function φt\varphi_{t} is injective as a consequence of tt being injective upon SxSx. Choose any yASxy\in A\cap Sx. Since tSx=SxtSx=Sx, there exists ySxy^{\prime}\in Sx with ty=yAt\cdot y^{\prime}=y\in A, so yt1Ay^{\prime}\in t^{-1}A as well, showing surjectivity of φt\varphi_{t}. Thus, μ\mu satisfies

μ(t1A)=|t1ASx||Sx|=|ASx||Sx|=μ(A),\mu(t^{-1}A)=\frac{|t^{-1}A\cap Sx|}{|Sx|}=\frac{|A\cap Sx|}{|Sx|}=\mu(A),

as we wanted. ∎

We have the following characterizations of periodic measures.

Proposition 2.9.

Let SXS\curvearrowright X be a continuous action and μS(X)\mu\in\mathcal{M}_{S}(X). Then,

  1. (i)

    μ𝒫(X,S)\mu\in\mathcal{P}(X,S) if and only if supp(μ)\operatorname{supp}(\mu) is a finite disjoint union of periodic SS-orbits,

  2. (ii)

    μ𝒫erg(X,S)\mu\in\mathcal{P}^{\mathrm{erg}}(X,S) if and only if supp(μ)\operatorname{supp}(\mu) corresponds to a single periodic SS-orbit.

Proof.

To prove (i), first assume that μ𝒫(X,S)\mu\in\mathcal{P}(X,S). Then, by Remark 1.5 we have that supp(μ)\operatorname{supp}(\mu) is SS-invariant, which means Sxsupp(μ)Sx\subseteq\operatorname{supp}(\mu) for all xsupp(μ)x\in\operatorname{supp}(\mu). If it were the case that an element xsupp(μ)x\in\operatorname{supp}(\mu) does not belong to the orbit of an element of supp(μ)\operatorname{supp}(\mu), then s1(x)Xsupp(μ)s^{-1}(x)\subseteq X-\operatorname{supp}(\mu) for every sSs\in S, yielding μ({x})=0\mu(\{x\})=0, a contradiction. This means that supp(μ)=xsupp(μ)Sx\operatorname{supp}(\mu)=\bigcup_{x\in\operatorname{supp}(\mu)}Sx, which implies the statement, since periodic orbits are either equal or disjoint by Proposition 2.6. The converse is direct.

To prove (ii), first assume μ𝒫erg(X,S)\mu\in\mathcal{P}^{\mathrm{erg}}(X,S). Then, by part (i) there exist SS-periodic elements x1,,xmXx_{1},\dots,x_{m}\in X such that

supp(μ)=i=1mSxi.\operatorname{supp}(\mu)=\bigsqcup_{i=1}^{m}Sx_{i}.

Given tSt\in S and iji\neq j, if xt1(Sxi)Sxjx\in t^{-1}(Sx_{i})\cap Sx_{j}, then txSxitSxj=SxiSxj=t\cdot x\in Sx_{i}\cap tSx_{j}=Sx_{i}\cap Sx_{j}=\varnothing, so t1(Sxi)Sxj=t^{-1}(Sx_{i})\cap Sx_{j}=\varnothing. Also note that Sxit1(Sxi)Sx_{i}\subseteq t^{-1}(Sx_{i}) for every tSt\in S and ii, as xix_{i} is SS-periodic. Therefore,

μ(t1(Sxi)Sxi)=μ(t1(Sxi)Sxi)μ(Xsupp(μ))=0.\mu(t^{-1}(Sx_{i})\triangle Sx_{i})=\mu(t^{-1}(Sx_{i})-Sx_{i})\leq\mu(X-\operatorname{supp}(\mu))=0.

Hence, since μ\mu is ergodic, μ(Sxi){0,1}\mu(Sx_{i})\in\{0,1\}, so supp(μ)\operatorname{supp}(\mu) consists of a single orbit.

Conversely, if supp(μ)=Sx\operatorname{supp}(\mu)=Sx for an SS-periodic point xx and A(X)A\in\mathcal{B}(X) is such that μ(s1AA)=0\mu(s^{-1}A\triangle A)=0 for all sSs\in S, then Sx(As1A)=Sx\cap(A-s^{-1}A)=\varnothing for all sSs\in S, and we have two cases. First, if SxA=Sx\cap A=\varnothing, then μ(A)=0\mu(A)=0. Second, if txSxAt\cdot x\in Sx\cap A for some tSt\in S, necessarily we must have txs1At\cdot x\in s^{-1}A for every sSs\in S as well, obtaining stxAst\cdot x\in A for every sSs\in S. Since the action of SS upon SxSx is transitive, this implies SxASx\subseteq A, and so μ(A)=1\mu(A)=1. Thus, μ\mu is ergodic. ∎

As a consequence of last proposition, a periodic measure can always be written in the form

μ=i=1mμ(Sxi)|Sxi|xSxiδx,\mu=\sum_{i=1}^{m}\frac{\mu(Sx_{i})}{|Sx_{i}|}\sum_{x\in Sx_{i}}\delta_{x},

where i=1mμ(Sxi)=1\sum_{i=1}^{m}\mu(Sx_{i})=1, with m=1m=1 if and only if the measure is moreover ergodic.

2.2. Residual finiteness and periodic orbits

We aim to relate algebraic properties of semigroups and both periodic and pre-periodic orbits. In order to do this, let us start by introducing two properties that are semigroup analogues to residual finiteness in groups.

Definition 2.10.

A semigroup SS is residually a finite group (resp. residually a finite semigroup) if for every pair of distinct elements t,tSt,t^{\prime}\in S there is a finite group (resp. semigroup) FF and a semigroup morphism θ:SF\theta\colon S\rightarrow F such that θ(t)θ(t)\theta(t)\neq\theta(t^{\prime}).

Remark 2.11.

If SS is residually a finite group (resp. residually a finite semigroup) and QSQ\subseteq S, then for every t,tQt,t^{\prime}\in Q there exist a group (resp. semigroup) Ft,tF_{t,t^{\prime}} and a semigroup morphism θt,t:SFt,t\theta_{t,t^{\prime}}\colon S\to F_{t,t^{\prime}} such that θt,t(t)θt,t(t)\theta_{t,t^{\prime}}(t)\neq\theta_{t,t^{\prime}}(t^{\prime}) if ttt\neq t^{\prime}, and Ft,tF_{t,t^{\prime}} is the trivial group if t=tt=t^{\prime}. Thus, we may define a semigroup morphism θQ:St,tQFt,t\theta_{Q}\colon S\to\prod_{t,t^{\prime}\in Q}F_{t,t^{\prime}} such that θ(t)θ(t)\theta(t)\neq\theta(t^{\prime}) for any t,tQt,t^{\prime}\in Q, ttt\neq t^{\prime}, by sending any sSs\in S to the tuple (θt,t(s))t,tQ(\theta_{t,t^{\prime}}(s))_{t,t^{\prime}\in Q}.

As a consequence of this, SS is residually a finite group (resp. residually a finite semigroup) if and only if for every finite subset QSQ\subseteq S, there is a finite group (resp. semigroup) FF, namely t,tQFt,t\prod_{t,t^{\prime}\in Q}F_{t,t^{\prime}}, and a semigroup morphism θ:SF\theta\colon S\to F, such that θ|Q\theta\rvert_{Q} is injective.

Remark 2.12.

It is clear that being residually a finite group always implies being residually a finite semigroup. Furthermore, we have the converse in the case where SS is a group, since the image of a group via a semigroup morphism is always a group. In this situation, both notions coincide with the classic notion of residual finiteness for groups. However, for general semigroups these notions differ: every finite non-bicancellative semigroup is residually a finite semigroup but not residually a finite group. Indeed, if SS a non-bicancellative semigroup, it cannot be residually a finite group. Indeed, if θ:SF\theta\colon S\rightarrow F is a morphism to a group FF, and a,b,cSa,b,c\in S are such that ab=acab=ac and bcb\neq c, then θ(a)θ(b)=θ(a)θ(c)\theta(a)\theta(b)=\theta(a)\theta(c) so θ(b)=θ(c)\theta(b)=\theta(c), so bb and cc cannot be distinguished by such morphism.

The following result is analogous to the characterization of residually finite groups provided in Proposition 1.6 (iii).

Proposition 2.13.

A semigroup SS is residually a finite group (resp. residually a finite semigroup) if and only if it can be embedded into a Cartesian product of finite groups (resp. semigroups). In particular, a semigroup is residually a finite group if and only if it can be embedded into a residually finite group.

Proof.

Assume SS is residually a finite group (resp. residually a finite semigroup). The map θQ\theta_{Q} for Q=SQ=S, following the notation from Remark 2.11, is an embedding into a product of finite groups (resp. semigroups). ∎

Remark 2.14.

Notice that Proposition 2.13 implies that any semigroup that is residually a finite group is necessarily bicancellative, and moreover embeddable into some residually finite group.

A semigroup SS such that for every pair of distinct elements s,sSs,s^{\prime}\in S, there is some tSt\in S with tststs\neq ts^{\prime}, is called a left reductive semigroup. In particular, every monoid and every left cancellative semigroup is left reductive. On the other hand, a semigroup action SXS\curvearrowright X is said to be faithful if for every s,sSs,s^{\prime}\in S with sss\neq s^{\prime} there exists xXx\in X such that sxsxs\cdot x\neq s^{\prime}\cdot x. The following result relates these two notions.

Lemma 2.15.

A semigroup SS is left reductive if and only if the shift action S𝒜SS\curvearrowright\mathcal{A}^{S} is faithful for every (resp. any) alphabet 𝒜\mathcal{A} with |𝒜|2|\mathcal{A}|\geq 2.

Proof.

If SS is left reductive and s,sSs,s^{\prime}\in S are distinct elements, there is a tSt\in S with tststs\neq ts^{\prime}, so the configuration x𝒜Sx\in\mathcal{A}^{S} given by x(t)=1x(t)=1 and x(t)=0x(t^{\prime})=0 if ttt^{\prime}\neq t satisfies sxsxs\cdot x\neq s^{\prime}\cdot x. Conversely, if S𝒜SS\curvearrowright\mathcal{A}^{S} is faithful for some alphabet 𝒜\mathcal{A} with |𝒜|2|\mathcal{A}|\geq 2, given sss\neq s^{\prime} in SS there exist x𝒜Sx\in\mathcal{A}^{S} and tSt\in S with (sx)(t)(sx)(t)(s\cdot x)(t)\neq(s^{\prime}\cdot x)(t), so tsts and tsts^{\prime} cannot be equal. ∎

Our first main result is the following.

\firstproposition

*

Proof.

Let us prove (ii), as the proof of (i) is essentially the same.

Suppose that SS is residually a finite group, x𝒜Sx\in\mathcal{A}^{S} is an arbitrary point, and let QSQ\subseteq S be an arbitrary finite set. By Remark 2.11, there exist a finite group FF and a semigroup morphism θ:SF\theta\colon S\to F with θ|Q\theta\rvert_{Q} injective.

Note that every w𝒜Fw\in\mathcal{A}^{F} defines a point w¯𝒜S\bar{w}\in\mathcal{A}^{S} by w¯(s)=w(θ(s))\bar{w}(s)=w(\theta(s)) for all sSs\in S. It holds true that sw¯=θ(s)w¯s\cdot\bar{w}=\overline{\theta(s)\cdot w} for any sSs\in S, since for every tSt\in S we have

(sw¯)(t)=w¯(ts)=w(θ(ts))=w(θ(t)θ(s))=(θ(s)w)(θ(t))=(θ(s)w¯)(t).(s\cdot\bar{w})(t)=\bar{w}(ts)=w(\theta(ts))=w(\theta(t)\theta(s))=(\theta(s)\cdot w)(\theta(t))=\big{(}\overline{\theta(s)\cdot w}\big{)}(t).

Thus, for any w𝒜Fw\in\mathcal{A}^{F}, the element w¯𝒜S\bar{w}\in\mathcal{A}^{S} is SS-periodic. Indeed, the above proven equality shows that Sw¯{w¯:wFw}S\bar{w}\subseteq\{\bar{w^{\prime}}:w^{\prime}\in Fw\}, which together with the fact that FF is finite prove that |Sw¯|<|S\bar{w}|<\infty. Moreover, given tSt\in S, since θ(S)\theta(S) is a group, for every sSs\in S, there is an sSs^{\prime}\in S such that θ(t)θ(s)=θ(s)\theta(t)\theta(s^{\prime})=\theta(s), yielding

tsw¯=θ(ts)w¯=θ(s)w¯=sw¯.ts^{\prime}\cdot\bar{w}=\overline{\theta(ts^{\prime})\cdot w}=\overline{\theta(s)\cdot w}=s\cdot\bar{w}.

Thus, tSw¯=Sw¯tS\bar{w}=S\bar{w} for every tSt\in S. Define a configuration wx:F𝒜w_{x}\colon F\to\mathcal{A} by wx(θ(s))=x(s)w_{x}(\theta(s))=x(s) for all sQs\in Q, extending it arbitrarily to the rest of FF if needed. As the map θ|Q:QF\theta\rvert_{Q}\colon Q\to F is injective, this is well-defined. The corresponding point wx¯\overline{w_{x}} in 𝒜S\mathcal{A}^{S} is SS-periodic and satisfies the equality wx¯|Q=x|Q\overline{w_{x}}\rvert_{Q}=x\rvert_{Q}. This shows that, for any x𝒜Sx\in\mathcal{A}^{S} and any finite QSQ\subseteq S, the cylinder [x;Q][x;Q] must contain a periodic point, and thus the set of periodic points is dense.

Conversely, if 𝒜S\mathcal{A}^{S} has dense set of periodic points and s,sSs,s^{\prime}\in S are distinct elements, by left reductiveness, there exists tSt\in S such that tststs\neq ts^{\prime}. Choose two distinct elements from 𝒜\mathcal{A}, which, for simplicity, will be denoted by 0 and 11, and consider any x𝒜Sx\in\mathcal{A}^{S} such that x(ts)=0x(ts)=0 and x(ts)=1x(ts^{\prime})=1. By our hypothesis of denseness of periodic points, we may assume that xx is SS-periodic and consider F=Sym(Sx)F=\text{Sym}(Sx), which is a finite group. The definition of SS-periodicity yields a morphism θ:SF\theta\colon S\rightarrow F by sending aSa\in S to the function a|Sx:SxSxa\rvert_{Sx}\colon Sx\rightarrow Sx given by a|Sx(bx)=abxa\rvert_{Sx}(b\cdot x)=ab\cdot x for all bSb\in S. By our choices of xx and tt, we must have (sx)(t)=x(ts)=0(s\cdot x)(t)=x(ts)=0 and (sx)(t)=x(ts)=1(s^{\prime}\cdot x)(t)=x(ts^{\prime})=1. Hence, sxsxs\cdot x\neq s^{\prime}\cdot x, and thus s|Sxs\rvert_{Sx} and s|Sxs^{\prime}\rvert_{Sx} correspond to different permutations of SxSx, that is, θ(s)θ(s)\theta(s)\neq\theta(s^{\prime}).

The proof of (i) is identical, but replacing the finite group FF by a finite semigroup satisfying that θ|Q\theta\rvert_{Q} is injective, and omitting the proof of complete SS-invariance; and for the opposite direction, by replacing the symmetric group Sym(Sx)\textup{Sym}(Sx) by the monoid of all transformations SxSxSx\to Sx. ∎

Since a left cancellative semigroup is automatically left reductive, any semigroup that is residually a finite group is left reductive. In the non-left reductive case, however, the associated full SS-shift may have dense periodic points, but this does not imply that the semigroup is residually a finite group.

Example 2.16 (Non-reductive case).

Let L={a,b}L=\{a,b\} be the left zero semigroup, where st=sst=s for all s,tLs,t\in L. The semigroup LL is not left reductive, but satisfies that sx=xs\cdot x=x for every sLs\in L and x𝒜Lx\in\mathcal{A}^{L}, so every element of 𝒜L\mathcal{A}^{L} is periodic. Thus, it has a dense set of periodic points for the shift action L𝒜LL\curvearrowright\mathcal{A}^{L}, but LL is not residually a finite group since it is not bicancellative. Similarly, the semigroup S=×LS=\mathbb{Z}\times L provides an example of an infinite semigroup with dense set of periodic points in 𝒜S\mathcal{A}^{S}, which is not residually a finite group.

2.3. The periodic approximation property

In view of Theorem Introduction, it is natural to ask under what conditions upon the acting semigroup SS the periodic measures on 𝒜S\mathcal{A}^{S} are weak-* dense. Considering this, we now introduce a definition which will be fundamental throughout this work.

Definition 2.17.

A semigroup SS has the periodic approximation property, or simply the pa property, if 𝒫(X,S)\mathcal{P}(X,S) is weak-* dense in S(𝒜S)\mathcal{M}_{S}(\mathcal{A}^{S}) for every finite alphabet 𝒜\mathcal{A}. A semigroup SS has the ergodic periodic approximation property, or simply the epa property, if 𝒫erg(X,S)\mathcal{P}^{\mathrm{erg}}(X,S) is weak-* dense in S(𝒜S)\mathcal{M}_{S}(\mathcal{A}^{S}) for every finite alphabet 𝒜\mathcal{A}.

This definition extends the one discussed for groups in the Introduction. In the case of groups, the pa property is equivalent to the property md described by Kechris [7, 22] in the context of weak containment of group actions. It is known that all amenable, residually finite groups have the epa property [26]. Similarly, Bowen showed that free groups on finitely many generators have the pa property [5]. Examples of residually finite groups without the pa property include the special linear group SLn()\operatorname{SL}_{n}(\mathbb{Z}) for n3n\geq 3 [22] (in contrast to SL2()\operatorname{SL}_{2}(\mathbb{Z}), which has the pa property) and 𝔽2×𝔽2\mathbb{F}_{2}\times\mathbb{F}_{2} as a consequence of the negative answer to the Connes’ Embedding Problem (see [7, p. 27] and [20]).

We stress that, although the pa property has been attentively studied for groups, this has not been the case with general semigroups. As a first preliminary result, we show that residual finiteness turns out to be a key necessary condition for having the pa property, as seen below.

Proposition 2.18.

Let SS be a left reductive semigroup. If SS has the pa property, then it is residually a finite group.

Proof.

Assume SS is not residually a finite group. By Theorem Introduction, there is a configuration x𝒜Sx\in\mathcal{A}^{S} and a finite subset FSF\subseteq S such that [x;F]Per(𝒜S,S)=[x;F]\cap\operatorname{Per}(\mathcal{A}^{S},S)=\varnothing. Choose a measure μS(𝒜S)\mu\in\mathcal{M}_{S}(\mathcal{A}^{S}) such that μ([x;F])>0\mu([x;F])>0 (take, for instance, the Bernoulli measure associated to a positive probability vector), and let (μn)n(\mu_{n})_{n} be a sequence in 𝒫(X,S)\mathcal{P}(X,S) that weak-* converges to μ\mu. By Proposition 2.9, we can assume that for every n1n\geq 1,

μn=i=1mnμn(Sxin)|Sxin|ySxinδy,\mu_{n}=\sum_{i=1}^{m_{n}}\frac{\mu_{n}(Sx^{n}_{i})}{|Sx^{n}_{i}|}\sum_{y\in Sx^{n}_{i}}\delta_{y},

where each xinx^{n}_{i} is SS-periodic for every 1imn1\leq i\leq m_{n}. By weak-* convergence, we have that μn([x;F])μ([x;F])\mu_{n}([x;F])\to\mu([x;F]). However, for all n1n\geq 1 and 1imn1\leq i\leq m_{n}, the set [x;F]Sxin[x;F]\cap Sx^{n}_{i} is empty, meaning that μn([x;F])=0\mu_{n}([x;F])=0, a contradiction with our choice of μ\mu. ∎

Remark 2.19.

Based on Example 2.16, it is not difficult to construct an example of a semigroup with the pa property that is not left reductive, and thus cannot be residually a finite group.

3. A sufficient condition for the pa property

We want to understand the pa property in the more general landscape of semigroups, taking advantage of what is already known for groups. With this goal in mind, we restrict ourselves to the class of embeddable monoids. The purpose of this section is to establish a direct connection between the pa property of an embeddable monoid SS and the pa property of the corresponding free SS-group. In order to do this, the tool of choice will be the natural extension construction, which associates to an SS-action a corresponding invertible action of the free SS-group which extends the original action in a natural way.

3.1. Topological natural extensions

In [6], topological natural extensions were profusely discussed. Based on this work, we consider the following definition.

Definition 3.1.

Let SS be an embeddable monoid and S𝛼XS\overset{\alpha}{\curvearrowright}X a continuous SS-action. Given a receiving SS-group 𝐆=(G,η)\mathbf{G}=(G,\eta), the topological natural G\mathbf{G}-extension is the tuple (X𝐆,σ,π)(X_{\mathbf{G}},\sigma,\pi), where

X𝐆={(xh)hGXG:sxh=xη(s)h for all sS and hG}X_{\mathbf{G}}=\left\{(x_{h})_{h\in G}\in X^{G}:s\cdot x_{h}=x_{\eta(s)h}\textup{ for all }s\in S\textup{ and }h\in G\right\}

is endowed with the subspace topology of the product topology, σ\sigma denotes the restriction of the shift action G𝜎XGG\overset{\sigma}{\curvearrowright}X^{G} to X𝐆X_{\mathbf{G}}, and π:X𝐆X\pi\colon X_{\mathbf{G}}\rightarrow X is the projection π1G:XGX\pi_{1_{G}}\colon X^{G}\to X restricted to X𝐆X_{\mathbf{G}}. If π:X𝐆X\pi\colon X_{\mathbf{G}}\rightarrow X is surjective, α\alpha is said to be topologically 𝐆\textbf{topologically }\mathbf{G}-extensible, and if X𝐆X_{\mathbf{G}}\neq\varnothing, α\alpha is said to be topologically partially 𝐆\mathbf{G}-extensible.

Since X𝐆X_{\mathbf{G}} is closed in XGX^{G}, and XGX^{G} is compact—as we are assuming XX is a compact metric space—, we have that X𝐆X_{\mathbf{G}} compact as well. In addition, the map π\pi is continuous and η\eta-equivariant in the sense that

sπ(x¯)=π(η(s)x¯) for all x¯X𝐆 and sS.s\cdot\pi(\overline{x})=\pi(\eta(s)\cdot\overline{x})\quad\text{ for all }\overline{x}\in X_{\mathbf{G}}\text{ and }s\in S.

In [6] it was proven that the natural 𝐆\mathbf{G}-extension of SXS\curvearrowright X comes with a universal property: if (X^,β,τ)(\hat{X},\beta,\tau) is any tuple such that X^\hat{X} is a topological space, β\beta is a continuous action GX^G\curvearrowright\hat{X}, and τ:X^X\tau\colon\hat{X}\rightarrow X is surjective, continuous, and η\eta-equivariant, then there is a unique equivariant continuous function φ:X^X𝐆\varphi\colon\hat{X}\rightarrow X_{\mathbf{G}} satisfying πφ=τ\pi\circ\varphi=\tau, so that the following diagram commutes.

X^{\hat{X}}X𝐆{X_{\mathbf{G}}}X{X}φ\scriptstyle{\varphi}τ\scriptstyle{\tau}π\scriptstyle{\pi}

A key consequence of the main result in [6] is the following:

Theorem 3.2 ([6]).

If 𝐆\mathbf{G} is the free SS-group, then every surjective continuous SS-action is topologically 𝐆\mathbf{G}-extensible.

3.2. Measure-theoretical extensions

We will make use of topological natural extensions to define measure-theoretical natural extensions. Observe that, given a receiving SS-group 𝐆=(G,η)\mathbf{G}=(G,\eta) and a topologically 𝐆\mathbf{G}-extensible continuous action SXS\curvearrowright X, the projection map π:X𝐆X\pi\colon X_{\mathbf{G}}\rightarrow X induces a push-forward map π:(X𝐆)(X)\pi_{*}\colon\mathcal{M}(X_{\mathbf{G}})\rightarrow\mathcal{M}(X) given by πμ^(A)=μ(π1(A)),\pi_{*}\hat{\mu}(A)=\mu(\pi^{-1}(A)), for all A(X)A\in\mathcal{B}(X). Moreover, since π\pi is η\eta-equivariant, the image of a GG-invariant measure on X𝐆X_{\mathbf{G}} via π\pi_{*} is an SS-invariant measure on XX, so the operator π:G(X𝐆)S(X)\pi_{*}\colon\mathcal{M}_{G}(X_{\mathbf{G}})\rightarrow\mathcal{M}_{S}(X) is well-defined. From now on, π\pi_{*} will denote this restricted version of the push-forward. It is standard that π\pi_{*} is weak-* continuous (see, e.g., [14, Appendix B]).

Definition 3.3.

Let SXS\curvearrowright X be a continuous action and μS(X)\mu\in\mathcal{M}_{S}(X). If there exists μ¯G(X𝐆)\bar{\mu}\in\mathcal{M}_{G}(X_{\mathbf{G}}) such that πμ¯=μ\pi_{*}\bar{\mu}=\mu, then μ\mu will be called 𝐆\mathbf{G}-extensible and μ¯\bar{\mu} a 𝐆\mathbf{G}-extension of μ\mu. The set im(π)\operatorname{im}(\pi_{*}) will be denoted by Ext𝐆(X,S)\operatorname{Ext}_{\mathbf{G}}(X,S).

Remark 3.4.

Observe that if 𝐆=(G,η)\mathbf{G}=(G,\eta) is a receiving SS-group, S(X,μ)S\curvearrowright(X,\mu) is a p.m.p. action, and μS(X)\mu\in\mathcal{M}_{S}(X) is 𝐆\mathbf{G}-extensible, there is a GG-invariant measure μ¯\bar{\mu} on (X𝐆,(X𝐆))(X_{\mathbf{G}},\mathcal{B}(X_{\mathbf{G}})) such that π:(X𝐆,μ¯)(X,μ)\pi\colon(X_{\mathbf{G}},\bar{\mu})\to(X,\mu) is measure-preserving. Thus, the shift action G(X𝐆,μ¯)G\curvearrowright(X_{\mathbf{G}},\bar{\mu}) is a p.m.p. action, and the tuple (X𝐆,σ,μ¯,π)(X_{\mathbf{G}},\sigma,\bar{\mu},\pi) may be regarded as a measure-theoretical natural 𝐆\mathbf{G}-extension of S(X,μ)S\curvearrowright(X,\mu).

Remark 3.5.

Let 𝐆=(G,η)\mathbf{G}=(G,\eta) be a receiving SS-group. We may identify the topological natural 𝐆\mathbf{G}-extension of 𝒜S\mathcal{A}^{S} with (𝒜G,π)(\mathcal{A}^{G},\pi), where π:𝒜G𝒜S\pi\colon\mathcal{A}^{G}\rightarrow\mathcal{A}^{S} is given by x¯x¯η\overline{x}\mapsto\overline{x}\circ\eta (see [6]). With this identification in mind, a measure μS(𝒜S)\mu\in\mathcal{M}_{S}(\mathcal{A}^{S}) is 𝐆\mathbf{G}-extensible if and only if there is a measure μ¯G(𝒜G)\bar{\mu}\in\mathcal{M}_{G}(\mathcal{A}^{G}) such that πμ¯=μ\pi_{*}\bar{\mu}=\mu. In the forthcoming, whenever we deal with natural extensions in the symbolic case, we will proceed with this identification, so π\pi will denote the map x¯x¯η\overline{x}\mapsto\overline{x}\circ\eta.

Proposition 3.6.

The subset Ext𝐆(X,S)S(X)\operatorname{Ext}_{\mathbf{G}}(X,S)\subseteq\mathcal{M}_{S}(X) is weak-* closed and convex.

Proof.

Let μS(X)\mu\in\mathcal{M}_{S}(X), and let (μn)n(\mu_{n})_{n} be a sequence in Ext𝐆(X,S)\operatorname{Ext}_{\mathbf{G}}(X,S) converging to μ\mu. Consider, for every nn\in\mathbb{N}, a measure μ¯nG(X𝐆)\bar{\mu}_{n}\in\mathcal{M}_{G}(X_{\mathbf{G}}) such that πμ¯n=μn\pi_{*}\bar{\mu}_{n}=\mu_{n}. Since X𝐆X_{\mathbf{G}} is compact, G(X𝐆)\mathcal{M}_{G}(X_{\mathbf{G}}) is weak-* compact. Thus, we can find a subsequence (μ¯nk)k(\bar{\mu}_{n_{k}})_{k} of (μ¯n)n(\bar{\mu}_{n})_{n} weak-* converging to some μ¯G(X𝐆)\bar{\mu}\in\mathcal{M}_{G}(X_{\mathbf{G}}). By continuity of π\pi_{*}, we get

μ=limkμnk=limkπμ¯nk=π(limkμ¯nk)=πμ¯.\mu=\lim_{k\to\infty}\mu_{n_{k}}=\lim_{k\to\infty}\pi_{*}\bar{\mu}_{n_{k}}=\pi_{*}\left(\lim_{k\to\infty}\bar{\mu}_{n_{k}}\right)=\pi_{*}\bar{\mu}.

Therefore, μExt𝐆(X,S)\mu\in\operatorname{Ext}_{\mathbf{G}}(X,S). Finally, Ext𝐆(X,S)\operatorname{Ext}_{\mathbf{G}}(X,S) is convex, as it is the image of the convex set G(X𝐆)\mathcal{M}_{G}(X_{\mathbf{G}}) under the linear function π\pi_{*}. ∎

Proposition 3.7.

Let 𝐆=(G,η)\mathbf{G}=(G,\eta) be a receiving SS-group. Let S𝛼(X,μ),S𝛽(Y,ν)S\overset{\alpha}{\curvearrowright}(X,\mu),S\overset{\beta}{\curvearrowright}(Y,\nu) be two p.m.p. actions such that α\alpha is a factor of β\beta. Then, if ν\nu is 𝐆\mathbf{G}-extensible, so is μ\mu.

Proof.

Let XXX^{\prime}\subseteq X and YYY^{\prime}\subseteq Y be full measure SS-invariant sets and let φ:YX\varphi\colon Y^{\prime}\rightarrow X^{\prime} be a measure preserving equivariant map. Let (X𝐆,πα)(X_{\mathbf{G}},\pi_{\alpha}) and (Y𝐆,πβ)(Y_{\mathbf{G}},\pi_{\beta}) be the topological 𝐆\mathbf{G}-extensions of α\alpha and β\beta, respectively, and ν¯G(X𝐆)\bar{\nu}\in\mathcal{M}_{G}(X_{\mathbf{G}}) such that (πβ)ν¯=ν(\pi_{\beta})_{*}\bar{\nu}=\nu. The first thing we need to check is that X𝐆X_{\mathbf{G}} is non-empty (and hence α\alpha is topologically partially 𝐆\mathbf{G}-extensible). Indeed, since ν¯\bar{\nu} is GG-invariant, ν¯(gπβ1(Y))=μ(Y)=1\bar{\nu}(g\pi_{\beta}^{-1}(Y^{\prime}))=\mu(Y^{\prime})=1 for all gGg\in G, so the fact that GG is countable implies

ν¯(gGgπβ1(Y))=1.\bar{\nu}\left(\bigcap_{g\in G}g\pi_{\beta}^{-1}(Y^{\prime})\right)=1.

Thus, there is an element (yh)gGY𝐆(y_{h})_{g\in G}\in Y_{\mathbf{G}} with yhYy_{h}\in Y^{\prime} for all hGh\in G. The element (φ(yh))hG(\varphi(y_{h}))_{h\in G} is an element of X𝐆X_{\mathbf{G}}.

We want to construct a measure μ¯G(X𝐆)\bar{\mu}\in\mathcal{M}_{G}(X_{\mathbf{G}}) satisfying (πα)μ¯=μ(\pi_{\alpha})_{*}\bar{\mu}=\mu. Consider the function

φ¯:gGgπβ1(Y)\displaystyle\bar{\varphi}\colon\bigcap_{g\in G}g\pi_{\beta}^{-1}(Y^{\prime}) gGgπα1(X)\displaystyle\longrightarrow\bigcap_{g\in G}g\pi_{\alpha}^{-1}(X^{\prime})
(yh)hG\displaystyle(y_{h})_{h\in G} (φ(yh))hG,\displaystyle\longmapsto(\varphi(y_{h}))_{h\in G},

which is well-defined, as φ(Y)X\varphi(Y^{\prime})\subseteq X^{\prime} and, given (yh)hG(y_{h})_{h\in G} in gGgπβ1(Y)\bigcap_{g\in G}g\pi_{\beta}^{-1}(Y^{\prime}), yhYy_{h}\in Y^{\prime} for every hGh\in G. As we already mentioned, φ¯\bar{\varphi} is defined upon a full-measure subset of Y𝐆Y_{\mathbf{G}}. Also, φ¯\bar{\varphi} is clearly GG-equivariant. Define, for A¯(X𝐆)\bar{A}\in\mathcal{B}(X_{\mathbf{G}}),

μ¯(A¯)=ν¯(φ¯1(A¯gGgπα1(X))).\bar{\mu}\left(\bar{A}\right)=\bar{\nu}\left(\bar{\varphi}^{-1}\left(\bar{A}\cap\bigcap_{g\in G}g\pi_{\alpha}^{-1}(X^{\prime})\right)\right).

The set function μ¯\bar{\mu} is a probability measure on X𝐆X_{\mathbf{G}}. The GG-invariance of μ¯\bar{\mu} comes as a consequence of the GG-equivariance of φ¯\bar{\varphi} and the GG-invariance of gGgπα1(X)\bigcap_{g\in G}g\pi_{\alpha}^{-1}(X^{\prime}). To see that πα\pi_{\alpha} is a measure-preserving map, note that παφ¯=φπβ\pi_{\alpha}\circ\bar{\varphi}=\varphi\circ\pi_{\beta}, so for every A(X)A\in\mathcal{B}(X),

μ¯(πα1(A))\displaystyle\bar{\mu}(\pi_{\alpha}^{-1}(A)) =ν¯(φ¯1(πα1(A)gGgπα1(X)))\displaystyle=\bar{\nu}\left(\bar{\varphi}^{-1}\left(\pi_{\alpha}^{-1}(A)\cap\bigcap_{g\in G}g\pi_{\alpha}^{-1}(X^{\prime})\right)\right)
=ν¯(φ¯1(πα1(A))gGgφ¯1(πα1(X)))\displaystyle=\bar{\nu}\left(\bar{\varphi}^{-1}(\pi_{\alpha}^{-1}(A))\cap\bigcap_{g\in G}g\bar{\varphi}^{-1}(\pi_{\alpha}^{-1}(X^{\prime}))\right)
=ν¯(πβ1(φ1(A))gGgπβ1(Y))\displaystyle=\bar{\nu}\left(\pi_{\beta}^{-1}\left(\varphi^{-1}(A)\right)\cap\bigcap_{g\in G}g\pi_{\beta}^{-1}(Y^{\prime})\right)
=ν¯(πβ1(φ1(A)))=ν(φ1(A))=μ(A).\displaystyle=\bar{\nu}\left(\pi_{\beta}^{-1}\left(\varphi^{-1}(A)\right)\right)=\nu(\varphi^{-1}(A))=\mu(A).

In view of Proposition 3.7, if S(X,μ)S\curvearrowright(X,\mu) and S(Y,ν)S\curvearrowright(Y,\nu) are measure-theoretically conjugate, then μ\mu is 𝐆\mathbf{G}-extensible if and only if ν\nu is.

3.3. Periodicity and extensibility

A first key observation is that periodic measures are always 𝚪\bm{\Gamma}-extensible when 𝚪\bm{\Gamma} is the free SS-group.

Proposition 3.8.

Assume that SS is an embeddable monoid, that 𝚪=(Γ,γ)\bm{\Gamma}=(\Gamma,\gamma) is a realization of the free SS-group, and let SXS\curvearrowright X be a continuous action. Then, for every μ𝒫erg(X,S)\mu\in\mathcal{P}^{\mathrm{erg}}(X,S), there exists μ¯𝒫erg(X𝚪,Γ)\bar{\mu}\in\mathcal{P}^{\mathrm{erg}}(X_{\bm{\Gamma}},\Gamma) such that πμ¯=μ\pi_{*}\bar{\mu}=\mu. In particular, 𝒫(X,S)Ext𝚪(X,S)\mathcal{P}(X,S)\subseteq\operatorname{Ext}_{\bm{\Gamma}}(X,S).

Proof.

Let μ𝒫erg(X,S)\mu\in\mathcal{P}^{\mathrm{erg}}(X,S) be given by

μ=1|Sx|ySxδy for some xPer(X,S).\mu=\frac{1}{|Sx|}\sum_{y\in Sx}\delta_{y}\quad\text{ for some }x\in\operatorname{Per}(X,S).

Consider a disjoint copy S^={s^:sS}\hat{S}=\{\hat{s}:s\in S\} of SS, with the operation t^s^=st^\hat{t}\hat{s}=\hat{st} for s^,t^S^\hat{s},\hat{t}\in\hat{S}. We define an action of S^\hat{S} on SxSx as follows: for every s^S^\hat{s}\in\hat{S} and ySxy\in Sx, we let s^y\hat{s}\cdot y to be the only element of s1({y})Sxs^{-1}(\{y\})\cap Sx. Notice that this is indeed an action, because st^y\hat{st}\cdot y is the only element of

(st)1({y})Sx=t1(s1({y}))Sx=t1(s1({y})Sx)Sx,(st)^{-1}(\{y\})\cap Sx=t^{-1}(s^{-1}(\{y\}))\cap Sx=t^{-1}(s^{-1}(\{y\})\cap Sx)\cap Sx,

and the only element of the latter is exactly t^(s^y)\hat{t}\cdot(\hat{s}\cdot y).

Since both SS and S^\hat{S} act upon SxSx, we immediately get an action SS^SxS\ast\hat{S}\curvearrowright Sx by concatenation, where SS^S\ast\hat{S} denotes the free product semigroup, where the empty word ε\varepsilon acts as the identity permutation.

We want to see the action of SS^S\ast\hat{S} descends to an action of Γ\Gamma. To do this, we define, for every s^S^\hat{s}\in\hat{S}, the element γ^(s^)=γ(s)1Γ\hat{\gamma}(\hat{s})=\gamma(s)^{-1}\in\Gamma. Note that this defines a morphism γ^:S^Γ\hat{\gamma}\colon\hat{S}\to\Gamma, since γ^(t^s^)=γ^(st^)=γ(st)1=γ(t)1γ(s)1=γ^(t^)γ^(s^)\hat{\gamma}(\hat{t}\hat{s})=\hat{\gamma}(\hat{st})=\gamma(st)^{-1}=\gamma(t)^{-1}\gamma(s)^{-1}=\hat{\gamma}(\hat{t})\hat{\gamma}(\hat{s}) for all s^,t^S^\hat{s},\hat{t}\in\hat{S}. This induces a morphism γ:SS^Γ\gamma_{*}\colon S\ast\hat{S}\to\Gamma by

γ(s1t^1snt^n)=γ(s1)γ^(t^1)γ(sn)γ^(t^n).\gamma_{*}(s_{1}\hat{t}_{1}\cdots s_{n}\hat{t}_{n})=\gamma(s_{1})\hat{\gamma}(\hat{t}_{1})\cdots\gamma(s_{n})\hat{\gamma}(\hat{t}_{n}).

Given w1,w2SS^w_{1},w_{2}\in S\ast\hat{S}, declare w1w2w_{1}\sim w_{2} if and only if w1w_{1} and w2w_{2} induce the same element in Sym(Sx)\mathrm{Sym}(Sx), i.e., if w1y=w2yw_{1}\cdot y=w_{2}\cdot y for every ySxy\in Sx. We want to check that, if γ(w1)=γ(w2)\gamma_{*}(w_{1})=\gamma_{*}(w_{2}), then w1w2w_{1}\sim w_{2}.

First, observe that if w1w2w_{1}\sim w_{2}, then for any u,vSS^u,v\in S\ast\hat{S},

uw1vy=u(w1(vy))=u(w2(vy))=uw2vy for every ySx,uw_{1}v\cdot y=u\cdot(w_{1}\cdot(v\cdot y))=u\cdot(w_{2}\cdot(v\cdot y))=uw_{2}v\cdot y\quad\text{ for every }y\in Sx,

thus making \sim a congruence on SS^S\ast\hat{S}. Next, consider the quotient semigroup SS^/S\ast\hat{S}/\sim. Take an arbitrary element w=s1t^1snt^nw=s_{1}\hat{t}_{1}\cdots s_{n}\hat{t}_{n} in SS^S\ast\hat{S} and define w^=tns^nt1s^1\hat{w}=t_{n}\hat{s}_{n}\cdots t_{1}\hat{s}_{1}. Since ss^εs^ss\hat{s}\sim\varepsilon\sim\hat{s}s for every sSs\in S, inductively we have ww^εw^ww\hat{w}\sim\varepsilon\sim\hat{w}w. This means

[w][w^]=[ww^]=[ε]=[w^w]=[w^][w].[w]_{\sim}[\hat{w}]_{\sim}=[w\hat{w}]_{\sim}=[\varepsilon]_{\sim}=[\hat{w}w]_{\sim}=[\hat{w}]_{\sim}[w]_{\sim}.

Hence, G:=SS^/G:=S\ast\hat{S}/\sim is actually a group, where the inverse of [w][w]_{\sim} is just [w^][\hat{w}]_{\sim}.

Let q:SS^Gq\colon S\ast\hat{S}\to G be the quotient map associated to \sim. The map q|S:SGq\rvert_{S}\colon S\rightarrow G is a semigroup morphism, as equal elements of SS define equal elements in Sym(Sx)\textup{Sym}(Sx). Analogously, q|S^:S^Gq\rvert_{\hat{S}}\colon\hat{S}\rightarrow G is a semigroup morphism, too. Since [s1t^1snt^n]=[s1][t1]1[sn][tn]1[s_{1}\hat{t}_{1}\cdots s_{n}\hat{t}_{n}]_{\sim}=[s_{1}]_{\sim}[t_{1}]_{\sim}^{-1}\cdots[s_{n}]_{\sim}[t_{n}]_{\sim}^{-1} for every s1t^1snt^nSS^s_{1}\hat{t}_{1}\cdots s_{n}\hat{t}_{n}\in S\ast\hat{S}, we find that (G,q|S)(G,q\rvert_{S}) is an SS-group (note that, in general, the morphism q|Sq\rvert_{S} may not be injective). By the universal property of the free SS-group, we get a group morphism θ:ΓG\theta\colon\Gamma\rightarrow G such that θγ=q|S\theta\circ\gamma=q\rvert_{S}, and it is verified that θγ^=q|S^\theta\circ\hat{\gamma}=q\rvert_{\hat{S}} as well. Therefore, θγ=q\theta\circ\gamma_{*}=q, which implies w1w2w_{1}\sim w_{2} whenever γ(w1)=γ(w2)\gamma_{*}(w_{1})=\gamma_{*}(w_{2}), as desired.

We can now define an action ΓSx\Gamma\curvearrowright Sx by setting γ(w)y=wy\gamma_{*}(w)\cdot y=w\cdot y for every wSS^w\in S\ast\hat{S} and ySxy\in Sx, as a consequence of the fact that γ(SS^)=Γ\gamma_{*}(S\ast\hat{S})=\Gamma. For a given ySxy\in Sx, let y¯=(yh)hGX𝚪\bar{y}=(y_{h})_{h\in G}\in X_{\bm{\Gamma}} be given by yh=hyy_{h}=h\cdot y for every hΓh\in\Gamma. Note that if gΓg\in\Gamma, then

gx¯=g(hx)hG=(hgx)hG=(h(gx))hG=gx¯.g\cdot\bar{x}=g\cdot(h\cdot x)_{h\in G}=(hg\cdot x)_{h\in G}=(h\cdot(g\cdot x))_{h\in G}=\overline{g\cdot x}.

This shows Γx¯{y¯:ySx}\Gamma\bar{x}\subseteq\{\bar{y}:y\in Sx\}. Also, if ySxy\in Sx, then y=sxy=s\cdot x for some sSs\in S, which implies y¯=sx¯=γ(s)x¯=γ(s)x¯\bar{y}=\overline{s\cdot x}=\overline{\gamma(s)\cdot x}=\gamma(s)\cdot\bar{x}. Notice that in this last equation there are three different actions involved. Thus, {y¯:ySx}Γx¯\{\bar{y}:y\in Sx\}\subseteq\Gamma\bar{x}. Since y¯y¯\bar{y}\neq\bar{y}^{\prime} whenever y,ySxy,y^{\prime}\in Sx differ, we have that |Γx¯|=|{y¯:ySx}|=|Sx||\Gamma\bar{x}|=|\{\bar{y}:y\in Sx\}|=|Sx|. Thus, x¯\bar{x} is Γ\Gamma-periodic, and the measure

μ¯=1|Γx¯|y¯Γx¯δy¯=1|Sx|y¯Γx¯δy¯\bar{\mu}=\frac{1}{|\Gamma\bar{x}|}\sum_{\bar{y}\in\Gamma\,\bar{x}}\delta_{\bar{y}}=\frac{1}{|Sx|}\sum_{\bar{y}\in\Gamma\,\bar{x}}\delta_{\bar{y}}

is Γ\Gamma-periodic and ergodic, as it is supported on a single orbit. Finally, note that

μ¯(π1A)=1|Sx|y¯Γx¯δy¯(π1A)=1|Sx|y¯Γx¯δπ(y¯)(A)=1|Sx|ySxδy(A)=μ(A)\bar{\mu}(\pi^{-1}A)=\frac{1}{|Sx|}\sum_{\bar{y}\in\Gamma\,\bar{x}}\delta_{\bar{y}}(\pi^{-1}A)=\frac{1}{|Sx|}\sum_{\bar{y}\in\Gamma\,\bar{x}}\delta_{\pi(\bar{y})}(A)=\frac{1}{|Sx|}\sum_{y\in Sx}\delta_{y}(A)=\mu(A)

for every A(X)A\in\mathcal{B}(X), so πμ¯=μ\pi_{*}\overline{\mu}=\mu. Hence μExt𝚪(X,S)\mu\in\textup{Ext}_{\bm{\Gamma}}(X,S).

Finally, as every μ𝒫(X,S)\mu\in\mathcal{P}(X,S) is a convex combination of elements in 𝒫erg(X,S)\mathcal{P}^{\mathrm{erg}}(X,S), π\pi_{*} is linear, and, by Proposition 3.6, Ext𝚪(X,S)\operatorname{Ext}_{\bm{\Gamma}}(X,S) is a convex subset of S(X)\mathcal{M}_{S}(X), we conclude. ∎

Example 3.9.

Note that if we remove the assumption that 𝚪\bm{\Gamma} is the free SS-group then Proposition 3.8 does not necessarily hold, as the set X𝐆X_{\mathbf{G}} might be empty for an arbitrary receiving SS-group 𝐆\mathbf{G}.

It is important to know the behaviour of images of GG-orbits via SS-equivariant maps.

Lemma 3.10.

Let 𝐆=(G,η)\mathbf{G}=(G,\eta) be a receiving SS-group, and let GX^G\curvearrowright\hat{X} and SXS\curvearrowright X be two actions. If ϕ:X^X\phi:\hat{X}\rightarrow X is an η\eta-equivariant map and x^X^\hat{x}\in\hat{X} is GG-periodic (i.e., |Gx^|<|G\hat{x}|<\infty), then ϕ(x^)\phi(\hat{x}) is SS-periodic and ϕ(Gx^)=Sϕ(x^)\phi(G\hat{x})=S\phi(\hat{x}).

Proof.

If sSs\in S, it is clear that ϕ(η(s)x^)=sϕ(x^)Sϕ(x^)\phi(\eta(s)\cdot\hat{x})=s\cdot\phi(\hat{x})\in S\phi(\hat{x}). Also, η(s)\eta(s) has finite order as an element of Sym(Gx^)\textup{Sym}(G\hat{x}), so there exists n>1n>1 such that η(sn)\eta(s^{n}) acts as the identity of Gx^G\hat{x}. Thus,

ϕ(η(s)1x^)=ϕ(η(sn1)x^)=sn1ϕ(x^)Sϕ(x^).\phi(\eta(s)^{-1}\cdot\hat{x})=\phi(\eta(s^{n-1})\cdot\hat{x})=s^{n-1}\cdot\phi(\hat{x})\in S\phi(\hat{x}).

Since G=η(S)G=\langle\eta(S)\rangle, we conclude that ϕ(Gx^)Sϕ(x^)\phi(G\hat{x})\subseteq S\phi(\hat{x}), which directly implies that ϕ(Gx^)=Sϕ(x^)\phi(G\hat{x})=S\phi(\hat{x}).

It is now clear that ϕ(x^)\phi(\hat{x}) is pre-periodic. To see that Sϕ(x^)S\phi(\hat{x}) is completely SS-invariant, let s,tSs,t\in S. Since the set {η(t)nη(s)x^:n1}\{\eta(t)^{n}\eta(s)\cdot\hat{x}:n\geq 1\} is finite, there are k>jk>j such that η(t)kη(s)x^=η(t)jη(s)x^\eta(t)^{k}\eta(s)\cdot\hat{x}=\eta(t)^{j}\eta(s)\cdot\hat{x}. The action on X^\hat{X} is given by a group, so η(s)x^=η(t)kjη(s)x^\eta(s)\cdot\hat{x}=\eta(t)^{k-j}\eta(s)\cdot\hat{x} and tkjSt^{k-j}\in S. Then, due to the η\eta-equivariance of ϕ\phi,

sϕ(x^)=t(tkj1sϕ(x^))tSϕ(x^),s\cdot\phi(\hat{x})=t\cdot(t^{k-j-1}s\cdot\phi(\hat{x}))\in tS\phi(\hat{x}),

so t:Sϕ(x^)Sϕ(x^)t\colon S\phi(\hat{x})\rightarrow S\phi(\hat{x}) is surjective, and hence bijective by finiteness of Sϕ(x^)S\phi(\hat{x}).

Proposition 3.11.

Let SXS\curvearrowright X be a continuous action, and let 𝐆\mathbf{G} be a receiving SS-group. If μ¯G(X𝐆)\bar{\mu}\in\mathcal{M}_{G}(X_{\mathbf{G}}), then supp(πμ¯)=π(supp(μ¯))\operatorname{supp}(\pi_{*}\bar{\mu})=\pi(\operatorname{supp}(\bar{\mu})). In particular, we have the following:

  1. (i)

    π(𝒫erg(X𝐆,G))𝒫erg(X,S)\pi_{*}(\mathcal{P}^{\mathrm{erg}}(X_{\mathbf{G}},G))\subseteq\mathcal{P}^{\mathrm{erg}}(X,S) and π(𝒫(X𝐆,G))𝒫(X,S)\pi_{*}(\mathcal{P}(X_{\mathbf{G}},G))\subseteq\mathcal{P}(X,S).

  2. (ii)

    If 𝒫(X𝐆,G)¯=G(X𝐆)\overline{\mathcal{P}(X_{\mathbf{G}},G)}=\mathcal{M}_{G}(X_{\mathbf{G}}), then Ext𝐆(X,S)𝒫(X,S)¯\operatorname{Ext}_{\mathbf{G}}(X,S)\subseteq\overline{\mathcal{P}(X,S)}.

  3. (iii)

    If 𝒫erg(X𝐆,G)¯=G(X𝐆)\overline{\mathcal{P}^{\mathrm{erg}}(X_{\mathbf{G}},G)}=\mathcal{M}_{G}(X_{\mathbf{G}}), then Ext𝐆(X,S)𝒫erg(X,S)¯\operatorname{Ext}_{\mathbf{G}}(X,S)\subseteq\overline{\mathcal{P}^{\mathrm{erg}}(X,S)}.

Proof.

First, we prove that supp(πμ¯)=π(supp(μ¯))\operatorname{supp}(\pi_{*}\bar{\mu})=\pi(\operatorname{supp}(\bar{\mu})). Let xsupp(πμ¯)x\in\operatorname{supp}(\pi_{*}\bar{\mu}) and fix a decreasing sequence (Un)n(U_{n})_{n} of open neighborhoods of xx with nUn={x}\bigcap_{n}U_{n}=\{x\}. Since xsupp(πμ¯)x\in\operatorname{supp}(\pi_{*}\bar{\mu}), πμ¯(Un)>0\pi_{*}\bar{\mu}(U_{n})>0 for every nn. In particular, for every nn, π1(Un)supp(μ¯)\pi^{-1}(U_{n})\cap\operatorname{supp}(\bar{\mu}) is non-empty, so there exists x¯nsupp(μ¯)\bar{x}_{n}\in\operatorname{supp}(\bar{\mu}) with π(x¯n)Un\pi(\bar{x}_{n})\in U_{n}. By compactness of X𝐆X_{\mathbf{G}}, we can take a subsequence (x¯nk)k(\bar{x}_{n_{k}})_{k} converging to x¯supp(μ¯)\bar{x}\in\operatorname{supp}(\bar{\mu}) as kk\to\infty. By continuity of π\pi, we get π(x¯nk)π(x¯)\pi(\bar{x}_{n_{k}})\to\pi(\bar{x}), but since π(x¯nk)Unk\pi(\bar{x}_{n_{k}})\in U_{n_{k}}, we also have that π(x¯nk)x\pi(\bar{x}_{n_{k}})\to x. Thus, we conclude that π(x¯)=x\pi(\bar{x})=x, so xπ(supp(μ¯))x\in\pi(\operatorname{supp}(\bar{\mu})). For the opposite inclusion, let x¯supp(μ¯)\bar{x}\in\operatorname{supp}(\bar{\mu}). If UU is an open neighborhood of π(x¯)\pi(\bar{x}), then x¯π1U\bar{x}\in\pi^{-1}U and πμ¯(U)=μ¯(π1U)>0\pi_{*}\bar{\mu}(U)=\bar{\mu}(\pi^{-1}U)>0, so π(x)supp(πμ¯)\pi(x)\in\operatorname{supp}(\pi_{*}\bar{\mu}), following the desired inclusion.

To prove (i), let μ¯𝒫(X𝐆,G)\bar{\mu}\in\mathcal{P}(X_{\mathbf{G}},G). There exist elements x1,,xmPer(X𝐆,G)x_{1},\dots,x_{m}\in\operatorname{Per}(X_{\mathbf{G}},G) such that

supp(μ¯)=i=1mGxi,\operatorname{supp}(\bar{\mu})=\bigcup_{i=1}^{m}Gx_{i},

so by the equality proven above and Lemma 3.10, we have supp(πμ¯)=i=1mπ(Gxi)=i=1mSπ(xi)\operatorname{supp}(\pi_{*}\bar{\mu})=\bigcup_{i=1}^{m}\pi(Gx_{i})=\bigcup_{i=1}^{m}S\pi(x_{i}). Taking m=1m=1 proves the ergodic case.

To prove (ii) and (iii), by (i) and weak-* continuity of π\pi_{*} we have Ext𝐆(X,S)=im(π)=π(𝒫(X𝐆,G)¯)𝒫(X,S)¯\operatorname{Ext}_{\mathbf{G}}(X,S)=\operatorname{im}(\pi_{*})=\pi_{*}\left(\overline{\mathcal{P}(X_{\mathbf{G}},G)}\right)\subseteq\overline{\mathcal{P}(X,S)}, and the same argument shows that Ext𝐆(X,S)𝒫erg(X,S)¯\operatorname{Ext}_{\mathbf{G}}(X,S)\subseteq\overline{\mathcal{P}^{\mathrm{erg}}(X,S)}. ∎

Combining the previous results together yields the following theorem.

\secondtheorem

*

Proof.

We prove the periodic case; the ergodic periodic case is identical. By Proposition 3.8, 𝒫(X,S)Ext𝚪(X,S)\mathcal{P}(X,S)\subseteq\operatorname{Ext}_{\bm{\Gamma}}(X,S). By Proposition 3.11(ii), Ext𝚪(X,S)𝒫(X,S)¯\operatorname{Ext}_{\bm{\Gamma}}(X,S)\subseteq\overline{\mathcal{P}(X,S)}. By Proposition 3.6, Ext𝚪(X,S)\operatorname{Ext}_{\bm{\Gamma}}(X,S) is weak-* closed, so Ext𝚪(X,S)=𝒫(X,S)¯\operatorname{Ext}_{\bm{\Gamma}}(X,S)=\overline{\mathcal{P}(X,S)}.

Finally, since (𝒜S)𝚪(\mathcal{A}^{S})_{\bm{\Gamma}} is topologically conjugate to 𝒜Γ\mathcal{A}^{\Gamma} for every finite alphabet 𝒜\mathcal{A}, if (i) and (ii) are satisfied, then 𝒫(𝒜S,S)¯=Ext𝚪(𝒜S,S)=S(𝒜S)\overline{\mathcal{P}(\mathcal{A}^{S},S)}=\operatorname{Ext}_{\bm{\Gamma}}(\mathcal{A}^{S},S)=\mathcal{M}_{S}(\mathcal{A}^{S}), so SS has the pa property. ∎

4. Semigroups with the pa property

In this section we aim to prove that some particular families of semigroups have the (e)pa property. The strategy will be to take an embeddable semigroup SS and its free SS-group 𝚪=(Γ,γ)\bm{\Gamma}=(\Gamma,\gamma), check that Γ\Gamma has the (e)pa property and that every measure in S(𝒜S)\mathcal{M}_{S}(\mathcal{A}^{S}) is 𝚪\bm{\Gamma}-extensible, and then appeal to Theorem Introduction to conclude.

First, we observe that, due to Proposition 2.18, a necessary condition for a semigroup to have the pa property is to be residually a finite group. In particular, every group that has the pa property must be residually finite. Thus, we will focus on embeddable semigroups SS that are residually a finite group and such that the underlying group Γ\Gamma of the free SS-group is residually finite. To check these two conditions, it suffices to check that Γ\Gamma is residually finite, as this directly implies that SS is residually a finite group. However, the converse might not be the case. An example of this would be the Baumslag–Solitar semigroup

BS(2,3)+=a,bab2=b3a,\textup{BS}(2,3)^{+}=\langle a,b\mid ab^{2}=b^{3}a\rangle,

which is residually a finite group [19, Theorem 4.5], while the corresponding Baumslag-Solitar group BS(2,3)\textup{BS}(2,3) is known to be non-Hopfian, and hence non-residually finite [4].

4.1. The left amenable case

The following result was proven in [26].

Theorem 4.1 ([26, Theorem 1.1]).

Let GG be a discrete countable residually finite amenable group acting on a compact metric space XX with specification property. Then 𝒫erg(X,G)\mathcal{P}^{\mathrm{erg}}(X,G) is dense in G(X)\mathcal{M}_{G}(X) in the weak-* topology.

Since the shift action on the full GG-shift 𝒜G\mathcal{A}^{G} trivially satisfies the specification property, this result immediately implies that residually finite amenable groups have the epa property. We want to establish an analogous result for semigroups.

A bicancellative semigroup SS is left amenable if there exists a left Følner sequence, namely a sequence (Fn)n(F_{n})_{n} of finite subsets of SS such that

limn|sFnFn||Fn|=0 for every sS.\lim_{n\to\infty}\frac{|sF_{n}\triangle F_{n}|}{|F_{n}|}=0\quad\text{ for every }s\in S.

Check [2, 16, 24] for further details.

We need to check that when SS is left amenable, residually a finite group, and 𝚪=(Γ,γ)\bm{\Gamma}=(\Gamma,\gamma) is the free SS-group, Γ\Gamma has the epa property. In order to do this, we prove that Γ\Gamma is amenable and residually finite.

Proposition 4.2 ([12, Lemma 1], [6, Proposition 4.2]).

If SS is a bicancellative and left amenable semigroup, then SS is left reversible. In addition, if 𝚪S=(ΓS,γ)\bm{\Gamma}_{S}=(\Gamma_{S},\gamma) is the group of right fractions of SS, then ΓS\Gamma_{S} is amenable.

Similarly, residual finiteness translates in the reversible case as well.

Proposition 4.3.

Let SS be a left reversible semigroup, and (ΓS,γ)(\Gamma_{S},\gamma) be the group of right fractions of SS. If SS is residually a finite group, then ΓS\Gamma_{S} is residually finite.

Proof.

Assume SS is residually a finite group. By Proposition 2.13 we may assume SHS\leq H, where HH is a residually finite group. Hence, the subgroup of HH generated by SS, together with the natural embedding ι:SH\iota\colon S\to H, is a receiving SS-group and thus isomorphic to the group of right fractions of SS; see Remark 1.3. As a consequence, ΓS\Gamma_{S} is isomorphic to a subgroup of the residually finite group HH and is thus residually finite as well. ∎

Corollary 4.4.

If SS is left amenable, residually a finite group and 𝚪=(Γ,γ)\bm{\Gamma}=(\Gamma,\gamma) is the free SS-group, then Γ\Gamma has the epa property.

Proof.

By Remark 1.3, 𝚪\bm{\Gamma} is isomorphic to 𝚪S=(ΓS,γ)\bm{\Gamma}_{S}=(\Gamma_{S},\gamma). In particular, ΓΓS\Gamma\simeq\Gamma_{S}. By Proposition 4.2 and Proposition 4.3, we know ΓS\Gamma_{S} is amenable and residually finite, hence so is Γ\Gamma. By Theorem 4.1, we conclude Γ\Gamma has the epa property. ∎

Regarding 𝚪\bm{\Gamma}-extensibility of SS-invariant measures in the reversible case, we have the following.

Theorem 4.5.

Let SXS\curvearrowright X be a continuous action. If SS is left reversible and bicancellative, and 𝚪\bm{\Gamma} is the free SS-group, then Ext𝚪(X𝚪,S)=S(X)\textup{Ext}_{\bm{\Gamma}}(X_{\bm{\Gamma}},S)=\mathcal{M}_{S}(X).

Remark 4.6.

While this work was in preparation, a version of Theorem 4.5 was announced in [15, Theorem 2.9]. For completeness, we include our own proof in the Appendix.

A consequence of the previous discussion is the following.

\thirdtheorem

*

Proof.

By Corollary 4.4, the free SS-group 𝚪=(Γ,γ)\bm{\Gamma}=(\Gamma,\gamma) is such that Γ\Gamma has the epa property. Since SS is residually a finite group, it is embeddable and, in particular, bicancellative. By Proposition 4.2, SS is left reversible. Therefore, by Theorem 4.5, Ext𝚪(X𝚪,S)=S(X)\textup{Ext}_{\bm{\Gamma}}(X_{\bm{\Gamma}},S)=\mathcal{M}_{S}(X). We conclude by appealing to Theorem Introduction. ∎

4.2. The free case

Let dd be a positive integer and consider the free group 𝔽d\mathbb{F}_{d} on the generators {a1,,ad}\{a_{1},\dots,a_{d}\}. The following result was proven in [5].

Theorem 4.7 ([5, Theorem 3.4]).

For every d1d\geq 1, the group 𝔽d\mathbb{F}_{d} has the pa property.

Given Σ{a1±1,,ad±1}\Sigma\subseteq\{a_{1}^{\pm 1},\dots,a_{d}^{\pm 1}\}, consider the subsemigroup S=Σ+S=\left<\Sigma\right>^{+}. Notice that if Σ={a1,,ad}\Sigma=\{a_{1},\dots,a_{d}\}, SS is the free semigroup 𝔽d+\mathbb{F}_{d}^{+}. If |Σ|=1|\Sigma|=1, then SS is \mathbb{N}; if Σ=Σ1\Sigma=\Sigma^{-1}, then SS is 𝔽k\mathbb{F}_{k} with 2k=|Σ|2k=|\Sigma|; otherwise, SS is a non-reversible semigroup. Without loss of generality, we will assume that Σ{a1,,ad}\Sigma\supseteq\{a_{1},\dots,a_{d}\}, and therefore k=dk=d. Denote the set {a1±1,,ad±1}\{a_{1}^{\pm 1},\dots,a_{d}^{\pm 1}\} by Σ±\Sigma^{\pm}. The free SS-group is 𝐅d=(𝔽d,ι)\mathbf{F}_{d}=(\mathbb{F}_{d},\iota) with ι:S𝔽d\iota\colon S\to\mathbb{F}_{d} the inclusion. When dealing with elements of SS as elements of 𝔽d\mathbb{F}_{d}, we will omit any mention of ι\iota.

We want to use Theorem 4.7 and appeal Theorem Introduction to establish the pa property for the semigroups S=Σ+S=\left<\Sigma\right>^{+}. It remains to show that every SS-invariant measure on the full SS-shift is 𝔽d\mathbb{F}_{d}-extensible.

Let Cay(S,Σ)\text{Cay}(S,\Sigma) be the right Cayley graph of SS with respect to Σ\Sigma. The identity 1S1_{S} will be denoted by ε\varepsilon, and the ball of radius rr with center on ε\varepsilon will be denoted by BrB_{r}. A subset TST\subseteq S will be called a tree if it induces a connected subgraph (T,E(T))(T,E(T)) in Cay(S,Σ)\text{Cay}(S,\Sigma). The root of a tree TT will be the unique element r(T)Tr(T)\in T with minimal distance to ε\varepsilon.

Fix a finite alphabet 𝒜\mathcal{A}. A Markov Σ\Sigma-tree chain is a pair (𝐩,𝐏)(\mathbf{p},\mathbf{P}) consisting of a positive probability vector 𝐩\mathbf{p} (i.e., 𝐩>𝟎\mathbf{p}>\mathbf{0} and 𝒜𝐩=1\sum_{\ell\in\mathcal{A}}\mathbf{p}_{\ell}=1) and a family 𝐏={𝐏a:aΣ}\mathbf{P}=\{\mathbf{P}^{a}:a\in\Sigma\} of real |𝒜|×|𝒜||\mathcal{A}|\times|\mathcal{A}| stochastic matrices 𝐏a\mathbf{P}^{a} (i.e., 𝐏a𝟎\mathbf{P}^{a}\geq\mathbf{0} and 𝒜𝐏k,a=1\sum_{\ell\in\mathcal{A}}\mathbf{P}^{a}_{k,\ell}=1 for all k𝒜k\in\mathcal{A}).

Definition 4.8.

Given a Markov Σ\Sigma-tree chain (𝐩,𝐏)(\mathbf{p},\mathbf{P}), a measure μ(𝒜S)\mu\in\mathcal{M}(\mathcal{A}^{S}) is said to be (𝐩,𝐏)(\mathbf{p},\mathbf{P})-Markov if, for every x𝒜Sx\in\mathcal{A}^{S} and every finite tree TST\subseteq S with root ε\varepsilon, we have

μ([x;T])=𝐩x(ε)(t,at)E(T)𝐏x(t),x(at)a.\mu([x\,;T])=\mathbf{p}_{x(\varepsilon)}\cdot\prod_{(t,at)\in E(T)}\mathbf{P}^{a}_{x(t),x(at)}.

A measure μ(𝒜S)\mu\in\mathcal{M}(\mathcal{A}^{S}) is said to be Markov if it is (𝐩,𝐏)(\mathbf{p},\mathbf{P})-Markov for some Markov Σ\Sigma-tree chain (𝐩,𝐏)(\mathbf{p},\mathbf{P}).

Proposition 4.9.

For every Markov Σ\Sigma-tree chain (𝐩,𝐏)(\mathbf{p},\mathbf{P}), there exists a unique (𝐩,𝐏)(\mathbf{p},\mathbf{P})-Markov measure on 𝒜S\mathcal{A}^{S}. Moreover, this measure is SS-invariant if and only if

  1. (1)

    𝐩\mathbf{p} is a left eigenvector of each 𝐏a\mathbf{P}^{a}, i.e., 𝐩𝐏a=𝐩\mathbf{p}\mathbf{P}^{a}=\mathbf{p} for every aΣa\in\Sigma, and

  2. (2)

    if a,a1Σa,a^{-1}\in\Sigma, then 𝐩k𝐏k,a1=𝐩𝐏,ka\mathbf{p}_{k}\mathbf{P}^{a^{-1}}_{k,\ell}=\mathbf{p}_{\ell}\mathbf{P}^{a}_{\ell,k} for every k,𝒜k,\ell\in\mathcal{A}.

Proof.

See [17, p. 240]. ∎

Proposition 4.10.

For every Markov Σ\Sigma-tree chain (𝐩,𝐏)(\mathbf{p},\mathbf{P}) that induces an SS-invariant measure μ\mu, there exists a family 𝐏^={𝐏^a:aΣ±}\hat{\mathbf{P}}=\{\hat{\mathbf{P}}^{a}:a\in\Sigma^{\pm}\} such that the Markov Σ±\Sigma^{\pm}-tree chain (𝐩,𝐏^)(\mathbf{p},\hat{\mathbf{P}}) induces an 𝔽d\mathbb{F}_{d}-invariant measure μ¯𝔽d\bar{\mu}_{\mathbb{F}_{d}} that is an 𝐅d\mathbf{F}_{d}-extension of μ\mu.

Proof.

For aΣa\in\Sigma, let 𝐏^a=𝐏a\hat{\mathbf{P}}^{a}=\mathbf{P}^{a}, and for aΣa\notin\Sigma, let 𝐏^a\hat{\mathbf{P}}^{a} be defined coordinate-wise as

𝐏^k,a=𝐩𝐩k𝐏,ka1 for k,𝒜.\hat{\mathbf{P}}^{a}_{k,\ell}=\frac{\mathbf{p}_{\ell}}{\mathbf{p}_{k}}\mathbf{P}^{a^{-1}}_{\ell,k}\quad\text{ for }k,\ell\in\mathcal{A}.

First, let’s check that (𝐩,𝐏^)(\mathbf{p},\hat{\mathbf{P}}) is a Markov Σ±\Sigma^{\pm}-tree chain. For every aΣa\not\in\Sigma, we have a1Σa^{-1}\in\Sigma and 𝐩𝐏a1=𝐩\mathbf{p}\mathbf{P}^{a^{-1}}=\mathbf{p}. Thus, for all k𝒜k\in\mathcal{A},

𝒜𝐏^k,a=𝒜𝐩𝐩k𝐏,ka1=1𝐩k𝒜𝐩𝐏,ka1=1𝐩k𝐩k=1,\sum_{\ell\in\mathcal{A}}\hat{\mathbf{P}}^{a}_{k,\ell}=\sum_{\ell\in\mathcal{A}}\frac{\mathbf{p}_{\ell}}{\mathbf{p}_{k}}\mathbf{P}^{a^{-}1}_{\ell,k}=\frac{1}{\mathbf{p}_{k}}\sum_{\ell\in\mathcal{A}}\mathbf{p}_{\ell}\mathbf{P}^{a^{-}1}_{\ell,k}=\frac{1}{\mathbf{p}_{k}}\mathbf{p}_{k}=1,

showing that 𝐏^a\hat{\mathbf{P}}^{a} is stochastic. Let’s check that (𝐩,𝐏^)(\mathbf{p},\hat{\mathbf{P}}) induces an 𝔽d\mathbb{F}_{d}-invariant Markov measure. Indeed,

  1. (1)

    if aΣa\in\Sigma, then 𝐩𝐏^a=𝐩𝐏a=𝐩\mathbf{p}\hat{\mathbf{P}}^{a}=\mathbf{p}\mathbf{P}^{a}=\mathbf{p} and, if aΣa\notin\Sigma, then a1Σa^{-1}\in\Sigma and 𝐩𝐏^a=𝐩𝐏a1=𝐩\mathbf{p}\hat{\mathbf{P}}^{a}=\mathbf{p}\mathbf{P}^{a^{-1}}=\mathbf{p};

  2. (2)

    for every a𝔽da\in\mathbb{F}_{d}, 𝐩k𝐏^k,a1=𝐩𝐏^,ka\mathbf{p}_{k}\hat{\mathbf{P}}^{a^{-1}}_{k,\ell}=\mathbf{p}_{\ell}\hat{\mathbf{P}}^{a}_{\ell,k} for every k,𝒜k,\ell\in\mathcal{A}, by construction of 𝐏^\hat{\mathbf{P}}.

Let μ¯𝔽d\bar{\mu}_{\mathbb{F}_{d}} be the 𝔽d\mathbb{F}_{d}-invariant (𝐩,𝐏^)(\mathbf{p},\hat{\mathbf{P}})-Markov measure on 𝒜𝔽d\mathcal{A}^{\mathbb{F}_{d}}. It remains to show that μ¯𝔽d\bar{\mu}_{\mathbb{F}_{d}} is an 𝐅d\mathbf{F}_{d}-extension of μ\mu. To see this, it suffices to check that πμ¯𝔽d=μ\pi_{*}\bar{\mu}_{\mathbb{F}_{d}}=\mu on a cylinder set supported upon a finite tree TST\subseteq S, where π:𝒜𝔽d𝒜S\pi\colon\mathcal{A}^{\mathbb{F}_{d}}\rightarrow\mathcal{A}^{S} is the restriction map xx|Sx\mapsto x|_{S} (see Remark 3.5; precomposing with ι\iota is the same as restricting to SS). Notice that, due to SS-invariance of πμ¯𝔽d\pi_{*}\bar{\mu}_{\mathbb{F}_{d}} and μ\mu, we can assume that TT has root ε\varepsilon. Then, it is direct that πμ¯𝔽d=μ\pi_{*}\bar{\mu}_{\mathbb{F}_{d}}=\mu, since both μ\mu and μ¯𝔽d\bar{\mu}_{\mathbb{F}_{d}} assign the same measure to every cylinder [x;T][x;T]—as presented in Definition 4.8—, because 𝐏a\mathbf{P}^{a} and 𝐏^a\hat{\mathbf{P}}^{a} coincide for aΣa\in\Sigma. ∎

\displaystyle\dotscε\displaystyle\varepsilona1\displaystyle a_{1}a2\displaystyle a_{2}a11\displaystyle a_{1}^{-1}a21\displaystyle a_{2}^{-1}
Figure 2. The Cayley graph of 𝔽2\mathbb{F}_{2} for the definition of a Markov measure on 𝒜𝔽2\mathcal{A}^{\mathbb{F}_{2}}. Dashed lines represent inverses of the generators of 𝔽2\mathbb{F}_{2}.
Remark 4.11.

A feature worth highlighting from the last proof is that, if aΣa\in\Sigma, and 𝐏a\mathbf{P}^{a} is a real |𝒜|×|𝒜||\mathcal{A}|\times|\mathcal{A}| stochastic matrix with left eigenvector 𝐩\mathbf{p}, then the matrix 𝐏a1\mathbf{P}^{a^{-1}} defined by

𝐏k,a1=𝐩𝐩k𝐏,ka for k,𝒜.\mathbf{P}^{a^{-1}}_{k,\ell}=\frac{\mathbf{p}_{\ell}}{\mathbf{p}_{k}}\mathbf{P}^{a}_{\ell,k}\quad\text{ for }k,\ell\in\mathcal{A}.

is also a real |𝒜|×|𝒜||\mathcal{A}|\times|\mathcal{A}| stochastic matrix with left eigenvector 𝐩\mathbf{p}.

We want to use the fact that all SS-invariant Markov measures are 𝐅d\mathbf{F}_{d}-extensible to show that every SS-invariant measure is 𝐅d\mathbf{F}_{d}-extensible. To do this, we consider appropriate Markovizations and recodings of these measures.

Given alphabets 𝒜,\mathcal{A},\mathcal{B}, and SS-subshifts X𝒜SX\subseteq\mathcal{A}^{S} and YSY\subseteq\mathcal{B}^{S}, we say that φ:XY\varphi:X\to Y is a sliding block code if there exists a finite subset FSF\subseteq S and a map Φ:𝒜F\Phi:\mathcal{A}^{F}\to\mathcal{B} such that φ(x)(s)=Φ((sx)|F)\varphi(x)(s)=\Phi((s\cdot x)|_{F}) for every xXx\in X and sSs\in S. A sliding block code is always continuous and SS-equivariant and, due to a straightforward generalization of Curtis-Hedlund-Lyndon theorem, these two properties characterize them. Given a finite subset FSF\subseteq S, the higher FF-block code will be the particular sliding block code ϕF:𝒜F(𝒜F)S\phi_{F}:\mathcal{A}^{F}\to(\mathcal{A}^{F})^{S} given by Φ=Id𝒜F\Phi=\mathrm{Id}_{\mathcal{A}^{F}}. It is direct to check that ϕF\phi_{F} is injective and a conjugacy between 𝒜S\mathcal{A}^{S} and ϕF(𝒜S)\phi_{F}(\mathcal{A}^{S}). In particular, ϕF(𝒜S)\phi_{F}(\mathcal{A}^{S}) is an SS-subshift.

Proposition 4.12.

Every measure μS(𝒜S)\mu\in\mathcal{M}_{S}(\mathcal{A}^{S}) is 𝐅d\mathbf{F}_{d}-extensible.

Proof.

Fix μS(𝒜S)\mu\in\mathcal{M}_{S}(\mathcal{A}^{S}) and let X=supp(μ)X=\text{supp}(\mu), which is an SS-subshift of 𝒜S\mathcal{A}^{S}. For each m0m\geq 0, we abbreviate by ϕm\phi_{m} the higher BmB_{m}-block code ϕBm:𝒜S(𝒜Bm)S\phi_{B_{m}}:\mathcal{A}^{S}\to(\mathcal{A}^{B_{m}})^{S}. Fix mm, define X|Bm={x|Bm:xX}\left.X\right|_{B_{m}}=\{x\rvert_{B_{m}}:x\in X\}, which is a finite set and will play the role of an alphabet. Set for each α,βX|Bm\alpha,\beta\in\left.X\right|_{B_{m}} and aΣa\in\Sigma,

𝐩α=μ([α;Bm]) and 𝐏α,βa=μ([α;Bm]a1[β;Bm])𝐩α.\mathbf{p}_{\alpha}=\mu([\alpha\,;B_{m}])\quad\text{ and }\quad\mathbf{P}^{a}_{\alpha,\beta}=\dfrac{\mu\big{(}[\alpha\,;B_{m}]\cap a^{-1}[\beta\,;B_{m}]\big{)}}{\mathbf{p}_{\alpha}}.

Notice that 𝐩α>0\mathbf{p}_{\alpha}>0, since [α;Bm]supp(μ)[\alpha\,;B_{m}]\cap\text{supp}(\mu)\neq\emptyset. It is clear that 𝐩\mathbf{p} is a probability vector, and for each αX|Bm\alpha\in\left.X\right|_{B_{m}},

βX|Bm𝐏α,βa=1𝐩αβX|Bmμ([α;Bm]a1[β;Bm])=1,\sum_{\beta\in\left.X\right|_{B_{m}}}\mathbf{P}^{a}_{\alpha,\beta}=\frac{1}{\mathbf{p}_{\alpha}}\sum_{\beta\in\left.X\right|_{B_{m}}}\mu\big{(}[\alpha\,;B_{m}]\cap a^{-1}[\beta\,;B_{m}]\big{)}=1,

since the sets {a1[β;Bm]:βX|Bm}\big{\{}a^{-1}[\beta;B_{m}]:\beta\in\left.X\right|_{B_{m}}\big{\}} form a partition of (X|Bm)S(\left.X\right|_{B_{m}})^{S}. Thus, 𝐏a\mathbf{P}^{a} is a stochastic matrix for all aΣa\in\Sigma. For each βX|Bm\beta\in\left.X\right|_{B_{m}}, we get

αX|Bm𝐩α𝐏α,βa=αX|Bmμ([α;Bm]a1[β;Bm])=μ(a1[β;Bm])=𝐩β,\sum_{\alpha\in\left.X\right|_{B_{m}}}\mathbf{p}_{\alpha}\mathbf{P}^{a}_{\alpha,\beta}=\sum_{\alpha\in\left.X\right|_{B_{m}}}\mu\big{(}[\alpha;B_{m}]\cap a^{-1}[\beta;B_{m}]\big{)}=\mu\big{(}a^{-1}[\beta\,;B_{m}]\big{)}=\mathbf{p}_{\beta},

so 𝐩𝐏a=𝐩\mathbf{p}\mathbf{P}^{a}=\mathbf{p}. Finally, if a,a1Σa,a^{-1}\in\Sigma, then

𝐩α𝐏α,βa=μ([α;Bm]a1[β;Bm])=μ(a[α;Bm][β;Bm])=𝐩β𝐏β,αa1,\mathbf{p}_{\alpha}\mathbf{P}^{a}_{\alpha,\beta}=\mu\big{(}[\alpha\,;B_{m}]\cap a^{-1}[\beta\,;B_{m}]\big{)}=\mu\big{(}a[\alpha\,;B_{m}]\cap[\beta\,;B_{m}]\big{)}=\mathbf{p}_{\beta}\mathbf{P}^{a^{-1}}_{\beta,\alpha},

and, by Proposition 4.9, we conclude 𝐩\mathbf{p} and {𝐏a:aΣ}\{\mathbf{P}^{a}:a\in\Sigma\} define an SS-invariant Markov measure νm\nu_{m} on (X|Bm)S(\left.X\right|_{B_{m}})^{S}.

We want to see that supp(νm)ϕm(𝒜S)\operatorname{supp}(\nu_{m})\subseteq\phi_{m}(\mathcal{A}^{S}). Notice that, if y(X|Bm)Sϕm(𝒜S)y\in(\left.X\right|_{B_{m}})^{S}-\phi_{m}(\mathcal{A}^{S}), then there exist sSs\in S and aΣa\in\Sigma such that [y(s);Bm]a1[y(as);Bm]=[y(s)\,;B_{m}]\cap a^{-1}[y(as)\,;B_{m}]=\varnothing. Indeed, if we assume otherwise, for every sSs\in S and aΣa\in\Sigma, there is an x[y(s);Bm]a1[y(as);Bm]x\in[y(s);B_{m}]\cap a^{-1}[y(as);B_{m}], so for all <m\ell<m and tBt\in B_{\ell},

y(s)(ta)=x(ta)=(ax)(t)=y(as)(t).y(s)(ta)=x(ta)=(a\cdot x)(t)=y(as)(t).

Define z𝒜Sz\in\mathcal{A}^{S} by z(s)=y(s)(ε)z(s)=y(s)(\varepsilon). Applying iteratively the identity just proven implies that, for sSs\in S and tBmt\in B_{m},

(φm(z)(s))(t)=(sz)(t)=z(ts)=y(ts)(ε)=y(s)(t),(\varphi_{m}(z)(s))(t)=(s\cdot z)(t)=z(ts)=y(ts)(\varepsilon)=y(s)(t),

so φm(z)=y\varphi_{m}(z)=y, contradicting that yϕm(𝒜S)y\not\in\phi_{m}(\mathcal{A}^{S}). Hence, [y(s);Bm]a1[y(as);Bm]=[y(s)\,;B_{m}]\cap a^{-1}[y(as)\,;B_{m}]=\varnothing for some sSs\in S and aΣa\in\Sigma, which directly implies that νm([y;{s,as}])=𝐩y(s)𝐏y(s),y(as)a=0\nu_{m}([y\,;\{s,as\}])=\mathbf{p}_{y(s)}\mathbf{P}^{a}_{y(s),y(as)}=0, so there is an open set containing yy with null measure, i.e., ysupp(νm)y\not\in\operatorname{supp}(\nu_{m}). Thus supp(νm)ϕm(𝒜S)\operatorname{supp}(\nu_{m})\subseteq\phi_{m}(\mathcal{A}^{S}).

Notice that the Markov measure νmS((X|Bm)S)\nu_{m}\in\mathcal{M}_{S}((\left.X\right|_{B_{m}})^{S}) is 𝐅d\mathbf{F}_{d}-extensible (by Proposition 4.10) and trivially conjugate to a measure νmS(ϕm(𝒜S))\nu_{m}^{\prime}\in\mathcal{M}_{S}(\phi_{m}(\mathcal{A}^{S})). Then, since 𝐅d\mathbf{F}_{d}-extensibility is preserved under factors (by Proposition 3.7), the measure νm\nu^{\prime}_{m} is 𝐅d\mathbf{F}_{d}-extensible. Let μm=(ϕm1)νm\mu_{m}=(\phi_{m}^{-1})_{*}\nu^{\prime}_{m}, i.e., μm(A)=νm(ϕm(A))\mu_{m}(A)=\nu^{\prime}_{m}(\phi_{m}(A)) for all A(𝒜S)A\in\mathcal{B}(\mathcal{A}^{S}). Since ϕm1:ϕm(𝒜S)𝒜S\phi^{-1}_{m}:\phi_{m}(\mathcal{A}^{S})\to\mathcal{A}^{S} is a conjugacy and the measure νm\nu^{\prime}_{m} is 𝐅d\mathbf{F}_{d}-extensible, the measure μm\mu_{m} is 𝐅d\mathbf{F}_{d}-extensible (again by Proposition 3.7).

We just need to prove that μmμ\mu_{m}\to\mu in the weak-* topology. If FSF\subseteq S is finite and x𝒜Fx\in\mathcal{A}^{F}, there is some m00m_{0}\geq 0 such that FBmF\subseteq B_{m} for every mm0m\geq m_{0}, so

ϕm([x;F])=ϕm(y𝒜BmF[yx;Bm])=ϕm(𝒜S)y𝒜BmF[yx;ε],\displaystyle\phi_{m}([x\,;F])=\phi_{m}\left(\bigsqcup_{y\in\mathcal{A}^{B_{m}-F}}[y\wedge x\,;B_{m}]\right)=\phi_{m}(\mathcal{A}^{S})\cap\ \bigsqcup_{y\in\mathcal{A}^{B_{m}-F}}[y\wedge x\,;\varepsilon],

whence

μm([x;F])\displaystyle\mu_{m}([x\,;F]) =ν¯m(y𝒜BmF[yx;ε])=νm(y𝒜BmF:yxX|Bm[yx;ε])\displaystyle=\overline{\nu}_{m}\left(\bigsqcup_{y\in\mathcal{A}^{B_{m}-F}}[y\wedge x\,;\varepsilon]\right)=\nu_{m}\left(\bigsqcup_{y\in\mathcal{A}^{B_{m}-F}:y\wedge x\in\left.X\right|_{B_{m}}}[y\wedge x\,;\varepsilon]\right)
=y𝒜BmF:yxX|Bm𝐩yx=y𝒜BmFμ([yx;Bm])=μ([x;F]).\displaystyle=\sum_{y\in\mathcal{A}^{B_{m}-F}:y\wedge x\in\left.X\right|_{B_{m}}}\mathbf{p}_{y\wedge x}=\sum_{y\in\mathcal{A}^{B_{m}-F}}\mu([y\wedge x\,;B_{m}])=\mu([x\,;F]).

As a result, μm([x;F])\mu_{m}([x;F]) is eventually μ([x;F])\mu([x;F]) as mm\to\infty for every cylinder, implying that μmμ\mu_{m}\to\mu in the weak-* topology.

We have the following result.

\fourththeorem

*

Proof.

By Theorem 4.7, we know 𝔽d\mathbb{F}_{d} has the pa property. By Proposition 4.12, we have that Ext𝐅d(𝒜S,S)=S(𝒜S)\textup{Ext}_{\mathbf{F}_{d}}(\mathcal{A}^{S},S)=\mathcal{M}_{S}(\mathcal{A}^{S}) for every finite alphabet 𝒜\mathcal{A}. We conclude that SS has the pa property by appealing to Theorem Introduction. ∎

4.3. Non-extensibility

Another consequence of the main result in [6] is the following:

Theorem 4.13 ([6]).

Let 𝐆\mathbf{G} be a receiving SS-group, 𝚪=(Γ,γ)\bm{\Gamma}=(\Gamma,\gamma) be the free SS-group, and assume that Γ\Gamma is residually finite. Then, 𝐆\mathbf{G} is the free SS-group if and only if every surjective continuous SS-action is topologically 𝐆\mathbf{G}-extensible.

We want to establish a measure-theoretical analog of this theorem. Observe that in the case where SS is left amenable this is trivial, as there is only one receiving SS-group up to isomorphism of SS-groups, namely, the group of right fractions 𝚪S\bm{\Gamma}_{S}, and we have already shown that every SS-invariant measure is 𝚪S\bm{\Gamma}_{S}-extensible. We focus on the free case. Let us consider the following proposition.

Proposition 4.14.

Let S(X,μ)S\curvearrowright(X,\mu) be a p.m.p. action, and let 𝐆\mathbf{G} be a receiving SS-group. Then, for every μ¯G(X𝐆)\bar{\mu}\in\mathcal{M}_{G}(X_{\mathbf{G}}) such that πμ¯=μ\pi_{*}\bar{\mu}=\mu, we have that supp(μ¯)(supp(μ))𝐆\operatorname{supp}(\bar{\mu})\subseteq(\operatorname{supp}(\mu))_{\mathbf{G}}. In particular, if Ssupp(μ)S\curvearrowright\operatorname{supp}(\mu) is not partially 𝐆\mathbf{G}-extensible, then μ\mu cannot be 𝐆\mathbf{G}-extensible.

Proof.

Note that if x¯=(xh)hGsupp(μ¯)X𝐆\bar{x}=(x_{h})_{h\in G}\in\operatorname{supp}(\bar{\mu})\subseteq X_{\mathbf{G}} and UhXU_{h}\subseteq X is an open neighborhood of xhx_{h} for hGh\in G, then h1π1Uhh^{-1}\pi^{-1}U_{h} is an open neighborhood of x¯\bar{x}, which means

μ(Uh)=μ¯(π1Uh)=μ¯(h1π1Uh)>0,\mu(U_{h})=\bar{\mu}(\pi^{-1}U_{h})=\bar{\mu}(h^{-1}\pi^{-1}U_{h})>0,

so xhsupp(μ)x_{h}\in\operatorname{supp}(\mu) for all hGh\in G. Hence supp(μ¯)X𝐆(supp(μ))G=(supp(μ))𝐆\operatorname{supp}(\bar{\mu})\subseteq X_{\mathbf{G}}\cap(\operatorname{supp}(\mu))^{G}=(\operatorname{supp}(\mu))_{\mathbf{G}}. Finally, if Ssupp(μ)S\curvearrowright\operatorname{supp}(\mu) is not partially 𝐆\mathbf{G}-extensible and μ¯G(X𝐆)\bar{\mu}\in\mathcal{M}_{G}(X_{\mathbf{G}}) is such that πμ¯=μ\pi_{*}\bar{\mu}=\mu, then (supp(μ))𝐆=(\operatorname{supp}(\mu))_{\mathbf{G}}=\varnothing, so supp(μ¯)=\operatorname{supp}(\bar{\mu})=\varnothing, which is absurd. ∎

So far, we have not provided an example of a receiving SS-group 𝐆\mathbf{G} and an SS-invariant measure which is not 𝐆\mathbf{G}-extensible. In view of Proposition 4.14, taking a continuous action SXS\curvearrowright X that is not partially 𝐆\mathbf{G}-extensible and admits an SS-invariant measure would provide an example of this. This can be done, as for every SS such that the free SS-group 𝚪=(Γ,γ)\bm{\Gamma}=(\Gamma,\gamma) satisfies that Γ\Gamma is residually finite, if 𝐆≄𝚪\mathbf{G}\not\simeq\bm{\Gamma}, there exist a finite alphabet 𝒜\mathcal{A} and an SS-periodic point x𝒜Sx\in\mathcal{A}^{S} such that SSxS\curvearrowright Sx is not partially 𝐆\mathbf{G}-extensible, so we may take the periodic measure associated with xx (see [6]).

Our last main result characterizes 𝐆\mathbf{G}-extensibility of SS-invariant measures in the symbolic context. In particular, it tells us that if 𝐆≄𝐅d\mathbf{G}\not\simeq\mathbf{F}_{d}, there is a fully-supported Markov measure on 𝒜S\mathcal{A}^{S}—which is trivially topologically 𝐆\mathbf{G}-extensible—that is not 𝐆\mathbf{G}-extensible.

\fifththeorem

*

Before proving Theorem Introduction, we need the following general result, which in particular will allow us to translate 𝐅d\mathbf{F}_{d}-extensibility to arbitrary realizations of the free SS-group.

Proposition 4.15.

Let SXS\curvearrowright X be a surjective continuous action, and let θ:𝐆𝐆\theta\colon\mathbf{G}\to\mathbf{G}^{\prime} be an isomorphism between two receiving SS-groups 𝐆=(G,η)\mathbf{G}=(G,\eta) and 𝐆=(G,η)\mathbf{G}^{\prime}=(G^{\prime},\eta^{\prime}). Then, there is a θ\theta-equivariant homeomorphism φ:X𝐆X𝐆\varphi\colon X_{\mathbf{G}^{\prime}}\to X_{\mathbf{G}} such that πφ=π\pi\circ\varphi=\pi^{\prime}. Moreover, the push-forward φ:G(X𝐆)G(X𝐆)\varphi_{*}\colon\mathcal{M}_{G^{\prime}}(X_{\mathbf{G}^{\prime}})\to\mathcal{M}_{G}(X_{\mathbf{G}}) is a bijection and, in particular, Ext𝐆(X,S)=Ext𝐆(X,S)\textup{Ext}_{\mathbf{G}}(X,S)=\textup{Ext}_{\mathbf{G}^{\prime}}(X,S).

Proof.

Define φ:X𝐆X𝐆\varphi\colon X_{\mathbf{G}^{\prime}}\to X_{\mathbf{G}} by

(xh)hG(xθ(h))hG,(x_{h^{\prime}})_{h^{\prime}\in G^{\prime}}\mapsto(x_{\theta(h)})_{h\in G},

which is well-defined, since for all sSs\in S and (xh)hGX𝐆(x_{h^{\prime}})_{h^{\prime}\in G^{\prime}}\in X_{\mathbf{G}^{\prime}} we have

sxθ(h)=xη(s)θ(h)=xθ(η(s))θ(h)=xθ(η(s)h),s\cdot x_{\theta(h)}=x_{\eta^{\prime}(s)\theta(h)}=x_{\theta(\eta(s))\theta(h)}=x_{\theta(\eta(s)h)},

so (xθ(h))hGX𝐆(x_{\theta(h)})_{h\in G}\in X_{\mathbf{G}}. Continuity of φ\varphi is easily verified, since X𝐆X_{\mathbf{G}} and X𝐆X_{\mathbf{G}^{\prime}} have the topology of pointwise convergence: for every gGg\in G there is a unique hGh\in G^{\prime} with θ(g)=h\theta(g)=h and vice versa, and thus for any sequence (x(n))n(x^{(n)})_{n} in X𝐆X_{\mathbf{G}}, convergence of x(n)x^{(n)} at the coordinate hh is equivalent to the convergence of φ(x(n))\varphi(x^{(n)}) at coordinate gg, so φ\varphi maps convergent sequences to convergent sequences. Similarly, the map X𝐆X𝐆X_{\mathbf{G}}\to X_{\mathbf{G}^{\prime}} given by (xh)hG(xθ1(h))hG(x_{h})_{h\in G}\mapsto(x_{\theta^{-1}(h^{\prime})})_{h^{\prime}\in G^{\prime}} is well-defined, so it defines an inverse for φ\varphi, which is thus bijective. The map φ1\varphi^{-1} is, as well, continuous by the same argument as above; thus, φ\varphi is a homeomorphism. Since θ(1G)=1G\theta(1_{G^{\prime}})=1_{G}, we have π=πφ\pi^{\prime}=\pi\circ\varphi and π=πφ1\pi=\pi^{\prime}\circ\varphi^{-1}. For any gGg\in G and (xh)hGX𝐆(x_{h^{\prime}})_{h^{\prime}\in G}\in X_{\mathbf{G}^{\prime}}, we have

φ(θ(g)(xh)hG)=φ((xhθ(g))hG)=(xθ(h)θ(g))hG=g(xθ(h))hG=gφ((xh)hG),\varphi(\theta(g)\cdot(x_{h^{\prime}})_{h^{\prime}\in G^{\prime}})=\varphi((x_{h^{\prime}\theta(g)})_{h^{\prime}\in G^{\prime}})=(x_{\theta(h)\theta(g)})_{h\in G}=g\cdot(x_{\theta(h)})_{h\in G}=g\cdot\varphi((x_{h})_{h\in G}),

so φ\varphi is θ\theta-equivariant.

It is readily checked that the push-forward φ:G(X𝐆)G(X𝐆)\varphi_{*}\colon\mathcal{M}_{G^{\prime}}(X_{\mathbf{G}^{\prime}})\to\mathcal{M}_{G}(X_{\mathbf{G}}) is well-defined and satisfies πφ=π\pi_{*}\circ\varphi_{*}=\pi_{*}^{\prime}. The inverse of φ\varphi_{*} is (φ1)(\varphi^{-1})_{*}. ∎

𝐯\mathbf{v^{*}}Ainεn𝐯A_{i_{n}}^{\varepsilon_{n}}\mathbf{v^{*}}Ain1εn1Ainεn𝐯A_{i_{n-1}}^{\varepsilon_{n-1}}A_{i_{n}}^{\varepsilon_{n}}\mathbf{v^{*}}Ain2εn2Ain1εn1Ainεn𝐯A_{i_{n-2}}^{\varepsilon_{n-2}}A_{i_{n-1}}^{\varepsilon_{n-1}}A_{i_{n}}^{\varepsilon_{n}}\mathbf{v^{*}}\cdot\cdot\cdotAi3ε3Ainεn𝐯A_{i_{3}}^{\varepsilon_{3}}\cdots A_{i_{n}}^{\varepsilon_{n}}\mathbf{v^{*}}Ai2ε2Ainεn𝐯A_{i_{2}}^{\varepsilon_{2}}\cdots A_{i_{n}}^{\varepsilon_{n}}\mathbf{v^{*}}
Figure 3. In this cycle in GG, transitions corresponding to solid edges all have high probability. As Ai1ε1Ainεn𝐯𝐯A_{i_{1}}^{\varepsilon_{1}}\cdots A_{i_{n}}^{\varepsilon_{n}}\mathbf{v^{*}}\neq\mathbf{v^{*}}, the final transition is thus forced to be of very low probability.
Proof of Theorem Introduction.

It follows from Proposition 4.12 together with Proposition 4.15 that (i) implies (ii). That (ii) implies (iii) is direct. To see that (iii) implies (i), assume that (i) does not hold and let us show that (iii) does not hold either. Indeed, since 𝐆≄𝐅d\mathbf{G}\not\simeq\mathbf{F}_{d}, there is a surjective morphism θ:𝔽dG\theta\colon\mathbb{F}_{d}\to G with non-trivial kernel and such that θι=η\theta\circ\iota=\eta, so we may choose an element wker(θ){ε}w\in\text{ker}(\theta)-\{\varepsilon\}. Consider a 2×22\times 2 matrix representation of 𝔽d\mathbb{F}_{d} generated by dd matrices A1,,AdA_{1},\dots,A_{d}. Write w=Ai1ε1Ainεnw=A_{i_{1}}^{\varepsilon_{1}}\cdots A_{i_{n}}^{\varepsilon_{n}} for n1n\geq 1 and εj{1,1}\varepsilon_{j}\in\{-1,1\}, j=1,,nj=1,\dots,n, and choose a sufficiently large positive prime pp such that Ai1ε1AinεnIdmodpA_{i_{1}}^{\varepsilon_{1}}\cdots A_{i_{n}}^{\varepsilon_{n}}\neq I_{d}\bmod p. Let 𝒜=p2\mathcal{A}=\mathbb{Z}_{p}^{2} and 0<δ<1/(p21)0<\delta<1/(p^{2}-1). Define, for AΣA\in\Sigma and 𝐮,𝐯𝒜\mathbf{u},\mathbf{v}\in\mathcal{A},

𝐏𝐮,𝐯A={1(p21)δif 𝐯=A𝐮modp,δotherwise.\mathbf{P}^{A}_{\mathbf{u},\mathbf{v}}=\begin{cases}1-(p^{2}-1)\delta&\text{if }\mathbf{v}=A\mathbf{u}\bmod p,\\ \delta&\text{otherwise}.\end{cases}

and let 𝐩[0,1]p2\mathbf{p}\in[0,1]^{p^{2}} be the uniform distribution upon 𝒜\mathcal{A}, i.e., 𝐩𝐮=1/p2\mathbf{p}_{\mathbf{u}}=1/p^{2} for all 𝐮𝒜\mathbf{u}\in\mathcal{A}. Since AA is invertible and pp is prime, AA is invertible modulo pp as well. Then, for a fixed 𝐯𝒜\mathbf{v}\in\mathcal{A}, there is a unique 𝐮𝒜\mathbf{u}\in\mathcal{A} such that 𝐯=A𝐮modp\mathbf{v}=A\mathbf{u}\bmod p, so

(𝐩𝐏A)𝐯=𝐮𝒜𝐩𝐮𝐏𝐮,𝐯A=(p21)δp2+1(p21)δp2=1p2,(\mathbf{p}\mathbf{P}^{A})_{\mathbf{v}}=\sum_{\mathbf{u}\in\mathcal{A}}\mathbf{p}_{\mathbf{u}}\mathbf{P}^{A}_{\mathbf{u},\mathbf{v}}=\frac{(p^{2}-1)\delta}{p^{2}}+\frac{1-(p^{2}-1)\delta}{p^{2}}=\frac{1}{p^{2}},

yielding 𝐩𝐏A=𝐩\mathbf{p}\mathbf{P}^{A}=\mathbf{p}. It is also readily checked that 𝐏A1\mathbf{P}^{A^{-1}} is the transpose of 𝐏A\mathbf{P}^{A}, which implies that 𝐩𝐮𝐏𝐮,𝐯A1=𝐩𝐯𝐏𝐯,𝐮A\mathbf{p}_{\mathbf{u}}\mathbf{P}^{A^{-1}}_{\mathbf{u},\mathbf{v}}=\mathbf{p}_{\mathbf{v}}\mathbf{P}^{A}_{\mathbf{v},\mathbf{u}} for all 𝐮,𝐯𝒜\mathbf{u},\mathbf{v}\in\mathcal{A}, since 𝐩\mathbf{p} is uniform. Thus, letting 𝐏={𝐏A:AΣ}\mathbf{P}=\{\mathbf{P}^{A}:A\in\Sigma\}, the (𝐩,𝐏)(\mathbf{p},\mathbf{P})-Markov measure μ\mu on 𝒜S\mathcal{A}^{S} is SS-invariant by Proposition 4.9. Since 𝐩>0\mathbf{p}>0 and 𝐏A>0\mathbf{P}^{A}>0 for every AΣA\in\Sigma, μ\mu is fully supported.

Assume that there is a 𝐆\mathbf{G}-extension μ¯G(𝒜G)\bar{\mu}\in\mathcal{M}_{G}(\mathcal{A}^{G}) of μ\mu. If x¯𝒜G\overline{x}\in\mathcal{A}^{G}, gGg\in G, and AΣA\in\Sigma, then

μ¯([x¯;{g,η(A)g}])\displaystyle\bar{\mu}\big{(}[\overline{x}\,;\{g,\eta(A)g\}]\big{)} =μ¯(g1[gx¯;{1G,η(A)}])\displaystyle=\bar{\mu}\big{(}g^{-1}[g\cdot\overline{x}\,;\{1_{G},\eta(A)\}]\big{)}
=μ¯([gx¯;{1G,η(A)}])\displaystyle=\bar{\mu}\big{(}[g\cdot\overline{x}\,;\{1_{G},\eta(A)\}]\big{)}
=μ([π(gx¯);{ε,A}])\displaystyle=\mu\left(\left[\pi(g\cdot\overline{x})\,;\left\{\varepsilon,A\right\}\right]\right)
=𝐩x¯(g)𝐏x¯(g),x¯(η(A)g)A,\displaystyle=\mathbf{p}_{\hskip 1.0pt\overline{x}(g)}\mathbf{P}^{A}_{\overline{x}(g),\overline{x}(\eta(A)g)},

as π(gx¯)(t)=x¯(tg)\pi(g\cdot\overline{x})(t)=\overline{x}(tg) for all tSt\in S. Analogously, we have that

μ¯([x¯;{g,η(A)1g}])=𝐩x¯(η(A)1g)𝐏x¯(η(A)1g),x¯(g)A.\bar{\mu}\big{(}[\overline{x}\,;\{g,\eta(A)^{-1}g\}]\big{)}=\mathbf{p}_{\hskip 1.0pt\overline{x}(\eta(A)^{-1}g)}\mathbf{P}^{A}_{\overline{x}(\eta(A)^{-1}g),\overline{x}(g)}.

Since Ai1ε1AinεnIdmodpA_{i_{1}}^{\varepsilon_{1}}\cdots A_{i_{n}}^{\varepsilon_{n}}\neq I_{d}\bmod p, there exists 𝐯𝒜\mathbf{v^{*}}\in\mathcal{A} such that Ai1εi1Ainεin𝐯𝐯modpA_{i_{1}}^{\varepsilon_{i_{1}}}\cdots A_{i_{n}}^{\varepsilon_{i_{n}}}\mathbf{v^{*}}\neq\mathbf{v^{*}}\bmod p. Consider the finite subset C={η(Aik)εkη(Ain)εn:1kn}C=\{\eta(A_{i_{k}})^{\varepsilon_{k}}\cdots\eta(A_{i_{n}})^{\varepsilon_{n}}:1\leq k\leq n\} of GG. Since θι=η\theta\circ\iota=\eta, we have that η(Ai1)ε1η(Ain)εn=θ(w)=1G\eta(A_{i_{1}})^{\varepsilon_{1}}\cdots\eta(A_{i_{n}})^{\varepsilon_{n}}=\theta(w)=1_{G}, so CC can be viewed as a cycle in GG. Let τ𝒜C\tau\in\mathcal{A}^{C} be a configuration such that τ(1G)=𝐯\tau(1_{G})=\mathbf{v}^{*}, and assume that τ(η(Aik)εkη(Ain)εn)=Aikεkτ(η(Aik+1)εk+1η(Ain)εn)\tau(\eta(A_{i_{k}})^{\varepsilon_{k}}\cdots\eta(A_{i_{n}})^{\varepsilon_{n}})=A_{i_{k}}^{\varepsilon_{k}}\tau(\eta(A_{i_{k+1}})^{\varepsilon_{k+1}}\cdots\eta(A_{i_{n}})^{\varepsilon_{n}}) for every 0k<n0\leq k<n, where we consider Ai0=AinA_{i_{0}}=A_{i_{n}} and ε0=εn\varepsilon_{0}=\varepsilon_{n}. Iteratively, we see that

𝐯=τ(1G)=τ(η(Ai1)ε1η(Ain)εn)=Ai1ε1Ainεn𝐯𝐯,\mathbf{v^{*}}=\tau(1_{G})=\tau(\eta(A_{i_{1}})^{\varepsilon_{1}}\cdots\eta(A_{i_{n}})^{\varepsilon_{n}})=A_{i_{1}}^{\varepsilon_{1}}\cdots A_{i_{n}}^{\varepsilon_{n}}\mathbf{v}^{*}\neq\mathbf{v}^{*},

a contradiction, so it must be the case that there exist gCg\in C and 0k<n0\leq k<n such that η(Aik)εkgC\eta(A_{i_{k}})^{\varepsilon_{k}}g\in C and τ(η(Aik)εkg)Aikεkτ(g)\tau(\eta(A_{i_{k}})^{\varepsilon_{k}}g)\neq A_{i_{k}}^{\varepsilon_{k}}\tau(g). Hence,

μ¯([τ;C])μ¯([τ;{g,η(Aik)εkg}])=𝐩τ(g)𝐏τ(g),τ(η(Aik)g)Aik=δp2\bar{\mu}([\tau;C])\leq\bar{\mu}\big{(}[\tau\,;\{g,\eta(A_{i_{k}})^{\varepsilon_{k}}g\}]\big{)}=\mathbf{p}_{\tau(g)}\mathbf{P}^{A_{i_{k}}}_{\tau(g),\tau(\eta(A_{i_{k}})g)}=\frac{\delta}{p^{2}}

if εk=1\varepsilon_{k}=1, and similarly μ¯([τ;C])δ/p2\bar{\mu}([\tau;C])\leq\delta/p^{2} if εk=1\varepsilon_{k}=-1 as well. Therefore,

1p2=μ([𝐯;ε])=μ¯([𝐯;1G])=τ𝒜C:τ(1G)=𝐯μ¯([τ;C])(p2)n1δp2=p2n4δ,\frac{1}{p^{2}}=\mu([\mathbf{v^{*}}\,;\varepsilon])=\bar{\mu}([\mathbf{v^{*}}\,;1_{G}])=\sum_{\tau\in\mathcal{A}^{C}:\tau(1_{G})=\mathbf{v^{*}}}\bar{\mu}([\tau;C])\leq(p^{2})^{n-1}\frac{\delta}{p^{2}}=p^{2n-4}\delta,

which leads to a contradiction if we take δ<1p2n2\delta<\frac{1}{p^{2n-2}}.

References

  • [1] S. I. Adian. Defining relations and algorithmic problems for groups and semigroups. Tr. Mat. Inst. Steklova, 85:3–123, 1966.
  • [2] L. N. Argabright and C. O. Wilde. Semigroups satisfying a strong Følner condition. Proc. Amer. Math. Soc., 18:587–591, 1967.
  • [3] M. Artin and B. Mazur. On periodic points. Ann. of Math., 81(1):82–99, 1965.
  • [4] G. Baumslag and D. Solitar. Some two-generator one-relator non-Hopfian groups. Bull. Amer. Math. Soc., 68(3):199–201, 1962.
  • [5] L. Bowen. Periodicity and circle packings of the hyperbolic plane. Geometriae Dedicata, 102(1):213–236, 2003.
  • [6] R. Briceño, Á. Bustos-Gajardo, and M. Donoso-Echenique. Natural extensions of embeddable semigroup actions. arXiv preprint arXiv:2501.05536.
  • [7] P. J. Burton and A. S. Kechris. Weak containment of measure-preserving group actions. Ergodic Theory Dynam. Systems, 40(10):2681–2733, 2020.
  • [8] T. Ceccherini-Silberstein and M. Coornaert. Cellular Automata and Groups. Springer, 2010.
  • [9] A. H. Clifford and G. B. Preston. The algebraic theory of semigroups. Vol. I, volume No. 7 of Mathematical Surveys. American Mathematical Society, Providence, RI, 1961.
  • [10] A. H. Clifford and G. B. Preston. The algebraic theory of semigroups. Vol. II, volume No. 7 of Mathematical Surveys. American Mathematical Society, Providence, RI, 1967.
  • [11] R. Devaney. An Introduction to Chaotic Dynamical Systems. Addison-Wesley studies in nonlinearity. Benjamin/Cummings, 1986.
  • [12] J. Donnelly. Subsemigroups of cancellative amenable semigroups. Int. J. Contemp. Math. Sciences, 7(23):1131–1137, 2012.
  • [13] P. Dubreil. Sur les problemes d’immersion et la théorie des modules. C. R. Math. Acad. Sci. Paris, 216:625–627, 1943.
  • [14] M. Einsiedler and T. Ward. Ergodic Theory: with a view towards Number Theory. Graduate Texts in Mathematics. Springer London, 2010.
  • [15] S. Farhangi, S. Jackson, and B. Mance. Undecidability in the Ramsey theory of polynomial equations and Hilbert’s tenth problem. arXiv preprint arXiv:2412.14917.
  • [16] A. H. Frey, Jr. Studies on amenable semigroups. ProQuest LLC, Ann Arbor, MI, 1960. Thesis (Ph.D.)–University of Washington.
  • [17] H.-O. Georgii. Gibbs Measures and Phase Transitions, volume 9 of De Gruyter Studies in Mathematics. Berlin, 2 edition, 2011.
  • [18] E. Glasner and B. Weiss. Kazhdan’s property T and the geometry of the collection of invariant measures. Geom. Funct. Anal., 7(5):917–935, 1997.
  • [19] D. A. Jackson. Decision and separability problems for Baumslag–Solitar semigroups. International Journal of Algebra and Computation, 12(01n02):33–49, 2002.
  • [20] Z. Ji, A. Natarajan, T. Vidick, J. Wright, and H. Yuen. MIP* = RE. Commun. ACM, 64(11):131–138, Oct. 2021.
  • [21] T. Kaiser. A closer look at the non-Hopfianness of BS(2,3)\text{BS}(2,3). Bull. Belg. Math. Soc. Simon Stevin, 28(1):147–159, 2021.
  • [22] A. S. Kechris. Weak containment in the space of actions of a free group. Isr. J. Math., 189:461–507, 2012.
  • [23] A. Malcev. On the immersion of an algebraic ring into a field. Math. Ann., 113(1):686–691, 1937.
  • [24] I. Namioka. Følner’s conditions for amenable semi-groups. Math. Scand., 15:18–28, 1964.
  • [25] O. Ore. Linear equations in non-commutative fields. Ann. of Math., 32(3):463–477, 1931.
  • [26] X. Ren. Periodic measures are dense in invariant measures for residually finite amenable group actions with specification. Discrete Contin. Dyn. Syst., 38(4):1657–1667, 2018.
  • [27] D. Ruelle. Statistical mechanics on a compact set with ν\mathbb{Z}^{\nu} action satisfying expansiveness and specification. Trans. Amer. Math. Soc., 187:237–251, 1973.
  • [28] C. Shriver. Free energy, Gibbs measures, and Glauber dynamics for nearest-neighbor interactions. Comm. Math. Phys., 398(2):679–702, 2023.
  • [29] K. Sigmund. On dynamical systems with the specification property. Trans. Amer. Math. Soc., 190:285–299, 1974.

Appendix A Proof of Theorem 4.5

Assume that SS is left reversible and bicancellative, and 𝚪\bm{\Gamma} is the free SS-group. Recall that 𝚪\bm{\Gamma} is isomorphic to 𝚪S\bm{\Gamma}_{S}, the group of right fractions of SS. Define the preorder S\leq_{S} on Γ\Gamma by

gShhg1γ(S).g\leq_{S}h\iff hg^{-1}\in\gamma(S).

Denote by (Γ)\mathcal{F}(\Gamma) the collection of all finite subsets of Γ\Gamma. We say S\leq_{S} is downward directed if for every F(Γ)F\in\mathcal{F}(\Gamma) there is an mΓm\in\Gamma such that mStm\leq_{S}t for all gFg\in F.

Lemma A.1 ([6, Lemma 2.22]).

The pre-order S\leq_{S} is downward directed. Equivalently, the subset γ(S)\gamma(S) is thick, that is, for every F(Γ)F\in\mathcal{F}(\Gamma) there exists gΓg\in\Gamma such that Fgγ(S)Fg\subseteq\gamma(S).

Let S𝛼XS\overset{\alpha}{\curvearrowright}X be a continuous action, and μS(X)\mu\in\mathcal{M}_{S}(X). Given F(Γ)F\in\mathcal{F}(\Gamma), fix any lower bound mFm_{F} for FF (i.e., such that mFStm_{F}\leq_{S}t for all tFt\in F). Note that, by definition, this means st,Fγ(S)s_{t,F}\in\gamma(S) for each tFt\in F, so that there is a unique element st,FSs_{t,F}\in S with γ(st,F)=st,F\gamma(s_{t,F})=s_{t,F}. Define the following collection of subsets of (XF)\mathcal{B}(X^{F}):

𝒞F={tFAt:At(X) for every tF},\mathcal{C}_{F}=\left\{\prod_{t\in F}A_{t}:A_{t}\in\mathcal{B}(X)\text{ for every }t\in F\right\},

and the set function μF:𝒞F[0,1]\mu_{F}\colon\mathcal{C}_{F}\rightarrow[0,1] by

μF(tFAt)=μ(tF(st,F)1(At)).\mu_{F}\left(\prod_{t\in F}A_{t}\right)=\mu\left(\bigcap_{t\in F}(s_{t,F})^{-1}(A_{t})\right).

Observe that, due to the SS-invariance of μ\mu, the value of the function μF\mu_{F} does not depend on the choice of mFm_{F}.

Lemma A.2.

The function μF\mu_{F} extends to a finitely additive probability measure on the algebra of sets 𝒜F\mathcal{A}_{F} generated by 𝒞F\mathcal{C}_{F}.

Proof.

Since 𝒜F\mathcal{A}_{F} consists of finite disjoint unions of elements of 𝒞F\mathcal{C}_{F}, we just need to check that, for any C𝒜FC\in\mathcal{A}_{F} and finite partitions 𝒫,𝒬\mathcal{P},\mathcal{Q} of CC by elements of 𝒞F\mathcal{C}_{F},

A𝒫μF(A)=B𝒬μF(B).\sum_{A\in\mathcal{P}}\mu_{F}(A)=\sum_{B\in\mathcal{Q}}\mu_{F}(B).

First, if 𝒬\mathcal{Q} is a refinement of 𝒫\mathcal{P}, take any A𝒫A\in\mathcal{P} and write it as a union of a collection {B(i)}i=1n𝒬\{B^{(i)}\}_{i=1}^{n}\subseteq\mathcal{Q}:

A=tFAt=i=1nB(i),B(i)=tFBt(i).A=\prod_{t\in F}A_{t}=\bigsqcup_{i=1}^{n}B^{(i)},\quad B^{(i)}=\prod_{t\in F}B^{(i)}_{t}.

Since this union is disjoint, for 1i<jn1\leq i<j\leq n there must be a tijFt_{ij}\in F such that Btij(i)Btij(j)=B_{t_{ij}}^{(i)}\cap B_{t_{ij}}^{(j)}=\varnothing, which implies

(stij,F)1(Btij(i))(stij,F)1(Btij(j))=.(s_{t_{ij},F})^{-1}\left(B_{t_{ij}}^{(i)}\right)\cap(s_{t_{ij},F})^{-1}\left(B_{t_{ij}}^{(j)}\right)=\varnothing.

In particular, this means

[tF(st,F)1(Bt(i))][tF(st,F)1(Bt(j))]=,\left[\bigcap_{t\in F}(s_{t,F})^{-1}\left(B_{t}^{(i)}\right)\right]\cap\left[\bigcap_{t\in F}(s_{t,F})^{-1}\left(B_{t}^{(j)}\right)\right]=\varnothing,

and therefore the collection {tF(st,F)1(Bt(i)):1in}\left\{\bigcap_{t\in F}(s_{t,F})^{-1}\left(B_{t}^{(i)}\right):1\leq i\leq n\right\} is pairwise disjoint. Hence,

μ(i=1ntF(st,F)1(Bt(i)))=i=1nμF(B(i)).\mu\left(\bigsqcup_{i=1}^{n}\bigcap_{t\in F}(s_{t,F})^{-1}\left(B_{t}^{(i)}\right)\right)=\sum_{i=1}^{n}\mu_{F}\left(B^{(i)}\right).

Now, for every 1in1\leq i\leq n and tFt\in F, Bt(i)AtB^{(i)}_{t}\subseteq A_{t}, from where we obtain

i=1ntF(st,F)1(Bt(i))tF(st,F)1(At).\bigsqcup_{i=1}^{n}\bigcap_{t\in F}(s_{t,F})^{-1}\left(B_{t}^{(i)}\right)\subseteq\bigcap_{t\in F}(s_{t,F})^{-1}(A_{t}).

We want to see that this last inclusion is an equality. Take any xtF(st,F)1(At)x\in\bigcap_{t\in F}(s_{t,F})^{-1}(A_{t}). Then, the tuple (st,Fx)tF(s_{t,F}\cdot x)_{t\in F} belongs to AA, which means it belongs to some B(i)B^{(i)}. Thus,

xtF(st,F)1(Bt(i)),x\in\bigcap_{t\in F}(s_{t,F})^{-1}\left(B^{(i)}_{t}\right),

and we get the desired opposite inclusion. Putting all together yields

μF(A)=μ(tF(st,F)1(At))=μ(i=1ntF(st,F)1(Bt(i)))=i=1nμF(B(i)).\mu_{F}(A)=\mu\left(\bigcap_{t\in F}(s_{t,F})^{-1}(A_{t})\right)=\mu\left(\bigsqcup_{i=1}^{n}\bigcap_{t\in F}(s_{t,F})^{-1}\left(B_{t}^{(i)}\right)\right)=\sum_{i=1}^{n}\mu_{F}\big{(}B^{(i)}\big{)}.

Finally, summing over 𝒫\mathcal{P}:

A𝒫μF(A)=A𝒫B𝒬BAμF(B)=B𝒬μF(B).\sum_{A\in\mathcal{P}}\mu_{F}(A)=\sum_{A\in\mathcal{P}}\sum_{\begin{subarray}{c}B\in\mathcal{Q}\\ B\subseteq A\end{subarray}}\mu_{F}(B)=\sum_{B\in\mathcal{Q}}\mu_{F}(B).

The remaining case, where 𝒬\mathcal{Q} need not be a refinement of 𝒫\mathcal{P}, follows by considering a refinement 𝒫𝒬\mathcal{P}\lor\mathcal{Q} of both 𝒫\mathcal{P} and 𝒬\mathcal{Q}. ∎

We want to prove that μF\mu_{F} is σ\sigma-additive on 𝒜F\mathcal{A}_{F}. We recall a result which will help us.

Lemma A.3.

Let ν\nu be a finite measure on an algebra 𝒜\mathcal{A} which is finitely additive and continuous at \varnothing. Then, ν\nu is σ\sigma-additive in 𝒜\mathcal{A}.

Lemma A.4.

The measure μF:𝒜F[0,1]\mu_{F}\colon\mathcal{A}_{F}\rightarrow[0,1] is continuous at \varnothing, hence σ\sigma-additive.

Proof.

Let AnA_{n}\downarrow\varnothing in 𝒜F\mathcal{A}_{F}, that is, AnAn+1A_{n}\supseteq A_{n+1} and nAn=\bigcap_{n}A_{n}=\varnothing. For each nn\in\mathbb{N} write

An=i=1kntFAt(i,n).A_{n}=\bigsqcup_{i=1}^{k_{n}}\prod_{t\in F}A_{t}^{(i,n)}.

Observe that

μF(An)\displaystyle\mu_{F}(A_{n}) =i=1knμF(tFAt(i,n))=i=1knμ(tF(st,F)1(At(i,n)))\displaystyle=\sum_{i=1}^{k_{n}}\mu_{F}\left(\prod_{t\in F}A_{t}^{(i,n)}\right)=\sum_{i=1}^{k_{n}}\mu\left(\bigcap_{t\in F}(s_{t,F})^{-1}\left(A_{t}^{(i,n)}\right)\right)
=μ(i=1kntF(st,F)1(At(i,n))).\displaystyle=\mu\left(\bigsqcup_{i=1}^{k_{n}}\bigcap_{t\in F}(s_{t,F})^{-1}\left(A_{t}^{(i,n)}\right)\right).

These last sets are decreasing in nn. Suppose there exists an element

xn[i=1kntF(st,F)1(At(i,n))].x\in\bigcap_{n}\left[\bigsqcup_{i=1}^{k_{n}}\bigcap_{t\in F}(s_{t,F})^{-1}\left(A_{t}^{(i,n)}\right)\right].

Then, for each tFt\in F define xt:=st,Fxx_{t}:=s_{t,F}\cdot x. We would have that for all n1n\geq 1 there is a 1inkn1\leq i_{n}\leq k_{n} such that xtAt(in,n)x_{t}\in A_{t}^{(i_{n},n)} for all tFt\in F, meaning

(xt)tFtFAt(in,n)i=1kntFAt(i,n)(x_{t})_{t\in F}\in\prod_{t\in F}A_{t}^{(i_{n},n)}\subseteq\bigsqcup_{i=1}^{k_{n}}\prod_{t\in F}A_{t}^{(i,n)}

for all n1n\geq 1, which contradicts the fact that AnA_{n}\downarrow\varnothing. Thus, the intersection was empty, and by continuity of μ\mu we conclude

limnμF(An)=limnμ(i=1kntF(st,F)1(At(i,n)))=0.\lim_{n\to\infty}\mu_{F}(A_{n})=\lim_{n\to\infty}\mu\left(\bigsqcup_{i=1}^{k_{n}}\bigcap_{t\in F}(s_{t,F})^{-1}\left(A_{t}^{(i,n)}\right)\right)=0.

By applying Carathéodory’s Extension Theorem, we obtain a unique extension of μF\mu_{F} to the σ\sigma-algebra σ(𝒞F)=(XF)\sigma(\mathcal{C}_{F})=\mathcal{B}(X^{F}) generated by 𝒞F\mathcal{C}_{F}.

Corollary A.5.

Let F(Γ)F\in\mathcal{F}(\Gamma) and mFStm_{F}\leq_{S}t for all tFt\in F. Then, there is a unique probability measure μF:(XF)[0,1]\mu_{F}\colon\mathcal{B}(X^{F})\rightarrow[0,1] such that

μF(tFAt)=μ(tF(st,F)1(At))\mu_{F}\left(\prod_{t\in F}A_{t}\right)=\mu\left(\bigcap_{t\in F}(s_{t,F})^{-1}(A_{t})\right)

for every tFAt(XF)\prod_{t\in F}A_{t}\in\mathcal{B}(X^{F}).

We want to extend this collection of measures to a measure on XΓX^{\Gamma}, via Kolmogorov’s Extension Theorem. If FKΓF\subseteq K\subseteq\Gamma, we define πFK:XKXF\pi_{F}^{K}\colon X^{K}\rightarrow X^{F} as the canonical projection, and we omit the super-index if K=ΓK=\Gamma. Recall that a family of measures {νF:(XF)[0,1]F(Γ)}\{\nu_{F}\colon\mathcal{B}(X^{F})\rightarrow[0,1]\mid F\in\mathcal{F}(\Gamma)\} is called consistent if whenever FK(Γ)F\subseteq K\in\mathcal{F}(\Gamma), we have (πFK)μK=μF(\pi^{K}_{F})_{*}\mu_{K}=\mu_{F}.

Lemma A.6.

The family of probability measures {μF:F(Γ)}\{\mu_{F}:F\in\mathcal{F}(\Gamma)\} is consistent.

Proof.

Let FK(Γ)F\subseteq K\in\mathcal{F}(\Gamma). Then, a lower bound mKm_{K} for KK is a lower bound for FF as well. Let tFAt\prod_{t\in F}A_{t} be an arbitrary element of 𝒞F\mathcal{C}_{F}, and define, for gKFg\in K-F, Ag:=XA_{g}:=X. Then,

(πFK)μK(tFAt)\displaystyle\left(\pi_{F}^{K}\right)_{*}\mu_{K}\left(\prod_{t\in F}A_{t}\right) =μK(gKAg)=μ(gK(sg,K)1(Ag))\displaystyle=\mu_{K}\left(\prod_{g\in K}A_{g}\right)=\mu\left(\bigcap_{g\in K}(s_{g,K})^{-1}(A_{g})\right)
=μ(tF(st,K)1(At))=μF(tFAt).\displaystyle=\mu\left(\bigcap_{t\in F}(s_{t,K})^{-1}(A_{t})\right)=\mu_{F}\left(\prod_{t\in F}A_{t}\right).

Now, the sets in (XF)\mathcal{B}(X^{F}) which satisfy the formula (πFK)μK=μF\left(\pi_{F}^{K}\right)_{*}\mu_{K}=\mu_{F} form a σ\sigma-algebra, which implies the result for all sets in (XF)\mathcal{B}(X^{F}). ∎

By Kolmogorov’s Extension Theorem, we obtain a unique probability measure μ¯Γ\bar{\mu}_{\Gamma} on (XΓ)\mathcal{B}(X^{\Gamma}) satisfying the condition μF=(πFΓ)μ¯Γ\mu_{F}=\left(\pi_{F}^{\Gamma}\right)_{*}\bar{\mu}_{\Gamma} for every finite subset FΓF\subseteq\Gamma. This allows to establish the following result.

Proposition A.7.

The measure μ¯Γ\bar{\mu}_{\Gamma} is Γ\Gamma-invariant, and we have μ¯Γ(X𝚪)=1\bar{\mu}_{\Gamma}(X_{\bm{\Gamma}})=1. Therefore, μΓ:=μ¯Γ|(X𝚪)\mu_{\Gamma}:=\bar{\mu}_{\Gamma}|_{\mathcal{B}(X_{\bm{\Gamma}})} is a Γ\Gamma-invariant probability measure satisfying πμΓ=μ\pi_{*}\mu_{\Gamma}=\mu.

Proof.

To show Γ\Gamma-invariance of μ¯Γ\bar{\mu}_{\Gamma}, let gΓg\in\Gamma and define ν=gμ¯Γ\nu=g_{*}\bar{\mu}_{\Gamma}. It suffices to check that, for a finite subset FΓF\subseteq\Gamma, μF=(πFΓ)ν\mu_{F}=\left(\pi_{F}^{\Gamma}\right)_{*}\nu on cylinders, so that by the uniqueness granted by Kolmogorov’s Extension Theorem we get ν=μ¯Γ\nu=\overline{\mu}_{\Gamma}.

It can be easily verified that, if mFm_{F} is a lower bound for FF, then so is mFgm_{F}g for FgFg. Let tFAt\prod_{t\in F}A_{t} be any element of 𝒞F\mathcal{C}_{F}, and define Ag=XA_{g}=X for gΓFg\in\Gamma-F. We obtain the following.

(πFG)ν(tFAt)\displaystyle\left(\pi_{F}^{G}\right)_{*}\nu\left(\prod_{t\in F}A_{t}\right) =μ¯Γ(g1tΓAt)=μ¯Γ(tFgAtg1)=μFg(tFgAtg1)\displaystyle=\bar{\mu}_{\Gamma}\left(g^{-1}\prod_{t\in\Gamma}A_{t}\right)=\bar{\mu}_{\Gamma}\left(\prod_{t\in Fg}A_{tg^{-1}}\right)=\mu_{Fg}\left(\prod_{t\in Fg}A_{tg^{-1}}\right)
=μ(tFg(stg1,F)1Atg1)=μ(tF(st,F)1At)=μF(tFAt).\displaystyle=\mu\left(\bigcap_{t\in Fg}(s_{tg^{-1},F})^{-1}A_{tg^{-1}}\right)=\mu\left(\bigcap_{t\in F}(s_{t,F})^{-1}A_{t}\right)=\mu_{F}\left(\prod_{t\in F}A_{t}\right).

Now we prove XGX_{G} has full measure. We already know X𝚪X_{\bm{\Gamma}} is closed, thus measurable. As it can be written as

X𝚪\displaystyle X_{\bm{\Gamma}} =tGsS{(xt)tΓXΓ:sxt=xγ(s)t},\displaystyle=\bigcap_{t\in G}\bigcap_{s\in S}\left\{(x_{t})_{t\in\Gamma}\in X^{\Gamma}:s\cdot x_{t}=x_{\gamma(s)t}\right\},

defining, for each sSs\in S and tΓt\in\Gamma, As,t={(xt)tΓXΓ:sxt=xγ(s)t}A_{s,t}=\{(x_{t})_{t\in\Gamma}\in X^{\Gamma}:s\cdot x_{t}=x_{\gamma(s)t}\} (which is also closed, by continuity of SXS\curvearrowright X and of the coordinate projections), it suffices to check that each As,tA_{s,t} has full measure. To see this, fix sSs\in S, tΓt\in\Gamma. The set XΓAs,tX^{\Gamma}-A_{s,t} is open, and can hence be written as a countable union of cylinders of XΓX^{\Gamma}:

XΓAs,t=ngΓCg(n),X^{\Gamma}-A_{s,t}=\bigcup_{n}\prod_{g\in\Gamma}C_{g}^{(n)},

where for each nn, Cg(n)=XC^{(n)}_{g}=X if gΓFng\in\Gamma-F_{n}. We must have that Ct(n)s1(Cγ(s)t(n))=C_{t}^{(n)}\cap s^{-1}\big{(}C_{\gamma(s)t}^{(n)}\big{)}=\varnothing (in particular t,γ(s)tFnt,\gamma(s)t\in F_{n} for every n1n\geq 1), and therefore, for all n1n\geq 1,

(st,Fn)1[Ct(n)s1(Cγ(s)t(n))]=(st,Fn)1(Ct(n))(sγ(s)t,Fn)1(Cγ(s)t(n))=.\big{(}s_{t,F_{n}}\big{)}^{-1}\left[C_{t}^{(n)}\cap s^{-1}\big{(}C_{\gamma(s)t}^{(n)}\big{)}\right]=\big{(}s_{t,F_{n}}\big{)}^{-1}\left(C_{t}^{(n)}\right)\cap\big{(}s_{\gamma(s)t,F_{n}}\big{)}^{-1}\big{(}C_{\gamma(s)t}^{(n)}\big{)}=\varnothing.

This last fact directly implies that

μ¯Γ(X𝚪As,t)nμFn(hFnCh(n))=nμ(hFn(sh,Fn)1(Ch(n)))=0,\displaystyle\bar{\mu}_{\Gamma}\left(X_{\bm{\Gamma}}-A_{s,t}\right)\leq\sum_{n}\mu_{F_{n}}\left(\prod_{h\in F_{n}}C_{h}^{(n)}\right)=\sum_{n}\mu\left(\bigcap_{h\in F_{n}}(s_{h,F_{n}})^{-1}\left(C_{h}^{(n)}\right)\right)=0,

obtaining μ¯Γ(As,t)=1\bar{\mu}_{\Gamma}(A_{s,t})=1, as desired.

Finally, let μΓ:=μ¯Γ|(X𝚪)\mu_{\Gamma}:=\bar{\mu}_{\Gamma}|_{\mathcal{B}(X_{\bm{\Gamma}})}. This measure is Γ\Gamma-invariant, as μΓ(A)=μ¯Γ(A)\mu_{\Gamma}(A)=\bar{\mu}_{\Gamma}(A) for every A(X𝚪)A\in\mathcal{B}(X_{\bm{\Gamma}}). Since (πFΓ)μ¯Γ=μF(\pi_{F}^{\Gamma})_{*}\bar{\mu}_{\Gamma}=\mu_{F} and π1ΓΓ(x¯)=π(x¯)\pi^{\Gamma}_{1_{\Gamma}}(\overline{x})=\pi(\overline{x}) for every x¯X𝚪\overline{x}\in X_{\bm{\Gamma}}, we have

μΓ(π1(A))=μΓ([π1ΓΓ]1(A)X𝚪)=μ¯Γ([π1ΓΓ]1(A))=μ(A)\mu_{\Gamma}(\pi^{-1}(A))=\mu_{\Gamma}\big{(}[\pi^{\Gamma}_{1_{\Gamma}}]^{-1}(A)\cap X_{\bm{\Gamma}}\big{)}=\overline{\mu}_{\Gamma}\big{(}[\pi^{\Gamma}_{1_{\Gamma}}]^{-1}(A)\big{)}=\mu(A)

for all A(X𝚪)A\in\mathcal{B}(X_{\bm{\Gamma}}), i.e., πμΓ=μ\pi_{*}\mu_{\Gamma}=\mu. ∎

Theorem 4.5 follows directly from Proposition A.7.