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
[email protected]
Abstract

In this article we give characterizations of the notions of van der Corput (vdC) set, nice vdC set and set of nice recurrence (defined below) in countable amenable groups. This allows us to prove that nice vdC sets are sets of nice recurrence and that vdC sets are independent of the Følner sequence used to define them, answering questions from [BL] in the context of countable amenable groups. We also give a spectral characterization of vdC sets in abelian groups. The methods developed in this paper allow us to establish a converse to the Furstenberg correspondence principle. In addition, we introduce vdC sets in general non amenable groups and establish some basic properties of them, such as partition regularity.

Several results in this paper, including the converse to Furstenberg’s correspondence principle, have also been proved independently by Robin Tucker-Drob and Sohail Farhangi in the article [FT], which is being uploaded to arXiv simultaneously to this one.

1 Introduction

In this paper we obtain new results about van der Corput (vdC) sets. Among other things, we will provide answers to some questions posed in [BL] and [BF]. The notion of van der Corput set was introduced in [KM] in connection with the study 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 1.1 ([KM, 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} is itself u.d. mod 11.

In [Ru] the following characterization of vdC sets in terms of Cesaro averages of bounded sequences of complex numbers was established:

Theorem 1.2 (cf. [Ru, Theorem 1]).

A set HH\subseteq\mathbb{N} is a van der Corput 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 such that

hH,limN1Nn=1Nzh+nzn¯=0,\forall h\in H,\quad\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}z_{h+n}\overline{z_{n}}=0,

we have

limN1Nn=1Nzn=0.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}z_{n}=0.

Kamae and Mendès France proved in [KM, Theorem 2] that any vdC set HH\subseteq\mathbb{N} is a set of recurrence, meaning that for any m.p.s.111From now, by ‘m.p.s.’ we mean probability measure preserving system. (X,,μ,T)(X,\mathcal{B},\mu,T) and any AA\in\mathcal{B} with μ(A)>0\mu(A)>0 we have μ(AThA)>0\mu(A\cap T^{h}A)>0 for some hHh\in H. It was observed in [KM] that many classical examples of sets of recurrence, such as sets of differences, images of polynomial functions and shifted primes, are vdC sets. This led to the question of whether the notions of set of recurrence and vdC set coincide. It was Bourgain who answered this question in the negative by giving in [Bo] a rather complicated example of a set of recurrence which is not vdC (also see [Mou] for a more refined example).

In [BL] the notion of vdC set was extended to subsets of d\mathbb{Z}^{d} (d2d\geq 2), and a natural notion of ‘FF-vdC set’ was introduced for any Følner sequence222From now, by ‘Følner sequence’ we mean left-Følner sequence. F=(FN)NF=(F_{N})_{N} in \mathbb{Z}. The definition of FF-vdC given in [BL, Section 4.2] naturally extends to Følner sequences in any countable amenable group.

Definition 1.3.

Let F=(FN)NF=(F_{N})_{N\in\mathbb{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.
Definition 1.4 (cf. [BL, Page 44]).

Let F=(FN)NF=(F_{N})_{N\in\mathbb{N}} be a Følner sequence in a countable amenable group GG. We say a subset HH of GG 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 particular, any set HH containing 1G1_{G} is FF-vdC. Note that a subset HH of \mathbb{N} is vdC in the sense of Definition 1.1 iff it is (FN)N(F_{N})_{N}-vdC in \mathbb{Z}, where FN={1,,N}F_{N}=\{1,\dots,N\}. It was asked in [BL, Section 4.2] whether, for any Følner sequence FF in \mathbb{Z}, a subset HH\subseteq\mathbb{N} is FF-vdC if and only if it is vdC. We show that the answer is yes by giving the following characterization of FF-vdC sets in a countable amenable group GG, which does not depend on the Følner sequence:

Theorem 1.5 (See Theorem 3.1).

Let GG be a countably infinite amenable group with a Følner sequence F=(FN)NF=(F_{N})_{N}. A set HGH\subseteq G 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(Th(x))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.

The condition given in Theorem 1.5 makes sense for any countable group333The 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., not only for amenable ones. We introduce it below as a definition of vdC subset of a countable group. This makes the notation ‘vdC set’ independent of the Følner sequence and allows us to prove properties of vdC sets in a more general setting.

Definition 1.6.

Let GG be a (discrete) countable group. We will say that HGH\subseteq G 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(Th(x))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 1.6 it easily follows that any vdC set in a countable group GG is of 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 1.6 becomes the definition of set of recurrence. This supports the idea implied by [BL, Section 3.2] that sets of recurrence are a ‘positive version’ of vdC sets.

In Theorem 1.7 below we give a spectral characterization of vdC subsets of any countable abelian group GG. This generalizes similar characterizations which were given in [KM, BL, Ru] for vdC subsets of \mathbb{Z} or d\mathbb{Z}^{d} for d2d\geq 2 (see e.g. [BL, Theorem 1.8.]). These characterizations are useful both for proving properties of vdC sets and for finding (non-) examples of them. For example, Bourgain used a version of Theorem 1.7 for G=G=\mathbb{Z} to construct his set of recurrence which is not vdC.

For a countable abelian group GG we denote by 0 its identity element and by G^\widehat{G} its Pontryagin dual (which is compact and metrizable, see [RuFA, Theorems 1.2.5, 2.2.6]). Also, 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)\widehat{\mu}(h)=\int_{\widehat{G}}\gamma(h)d\mu(\gamma) (for hGh\in G).

Theorem 1.7 below is proved in this article; another very elegant proof of it can be found in [FT].

Theorem 1.7 (cf. [BL, Theorem 1.8]).

Let GG be a countable abelian group. A set HGH\subseteq G is vdC 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.

Taking definition Definition 1.6 as the starting point, we will prove in Section 5 several properties of vdC sets. Most of them were proved in [Ru] or [BL] for vdC sets in \mathbb{Z} and d\mathbb{Z}^{d} respectively. [Ru, BL] used the spectral criterion to prove many of these properties (which we avoid), so their proofs would only work for countable abelian groups. We now state some of the properties we prove in Section 5.

We show that the family of vdC sets in a countable group satisfies the Ramsey property444Also known as partition regularity., that is, if HGH\subseteq G is vdC and H=H1H2H=H_{1}\cup H_{2}, then either H1H_{1} or H2H_{2} is vdC. We also prove that, if a vdC set in a countable group does not contain the identity, then it contains two disjoint vdC sets. Proving that requires a finitistic criterion for the notion of vdC sets; letting 𝔻:={z;|z|1}\mathbb{D}:=\{z\in\mathbb{C};|z|\leq 1\}, we have

Proposition 1.8.

Let GG be a countably infinite group, let HGH\subseteq G. Then HH is vdC 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.

We give a proof of 1.8 at the end of Section 5. We also study the behaviour of vdC sets in subgroups and under group homomorphisms and give some easy examples and non-examples of vdC sets.

We now discuss another question in [BL]. The following definition was introduced in [Be2] for subsets of \mathbb{Z}.

Definition 1.9 (cf. [Be2, 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 [BL, Definition 10], a notion of ‘nice vdC set’ was introduced:

Definition 1.10 (cf. [BL, Definition 10]).

Let GG be a countable amenable group with a Følner sequence (FN)N(F_{N})_{N}. A subset HH of GG 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|

Note that nice vdC sets are vdC by Theorem 1.2. Also note that Definition 1.10 and Definition 1.9 are expressed in different settings: Definition 1.10 is about Cesaro averages of sequences of complex numbers, while Definition 1.9 is about integrals. We will translate each of these definitions to the setting of the other one in Propositions 1.12 and 1.13 below. First, we need to recall some notions of density.

Definition 1.11 ([BF, Definitions 2.1, 2.2]).

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

dF¯(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)N(F_{N})_{N}. The upper Banach density of EE, d(E)d^{*}(E), is defined by

d(E)=sup{dF¯(E);F Følner sequence in G}.d^{*}(E)=\sup\left\{\overline{d_{F}}(E);F\text{ F{\o}lner sequence in }G\right\}.
Proposition 1.12.

Let GG be a countable amenable group with a Følner sequence F=(FN)NF=(F_{N})_{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 1.13.

Let GG be a countable amenable group with a Følner sequence (FN)N(F_{N})_{N}. Then a subset HH of GG 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|.

In [BL, 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 [BL] 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). For subsets of \mathbb{N}, it was proved by S. Farhangi that nice vdC implies nice recurrence in [Fa, Theorem 5.2.4]; both 1.13 and 1.12 generalize this fact to all amenable groups.

Note also that 1.13 also implies that the notion of nice FF-vdC set is independent of the Følner sequence.

As we mentioned above, both 1.12 and 1.13 are just translations of Definitions 1.9 and 1.10 between the language of Cesaro averages and the language of integrals. Both of them, as well as Theorem 1.5, are easy corollaries of the following theorem, which is the machinery behind most results in this article:

Theorem 1.14.

Let GG be a countably infinite amenable group with a Følner sequence (FN)N(F_{N})_{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. (1)

Conversely, given a sequence (zg)gG(z_{g})_{g\in G} in DD there is a m.p.s. (X,,μ,T)(X,\mathcal{B},\mu,T) 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 1 holds if the limit in the left hand side exists.

Note that Theorem 1.14 is about bounded sequences in \mathbb{C}; in some cases one may adapt it to sequences (zg)gG(z_{g})_{g\in G} such that lim supN1|FN|gFNzg2<\limsup_{N}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}\|z_{g}\|^{2}<\infty, see [FT, Theorem 3.3].

We remark that in our study of vdC sets we only need cases j=1,2j=1,2 of Theorem 1.14. However, the proof is essentially the same for all jj, and allowing jj to take arbitrary values allows us to obtain applications to the theory of multiple recurrence. In particular, we can use Theorem 1.14 to prove Theorem 1.16 below, a converse to the Furstenberg correspondence principle which we can use to answer a question from [BF]. We state the Furstenberg correspondence principle and Theorem 1.16 below.

Theorem 1.15 (Furstenberg correspondence principle, cf. [Be5, Theorem 1.8]).

Let GG be a countable amenable group, let AGA\subseteq G. Then there exists 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(A)\mu(B)=d^{*}(A) such that, for all kk\in\mathbb{N} and h1,,hkGh_{1},\dots,h_{k}\in G one has:

d(h1AhkA)μ(Th1(B)Thk(B))d^{*}(h_{1}A\cap\dots\cap h_{k}A)\geq\mu(T_{h_{1}}(B)\cap\dots\cap T_{h_{k}}(B))
Theorem 1.16.

Let GG be a countably infinite amenable group with a Følner sequence (FN)N(F_{N})_{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)

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 2 for all k,h1,,hkk,h_{1},\dots,h_{k} as above whenever dF(h1AhkA)d_{F}\left(h_{1}A\cap\dots\cap h_{k}A\right) exists.

Theorem 1.16 can be viewed as a strengthening of [BF2, Theorem 5.1]: in our case, we do not need the m.p.s. to be ergodic, and we do not need to take a subsequence of (FN)N(F_{N})_{N}. In [Av, Theorem 3.1] they also give a (weaker) inverse of the Furstenberg correspondence in the case G=G=\mathbb{Z}.

As an application of Theorem 1.16 we prove Theorem 1.17 below, which answers a question asked in [BF] after Remark 3.6. The question is 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}.

Theorem 1.17.

Let GG be a countably infinite amenable group with a Følner sequence F=(FN)NF=(F_{N})_{N}. Then there is 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}.

The structure of the article is as follows. In Section 2 we prove Theorem 1.14, which is the machinery behind Theorem 1.5. At the end of the section we prove several corollaries of Theorem 1.14: Theorem 2.14 (a generalization of Theorem 1.16), Propositions 1.12 and 1.13, and Theorem 1.17. In Section 3 we give a characterization theorem for FF-vdC sets in countable amenable groups, Theorem 3.1, which expands upon Theorem 1.5. We also prove in 3.5 that there is a close relationship between Cesaro averages of sequences taking values in a compact set DD\subseteq\mathbb{C}, and Cesaro averages of sequences in the convex hull of DD. In Section 4 we prove the spectral characterization of vdC sets in countable abelian groups, Theorem 1.7. In Section 5 we prove the properties of vdC sets mentioned above and 1.8.

Acknowledgements.

Special thanks to Vitaly Bergelson for his guidance while writing this article, and for some interesting discussions and suggestions. The author gratefully acknowledges support from the grants BSF 2020124 and NSF CCF AF 2310412.

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 [FT] contains a long list of characterizations of vdC sets, including the ones from Theorem 1.5 and Theorem 1.7. Thanks to Sohail Farhangi for his input and suggestions, and especially for bringing [DHZ, Theorem 5.2] to my attention; this result allowed the author to simplify the proof of Theorem 1.7 and to state Theorem 1.14 in full generality (for all amenable groups instead of only monotileable ones).

Some of the results contained in this article (in particular Theorem 1.14 and Theorem 1.16) and in [FT] were announced by Sohail Farhangi in the Conference on Ergodic group actions and unitary representations at IMPAN, June 2024.

After this article was completed, the author was made aware of the existence of the article [FS], which also touches upon an inverse Furstenberg correspondence principle. In this article, among other things, Fish and Skinner obtain the particular case of Theorem 1.16 which corresponds to the Følner sequence FN={1,,N}F_{N}=\{1,\dots,N\} in \mathbb{Z} (see [FS, Theorem 1.4]).

2 Correspondence between Cesaro and integral averages

The main objective of this section is proving Theorem 1.14. It will be deduced from 2.1, a more technical version of Theorem 1.14 which also includes a finitistic criterion for the existence of sequences with given Cesaro averages.

At the end of the section we prove Theorem 2.14, the converse of a general converse of the Furstenberg correspondence principle. We also give proofs of Propositions 1.12 and 1.13, answering a question of [BL] by proving that nice vdC sets are sets of nice recurrence, and of Theorem 1.17, answering a question of [BF].

Proposition 2.1.

Let GG be a countably infinite amenable group with a Følner sequence (FN)N(F_{N})_{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). (3)
  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,1(x)),,f(Thl,jl(x))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).
  3. 3.

    (Finitistic criterion) For all AGA\subseteq G finite and for all L,δ>0L\in\mathbb{N},\delta>0 there exists 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. (4)
Remark 2.2.

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; to do it, one can let NN be big enough, let K=|FN|K=|F_{N}| and for all kFNk\in F_{N} let (zg,k)g(z_{g,k})_{g} be given by zg,k=zgkz_{g,k}=z_{gk}, where (zg)g(z_{g})_{g} satisfies 1.

Proof of Theorem 1.14 from 2.1.

Let G,F=(FN)N,D,(X,,μ,(Tg)gG)G,F=(F_{N})_{N},D,(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and f:XDf:X\to D be as in Theorem 1.14.

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 2.1 to the sequence

γ(l)=Xpl(f(Thl,1(x)),,f(Thl,jl(x))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,1(x)),,f(Thl,jl(x))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). (5)

This implies that Equation 1 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.

For the converse we can use the same argument with the dense sequence (pl)l(p_{l})_{l\in\mathbb{N}} from above, except that we first need to take a Følner subsequence (FN)N(F_{N}^{\prime})_{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}).\qed

We will now prove 2.1. We will show 1\implies2\implies3\implies1. The complicated implication is 3\implies1. During the rest of the section we fix an amenable group GG, a Følner sequence (FN)N(F_{N})_{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}.

Now, for each NN consider the average of Dirac measures νN=1|FN|gFNδRg𝐳\nu_{N}=\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 the function f:DGf:D^{G}\to\mathbb{C} by f((za)aG)=z1f((z_{a})_{a\in G})=z_{1}. Then for all ll\in\mathbb{N} we have

DGpl(f(Rhl,1(x)),,f(Rhl,jl(x))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,1(x)),,f(Rhl,jl(x))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.

Let (X,,μ,(Tg)gG)(X,\mathcal{B},\mu,(T_{g})_{g\in G}) and ff be as in 2, and let (DG,(DG),(Rg)gG)(D^{G},\mathcal{B}(D^{G}),(R_{g})_{g\in G}) be 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 in a compact metric space, ν\nu can be weakly approximated by a sequence of finitely supported measures (νN)N(\nu_{N})_{N}, which we may assume are of the form

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

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)a)\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})_{a})
=\displaystyle= 1|A|gADGpl(zhl,1g,,zhl,jlg)𝑑ν((za)a)\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})_{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=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=\frac{1}{|A|}\sum_{g\in A}\gamma(l)=\gamma(l).

So for any LL\in\mathbb{N} and δ>0\delta>0, Equation 4 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 2.3.

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.

Indeed, the LHS above is equal to

||E||E||E||E|gEzg+1|E|gEEzg||E||E||E||E|M|E|+1|E|M|EE|<Mδ+Mδ.\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|\\ \leq\frac{|E|-|E^{\prime}|}{|E|\cdot|E^{\prime}|}\cdot M|E^{\prime}|+\frac{1}{|E|}\cdot M|E\setminus E^{\prime}|<M\delta+M\delta.\qed
Definition 2.4.

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 2.5.

Note that if (FN)N(F_{N})_{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 2.1. 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 be a weighted average of the average values of the fif_{i}.

Lemma 2.6.

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|1Mδ\frac{|\partial_{S}T_{k}|}{|T_{k}|}\leq\frac{1}{M\delta} 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}|}\cdot\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.

To abbreviate, for any sequence (za)aG(z_{a})_{a\in G} we will denote p((za)a)=p(zh1,,zhj)p((z_{a})_{a})=p(z_{h_{1}},\dots,z_{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| =|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|
=|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|<1Mδ\frac{|T_{k}c_{k}\cap\partial_{S}T_{k}c_{k}|}{|T_{k}|}<\frac{1}{M\delta} 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)-\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 2.3 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 main tool we will need in the proof of 3\implies1 is [DHZ, Theorem 5.2]; in order to state it we first recall some definitions:

Definition 2.7 (cf. [DHZ, 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 1G1_{G}.

  • 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 2.8 (cf. [DHZ, Page 17]).

We say that 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 2.9.

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.666This is not the definition of (B,ε)(B,\varepsilon)-invariant used in [DHZ], but it will be more convenient for our purposes.

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

Theorem 2.10 (cf. [DHZ, 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 2.11.

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 2.12.

Let GG be a countable amenable group with a tiling 𝒯\mathcal{T} and a Følner sequence (FN)N(F_{N})_{N}. As 𝒯FNS𝒮(𝒯)SFN\partial_{\mathcal{T}}F_{N}\subseteq\cup_{S\in\mathcal{S}(\mathcal{T})}\partial_{S}F_{N}, Remark 2.5 implies that

limN|𝒯FN||FN|=0.\lim_{N\to\infty}\frac{|\partial_{\mathcal{T}}F_{N}|}{|F_{N}|}=0.
Lemma 2.13.

Let GG be a countably infinite amenable group with a Følner sequence (FN)N(F_{N})_{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-invariant 777By 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. (6)

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} (7)

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 7, 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 number in {1,2,,KS}\{1,2,\dots,K_{S}\}, and such that for each SS, the same number of right-translates of SS have value each of the numbers 1,2,,KS1,2,\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 2.6 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}}})-\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 6, 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 2.10, 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}. (8)

Now, letting (FN)N(F_{N})_{N} be the Følner sequence of Item 1, we let 𝒫\mathcal{P} and (AN)N(A_{N})_{N} be as in 2.13, 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 2.6 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})-\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}. (9)

However, the double sum in Equation 9 is at distance ε10\leq\frac{\varepsilon}{10} of γ(l)\gamma(l) (this follows from taking an affine combination of Equation 8 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 the Furstenberg correspondence principle. Theorem 2.14 is a slight generalization of Theorem 1.16 in which we also allow intersections with complements of the sets giAg_{i}A. This makes Theorem 2.14 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\}. Indeed, any element of 𝒜\mathcal{A} is a finite disjoint union of elements of the form (i=1kgiA)(i=k+1nGgiA)\left(\bigcap_{i=1}^{k}g_{i}A\right)\cap\left(\bigcap_{i=k+1}^{n}G\setminus g_{i}A\right), where nn\in\mathbb{N}, k{0,1,,n}k\in\{0,1,\dots,n\} and g1,,gnGg_{1},\dots,g_{n}\in G.

Theorem 2.14.

Let GG be a countably infinite amenable group with a Følner sequence (FN)N(F_{N})_{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}, k{0,1,,n}k\in\{0,1,\dots,n\} and g1,,gnGg_{1},\dots,g_{n}\in G we have

dF((i=1kgiA)(i=k+1nGgiA))=μ((i=1kTgiB)(i=k+1nXTgiB)).d_{F}\left(\left(\bigcap_{i=1}^{k}g_{i}A\right)\cap\left(\bigcap_{i=k+1}^{n}G\setminus g_{i}A\right)\right)=\mu\left(\left(\bigcap_{i=1}^{k}T_{g_{i}}B\right)\cap\left(\bigcap_{i=k+1}^{n}X\setminus T_{g_{i}}B\right)\right). (10)

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 10 for all n,k,g1,,gnn,k,g_{1},\dots,g_{n} as above such that the density in Equation 10 exists.

Proof.

Note that for all nn\in\mathbb{N}, k{0,1,,n}k\in\{0,1,\dots,n\} and g1,,gnGg_{1},\dots,g_{n}\in G we have

μ((i=1kTgiB)(i=k+1nXTgiB))=X(i=1kχB(Tgi1x))(i=k+1n(1χB(Tgi1x)))𝑑μ.\displaystyle\mu\left(\left(\bigcap_{i=1}^{k}T_{g_{i}}B\right)\cap\left(\bigcap_{i=k+1}^{n}X\setminus T_{g_{i}}B\right)\right)=\int_{X}\left(\prod_{i=1}^{k}\chi_{B}\left(T_{g_{i}^{-1}}x\right)\right)\cdot\left(\prod_{i=k+1}^{n}\left(1-\chi_{B}\left(T_{g_{i}^{-1}}x\right)\right)\right)d\mu.

So by Theorem 1.14 applied to the polynomials of the form p(x1,,xj)=x1x2xk(1xk+1)(1xn)p(x_{1},\dots,x_{j})=x_{1}x_{2}\dots x_{k}(1-x_{k+1})\dots(1-x_{n}) 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 nn\in\mathbb{N}, k{0,1,,n}k\in\{0,1,\dots,n\} and g1,,gnGg_{1},\dots,g_{n}\in G we have

μ((i=1kTgiB)(i=k+1nXTgiB))\displaystyle\mu\left(\left(\bigcap_{i=1}^{k}T_{g_{i}}B\right)\cap\left(\bigcap_{i=k+1}^{n}X\setminus T_{g_{i}}B\right)\right)
=limN1|FN|gFN(i=1kχA(gi1g))(i=k+1n(1χA(gi1g)))\displaystyle=\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}\left(\prod_{i=1}^{k}\chi_{A}(g_{i}^{-1}g)\right)\cdot\left(\prod_{i=k+1}^{n}(1-\chi_{A}(g_{i}^{-1}g))\right)
=dF((i=1kgiA)(i=k+1nGgiA)).\displaystyle=d_{F}\left(\left(\bigcap_{i=1}^{k}g_{i}A\right)\cap\left(\bigcap_{i=k+1}^{n}G\setminus g_{i}A\right)\right).

The other implication is proved similarly, applying Theorem 1.14 in the other direction. ∎

Remark 2.15.

It is possible to give a version of Theorem 1.16 that involves translates of several sets B1,,BlXB_{1},\dots,B_{l}\subseteq X, as in [BF2, Definition A.3], or even countably many sets (Bk)k(B_{k})_{k\in\mathbb{N}}. But we will not do so as it is not the main theme of this article and it would involve proving a modified version of Theorem 1.14.

Proof of 1.12.
  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),dF(EhE)d_{F^{\prime}}(E),d_{F^{\prime}}(E\cap hE) exist 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 1.14 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 1.14 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 1.13 is essentially the same as that of 1.12, using Theorem 1.14 to prove both directions.

Proof of Theorem 1.17.

We can change the set gAgE\cup_{g\in A}gE by gAgE\cap_{g\in A}gE by considering the complement of EE. Also, it is enough to prove Theorem 1.17 for just one Følner sequence FF in GG. Indeed, suppose there is EGE\subseteq G such that for all finite AG\varnothing\neq A\subseteq G we have dF(E)=dF(gAgE)=12d_{F}(E)=d_{F}\left(\cap_{g\in A}gE\right)=\frac{1}{2}. Then for any other Følner sequence FF^{\prime}, by Theorem 1.16 applied two times we can find a set EE^{\prime} such that dF(E)=dF(E)d_{F^{\prime}}(E)=d_{F}(E) and dF(gAgE)=dF(gAgE)d_{F^{\prime}}(\cap_{g\in A}gE)=d_{F}(\cap_{g\in A}gE) for all finite AGA\subseteq G.

So we just need to choose a Følner sequence (FN)N(F_{N})_{N} for which the theorem is true. Let (AN)N(A_{N})_{N} be a Følner sequence in GG and let FN=ANgN,1ANgN,2F_{N}=A_{N}g_{N,1}\cup A_{N}g_{N,2}, where gN,1,gN,2Gg_{N,1},g_{N,2}\in G are chosen so that the sets ANgN,iA_{N}g_{N,i}, for i=1,2i=1,2 and NN\in\mathbb{N}, are pairwise disjoint. And let E=NANgN,1E=\cup_{N}A_{N}g_{N,1}. We then have dF(E)=dF(gAgE)=12d_{F}(E)=d_{F}\left(\cap_{g\in A}gE\right)=\frac{1}{2} for all finite AGA\subseteq G, as we wanted. ∎

3 Characterizations of vdC sets in amenable groups

In this section we prove Theorem 3.1 below, our main characterization theorem for vdC sets in countable amenable groups. Theorem 3.1 gives a characterization of FF-vdC sets analogous to Theorem 1.2 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 [BL, Section 4.2] of whether FF-vdC implies vdC.

Theorem 3.1.

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

  1. 1.

    HH is an FF-vdC set.

  2. 2.

    If a sequence (zg)gG(z_{g})_{g\in G} of complex numbers in the unit disk 𝔻\mathbb{D} satisfies

    limN1|FN|gFNzhgzg¯=0 for all hH,\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,

    then

    limN1|FN|gFNzg=0.\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{g}=0.
  3. 3.

    If a sequence (zg)gG(z_{g})_{g\in G} of complex numbers in 𝕊1\mathbb{S}^{1} satisfies

    limN1|FN|gFNzhgzg¯=0 for all hH,\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,

    then

    limN1|FN|gFNzg=0.\lim_{N\to\infty}\frac{1}{|F_{N}|}\sum_{g\in F_{N}}z_{g}=0.
  4. 4.

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

    X(f(Th(x)))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. (11)
  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 3.2.

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 1.6, the resulting definition of vdC set in GG is equivalent. A similar question is posed in [FT, Conjecture 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 3.4 below, which was introduced by Weyl in [We].

Definition 3.3.

Let GG be a countable amenable group with a Følner sequence (FN)N(F_{N})_{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}.

Thus, a sequence (xg)gG(x_{g})_{g\in G} in 𝕋\mathbb{T} is u.d. mod 11 iff (e2πixg)gG(e^{2\pi ix_{g}})_{g\in G} is FF-u.d. in 𝕊1\mathbb{S}^{1}, where F=({1,,N})NF=(\{1,\dots,N\})_{N}.

Proposition 3.4 (Weyl’s criterion for uniform distribution).

Let GG be a countable amenable group with a Følner sequence (FN)N(F_{N})_{N}. A sequence (zg)gG(z_{g})_{g\in G} in 𝕊1\mathbb{S}^{1} is FF-u.d. in 𝕊1\mathbb{S}^{1} if and only if 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 3.4 see e.g. [KN, Theorem 2.1] (the proof works for any Følner sequence).

Proof of Theorem 3.1.

We prove 2\implies2\implies2\implies2\implies2\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} 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 1.14 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 1.14 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 5, 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

X(f(Th(x)))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. (12)

So by Theorem 1.14Item 1\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 ¬\neg5 holds 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 G888It 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. ¬\neg4’.

    There is ε>0\varepsilon>0 such that, for any finite sets AGA\subseteq G and H0HH_{0}\subseteq H and for any L,δ>0L\in\mathbb{N},\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. (13)

    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. (14)

First we note that ¬\neg4’. implies ¬\neg1, due to Theorem 1.14 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 ¬\neg4’. using a probabilistic trick from [Ru, 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. (15)

Equation 15 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. (16)

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. (17)

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 17 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 16, 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 13 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 14 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. ∎

As a consequence of Theorem 3.12\iff3, we see that there are two equivalent definitions of FF-vdC sets, one using sequences of complex numbers in 𝕊1\mathbb{S}^{1} and one using sequences in 𝔻\mathbb{D}. This can be seen as a consequence of the fact that 𝔻\mathbb{D} is the convex hull of 𝕊1\mathbb{S}^{1}. In fact, using the same idea of the proof of 1\implies5 in Theorem 3.1, we prove in 3.5 below that there is a close relationship between Cesaro averages of sequences taking values in a compact set DD\subseteq\mathbb{C}, and Cesaro averages of sequences in the convex hull of DD.

Proposition 3.5.

Let GG be a countably infinite amenable group with a Følner sequence F=(FN)NF=(F_{N})_{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 2.1, 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. (18)
Claim 3.6.

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.

Suppose that 3.6 is true. It then follows easily from 3.6 and Equation 18 that 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 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 2.1 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 3.6, 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 central limit theorem 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 3.5 to the set D={0,1}D=\{0,1\} and Theorem 1.14, one obtains the following:

Corollary 3.7.

Let GG be a countably infinite amenable group with a Følner sequence (FN)N(F_{N})_{N}. 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 subset AGA\subseteq G such that for all kk\in\mathbb{N} and all distinct h1,,hkGh_{1},\dots,h_{k}\in G we have

dF(h11Ahk1A)=Xf(Th1(x))f(Thk(x))𝑑μ.d_{F}\left(h_{1}^{-1}A\cap\dots\cap h_{k}^{-1}A\right)=\int_{X}f(T_{h_{1}}(x))\cdots f(T_{h_{k}}(x))d\mu.\qed

4 Spectral Characterization of vdC sets in abelian groups

In this section we prove a spectral criterion for vdC sets in countable abelian groups (Theorem 1.7), which is a direct generalization of the spectral criterion obtained in [Ru, Theorem 1]; we state it in a fashion similar to [BL, Theorem 8]. Theorem 1.7 implies that the notion of vdC set in d\mathbb{Z}^{d} defined in [BL] is equivalent to our notion of vdC set, even if it is not defined in terms of a Følner sequence (they use a Følner net instead). Also see [FT, Theorem 4.3] for a different proof of Theorem 1.7, shorter than the one included here.

Theorem 1.7.

Let GG be a countable abelian group. A set HGH\subseteq G is vdC 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.

As in [BL, Theorem 1.8] it follows from Theorem 1.7 that, if HH is vdC 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({0})=λ>0\mu(\{0\})=\lambda>0.

    Consider a sequence of finitely supported measures (μN)N(\mu_{N})_{N} which converge weakly to μ\mu; we can suppose μN({1})λ\mu_{N}(\{1\})\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}\to\infty and xN,1,,xN,uNG^x_{N,1},\dots,x_{N,u_{N}}\in\widehat{G}, so that xN,i=1G^x_{N,i}=1\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. (19)

    We will prove that, letting ε<λ\varepsilon<\lambda, Item 5 of Theorem 3.1 is not satisfied. So let AGA\subseteq G and H0HH_{0}\subseteq H be finite and let (FN)N(F_{N})_{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, (20)

    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. (21)

    Equation 21 is a direct consequence of Equation 19 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 20 first note that, as (FN)N(F_{N})_{N} is a Følner sequence, Equation 20 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. (22)

    Now, let δ>0\delta>0 and consider a neighborhood UU of 1G^1\in\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 4.1.

    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.

    4.1 implies Equation 22, because δ\delta is arbitrary and by 4.1 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 4.1 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)|=1|FN||gFNx(gig)gFNx(g)|1|FN|gFNΔgiFN|x(g)|.|x(g_{i})-1|\cdot\left|\frac{1}{|F_{N}|}\sum_{g\in F_{N}}x(g)\right|=\frac{1}{|F_{N}|}\left|\sum_{g\in F_{N}}x(g_{i}g)-\sum_{g\in F_{N}}x(g)\right|\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 4.1.

  2. \impliedby

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

    Lemma 4.2.

    Let (ug)gG(u_{g})_{g\in G} be a sequence of complex numbers in 𝔻\mathbb{D} and let (FN)N(F_{N})_{N} be a Følner sequence such that, for all hGh\in G, 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 [BL, Lemma 1.9], changing the functions gN,hNg_{N},h_{N} from [BL] 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 [CKM, Theorem 2] works for any countable abelian group:

    Lemma 4.3 (Cf. [CKM, 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},(\nu_{n})_{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}). (23)

    The same proof of [CKM, 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 [RuRC, Theorem 3.14]), so they exist for any Borel probability measure in G^\widehat{G}.∎

5 Properties of vdC sets

In [BL] and [Ru], 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.

We also prove 1.8, a finitistic criterion for the notion of vdC set which will be needed to prove 5.4 below.

Remark 5.1.

All the properties below are also satisfied for sets of operatorial recurrence, as defined in [FT] (see [FT, Section 5]), with the proofs of the properties being essentially the same as for vdC sets.

Proposition 5.2 (Ramsey Property of vdC sets, cf. [Ru, Cor. 1]).

Let GG be a countably infinite group and let H,H1,H2GH,H_{1},H_{2}\subseteq G 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 vdC in GG. ∎

We will need the following in order to prove 5.4.

Proposition 5.3.

Let GG be a countably infinite group, and let HG{1}H\subseteq G\setminus\{1\} 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)a;x1=0}{0,1}GA=\{(x_{a})_{a};x_{1}=0\}\subseteq\{0,1\}^{G} and let B=AhHThAB=A\setminus\cup_{h\in H}T_{-h}A. As HH is finite, BB has positive measure, but clearly μ(BThB)=0\mu(B\cap T_{-h}B)=0 for all hHh\in H, so we are done. ∎

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

Proposition 5.4.

Let GG be a countably infinite group and let HG{1}H\subseteq G\setminus\{1\} be a vdC set in GG. Then we can find infinitely many disjoint vdC subsets of HH.

Proof of 5.4.

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 5.2, Hi=1n1HiH\setminus\bigcup_{i=1}^{n-1}H_{i} is a vdC set. Then by 1.8 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 [BL, Corollary 1.15.2] to countable amenable groups.

Theorem 5.5.

Let SS be a subgroup of a countable amenable group GG, let HSH\subseteq S. Then HH is vdC in SS iff it is vdC in GG.

Proof.

If HH is not vdC in GG, then it is not vdC 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 vdC in SS then it is not vdC in GG can be deduced from Item 5 of Theorem 3.1: 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}, (24)

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. (25)

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 3.1 (by taking averages of Equations 24 and 25 for k=1,,Kk=1,\dots,K), so we are done. ∎

The following is proved in [BL, Cor. 1.15.1] for vdC sets in d\mathbb{Z}^{d} using the spectral criterion; using Definition 1.6 instead we prove it for any countable group.

Proposition 5.6.

Let π:GS\pi:G\to S be a group homomorphism, let HH be a vdC set in GG. 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 mps (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 5.7.

In 5.6, 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 5.8 below), but its projection to the first coordinate is a vdC set.

The following generalizes [BL, Corollary 1.16].

Corollary 5.8.

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 vdC 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 (see e.g. [Fa, Lemma 5.2.8] for the case G=G=\mathbb{Z}). The proof of 5.9 is just the proof that any set of differences is a set of recurrence999A slight modification of this proof shows that AA1AA^{-1} is a set of nice recurrence (see Definition 1.9), which is already found in [Fu, Page 74] for the case G=G=\mathbb{Z}.

Proposition 5.9.

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|. (26)

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(Ta(x))f(Tb(x))¯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, we have

N2|Xf𝑑μ|2=|aA0Taf,1|21,1aA0Taf,aA0TafN|B|fL2(X,μ)2+N2λ.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}\leq\left\langle\sum_{a\in A_{0}}T_{a}f,\sum_{a\in A_{0}}T_{a}f\right\rangle\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 vdC in GG if and only if HH1H\cup H^{-1} is vdC 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 5.10.

If GG is a countable group and a subset HGH\subseteq G 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} 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 vdC, so HH is vdC. ∎

Proof of 1.8.

The most natural proof of this fact uses Loeb measures. In order to avoid using non-standard analysis, we will adapt the proof of [Fo, 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 then prove that HH is not vdC 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. (27)

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 Tg(x)=Tn,g(x)T_{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)nLlimnanL:l^{\infty}(\mathbb{N})\to\mathbb{C};(a_{n})_{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)(Sg(y))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 27: for all hHh\in H,

YΦ(f)(Shy)Φ(f)(y)¯𝑑μ(y)=YΦ(fTh)(y)Φ(f)(y)¯𝑑μ(y)=F((fTh)f¯)=LlimnXnf(Tn,h(x))fn(x)¯𝑑μn(x)=0.\int_{Y}\Phi(f)(S_{h}y)\overline{\Phi(f)(y)}d\mu(y)=\int_{Y}\Phi(f\circ T_{h})(y)\overline{\Phi(f)(y)}d\mu(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.

References

  • [Av] J. Avigad. Inverting the Furstenberg correspondence. Discrete and Continuous Dynamical Systems, 32(10), 3421–3431, 2012.
  • [Be2] V. Bergelson. A Density Statement Generalizing Schur’s Theorem. Journal of Combinatorial Theory, Series A 43, pp. 338-343, 1986.
  • [Be5] V. Bergelson. Ergodic Ramsey Theory - An update. Ergodic Theory and d\mathbb{Z}^{d} Actions. London Mathematical Society Lecture Note Series. Cambridge University Press; 1996:1-62.
  • [BF] V. Bergelson, A. Ferré Moragues. An ergodic correspondence principle, invariant means and applications. Israel J. Math. 245 (2021), pp. 921–962.
  • [BF2] V. Bergelson, A. Ferré Moragues. Uniqueness of a Furstenberg system. Proceedings of the American Mathematical Society, 149(7), 2983–2997 (2021).
  • [BL] V. Bergelson, E. Lesigne. Van der Corput sets in d\mathbb{Z}^{d}. Colloq. Math., 110 (1) (2008), pp. 1-49.
  • [Bo] J. Bourgain. Ruzsa’s problem on sets of recurrence. Israel J. Math. 59 (1987), 151–166.
  • [CKM] J. Coquet, T. Kamae, M. Mendès France Sur la mesure spectrale de certaines suites arithmétiques. Bull. Soc. Math. France 105 (1977), 369–384.
  • [DHZ] T. Downarowicz, D. Huczek, G. Zhang. Tilings of amenable groups. J. Reine Angew. Math. 747 277–98, 2019.
  • [Fa] S. Farhangi. Topics in Ergodic Theory and Ramsey Theory. Ph.D thesis, The Ohio State University, 2022.
  • [Fo] A. H. Forrest. Recurrence in Dynamical Systems: A Combinatorial Approach. Ph.D thesis, The Ohio State University, 1990.
  • [FS] A. Fish, S. Skinner. An inverse of Furstenberg’s correspondence principle and applications to nice recurrence. arXiv preprint, arXiv:2407.19444.
  • [FT] S. Farhangi, R. Tucker-Drob. Van der Corput sets in amenable groups and beyond. To be uploaded to arXiv.
  • [Fu] H. Furstenberg. Recurrence in Ergodic Theory and Combinatorial Number Theory. Princeton Univ Press (1981).
  • [KM] T. Kamae, M. Mendès France. Van der Corput’s difference theorem. Israel Journal of Mathematics, Vol. 31, Nos. 3-4, 1978.
  • [KN] L. Kuipers, H. Niederreiter. Uniform Distribution of Sequences. John Wiley (1974).
  • [Lin] E. Lindenstrauss. Pointwise theorems for amenable groups. Electronic Research Announcements of the American Mathematical Society, vol. 5, no. 12, June 1999, pp. 82–90.
  • [Mou] A. Mountakis. Distinguishing sets of strong recurrence from van der Corput Sets . Israel Journal of Mathematics TBD, 2024.
  • [RuFA] W. Rudin. Fourier Analysis on Groups. Interscience, New York, 1962.
  • [RuRC] W. Rudin. Real and Complex Analysis. 3rd edn.(McGraw-Hill Book Co., New York,1987).
  • [Ru] Ruzsa. Connections between the uniform distribution of a sequence and its differences. Topics in classical number theory Vol. 2 pg. 1419-1443 (1984).
  • [We] H. Weyl. Über die Gleichverteilung von Zahlen mod. Eins. Mathematische Annalen 77, 313-352 (1916).