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

An inverse of Furstenberg’s correspondence principle and applications to van der Corput sets

Saúl Rodríguez Martín The Ohio State University
rodriguezmartin.1@osu.edu
Abstract

We obtain an inverse of Furstenberg’s correspondence principle in the setting of countable cancellative, amenable semigroups. Besides being of intrinsic interest on its own, this result allows us to answer a variety of questions concerning sets of recurrence and van der Corput (vdC) sets, which were posed by Bergelson and Lesigne [BL08], Bergelson and Ferré Moragues [BF21a], Kelly and Lê [KL18], and Moreira [Mor]. We also prove a spectral characterization of vdC sets and prove some of their basic properties in the context of countable amenable groups.

Several results in this article were independently found by Sohail Farhangi and Robin Tucker-Drob, see [FT24].

1 Introduction

In this article, we establish an inverse of Furstenberg’s correspondence principle in the framework of countable, discrete amenable semigroups. Beyond its intrinsic significance, this result enables us to answer a range of open questions posed by Bergelson and Lesigne [BL08], Bergelson and Ferré Moragues [BF21a], Kelly and Lê [KL18], and Moreira [Mor].

We first give some context for Furstenberg’s correspondence principle.

Definition 1.1.

Let GG be a countable group. A (left) Følner sequence F=(FN)NF=(F_{N})_{N\in\mathbb{N}} in GG is a sequence of finite sets FNGF_{N}\subseteq G such that, for all gGg\in G, limN|gFNΔFN||FN|=0\lim_{N\to\infty}\frac{|gF_{N}\Delta F_{N}|}{|F_{N}|}=0. We say GG is (left) amenable if it has a Følner sequence.111Throughout this article we will use only left amenability and left Følner sequences, so we will omit the adjective ‘left’.

Definition 1.2.

Let EE be a subset of a countable amenable group GG. The upper density of EE along a Følner sequence F=(FN)F=(F_{N}) is defined by

d¯F(E):=lim supN|EFN||FN|.\overline{d}_{F}(E):=\limsup_{N\to\infty}\frac{|E\cap F_{N}|}{|F_{N}|}.

If the lim sup\limsup is actually a limit, we call it dF(E)d_{F}(E), the density of EE along (FN)(F_{N}). The upper Banach density of EE, d(E)d^{*}(E), is defined by

d(E)=sup{d¯F(E);F Følner sequence in G}.d^{*}(E)=\sup\left\{\overline{d}_{F}(E);F\text{ F{\o}lner sequence in }G\right\}.

If A={1,2,}A\subseteq\mathbb{N}=\{1,2,\dots\}, we denote by d¯(A)\overline{d}(A) the upper density of AA with respect to the Følner sequence FN={1,2,,N}F_{N}=\{1,2,\dots,N\} in \mathbb{Z}.

Szemerédi’s theorem on arithmetic progressions states that any set AA of natural numbers with d¯(A)>0\overline{d}(A)>0 contains arithmetic progressions of length kk for all kk\in\mathbb{N}. Szemerédi proved this theorem in [Sze75] using combinatorial methods. In [Fur77], Furstenberg gave a new, ergodic proof of Szemerédi’s theorem, see Theorem 1.3 below. Throughout this article, we say that (X,,μ,(Ts)sS)(X,\mathcal{B},\mu,(T_{s})_{s\in S}) is a measure preserving system, m.p.s. for short, if (X,,μ)(X,\mathcal{B},\mu) is a probability space and (Ts)sS(T_{s})_{s\in S} is an action of a semigroup SS on XX by measure preserving maps Ts:XXT_{s}:X\to X (so TsTt=TstT_{s}\circ T_{t}=T_{st}). Similarly, we say (X,,μ,T)(X,\mathcal{B},\mu,T) is a m.p.s. when T:XXT:X\to X is a measure preserving map of the probability space (X,,μ)(X,\mathcal{B},\mu).

Theorem 1.3.

Let (X,,μ,T)(X,\mathcal{B},\mu,T) be a m.p.s. and let CC\in\mathcal{B} satisfy μ(C)>0\mu(C)>0. Then for all kk\in\mathbb{N} there is some nn\in\mathbb{N} such that

μ(TnCT2nCTknC)>0.\mu(T^{-n}C\cap T^{-2n}C\cap\cdots\cap T^{-kn}C)>0.

The method that Furstenberg used to derive Szemerédi’s theorem from Theorem 1.3 is nowadays called Furstenberg’s correspondence principle. We state it in the setting of amenable groups:

Theorem 1.4 (Furstenberg’s correspondence principle, cf. [Ber96, Theorem 1.8]).

Let GG be a countable amenable group with a Følner sequence F=(FN)F=(F_{N}). For any AGA\subseteq G there is a m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and BB\in\mathcal{B} with μ(B)=d¯F(A)\mu(B)=\overline{d}_{F}(A) such that, for all kk\in\mathbb{N} and h1,,hkGh_{1},\dots,h_{k}\in G,

d¯F(h1AhkA)μ(Th1(B)Thk(B)),\overline{d}_{F}(h_{1}A\cap\dots\cap h_{k}A)\geq\mu(T_{h_{1}}(B)\cap\dots\cap T_{h_{k}}(B)),

and for all h1,,hkh_{1},\dots,h_{k} such that dF(h1AhkA)d_{F}(h_{1}A\cap\dots\cap h_{k}A) exists,

dF(h1AhkA)=μ(Th1(B)Thk(B)).d_{F}(h_{1}A\cap\dots\cap h_{k}A)=\mu(T_{h_{1}}(B)\cap\dots\cap T_{h_{k}}(B)). (1)

In particular, when G=G=\mathbb{Z} and FN={1,,N}F_{N}=\{1,\dots,N\}, Theorem 1.4 and Theorem 1.3 imply the following result, which in turn implies Szemerédi’s theorem.

Theorem 1.5.

Let AA\subseteq\mathbb{N} satisfy d¯(A)>0\overline{d}(A)>0. Then for all kk\in\mathbb{N} there is some nn\in\mathbb{N} such that

d¯((An)(A2n)(Akn))>0.\overline{d}((A-n)\cap(A-2n)\cap\cdots\cap(A-kn))>0.

Equation 1 naturally leads to the question of whether given a m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}), a set BB\in\mathcal{B} and a Følner sequence FF in GG, there is some AGA\subseteq G satisfying (1) for all k,h1,,hkk,h_{1},\dots,h_{k}. The answer is yes:

Theorem 1.6 (Inverse Furstenberg correspondence principle).

Let GG be a countably infinite amenable group with a Følner sequence (FN)(F_{N}). For every m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and every BB\in\mathcal{B} there exists a subset AGA\subseteq G such that for all kk\in\mathbb{N} and h1,,hkGh_{1},\dots,h_{k}\in G we have

dF(h1AhkA)=μ(Th1BThkB).d_{F}\left(h_{1}A\cap\dots\cap h_{k}A\right)=\mu\left(T_{h_{1}}B\cap\dots\cap T_{h_{k}}B\right). (2)

In Section 3 we prove that Theorem 1.6 actually holds for cancellative amenable semigroups (see Theorem 3.20).

Remark 1.7.

A special case of Theorem 1.6, which deals with the case G=G=\mathbb{Z} and FN={1,,N}F_{N}=\{1,\dots,N\}, was obtained by Fish and Skinner in [FS24, Theorem 1.4]. Theorem 1.6 answers a question of Moreira [Mor, Section 6], which was formulated for countable abelian groups. Another special case of Theorem 1.6 is [BF21b, Theorem 5.1], where the authors assume that the action (Tg)gG(T_{g})_{g\in G} is ergodic and obtain a variant of Equation 2 by passing to a subsequence of (FN)(F_{N}). Farhangi and Tucker-Drob have independently obtained (by a different method) a version of Theorem 1.6, and its generalization, Theorem 2.7 (see [FT24, Theorem 1.2]).

Remark 1.8.

Due to the algebraic nature of Furstenberg’s correspondence principle, Theorem 1.6 admits a more general version where some of the involved sets are replaced by their complements, or some intersections are replaced by unions (a version of Furstenberg’s correspondence principle dealing with unions and complements was established in [BF21a, Theorem 2.3], see also [BBF10, Theorem 2.3]). See Theorems 3.14 and 3.15 for more details.

Our Theorem 1.6 was motivated by some open questions in the theory of van der Corput (vdC) sets, which Theorem 1.6 allows us to resolve. The notion of vdC set was introduced by Kamae and Mendès France in [KM78], in connection with the theory of uniform distribution of sequences in 𝕋=/\mathbb{T}=\mathbb{R}/\mathbb{Z}. Recall that a sequence (xn)n(x_{n})_{n\in\mathbb{N}} in 𝕋\mathbb{T} is uniformly distributed mod 11 (u.d. mod 11) if for any continuous function f:𝕋f:\mathbb{T}\to\mathbb{C} we have

limN1Nn=1Nf(xn)=𝕋f𝑑m,\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}f(x_{n})=\int_{\mathbb{T}}fdm,

where mm is Lebesgue measure.

Definition 1.9 ([KM78, Page 1]).

A set H:={1,2,}H\subseteq\mathbb{N}:=\{1,2,\dots\} is a van der Corput set (vdC set) if, for any sequence (xn)n(x_{n})_{n\in\mathbb{N}} in 𝕋=/\mathbb{T}=\mathbb{R}/\mathbb{Z} such that (xn+hxn)n(x_{n+h}-x_{n})_{n\in\mathbb{N}} is u.d. mod 11 for all hHh\in H, the sequence (xn)n(x_{n})_{n\in\mathbb{N}} is itself u.d. mod 11.

Both the notions of u.d. mod 11 and of vdC set naturally extend to more general averaging schemes and indeed to amenable semigroups, as suggested in [BL08, Section 4.2].

Definition 1.10.

Let F=(FN)F=(F_{N}) be a Følner sequence in a countable amenable group GG. We say that a sequence (xg)gG(x_{g})_{g\in G} in 𝕋\mathbb{T} is FF-u.d. mod 11 if for any continuous function f:𝕋f:\mathbb{T}\to\mathbb{\mathbb{C}} we have

limN1|FN|gFNf(xg)=𝕋f𝑑m.\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}f(x_{g})=\int_{\mathbb{T}}fdm.

For general (not necessarily abelian) groups GG, we will denote their identity element by eGe_{G}, or just ee if the group is clear from the context.

Definition 1.11 (cf. [BL08, Page 44]).

Let F=(FN)F=(F_{N}) be a Følner sequence in a countable amenable group GG. We say a subset HH of G{e}G\setminus\{e\} is FF-vdC if any sequence (xg)gG(x_{g})_{g\in G} in 𝕋\mathbb{T} such that (xhgxg)gG(x_{hg}-x_{g})_{g\in G} is FF-u.d. mod 11 for all hHh\in H, is itself FF-u.d. mod 11.

In [BL08, Section 4.2] the authors posed the question whether, for any Følner sequence FF in \mathbb{Z}, a subset HH\subseteq\mathbb{N} is FF-vdC if and only if it is a vdC set. By using an amplified version of Theorem 1.6 (3.1) we show that the answer is yes by giving the following characterization of FF-vdC sets which does not depend on the Følner sequence:

Theorem 1.12.

Let GG be a countably infinite amenable group with a Følner sequence F=(FN)F=(F_{N}). A set HG{e}H\subseteq G\setminus\{e\} is FF-vdC in GG if and only if for any m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and for any function fL(μ)f\in L^{\infty}(\mu),

Xf(Thx)f(x)¯𝑑μ(x)=0 for all hH implies Xf𝑑μ=0.\int_{X}f(T_{h}x)\cdot\overline{f(x)}d\mu(x)=0\text{ for all }h\in H\textup{ implies }\int_{X}fd\mu=0.

It is worth mentioning that the condition given in Theorem 1.12 makes sense for any countable (not necessarily amenable) group222The definition makes sense for any (discrete) group, even if it is not countable. We will not be interested in uncountable discrete groups in this article, but many of the properties of vdC sets generalize to this setting.. This leads to the following general definition:

Definition 1.13.

Let GG be a countable group. We will say that HG{e}H\subseteq G\setminus\{e\} is vdC in GG if, for every m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and every fL(μ)f\in L^{\infty}(\mu),

Xf(Thx)f(x)¯𝑑μ(x)=0 for all hH implies Xf𝑑μ=0.\int_{X}f(T_{h}x)\cdot\overline{f(x)}d\mu(x)=0\text{ for all }h\in H\text{ implies }\int_{X}fd\mu=0.

From Definition 1.13 it follows that any vdC set in a countable group GG is of (measurable) recurrence, in the sense that for every m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and for every BB\in\mathcal{B} such that μ(B)>0\mu(B)>0, we have μ(BThB)>0\mu(B\cap T_{h}B)>0 for some hHh\in H. Indeed, if instead of allowing any fL(μ)f\in L^{\infty}(\mu) we restrict our attention to characteristic functions (or positive functions), Definition 1.13 becomes the definition of set of recurrence. This supports the idea implied by [BL08, Section 3.2] that sets of recurrence are a ‘positive version’ of vdC sets.

Taking definition Definition 1.13 as the starting point, we will prove in Section 6, with the help of Theorem 1.12, several properties of vdC sets. For example, we show that the family of vdC sets in a countable group GG has the partition regularity property, that is, if HGH\subseteq G is a vdC set and H=H1H2H=H_{1}\cup H_{2}, then either H1H_{1} or H2H_{2} is a vdC set. We also prove that any vdC set in an amenable group contains two disjoint vdC sets. And finally, we study the behaviour of vdC sets in subgroups and under group homomorphisms and give some (non-)examples of vdC sets. Some of the results of Section 6 are generalizations of results which were obtained for \mathbb{Z} and d\mathbb{Z}^{d} in [Ruz81] and [BL08].

In Theorem 1.14, which will be presently formulated, we give a spectral characterization of vdC subsets of any countable abelian group GG (another, very elegant, proof of the result can be found in [FT24]). This provides a generalization of similar results obtained in [KM78, BL08, Ruz81] for vdC subsets of \mathbb{Z} or d\mathbb{Z}^{d} for d2d\geq 2 (see e.g. [BL08, Theorem 1.8.]). Spectral characterizations (in the setup of abelian groups) are useful both for proving properties of vdC sets and for finding (non-)examples of them. For example, Bourgain used in [Bou87] a version of Theorem 1.14 for G=G=\mathbb{Z} to construct a set of recurrence which is not a vdC set.

Given a discrete, countable abelian group GG we denote by 0 (instead of ee) the identity element of GG and by G^\widehat{G} its Pontryagin dual (which is compact and metrizable, see [Rud62, Theorems 1.2.5, 2.2.6]), with identity 1G^1_{\widehat{G}}. For any Borel probability measure μ\mu in G^\widehat{G} we denote the Fourier coefficients of μ\mu by μ^(h)=G^γ(h)𝑑μ(γ),hG\widehat{\mu}(h)=\int_{\widehat{G}}\gamma(h)d\mu(\gamma),h\in G.

Theorem 1.14 (cf. [BL08, Theorem 1.8]).

Let GG be a countable abelian group. A set HG{0}H\subseteq G\setminus\{0\} is a vdC set in GG iff any Borel probability measure μ\mu in G^\widehat{G} with μ^(h)=0hH\widehat{\mu}(h)=0\;\forall h\in H satisfies μ({0})=0\mu(\{0\})=0.

The structure of the article is as follows. In Section 2 we use Theorem 1.6 (and a more general version of it) to answer some questions from [BL08, BF21a, Mor]. In Section 3 we first prove 3.1, an amplified version of Theorem 1.6. We also obtain Theorem 3.20, a general version of Theorem 1.6 for cancellative amenable semigroups. In Section 4 we prove Theorem 4.1, an amplified version of Theorem 1.12 which contains several characterizations of vdC sets. In Section 5 we establish a spectral characterization of vdC sets in countable abelian groups, Theorem 1.14. In Section 6 we prove fundamental properties of vdC sets in amenable groups, such as for example partition regularity. The main result in Section 7 is 7.1, which concerns the relationship between the set of possible Cesaro averages of sequences taking values in a compact set DD\subseteq\mathbb{C}, and in the convex hull of DD. We deduce several results from 7.1, including an affirmative answer to a question of Kelly and Lê.

Acknowledgements.

Thanks to Vitaly Bergelson for his guidance while writing this article, and for some interesting discussions and suggestions.

Several months before uploading this article, it came to our attention that Sohail Farhangi and Robin Tucker-Drob had been independently studying the topic of vdC sets. Their paper [FT24] contains a long list of characterizations of vdC sets, including the ones from Theorem 1.12 and Theorem 1.14.

We appreciate Farhangi’s input, suggestions, and quick review of the paper. He brought [DHZ19, Theorem 5.2] to our attention, which allowed us to simplify the proof of Theorem 1.14 and to state 3.1 for all amenable groups instead of only monotileable ones. He also noticed how 7.1 can be used to answer a question of Kelly and Lê.

We gratefully acknowledge support from the grants BSF 2020124 and NSF CCF AF 2310412.

2 Some applications of the inverse correspondence principle

In this section we use Theorem 1.6 (and a more general result, Theorem 2.7), to answer several questions from the literature. We first answer333S. Farhangi proved in his dissertation [Far22] that every nice vdC set is a set of nice recurrence for the Følner sequence FN={1,,N}F_{N}=\{1,\dots,N\}, thereby addressing Bergelson and Lesigne’s original question. We give a different proof and generalize this result to amenable groups. a question of Bergelson and Lesigne in the general context of countable amenable groups, by proving that every nice vdC set is a set of nice recurrence. The second of these two notions was introduced in [Ber86] for subsets of \mathbb{Z}, although we use the slightly different definition given in [BL08] (we check that both are equivalent in 2.10).

Definition 2.1 (cf. [Ber86, Definition 2.2]).

Let GG be a group. A subset HH of GG is a set of nice recurrence if for any m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and for any BB\in\mathcal{B},

μ(B)2lim suphHμ(BThB).\mu(B)^{2}\leq\limsup_{h\in H}\mu(B\cap T_{h}B).

In order to motivate the definition of nice vdC sets, we first state a theorem of Ruzsa which characterizes vdC sets in terms of Cesaro averages:

Theorem 2.2 (cf. [Ruz81, Theorem 1]).

A set HH\subseteq\mathbb{N} is a vdC set iff for any sequence (zn)n(z_{n})_{n\in\mathbb{N}} of complex numbers with |zn|1|z_{n}|\leq 1 for all nn,

limN1Nn=1Nzh+nzn¯=0 for all hH implies limN1Nn=1Nzn=0.\quad\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}z_{h+n}\overline{z_{n}}=0\textup{ for all }h\in H\textup{ implies }\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}z_{n}=0.
Definition 2.3 (cf. [BL08, Definition 10]).

Let GG be a countable amenable group with a Følner sequence (FN)(F_{N}). A subset HH of G{e}G\setminus\{e\} is nice FF-vdC if for any sequence (zg)gG(z_{g})_{g\in G} in 𝔻\mathbb{D},

lim supN|1|FN|gFNzg|2lim suphHlim supN|1|FN|gFNzhgzg¯|\limsup_{N}\left|\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{g}\right|^{2}\leq\limsup_{h\in H}\limsup_{N}\left|\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{hg}\overline{z_{g}}\right|

Theorem 2.2 implies that in \mathbb{Z} (or \mathbb{N}), nice vdC sets are vdC sets. This result is true for any amenable group, as implied by Theorem 1.12 and 2.5 below. Note that Definition 2.3 and Definition 2.1 are expressed in different settings: Definition 2.3 is about Cesaro averages of sequences of complex numbers, while Definition 2.1 is about integrals. We now translate each of these definitions to the setting of the other one:

Proposition 2.4.

Let GG be a countable amenable group with a Følner sequence F=(FN)F=(F_{N}). Then a subset HGH\subseteq G is a set of nice recurrence iff for any EGE\subseteq G we have

dF¯(E)2lim suphHdF¯(EhE).\overline{d_{F}}(E)^{2}\leq\limsup_{h\in H}\overline{d_{F}}(E\cap hE).
Proposition 2.5.

Let GG be a countable amenable group with a Følner sequence (FN)(F_{N}). Then a subset HH of G{e}G\setminus\{e\} is nice FF-vdC iff for any m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and any fL(X,μ)f\in L^{\infty}(X,\mu) we have

|Xf𝑑μ|2lim suphH|Xf(Thx)f(x)¯𝑑μ(x)|.\left|\int_{X}fd\mu\right|^{2}\leq\limsup_{h\in H}\left|\int_{X}f(T_{h}x)\overline{f(x)}d\mu(x)\right|.
Remark 2.6.

2.5 implies that the notion of nice FF-vdC set is independent of the Følner sequence.

Proof of 2.4.
  1. \implies

    Suppose there is some set EGE\subseteq G such that dF¯(E)2>lim suphHdF¯(EhE)\overline{d_{F}}(E)^{2}>\limsup_{h\in H}\overline{d_{F}}(E\cap hE). Then there is a Følner subsequence FF^{\prime} of FF such that dF(E)d_{F^{\prime}}(E) exists, dF(EhE)d_{F^{\prime}}(E\cap hE) exists for all hh and dF(E)2>lim suphHdF(EhE)d_{F^{\prime}}(E)^{2}>\limsup_{h\in H}d_{F^{\prime}}(E\cap hE). But by Theorem 1.4 there is some m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and some BB\in\mathcal{B} such that μ(B)=dF(E)\mu(B)=d_{F^{\prime}}(E) and μ(BThB)=dF(EhE)\mu(B\cap T_{h}B)=d_{F^{\prime}}(E\cap hE) for all hGh\in G, thus μ(B)2>lim suphHμ(BThB)\mu(B)^{2}>\limsup_{h\in H}\mu(B\cap T_{h}B).

  2. \impliedby

    Suppose there is a m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and some BB\in\mathcal{B} such that μ(B)2>lim suphHμ(BThB)\mu(B)^{2}>\limsup_{h\in H}\mu(B\cap T_{h}B). By Theorem 1.6 there is some set EGE\subseteq G such that μ(B)=dF(E)\mu(B)=d_{F}(E) and μ(BThB)=dF(EhE)\mu(B\cap T_{h}B)=d_{F}(E\cap hE) for all hGh\in G, thus dF(E)2>lim suphHdF(EhE)d_{F^{\prime}}(E)^{2}>\limsup_{h\in H}d_{F^{\prime}}(E\cap hE).∎

The proof of 2.5 is completely analogous to that of 2.4, except that we will need to use a result slightly more general than Theorem 1.6. Note that the sets A,BA,B from Theorem 1.6 can be identified with their characteristic functions, that is, {0,1}\{0,1\}-valued functions. Theorem 1.6 corresponds to the specific case of Theorem 2.7 corresponding to D={0,1}D=\{0,1\} and functions pp of the form p(z1,,zn)=i=1nzip(z_{1},\dots,z_{n})=\prod_{i=1}^{n}z_{i}.

Theorem 2.7.

Let GG be a countably infinite amenable group with a Følner sequence (FN)(F_{N}) and let DD\subseteq\mathbb{C} be compact. Then for any m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and any measurable function f:XDf:X\to D there is a sequence (zg)gG(z_{g})_{g\in G} of complex numbers in DD such that, for all jj\in\mathbb{N}, h1,,hjGh_{1},\dots,h_{j}\in G and all continuous functions p:Djp:D^{j}\to\mathbb{C},

limN1|FN|gFNp(zh1g,,zhjg)=Xp(f(Th1x),,f(Thjx))𝑑μ.\lim_{N}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}p(z_{h_{1}g},\dots,z_{h_{j}g})=\int_{X}p(f(T_{h_{1}}x),\dots,f(T_{h_{j}}x))d\mu. (3)

Conversely, given a sequence (zg)gG(z_{g})_{g\in G} in DD there is a m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and a measurable function f:XDf:X\to D such that, for all j,h1,,hjGj\in\mathbb{N},h_{1},\dots,h_{j}\in G and p:Djp:D^{j}\to\mathbb{C} continuous, Equation 3 holds if the left hand side limit exists.

Theorem 2.7 pertains to sequences with values in a compact subset of \mathbb{C}, but in some cases one may adapt it to unbounded sequences, see [FT24, Theorem 3.3].

Proof of 2.5.
  1. \implies

    Suppose for contradiction that there is a m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and some measurable function f:X𝔻f:X\to\mathbb{D} such that

    |Xf𝑑μ|2>lim suphH|Xf(Thx)f(x)¯𝑑μ(x)|.\left|\int_{X}fd\mu\right|^{2}>\limsup_{h\in H}\left|\int_{X}f(T_{h}x)\overline{f(x)}d\mu(x)\right|.

    By Theorem 2.7 there is some sequence (zg)gG(z_{g})_{g\in G} such that

    limN1|FN|gFNzg\displaystyle\lim_{N}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{g} =Xf(x)𝑑μ\displaystyle=\int_{X}f(x)d\mu
    limN1|FN|gFNzhgzg¯\displaystyle\lim_{N}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{hg}\overline{z_{g}} =Xf(Thx)f(x)¯𝑑μ for all hH,\displaystyle=\int_{X}f(T_{h}x)\overline{f(x)}d\mu\textup{ for all }h\in H,

    so HH is not a nice FF-vdC set.

  2. \impliedby

    If HH is not nice FF-vdC then for an adequate sequence (zg)gG(z_{g})_{g\in G} in 𝔻\mathbb{D} and some subsequence F=(FN)NF^{\prime}=(F_{N}^{\prime})_{N\in\mathbb{N}} of FF we have

    limN|1|FN|gFNzg|2>lim suphHlimN|1|FN|gFNzhgzg¯|.\lim_{N}\left|\frac{1}{|F_{N}^{\prime}|}\sum_{g\in F_{N}^{\prime}}z_{g}\right|^{2}>\limsup_{h\in H}\lim_{N}\left|\frac{1}{|F_{N}^{\prime}|}\sum_{g\in F_{N}^{\prime}}z_{hg}\overline{z_{g}}\right|.

    Applying again Theorem 2.7, we obtain a m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and some measurable function f:X𝔻f:X\to\mathbb{D} such that

    |Xf𝑑μ|2>lim suphH|Xf(Thx)f(x)¯𝑑μ(x)|.\left|\int_{X}fd\mu\right|^{2}>\limsup_{h\in H}\left|\int_{X}f(T_{h}x)\overline{f(x)}d\mu(x)\right|.\qed

In [BL08, Question 8] it is asked whether there is any implication between the notions ‘nice vdC set’ and ‘set of nice recurrence’. It was also proved in [BL08] that if HH is a nice vdC set, then HH satisfies the following, weak version of nice recurrence: for any m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and for any BB\in\mathcal{B}, we have μ(B)4lim suphHμ(BThB)\mu(B)^{4}\leq\limsup_{h\in H}\mu(B\cap T_{h}B). Both 2.5 and 2.4 imply that nice vdC sets are sets of nice recurrence in any amenable group.

We now recall a combinatorial version of sets of nice recurrence:

Definition 2.8.

Let GG be a countable amenable group with a Følner sequence F=(FN)F=(F_{N}). A set HGH\subseteq G is nicely FF-intersective if for all EGE\subseteq G and ε>0\varepsilon>0 there is some gRg\in R such that

d¯F(EgE)>d¯F(E)2ε.\overline{d}_{F}(E\cap gE)>\overline{d}_{F}(E)^{2}-\varepsilon.

In his blog post [Mor, Section 6], Moreira asked whether, for a given Følner sequence FF, every subset of the natural numbers is of nice recurrence iff it is nicely FF-intersective. Fish and Skinner very recently established this result in [FS24, Theorem 1.3] for the Følner sequence FN={1,,N}F_{N}=\{1,\dots,N\} but left open the question of whether it holds for all Følner sequences in \mathbb{N}. We generalize the result to Følner sequences in amenable groups:

Proposition 2.9.

For any Følner sequence F=(FN)F=(F_{N}) in a countable amenable group GG, a set HG{e}H\subseteq G\setminus\{e\} is of nice recurrence iff it is nicely FF-intersective.

2.9 can be proved using Furstenberg’s correspondence principle, once we check the following characterization of sets of nice recurrence (a general version of [BL08, Proposition 3.8]):

Lemma 2.10.

Let GG be a countable group. A set HG{e}H\subseteq G\setminus\{e\} is of nice recurrence iff for any m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}), BB\in\mathcal{B} and ε>0\varepsilon>0 there exists hHh\in H such that μ(BThB)>μ(B)2ε\mu(B\cap T_{h}B)>\mu(B)^{2}-\varepsilon.

Proof.

The forward implication is clear. So suppose HG{e}H\subseteq G\setminus\{e\} is not of nice recurrence. That means that there is a m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}), BB\in\mathcal{B} and ε>0\varepsilon>0 such that for all hHh\in H except finitely many elements h1,,hjh_{1},\dots,h_{j} we have μ(BThB)<μ(B)2ε\mu(B\cap T_{h}B)<\mu(B)^{2}-\varepsilon.

Now consider the uniform Bernoulli shift (Y,𝒞,ν,(Sg)gG)(Y,\mathcal{C},\nu,(S_{g})_{g\in G}), where YY is the product {0,1,2,,j}G\{0,1,2,\dots,j\}^{G} and Sh((xg)gG)=(Sgh)gGS_{h}((x_{g})_{g\in G})=(S_{gh})_{g\in G}. Letting ee be the identity in GG, the set C:={(xg)Y;xe=0 and xhi=i for i=1,,j}𝒞C:=\{(x_{g})\in Y;x_{e}=0\textup{ and }x_{h_{i}}=i\textup{ for }i=1,\dots,j\}\in\mathcal{C} satisfies that ν(CSgC){0,μ(C)2}\nu(C\cap S_{g}C)\in\{0,\mu(C)^{2}\} for all gG{e}g\in G\setminus\{e\}, and in particular ν(CShiC)=0\nu(C\cap S_{h_{i}}C)=0 for i=1,,ji=1,\dots,j. Thus, in the product m.p.s. (X×Y,×𝒞,μ×ν,(Tg×Sg)gG)(X\times Y,\mathcal{B}\times\mathcal{C},\mu\times\nu,(T_{g}\times S_{g})_{g\in G}), we have (μ×ν)((B×C)(Th×Sh)(BC))<ν(C)2(μ(B)2ε)(\mu\times\nu)\,((B\times C)\cap(T_{h}\times S_{h})(B\cap C))<\nu(C)^{2}(\mu(B)^{2}-\varepsilon) for all hHh\in H, concluding the proof. ∎

Proof of 2.9.

If HG{e}H\subseteq G\setminus\{e\} is not of nice recurrence, by 2.10 there is some m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}), BB\in\mathcal{B} and ε>0\varepsilon>0 such that, for all hHh\in H, μ(BThB)<μ(B)2ε\mu(B\cap T_{h}B)<\mu(B)^{2}-\varepsilon. Thus, by Theorem 1.6 there is some subset EE of GG such that dF(E)=μ(B)d_{F}(E)=\mu(B) and for all hHh\in H, dF(EhE)=μ(BThB)<dF(E)2εd_{F}(E\cap hE)=\mu(B\cap T_{h}B)<d_{F}(E)^{2}-\varepsilon. So HH is not a nicely intersective set. The other implication can be proved similarly, using the usual Furstenberg correspondence principle instead of the inverse one. ∎

Bergelson and Moragues asked in [BF21a] (after Remark 3.6) whether for all countable amenable groups GG and all Følner sequences FF in GG there is a set EGE\subseteq G such that d¯F(E)>0\overline{d}_{F}(E)>0 but for all finite AGA\subseteq G, d¯F(gAg1E)<34\overline{d}_{F}\left(\cup_{g\in A}g^{-1}E\right)<\frac{3}{4}. Well, if we apply a version of Theorem 1.6 with unions instead of intersections (Theorem 3.15) to a m.p.s. with Tg=IdXT_{g}=\textup{Id}_{X} for all gg and μ(B)=12\mu(B)=\frac{1}{2}, we obtain the following:

Proposition 2.11.

Let GG be a countably infinite amenable group with a Følner sequence FF. Then there is a set EGE\subseteq G such that, for all finite AG\varnothing\neq A\subseteq G,

dF(E)=dF(gAgE)=12.d_{F}(E)=d_{F}\left(\cup_{g\in A}gE\right)=\frac{1}{2}.\qed

3 Correspondence between Cesaro and integral averages

The main objective of this section is proving Theorem 2.7. It will be deduced from 3.1, a more technical version of Theorem 2.7 which also includes a finitistic criterion for the existence of sequences with given Cesaro averages. At the end of the section we prove Theorem 3.14, a general converse of the Furstenberg correspondence principle, and Theorem 3.20, a version of Theorem 1.6 for semigroups.

Proposition 3.1.

Let GG be a countably infinite amenable group with a Følner sequence (FN)(F_{N}), and let DD\subseteq\mathbb{C} be compact. For each ll\in\mathbb{N} let jlj_{l}\in\mathbb{N}, hl,1,,hl,jlGh_{l,1},\dots,h_{l,j_{l}}\in G and let pl:Djlp_{l}:D^{j_{l}}\to\mathbb{C} be continuous. Finally, let γ:\gamma:\mathbb{N}\to\mathbb{C} be a sequence of complex numbers. The following are equivalent:

  1. 1.

    There exists a sequence (zg)gG(z_{g})_{g\in G} of elements of DD such that, for all ll\in\mathbb{N},

    limN1|FN|gFNpl(zhl,1g,,zhl,jlg)=γ(l).\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}p_{l}(z_{h_{l,1}g},\dots,z_{h_{l,j_{l}}g})=\gamma(l). (4)
  2. 2.

    There exists a m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and a measurable function f:XDf:X\to D such that, for all ll\in\mathbb{N},

    Xpl(f(Thl,1x),,f(Thl,jlx)dμ(x)=γ(l).\int_{X}p_{l}(f(T_{h_{l,1}}x),\dots,f(T_{h_{l,j_{l}}}x)d\mu(x)=\gamma(l). (5)
  3. 3.

    (Finitistic criterion) For all AGA\subseteq G finite and for all L,δ>0L\in\mathbb{N},\delta>0 there exist some KK\in\mathbb{N} and sequences (zg,k)gG(z_{g,k})_{g\in G} in DD, for k=1,,Kk=1,\dots,K, such that for all l=1,,Ll=1,\dots,L we have

    |γ(l)1K|A|k=1KgApl(zhl,1g,k,,zhl,jlg,k)|<δ.\left|\gamma(l)-\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{g\in A}p_{l}(z_{h_{l,1}g,k},\dots,z_{h_{l,j_{l}}g,k})\right|<\delta. (6)
Remark 3.2.

For the equivalence 1\iff3 to hold DD need not be compact, and pl:Djlp_{l}:D^{j_{l}}\to\mathbb{C} can be any bounded function (not necessarily continuous). Indeed, all we will use in the proof of 3\implies1 is that the functions plp_{l} are bounded, and it is not hard to show 1\implies3 if the functions plp_{l} are bounded: suppose that (zg)(z_{g}) satisfies 1 for some Følner sequence (FN)(F_{N}). Then one can check that given L,δ>0L\in\mathbb{N},\delta>0, for big enough NN\in\mathbb{N}, the constant K=|FN|K=|F_{N}| and the sequences (zg,k)g(z_{g,k})_{g} given by zg,k=zgkz_{g,k}=z_{gk}, for kFNk\in F_{N}, satisfy Item 3.

Proof of Theorem 2.7 from 3.1.

Let G,(FN),(X,,μ,(Tg)gG),G,(F_{N}),(X,\mathcal{B},\mu,(T_{g})_{g\in G}), DD and f:XDf:X\to D be as in Theorem 2.7.

Now, for each kk\in\mathbb{N}, the set C(Dk):={p:Dk;p continuous}C(D^{k}):=\{p:D^{k}\to\mathbb{C};p\text{ continuous}\} is separable in the supremum norm. So we can consider for each ll\in\mathbb{N} elements hl,1,,hl,jlh_{l,1},\dots,h_{l,j_{l}} and functions plC(Djl)p_{l}\in C(D^{j_{l}}) such that for any h1,,hjGh_{1},\dots,h_{j}\in G, for any pC(Dj)p\in C(D^{j}) continuous and for any ε>0\varepsilon>0 there exists ll\in\mathbb{N} such that jl=jj_{l}=j, (hl,1,,hl,jl)=(h1,,hj)(h_{l,1},\dots,h_{l,j_{l}})=(h_{1},\dots,h_{j}) and ppl<ε\|p-p_{l}\|_{\infty}<\varepsilon.

If we now apply 3.1 to the sequence

γ(l)=Xpl(f(Thl,1x),,f(Thl,jlx)dμ(x),\gamma(l)=\int_{X}p_{l}(f(T_{h_{l,1}}x),\dots,f(T_{h_{l,j_{l}}}x)d\mu(x),

we obtain a sequence (zg)gG(z_{g})_{g\in G} of elements of DD such that, for all ll\in\mathbb{N},

limN1|FN|gFNpl(zhl,1g,,zhl,jlg)=Xpl(f(Thl,1x),,f(Thl,jlx)dμ(x).\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}p_{l}(z_{h_{l,1}g},\dots,z_{h_{l,j_{l}}g})=\int_{X}p_{l}(f(T_{h_{l,1}}x),\dots,f(T_{h_{l,j_{l}}}x)d\mu(x). (7)

This implies that Equation 3 holds for any h1,,hjGh_{1},\dots,h_{j}\in G and pC(Dk)p\in C(D^{k}), so we are done proving the first implication.

The converse implication of Theorem 2.7 follows from standard proofs of the the Furstenberg correspondence principle, see e.g. [BF21a, Pages 922-923]. We give a short proof using 3.1. Given a sequence (zg)gG(z_{g})_{g\in G} in DD, first choose a Følner subsequence (FN)n(F_{N}^{\prime})_{n\in\mathbb{N}} such that the following limit exists for all ll.

γ(l)=limN1|FN|gFNpl(zhl,1g,,zhl,jlg).\gamma(l)=\lim_{N\to\infty}\frac{1}{|F_{N}^{\prime}|}\sum_{g\in F_{N}^{\prime}}p_{l}(z_{h_{l,1}g},\dots,z_{h_{l,j_{l}}g}).

So, by 3.1 there is a m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and a measurable function f:XDf:X\to D such that, for all ll\in\mathbb{N}, Equation 5 holds. In particular, Equation 3 holds if the left hand side limit exists (as the limit along the sequence (FN)(F_{N}) is the same as along the sequence (FN)(F_{N}^{\prime})), as we wanted. ∎

We will now prove 3.1. We will show 1\implies2\implies3\implies1. The complicated implication is 3\implies1. During the rest of the proof we fix an amenable group GG, a Følner sequence (FN)(F_{N}) and a compact set DD\subseteq\mathbb{C}.

Proof of 1\implies2.

Let 𝐳=(za)aG\mathbf{z}=(z_{a})_{a\in G} be as in 1. Consider the (compact, metrizable) product space DGD^{G} of sequences (zg)gG(z_{g})_{g\in G} in DD, with the Borel σ\sigma-algebra. We have an action (Rg)gG(R_{g})_{g\in G} of GG on DGD^{G} by Rg((za)aG)=(zag)aGR_{g}((z_{a})_{a\in G})=(z_{ag})_{a\in G}.

For each NN let νN\nu_{N} be the average of Dirac measures 1|FN|gFNδRg𝐳\frac{1}{|F_{N}|}\sum_{g\in F_{N}}\delta_{R_{g}\mathbf{z}}. Let ν\nu be the weak limit of some subsequence (νNm)m(\nu_{N_{m}})_{m}; then ν\nu is invariant by RhR_{h} for all hGh\in G, because limN|FNΔhFN||FN|=0\lim_{N}\frac{|F_{N}\Delta hF_{N}|}{|F_{N}|}=0 and for all NN\in\mathbb{N},

νN(Rh)νN=1|FN|(gFNhFNδRg𝐳ghFNFNδRg𝐳).\nu_{N}-(R_{h})_{*}\nu_{N}=\frac{1}{|F_{N}|}\left(\sum_{g\in F_{N}\setminus hF_{N}}\delta_{R_{g}\mathbf{z}}-\sum_{g\in hF_{N}\setminus F_{N}}\delta_{R_{g}\mathbf{z}}\right).

Finally, we define f:DGf:D^{G}\to\mathbb{C} by f((za)aG)=zef((z_{a})_{a\in G})=z_{e} (where eGe\in G is the identity). Then for all ll\in\mathbb{N} we have

DGpl(f(Rhl,1x),,f(Rhl,jlx))𝑑ν(x)\displaystyle\int_{D^{G}}p_{l}(f(R_{h_{l,1}}x),\dots,f(R_{h_{l,j_{l}}}x))d\nu(x)
=\displaystyle= limmDGpl(f(Rhl,1x),,f(Rhl,jlx))𝑑νNm(x)\displaystyle\lim_{m}\int_{D^{G}}p_{l}(f(R_{h_{l,1}}x),\dots,f(R_{h_{l,j_{l}}}x))d\nu_{N_{m}}(x)
=\displaystyle= limm1|FNm|gFNmpl(f(Rhl,1(Rg𝐳)),,f(Rhl,jl(Rg(𝐳)))\displaystyle\lim_{m}\frac{1}{|F_{N_{m}}|}\sum_{g\in F_{N_{m}}}p_{l}(f(R_{h_{l,1}}(R_{g}\mathbf{z})),\dots,f(R_{h_{l,j_{l}}}(R_{g}(\mathbf{z})))
=\displaystyle= limm1|FNm|gFNmpl(zhl,1g,,zhl,jlg)=γ(l).\displaystyle\lim_{m}\frac{1}{|F_{N_{m}}|}\sum_{g\in F_{N_{m}}}p_{l}(z_{h_{l,1}g},\dots,z_{h_{l,j_{l}}g})=\gamma(l).\qed
Proof of 2\implies3.

We let (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and ff be as in 2, and we define the m.p.s. (DG,(DG),(Rg)gG)(D^{G},\mathcal{B}(D^{G}),(R_{g})_{g\in G}) as in the proof of 1\implies2.

Consider the function Φ:XDG\Phi:X\to D^{G} given by Φ(x)=(f(Tax))aG\Phi(x)=(f(T_{a}x))_{a\in G} and let ν\nu be the measure Φμ\Phi_{*}\mu in DGD^{G}. As a Borel probability measure on a compact metric space, ν\nu can be weakly approximated by a sequence of finitely supported measures (νN)n(\nu_{N})_{n\in\mathbb{N}}, which we may assume are of the form

νN=1|KN|kKNδ(za,k)aG.\nu_{N}=\frac{1}{|K_{N}|}\sum_{k\in K_{N}}\delta_{(z_{a,k})_{a\in G}}.

For some finite index sets KNK_{N} and some sequences (za,k)aG(z_{a,k})_{a\in G}, for kKNk\in K_{N}. Now, the fact that νNν\nu_{N}\to\nu weakly means that for all ll\in\mathbb{N} and AGA\subseteq G finite we have

limN1|KN||A|kKNgApl(zhl,1g,k,,zhl,jlg,k)\displaystyle\lim_{N}\frac{1}{|K_{N}|\cdot|A|}\sum_{k\in K_{N}}\sum_{g\in A}p_{l}(z_{h_{l,1}g,k},\dots,z_{h_{l,j_{l}}g,k})
=\displaystyle= limN1|A|gADGpl(zhl,1g,,zhl,jlg)𝑑νN((za))\displaystyle\lim_{N}\frac{1}{|A|}\sum_{g\in A}\int_{D^{G}}p_{l}(z_{h_{l,1}g},\dots,z_{h_{l,j_{l}}g})d\nu_{N}((z_{a}))
=\displaystyle= 1|A|gADGpl(zhl,1g,,zhl,jlg)𝑑ν((za))\displaystyle\frac{1}{|A|}\sum_{g\in A}\int_{D^{G}}p_{l}(z_{h_{l,1}g},\dots,z_{h_{l,j_{l}}g})d\nu((z_{a}))
=\displaystyle= 1|A|gAXpl(f(Thl,1gx),,f(Thl,jlgx))𝑑μ(x)\displaystyle\frac{1}{|A|}\sum_{g\in A}\int_{X}p_{l}(f(T_{h_{l,1}g}x),\dots,f(T_{h_{l,j_{l}}g}x))d\mu(x)
=\displaystyle= 1|A|gAXpl(f(Thl,1x),,f(Thl,jlx))𝑑μ(x)=1|A|gAγ(l)=γ(l).\displaystyle\frac{1}{|A|}\sum_{g\in A}\int_{X}p_{l}(f(T_{h_{l,1}}x),\dots,f(T_{h_{l,j_{l}}}x))d\mu(x)=\frac{1}{|A|}\sum_{g\in A}\gamma(l)=\gamma(l).

So for any LL\in\mathbb{N} and δ>0\delta>0, Equation 6 will be satisfied for big enough NN. ∎

In order to prove 3\implies1, we will first need some lemmas about averages of sequences of complex numbers.

Proposition 3.3.

Let M>0M>0 and δ(0,1)\delta\in(0,1). For any finite sets EEGE^{\prime}\subseteq E\subseteq G such that |EE||E|<δ\frac{|E\setminus E^{\prime}|}{|E|}<\delta and any complex numbers (zg)gG(z_{g})_{g\in G} with |zg|MgG|z_{g}|\leq M\;\forall g\in G, we have

|1|E|gEzg1|E|gEzg|<2Mδ.\left|\frac{1}{|E|}\sum_{g\in E}z_{g}-\frac{1}{|E^{\prime}|}\sum_{g\in E^{\prime}}z_{g}\right|<2M\delta.
Proof.

The left hand side above is equal to

||E||E||E||E|gEzg+1|E|gEEzg|\displaystyle\left|\frac{|E^{\prime}|-|E|}{|E|\cdot|E^{\prime}|}\sum_{g\in E^{\prime}}z_{g}+\frac{1}{|E|}\sum_{g\in E\setminus E^{\prime}}z_{g}\right| |E||E||E||E|M|E|+1|E|M|EE|\displaystyle\leq\frac{|E|-|E^{\prime}|}{|E|\cdot|E^{\prime}|}\cdot M|E^{\prime}|+\frac{1}{|E|}\cdot M|E\setminus E^{\prime}|
<Mδ+Mδ.\displaystyle<M\delta+M\delta.\qed
Definition 3.4.

Let S,TS,T be subsets of a group GG. We denote

ST:={gG;SgT and Sg(GT)}=(sSs1T)(sSs1T).\partial_{S}T:=\{g\in G;Sg\cap T\neq\varnothing\text{ and }Sg\cap(G\setminus T)\neq\varnothing\}=\left(\cup_{s\in S}s^{-1}T\right)\setminus\left(\cap_{s\in S}s^{-1}T\right).
Remark 3.5.

Note that if (FN)(F_{N}) is a Følner sequence in GG and SS is finite, then limN|SFN||FN|=0\lim_{N}\frac{|\partial_{S}F_{N}|}{|F_{N}|}=0, because limN|s1FNΔFN||FN|=0\lim_{N}\frac{|s^{-1}F_{N}\Delta F_{N}|}{|F_{N}|}=0 for all sGs\in G. Also note that if cGc\in G, then S(Tc)=(ST)c\partial_{S}(Tc)=(\partial_{S}T)c. Finally, if eSe\in S, then for all gTSTg\in T\setminus\partial_{S}T we have SgTSg\subseteq T.

The following lemma is a crucial part of the proof of 3.1. Intuitively, it says that if you have a finite set of bounded sequences fi:Tif_{i}:T_{i}\to\mathbb{C} defined in finite subsets T1,,TkT_{1},\dots,T_{k} of a group GG, you can reassemble them into a sequence f:Af:A\to\mathbb{C} defined in a bigger finite set AGA\subseteq G which is a union of right translates of the sets TiT_{i}, and the average value of ff will approximately be the weighted average of the average values of the fif_{i}.

Lemma 3.6.

Let GG be a group, δ,M>0,K\delta,M>0,K\in\mathbb{N} and c1,,cKGc_{1},\dots,c_{K}\in G. Let T1,,TK,AT_{1},\dots,T_{K},A be finite subsets of GG such that T1c1,,TKcKT_{1}c_{1},\dots,T_{K}c_{K} are pairwise disjoint, contained in AA and k=1K|Tk||A|>1δM\frac{\sum_{k=1}^{K}|T_{k}|}{|A|}>1-\frac{\delta}{M}. For k=1,,Kk=1,\dots,K let (zg,k)gG(z_{g,k})_{g\in G} be sequences of complex numbers in DD, and consider a sequence (zg)gG(z_{g})_{g\in G} such that

zg=zgck1,k for all gTkck,z_{g}=z_{gc_{k}^{-1},k}\text{ for all }g\in T_{k}c_{k},

Finally, let S={h1,,hj}GS=\{h_{1},\dots,h_{j}\}\subseteq G be finite with eSe\in S and let p:Djp:D^{j}\to\mathbb{C} satisfy pM\|p\|_{\infty}\leq M. If we have |STk||Tk|δM\frac{|\partial_{S}T_{k}|}{|T_{k}|}\leq\frac{\delta}{M} for all kk, then

|1|A|gAp(zh1g,,zhjg)k=1K|Tk|i=1K|Ti|(1|Tk|gTkp(zh1g,k,,zhjg,k))|4δ.\left|\frac{1}{|A|}\sum_{g\in A}p(z_{h_{1}g},\dots,z_{h_{j}g})-\sum_{k=1}^{K}\frac{|T_{k}|}{\sum_{i=1}^{K}|T_{i}|}\left(\frac{1}{|T_{k}|}\sum_{g\in T_{k}}p(z_{h_{1}g,k},\dots,z_{h_{j}g,k})\right)\right|\leq 4\delta.
Proof.

Considering sequences of complex numbers as functions f:G;af(a)f:G\to\mathbb{C};a\mapsto f(a), which we write during this proof as (f(a))a(f(a))_{a}, we will denote

p(f):=p(f(h1),,f(hj)).p(f):=p(f(h_{1}),\dots,f(h_{j})).

Note that, given gGg\in G, we will have p((zagck1,k)a)=p((zag)a)p((z_{agc_{k}^{-1},k})_{a})=p((z_{ag})_{a}) whenever zagck1,k=zagz_{agc_{k}^{-1},k}=z_{ag} for all aSa\in S, which happens when agTkckag\in T_{k}c_{k} for all aSa\in S, that is, when gTkckSTkckg\in T_{k}c_{k}\setminus\partial_{S}T_{k}c_{k}. Thus, for each k=1,,Kk=1,\dots,K we have

|1|Tk|gTkp((zag,k)a)1|Tk|gTkckp((zag)a)|\displaystyle\left|\frac{1}{|T_{k}|}\sum_{g\in T_{k}}p((z_{ag,k})_{a})-\frac{1}{|T_{k}|}\sum_{g\in T_{k}c_{k}}p((z_{ag})_{a})\right|
=\displaystyle= |1|Tk|gTkckp((zagck1,k)a)1|Tk|gTkckp((zag)a)|\displaystyle\left|\frac{1}{|T_{k}|}\sum_{g\in T_{k}c_{k}}p((z_{agc_{k}^{-1},k})_{a})-\frac{1}{|T_{k}|}\sum_{g\in T_{k}c_{k}}p((z_{ag})_{a})\right|
=\displaystyle= |1|Tk|gTkckSTkck(p((zagck1,k)a)p((zag)a))|2δ,\displaystyle\left|\frac{1}{|T_{k}|}\sum_{g\in T_{k}c_{k}\cap\partial_{S}T_{k}c_{k}}\left(p((z_{agc_{k}^{-1},k})_{a})-p((z_{ag})_{a})\right)\right|\leq 2\delta,

the last inequality being because |TkckSTkck||Tk|<δM\frac{|T_{k}c_{k}\cap\partial_{S}T_{k}c_{k}|}{|T_{k}|}<\frac{\delta}{M} and the summands have norm 2M\leq 2M. Taking weighted averages over k=1,,Kk=1,\dots,K, we obtain

|k=1K|Tk|i=1K|Ti|(1|Tk|gTkp((zag,k)a))k=1K|Tk|i=1K|Ti|(1|Tk|gTkckp((zag)a))|2δ.\left|\sum_{k=1}^{K}\frac{|T_{k}|}{\sum_{i=1}^{K}|T_{i}|}\cdot\left(\frac{1}{|T_{k}|}\sum_{g\in T_{k}}p((z_{ag,k})_{a})\right)\right.\\ \left.-\sum_{k=1}^{K}\frac{|T_{k}|}{\sum_{i=1}^{K}|T_{i}|}\cdot\left(\frac{1}{|T_{k}|}\sum_{g\in T_{k}c_{k}}p((z_{ag})_{a})\right)\right|\leq 2\delta.

But applying 3.3 with E=AE=A, E=iTiciE^{\prime}=\cup_{i}T_{i}c_{i} and the sequence gp((zag)a)g\mapsto p((z_{ag})_{a}) we also have

|1|A|gAp((zag)a)k=1K|Tk|i=1K|Ti|(1|Tk|gTkckp((zag)a))|2δ.\left|\frac{1}{|A|}\sum_{g\in A}p((z_{ag})_{a})-\sum_{k=1}^{K}\frac{|T_{k}|}{\sum_{i=1}^{K}|T_{i}|}\cdot\left(\frac{1}{|T_{k}|}\sum_{g\in T_{k}c_{k}}p((z_{ag})_{a})\right)\right|\leq 2\delta.

So by the triangle inequality, we are done. ∎

The only non elementary fact that we will need in the proof of 3\implies1 is [DHZ19, Theorem 5.2]; in order to state it we first recall some definitions:

Definition 3.7 (cf. [DHZ19, Defs 3.1,3.2]).

A tiling 𝒯\mathcal{T} of a group GG consists of two objects:

  • A finite family 𝒮(𝒯)\mathcal{S}(\mathcal{T}) (the shapes) of finite subsets of GG containing the identity ee.

  • A finite collection C(𝒯)={C(S);S𝒮(𝒯)}C(\mathcal{T})=\{C(S);S\in\mathcal{S}(\mathcal{T})\} of subsets of GG, the center sets, such that the family of right translates of form ScSc with S𝒮(𝒯)S\in\mathcal{S}(\mathcal{T}) and cC(S)c\in C(S) (such sets ScSc are the tiles of 𝒯\mathcal{T}) form a partition of GG.

Definition 3.8 (cf. [DHZ19, Page 17]).

We say a sequence of tilings (𝒯k)k(\mathcal{T}_{k})_{k\in\mathbb{N}} of a group GG is congruent if every tile of 𝒯k+1\mathcal{T}_{k+1} equals a union of tiles of 𝒯k\mathcal{T}_{k}.

Definition 3.9.

If AA and BB are finite subsets of a group GG, we say that AA is (B,ε)(B,\varepsilon)-invariant if |BA||A|<ε\frac{|\partial_{B}A|}{|A|}<\varepsilon.555This is not the definition of (B,ε)(B,\varepsilon)-invariant used in [DHZ19], but it will be more convenient for our purposes.

We will need the following weak version of [DHZ19, Theorem 5.2]:

Theorem 3.10 (cf. [DHZ19, Theorem 5.2]).

For any countable amenable group GG there exists a congruent sequence of tilings (𝒯k)k(\mathcal{T}_{k})_{k\in\mathbb{N}} such that, for every AGA\subseteq G finite and δ>0\delta>0, all the tiles of 𝒯k\mathcal{T}_{k} are (A,δ)(A,\delta)-invariant for big enough kk.

The following notation will be useful.

Definition 3.11.

For any tiling of a group GG and any finite set BGB\subseteq G, we let 𝒯B\partial_{\mathcal{T}}B denote the union of all tiles of 𝒯\mathcal{T} which intersect both BB and GBG\setminus B.

Remark 3.12.

Let GG be a countable amenable group with a tiling 𝒯\mathcal{T} and a Følner sequence (FN)(F_{N}). As 𝒯FNS𝒮(𝒯)SFN\partial_{\mathcal{T}}F_{N}\subseteq\cup_{S\in\mathcal{S}(\mathcal{T})}\partial_{S}F_{N}, Remark 3.5 implies that

limN|𝒯FN||FN|=0.\lim_{N\to\infty}\frac{|\partial_{\mathcal{T}}F_{N}|}{|F_{N}|}=0.
Lemma 3.13.

Let GG be a countably infinite amenable group with a Følner sequence (FN)(F_{N}), and let (𝒯k)k(\mathcal{T}_{k})_{k\in\mathbb{N}} be a congruent sequence of tilings of GG. Then there is a partition 𝒫\mathcal{P} of GG into tiles of the tilings 𝒯k\mathcal{T}_{k} such that

  • For each kk\in\mathbb{N}, 𝒫\mathcal{P} contains only finitely many tiles of 𝒯k\mathcal{T}_{k}.

  • If for each NN\in\mathbb{N} we let AN={T𝒫;TFN}FNA_{N}=\bigcup\{T\in\mathcal{P};T\subseteq F_{N}\}\subseteq F_{N}, then we have

    limN|AN||FN|=1.\lim_{N\to\infty}\frac{|A_{N}|}{|F_{N}|}=1.
Proof.

We can assume that NFN=G\cup_{N}F_{N}=G, adding some Følner sets to the sequence if necessary. In the following, for each finite set BGB\subseteq G and kk\in\mathbb{N} we will denote kB:=j=1k𝒯jB\partial_{k}B:=\cup_{j=1}^{k}\partial_{\mathcal{T}_{j}}B, so that for all kk\in\mathbb{N},

limN|kFN||FN|=0.\lim_{N\to\infty}\frac{|\partial_{k}F_{N}|}{|F_{N}|}=0.

Now for each kk\in\mathbb{N} let NkN_{k} be a big enough number that |k+1FN||FN|1k+1\frac{|\partial_{k+1}F_{N}|}{|F_{N}|}\leq\frac{1}{k+1} for all NNkN\geq N_{k}.

Let D0=D_{0}=\varnothing and for each kk\in\mathbb{N} let DkGD_{k}\subseteq G be the union of all tiles of 𝒯k+1\mathcal{T}_{k+1} intersecting some element of N=1NkFN\bigcup_{N=1}^{N_{k}}F_{N}. Thus, DkDk+1D_{k}\subseteq D_{k+1} for all kk, G=kDkG=\cup_{k}D_{k} and DkDk1D_{k}\setminus D_{k-1} is a union of tiles of 𝒯k\mathcal{T}_{k}. We define 𝒫\mathcal{P} to be the partition of GG formed by all tiles of 𝒯k\mathcal{T}_{k} contained in DkDk1D_{k}\setminus D_{k-1}, for all kk\in\mathbb{N}.

Now, fix NN and let kk be the smallest natural number such that NNkN\leq N_{k} (note that kk\to\infty when NN\to\infty). Note that all tiles T𝒫T\in\mathcal{P} intersecting FNF_{N} must be in 𝒯j\mathcal{T}_{j} for some jkj\leq k. Thus, the set ANA_{N} of all tiles of 𝒫\mathcal{P} which are contained in FNF_{N} must contain |FNkFN||F_{N}\setminus\partial_{k}F_{N}|. But we have N>Nk1N>N_{k-1}, so |kFN||FN|1k\frac{|\partial_{k}F_{N}|}{|F_{N}|}\leq\frac{1}{k}. So |AN||FN|11k\frac{|A_{N}|}{|F_{N}|}\geq 1-\frac{1}{k}, and we are done.∎

Proof of Item 3\impliesItem 1.

Let Sl={hl,1,,hl,jl}S_{l}=\{h_{l,1},\dots,h_{l,j_{l}}\} and MlplM_{l}\geq\|p_{l}\|_{\infty} for all ll\in\mathbb{N} (we may assume eSle\in S_{l} and Ml+1MlM_{l+1}\geq M_{l} for all ll). We prove first that 3 implies the following:

  1. 3’.

    Let δ>0\delta>0 and LL\in\mathbb{N}. Then for any sufficiently left-invariant666By this we mean that there is a finite set AGA\subseteq G and some ε>0\varepsilon>0 such that the property stated below is satisfied for all (A,ε)(A,\varepsilon)-invariant sets. subset BB of GG there exists a sequence (wg)gG(w_{g})_{g\in G} in DD such that, for all l=1,,Ll=1,\dots,L,

    |γ(l)1|B|gBpl(whl,1g,,whl,jlg)|<δ.\left|\gamma(l)-\frac{1}{|B|}\sum_{g\in B}p_{l}(w_{h_{l,1}g},\dots,w_{h_{l,j_{l}}g})\right|<\delta.

In order to prove 3’. from 3, fix δ,L\delta,L and consider a tiling 𝒯\mathcal{T} of GG such that |SlS||S|<δ10ML\frac{|\partial_{S_{l}}S|}{|S|}<\frac{\delta}{10M_{L}} for all l=1,,Ll=1,\dots,L and S𝒮(𝒯)S\in\mathcal{S}(\mathcal{T}). For each S𝒮(𝒯)S\in\mathcal{S}(\mathcal{T}) there is by 3 some KSK_{S}\in\mathbb{N} and sequences (zS,g,k)gG(z_{S,g,k})_{g\in G} in DD (k=1,,KSk=1,\dots,K_{S}) satisfying

|γ(l)1KS|S|k=1KSgSpl(zS,hl,1g,k,,zS,hl,jlg,k)|<δ5 for l=1,,L.\left|\gamma(l)-\frac{1}{K_{S}|S|}\sum_{k=1}^{K_{S}}\sum_{g\in S}p_{l}(z_{S,h_{l,1}g,k},\dots,z_{S,h_{l,j_{l}}g,k})\right|<\frac{\delta}{5}\text{ for }l=1,\dots,L. (8)

Then any finite subset BB of GG such that |TB||B|<δ10ML\frac{|\partial_{T}B|}{|B|}<\frac{\delta}{10M_{L}} and

|B|10MLδS𝒮(𝒯)|S|KS|B|\geq\frac{10M_{L}}{\delta}\sum_{S\in\mathcal{S}(\mathcal{T})}|S|K_{S} (9)

will satisfy 3’.; to see why, first note that the union B0B_{0} of all tiles of 𝒯\mathcal{T} contained in BB satisfies |B0||B|>1δ10ML\frac{|B_{0}|}{|B|}>1-\frac{\delta}{10M_{L}}; now obtain a set B1B0B_{1}\subseteq B_{0} by removing finitely many tiles from B0B_{0} in such a way that, for each S𝒮(𝒯)S\in\mathcal{S}(\mathcal{T}), the number of tiles of shape SS contained in B1B_{1} is a multiple of KSK_{S}. Note that, due to Equation 9, this can be done in such a way that |B1||B|1δ5ML\frac{|B_{1}|}{|B|}\geq 1-\frac{\delta}{5M_{L}}.

So letting 𝒫\mathcal{P} be the set of tiles of 𝒯\mathcal{T} contained in B1B_{1}, we can define a function f:𝒫f:\mathcal{P}\to\mathbb{N} that associates for each translate ScSc some value in {1,2,,KS}\{1,2,\dots,K_{S}\}, and such that for each SS, the number of right-translates of SS having value kk is the same for all k=1,2,,KSk=1,2,\dots,K_{S}. Finally, define a sequence (wg)gG(w_{g})_{g\in G} by

wg=zS,gc1,f(Sc) if gSc and Sc is a tile of 𝒯 contained in B1.w_{g}=z_{S,gc^{-1},f(Sc)}\text{ if }g\in Sc\text{ and }Sc\text{ is a tile of }\mathcal{T}\text{ contained in }B_{1}.

The rest of values of wgw_{g} are not important, we can let them be some fixed value of DD. Then, by 3.6 applied to the set BB and the tiles of 𝒯\mathcal{T} contained in B1B_{1}, we obtain for all l=1,,Ll=1,\dots,L that

|1|B|gBp(whl,1g,,whl,jl)Sc𝒫|Sc||B1|(1|S|gSp(zS,hl,1g,f(Sc),,zS,hl,jlg,f(Sc)))|4(δ5).\left|\frac{1}{|B|}\sum_{g\in B}p(w_{h_{l,1}g},\dots,w_{h_{l,j_{l}}})\right.\\ \left.-\sum_{Sc\in\mathcal{P}}\frac{|Sc|}{|B_{1}|}\cdot\left(\frac{1}{|S|}\sum_{g\in S}p(z_{S,h_{l,1}g,f(Sc)},\dots,z_{S,h_{l,j_{l}}g,f(Sc)})\right)\right|\leq 4\left(\frac{\delta}{5}\right).

But as the sequences (zS,g,k)gG(z_{S,g,k})_{g\in G} in DD (k=1,,KSk=1,\dots,K_{S}) satisfy Equation 8, we have for all l=1,,Ll=1,\dots,L that

|γ(l)Sc𝒫|Sc||B1|(1|S|gSp(zS,hl,1g,f(Sc),,zS,hl,jlg,f(Sc)))|δ5,\left|\gamma(l)-\sum_{Sc\in\mathcal{P}}\frac{|Sc|}{|B_{1}|}\cdot\left(\frac{1}{|S|}\sum_{g\in S}p(z_{S,h_{l,1}g,f(Sc)},\dots,z_{S,h_{l,j_{l}}g,f(Sc)})\right)\right|\leq\frac{\delta}{5},

so we are done proving 3’. by the triangle inequality.

Now we use 3’. to prove 1. By 3’. and Theorem 3.10, there is a congruent sequence of tilings (𝒯L)L(\mathcal{T}_{L})_{L\in\mathbb{N}} of GG (we may also assume 𝒮(𝒯L)𝒮(𝒯L)=\mathcal{S}(\mathcal{T}_{L})\cap\mathcal{S}(\mathcal{T}_{L^{\prime}})=\varnothing if LLL\neq L^{\prime}) such that

  1. 1.

    For all S𝒮(𝒯L)S\in\mathcal{S}(\mathcal{T}_{L}) and for l=1,,Ll=1,\dots,L we have |SlS||S|<1L\frac{|\partial_{S_{l}}S|}{|S|}<\frac{1}{L}.

  2. 2.

    For each S𝒮(𝒯L)S\in\mathcal{S}(\mathcal{T}_{L}) there is a sequence (wS,g)gG(w_{S,g})_{g\in G} such that for all l=1,,Ll=1,\dots,L,

    |γ(l)1|S|gSpl(wS,hl,1g,,wS,hl,jlg)|<1L.\left|\gamma(l)-\frac{1}{|S|}\sum_{g\in S}p_{l}(w_{S,h_{l,1}g},\dots,w_{S,h_{l,j_{l}}g})\right|<\frac{1}{L}. (10)

Now, letting (FN)(F_{N}) be the Følner sequence of Item 1, we let 𝒫\mathcal{P} and (AN)n(A_{N})_{n\in\mathbb{N}} be as in 3.13, with 𝒫L𝒫\mathcal{P}_{L}\subseteq\mathcal{P} (L=1,2,L=1,2,\dots) being finite sets of tiles of 𝒯L\mathcal{T}_{L} such that 𝒫=L𝒫L\mathcal{P}=\sqcup_{L}\mathcal{P}_{L}. We define the sequence (zg)gG(z_{g})_{g\in G} by

zg=wS,gc1 if gSc, where S𝒮(𝒯L) for some L and Sc𝒫L.z_{g}=w_{S,gc^{-1}}\text{ if }g\in Sc,\text{ where }S\in\mathcal{S}(\mathcal{T}_{L})\text{ for some $L$ and }Sc\in\mathcal{P}_{L}.

All that is left is proving that (zg)g(z_{g})_{g} satisfies 1 for all ll\in\mathbb{N}. So let ε>0\varepsilon>0 and ll\in\mathbb{N}. Note that for big enough NN we have |AN||FN|>1ε10Ml\frac{|A_{N}|}{|F_{N}|}>1-\frac{\varepsilon}{10M_{l}}. Fix some natural number L>10εL>\frac{10}{\varepsilon} (also suppose LlL\geq l). Letting ANA_{N}^{\prime} be obtained from ANA_{N} by removing the (finitely many) tiles which are in 𝒫i\mathcal{P}_{i} for some i=1,,Li=1,\dots,L, then for big enough NN we have |AN||FN|>1ε5Ml\frac{|A_{N}^{\prime}|}{|F_{N}|}>1-\frac{\varepsilon}{5M_{l}}. Thus, applying 3.6 to the set FNF_{N} and to all the tiles contained in ANA_{N}^{\prime}, and for each S𝒮(𝒯L)S\in\mathcal{S}(\mathcal{T}_{L}) letting nSn_{S} be the number of tiles of shape SS in 𝒫L\mathcal{P}_{L} which are contained in ANA_{N}^{\prime}, we obtain

|1|FN|gFNpl(zhl,1g,,zhl,jlg)LS𝒮(𝒯L)nS|S||AN|(1|S|gSpl(wS,hl,1g,L,,wS,hl,jlg,L))|4ε5.\left|\frac{1}{|F_{N}|}\sum_{g\in F_{N}}p_{l}(z_{h_{l,1}g},\dots,z_{h_{l,j_{l}}g})\right.\\ -\left.\sum_{L\in\mathbb{N}}\sum_{S\in\mathcal{S}(\mathcal{T}_{L})}\frac{n_{S}|S|}{|A_{N}^{\prime}|}\cdot\left(\frac{1}{|S|}\sum_{g\in S}p_{l}(w_{S,h_{l,1}g,L},\dots,w_{S,h_{l,j_{l}}g,L})\right)\right|\leq 4\frac{\varepsilon}{5}. (11)

However, the double sum in Equation 11 is at distance ε10\leq\frac{\varepsilon}{10} of γ(l)\gamma(l) (this follows from taking an affine combination of Equation 10 applied to the tiles SS of ANA_{N}^{\prime}, with constant 1L<ε10\frac{1}{L}<\frac{\varepsilon}{10}). So by the triangle inequality, for big enough NN we have

|γ(l)1|FN|gFNpl(zhl,1g,,zhl,jlg)|<ε.\left|\gamma(l)-\frac{1}{|F_{N}|}\sum_{g\in F_{N}}p_{l}(z_{h_{l,1}g},\dots,z_{h_{l,j_{l}}g})\right|<\varepsilon.

As ε\varepsilon is arbitrary, we are done. ∎

As promised in the introduction, we now prove a converse to Furstenberg’s correspondence principle. Theorem 3.14 is a slight generalization of Theorem 1.6 in which we also allow intersections with complements of the sets giAg_{i}A. In the following, for any AGA\subseteq G and BXB\subseteq X, we denote A1=A,A0=GAA^{1}=A,A^{0}=G\setminus A, B1=BB^{1}=B and B0=XBB^{0}=X\setminus B.

Theorem 3.14.

Let GG be a countably infinite amenable group with a Følner sequence (FN)(F_{N}). For every m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and every BB\in\mathcal{B} there exists a subset AGA\subseteq G such that, for all nn\in\mathbb{N}, ε1,,εn{0,1}\varepsilon_{1},\dots,\varepsilon_{n}\in\{0,1\} and g1,,gnGg_{1},\dots,g_{n}\in G we have

dF(i=1n(giA)εi)=μ(i=1k(TgiB)εi).d_{F}\left(\bigcap_{i=1}^{n}(g_{i}A)^{\varepsilon_{i}}\right)=\mu\left(\bigcap_{i=1}^{k}(T_{g_{i}}B)^{\varepsilon_{i}}\right). (12)

Reciprocally, for any subset AGA\subseteq G there is a m.p.s. (X,,μ,T)(X,\mathcal{B},\mu,T) and BB\in\mathcal{B} satisfying Equation 12 for all n,ε1,,εn,g1,,gnn,\varepsilon_{1},\dots,\varepsilon_{n},g_{1},\dots,g_{n} as above such that the density in Equation 12 exists.

Proof.

Let p0,p1:{0,1}p_{0},p_{1}:\{0,1\}\to\mathbb{C} be given by p0(x)=1xp_{0}(x)=1-x and p1(x)=xp_{1}(x)=x. Note that for all n,ε1,,εn,g1,,gnn,\varepsilon_{1},\dots,\varepsilon_{n},g_{1},\dots,g_{n} we have

μ(i=1k(TgiB)εi)=Xi=1kpεi(χB(Tgi1x))dμ.\displaystyle\mu\left(\bigcap_{i=1}^{k}(T_{g_{i}}B)^{\varepsilon_{i}}\right)=\int_{X}\prod_{i=1}^{k}p_{\varepsilon_{i}}\left(\chi_{B}\left(T_{g_{i}^{-1}}x\right)\right)d\mu.

So by Theorem 2.7 applied to the polynomials of the form p(x1,,xn)=i=1kpεi(xi)p(x_{1},\dots,x_{n})=\prod_{i=1}^{k}p_{\varepsilon_{i}}(x_{i}) (for all n,εi,gin,\varepsilon_{i},g_{i}) and with D={0,1}D=\{0,1\}, there exists some characteristic function χA:G{0,1}\chi_{A}:G\to\{0,1\} such that for all n,(εi)i=1nn,(\varepsilon_{i})_{i=1}^{n} and (gi)i=1n(g_{i})_{i=1}^{n} we have

μ(i=1k(TgiB)εi)=limN1|FN|gFNi=1kpεi(χA(gi1g))=dF(i=1k(giA)εi).\displaystyle\mu\left(\bigcap_{i=1}^{k}(T_{g_{i}}B)^{\varepsilon_{i}}\right)=\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}\prod_{i=1}^{k}p_{\varepsilon_{i}}\left(\chi_{A}(g_{i}^{-1}g)\right)=d_{F}\left(\bigcap_{i=1}^{k}(g_{i}A)^{\varepsilon_{i}}\right).

The other implication is proved similarly, applying Theorem 2.7 in the other direction. ∎

Theorem 3.14 is a statement about densities of any sets in the algebra 𝒜𝒫(G)\mathcal{A}\subseteq\mathcal{P}(G) generated by the family {gA;gG}\{gA;g\in G\}. For example, using that for any g1,,gnGg_{1},\dots,g_{n}\in G the set i=1ngiA\cup_{i=1}^{n}g_{i}A is the union for all ε1,,εn\varepsilon_{1},\dots,\varepsilon_{n} not all equal to 11 of i=1n(giA)εi\cap_{i=1}^{n}(g_{i}A)^{\varepsilon_{i}}, we obtain a version of Furstenberg’s correspondence principle with unions:

Theorem 3.15 (Inverse Furstenberg correspondence principle with unions).

Let GG be a countably infinite amenable group with a Følner sequence (FN)(F_{N}). For every m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and every BB\in\mathcal{B} there exists a subset AGA\subseteq G such that for all kk\in\mathbb{N} and h1,,hkGh_{1},\dots,h_{k}\in G we have

dF(h1AhkA)=μ(Th1BThkB).d_{F}\left(h_{1}A\cup\dots\cup h_{k}A\right)=\mu\left(T_{h_{1}}B\cup\dots\cup T_{h_{k}}B\right).\qed
Remark 3.16.

It is possible to give a version of Theorem 1.6 that involves translates of several sets B1,,BlXB_{1},\dots,B_{l}\subseteq X, as in [BF21b, Definition A.3], or even countably many sets (Bk)k(B_{k})_{k\in\mathbb{N}}. We will not do so as it falls outside the scope of this article and it would involve proving a modified version of 3.1.

We include a version of Theorem 1.6 for semigroups.

Definition 3.17.

Let (S,)(S,\cdot) be a semigroup. For ASA\subseteq S and sSs\in S, we let s1A={xS;sxA}s^{-1}A=\{x\in S;sx\in A\}. We say SS is (left) amenable777This definition is known to be equivalent to Definition 1.1 for countable groups. if there is a finitely additive probability measure μ:𝒫(S)[0,1]\mu:\mathcal{P}(S)\to[0,1] such that μ(s1A)=μ(A)\mu(s^{-1}A)=\mu(A) for all sS,ASs\in S,A\subseteq S. We say that SS is cancellative when for all a,b,cSa,b,c\in S, ab=acab=ac implies b=cb=c and ba=caba=ca implies b=cb=c. A Følner sequence in SS is a sequence (FN)N(F_{N})_{N\in\mathbb{N}} of finite subsets of SS such that, for all sSs\in S, limN|FNΔsFN||FN|=0\lim_{N\to\infty}\frac{|F_{N}\Delta sF_{N}|}{|F_{N}|}=0.

For the rest of the section we fix a countable, (left) amenable, (two-sided) cancellative semigroup SS. In particular, SS is left-reversible, that is, aSbSaS\cap bS\neq\varnothing for all a,bSa,b\in S (in fact, μ(aSbS)=1\mu(aS\cap bS)=1 for any left-invariant mean μ\mu in SS). An argument dating back to [Ore31] implies that SS can be imbedded into a group GG, the group of right quotients of SS, such that G={st1;s,tS}G=\{st^{-1};s,t\in S\}, see [CP61, Theorem 1.23].

We will use the fact that any SS-m.p.s. can be extended to a GG-m.p.s. Extensions of measure-theoretic and topological semigroup actions to groups have quite recently been studied in [FJM24, Don24, BBD25a, BBD25b].

Definition 3.18.

We say a measurable space (X,)(X,\mathcal{B}) is standard Borel if there is a metric dd on XX such that (X,d)(X,d) is a complete, separable metric space with Borel σ\sigma-algebra \mathcal{B}. We say a m.p.s. (X,,μ,(Ts)sS)(X,\mathcal{B},\mu,(T_{s})_{s\in S}) is standard Borel if (X,)(X,\mathcal{B}) is standard Borel.

Theorem 3.19 (See [Don24, Theorem 2.7.7] or [FJM24, Theorem 2.9]).

For any standard Borel m.p.s. (X,,μ,(Ts)sS)(X,\mathcal{B},\mu,(T_{s})_{s\in S}) there is a m.p.s. (Y,𝒞,ν,(Sg)gG)(Y,\mathcal{C},\nu,(S_{g})_{g\in G}) and a measure preserving map π:YX\pi:Y\to X with πSs=Tsπ\pi\circ S_{s}=T_{s}\circ\pi for all sSs\in S.

Theorem 3.20.

Let SS be a countably infinite, amenable, cancellative semigroup with a Følner sequence (FN)(F_{N})888Any amenable, (two-sided) cancellative semigroup SS has a Følner sequence: let GG be the group of right fractions of SS. Then from any Følner sequence (FN)(F_{N}) in GG and any elements cNgFNg1Sc_{N}\in\cap_{g\in F_{N}}g^{-1}S we obtain a Følner sequence (FNcN)(F_{N}c_{N}) in SS.. For every m.p.s. (X,,μ,(Ts)sS)(X,\mathcal{B},\mu,(T_{s})_{s\in S}) and every BB\in\mathcal{B} there exists a subset ASA\subseteq S such that for all kk\in\mathbb{N} and h1,,hkSh_{1},\dots,h_{k}\in S we have

dF(h11Ahk1A)=μ(Th11BThk1B).d_{F}\left(h_{1}^{-1}A\cap\dots\cap h_{k}^{-1}A\right)=\mu\left(T_{h_{1}}^{-1}B\cap\dots\cap T_{h_{k}}^{-1}B\right). (13)
Remark 3.21.

We only generalize Theorem 1.6 to semigroup actions, but the same argument can be used to generalize Theorem 2.7.

Proof.

Let GG be the group of right quotients of SS. Note that as SS generates GG and (FN)(F_{N}) is a Følner sequence in SS, (FN)(F_{N}) is also a Følner sequence in GG.

We first suppose that (X,)(X,\mathcal{B}) is standard Borel, so by Theorem 3.19 there is a m.p.s. (Y,𝒞,ν,(Sg)gG)(Y,\mathcal{C},\nu,(S_{g})_{g\in G}) and a measure preserving map π:YX\pi:Y\to X such that πSs=Tsπ\pi\circ S_{s}=T_{s}\circ\pi for all sSs\in S. Letting D=π1(B)D=\pi^{-1}(B), we have

ν(Sh11DShk1D)=μ(Th11BThk1B)\nu\left(S_{h_{1}}^{-1}D\cap\dots\cap S_{h_{k}}^{-1}D\right)=\mu\left(T_{h_{1}}^{-1}B\cap\dots\cap T_{h_{k}}^{-1}B\right)

for all h1,,hkSh_{1},\dots,h_{k}\in S. We finish the proof be applying Theorem 1.6 to obtain a set AGA\subseteq G such that for all kk\in\mathbb{N} and h1,,hkGh_{1},\dots,h_{k}\in G we have

dF(h1AhkA)=ν(Sh1DShkD)=μ(Th1BThkB).d_{F}\left(h_{1}A\cap\dots\cap h_{k}A\right)=\nu(S_{h_{1}}D\cap\cdots\cap S_{h_{k}}D)=\mu\left(T_{h_{1}}B\cap\dots\cap T_{h_{k}}B\right).

We now tackle the general case, where (X,,μ)(X,\mathcal{B},\mu) is an arbitrary measure space. Our construction is similar to the ones in [Par05, Chapter 5, Section 3]. We may assume that SS contains the identity eGe\in G (if not, we work with S{e}S\cup\{e\}). Consider the compact metric space M={0,1}SM=\{0,1\}^{S} with the product topology, and let M\mathcal{B}_{M} be the Borel σ\sigma-algebra in MM, which is generated by the sets Mh={(xs)M;xh=1}M_{h}=\{(x_{s})\in M;x_{h}=1\}, hSh\in S. The map Φ:XM;x(χB(Tsx))sS\Phi:X\to M;x\mapsto(\chi_{B}(T_{s}x))_{s\in S} is measurable, and if we consider the continuous, measurable maps Uh:MM;(xs)sS(xsh)sSU_{h}:M\to M;(x_{s})_{s\in S}\mapsto(x_{sh})_{s\in S}, then we have ΦTh=UhΦ\Phi\circ T_{h}=U_{h}\circ\Phi for all hSh\in S. Thus, the pushforward probability measure ν=Φμ\nu=\Phi_{*}\mu in (M,M)(M,\mathcal{B}_{M}) is invariant by UhU_{h} for all hSh\in S. Letting C={(xs)M;xe=1}C=\{(x_{s})\in M;x_{e}=1\}, for all hSh\in S we have Φ1(Us1C)=Ts1B\Phi^{-1}(U_{s}^{-1}C)=T_{s}^{-1}B. So for all h1,,hkSh_{1},\dots,h_{k}\in S,

ν(Uh11CUhk1C)=μ(Th11BThk1B).\nu\left(U_{h_{1}}^{-1}C\cap\cdots\cap U_{h_{k}}^{-1}C\right)=\mu\left(T_{h_{1}}^{-1}B\cap\dots\cap T_{h_{k}}^{-1}B\right).

As (M,M)(M,\mathcal{B}_{M}) is standard Borel, we are done. ∎

4 Characterizations of vdC sets in amenable groups

In this section we prove Theorem 4.1 below, our main characterization theorem for vdC sets in countable amenable groups. Theorem 4.1 gives a characterization of FF-vdC sets analogous to Theorem 2.2 but for any Følner sequence. In particular, it implies that the notion of FF-vdC set is independent of the Følner sequence, thus answering the question in [BL08, Section 4.2] of whether FF-vdC implies vdC. See 7.14 for yet another characterization of FF-vdC sets.

Theorem 4.1.

Let GG be a countably infinite amenable group with a Følner sequence (FN)(F_{N}). For a set HG{e}H\subseteq G\setminus\{e\}, the following are equivalent:

  1. 1.

    HH is an FF-vdC set.

  2. 2.

    For all sequences (zg)gG(z_{g})_{g\in G} of complex numbers in the unit disk 𝔻\mathbb{D},

    limN1|FN|gFNzhgzg¯=0 for all hH implies limN1|FN|gFNzg=0.\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{hg}\overline{z_{g}}=0\text{ for all }h\in H\textup{ implies }\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{g}=0.
  3. 3.

    For all sequences (zg)gG(z_{g})_{g\in G} of complex numbers in 𝕊1\mathbb{S}^{1},

    limN1|FN|gFNzhgzg¯=0 for all hH implies limN1|FN|gFNzg=0.\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{hg}\overline{z_{g}}=0\text{ for all }h\in H\textup{ implies }\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{g}=0.
  4. 4.

    HH is a vdC set in GG: for all m.p.s. (X,𝒜,μ,(Tg)gG)(X,\mathcal{A},\mu,(T_{g})_{g\in G}) and fL(μ)f\in L^{\infty}(\mu),

    Xf(Thx)f(x)¯𝑑μ(x)=0 for all hH implies Xf𝑑μ=0.\int_{X}f(T_{h}x)\cdot\overline{f(x)}d\mu(x)=0\text{ for all }h\in H\text{ implies }\int_{X}fd\mu=0. (14)
  5. 5.

    (Finitistic characterization) For every ε>0\varepsilon>0 there is δ>0\delta>0 and finite sets AG,H0HA\subseteq G,H_{0}\subseteq H such that for any KK\in\mathbb{N} and sequences (za,k)aG(z_{a,k})_{a\in G} in 𝔻\mathbb{D}, for k=1,,Kk=1,\dots,K, such that

    |1K|A|k=1KaAzha,kza,k¯|<δ for all hH0,\left|\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{a\in A}z_{ha,k}\overline{z_{a,k}}\right|<\delta\text{ for all }h\in H_{0},

    we have

    |1K|A|k=1KaAza,k|<ε.\left|\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{a\in A}z_{a,k}\right|<\varepsilon.
Remark 4.2.

It is an interesting question whether, if we change ‘fL(μ)f\in L^{\infty}(\mu)’ by ‘fL2(μ)f\in L^{2}(\mu)’ in Definition 1.13, the resulting definition of vdC set in GG is equivalent. A similar question is posed in [FT24, Question 3.7], where they call the sets defined by the L2L^{2} definition ‘sets of operatorial recurrence’.

The relationship between equidistribution and Cesaro averages is explained by 4.4 below, which was introduced by Weyl in [Wey16].

Definition 4.3.

Let GG be a countable amenable group with a Følner sequence (FN)(F_{N}). We say that a sequence (zg)gG(z_{g})_{g\in G}, with zg𝕊1z_{g}\in\mathbb{S}^{1} for all gGg\in G, is FF-u.d. in 𝕊1\mathbb{S}^{1} if for every continuous function f:𝕊1f:\mathbb{S}^{1}\to\mathbb{C} we have

limN1|FN|gFNf(zg)=𝕊1f𝑑m,\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}f(z_{g})=\int_{\mathbb{S}^{1}}fdm,

where mm is the uniform probability measure in 𝕊1\mathbb{S}^{1}.

That is, a sequence (xg)gG(x_{g})_{g\in G} in 𝕋\mathbb{T} is FF-u.d. mod 11 (as in Definition 1.10) iff (e2πixg)gG(e^{2\pi ix_{g}})_{g\in G} is FF-u.d. in 𝕊1\mathbb{S}^{1}.

Proposition 4.4 (Weyl’s criterion for uniform distribution).

Let GG be a countable amenable group with a Følner sequence (FN)(F_{N}). A sequence (zg)gG(z_{g})_{g\in G} in 𝕊1\mathbb{S}^{1} is FF-u.d. in 𝕊1\mathbb{S}^{1} iff for all l{0}l\in\mathbb{Z}\setminus\{0\} (or equivalently, for all ll\in\mathbb{N}) we have

limN1|FN|gFNzgl=0.\lim_{N}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{g}^{l}=0.

For a proof of 4.4 see e.g. [KN74, Theorem 2.1] (the proof works for any Følner sequence).

Proof of Theorem 4.1.

We prove 2\implies3\implies1\implies5\implies4\implies2

  1. 2\implies3

    Obvious.

  2. 3\implies1

    If HH is not FF-vdC then there is a sequence of complex numbers (zg)gG(z_{g})_{g\in G} in 𝕊1\mathbb{S}^{1} which is not FF-u.d. in 𝕊1\mathbb{S}^{1} but such that (zhgzg¯)g(z_{hg}\overline{z_{g}})_{g} is FF-u.d. in 𝕊1\mathbb{S}^{1} for all hHh\in H. By Weyl’s criterion, that means that for all hHh\in H we have

    limN1|FN|gFNzhglzgl¯=0 for all l{0},\lim_{N}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{hg}^{l}\overline{z_{g}^{l}}=0\text{ for all }l\in\mathbb{Z}\setminus\{0\},

    but there is some l0{0}l_{0}\in\mathbb{Z}\setminus\{0\} such that

    lim supN|1|FN|gFNzgl0|>0.\limsup_{N}\left|\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{g}^{l_{0}}\right|>0.

    The sequence (zgl0)gG(z_{g}^{l_{0}})_{g\in G} contradicts 3, so we are done.

Suppose a sequence (zg)gG(z_{g})_{g\in G} does not satisfy 2. Taking a Følner subsequence (FN)n(F_{N}^{\prime})_{n\in\mathbb{N}} if necessary, we can assume that we have

limN1|FN|gFNzg=λ0\displaystyle\lim_{N\to\infty}\frac{1}{|F_{N}^{\prime}|}\sum_{g\in F_{N}^{\prime}}z_{g}=\lambda\neq 0
limN1|FN|gFNzhgzg¯=0 for all hH.\displaystyle\lim_{N\to\infty}\frac{1}{|F_{N}^{\prime}|}\sum_{g\in F_{N}^{\prime}}z_{hg}\overline{z_{g}}=0\text{ for all }h\in H.

This contradicts 4 by Theorem 2.7 applied with D=𝔻D=\mathbb{D} to the Cesaro averages of (zg)gG(z_{g})_{g\in G} and (zhgzg¯)gG(z_{hg}\overline{z_{g}})_{g\in G}.

The contrapositive of this implication follows from Theorem 2.7 applied to Cesaro averages of the functions (zg)gG(z_{g})_{g\in G} and (zhgzg¯)gG(z_{hg}\overline{z_{g}})_{g\in G}, with D=𝔻D=\mathbb{D}. Indeed, if HGH\subseteq G does not satisfy 4, then there is a m.p.s. (X,𝒜,μ,(Tg)gG)(X,\mathcal{A},\mu,(T_{g})_{g\in G}) and fL(μ)f\in L^{\infty}(\mu) such that, for some λ0\lambda\neq 0

Xf(Thx)f(x)¯𝑑μ(x)=0 for all hH but Xf𝑑μ=λ.\int_{X}f(T_{h}x)\cdot\overline{f(x)}d\mu(x)=0\text{ for all }h\in H\text{ but }\int_{X}fd\mu=\lambda.

So by Theorem 2.7Item 2\impliesItem 3, for any finite sets AG,H0HA\subseteq G,H_{0}\subseteq H there is KK\in\mathbb{N} and sequences (za,k)aG(z_{a,k})_{a\in G} in 𝔻\mathbb{D}, for k=1,,Kk=1,\dots,K, such that

|1K|A|k=1KaAzha,kza,k¯|<δ for all hH0,\left|\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{a\in A}z_{ha,k}\overline{z_{a,k}}\right|<\delta\text{ for all }h\in H_{0},

but

|1K|A|k=1KaAza,kλ|<δ,\left|\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{a\in A}z_{a,k}-\lambda\right|<\delta,

contradicting 4.

Suppose that 5 does not hold for some ε>0\varepsilon>0. So for any finite sets AGA\subseteq G and H0HH_{0}\subseteq H and for any δ>0\delta>0 there exist KK\in\mathbb{N} and sequences (za,k)aG(z_{a,k})_{a\in G} in 𝔻\mathbb{D}, for k=1,,Kk=1,\dots,K, such that

|1K|A|k=1KaAzha,kza,k¯|<δ for all hH0,\left|\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{a\in A}z_{ha,k}\overline{z_{a,k}}\right|<\delta\text{ for all }h\in H_{0},

but (we can assume that the following average is a positive real number)

1K|A|k=1KaAza,k>ε.\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{a\in A}z_{a,k}>\varepsilon.

In fact, we can assume the even stronger

|1K|A|k=1KaAza,kε|<δ;\left|\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{a\in A}z_{a,k}-\varepsilon\right|<\delta;

This can be achieved by adding some sequences (za,k)aG(z_{a,k})_{a\in G} (k=K+1,,K+Kk=K+1,\dots,K+K^{\prime}) with za,k=0z_{a,k}=0 for all aGa\in G.999It may also be necessary to change KK by a multiple of KK, by repeating each sequence (za,k)a(z_{a,k})_{a} (k=1,,Kk=1,\dots,K) several times We will now prove the following:

  1. ¬\neg5’

    There is ε>0\varepsilon>0 such that, for any finite sets AGA\subseteq G and H0HH_{0}\subseteq H and for any LL\in\mathbb{N} and δ>0\delta>0 there exist JJ\in\mathbb{N} and sequences (wa,j)aG(w_{a,j})_{a\in G} in 𝕊1\mathbb{S}^{1}, for j=1,,Jj=1,\dots,J, such that

    |1J|A|j=1JaAwha,jlwa,jl¯|<δ for all hH0,l=1,,L.\left|\frac{1}{J|A|}\sum_{j=1}^{J}\sum_{a\in A}w_{ha,j}^{l}\overline{w_{a,j}^{l}}\right|<\delta\text{ for all }h\in H_{0},l=1,\dots,L. (15)

    but

    |ε21J|A|j=1JaAwa,j|<δ.\left|\frac{\varepsilon}{2}-\frac{1}{J|A|}\sum_{j=1}^{J}\sum_{a\in A}w_{a,j}\right|<\delta. (16)

First we note that ¬\neg5’ implies ¬\neg1, due to Theorem 2.7 applied to the Cesaro averages of (zg)g(z_{g})_{g} and (zhglzgl¯)g(z_{hg}^{l}\overline{z_{g}^{l}})_{g}, for hHh\in H and ll\in\mathbb{Z}, and Weyl’s criterion. Let us now prove ¬\neg5’ using a probabilistic trick from [Ruz81, Section 6]

Let (za,k)aG(z_{a,k})_{a\in G}, k=1,,Kk=1,\dots,K, be as above, and let δ1>0\delta_{1}>0. We will consider a family of independent random variables (ξa,k)aG,k=1,,K(\xi_{a,k})_{a\in G,k=1,\dots,K} supported in 𝕊1\mathbb{S}^{1}, with ξa,k\xi_{a,k} having density function da,k:𝕊1[0,1];z1+Re(zza,k¯)d_{a,k}:\mathbb{S}^{1}\to[0,1];\;z\mapsto 1+\text{Re}(z\overline{z_{a,k}}). Then,

𝔼(ξa,k)=za,k2 and 𝔼(ξa,kn)=0 if n=±2,±3,.\mathbb{E}(\xi_{a,k})=\frac{z_{a,k}}{2}\text{ and }\mathbb{E}(\xi_{a,k}^{n})=0\text{ if }n=\pm 2,\pm 3,\dots. (17)

Equation 17 can be proved when za,k=1z_{a,k}=1 integrating, as we have 01e2πix(1+cos(2πx))=12\int_{0}^{1}e^{2\pi ix}(1+\cos(2\pi x))=\frac{1}{2} and for n=±2,±3,n=\pm 2,\pm 3,\dots, 01e2πinx(1+cos(2πx))=0\int_{0}^{1}e^{2\pi inx}(1+\cos(2\pi x))=0. For other values of za,kz_{a,k} one can change variables to w=zza,k¯w=z\overline{z_{a,k}}. So we have

𝔼(ξha,kξa,k¯)=zha,kza,k¯4 and 𝔼(ξha,klξa,kl¯)=0 if l=±2,±3,.\mathbb{E}(\xi_{ha,k}\overline{\xi_{a,k}})=\frac{z_{ha,k}\overline{z_{a,k}}}{4}\text{ and }\mathbb{E}(\xi_{ha,k}^{l}\overline{\xi_{a,k}^{l}})=0\text{ if }l=\pm 2,\pm 3,\dots. (18)

Now, for each mm\in\mathbb{N} we define a sequence (za,k,m)aG(z_{a,k,m})_{a\in G} by choosing, independently for all k=1,,K,aGk=1,\dots,K,a\in G and mm\in\mathbb{N}, a complex number za,k,m𝕊1z_{a,k,m}\in\mathbb{S}^{1} according to the distribution of ξa,k\xi_{a,k}. Then the strong law of large numbers implies that with probability 11 we will have, for all k=1,,K,l,hHk=1,\dots,K,l\in\mathbb{Z},h\in H and aGa\in G,

limM|𝔼(ξha,klξa,kl¯)1Mm=1Mzha,k,mlza,k,ml¯|=0.\lim_{M\to\infty}\left|\mathbb{E}\left(\xi_{ha,k}^{l}\overline{\xi_{a,k}^{l}}\right)-\frac{1}{M}\sum_{m=1}^{M}z_{ha,k,m}^{l}\overline{z_{a,k,m}^{l}}\right|=0. (19)

So we can fix a family (za,k,m)aG;k=1,,K;m(z_{a,k,m})_{a\in G;k=1,\dots,K;m\in\mathbb{N}} such that Equation 19 holds for all k,l,h,ak,l,h,a as above. Then, taking averages over all aAa\in A and k=1,,Kk=1,\dots,K, we obtain that for all hH0h\in H_{0} and ll\in\mathbb{Z}

limM|1K|A|k=1KaA𝔼(ξha,klξa,kl¯)1MK|A|k=1KaAm=1Mzha,k,mlza,k,ml¯|=0.\lim_{M\to\infty}\left|\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{a\in A}\mathbb{E}\left(\xi_{ha,k}^{l}\overline{\xi_{a,k}^{l}}\right)-\frac{1}{MK|A|}\sum_{k=1}^{K}\sum_{a\in A}\sum_{m=1}^{M}z_{ha,k,m}^{l}\overline{z_{a,k,m}^{l}}\right|=0.

Notice that, due to Equation 18, for l>1l>1 the expression in the LHS is 0, and for l=1l=1 it is 1K|A|k=1KaAzha,kza,k¯4\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{a\in A}\frac{z_{ha,k}\overline{z_{a,k}}}{4}, which has norm <δ4<\frac{\delta}{4}. So taking some big enough value MM, taking the sequences (wa,j)aG(w_{a,j})_{a\in G} to be the sequences (za,k,m)aG(z_{a,k,m})_{a\in G}, for k=1,,Kk=1,\dots,K and m=1,,Mm=1,\dots,M, Equation 15 will be satisfied with J=KMJ=KM for all hH0h\in H_{0} and l=1,,Ll=1,\dots,L. We can check similarly that, for a big enough value of MM, Equation 16 will be satisfied by the sequences (za,k,m)aG(z_{a,k,m})_{a\in G}, for k=1,,Kk=1,\dots,K and m=1,,Mm=1,\dots,M, so we are done. ∎

We will need a finitistic criterion for the notion of vdC sets in order to prove a property of vdC sets (6.4). Letting 𝔻:={z;|z|1}\mathbb{D}:=\{z\in\mathbb{C};|z|\leq 1\}, we have

Proposition 4.5.

Let GG be a countably infinite group, let HGH\subseteq G. Then HH is a vdC set in GG if and only if for any ε>0\varepsilon>0 there exists δ>0\delta>0 and a finite subset H0H_{0} of HH such that, for any m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and any measurable f:X𝔻f:X\to\mathbb{D} we have

|Xf(Thx)f(x)¯𝑑μ(x)|<δhH0 implies |Xf𝑑μ|<ε.\left|\int_{X}f(T_{h}x)\overline{f(x)}d\mu(x)\right|<\delta\;\forall h\in H_{0}\text{ implies }\left|\int_{X}fd\mu\right|<\varepsilon.
Proof.

A natural proof of this fact uses Loeb measures. In order to avoid using non-standard analysis, we will adapt the proof of [For90, Lemma 6.4].

Let HnH_{n} be an increasing sequence of finite sets with H=nHnH=\cup_{n}H_{n}. Suppose we have ε>0\varepsilon>0 and a sequence of m.p.s.’s (Xn,n,μn,(Tn,g)gG)(X_{n},\mathcal{B}_{n},\mu_{n},(T_{n,g})_{g\in G}) and measurable functions fn:X𝔻f_{n}:X\to\mathbb{D} such that

|Xnfn(Tn,hx)fn(x)¯𝑑μn(x)|<1nhHn, but |Xnfn(x)𝑑μn(x)|>ε.\left|\int_{X_{n}}f_{n}(T_{n,h}x)\overline{f_{n}(x)}d\mu_{n}(x)\right|<\frac{1}{n}\;\forall h\in H_{n},\text{ but }\left|\int_{X_{n}}f_{n}(x)d\mu_{n}(x)\right|>\varepsilon.

We will prove that HH is not a vdC set by constructing a m.p.s. (Y,𝒞,ν,(Sg)gG)(Y,\mathcal{C},\nu,(S_{g})_{g\in G}) and some measurable f:Y𝔻f_{\infty}:Y\to\mathbb{D} such that

|Yf(Shy)f(y)¯𝑑ν(y)|=0hH, but |Yf(y)𝑑ν(y)|ε.\left|\int_{Y}f_{\infty}(S_{h}y)\overline{f_{\infty}(y)}d\nu(y)\right|=0\;\forall h\in H,\text{ but }\left|\int_{Y}f_{\infty}(y)d\nu(y)\right|\geq\varepsilon. (20)

To do it, first consider the (infinite) measure space (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) with XX being the disjoint union nXn\sqcup_{n\in\mathbb{N}}X_{n}, :={nBn;Bnnn}\mathcal{B}:=\left\{\sqcup_{n}B_{n};B_{n}\in\mathcal{B}_{n}\;\forall n\right\}, μ(nBn):=nμn(Bn)\mu(\sqcup_{n}B_{n}):=\sum_{n}\mu_{n}(B_{n}) and Tgx=Tn,gxT_{g}x=T_{n,g}x if xXnx\in X_{n}. And let f:X𝔻f:X\to\mathbb{D} be given by f|Xn=fnf|_{X_{n}}=f_{n}.

Let KK be the smallest sub-CC^{*}-algebra of L(X)L^{\infty}(X) containing 11 and TgfT_{g}f for all gGg\in G. Then by the Gelfand representation theorem there is a CC^{*}-algebra isomorphism Φ:KC(Y)\Phi:K\to C(Y), where YY is the spectrum of KK, a compact metrizable space whose elements are non-zero *-homomorphisms y:Ky:K\to\mathbb{C}. The isometry Φ\Phi is given by Φ(φ)(y)=y(φ)\Phi(\varphi)(y)=y(\varphi).

We let 𝒞\mathcal{C} be the Borel σ\sigma-algebra of YY. Now consider a Banach limit L:l();(an)LlimnanL:l^{\infty}(\mathbb{N})\to\mathbb{C};(a_{n})\mapsto L-\lim_{n}a_{n} and define a norm 11 positive functional F:KF:K\to\mathbb{C} by

F(φ)=LlimnXnφ(x)𝑑μn(x).F(\varphi)=L-\lim_{n}\int_{X_{n}}\varphi(x)d\mu_{n}(x).

This induces a probability measure ν\nu on YY by YΦ(φ)𝑑ν=F(φ)\int_{Y}\Phi(\varphi)d\nu=F(\varphi). Also, the action of GG on KK induces an action (Sg)gG(S_{g})_{g\in G} of GG on YY by homeomorphisms by (Sgy)(φ)=y(φTg)(S_{g}y)(\varphi)=y(\varphi\circ T_{g}). SgS_{g} is ν\nu-preserving for all gGg\in G, as for any φK\varphi\in K we have

YΦ(φ)(Sgy)𝑑ν(y)=YΦ(φTg)(y)𝑑ν(y)=F(φTg)=F(φ)=YΦ(φ)(y)𝑑ν(y).\int_{Y}\Phi(\varphi)(S_{g}y)d\nu(y)=\int_{Y}\Phi(\varphi\circ T_{g})(y)d\nu(y)\\ =F(\varphi\circ T_{g})=F(\varphi)=\int_{Y}\Phi(\varphi)(y)d\nu(y).

We now let f:=Φ(f)C(Y)f_{\infty}:=\Phi(f)\in C(Y) and we check Equation 20: for all hHh\in H,

YΦ(f)(Shy)Φ(f)(y)¯𝑑ν(y)=YΦ(fTh)(y)Φ(f)(y)¯𝑑ν(y)=F((fTh)f¯)=LlimnXnf(Tn,hx)fn(x)¯𝑑μn(x)=0.\int_{Y}\Phi(f)(S_{h}y)\overline{\Phi(f)(y)}d\nu(y)=\int_{Y}\Phi(f\circ T_{h})(y)\overline{\Phi(f)(y)}d\nu(y)\\ =F((f\circ T_{h})\cdot\overline{f})=L-\lim_{n}\int_{X_{n}}f(T_{n,h}x)\overline{f_{n}(x)}d\mu_{n}(x)=0.
|YΦ(f)(y)𝑑ν(y)|=|F(f)|=|LlimnXnf(x)𝑑μn(x)|=Llimn|Xnf(x)𝑑μn(x)|ε.\left|\int_{Y}\Phi(f)(y)d\nu(y)\right|=\left|F(f)\right|=\left|L-\lim_{n}\int_{X_{n}}f(x)d\mu_{n}(x)\right|\\ =L-\lim_{n}\left|\int_{X_{n}}f(x)d\mu_{n}(x)\right|\geq\varepsilon.

5 Spectral Characterization of vdC sets in abelian groups

In this section we prove a spectral criterion for vdC sets in countable abelian groups (Theorem 1.14), which is a direct generalization of the spectral criterion obtained in [Ruz81, Theorem 1]; we state it in a fashion similar to [BL08, Theorem 8]. Theorem 1.14 implies that the notion of vdC set in d\mathbb{Z}^{d} defined in [BL08] is equivalent to our notion of vdC set, even if it is not defined in terms of a Følner sequence (their Definition 2 uses instead the Følner net of rectangles RM,N=[0,M]×[0,N]R_{M,N}=[0,M]\times[0,N], M,NM,N\in\mathbb{N}, ordered by RM,NRM,NR_{M,N}\geq R_{M^{\prime},N^{\prime}} if MMM\geq M^{\prime} and NNN\geq N^{\prime}). Also see [FT24, Theorem 4.3] for a different proof of Theorem 1.14, shorter than the one included here.

Theorem 1.14.

Let GG be a countable abelian group. A set HG{0}H\subseteq G\setminus\{0\} is a vdC set in GG iff any Borel probability measure μ\mu in G^\widehat{G} with μ^(h)=0hH\widehat{\mu}(h)=0\;\forall h\in H satisfies μ({1G^})=0\mu\left(\left\{1_{\widehat{G}}\right\}\right)=0.

As in [BL08, Theorem 1.8] it follows from Theorem 1.14 that, if HH is a vdC set in GG and a Borel probability measure μ\mu in G^\widehat{G} satisfies μ^(h)=0\widehat{\mu}(h)=0 for all hHh\in H, then μ({γ})=0\mu(\{\gamma\})=0 for all γG^\gamma\in\widehat{G}.

Proof.
  1. \implies

    Suppose there is a probability measure μ\mu in G^\widehat{G} such that μ^(h)=0\widehat{\mu}(h)=0 for all hHh\in H but μ({1G^})=λ>0\mu\left(\left\{1_{\widehat{G}}\right\}\right)=\lambda>0.

    Consider a sequence of finitely supported measures (μN)n(\mu_{N})_{n\in\mathbb{N}} which converge weakly to μ\mu; we can suppose μN({1G^})λ\mu_{N}\left(\left\{1_{\widehat{G}}\right\}\right)\to\lambda. Specifically, μN\mu_{N} will be an average of Dirac measures

    μN:=1uNi=1uNδxN,i,\mu_{N}:=\frac{1}{u_{N}}\sum_{i=1}^{u_{N}}\delta_{x_{N,i}},

    for some natural numbers (uN)n(u_{N})_{n\in\mathbb{N}}\to\infty and xN,1,,xN,uNG^x_{N,1},\dots,x_{N,u_{N}}\in\widehat{G}, so that xN,i=1G^G^x_{N,i}=1_{\widehat{G}}\in\widehat{G} iff iuNλi\leq u_{N}\lambda. The fact that μNμ\mu_{N}\to\mu weakly implies that for all hHh\in H (seeing hh as a map h:G^𝕊1h:\widehat{G}\to\mathbb{S}^{1}) we have

    limN1uNi=1uNxN,i(h)=limNG^h𝑑μN=G^h𝑑μ=μ^(h)=0.\lim_{N}\frac{1}{u_{N}}\sum_{i=1}^{u_{N}}x_{N,i}(h)=\lim_{N}\int_{\widehat{G}}hd\mu_{N}=\int_{\widehat{G}}hd\mu=\widehat{\mu}(h)=0. (21)

    We will prove that, letting ε<λ\varepsilon<\lambda, Item 5 of Theorem 4.1 is not satisfied. So let AGA\subseteq G and H0HH_{0}\subseteq H be finite and let (FN)(F_{N}) be a Følner sequence in GG. For each NN we consider the sequences (xN,i(ag))aA(x_{N,i}(ag))_{a\in A} for all i=1,,uNi=1,\dots,u_{N} and gFNg\in F_{N}. It will be enough to prove that

    limN1uN|FN||A|aA,gFN,i=1,,uNxN,i(ag)=λ,\lim_{N\to\infty}\frac{1}{u_{N}|F_{N}|\cdot|A|}\sum_{a\in A,g\in F_{N},i=1,\dots,u_{N}}x_{N,i}(ag)=\lambda, (22)

    and for all hH0h\in H_{0},

    limN1uN|FN||A|aA,gFN,i=1,,uNxN,i(hag)xN,i(ag)¯=0.\lim_{N\to\infty}\frac{1}{u_{N}|F_{N}|\cdot|A|}\sum_{a\in A,g\in F_{N},i=1,\dots,u_{N}}x_{N,i}(hag)\overline{x_{N,i}(ag)}=0. (23)

    Equation 23 is a direct consequence of Equation 21 and the fact that xN,i(hag)xN,i(ag)¯=xN,i(h)x_{N,i}(hag)\overline{x_{N,i}(ag)}=x_{N,i}(h). To prove Equation 22 first note that, as (FN)(F_{N}) is a Følner sequence, Equation 22 is equivalent to

    limN1uN|FN|gFN,i=1,,uNxN,i(g)=λ.\lim_{N\to\infty}\frac{1}{u_{N}|F_{N}|}\sum_{g\in F_{N},i=1,\dots,u_{N}}x_{N,i}(g)=\lambda. (24)

    Now, let δ>0\delta>0 and consider a neighborhood UU of 1G^1_{\widehat{G}} with μ(U¯)<λ+δ\mu\left(\overline{U}\right)<\lambda+\delta. After a reordering, we can assume that for big enough NN the points xN,ix_{N,i} are in G^U\widehat{G}\setminus U for all iuN(λ+2δ)i\geq u_{N}(\lambda+2\delta).

    Claim 5.1.

    There exists MM\in\mathbb{N} such that, for all xG^Ux\in\widehat{G}\setminus U and for all NMN\geq M, |1|FN|gFNx(g)|<δ\left|\frac{1}{|F_{N}|}\sum_{g\in F_{N}}x(g)\right|<\delta.

    5.1 implies Equation 24, because δ\delta is arbitrary and by 5.1 we have that for all iuNλi\leq u_{N}\lambda, the average 1|FN|gFNxN,i(g)\frac{1}{|F_{N}|}\sum_{g\in F_{N}}x_{N,i}(g) is exactly 11 (as xN,i=1x_{N,i}=1), and for all iuN(λ+2δ)i\geq u_{N}(\lambda+2\delta) and big enough NN, 1|FN|gFNxN,i(g)\frac{1}{|F_{N}|}\sum_{g\in F_{N}}x_{N,i}(g) has norm at most δ\delta.

    To prove 5.1 let ε>0\varepsilon>0 and g1,,gkGg_{1},\dots,g_{k}\in G be such that for all xG^Ux\in\widehat{G}\setminus U we have |x(gi)1|>ε|x(g_{i})-1|>\varepsilon for some i{1,,k}i\in\{1,\dots,k\}. Now note that for all NN\in\mathbb{N}, xG^Ux\in\widehat{G}\setminus U and i=1,,ki=1,\dots,k,

    |x(gi)1||1|FN|gFNx(g)|\displaystyle|x(g_{i})-1|\cdot\left|\frac{1}{|F_{N}|}\sum_{g\in F_{N}}x(g)\right| =1|FN||gFNx(gig)gFNx(g)|\displaystyle=\frac{1}{|F_{N}|}\left|\sum_{g\in F_{N}}x(g_{i}g)-\sum_{g\in F_{N}}x(g)\right|
    1|FN|gFNΔgiFN|x(g)|.\displaystyle\leq\frac{1}{|F_{N}|}\sum_{g\in F_{N}\Delta g_{i}F_{N}}\left|x(g)\right|.

    As limN|FNΔgiFN||FN|=0\lim_{N}\frac{|F_{N}\Delta g_{i}F_{N}|}{|F_{N}|}=0 for all ii, there is some MM such that for all NMN\geq M and for all ii, the right hand side is <δε<\delta\varepsilon for all xG^Ux\in\widehat{G}\setminus U. Thus, MM satisfies 5.1.

  2. \impliedby

    The proof of [BL08, Theorem 1.8, S2\impliesS1] can be adapted to any countable abelian group; instead of [BL08, Lemma 1.9] one needs to prove a statement of the form

    Lemma 5.2.

    Let (ug)gG(u_{g})_{g\in G} be a sequence of complex numbers in 𝔻\mathbb{D} and let (FN)(F_{N}) be a Følner sequence such that, for all hGh\in G, the value

    γ(h)=limN1|FN|gFNuhgug¯\gamma(h)=\lim_{N}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}u_{hg}\overline{u_{g}}

    is defined. Then there is a positive measure σ\sigma on G^\widehat{G} such that σ^(h)=γ(h)\widehat{\sigma}(h)=\gamma(h) for all hh, and

    lim supN1|FN||gFNug|σ({0}).\limsup_{N}\frac{1}{|F_{N}|}\left|\sum_{g\in F_{N}}u_{g}\right|\leq\sqrt{\sigma(\{0\})}.

    And the lemma can also be proved similarly to [BL08, Lemma 1.9], changing the functions gN,hNg_{N},h_{N} from [BL08] by

    gN(x)=1|FN||gFNzgx(g)|2 and hN(x)=1|FN||gFNg(x)|2,g_{N}(x)=\frac{1}{|F_{N}|}\left|\sum_{g\in F_{N}}z_{g}x(g)\right|^{2}\text{ and }h_{N}(x)=\frac{1}{|F_{N}|}\left|\sum_{g\in F_{N}}g(x)\right|^{2},

    and one also needs to check that [CKM77, Theorem 2] works for any countable abelian group:

    Lemma 5.3 (Cf. [CKM77, Theorem 2]).

    Let GG be a countable abelian group, let G^\widehat{G} be its dual. Let (μn)n,(νn)n,μ,ν(\mu_{n})_{n\in\mathbb{N}},(\nu_{n})_{n\in\mathbb{N}},\mu,\nu be Borel probability measures in G^\widehat{G} such that μnμ\mu_{n}\to\mu weakly, νnν\nu_{n}\to\nu weakly. Then

    ρ(μ,ν)lim supnρ(μn,νn),\rho(\mu,\nu)\geq\limsup_{n}\rho(\mu_{n},\nu_{n}),

    where ρ(μ,ν)\rho(\mu,\nu), the affinity between μ\mu and ν\nu, is given by, for any measure mm such that μ,ν\mu,\nu are absolutely continuous with respect to mm,

    ρ(μ,ν)=G^(dμdm)12(dνdm)12𝑑m.\rho(\mu,\nu)=\int_{\widehat{G}}\left(\frac{d\mu}{dm}\right)^{\frac{1}{2}}\left(\frac{d\nu}{dm}\right)^{\frac{1}{2}}dm.

    The same proof of [CKM77, Theorem 2] is valid; the proof uses the existence of Radon-Nikodym derivatives and a countable partition of unity fj:Tf_{j}:T\to\mathbb{R}, for jj\in\mathbb{Z}. These partitions of unity always exist for outer regular Radon measures (see [Rud87, Theorem 3.14]), so they exist for any Borel probability measure in G^\widehat{G}.∎

6 Properties of vdC sets

In [BL08] and [Ruz81], several properties of the family of vdC subsets of d\mathbb{Z}^{d} were proved. In this section we check that many of these properties hold for vdC sets in any countable group. Some of the statements about vdC sets follow from statements about sets of recurrence and the fact that any vdC set is a set of recurrence. Other properties can be proved in the same way as their analogs for sets of recurrence, even if they are not directly implied by them.

Remark 6.1.

All the properties below are also satisfied for sets of operatorial recurrence, as defined in [FT24] (see [FT24, Theorem 4.4]), with the proofs of the properties being essentially the same as for vdC sets.

Proposition 6.2 (Partition regularity of vdC sets, cf. [Ruz81, Cor. 1]).

Let GG be a countably infinite group and let H,H1,H2G{e}H,H_{1},H_{2}\subseteq G\setminus\{e\} satisfy H=H1H2H=H_{1}\cup H_{2}. If HH is a vdC set in GG, then either H1H_{1} or H2H_{2} are vdC sets in GG.

Proof.

Suppose that H1,H2H_{1},H_{2} are not vdC sets in GG. Then for i=1,2i=1,2 there are measure preserving systems (Xi,i,μi,(Tig)gG)(X_{i},\mathcal{B}_{i},\mu_{i},(T_{i}^{g})_{g\in G}), and functions fiL(μi)f_{i}\in L^{\infty}(\mu_{i}) such that

Xifi(Tihx)fi(x)¯𝑑μi(x)=0 for all hHi,\int_{X_{i}}f_{i}(T_{i}^{h}x)\cdot\overline{f_{i}(x)}d\mu_{i}(x)=0\text{ for all }h\in H_{i},

but

Xifi(x)𝑑μi(x)0.\int_{X_{i}}f_{i}(x)d\mu_{i}(x)\neq 0.

But then, considering the function f:X1×X2;f(x1,x2)=f1(x1)f2(x2)f:X_{1}\times X_{2}\to\mathbb{C};f(x_{1},x_{2})=f_{1}(x_{1})\cdot f_{2}(x_{2}), we have by Fubini’s theorem

X1×X2f1(T1hx1)f2(T2hx2)f1(x1)¯f2(x2)¯d(μ1×μ2)(x1,x2)=0 for all hH1H2,\int_{X_{1}\times X_{2}}f_{1}(T_{1}^{h}x_{1})f_{2}(T_{2}^{h}x_{2})\overline{f_{1}(x_{1})}\overline{f_{2}(x_{2})}d(\mu_{1}\times\mu_{2})(x_{1},x_{2})=0\text{ for all }h\in H_{1}\cup H_{2},

but

X1×X2f1(x1)f2(x2)d(μ1×μ2)(x1,x2)0,\int_{X_{1}\times X_{2}}f_{1}(x_{1})f_{2}(x_{2})d(\mu_{1}\times\mu_{2})(x_{1},x_{2})\neq 0,

so H1H2H_{1}\cup H_{2} is not a vdC set in GG. ∎

We will need the following in order to prove 6.4.

Proposition 6.3.

Let GG be a countably infinite group, and let HG{e}H\subseteq G\setminus\{e\} be finite. Then HH is not a set of recurrence in GG, so it is not vdC.

Proof.

Consider the (12,12)\left(\frac{1}{2},\frac{1}{2}\right)-Bernoulli scheme in {0,1}G\{0,1\}^{G} with the action (Tg)gG(T_{g})_{g\in G} of GG given by Tg((xa)aG)=(xag)aGT_{g}((x_{a})_{a\in G})=(x_{ag})_{a\in G}. Let A={(xa);xe=0}{0,1}GA=\{(x_{a});x_{e}=0\}\subseteq\{0,1\}^{G} and let B=AhHTh1AB=A\setminus\cup_{h\in H}T_{h^{-1}}A. As HH is finite, BB has positive measure, but clearly μ(BTh1B)=0\mu(B\cap T_{h^{-1}}B)=0 for all hHh\in H, so we are done. ∎

The following generalizes [Ruz81, Cor. 3], and has a similar proof.

Proposition 6.4.

Let GG be a countably infinite group and let HG{e}H\subseteq G\setminus\{e\} be a vdC set in GG. Then we can find infinitely many disjoint vdC subsets of HH.

Proof of 6.4.

It will be enough to prove that there are two disjoint vdC subsets H,H′′H^{\prime},H^{\prime\prime} of HH.

We will define a disjoint sequence H1,H2,H_{1},H_{2},\dots of finite subsets of HH by recursion. Suppose H1,,Hn1H_{1},\dots,H_{n-1} are given; notice that i=1n1Hi\bigcup_{i=1}^{n-1}H_{i} is finite, so it is not a vdC set. So by 6.2, Hi=1n1HiH\setminus\bigcup_{i=1}^{n-1}H_{i} is a vdC set. Then by 4.5 we can then define HnH_{n} to be a finite subset of Hi=1n1HiH\setminus\bigcup_{i=1}^{n-1}H_{i} such that for some constant δn>0\delta_{n}>0 and for any m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}), fL(X,μ)f\in L^{\infty}(X,\mu) we have

|Xf(Thx)f(x)¯𝑑μ(x)|<δnhHn implies |Xf𝑑μ|<1n.\left|\int_{X}f(T_{h}x)\overline{f(x)}d\mu(x)\right|<\delta_{n}\;\forall h\in H_{n}\text{ implies }\left|\int_{X}fd\mu\right|<\frac{1}{n}.

Now let H=nH2n1H^{\prime}=\bigcup_{n\in\mathbb{N}}H_{2n-1} and H′′=nH2nH^{\prime\prime}=\bigcup_{n\in\mathbb{N}}H_{2n}. It is then clear that both HH^{\prime} and H′′H^{\prime\prime} satisfy the definition of vdC set in GG, so we are done. ∎

The following generalizes [BL08, Corollary 1.15.2] to countable amenable groups.

Theorem 6.5.

Let SS be a subgroup of a countable amenable group GG, let HS{e}H\subseteq S\setminus\{e\}. Then HH is a vdC set in SS iff it is a vdC set in GG.

Proof.

If HH is not a vdC set in GG, then it is not a vdC set in SS, as any measure preserving action (Tg)gG(T_{g})_{g\in G} on a measure space restricts to a measure preserving action (Tg)gS(T_{g})_{g\in S} on the same measure space.

The fact that if HH is not a vdC set in SS then it is not a vdC set in GG can be deduced from Item 5 of Theorem 4.1: indeed, let ε\varepsilon be as in Item 5 (applied to the group SS) and consider any AGA\subseteq G and H0HH_{0}\subseteq H finite and any δ>0\delta>0.

We can express A=A1AmA=A_{1}\cup\dots\cup A_{m}, where the AiA_{i} are pairwise disjoint and of the form Ai=A(Sgi)A_{i}=A\cap(Sg_{i}), for some giGg_{i}\in G. Thus, Agi1SAg_{i}^{-1}\subseteq S for all ii.

Now, for each ii we know that there exist KiK_{i}\in\mathbb{N} and GG-sequences (zi,a,k)aS(z_{i,a,k})_{a\in S} in 𝔻\mathbb{D}, for k=1,,Kik=1,\dots,K_{i}, such that

|1Ki|Ai|k=1KiaAigi1zi,ha,kzi,a,k¯|<δ for all hH0,\left|\frac{1}{K_{i}|A_{i}|}\sum_{k=1}^{K_{i}}\sum_{a\in A_{i}g_{i}^{-1}}z_{i,ha,k}\overline{z_{i,a,k}}\right|<\delta\text{ for all }h\in H_{0}, (25)

but

|1Ki|Ai|k=1KiaAigi1zi,a,k|>ε.\left|\frac{1}{K_{i}|A_{i}|}\sum_{k=1}^{K_{i}}\sum_{a\in A_{i}g_{i}^{-1}}z_{i,a,k}\right|>\varepsilon. (26)

Note that we can assume K1,K2,,KmK_{1},K_{2},\dots,K_{m} are all equal to some number KK (e.g. taking KK to be the least common multiple of all of them) and that 1K|Ai|k=1KaAigi1zi,a,k\frac{1}{K|A_{i}|}\sum_{k=1}^{K}\sum_{a\in A_{i}g_{i}^{-1}}z_{i,a,k} is a positive real number for all i=1,,mi=1,\dots,m (multiplying the sequences (zi,a,k)(z_{i,a,k}) by some complex number of norm 11 if needed).

Finally, define for each k=1,,Kk=1,\dots,K a sequence (za,k)aG(z_{a,k})_{a\in G} by za,k=zi,agi1,kz_{a,k}=z_{i,ag_{i}^{-1},k} for aSgia\in Sg_{i} and by zg=0z_{g}=0 elsewhere. This sequence will satisfy Item 5 of Theorem 4.1 (by taking averages of Equations 25 and 26 for k=1,,Kk=1,\dots,K), so we are done. ∎

The following is proved in [BL08, Cor. 1.15.1] for vdC sets in d\mathbb{Z}^{d} using the spectral criterion; using Definition 1.13 instead we prove it for any countable group.

Proposition 6.6.

Let π:GS\pi:G\to S be a group homomorphism, let HH be a vdC set in GG. If eSπ(H)e_{S}\not\in\pi(H), then π(H)\pi(H) is a vdC set in SS.

Proof.

The contra-positive of the claim follows easily from the fact that any measure preserving action (Ts)sS(T_{s})_{s\in S} on a probability space (X,𝒜,μ)(X,\mathcal{A},\mu) induces a measure preserving action (Sg)gG(S_{g})_{g\in G} on (X,𝒜,μ)(X,\mathcal{A},\mu) by Sg=Tπ(g)S_{g}=T_{\pi(g)}. ∎

Remark 6.7.

In 6.6, π(H)\pi(H) may be a vdC set even if HH is not. Indeed, the set {(n,1);n}\{(n,1);n\in\mathbb{N}\} is not a set of recurrence in 2\mathbb{Z}^{2} (this follows from 6.8 below), but its projection to the first coordinate is a vdC set.

The following generalizes [BL08, Corollary 1.16].

Corollary 6.8.

Let GG be a countable group and let SS be a finite index subgroup of GG. Then GSG\setminus S is not a set of recurrence in GG, so it is not a vdC set in GG.

Proof.

Consider the action of GG on the finite set G/SG/S of left-cosets of SS, where we give G/SG/S the uniform probability measure μ\mu. The set {S}G/S\{S\}\subseteq G/S has positive measure, but for all gGSg\in G\setminus S we have g{S}{S}=g\{S\}\cap\{S\}=\varnothing. ∎

We finally prove that difference sets are nice vdC sets (see e.g. [Far22, Lemma 5.2.8] for the case G=G=\mathbb{Z}). The proof of 6.9 is just the proof that any set of differences is a set of recurrence101010A slight modification of this proof shows that AA1AA^{-1} is a set of nice recurrence (see Definition 2.1), which is already found in [Fur81, Page 74] for the case G=G=\mathbb{Z}.

Proposition 6.9.

Let GG be a countable group and let AGA\subseteq G be infinite. Then the difference set AA1={ba1;a,bA,ab}AA^{-1}=\{ba^{-1};a,b\in A,a\neq b\} is a nice vdC set in GG.

Proof.

Suppose that for some m.p.s. (X,𝒜,μ,(Tg)gG)(X,\mathcal{A},\mu,(T_{g})_{g\in G}) and some function fL(μ)f\in L^{\infty}(\mu) we have

|Xf𝑑μ|2>lim suphAA1|Xf(Thx)f(x)¯𝑑μ(x)|.\left|\int_{X}fd\mu\right|^{2}>\limsup_{h\in AA^{-1}}\left|\int_{X}f(T_{h}x)\overline{f(x)}d\mu(x)\right|. (27)

So for some λ<|Xf𝑑μ|2\lambda<\left|\int_{X}fd\mu\right|^{2} and some finite subset BAA1B\subseteq AA^{-1} we have

|Xf(Tax)f(Tbx)¯𝑑μ(x)|<λ for all a,bA with ab1B.\left|\int_{X}f(T_{a}x)\cdot\overline{f(T_{b}x)}d\mu(x)\right|<\lambda\text{ for all }a,b\in A\text{ with }ab^{-1}\not\in B.

Then for any NN\in\mathbb{N}, letting A0A_{0} be a subset of AA with NN elements and denoting Taf=fTaT_{a}f=f\circ T_{a}, we have

N2|Xf𝑑μ|2=|aA0Taf,1|21,1\displaystyle N^{2}\left|\int_{X}fd\mu\right|^{2}=\frac{\left|\left\langle\sum_{a\in A_{0}}T_{a}f,1\right\rangle\right|^{2}}{\langle 1,1\rangle} aA0Taf,aA0Taf\displaystyle\leq\left\langle\sum_{a\in A_{0}}T_{a}f,\sum_{a\in A_{0}}T_{a}f\right\rangle
N|B|fL2(X,μ)2+N2λ.\displaystyle\leq N\cdot|B|\cdot\|f\|^{2}_{L^{2}(X,\mu)}+N^{2}\lambda.

This is a contradiction for big enough NN, because λ<|Xf𝑑μ|2\lambda<|\int_{X}fd\mu|^{2}. ∎

The corollary below was suggested by V. Bergelson. Before stating it, recall that a subset HH of a countable group GG is thick when for any finite AGA\subseteq G, HH contains a right translate of AA. In particular, a subset of a countable amenable group has upper Banach density 11 iff it is thick. Also note that a set HH is a vdC set in GG if and only if HH1H\cup H^{-1} is a vdC set in GG; this follows from the fact that for any m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and any fL(X,μ)f\in L^{\infty}(X,\mu) we have Xf(Thx)f(x)¯𝑑μ=Xf(Th1x)f(x)¯𝑑μ¯\int_{X}f(T_{h}x)\overline{f(x)}d\mu=\overline{\int_{X}f(T_{h^{-1}}x)\overline{f(x)}d\mu}.

Corollary 6.10.

If GG is a countable group and a subset HG{e}H\subseteq G\setminus\{e\} is thick, then HH is a nice vdC set.

Proof.

If HGH\subseteq G is thick, then there is an infinite set A={a1,a2,}A=\{a_{1},a_{2},\dots\} such that HH contains the set {anam1;m>n}\{a_{n}a_{m}^{-1};m>n\}. Indeed, we can define (an)n(a_{n})_{n\in\mathbb{N}} be recursion by letting ana1,,an1a_{n}\neq a_{1},\dots,a_{n-1} be such that Anan1HA_{n}a_{n}^{-1}\subseteq H, where An:={a1,,an1}A_{n}:=\{a_{1},\dots,a_{n-1}\}.

Thus HH1H\cup H^{-1} contains the set AA1AA^{-1}, so HH1H\cup H^{-1} is a vdC set, so HH is a vdC set. ∎

7 Convexity and Cesaro averages

We prove in 7.1 below that there is a close relationship between the correlation functions of sequences taking values in a compact set DD\subseteq\mathbb{C}, and in the convex hull of DD. We then explain several applications of this result, including an answer to a question by Kelly and Lê. The arguments involving the law of large numbers which we use to prove 7.1 are based on [Ruz81, Section 6].

Proposition 7.1.

Let GG be a countably infinite amenable group with a Følner sequence F=(FN)F=(F_{N}). Let DD\subseteq\mathbb{C} be compact and let CC\subseteq\mathbb{C} be the convex hull of DD. Then for any sequence (zg)gG(z_{g})_{g\in G} of complex numbers in CC there is a sequence (wg)gG(w_{g})_{g\in G} in DD such that, for any kk\in\mathbb{N} and any pairwise distinct elements h1,,hkGh_{1},\dots,h_{k}\in G, we have

limN1|FN|gFNwh1gwhkg=limN1|FN|gFNzh1gzhkg,\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}w_{h_{1}g}\cdots w_{h_{k}g}=\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{h_{1}g}\cdots z_{h_{k}g},

whenever the right hand side is defined.

Proof.

Consider a list (hl,1,,hl,jl)(h_{l,1},\dots,h_{l,j_{l}}), ll\in\mathbb{N}, of all the finite sequences of pairwise distinct elements of GG, and let pl:jlp_{l}:\mathbb{C}^{j_{l}}\to\mathbb{C}; pl(z1,,zjl)=z1z2zjlp_{l}(z_{1},\dots,z_{j_{l}})=z_{1}z_{2}\cdots z_{j_{l}}.

Consider a Følner subsequence (FNi)i(F_{N_{i}})_{i} of FF such that for all ll\in\mathbb{N}, the limit

γ(l):=limi1|FNi|gFNipl(zhl,1g,,zhl,jlg)\gamma(l):=\lim_{i\to\infty}\frac{1}{|F_{N_{i}}|}\sum_{g\in F_{N_{i}}}p_{l}(z_{h_{l,1}g},\dots,z_{h_{l,j_{l}}g})

is defined. Then by 3.1, for all AGA\subseteq G finite and for all L,δ>0L\in\mathbb{N},\delta>0 there exist some KK\in\mathbb{N} and some sequences (zg,k)gG(z_{g,k})_{g\in G} in CC, for k=1,,Kk=1,\dots,K, such that for all l=1,,Ll=1,\dots,L we have

|γ(l)1K|A|k=1KgAzhl,1g,kzhl,jlg,k|<δ.\left|\gamma(l)-\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{g\in A}z_{h_{l,1}g,k}\dots z_{h_{l,j_{l}}g,k}\right|<\delta. (28)
Claim 7.2.

For every sequence 𝐱=(xg)gG\mathbf{x}=(x_{g})_{g\in G} taking values in CC, any finite AGA\subseteq G and any L,δ>0L\in\mathbb{N},\delta>0 there is some K𝐱K_{\mathbf{x}}\in\mathbb{N} and sequences (xg,k)gG(x_{g,k})_{g\in G} in DD, k=1,,K𝐱k=1,\dots,K_{\mathbf{x}}, such that for all gAg\in A and all l=1,,Ll=1,\dots,L we have

|1|A|gAxhl,1gxhl,jlg1K𝐱|A|k=1K𝐱xhl,1g,kxhl,jlg,k|<δ.\left|\frac{1}{|A|}\sum_{g\in A}x_{h_{l,1}g}\cdots x_{h_{l,j_{l}}g}-\frac{1}{K_{\mathbf{x}}|A|}\sum_{k=1}^{K_{\mathbf{x}}}x_{h_{l,1}g,k}\cdots x_{h_{l,j_{l}}g,k}\right|<\delta.

We prove 7.2 below. It follows from 7.2 and Equation 28 that for all AGA\subseteq G finite, LL\in\mathbb{N} and δ>0\delta>0 there exist some KK\in\mathbb{N} and sequences (zg,k)gG(z_{g,k})_{g\in G} in DD, for k=1,,Kk=1,\dots,K, such that for all l=1,,Ll=1,\dots,L we have

|γ(l)1K|A|k=1KgAzhl,1g,kzhl,jlg,k|<δ.\left|\gamma(l)-\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{g\in A}z_{h_{l,1}g,k}\dots z_{h_{l,j_{l}}g,k}\right|<\delta.

Thus, by 3.1 there exists a sequence (wg)gG(w_{g})_{g\in G} of elements of DD such that, for all ll\in\mathbb{N},

limN1|FN|gFNwhl,1gwhl,jlg=limN1|FN|gFNpl(whl,1g,,whl,jlg)=γ(l),\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}w_{h_{l,1}g}\cdots w_{h_{l,j_{l}}g}=\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}p_{l}(w_{h_{l,1}g},\dots,w_{h_{l,j_{l}}g})=\gamma(l),

so we are done.

It only remains to prove 7.2, so let 𝐱=(xg)gG\mathbf{x}=(x_{g})_{g\in G} take values in CC. Note that the set of extreme points of CC is contained in DD, so by Choquet’s theorem, for every xCx\in C there is a probability measure μ\mu supported in DD and with average xx. So we can consider for each gGg\in G a random variable ξg\xi_{g} supported in DD and satisfying 𝔼(ξg)=xg\mathbb{E}(\xi_{g})=x_{g}. Note that, if the variables (ξg)gG(\xi_{g})_{g\in G} are pairwise independent and h1,,hkGh_{1},\dots,h_{k}\in G are distinct, then we have

𝔼(ξh1ξhk)=𝔼(ξh1)𝔼(ξhk)=xh1xhk.\mathbb{E}(\xi_{h_{1}}\cdots\xi_{h_{k}})=\mathbb{E}(\xi_{h_{1}})\cdots\mathbb{E}(\xi_{h_{k}})=x_{h_{1}}\cdots x_{h_{k}}.

Now for each kk\in\mathbb{N} we will choose a sequence (xg,k)gG(x_{g,k})_{g\in G}, where the variables xg,kx_{g,k} are chosen pairwise independently and with distribution ξg\xi_{g}. Then, for all AGA\subseteq G finite, LL\in\mathbb{N} and δ>0\delta>0, by the strong law of large numbers we will have with probability 11 that, for big enough KK,

|1|A|gAxhl,1gxhl,jlg1K|A|k=1Kxhl,1g,kxhl,jlg,k|<δ,\left|\frac{1}{|A|}\sum_{g\in A}x_{h_{l,1}g}\cdots x_{h_{l,j_{l}}g}-\frac{1}{K|A|}\sum_{k=1}^{K}x_{h_{l,1}g,k}\cdots x_{h_{l,j_{l}}g,k}\right|<\delta,

concluding the proof.∎

Applying 7.1 to the set D={0,1}D=\{0,1\} gives the following result.

Proposition 7.3.

Let GG be a countably infinite amenable group with a Følner sequence F=(FN)F=(F_{N}). Then for any sequence (zg)gG(z_{g})_{g\in G} of numbers in [0,1][0,1] there is a set BGB\subseteq G such that, for any kk\in\mathbb{N} and any pairwise distinct elements h1,,hkGh_{1},\dots,h_{k}\in G, we have

dF(h11Bhk1B)=limN1|FN|gFNzh1gzhkg,d_{F}({h_{1}}^{-1}B\cap\dots\cap{h_{k}}^{-1}B)=\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{h_{1}g}\cdots z_{h_{k}g},

whenever the right hand side is defined.∎

It follows that, given a measurable function f:X[0,1]f:X\to[0,1], we can obtain a function g:X{0,1}g:X\to\{0,1\} (that is, a measurable set) with the same correlation functions:

Proposition 7.4 (Turning functions into sets).

Let GG be a countably infinite amenable group. For every m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and every measurable f:X[0,1]f:X\to[0,1] there exists a m.p.s. (Y,𝒞,ν,(Sg)gG)(Y,\mathcal{C},\nu,(S_{g})_{g\in G}) and BB\in\mathcal{B} such that for all kk\in\mathbb{N} and all distinct h1,,hkGh_{1},\dots,h_{k}\in G we have

ν(Th11BThk1B)=Xf(Th1x)f(Thkx)𝑑μ.\nu\left(T_{h_{1}^{-1}}B\cap\dots\cap T_{h_{k}^{-1}}B\right)=\int_{X}f(T_{h_{1}}x)\cdots f(T_{h_{k}}x)d\mu.
Proof.

Let (X,,μ,(Tg)gG),f(X,\mathcal{B},\mu,(T_{g})_{g\in G}),f be as in 7.4, and fix a Følner sequence F=(FN)F=(F_{N}) in GG. Thanks to Theorem 1.6 there is a [0,1][0,1]-valued sequence (zg)gG(z_{g})_{g\in G} such that, for all kk\in\mathbb{N} and h1,,hkGh_{1},\dots,h_{k}\in G,

limN1|FN|gFNzh1g,,zhjg=Xf(Th1x),,f(Thjx)dμ.\lim_{N}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{h_{1}g},\dots,z_{h_{j}g}=\int_{X}f(T_{h_{1}}x),\dots,f(T_{h_{j}}x)d\mu.

Thus, by 7.3 there is a set BGB\subseteq G such that, for all kk\in\mathbb{N} and h1,,hkGh_{1},\dots,h_{k}\in G,

dF(h11Bhk1B)=Xf(Th1x),,f(Thjx)dμ.d_{F}({h_{1}}^{-1}B\cap\dots\cap{h_{k}}^{-1}B)=\int_{X}f(T_{h_{1}}x),\dots,f(T_{h_{j}}x)d\mu.

Applying Theorem 1.4 to the set BB, we are done. ∎

Remark 7.5.

The elements h1,,hkh_{1},\dots,h_{k} need to be distinct in 7.4; if we applied the result to h1=h2=eh_{1}=h_{2}=e, we would obtain ν(B)=Xf𝑑μ=Xf2𝑑μ\nu(B)=\int_{X}fd\mu=\int_{X}f^{2}d\mu, which only happens if ff is essentially a characteristic function.

Note that the statement of 7.4 makes sense for non-amenable groups. It seems plausible to us that 7.4 could be proved for any group GG using some purely measure-theoretic construction:

Question 7.6.

Is 7.4 true for any countable group GG?

The last result in this section, which we prove after 7.1, is about sequences of norm 11 which are ‘very well distributed’ in 𝕊1\mathbb{S}^{1}:

Proposition 7.7 (White noise).

Let GG be a countable amenable group with a Følner sequence (FN)(F_{N}). There is a sequence (zg)gG(z_{g})_{g\in G} of complex numbers in 𝕊1\mathbb{S}^{1} such that, for all k,n1,,nk{0}k\in\mathbb{N},n_{1},\dots,n_{k}\in\mathbb{Z}\setminus\{0\} and distinct elements h1,,hkGh_{1},\dots,h_{k}\in G we have

limN1|FN|gFNzh1gn1zhkgnk=0.\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{h_{1}g}^{n_{1}}\cdots z_{h_{k}g}^{n_{k}}=0. (29)
Proof.

By 3.1, 3\implies1, it is enough to prove that for all AGA\subseteq G finite, h1,,hjGh_{1},\dots,h_{j}\in G and for all L,δ>0L\in\mathbb{N},\delta>0 there is some KK\in\mathbb{N} and sequences (zg,k)gG(z_{g,k})_{g\in G} in DD, for k=1,,Kk=1,\dots,K, such that we have

|1K|A|k=1KgAzh1g,kl1,,zhjg,klj|<δ for all l1,,lj{0},|li|L.\left|\frac{1}{K|A|}\sum_{k=1}^{K}\sum_{g\in A}z_{h_{1}g,k}^{l_{1}},\dots,z_{h_{j}g,k}^{l_{j}}\right|<\delta\textup{ for all }l_{1},\dots,l_{j}\in\mathbb{Z}\setminus\{0\},|l_{i}|\leq L. (30)

Consider an independent sequence of random variables (ξg)gG(\xi_{g})_{g\in G}, such that ξg\xi_{g} is uniformly distributed in 𝕊1\mathbb{S}^{1} for all gGg\in G. Clearly, for all k,n1,,nk{0}k\in\mathbb{N},n_{1},\dots,n_{k}\in\mathbb{Z}\setminus\{0\} and distinct h1,,hkGh_{1},\dots,h_{k}\in G we have

𝔼(ξh1n1ξhknk)=0.\mathbb{E}(\xi_{h_{1}}^{n_{1}}\cdots\xi_{h_{k}}^{n_{k}})=0.

So, as in the proof of 7.1, for each kk\in\mathbb{N} we choose a sequence (xg,k)gG(x_{g,k})_{g\in G}, where the variables xg,kx_{g,k} are chosen pairwise independently and with distribution ξg\xi_{g}. The strong law of large numbers then implies that for big enough KK, Equation 30 will be satisfied with high probability, concluding the proof. ∎

Remark 7.8.

If the sequence (FN)(F_{N}) from 7.7 does not grow very slowly (e.g. if we have Nα|FN|<\sum_{N}\alpha^{|F_{N}|}<\infty for all α<1\alpha<1), then we can choose each of the elements zgz_{g} from the sequence (zg)gG(z_{g})_{g\in G} independently and uniformly from 𝕊1\mathbb{S}^{1}. The sequence we obtain will satisfy 7.7 with probability 11.

Remark 7.9.

Let G=G=\mathbb{Z}. If we consider only finitely many values of kk, then we can construct sequences (zn)n(z_{n})_{n\in\mathbb{Z}} which satisfy Equation 29 for all Følner sequences simultaneously. For example, letting zn=en8αz_{n}=e^{n^{8}\alpha}, where α\alpha\in\mathbb{R}\setminus\mathbb{Q}, one can check that Equation 29 is satisfied for all k8k\leq 8, for all ni,hin_{i},h_{i} as above and, most importantly, for all Følner sequences (FN)(F_{N}). However, there is no sequence (zn)n(z_{n})_{n\in\mathbb{Z}} which satisfies Equation 29 for all kk\in\mathbb{N} and for all Følner sequences FF and (ni)i=1k,(hi)i=1k(n_{i})_{i=1}^{k},(h_{i})_{i=1}^{k} as above. Indeed, suppose that such a sequence (zn)n(z_{n})_{n\in\mathbb{Z}} exists. Then for all NN\in\mathbb{N}, the sequence n(zn+1,,zn+N)n\mapsto(z_{n+1},\dots,z_{n+N}) has to be u.d. in (𝕊1)N(\mathbb{S}^{1})^{N} (see [KN74, Chapter 1, Theorem 6.2]). But that implies that for some nNn_{N}\in\mathbb{N}, the numbers znN+1,,znN+Nz_{n_{N}+1},\dots,z_{n_{N}+N} all have positive real part. Thus, letting FN={nN+1,,nN+N}F_{N}=\{n_{N}+1,\dots,n_{N}+N\}, the sequence (zn)(z_{n}) is not (FN)(F_{N})-u.d., a contradiction.

As was pointed out to us by S. Farhangi, we can use 7.1 to answer a question of Kelly and Lê. We first need some definitions.

Definition 7.10.

We say a sequence (xn)n(x_{n})_{n\in\mathbb{N}} of elements of a compact topological group GG, with Haar measure μG\mu_{G}, is u.d. in GG if

limN1N|{n{1,,N};xnC}|=μG(C)\lim_{N}\frac{1}{N}|\{n\in\{1,\dots,N\};x_{n}\in C\}|=\mu_{G}(C)

for all open sets CGC\subseteq G with boundary of measure 0.

Definition 7.11.

Let GG be a compact topological group. We say a set HH\subseteq\mathbb{N} is GG-u.d.vdC111111We use the notation ‘GG-u.d.vdC’ to avoid confusion with ‘vdC in GG’, as in Definition 1.13. if for any sequence (xn)n=1(x_{n})_{n=1}^{\infty} in GG such that (xn+hxn1)n(x_{n+h}x_{n}^{-1})_{n\in\mathbb{N}} is u.d. in GG for all hHh\in H, the sequence (xn)n(x_{n})_{n\in\mathbb{N}} is also u.d. in GG.

In [KL18, Page 2] the authors mention that it is an interesting (and perhaps difficult) problem to determine whether all 2\mathbb{Z}_{2}-u.d.vdC sets are vdC.

Proposition 7.12.

A set HH\subseteq\mathbb{N} is vdC iff it is 2\mathbb{Z}_{2}-u.d.vdC.

7.12 follows from 7.14 and the observation that a sequence (xn)n(x_{n})_{n\in\mathbb{N}} is u.d. in 2\mathbb{Z}_{2} iff limN1Nn=1N(1)xn=0\lim_{N}\frac{1}{N}\sum_{n=1}^{N}(-1)^{x_{n}}=0.

Proposition 7.13.

Let HH\subseteq\mathbb{N}. HH is vdC iff for every sequence (xn)n(x_{n})_{n\in\mathbb{Z}} with xn[1,1]x_{n}\in[-1,1] for all nn we have

limN1Nn=1Nxn+hxn=0 for all hH implies limN1Nn=1Nxn=0.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}x_{n+h}x_{n}=0\text{ for all }h\in H\textup{ implies }\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}x_{n}=0. (31)
Proof.

The following argument was provided by S. Farhangi. \implies follows from Item 2 in Theorem 4.1. So let us prove that, if every sequence (xn)n(x_{n})_{n\in\mathbb{Z}} in [1,1][-1,1] satisfies Equation 31, then every sequence (xn)n(x_{n})_{n\in\mathbb{Z}} in 𝔻\mathbb{D} satisfies Equation 31, with xn+hxn¯x_{n+h}\overline{x_{n}} instead of xn+hxnx_{n+h}x_{n} (so HH is vdC by Theorem 4.1). Thanks to Theorem 2.7, it will be enough to prove the measure theoretic analog of our result: we will assume that for all m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and measurable f:X[1,1]f:X\to[-1,1] we have

Xf(Thx)f(x)¯𝑑μ(x)=0 for all hH implies Xf𝑑μ=0,\int_{X}f(T_{h}x)\cdot\overline{f(x)}d\mu(x)=0\text{ for all }h\in H\text{ implies }\int_{X}fd\mu=0, (32)

and we will prove that Equation 32 holds true for all m.p.s. (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and measurable f:X𝔻f:X\to\mathbb{D}.

So, fix (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and f:X𝔻f:X\to\mathbb{D} such that Xf(Thx)f(x)¯𝑑μ(x)=0\int_{X}f(T_{h}x)\cdot\overline{f(x)}d\mu(x)=0 for all hHh\in H. We let f0,f1:X[1,1]f_{0},f_{1}:X\to[-1,1] be the real and imaginary parts of ff respectively. Consider the set Y=X×{0,1}Y=X\times\{0,1\} with the product σ\sigma-algebra 𝒞=×𝒫({0,1})\mathcal{C}=\mathcal{B}\times\mathcal{P}(\{0,1\}), measure ν(A0×{0}A1×{1})=μ(A0)+μ(A1)2\nu(A_{0}\times\{0\}\cup A_{1}\times\{1\})=\frac{\mu(A_{0})+\mu(A_{1})}{2} for A0,A1A_{0},A_{1}\in\mathcal{B}, and action (Sg)gG(S_{g})_{g\in G} given by Sg(x,i)=(Tgx,i)S_{g}(x,i)=(T_{g}x,i) for xX,i{0,1}x\in X,i\in\{0,1\}.

Finally, let F:Y[1,1];F(x,i)=fi(x)F:Y\to[-1,1];F(x,i)=f_{i}(x). Then for all hHh\in H,

YF(Shy)F(y)𝑑ν(y)\displaystyle\int_{Y}F(S_{h}y)F(y)d\nu(y) =12Xf0(Thx)f0(x)+f1(Thx)f1(x)dμ(x)\displaystyle=\frac{1}{2}\int_{X}f_{0}(T_{h}x)f_{0}(x)+f_{1}(T_{h}x)f_{1}(x)d\mu(x)
=12Re(Xf(Thx)f(x)¯𝑑μ(x))=0.\displaystyle=\frac{1}{2}\textup{Re}\left(\int_{X}f(T_{h}x)\overline{f(x)}d\mu(x)\right)=0.

Equation 32 then implies YF𝑑μ=0\int_{Y}Fd\mu=0, so Xf0+f1dμ=0\int_{X}f_{0}+f_{1}d\mu=0. Similarly, we can conclude Xf1+f0dμ=0\int_{X}-f_{1}+f_{0}d\mu=0 by applying the same reasoning to the function g=if=f1+if0g=if=-f_{1}+if_{0} instead of ff. Thus, Xf0𝑑μ=Xf1𝑑μ=0\int_{X}f_{0}d\mu=\int_{X}f_{1}d\mu=0, so Xf𝑑μ=0\int_{X}fd\mu=0. ∎

Proposition 7.14.

A set HH\subseteq\mathbb{N} is vdC iff for every sequence (xn)n(x_{n})_{n\in\mathbb{Z}} with xn{1,1}x_{n}\in\{-1,1\} for all nn we have

limN1Nn=1Nxn+hxn=0 for all hH implies limN1Nn=1Nxn=0.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}x_{n+h}x_{n}=0\text{ for all }h\in H\textup{ implies }\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}x_{n}=0.
Proof.

\implies follows from Item 2 in Theorem 4.1. So suppose HH is not vdC. Then by 7.13 there is a sequence (xn)n(x_{n})_{n\in\mathbb{Z}} in [1,1][-1,1], some λ0\lambda\neq 0 and some Følner subsequence (FN)(F_{N}) of ({1,,N})(\{1,\dots,N\}), such that

limN1|FN|nFNxn+hxn=0 for all hH but limN1|FN|gFNxg=λ.\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{n\in F_{N}}x_{n+h}x_{n}=0\text{ for all }h\in H\textup{ but }\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}x_{g}=\lambda.

Thus, by 7.1 there is a sequence (wn)n(w_{n})_{n\in\mathbb{Z}} in {1,1}\{-1,1\} such that

limN1|FN|nFNwn+hwn=0 for all hH but limN1|FN|nFNwn=λ.\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{n\in F_{N}}w_{n+h}w_{n}=0\text{ for all }h\in H\textup{ but }\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{n\in F_{N}}w_{n}=\lambda.

So by 3.1 (which implies that the set of correlation functions of sequences in a given compact set is independent of the Følner sequence), there is a sequence (un)n(u_{n})_{n\in\mathbb{Z}} in {1,1}\{-1,1\} such that

limN1Nn=1Nun+hun=0 for all hH but limN1Nn=1Nun=λ.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}u_{n+h}u_{n}=0\text{ for all }h\in H\textup{ but }\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}u_{n}=\lambda.\qed

References

  • [BBD25a] Raimundo Briceño, Álvaro Bustos-Gajardo, and Miguel Donoso-Echenique. Extensibility and denseness of periodic semigroup actions. arXiv preprint arXiv:2502.00312, 2025.
  • [BBD25b] Raimundo Briceño, Álvaro Bustos-Gajardo, and Miguel Donoso-Echenique. Natural extensions of embeddable semigroup actions. arXiv preprint arXiv:2501.05536, 2025.
  • [BBF10] Mathias Beiglböck, Vitaly Bergelson, and Alexander Fish. Sumset phenomenon in countable amenable groups. Advances in Mathematics, 223(2):416–432, 2010.
  • [Ber86] Vitaly Bergelson. A density statement generalizing Schur’s theorem. Journal of Combinatorial Theory, Series A, 43(2):338–343, 1986.
  • [Ber96] Vitaly Bergelson. Ergodic Ramsey theory—an update. Ergodic theory of d\mathbb{Z}^{d} actions, 228:1–61, 1996.
  • [BF21a] Vitaly Bergelson and Andreu Ferré-Moragues. An ergodic correspondence principle, invariant means and applications. Israel Journal of Mathematics, pages 1–42, 2021.
  • [BF21b] Vitaly Bergelson and Andreu Ferré-Moragues. Uniqueness of a Furstenberg system. Proceedings of the American Mathematical Society, 149(7):2983–2997, 2021.
  • [BL08] Vitaly Bergelson and Emmanuel Lesigne. Van der Corput sets in d\mathbb{Z}^{d}. Colloquium Mathematicum, 110:1–49, 2008.
  • [Bou87] Jean Bourgain. Ruzsa’s problem on sets of recurrence. Israel Journal of Mathematics, 59(2):150–166, 1987.
  • [CKM77] Jean Coquet, Teturo Kamae, and Michel Mendès-France. Sur la mesure spectrale de certaines suites arithmétiques. Bulletin de la Société Mathématique de France, 105:369–384, 1977.
  • [CP61] Alfred H Clifford and Gordon B Preston. The algebraic theory of semigroups, vol. 1. American Mathematical Society, Providence, 2(7), 1961.
  • [DHZ19] Tomasz Downarowicz, Dawid Huczek, and Guohua Zhang. Tilings of amenable groups. Journal für die reine und angewandte Mathematik, 2019(747):277–298, 2019.
  • [Don24] Miguel Donoso-Echenique. Natural extensions and periodic approximations of semigroup actions. Master Dissertation, Pontificia Universidad Católica de Chile, 2024.
  • [Far22] Sohail Farhangi. Topics in ergodic theory and Ramsey theory. Doctoral dissertation, The Ohio State University, 2022.
  • [FJM24] Sohail Farhangi, Steve Jackson, and Bill Mance. Undecidability in the ramsey theory of polynomial equations and hilbert’s tenth problem. arXiv preprint arXiv:2412.14917, 2024.
  • [For90] Alan Hunter Forrest. Recurrence in dynamical systems: a combinatorial approach. Doctoral dissertation, The Ohio State University, 1990.
  • [FS24] Alexander Fish and Sean Skinner. An inverse of Furstenberg’s correspondence principle and applications to nice recurrence. arXiv preprint arXiv:2407.19444, 2024.
  • [FT24] Sohail Farhangi and Robin Tucker-Drob. Asymptotic dynamics on amenable groups and van der Corput sets. arXiv preprint arXiv:2409.00806, 2024.
  • [Fur77] Harry Furstenberg. Ergodic behavior of diagonal measures and a theorem of Szemerédi on arithmetic progressions. Journal d’Analyse Mathématique, 31(1):204–256, 1977.
  • [Fur81] Harry Furstenberg. Recurrence in Ergodic Theory and Combinatorial Number Theory. Princeton University Press, 1981.
  • [KL18] Michael Kelly and Thái Hoang Lê. Van der Corput sets with respect to compact groups. Archiv der Mathematik, 110(4), 2018.
  • [KM78] Teturo Kamae and Michel Mendès-France. Van der Corput’s difference theorem. Israel Journal of Mathematics, 31:335–342, 1978.
  • [KN74] Lauwerens Kuipers and Harald Niederreiter. Uniform Distribution of Sequences. John Wiley, 1974.
  • [Mor] Joel Moreira. Sets of nice recurrence. I Can’t Believe It’s Not Random! https://joelmoreira.wordpress.com/2013/03/04/323/.
  • [Ore31] Oystein Ore. Linear equations in non-commutative fields. Annals of Mathematics, 32(3):463–477, 1931.
  • [Par05] Kalyanapuram Rangachari Parthasarathy. Probability measures on metric spaces, volume 352. American Mathematical Soc., 2005.
  • [Rud62] Walter Rudin. Fourier analysis on groups. Interscience, 1962.
  • [Rud87] Walter Rudin. Real and complex analysis. McGraw-Hill international editions: Mathematics series, 1987.
  • [Ruz81] Imre Ruzsa. Connections between the uniform distribution of a sequence and its differences. Colloquia Mathematica Societatis János Bolyai, 34:1419–1443, 1981.
  • [Sze75] Endre Szemerédi. On sets of integers containing no k elements in arithmetic progression. Acta Arithmetica, 27(199-245):2, 1975.
  • [Wey16] Hermann Weyl. Über die Gleichverteilung von Zahlen mod. Eins. Mathematische Annalen, 77(3):313–352, 1916.