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

A generalization of van der Corput’s Difference Theorem

Sohail Farhangi Adam Mickiewicz University, Poznań, Poland Sohail Farhangi
Department of Mathematics and Informatics
University of Adam Mickiewicz
ul. Wieniawskiego 1
61-712–Poznań
Poland
sohail.farhangi@gmail.com
Abstract.

We prove a generalization of van der Corput’s Difference Theorem in the theory of uniform distribution by establishing a connection with unitary operators that have Lebesgue spectrum. This allows us to show, for example, that if (xn)n=1[0,1](x_{n})_{n=1}^{\infty}\subseteq[0,1] is such that (xn+hxn)n=1(x_{n+h}-x_{n})_{n=1}^{\infty} is uniformly distributed for all hh\in\mathbb{N}, then (xnk)k=1(x_{n_{k}})_{k=1}^{\infty} is uniformly distributed, where (nk)k=1(n_{k})_{k=1}^{\infty} is an enumeration of the 1s1s in the classical Thue-Morse sequence. We also establish a variant of van der Corput’s Difference Theorem that is connected to unitary operators with continuous spectrum. Lastly, we obtain a new characterization of those sequence (xn)n=1[0,1](x_{n})_{n=1}^{\infty}\subseteq[0,1] for which (xn+h,xn)n=1(x_{n+h},x_{n})_{n=1}^{\infty} is uniformly distributed in [0,1]2[0,1]^{2} for all hh\in\mathbb{N}.

Key words and phrases:
van der Corput’s Difference Theorem, discrepancy, ergodic hierarchy of mixing properties,
1991 Mathematics Subject Classification:
37A25, 11K06, 11K38, 37A30
Supported by the Grant “Set theoretic methods in dynamics and number theory”, n. 2019/34/E/ST1/00082.

1. Introduction

In [16] van der Corput proved Theorem 1.1, which is now known as van der Corput’s Difference Theorem (henceforth abbreviated as vdCDT).

Theorem 1.1 ([12, Theorem 1.3.1]).

If (xn)n=1[0,1](x_{n})_{n=1}^{\infty}\subseteq[0,1] is a sequence for which (xn+hxn)n=1(x_{n+h}-x_{n})_{n=1}^{\infty} is uniformly distributed for all hh\in\mathbb{N}, then (xn)n=1(x_{n})_{n=1}^{\infty} is uniformly distributed.

It is clear that there exist uniformly distributed sequences (xn)n=1[0,1](x_{n})_{n=1}^{\infty}\subseteq[0,1] for which (xn+hxn)n=1(x_{n+h}-x_{n})_{n=1}^{\infty} is not uniformly distributed for any hh\in\mathbb{N}. A natural example of such a sequence is found by taking some α\alpha\in\mathbb{R}\setminus\mathbb{Q} and letting xn=In:=nαx_{n}=I_{n}:=\lfloor\sqrt{n}\rfloor\alpha, and it is worth noting that in this example, for any hh\in\mathbb{N} we have xn+hxn=hαx_{n+h}-x_{n}=h\alpha for a full density set of nn\in\mathbb{N}. It is therefore natural to ask what additional properties are satisfied by those (xn)n=1[0,1](x_{n})_{n=1}^{\infty}\subseteq[0,1] that satisfy the hypothesis of vdCDT, and it is the goal of the current paper to answer this question.

For reasons that will become apparant later, let us call a sequence (xn)n=1[0,1](x_{n})_{n=1}^{\infty}\subseteq[0,1] a sL-sequence (spectrally Lebesgue sequence) if (xn+hxn)n=1(x_{n+h}-x_{n})_{n=1}^{\infty} is uniformly distributed for all hh\in\mathbb{N}. It had already been observed in [10] that if (xn)n=1(x_{n})_{n=1}^{\infty} is a sL-sequence, then so is (xn+nα)n=1(x_{n}+n\alpha)_{n=1}^{\infty} for any α\alpha\in\mathbb{R}, and this can be used to show that (xnk)k=1(x_{n_{k}})_{k=1}^{\infty} is uniformly distributed when (nk)k=1(n_{k})_{k=1}^{\infty}\subseteq\mathbb{N} is almost periodic (see [3, Theorem 4.4]), or that (xn,nα)n=1(x_{n},n\alpha)_{n=1}^{\infty} is uniformly distributed in [0,1]2[0,1]^{2} when α\alpha\in\mathbb{R}\setminus\mathbb{Q}. However, we see that these properties are also satisfied by the sequence (In)n=1(I_{n})_{n=1}^{\infty} that we constructed above, so they cannot possibly characterize sL-sequences. One main goal of this paper is to better understand sL-sequences through similar corollaries. In particular, we will give a sufficient condition on those (nk)k=1(n_{k})_{k=1}^{\infty}\subseteq\mathbb{N} for which (xnk)k=1(x_{n_{k}})_{k=1}^{\infty} is uniformly distributed whenever (xn)n=1(x_{n})_{n=1}^{\infty} is a sL-sequence, and we will give a sufficient condition on those (yn)n=1[0,1](y_{n})_{n=1}^{\infty}\subseteq[0,1] for which (xn,yn)n=1(x_{n},y_{n})_{n=1}^{\infty} is uniformly distributed in [0,1]2[0,1]^{2} whenever (xn)n=1(x_{n})_{n=1}^{\infty} is a sL-sequence.

In order to give some context for our main results, let us recall the following Hilbertian analogues of Theorem 1.1 that were introduced by Bergelson in [1] and are of great use in Ergodic Theory.

Theorem 1.2.

Let \mathcal{H} be a Hilbert space and (xn)n=1(x_{n})_{n=1}^{\infty}\subseteq\mathcal{H} a bounded sequence of vectors.

  1. (i)

    If for every hh\in\mathbb{N} we have

    limN1Nn=1Nxn+h,xn=0, then limN1Nn=1Nxn=0.\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}\langle x_{n+h},x_{n}\rangle=0\text{, then }\lim_{N\rightarrow\infty}\left|\left|\frac{1}{N}\sum_{n=1}^{N}x_{n}\right|\right|=0. (1)
  2. (ii)

    If

    limhlim supN|1Nn=1Nxn+h,xn|=0, then limN1Nn=1Nxn=0.\lim_{h\rightarrow\infty}\limsup_{N\rightarrow\infty}\left|\frac{1}{N}\sum_{n=1}^{N}\langle x_{n+h},x_{n}\rangle\right|=0\text{, then }\lim_{N\rightarrow\infty}\left|\left|\frac{1}{N}\sum_{n=1}^{N}x_{n}\right|\right|=0. (2)
  3. (iii)

    If

    limH1Hh=1Hlim supN|1Nn=1Nxn+h,xn|=0, then limN1Nn=1Nxn=0.\lim_{H\rightarrow\infty}\frac{1}{H}\sum_{h=1}^{H}\limsup_{N\rightarrow\infty}\left|\frac{1}{N}\sum_{n=1}^{N}\langle x_{n+h},x_{n}\rangle\right|=0\text{, then }\lim_{N\rightarrow\infty}\left|\left|\frac{1}{N}\sum_{n=1}^{N}x_{n}\right|\right|=0. (3)

It is natural to ask why there are 3 variations of vdCDT when working with Hilbert spaces, but only 1 vdCDT in the Theory of Uniform Distribution. This question is addressed in [5, Chapter 2] by establishing connections between variations of vdCDT with various levels of the ergodic hierarchy of mixing properties of a unitary operator. In particular, [5, Corollary 2.2.11] shows that 1.2(iii) corresponds to weak mixing, [5, Corollary 2.2.15] shows that 1.2(ii) corresponds to strong mixing, and [5, Corollary 2.2.17] and [6] show that 1.2(i) corresponds to Lebesgue spectrum.

Building upon these results, [5, Corollary 2.4.18] is a variation of Theorem 1.1 which resembles Theorem 1.2(iii) and is connected to a new notion in the Theory of Uniform Distribution, wm-sequences, that is related to weak mixing. Similarly, [5, Corollary 2.4.22] is a variation of Theorem 1.1 which resembles Theorem 1.2(ii) and is connected to another new notion in the Theory of Uniform Distribution, sm-sequences, that is related to strong mixing. Interestingly, a natural generalization of Theorem 1.1 was not obtained in [5, Chapter 2.4], so it is one of the (aforementioned) goals of the present paper to prove such a result. Another main goal of this paper is to give new characterizations of wm-sequences. In particular, we identify the classes of c-sequences111We remark the the sequence (In)n=1(I_{n})_{n=1}^{\infty} is a uniformly distributed c-sequence. and compact sequences, which are related to almost periodicity. Then we show that (xn)n=1[0,1](x_{n})_{n=1}^{\infty}\subseteq[0,1] is a wm-sequence if and only if (xn,yn)n=1(x_{n},y_{n})_{n=1}^{\infty} is uniformly distributed in [0,1]2[0,1]^{2} for any uniformly distributed c-sequences (yn)n=1[0,1](y_{n})_{n=1}^{\infty}\subseteq[0,1]. Similarly, we show that (xn)n=1(x_{n})_{n=1}^{\infty} is a wm-sequence if and only if (xnk)k=1(x_{n_{k}})_{k=1}^{\infty} is uniformly distributed for any compact sequence (nk)k=1(n_{k})_{k=1}^{\infty}\subseteq\mathbb{N}.

Lastly, we define a class of mixing sequences (xn)n=1[0,1](x_{n})_{n=1}^{\infty}\subseteq[0,1] called o-sequences, and we show that they are the same as those sequences for which (xn+h,xn)n=1(x_{n+h},x_{n})_{n=1}^{\infty} is uniformly distributed in [0,1]2[0,1]^{2} for all hh\in\mathbb{N}. While the definition of an o-sequence appears to be intimately related to Theorem 1.1, we give two examples of sL-sequences that are not o-sequences.

We must also mention here that variations of Theorem 1.1 and 1.2 corresponding to ergodicity and mild mixing were already established in [5, Chapter 2], and connections between other forms of vdCDT in Hilbert spaces and the ergodic hierarchy of mixing were established in [15].

The structure of the paper is as follows. In the next Section 2 we gather facts discrepancy, spectral theory, and mixing/rigid sequence that we will need later on. In Section 3 we state and prove our main results. In Section 4 we list some examples that give more context to our main results

2. Preliminaries

2.1. Discrepancy

Definition 2.1.

Let dd\in\mathbb{N}. The discrepancy of (xn)n=1N[0,1]d(x_{n})_{n=1}^{N}\subseteq[0,1]^{d} is given by

DN((xn)n=1N):=sup B|1N|{1nN|xnB}|md(B)|,D_{N}\left((x_{n})_{n=1}^{N}\right):=\underset{B\in\mathcal{R}}{\text{sup }}\left|\frac{1}{N}|\{1\leq n\leq N\ |\ x_{n}\in B\}|-m^{d}(B)\right|, (4)

where \mathcal{R} denotes the collection of all rectangular prisms contained in [0,1]d[0,1]^{d}. For an infinite sequence (xn)n=1[0,1]d(x_{n})_{n=1}^{\infty}\subseteq[0,1]^{d}, we let

D¯((xn)n=1):=lim supNDN((xn)n=1N),\overline{D}((x_{n})_{n=1}^{\infty}):=\limsup_{N\rightarrow\infty}D_{N}((x_{n})_{n=1}^{N}), (5)

and we let

D((xn)n=1,(Nq)q=1):=limqDNq((xn)n=1Nq),D((x_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}):=\lim_{q\rightarrow\infty}D_{N_{q}}((x_{n})_{n=1}^{N_{q}}), (6)

provided that the limit exists.

It is worth noting that a sequence (xn)n=1[0,1]d(x_{n})_{n=1}^{\infty}\subseteq[0,1]^{d} is uniformly distributed if and only if D¯((xn)n=1)=0\overline{D}((x_{n})_{n=1}^{\infty})=0 (cf. Theorem 2.1.1 in [12]). We will also be needing the discrepancy of (nα)n=1N(n\alpha)_{n=1}^{N} for some α\alpha, so we record the following result that is a consequence of Theorem 2.3.2 and Example 2.3.1 of [12].

Theorem 2.2.

Let α\alpha\in\mathbb{R} be an algebraic irrational. For any ϵ>0\epsilon>0, there exists a constant C=C(α,ϵ)C=C(\alpha,\epsilon) such that for all NN\in\mathbb{N} we have

DN((xn)n=1N)CN1+ϵ.D_{N}\left((x_{n})_{n=1}^{N}\right)\leq CN^{-1+\epsilon}. (7)

The following is a corollary of the Koksma-Hlawka inequality, and is a special case of [12, Theorem 2.5.6].

Theorem 2.3.

Let dd\in\mathbb{N}, let m:=(m1,,md)d{0}\vec{m}:=(m_{1},\cdots,m_{d})\in\mathbb{Z}^{d}\setminus\{0\}, and let M=max1id(|mi|)M=\max_{1\leq i\leq d}(|m_{i}|). For all (xn)n=1N[0,1]d(x_{n})_{n=1}^{N}\in[0,1]^{d} we have

|1Nn=1Ne(mxn)|(4πM)dDN((xn)n=1N)\left|\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot x_{n})\right|\leq(4\pi M)^{d}D_{N}\left((x_{n})_{n=1}^{N}\right) (8)

2.2. Spectral theory

We give a brief review of spectral theory in order to give the reader context for the definitions and results in later sections. For a more detailed treatment we refer the reader to [7, Chapter 18].

Suppose that \mathcal{H} is a Hilbert space, U:U:\mathcal{H}\rightarrow\mathcal{H} is a unitary operator, and ξ\xi\in\mathcal{H} is a cyclic vector for UU. The Spectral Theorem tells us that there is a Hilbert space isomorphism ϕ:HL2([0,1],νξ)\phi:H\rightarrow L^{2}([0,1],\nu_{\xi}) for which ϕU=Meϕ\phi U=M_{e}\phi, where Me:L2([0,1],νξ)L2([0,1],νξ)M_{e}:L^{2}([0,1],\nu_{\xi})\rightarrow L^{2}([0,1],\nu_{\xi}) is the unitary operator given by (Mef)(x)=e(x)f(x)(M_{e}f)(x)=e(-x)f(x), and ϕ(ξ)=1\phi(\xi)=1. The measure νξ\nu_{\xi} is the spectral measure of ξ\xi (with respect to UU) and is seen to satisfy

ν^ξ(n)=01e(nx)𝑑νξ(x)=Men1,1L2=Unξ,ξ.\hat{\nu}_{\xi}(n)=\int_{0}^{1}e(-nx)d\nu_{\xi}(x)=\langle M_{e}^{n}1,1\rangle_{L^{2}}=\langle U^{n}\xi,\xi\rangle_{\mathcal{H}}. (9)

Furthermore, ξ1,ξ2\xi_{1},\xi_{2}\in\mathcal{H} are such that νξ1νξ2\nu_{\xi_{1}}\perp\nu_{\xi_{2}}, then ξ1,ξ2=0\langle\xi_{1},\xi_{2}\rangle_{\mathcal{H}}=0. Some classical results in ergodic theory can be understood as a corollary of this fact.

For example, if ξ1,ξ2\xi_{1},\xi_{2}\in\mathcal{H} are such that Uξ1=ξ1U\xi_{1}=\xi_{1} and ξ2c({Uηη|η})\xi_{2}\in c\ell(\{U\eta-\eta\ |\ \eta\in\mathcal{H}\}), then νξ1=ξ1𝟙{0}\nu_{\xi_{1}}=||\xi_{1}||\mathbbm{1}_{\{0\}} and νξ2({0})=0\nu_{\xi_{2}}(\{0\})=0, hence νξ1νξ2\nu_{\xi_{1}}\perp\nu_{\xi_{2}}, and we recover the the decomposition of Von Neumann of \mathcal{H} into the invariant space and the span closure of cocycles. Another example is when ξ1\xi_{1} is in the span closure of eigenvectors of UU and ξ2\xi_{2} is a weakly mixing vector for UU, in which case νξ1\nu_{\xi_{1}} is a discrete measure and νξ2\nu_{\xi_{2}} is a continuous measure. We again see that νξ1νξ2\nu_{\xi_{1}}\perp\nu_{\xi_{2}}, and we recover the Jacobs-de Leeuw-Glicksberg decomposition for a unitary operator. Another example that we will be using later on is when νξ1m\nu_{\xi_{1}}\perp m and νξ2<<m\nu_{\xi_{2}}<<m, in which case we immediately see that νξ1νξ2\nu_{\xi_{1}}\perp\nu_{\xi_{2}}, hence ξ1,ξ2=0\langle\xi_{1},\xi_{2}\rangle_{\mathcal{H}}=0.

2.3. Mixing and rigidity of sequences

2.3.1. Definitions

We begin with the notion of permissible triples and permissible pairs, which are notions used to study bounded sequences by considering all possible subsequences in which relevant limits exists. Permissible triples were used in [5, Chapter 2] and [6] to construct Hilbert spaces associated to a given pair of sequences, and we record the results of this construction as Lemma 2.9. The usage of permissible triples is similar to the construction of Furstenberg Systems discussed in [8] and the references therein. Lemma 4.2 ([5, Lemma 2.2.20]) will indicate that it is indeed necessary to consider permissible pairs (triples) instead of some simpler notion.

Definition 2.4.

Let (xn)n=1(x_{n})_{n=1}^{\infty} and (yn)n=1(y_{n})_{n=1}^{\infty} be bounded sequences of complex numbers and let (Nq)q=1(N_{q})_{q=1}^{\infty}\subseteq\mathbb{N} be an increasing sequence. The triple ((xn)n=1,((x_{n})_{n=1}^{\infty}, (yn)n=1,(Nq)q=1)(y_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) is a permissible triple if

limq1Nqn=1Nqcn+hdn¯\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}c_{n+h}\overline{d_{n}} (10)

exists for all hh\in\mathbb{N} and (cn)n=1,(dn)n=1{(xn)n=1,(c_{n})_{n=1}^{\infty},(d_{n})_{n=1}^{\infty}\in\{(x_{n})_{n=1}^{\infty}, (yn)n=1}(y_{n})_{n=1}^{\infty}\}. Given a family {(xn,h)n=1}h=1\{(x_{n,h})_{n=1}^{\infty}\}_{h=1}^{\infty} of sequences in [0,1]d[0,1]^{d} and an increasing sequence (Nq)q=1(N_{q})_{q=1}^{\infty}\subseteq\mathbb{N}, we define ({(xn,h)n=1}h=1,(Nq)q=1)(\{(x_{n,h})_{n=1}^{\infty}\}_{h=1}^{\infty},(N_{q})_{q=1}^{\infty}) to be a permissible pair if for all hh\in\mathbb{N} D((xn,h)n=1,(Nq)q=1)D((x_{n,h})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) is well defined.

Definition 2.5.

Let (xn)n=1(x_{n})_{n=1}^{\infty} be a bounded sequence of complex numbers.

  1. (i)

    If for any permissible triple ((xn)n=1,((x_{n})_{n=1}^{\infty}, (yn)n=1,(Nq)q=1)(y_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) we have

    limH1Hh=1Hlimq|1Nqn=1Nqxn+hyn¯|=0,\lim_{H\rightarrow\infty}\frac{1}{H}\sum_{h=1}^{H}\lim_{q\rightarrow\infty}\left|\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}x_{n+h}\overline{y_{n}}\right|=0, (11)

    then (xn)n=1(x_{n})_{n=1}^{\infty} is a nearly weakly mixing sequence.

  2. (ii)

    If for any permissible triple ((xn)n=1,(xn)n=1,(Nq)q=1)((x_{n})_{n=1}^{\infty},(x_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) there is a function fL1([0,1],m)f\in L^{1}([0,1],m) for which

    limq|1Nqn=1Nqxn+hxn¯|=01f(x)e2πihx𝑑x,\lim_{q\rightarrow\infty}\left|\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}x_{n+h}\overline{x_{n}}\right|=\int_{0}^{1}f(x)e^{2\pi ihx}dx, (12)

    then (xn)n=1(x_{n})_{n=1}^{\infty} is a spectrally Lebesgue sequence.

  3. (iii)

    If for any permissible triple ((xn)n=1,(xn)n=1,(Nq)q=1)((x_{n})_{n=1}^{\infty},(x_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) we have

    limq|1Nqn=1Nqxn+hxn¯|=0,\lim_{q\rightarrow\infty}\left|\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}x_{n+h}\overline{x_{n}}\right|=0, (13)

    then (xn)n=1(x_{n})_{n=1}^{\infty} is a nearly orthogonal sequence.

  4. (iv)

    If for any permissible triple of the form ((xn)n=1,(xn)n=1,(Nq)q=1)((x_{n})_{n=1}^{\infty},(x_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) we have

    supmmin 1kKlimq1Nqn=1Nq|cn+mcn+k|2<ϵ.\underset{m\in\mathbb{N}}{\sup}\ \underset{1\leq k\leq K}{\text{min }}\ \lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}|c_{n+m}-c_{n+k}|^{2}<\epsilon. (14)

    then (xn)n=1(x_{n})_{n=1}^{\infty} is a compact sequence222This definition was motivated by Definition 3.13 in [14]..

  5. (v)

    If for any permissible triple of the form ((xn)n=1,(xn)n=1,(Nq)q=1)((x_{n})_{n=1}^{\infty},(x_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) we have

    limq|1Nqn=1Nqxn+hxn¯|=01e2πihx𝑑μ(x),\lim_{q\rightarrow\infty}\left|\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}x_{n+h}\overline{x_{n}}\right|=\int_{0}^{1}e^{2\pi ihx}d\mu(x), (15)

    where μm\mu\perp m is a positive measure, then (xn)n=1(x_{n})_{n=1}^{\infty} is a spectrally singular sequence.

It is worth observing that if (xn)n=1(x_{n})_{n=1}^{\infty} satisfies any of items (i)(v)(i)-(v) in Definition 2.5, then (x¯n)n=1(\overline{x}_{n})_{n=1}^{\infty} satisfies the same list of items.

Definition 2.6.

Let d,𝒞d={fC([0,1]d)|[0,1]df𝑑md=0}d\in\mathbb{N},\mathcal{C}_{d}=\{f\in C([0,1]^{d})\ |\ \int_{[0,1]^{d}}fdm^{d}=0\}, and (xn)n=1[0,1]d(x_{n})_{n=1}^{\infty}\subseteq[0,1]^{d}.

  1. (i)

    If for every f𝒞df\in\mathcal{C}_{d}, (f(xn)n=1)(f(x_{n})_{n=1}^{\infty}) is a nearly weakly mixing sequence, then (xn)n=1(x_{n})_{n=1}^{\infty} is a wm-sequence.

  2. (ii)

    If for every f𝒞df\in\mathcal{C}_{d}, (f(xn)n=1)(f(x_{n})_{n=1}^{\infty}) is a spectrally Lebesgue sequence, then (xn)n=1(x_{n})_{n=1}^{\infty} is a sL-sequence.

  3. (iii)

    If for every f𝒞df\in\mathcal{C}_{d}, (f(xn)n=1)(f(x_{n})_{n=1}^{\infty}) is a nearly orthogonal sequence, then (xn)n=1(x_{n})_{n=1}^{\infty} is a o-sequence.

  4. (iv)

    If for every fC([0,1]d)f\in C([0,1]^{d}), (f(xn)n=1)(f(x_{n})_{n=1}^{\infty}) is a compact sequence, then (xn)n=1(x_{n})_{n=1}^{\infty} is a c-sequence.

  5. (v)

    If for every fC([0,1]d)f\in C([0,1]^{d}), (f(xn)n=1)(f(x_{n})_{n=1}^{\infty}) is a spectrally singular sequence, then (xn)n=1(x_{n})_{n=1}^{\infty} is a ss-sequence.

Definition 2.7.

For a sequence of natural numbers A=(nk)k=1A=(n_{k})_{k=1}^{\infty} let

d¯(A):=lim infN1N|{1nN|nA}|=lim infN1Nn=1N𝟙A(n), and\underline{d}(A):=\liminf_{N\rightarrow\infty}\frac{1}{N}\left|\{1\leq n\leq N\ |\ n\in A\}\right|=\liminf_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}\mathbbm{1}_{A}(n)\text{, and} (16)
d¯(A):=lim supN1N|{1nN|nA}|=lim supN1Nn=1N𝟙A(n).\overline{d}(A):=\limsup_{N\rightarrow\infty}\frac{1}{N}\left|\{1\leq n\leq N\ |\ n\in A\}\right|=\limsup_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}\mathbbm{1}_{A}(n). (17)

d¯(A)\underline{d}(A) is the natural lower density of A and d¯(A)\overline{d}(A) is the natural upper density of A. If d¯(A)=d¯(A)\overline{d}(A)=\underline{d}(A), then we let d(A)d(A) denote the common value which is the natural density of A.

Definition 2.8.

Let A=(nk)k=1A=(n_{k})_{k=1}^{\infty}\subseteq\mathbb{N} be a strictly increasing sequence satisfying d¯(A)>0\underline{d}(A)>0. If (𝟙A(n))n=1(\mathbbm{1}_{A}(n))_{n=1}^{\infty} is a compact sequence as a sequence of bounded complex numbers, then AA is compact as a sequence of natural numbers. Similarly, if (𝟙A(n))n=1(\mathbbm{1}_{A}(n))_{n=1}^{\infty} is a spectrally singular as a sequence of bounded complex numbers, then AA is spectrally singular as a sequence of natural numbers.

2.3.2. A potpourri of lemmas

Our first lemma follows from [6, Theorem 2.1] and the preceding discussion.

Lemma 2.9.

Let ((xn)n=1,((x_{n})_{n=1}^{\infty}, (yn)n=1,(Nq)q=1)(y_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) be a permissible triple. Then there exists a Hilbert space \mathcal{H} and an injective map ı\i from \mathcal{H} to the space of sequence of complex numbers satisfying the following properties:

  1. (i)

    For any ξ1,ξ2\xi_{1},\xi_{2}\in\mathcal{H} and (zi,n)n=1:=ı(ξi)(z_{i,n})_{n=1}^{\infty}:=\i (\xi_{i}), we have

    ξ1,ξ2=limq1Nqn=1Nqz1,nz2,n¯.\langle\xi_{1},\xi_{2}\rangle=\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}z_{1,n}\overline{z_{2,n}}. (18)
  2. (ii)

    The map ı\i is linear in the sense that for any cc\in\mathbb{C} and (z3,n)n=1:=ı(ξ1+cξ2)(z_{3,n})_{n=1}^{\infty}:=\i (\xi_{1}+c\xi_{2}), we have

    limq1Nqn=1Nq|z3,nz1,ncz2,n|2=0.\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}|z_{3,n}-z_{1,n}-cz_{2,n}|^{2}=0. (19)
  3. (iii)

    There exists a unitary operator S:S:\mathcal{H}\rightarrow\mathcal{H} for which

    Shξ1,ξ2=limq1Nqn=1Nqz1,n+hz2,n¯.\langle S^{h}\xi_{1},\xi_{2}\rangle=\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}z_{1,n+h}\overline{z_{2,n}}. (20)
  4. (iv)

    \mathcal{H} is the smallest SS-invariant Hilbert space containing the vectors x\vec{x} and y\vec{y} satisfying ı(x)=(xn)n=1\i (\vec{x})=(x_{n})_{n=1}^{\infty} and ı(y)=(yn)n=1\i (\vec{y})=(y_{n})_{n=1}^{\infty}.

Remark 1.

Lemma 2.9 not only shows why we like to work with permissible triples, but it also justifies the terminology used in Definition 2.5. To illustrate the latter claim with an example, let (xn)n=1(x_{n})_{n=1}^{\infty} be a spectrally singular sequence. We see that if ((xn)n=1,(xn)n=1,(Nq)q=1)((x_{n})_{n=1}^{\infty},(x_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) is a permissible triple and x\vec{x} is as in Lemma 2.9(iv), then

Shx,x=limq1Nqn=1Nqxn+hxn¯,\langle S^{h}\vec{x},\vec{x}\rangle=\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}x_{n+h}\overline{x_{n}}, (21)

hence the spectral measure of x\vec{x} with respect to SS is singular with respect to mm.

Our next lemma is a simple extension of [14, Lemma 3.26].

Lemma 2.10.

Let (cn)n=1(c_{n})_{n=1}^{\infty}\subseteq\mathbb{C} be bounded. There exists a compact metric space YY, a continuous map S:XXS:X\rightarrow X, a continuous function F:XF:X\rightarrow\mathbb{C}, and a point xXx\in X with a dense orbit under SS such that cn=F(Snx)c_{n}=F(S^{n}x) for all nn\in\mathbb{N}. Furthermore, if (cn)n=1(c_{n})_{n=1}^{\infty} is compact, then (X,S)(X,S) is such that for any weak limit μ\mu of {1Nn=1NδSnx}N=1\{\frac{1}{N}\sum_{n=1}^{N}\delta_{S^{n}x}\}_{N=1}^{\infty}, the measure preserving system (X,S,μ)(X,S,\mu) has discrete spectrum.

Proof.

We may assume without loss of generality that |cn|1|c_{n}|\leq 1 for all nn\in\mathbb{N}. Let BB\subseteq\mathbb{C} denote the closed unit ball and let X=BX^{\prime}=B^{\mathbb{N}} endowed with the product topology. Since BB is a compact metric space we see that XX^{\prime} is also a compact metric space. Let S:XXS:X^{\prime}\rightarrow X^{\prime} denote the left shift. Let F:X(C)F:X^{\prime}\rightarrow\mathbb{(}C) denote the projection onto the first coordinate, which is seen to be continuous. Let x=(cn)n=1Xx=(c_{n})_{n=1}^{\infty}\in X^{\prime}, and let X=c({Snx}n=1)X=c\ell(\{S^{n}x\}_{n=1}^{\infty}). Since XX is a closed subset of XX^{\prime} we see that XX is also a compact metric space, and it is clear that xx has a dense orbit in XX by construction. Since F(Snx)=cnF(S^{n}x)=c_{n}, the first part of the lemma is proven. To see the latter claim, let (Nq)q=1(N_{q})_{q=1}^{\infty} be such that

limq1Nqn=1NqδSnx=μ,\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}\delta_{S^{n}x}=\mu, (22)

with convergence taking place in the weak topology. We see that for any hh\in\mathbb{N}, we have

limq1Nqn=1Nqcn+hcn¯=XShFF𝑑μ,\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}c_{n+h}\overline{c_{n}}=\int_{X}S^{h}F\cdot Fd\mu, (23)

so ((cn)n=1,(cn)n=1,(Nq)q=1)((c_{n})_{n=1}^{\infty},(c_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) is a permissible triple. It follows that for any ϵ>0\epsilon>0, there exists KK\in\mathbb{N} such that

supmmin1kKSmFSkF2=supmmin1kKX|SmFSkF|2𝑑μ\displaystyle\sup_{m\in\mathbb{N}}\min_{1\leq k\leq K}||S^{m}F-S^{k}F||_{2}=\sup_{m\in\mathbb{N}}\min_{1\leq k\leq K}\int_{X}\left|S^{m}F-S^{k}F\right|^{2}d\mu
=\displaystyle= supmmin 1kKlimq1Nqn=1Nq|F(Sn+mx)F(Sn+kx)|2\displaystyle\underset{m\in\mathbb{N}}{\sup}\ \underset{1\leq k\leq K}{\text{min }}\ \lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}|F(S^{n+m}x)-F(S^{n+k}x)|^{2}
=\displaystyle= supmmin 1kKlimq1Nqn=1Nq|cn+mcn+k|2<ϵ,\displaystyle\underset{m\in\mathbb{N}}{\sup}\ \underset{1\leq k\leq K}{\text{min }}\ \lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}|c_{n+m}-c_{n+k}|^{2}<\epsilon,

so FF is a compact vector. Since the set of bounded compact vectors is a norm-closed SS invariant algebra, we see that C(X)C(X) consists of compact vectors, hence all of L2(X,μ)L^{2}(X,\mu) consists of compact vectors, which yields the desired result. ∎

Lemma 2.11.

Let (xn)n=1(x_{n})_{n=1}^{\infty} be a bounded sequence of complex numbers.

  1. (i)

    If

    h=1lim supN|1Nn=1Nxn+hxn¯|2<,\sum_{h=1}^{\infty}\limsup_{N\rightarrow\infty}\left|\frac{1}{N}\sum_{n=1}^{N}x_{n+h}\overline{x_{n}}\right|^{2}<\infty, (24)

    then (xn)n=1(x_{n})_{n=1}^{\infty} is spectrally Lebesgue.

  2. (ii)

    If

    limHh=1Hlim supN|1Nn=1Nxn+hxn¯|=0,\lim_{H\rightarrow\infty}\sum_{h=1}^{H}\limsup_{N\rightarrow\infty}\left|\frac{1}{N}\sum_{n=1}^{N}x_{n+h}\overline{x_{n}}\right|=0, (25)

    then (xn)n=1(x_{n})_{n=1}^{\infty} is a nearly weakly mixing sequence.

Proof of (i).

Item (i) is a special case of [6, Theorem 2.7] and item (ii) is a special case of [5, Corollary 2.2.11]. ∎

Lemma 2.12.

Let 𝒞\mathcal{C} denote one of the following classes of bounded sequences of complex numbers: nearly weakly mixing, spectrally Lebesgue, compact, spectrally singular.

  1. (i)

    If (xn)n=1,(x_{n})_{n=1}^{\infty}, (yn)n=1𝒞(y_{n})_{n=1}^{\infty}\in\mathcal{C} and cc\in\mathbb{C}, then for zn=xn+cynz_{n}=x_{n}+cy_{n} we have that (zn)n=1𝒞(z_{n})_{n=1}^{\infty}\in\mathcal{C}.

  2. (ii)

    If {(zn,k)n=1}k=1𝒞\{(z_{n,k})_{n=1}^{\infty}\}_{k=1}^{\infty}\subseteq\mathcal{C} and there exists a bounded sequence (zn)n=1(z_{n})_{n=1}^{\infty} satisfying

    limksupn|zn,kzn|=0,\lim_{k\rightarrow\infty}\sup_{n\in\mathbb{N}}|z_{n,k}-z_{n}|=0, (26)

    then (zn)n=1𝒞(z_{n})_{n=1}^{\infty}\in\mathcal{C}.

Proof.

We first prove (i). Let ((zn)n=1,(zn)n=1,(Nq)q=1)((z_{n})_{n=1}^{\infty},(z_{n})_{n=1}^{\infty},(N_{q}^{\prime})_{q=1}^{\infty}) be a permissible triple, let (Nq)q=1(N_{q})_{q=1}^{\infty} be a subsequence of (Nq)q=1(N_{q}^{\prime})_{q=1}^{\infty} for which ((xn)n=1,((x_{n})_{n=1}^{\infty}, (yn)n=1,(y_{n})_{n=1}^{\infty}, (Nq)q=1)(N_{q})_{q=1}^{\infty}) is a permissible triple, and let ,x,y,\mathcal{H},\vec{x},\vec{y}, and SS be as in Lemma 2.9. We see that the spectral measures of x\vec{x} and y\vec{y} are continuous if CC denotes nearly weakly mixing sequences, absolutely continuous to Lebesgue if 𝒞\mathcal{C} denotes spectrally Lebesgue sequences, discrete if 𝒞\mathcal{C} denotes compact sequences, and mutually singular with Lebesgue if 𝒞\mathcal{C} denotes spectrally singular sequences. Consequently, the spectral measure of x+cy\vec{x}+c\vec{y} is of the same class as that of x\vec{x} (and y\vec{y}).

We see that (ii) is immediate if 𝒞\mathcal{C} denotes either the class of nearly weakly mixing sequences, or the class of compact sequences. Now we prove (ii) by way of contradiction when 𝒞\mathcal{C} denotes the class of spectrally Lebesgue sequences. Since (zn)n=1(z_{n})_{n=1}^{\infty} is not spectrally Lebesgue, let ((zn)n=1,(zn)n=1,(Nq)q=1)((z_{n})_{n=1}^{\infty},(z_{n})_{n=1}^{\infty},(N_{q}^{\prime})_{q=1}^{\infty}) be a permissible triple for which

γ(h):=limq1Nqn=1Nqzn+hzn¯\gamma(h):=\lim_{q\rightarrow\infty}\frac{1}{N_{q}^{\prime}}\sum_{n=1}^{N_{q}^{\prime}}z_{n+h}\overline{z_{n}} (27)

form the Fourier coefficients of some measure μ\mu that is not absolutely continuous with respect to mm. Let us apply Lemma 2.9 to the permissible triple ((zn)n=1,((z_{n})_{n=1}^{\infty}, (zn)n=1,(z_{n})_{n=1}^{\infty}, (Nq)q=1)(N_{q}^{\prime})_{q=1}^{\infty}) to obtain \mathcal{H}^{\prime}, z\vec{z}, and SS^{\prime}. The spectral theorem lets us write z=zL+zs\vec{z}=\vec{z}_{L}+\vec{z}_{s} where νzL<<m\nu_{\vec{z}_{L}}<<m and 0νzsm0\neq\nu_{\vec{z}_{s}}\perp m. Let 0<ϵ<zs0<\epsilon<||\vec{z}_{s}|| be arbitrary, let kk\in\mathbb{N} be such that supn|zn,kzn|<ϵ\sup_{n\in\mathbb{N}}|z_{n,k}-z_{n}|<\epsilon, and let (Nq)q=1(N_{q})_{q=1}^{\infty} be a subsequence of (Nq)q=1(N_{q}^{\prime})_{q=1}^{\infty} for which ((zn)n=1,(zn,k)n=1,(Nq)q=1)((z_{n})_{n=1}^{\infty},(z_{n,k})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) is a permissible triple. Let ,z,zk,\mathcal{H},\vec{z},\vec{z}_{k}, and SS be given by Lemma 2.9. We see that \mathcal{H}^{\prime} is naturally identified with a subspace of \mathcal{H} and SS^{\prime} is the restriction of SS to \mathcal{H}^{\prime}, which is justifies our abuse of notation for the term z\vec{z} in both situations. Since νzk<<m\nu_{\vec{z}_{k}}<<m, we see that

ϵ<zszzk<ϵ,\epsilon<||\vec{z}_{s}||\leq||\vec{z}-\vec{z}_{k}||<\epsilon, (28)

which yields the desired contradiction. The proof in the case that 𝒞\mathcal{C} denotes the class of spectrally singular sequences is similar. ∎

Corollary 2.13.

Let dd\in\mathbb{N} and (xn)n=1[0,1]d(x_{n})_{n=1}^{\infty}\subseteq[0,1]^{d}.

  1. (i)

    If for every md{0}\vec{m}\in\mathbb{Z}^{d}\setminus\{0\} the sequence (e(mxn))n=1(e(\vec{m}\cdot x_{n}))_{n=1}^{\infty} is nearly weakly mixing, then (xn)n=1(x_{n})_{n=1}^{\infty} is a wm-sequence.

  2. (ii)

    If for every md{0}\vec{m}\in\mathbb{Z}^{d}\setminus\{0\} the sequence (e(mxn))n=1(e(\vec{m}\cdot x_{n}))_{n=1}^{\infty} is spectrally Lebesgue, then (xn)n=1(x_{n})_{n=1}^{\infty} is a sL-sequence.

  3. (iii)

    If for every md{0}\vec{m}\in\mathbb{Z}^{d}\setminus\{0\} the sequence (e(mxn))n=1(e(\vec{m}\cdot x_{n}))_{n=1}^{\infty} is compact, then (xn)n=1(x_{n})_{n=1}^{\infty} is a c-sequence.

  4. (iv)

    If for every md{0}\vec{m}\in\mathbb{Z}^{d}\setminus\{0\} the sequence (e(mxn))n=1(e(\vec{m}\cdot x_{n}))_{n=1}^{\infty} is spectrally singular, then (xn)n=1(x_{n})_{n=1}^{\infty} is a ss-sequence.

Lemma 2.14.

Let (xn)n=1(x_{n})_{n=1}^{\infty} and (yn)n=1(y_{n})_{n=1}^{\infty} be bounded sequences of complex numbers.

  1. (i)

    If (xn)n=1(x_{n})_{n=1}^{\infty} is spectrally Lebesgue and (yn)n=1(y_{n})_{n=1}^{\infty} is spectrally singular, then

    limN1Nn=1Nxnyn=0.\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}x_{n}y_{n}=0. (29)
  2. (ii)

    If (xn)n=1(x_{n})_{n=1}^{\infty} is nearly weakly mixing and (yn)n=1(y_{n})_{n=1}^{\infty} is compact, then

    limN1Nn=1Nxnyn=0.\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}x_{n}y_{n}=0. (30)
Proof.

Let (Nq)q=1(N_{q}^{\prime})_{q=1}^{\infty}\subseteq\mathbb{N} be any sequence for which

I:=limq1Nqn=1Nqxnyn=0I:=\lim_{q\rightarrow\infty}\frac{1}{N_{q}^{\prime}}\sum_{n=1}^{N_{q}^{\prime}}x_{n}y_{n}=0 (31)

exists. Let (Nq)q=1(N_{q})_{q=1}^{\infty} be any subsequence of (Nq)q=1(N_{q}^{\prime})_{q=1}^{\infty} for which ((xn)n=1,((x_{n})_{n=1}^{\infty}, (yn)n=1,(y_{n})_{n=1}^{\infty}, (Nq)q=1)(N_{q})_{q=1}^{\infty}) is a permissible triple. Let ,x,y\mathcal{H},\vec{x},\vec{y}, and SS be as in Lemma 2.9. As mentioned in Section 2.2, we see that νxνy\nu_{\vec{x}}\perp\nu_{\vec{y}}, so I=x,y=0I=\langle\vec{x},\vec{y}\rangle=0. ∎

Lemma 2.15.

Let (cn)n=1(c_{n})_{n=1}^{\infty} be a bounded and compact sequence of complex numbers.

  1. (i)

    For all \ell\in\mathbb{N}, (cn)n=1(c_{n}^{\ell})_{n=1}^{\infty} is a compact sequence.

  2. (ii)

    If (cn)n=1(c_{n})_{n=1}^{\infty} takes nonnegative values, then (cn)n=1(\sqrt{c_{n}})_{n=1}^{\infty} is a compact sequence.

  3. (iii)

    The sequence (cn)n=1(c_{n}^{\prime})_{n=1}^{\infty} given by cncn¯=|cn|c_{n}^{\prime}\overline{c_{n}}=|c_{n}| is a compact sequence.

Proof.

In all cases we may assume without loss of generality that |cn|1|c_{n}|\leq 1 for all nn. To prove (i), it suffices to observe that |ab|=|ab||a1+a2b++ab2+b1|max(|a|,|b|)1|ab||a^{\ell}-b^{\ell}|=|a-b|\cdot|a^{\ell-1}+a^{\ell-2}b+\cdots+ab^{\ell-2}+b^{\ell-1}|\leq\ell\max(|a|,|b|)^{\ell-1}|a-b|, so

limq1Nqn=1Nq|cn+mcn+k|22limq1Nqn=1Nq|cn+mcn+k|2.\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}|c_{n+m}^{\ell}-c_{n+k}^{\ell}|^{2}\leq\ell^{2}\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}|c_{n+m}-c_{n+k}|^{2}. (32)

To prove (ii), let ϵ>0\epsilon>0 be arbitrary, let ((cn)n=1,(cn)n=1,(Nq)q=1)((c_{n})_{n=1}^{\infty},(c_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) be a permissible trip be a permissible triple, and let KK\in\mathbb{N} be such that

supmmin1kKlimq1Nqn=1Nq|cn+mcn+k|2<ϵ2.\sup_{m\in\mathbb{N}}\min_{1\leq k\leq K}\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}|c_{n+m}-c_{n+k}|^{2}<\epsilon^{2}. (33)

For mm\in\mathbb{N}, let k=k(m)[1,K]k=k(m)\in[1,K] attain the minimum, and let A={n||cn+m+cn+k|ϵ}A=\{n\in\mathbb{N}\ |\ |\sqrt{c_{n+m}}+\sqrt{c_{n+k}}|\geq\sqrt{\epsilon}\}. We see that

limq1Nqn=1Nq|cn+mcn+k|2\displaystyle\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}|\sqrt{c_{n+m}}-\sqrt{c_{n+k}}|^{2}
\displaystyle\leq limq1Nq(n[1,Nq]A|cn+mcn+k|2+n[1,Nq]Ac|cn+mcn+k|2)\displaystyle\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\left(\sum_{n\in[1,N_{q}]\cap A}|\sqrt{c_{n+m}}-\sqrt{c_{n+k}}|^{2}+\sum_{n\in[1,N_{q}]\cap A^{c}}|\sqrt{c_{n+m}}-\sqrt{c_{n+k}}|^{2}\right)
\displaystyle\leq limq1Nq(n[1,Nq]A|cn+mcn+kcn+m+cn+k|2+n[1,Nq]Ac|cn+m+cn+k|2)\displaystyle\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\left(\sum_{n\in[1,N_{q}]\cap A}\left|\frac{c_{n+m}-c_{n+k}}{\sqrt{c_{n+m}}+\sqrt{c_{n+k}}}\right|^{2}+\sum_{n\in[1,N_{q}]\cap A^{c}}|\sqrt{c_{n+m}}+\sqrt{c_{n+k}}|^{2}\right)
\displaystyle\leq 1ϵlimq1Nqn=1Nq|cn+mcn+k|2+ϵ2ϵ.\displaystyle\frac{1}{\epsilon}\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}|c_{n+m}-c_{n+k}|^{2}+\epsilon\leq 2\epsilon.

To prove (iii), we recall that (cn¯)n=1(\overline{c_{n}})_{n=1}^{\infty} is a compact sequence, so (|cn|2=cncn¯)n=1(|c_{n}|^{2}=c_{n}\overline{c_{n}})_{n=1}^{\infty} is a compact sequence by [5, Lemma 2.3.8(ii)], so the desrired result now follows from Lemma 2.15(ii). ∎

3. Main Results

Theorem 3.1.

Let d,d1d,d_{1}\in\mathbb{N} and let (xn)n=1[0,1]d(x_{n})_{n=1}^{\infty}\subseteq[0,1]^{d}.

  1. (i)

    If (xn)n=1(x_{n})_{n=1}^{\infty} is a sL-sequence (wm-sequence) and (nk)k=1(n_{k})_{k=1}^{\infty}\subseteq\mathbb{N} is a spectrally singular (compact) sequence, then (xnk)k=1(x_{n_{k}})_{k=1}^{\infty} is uniformly distributed.

  2. (ii)

    If for any compact sequence (nk)k=1(n_{k})_{k=1}^{\infty}\subseteq\mathbb{N}, the sequence (xnk)k=1(x_{n_{k}})_{k=1}^{\infty} is uniformly distributed, then (xn)n=1(x_{n})_{n=1}^{\infty} is a wm-sequence.

  3. (iii)

    If (xn)n=1(x_{n})_{n=1}^{\infty} is a sL-sequence (wm-sequence) and (yn)n=1[0,1]d1(y_{n})_{n=1}^{\infty}\subseteq[0,1]^{d_{1}} is a uniformly distributed ss-sequence (c-sequence), then (xn,yn)n=1(x_{n},y_{n})_{n=1}^{\infty} is uniformly distributed.

  4. (iv)

    If (xn+yn)n=1(x_{n}+y_{n})_{n=1}^{\infty} is uniformly distributed for all c-sequences (yn)n=1[0,1]d(y_{n})_{n=1}^{\infty}\subseteq[0,1]^{d}, then (xn)n=1(x_{n})_{n=1}^{\infty} is a wm-sequence.

Proof of (i).

Let md{0}\vec{m}\in\mathbb{Z}^{d}\setminus\{\vec{0}\} be arbitrary, and note that (e2πimxn)n=1(e^{2\pi i\vec{m}\cdot x_{n}})_{n=1}^{\infty} is a spectrally Lebesgue (nearly weakly mixing) sequence . Letting B=(nk)k=1B=(n_{k})_{k=1}^{\infty}, we see that (𝟙B(n))n=1(\mathbbm{1}_{B}(n))_{n=1}^{\infty} is a spectrally singular sequence (compact sequence), so by Lemma 2.14 we see that

0\displaystyle 0 =limN1N|n=1Ne(mxn)𝟙B(n)|=limN1N|nk[1,N]e(mxnk)|\displaystyle=\lim_{N\rightarrow\infty}\frac{1}{N}\left|\sum_{n=1}^{N}e(\vec{m}\cdot x_{n})\mathbbm{1}_{B}(n)\right|=\lim_{N\rightarrow\infty}\frac{1}{N}\left|\sum_{n_{k}\in[1,N]}e(\vec{m}\cdot x_{n_{k}})\right|
limKd¯(B)K|k=1Ke(mxnk)|.\displaystyle\geq\lim_{K\rightarrow\infty}\frac{\underline{d}(B)}{K}\left|\sum_{k=1}^{K}e(\vec{m}\cdot x_{n_{k}})\right|.

Proof of (ii).

Let m=(m1,,md)d{0}\vec{m}=(m_{1},\cdots,m_{d})\in\mathbb{Z}^{d}\setminus\{0\} be arbitrary, let ((e(mxn))n=1,(e(mxn))n=1,(Nq)q=1)((e(\vec{m}\cdot x_{n}))_{n=1}^{\infty},(e(\vec{m}\cdot x_{n}))_{n=1}^{\infty},(N_{q})_{q=1}^{\infty}) be a permissible triple, and let \mathcal{H}, SS, and ı\i be as in Lemma 2.9. Let ξ:=ı1((e(mxn))n=1)\xi:=\i ^{-1}((e(\vec{m}\cdot x_{n}))_{n=1}^{\infty}), and let ξ=ξw+ξc\xi=\xi_{w}+\xi_{c} where ξc\xi_{c} has a pre-compact orbit under SS in the norm topology, and ξw\xi_{w} is weakly mixing with respect to SS. Letting (wn)n=1=ı(ξw)(w_{n})_{n=1}^{\infty}=\i (\xi_{w}) and (Cn)n=1=ı(ξc)(C_{n})_{n=1}^{\infty}=\i (\xi_{c}), we can assume without loss of generality that e(mxn)=Cn+dne(\vec{m}\cdot x_{n})=C_{n}+d_{n}. Using [5, Lemma 2.2.2] and [2, Theorem 2.25], we see that supn|Cn|supn|e(mxn)|=1\sup_{n}|C_{n}|\leq\sup_{n}|e(\vec{m}\cdot x_{n})|=1. Let (cn)n=1𝕊1(c_{n})_{n=1}^{\infty}\subseteq\mathbb{S}^{1} be such that cnCn¯=|Cn|c_{n}\overline{C_{n}}=|C_{n}|. We see that (cn)n=1(c_{n})_{n=1}^{\infty} is a compact sequence by Lemma 2.15(iii). Let x,X,S,x,X,S, and FF be as in Lemma 2.10 and satisfy F(Snx)=cnF(S^{n}x)=c_{n}. Now let (Nq)q=1(N_{q})_{q=1}^{\infty} be such that

limq1Nqn=1NqδSnx=μ,\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}\delta_{S^{n}x}=\mu, (34)

and recall that (X,S,μ)(X,S,\mu) has discrete spectrum by Lemma 2.10. Now let ϵ>0\epsilon>0 be arbitrary, and let g:=i=1Mdi𝟙Aig:=\sum_{i=1}^{M}d_{i}\mathbbm{1}_{A_{i}} be such that Fg<ϵ||F-g||_{\infty}<\epsilon and each AiA_{i} is open. For 1iM1\leq i\leq M, let (ci,n)n=1(c_{i,n})_{n=1}^{\infty} be given by ci,n=𝟙Ai(Snx)c_{i,n}=\mathbbm{1}_{A_{i}}(S^{n}x). To see that (ci,n)n=1(c_{i,n})_{n=1}^{\infty} is compact, let ((ci,n)n=1,(ci,n)n=1,(Nq)q=1)((c_{i,n})_{n=1}^{\infty},(c_{i,n})_{n=1}^{\infty},(N_{q}^{\prime})_{q=1}^{\infty}) be a permissible triple, and let (Nq′′)q=1(N_{q}^{\prime\prime})_{q=1}^{\infty} be any subsequence of (Nq)q=1(N_{q}^{\prime})_{q=1}^{\infty} for which

limq1Nq′′n=1Nq′′δSnx=μ,\lim_{q\rightarrow\infty}\frac{1}{N_{q}^{\prime\prime}}\sum_{n=1}^{N_{q}^{\prime\prime}}\delta_{S^{n}x}=\mu^{\prime}, (35)

with convergence taking place in the weak topology. Lemma 2.10 tells us that (X,S,μ)(X,S,\mu^{\prime}) has discrete spectrum, so 𝟙Ai\mathbbm{1}_{A_{i}} has precompact orbit under SS, which yields the desired result. It follows that if we let (ni,k)k=1(n_{i,k})_{k=1}^{\infty} be an increasing enumeration of those (ci,k)k=1(c_{i,k})_{k=1}^{\infty} that are 11, then each (ni,k)k=1(n_{i,k})_{k=1}^{\infty} would be a compact sequence of natural numbers if they have positive lower density, which is not necessarily the case. To overcome this difficulty, we let (ni,k,1)k=1(n_{i,k,1})_{k=1}^{\infty} be the increasing enumeration of (ni,k)k=12(n_{i,k})_{k=1}^{\infty}\cup 2\mathbb{N} and (ni,k,2)k=1(n_{i,k,2})_{k=1}^{\infty} be the increasing enumeration of (ni,k)k=1(2+1)(n_{i,k})_{k=1}^{\infty}\cup(2\mathbb{N}+1), so that (ni,k,j)k=1(n_{i,k,j})_{k=1}^{\infty} is a compact sequence of natural numbers for j=1,2j=1,2. Since (xnk)k=1(x_{n_{k}})_{k=1}^{\infty} is uniformly distributed for any compact sequence (nk)k=1(n_{k})_{k=1}^{\infty}, we see that for Ai=(ni,k)k=1A_{i}=(n_{i,k})_{k=1}^{\infty} and Ai,j:=(ni,k,j)k=1A_{i,j}:=(n_{i,k,j})_{k=1}^{\infty} we have

0\displaystyle 0 =limN1Nn=1Ne(mxn)𝟙Ai,j(n)=limN1Nn=1Ne(mxn), hence\displaystyle=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot x_{n})\mathbbm{1}_{A_{i,j}}(n)=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot x_{n})\text{, hence}
0\displaystyle 0 =limN1Nn=1Ne(mxn)(𝟙Ai,1(n)+𝟙Ai,2(n)1)\displaystyle=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot x_{n})\left(\mathbbm{1}_{A_{i,1}}(n)+\mathbbm{1}_{A_{i,2}}(n)-1\right)
=limN1Nn=1Ne(mxn)𝟙Ai(n), hence\displaystyle=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot x_{n})\mathbbm{1}_{A_{i}}(n)\text{, hence}
0\displaystyle 0 =limN1Nn=1Ne(mxn)i=1mdi𝟙Ai(n).\displaystyle=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot x_{n})\sum_{i=1}^{m}d_{i}\mathbbm{1}_{A_{i}}(n).

After recalling that i=1mdi𝟙Ai(n)=g(Snx)\sum_{i=1}^{m}d_{i}\mathbbm{1}_{A_{i}}(n)=g(S^{n}x), |g(Snx)F(Snx)|<ϵ|g(S^{n}x)-F(S^{n}x)|<\epsilon, and that ϵ>0\epsilon>0 was arbitrary, we see that

0\displaystyle 0 =limN1Nn=1Ne(mxn)cn=limN1Nn=1N(Cncn+dncn)\displaystyle=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot x_{n})c_{n}=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}(C_{n}c_{n}+d_{n}c_{n})
=limN1Nn=1N|Cn|lim supN1Nn=1N|Cn|2,\displaystyle=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}|C_{n}|\geq\limsup_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}|C_{n}|^{2},

so we may assume without loss of generality that Cn=0C_{n}=0 for all nn, so (e(mxn))n=1(e(\vec{m}\cdot x_{n}))_{n=1}^{\infty} is a nearly weakly mixing sequence. ∎

Proof of (iii).

Let (m,m1)d+d1{0}(\vec{m},\vec{m}_{1})\in\mathbb{Z}^{d+d_{1}}\setminus\{\vec{0}\} be arbitrary. Note that (e(mxn))n=1(e(\vec{m}\cdot x_{n}))_{n=1}^{\infty} is a spectrally Lebesgue (nearly weakly mixing) sequence when m0\vec{m}\neq\vec{0}, while (e(m1yn))n=1(e(\vec{m}_{1}\cdot y_{n}))_{n=1}^{\infty} is a spectrally singular (compact) sequence. Since (yn)n=1(y_{n})_{n=1}^{\infty} is uniformly distributed, it is clear that if m=0\vec{m}=\vec{0}, then m10\vec{m}_{1}\neq\vec{0} and

limN1Nn=1Ne((m+m1)(xn,yn))=limN1Nn=1Ne(m1yn)=0.\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e((\vec{m}+\vec{m_{1}})\cdot(x_{n},y_{n}))=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m_{1}}\cdot y_{n})=0. (36)

If m0\vec{m}\neq\vec{0}, then we can use Lemma 2.14 to see that

limN1Nn=1Ne((m+m1)(xn,yn))=limN1Nn=1Ne(mxn)e(m1yn)=0.\displaystyle\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e((\vec{m}+\vec{m_{1}})\cdot(x_{n},y_{n}))=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot x_{n})e(\vec{m}_{1}\cdot y_{n})=0.

Proof of (iv).

Firstly, we see that (0)n=1(0)_{n=1}^{\infty} is a c-sequence, so (xn)n=1(x_{n})_{n=1}^{\infty} is uniformly distributed. Since we have already proven (ii), it suffices to show that (xnk)k=1(x_{n_{k}})_{k=1}^{\infty} is uniformly distributed for an arbitrary compact sequence A=(nk)k=1A=(n_{k})_{k=1}^{\infty}\subseteq\mathbb{N}. For t[0,1]dt\in[0,1]^{d}, let (yn(t))n=1(y_{n}(t))_{n=1}^{\infty} be given by yn(t)=t𝟙A(t)y_{n}(t)=t\mathbbm{1}_{A}(t). Since (yn(t))n=1(y_{n}(t))_{n=1}^{\infty} is always a c-sequence, we see that (xn+yn(t))n=1(x_{n}+y_{n}(t))_{n=1}^{\infty} is uniformly distributed for any t(0,1)t\in(0,1). It follows that for all t[0,1]dt\in[0,1]^{d} we have

0=limN1Nn=1Nf(xn+yn(t))=limN1Nn=1Nf(xn+t𝟙A(n)), hence\displaystyle 0=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}f(x_{n}+y_{n}(t))=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}f(x_{n}+t\mathbbm{1}_{A}(n))\text{, hence}
0=limN1Nn=1Nf((xn+t)f(xn))𝟙A(n).\displaystyle 0=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}f((x_{n}+t)-f(x_{n}))\mathbbm{1}_{A}(n).

Consequently, if μ\mu is any weak limit point of {1Nn=1Nδxn𝟙A(n)}N=1\{\frac{1}{N}\sum_{n=1}^{N}\delta_{x_{n}}\mathbbm{1}_{A}(n)\}_{N=1}^{\infty}, then μ\mu is translation invariant, hence μ\mu is a constant multiple of the Lebesgue measure, which shows that (xnk)k=1(x_{n_{k}})_{k=1}^{\infty} is uniformly distributed. ∎

Remark 2.

It is natural to ask if items (ii) and (iv) of Theorem 3.1 have analogues for sL-sequences. While we do not have counterexamples showing that such analogues do not exist, we can explain why our current methods of proof do not extend to this situation. The proof of Theorem 3.1(ii) relies on the fact that if (X,,μ,T)(X,\mathscr{B},\mu,T) is a measure preserving system, then there is a σ\sigma-algebra 𝒦\mathcal{K}\subseteq\mathscr{B} with respect to which all eigenfunctions333We recall that the span closure of eigenfunctions is precisely those functions whose orbit under TT is pre-compact in the norm topology of L2(X,μ)L^{2}(X,\mu), which is why compact sequences can be associated to eigenfunctions. of TT are measurable. However, there need not exists σ\sigma-algebra 𝒮\mathcal{S}\subseteq\mathscr{B} with respect to which all fL2(X,μ)f\in L^{2}(X,\mu) that have singular spectrum are measurable. Similarly, the proof of Theorem 3.1(iv) (intuitively) uses the fact that if ff is in the span-closure of eigenfunctions of TT, then so it fkf^{k} for any kk\in\mathbb{N}, but the analogous statement for functions of singular spectrum is not true.

Our next two results can be seen as generalizations of Theorem 1.1.

Theorem 3.2.

Let dd\in\mathbb{N}. If (xn)n=1[0,1]d(x_{n})_{n=1}^{\infty}\subseteq[0,1]^{d} is such that

h=1D¯((xn+hxn)n=1)2<,\sum_{h=1}^{\infty}\overline{D}((x_{n+h}-x_{n})_{n=1}^{\infty})^{2}<\infty, (37)

then (xn)n=1(x_{n})_{n=1}^{\infty} is a sL-sequence.

Proof.

Let m=(m1,,md)d\vec{m}=(m_{1},\cdots,m_{d})\in\mathbb{Z}^{d} be arbitrary and let M=max1id(|mi|)M=\max_{1\leq i\leq d}(|m_{i}|). Using Theorem 2.3, we see that for hh\in\mathbb{N} we have

|lim supN1Nn=1Ne(m(xn+hxn))|(4πM)dD¯((xn+hxn)n=1), hence\displaystyle\left|\limsup_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot(x_{n+h}-x_{n}))\right|\leq(4\pi M)^{d}\overline{D}\left((x_{n+h}-x_{n})_{n=1}^{\infty}\right)\text{, hence}
h=1|lim supN1Nn=1Ne(m(xn+hxn))|2\displaystyle\sum_{h=1}^{\infty}\left|\limsup_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot(x_{n+h}-x_{n}))\right|^{2}
\displaystyle\leq (4πM)2dh=1D¯((xn+hxn)n=1)2<,\displaystyle(4\pi M)^{2d}\sum_{h=1}^{\infty}\overline{D}\left((x_{n+h}-x_{n})_{n=1}^{\infty}\right)^{2}<\infty,

so Lemma 2.11(i) tells us that (e(mxn))n=1(e(\vec{m}\cdot x_{n}))_{n=1}^{\infty} is a spectrally Lebesgue sequence. Since m\vec{m} was arbitrary, the desired result now follows from Corollary 2.13(ii). ∎

Theorem 3.3.

Let dd\in\mathbb{N}. If (xn)n=1[0,1]d(x_{n})_{n=1}^{\infty}\subseteq[0,1]^{d} is such that

limH1Hh=1HD¯((xn+hxn)n=1)=0,\lim_{H\rightarrow\infty}\frac{1}{H}\sum_{h=1}^{H}\overline{D}((x_{n+h}-x_{n})_{n=1}^{\infty})=0, (38)

then (xn)n=1(x_{n})_{n=1}^{\infty} is a wm-sequence.

Proof.

Let m=(m1,,md)d\vec{m}=(m_{1},\cdots,m_{d})\in\mathbb{Z}^{d} be arbitrary and let M=max1id(|mi|)M=\max_{1\leq i\leq d}(|m_{i}|). Using Theorem 2.3, we see that for hh\in\mathbb{N} we have

|lim supN1Nn=1Ne(m(xn+hxn))|(4πM)dD¯((xn+hxn)n=1), hence\displaystyle\left|\limsup_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot(x_{n+h}-x_{n}))\right|\leq(4\pi M)^{d}\overline{D}\left((x_{n+h}-x_{n})_{n=1}^{\infty}\right)\text{, hence}
limH1Hh=1H|lim supN1Nn=1Ne(m(xn+hxn))|\displaystyle\lim_{H\rightarrow\infty}\frac{1}{H}\sum_{h=1}^{H}\left|\limsup_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{m}\cdot(x_{n+h}-x_{n}))\right|
\displaystyle\leq (4πM)dlimH1Hh=1HD¯((xn+hxn)n=1)=0,\displaystyle(4\pi M)^{d}\lim_{H\rightarrow\infty}\frac{1}{H}\sum_{h=1}^{H}\overline{D}\left((x_{n+h}-x_{n})_{n=1}^{\infty}\right)=0,

so Lemma 2.11(ii) tells us that (e(mxn))n=1(e(\vec{m}\cdot x_{n}))_{n=1}^{\infty} is a spectrally Lebesgue sequence. Since m\vec{m} was arbitrary, the desired result now follows from Corollary 2.13(i). ∎

Remark 3.

In light of Lemma 4.2, we see that Theorem 3.2 does not characterize sL-sequences, and that Theorem 3.3 does not characterize wm-sequences. A characterization of wm-sequences using permissible pairs has already been obtained in [5], which we state below as Theorem 3.4. In light of Lemma 4.2, we see, intuitively speaking, that the reason only reason Theorem 3.3 is not a characterization of wm-sequences is that the lim sup\limsup that is in the definition of D¯((xn+hxn)n=1)\overline{D}((x_{n+h}-x_{n})_{n=1}^{\infty}) is attained by a sequence (Nq(h))q=1(N_{q}(h))_{q=1}^{\infty} that depends on hh. This also indicates that the utility of permissible pairs is that they do not allow the sequence (Nq)q=1(N_{q})_{q=1}^{\infty} to change with hh.

The situation for sL-sequences is similar, in that any potential characterization would have to use permissible pairs rather than D¯\overline{D}. However, we like to have sufficient conditions using D¯\overline{D} rather than permissible pairs for the sake of application. Let us now give an example to show why we do not pursue a more general sufficient condition for sL-sequences than that present in Theorem 3.2. Recall that there exist positive measure νm\nu\perp m on [0,1][0,1] whose Fourier coefficients tend to 0. Consequently, we may pick a convex decreasing sequence of real numbers (cn)n=1(c_{n})_{n=1}^{\infty} satisfying cn|ν^(n)|c_{n}\geq|\hat{\nu}(n)|, and it is a classical result [11, Theorem 4.1] that there exists a positive measure μ<<m\mu<<m for which μ^(n)=cn\hat{\mu}(n)=c_{n}. In particular, it is possible to decrease the Fourier coefficients of a measure μ<<m\mu<<m and as a result obtain the Fourier coefficients of a measure νm\nu\perp m. However, this phenomenon does not happen if we start with a measure μ\mu for which n=1|μ^(n)|2<.\sum_{n=1}^{\infty}|\hat{\mu}(n)|^{2}<\infty.

Theorem 3.4 ([5, Theorem 2.4.17]).

For (xn)n=1[0,1]d1(x_{n})_{n=1}^{\infty}\subseteq[0,1]^{d_{1}} the following are equivalent:

  1. (i)

    (xn)n=1(x_{n})_{n=1}^{\infty} is a wm-sequence.

  2. (ii)

    For all uniformly distributed (yn)n=1[0,1]d2(y_{n})_{n=1}^{\infty}\subseteq[0,1]^{d_{2}} and (Nq)q=1(N_{q})_{q=1}^{\infty}\subseteq\mathbb{N} for which
    ({(xn,yn+h)n=1}h=1,(Nq)q=1)(\{(x_{n},y_{n+h})_{n=1}^{\infty}\}_{h=1}^{\infty},(N_{q})_{q=1}^{\infty}) is a permissible pair, we have

    limH1Hh=1HD((xn,yn+h)n=1,(Nq)q=1)=0.\lim_{H\rightarrow\infty}\frac{1}{H}\sum_{h=1}^{H}D((x_{n},y_{n+h})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty})=0. (39)
  3. (iii)

    For all (Nq)q=1(N_{q})_{q=1}^{\infty}\subseteq\mathbb{N} for which ({(xn,xn+h)n=1}h=1,(Nq)q=1)(\{(x_{n},x_{n+h})_{n=1}^{\infty}\}_{h=1}^{\infty},(N_{q})_{q=1}^{\infty}) is a permissible pair, we have

    limH1Hh=1HD((xn,xn+h)n=1,(Nq)q=1)=0.\lim_{H\rightarrow\infty}\frac{1}{H}\sum_{h=1}^{H}D((x_{n},x_{n+h})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty})=0. (40)
  4. (iv)

    For all (Nq)q=1(N_{q})_{q=1}^{\infty}\subseteq\mathbb{N} for which ({(xn+hxn)n=1}h=1,(Nq)q=1))(\{(x_{n+h}-x_{n})_{n=1}^{\infty}\}_{h=1}^{\infty},(N_{q})_{q=1}^{\infty})) is a permissible pair, we have

    limH1Hh=1HD((xn+hxn)n=1,(Nq)q=1)=0.\lim_{H\rightarrow\infty}\frac{1}{H}\sum_{h=1}^{H}D((x_{n+h}-x_{n})_{n=1}^{\infty},(N_{q})_{q=1}^{\infty})=0. (41)

In light of Remark 3, it is surprising that our next theorem characterizes o-sequences without the use of permissible pairs.

Theorem 3.5.

(xn)n=1[0,1]d(x_{n})_{n=1}^{\infty}\subseteq[0,1]^{d} is an oo-sequence if and only if for each hh\in\mathbb{N},
(xn,xn+h)n=1[0,1]2d(x_{n},x_{n+h})_{n=1}^{\infty}\subseteq[0,1]^{2d} is uniformly distributed.

Proof.

For the first direction, let us assume that (xn,xn+h)n=1(x_{n},x_{n+h})_{n=1}^{\infty} is uniformly distributed in [0,1]2d[0,1]^{2d} for all hh\in\mathbb{N}. We see that for all k1,k2d\vec{k}_{1},\vec{k}_{2}\in\mathbb{Z}^{d} that are not both (0,0,,0)(0,0,\cdots,0) and any hh\in\mathbb{N} we have

limN1Nn=1Ne(k1xn+k2xn+h)=0.\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}e(\vec{k_{1}}\cdot x_{n}+\vec{k}_{2}\cdot x_{n+h})=0. (42)

Now let fC([0,1]d)f\in C([0,1]^{d}) satisfy [0,1]df𝑑md=0\int_{[0,1]^{d}}fdm^{d}=0, let ϵ(0,f)\epsilon\in(0,||f||_{\infty}) be arbitrary, and let KK be such that

f(x)k[K,K]dcke(kx)<ϵ,||f(x)-\sum_{k\in[-K,K]^{d}}c_{\vec{k}}e(\vec{k}\cdot x)||_{\infty}<\epsilon, (43)

for some (ck)k[K,K]d(c_{\vec{k}})_{\vec{k}\in[-K,K]^{d}}. Since we may assume without loss of generality that c(0,0,,0)=0c_{(0,0,\cdots,0)}=0, we see that for all hh\in\mathbb{N} we have

limN|1Nn=1Nf(xn+h)f(xn)¯|3ϵf+\displaystyle\lim_{N\rightarrow\infty}\left|\frac{1}{N}\sum_{n=1}^{N}f(x_{n+h})\overline{f(x_{n})}\right|\leq 3\epsilon||f||_{\infty}+
limN|1Nn=1N(k[K,K]dcke(kxn+h))(k[K,K]dcke(kxn))|\displaystyle\qquad\lim_{N\rightarrow\infty}\left|\frac{1}{N}\sum_{n=1}^{N}\left(\sum_{k\in[-K,K]^{d}}c_{k}e\left(\vec{k}\cdot x_{n+h}\right)\right)\left(\sum_{k\in[-K,K]^{d}}c_{k}e\left(-\vec{k}\cdot x_{n}\right)\right)\right|
=3ϵf+k1,k2[K,K]dlimN|1Nn=1Nck1ck2¯e((k1,k2)(xn+h,xn))|\displaystyle=3\epsilon||f||_{\infty}+\sum_{k_{1},k_{2}\in[-K,K]^{d}}\lim_{N\rightarrow\infty}\left|\frac{1}{N}\sum_{n=1}^{N}c_{k_{1}}\overline{c_{k_{2}}}e\left((\vec{k}_{1},-\vec{k}_{2})\cdot(x_{n+h},x_{n})\right)\right|
=3ϵf.\displaystyle=3\epsilon||f||_{\infty}.

Since ϵ>0\epsilon>0 was arbitrary, we are done with the first direciton. For the reverse direction, let us assume that (xn)n=1(x_{n})_{n=1}^{\infty} is an o-sequence. We will first show that (xn+hxn)n=1(x_{n+h}-x_{n})_{n=1}^{\infty} is uniformly distributed for all hh\in\mathbb{N}. To this end, let kd{(0,0,0)}\vec{k}\in\mathbb{Z}^{d}\setminus\{(0,0\cdots,0)\} and hh\in\mathbb{N} both be arbitrary and note that (e(kxn))n=1(e(\vec{k}\cdot x_{n}))_{n=1}^{\infty} is a nearly orthogonal sequence. Let (Nq)q=1(N_{q})_{q=1}^{\infty} be any sequence for which

limq1Nqn=1Nqe(k(xn+hxn))\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}e\left(\vec{k}\cdot(x_{n+h}-x_{n})\right) (44)

exists. By passing to a subsequence of (Nq)q=1(N_{q})_{q=1}^{\infty} if necessary, we may assume without loss of generality that ((e(kxn))n=1,(e(kxn))n=1,(Nq)q=1)\left(\left(e\left(\vec{k}\cdot x_{n}\right)\right)_{n=1}^{\infty},\left(e\left(\vec{k}\cdot x_{n}\right)\right)_{n=1}^{\infty},\left(N_{q}\right)_{q=1}^{\infty}\right) is a permissible triple. Since (e2πik,xn)n=1(e^{2\pi i\langle k,x_{n}\rangle})_{n=1}^{\infty} is a nearly orthogonal sequence it follows that

limq1Nqn=1Nqe(k(xn+hxn))=0,\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}e\left(\vec{k}\cdot(x_{n+h}-x_{n})\right)=0, (45)

from which it follows that (xn+hxn)n=1(x_{n+h}-x_{n})_{n=1}^{\infty} is indeed uniformly distributed for all hh\in\mathbb{N}. Now let hh\in\mathbb{N} be arbitrary, let k1,k2d\vec{k}_{1},\vec{k}_{2}\in\mathbb{Z}^{d} be such that k1\vec{k}_{1} and k2\vec{k}_{2} are not both (0,0,,0)(0,0,\cdots,0) and let (Nq)q=1(N_{q})_{q=1}^{\infty}\subseteq\mathbb{N} be such that

limq1Nqn=1Nqe((k1,k2)(xn,xn+h))\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{q}}e\left((\vec{k}_{1},\vec{k}_{2})\cdot(x_{n},x_{n+h})\right) (46)

exists. If k1\vec{k}_{1} or k2\vec{k}_{2} is (0,0,,0)(0,0,\cdots,0), then the limit in Equation (46) is 0 since the o-sequence (xn)n=1(x_{n})_{n=1}^{\infty} is uniformly distributed, so let us assume that neither of k1\vec{k}_{1} and k2\vec{k}_{2} are (0,0,,0)(0,0,\cdots,0). Note that for all cc\in\mathbb{C} we have that (e(k1xn)+ce(k2xn))n=1(e(\vec{k}_{1}\cdot x_{n})+ce(\vec{k}_{2}\cdot x_{n}))_{n=1}^{\infty} is a nearly orthogonal sequence since (xn)n=1(x_{n})_{n=1}^{\infty} is an o-sequence, so we once again see that

0\displaystyle 0 =limq1Nqq=1Nq(e(k1xn+h)+ce(k2xn+h)(e(k1xn)+c¯e(k2xn)\displaystyle=\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}(e(\vec{k}_{1}\cdot x_{n+h})+ce(\vec{k}_{2}\cdot x_{n+h})(e(-\vec{k}_{1}\cdot x_{n})+\overline{c}e(-\vec{k}_{2}\cdot x_{n})
=limq1Nqq=1Nqe(k1(xn+hxn))+|c|2limq1Nqq=1Nqe(k2(xn+hxn))\displaystyle=\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}e(\vec{k}_{1}\cdot(x_{n+h}-x_{n}))+|c|^{2}\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}e(\vec{k}_{2}\cdot(x_{n+h}-x_{n}))
=+climq1Nqq=1Nqe(k2xn+hk1xn)+c¯limq1Nqq=1Nqe(k1xn+hk2xn)\displaystyle{\color[rgb]{1,1,1}=}+c\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}e(\vec{k}_{2}\cdot x_{n+h}-\vec{k}_{1}\cdot x_{n})+\overline{c}\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}e(\vec{k}_{1}\cdot x_{n+h}-\vec{k}_{2}\cdot x_{n})
=climq1Nqq=1Nqe(k2xn+hk1xn)+c¯limq1Nqq=1Nqe(k1xn+hk2xn).\displaystyle=c\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}e(\vec{k}_{2}\cdot x_{n+h}-\vec{k}_{1}\cdot x_{n})+\overline{c}\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}e(\vec{k}_{1}\cdot x_{n+h}-\vec{k}_{2}\cdot x_{n}).

Letting A(c)A(c) represent the final quantity in the previous calculation, we observe that

0\displaystyle 0 =A(1)iA(i)\displaystyle=A(1)-iA(i)
=limq1Nqq=1Nqe(k2xn+hk1xn)+limq1Nqq=1Nqe(k1xn+hk2xn)\displaystyle=\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}e(\vec{k}_{2}\cdot x_{n+h}-\vec{k}_{1}\cdot x_{n})+\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}e(\vec{k}_{1}\cdot x_{n+h}-\vec{k}_{2}\cdot x_{n})
=i(ilimq1Nqq=1Nqe(k2xn+hk1xn)\displaystyle{\color[rgb]{1,1,1}=}-i\Bigg{(}i\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}e(\vec{k}_{2}\cdot x_{n+h}-\vec{k}_{1}\cdot x_{n})
ilimq1Nqq=1Nqe(k1xn+hk2xn))\displaystyle\qquad\qquad\qquad-i\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}e(\vec{k}_{1}\cdot x_{n+h}-\vec{k}_{2}\cdot x_{n})\Bigg{)}
=2limq1Nqq=1Nqe(k2xn+hk1xn).\displaystyle=2\lim_{q\rightarrow\infty}\frac{1}{N_{q}}\sum_{q=1}^{N_{q}}e(\vec{k}_{2}\cdot x_{n+h}-\vec{k}_{1}\cdot x_{n}).

Since (k1,k2)(k1,k2)(k_{1},k_{2})\mapsto(-k_{1},k_{2}) is a bijection from (d{(0,,0)})2\left(\mathbb{Z}^{d}\setminus\{(0,\cdots,0)\}\right)^{2} to itself, we see that (xn,xn+h)n=1(x_{n},x_{n+h})_{n=1}^{\infty} is uniformly distributed for all hh\in\mathbb{N}. ∎

It is worth noting that if p(x)=anxn++a1x+a0[x]p(x)=a_{n}x^{n}+\cdots+a_{1}x+a_{0}\in\mathbb{R}[x] is such that n2n\geq 2 and aja_{j} is irrational for some j2j\geq 2, then (p(n))n=1(p(n))_{n=1}^{\infty} is an o-sequence. In [5, Theorem 2.4.30] is it shown that (g(n)α)n=1(\lfloor g(n)\rfloor\alpha)_{n=1}^{\infty} is an o-sequence for a large class of tempered functions gg, such that g(n)=ntg(n)=n^{t} with t(1,)t\in(1,\infty)\setminus\mathbb{N}, and any irrational α\alpha.

4. Examples

Theorems 3.2 and 3.3 allows us to construct examples of sL-sequences and wm-sequences. Our next result shows us that we can use measure preserving systems to construct examples of sequences that appear in Definition 2.6.

Lemma 4.1.

Let 𝒳:=([0,1]d,,md,T)\mathcal{X}:=([0,1]^{d},\mathscr{B},m^{d},T) be a measure preserving system, x[0,1]dx\in[0,1]^{d}, and xn=Tnxx_{n}=T^{n}x. The following hold for a.e. xXx\in X:

  1. (i)

    If 𝒳\mathcal{X} is weakly mixing, then (xn)n=1(x_{n})_{n=1}^{\infty} is a wm-sequence.

  2. (ii)

    If 𝒳\mathcal{X} has Lebesgue spectrum, then (xn)n=1(x_{n})_{n=1}^{\infty} is a sL-sequence.

  3. (iii)

    If 𝒳\mathcal{X} has discrete spectrum, then (xn)n=1(x_{n})_{n=1}^{\infty} is a c-sequence.

  4. (iv)

    If 𝒳\mathcal{X} has discrete spectrum and AA\in\mathscr{B}, then {nk|xnkA}\{n_{k}\ |\ x_{n_{k}}\in A\} is a compact sequence of natural numbers.

  5. (v)

    If 𝒳\mathcal{X} has singular spectrum, then (xn)n=1(x_{n})_{n=1}^{\infty} is a ss-sequence.

  6. (vi)

    If 𝒳\mathcal{X} has singular spectrum and AA\in\mathscr{B}, then {nk|xnkA}\{n_{k}\ |\ x_{n_{k}}\in A\} is a spectrally singular sequence of natural numbers.

Proof.

To prove (i) and (ii), we first observe that the system 𝒳\mathcal{X} is ergodic. Using the pointwise ergodic theorem, we see that for all hh\in\mathbb{N} and fC([0,1]d){𝟙A|A}f\in C([0,1]^{d})\cup\{\mathbbm{1}_{A}\ |\ A\in\mathscr{B}\} we have

limN1Nn=1Nf(Tn+hx)f(Tnx)¯=01Thff¯𝑑md,\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}f(T^{n+h}x)\overline{f(T^{n}x)}=\int_{0}^{1}T^{h}f\overline{f}dm^{d}, (47)

for a.e. xXx\in X, from which the desired result follows.

We now proceed to prove items (iii)-(vi). Let {𝒳u:=(Xu,u,mu,Tu)|uU}\{\mathcal{X}_{u}:=(X_{u},\mathscr{B}_{u},m_{u},T_{u})\ |\ u\in U\} denote the ergodic components of 𝒳\mathcal{X}. We note that if 𝒳\mathcal{X} has discrete (singular) spectrum, then mdm^{d}-a.e. x[0,1]dx\in[0,1]^{d} is contained in an ergodic component 𝒳u\mathcal{X}_{u} that also has discrete (singular) spectrum. We observe that for mum_{u}-a.e. xXux\in X_{u}, we have

γf(h):=limN1Nn=1Nf(Tn+hx)f(Tnx)¯=XThff¯𝑑mu\gamma_{f}(h):=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{n=1}^{N}f(T^{n+h}x)\overline{f(T^{n}x)}=\int_{X}T^{h}f\overline{f}dm_{u} (48)

for all fC([0,1]d)f\in C([0,1]^{d}). Since (γf(h))h=1(\gamma_{f}(h))_{h=1}^{\infty} are the Fourier coefficients of a measure that is absolutely continuous to the maximal spectral type of 𝒳u\mathcal{X}_{u}, we have proven (iii) and (v). To show (iv) and (vi), we repeat the previous argument with f=𝟙Af=\mathbbm{1}_{A} in Equation (48). ∎

For a discussion of the prevalence of measure preserving systems with singular spectrum as well as examples, we refer the reader to [6, Remark 1.16]. For now, we only mention one particularly aesthetic family of examples. Recall that the Thue-Morse sequence (wn)n=1{1,1}(w_{n})_{n=1}^{\infty}\in\{-1,1\}^{\mathbb{N}} is given by wn=(1)enw_{n}=(-1)^{e_{n}} where ene_{n} is the number of times the digit 11 appears in the base 22 expansion of nn. Let (nk)k=1(n_{k})_{k=1}^{\infty} be an increasing enumeration of those nn for which wn=1w_{n}=1. The sequence (nk)k=1(n_{k})_{k=1}^{\infty} is a spectrally singular sequence as a consequence of the work of Mahler [13] (see also [9]). Consequently, if (xn)n=1[0,1]d(x_{n})_{n=1}^{\infty}\subseteq[0,1]^{d} is a sL-sequence, then (xnk)k=1(x_{n_{k}})_{k=1}^{\infty} is uniformly distributed.

More generally, for r2r\in\mathbb{N}_{\geq 2}, we define sr(n)s_{r}(n) to be the sum of the digits of the base rr expansion of nn. It was shown in [4] that if cc\in\mathbb{R} is such that (r1)c(r-1)c\notin\mathbb{Z}, then ((e(csr(n)))n=1,(e(csr(n)))n=1,(q)q=1)((e(cs_{r}(n)))_{n=1}^{\infty},(e(cs_{r}(n)))_{n=1}^{\infty},(q)_{q=1}^{\infty}) is a permissible triple, any that (e(csr(n)))n=1(e(cs_{r}(n)))_{n=1}^{\infty} is a spectrally singular sequence. We observe that wn=e(12s2(n))w_{n}=e\left(\frac{1}{2}s_{2}(n)\right). A consequence of this is the following. If r2r\in\mathbb{N}_{\geq 2} and 0j<r0\leq j<r, and (nk(j,r))k=1(n_{k}(j,r))_{k=1}^{\infty} is an increasing enumeration of the nn\in\mathbb{N} for which sr(n)j(modr)s_{r}(n)\equiv j\pmod{r}, then (nk(j,r))k=1(n_{k}(j,r))_{k=1}^{\infty} is a spectrally singular sequence. Consequently, if (xn)n=1[0,1]d(x_{n})_{n=1}^{\infty}\subseteq[0,1]^{d} is a sL-sequence, then (xnk(j,r))k=1(x_{n_{k}(j,r)})_{k=1}^{\infty} is uniformly distributed.

Our next example justifies the usage of permissible pairs in Theorem 3.4, and shows that Theorems 3.2 and 3.3 are not characterizations of sL-sequences and wm-sequences respectively.

Lemma 4.2.

There exists a sL-sequence (xn)n=1[0,1](x_{n})_{n=1}^{\infty}\subseteq[0,1] for which

D¯((xn+hxn)n=1)=12\overline{D}((x_{n+h}-x_{n})_{n=1}^{\infty})=\frac{1}{2} (49)

for all hh\in\mathbb{N}.

Proof.

Let (Nq)q=0(N_{q})_{q=0}^{\infty} be given by N0=0N_{0}=0 and Nq=(q!)2+Nq1N_{q}=(q!)^{2}+N_{q-1}. Let f:2f:\mathbb{N}\rightarrow\mathbb{N}^{2} the bijection corresponding to {(1,1),(1,2),(2,1),(3,1),(2,2),(1,3),(1,4),}\{(1,1),(1,2),(2,1),(3,1),(2,2),(1,3),(1,4),\cdots\}, let f1(q)f_{1}(q) denote the first coordinate of f(q)f(q), and observe that the first coordinate of f(q)f(q) always divides q!q!. For all qq\in\mathbb{N}, let us divide (Nq1,Nq](N_{q-1},N_{q}] into MqM_{q} consecutive intervals of the form {(ai,q,ai,q+2f1(q)]}\{(a_{i,q},a_{i,q}+2f_{1}(q)]\}. We now define

xn:={n22 if n(ai,q,ai,q+f1(q)](nf1(q))22 if n(ai,q+f1(q),ai,q+2f1(q)].x_{n}:=\begin{cases}n^{2}\sqrt{2}&\text{ if }n\in(a_{i,q},a_{i,q}+f_{1}(q)]\\ (n-f_{1}(q))^{2}\sqrt{2}&\text{ if }n\in(a_{i,q}+f_{1}(q),a_{i,q}+2f_{1}(q)].\end{cases} (50)

We will now show that Equation (49) holds. Now let us fix some hh\in\mathbb{N} and let (qh,k)k=1(q_{h,k})_{k=1}^{\infty} be an increasing enumeration of f11(h)f_{1}^{-1}(h). We see that for n(Nqh,k1,Nqh,kh]n\in(N_{q_{h,k}-1},N_{q_{h,k}}-h]

xn+hxn={0 if n(ai,qh,k,ai,qh,k+h]4nh2 if n(ai,qh,k+h,ai,qh,k+2h].x_{n+h}-x_{n}=\begin{cases}0&\text{ if }n\in(a_{i,q_{h,k}},a_{i,q_{h,k}}+h]\\ 4nh\sqrt{2}&\text{ if }n\in(a_{i,q_{h,k}}+h,a_{i,q_{h,k}}+2h].\end{cases} (51)

Now let

I1,k=i=1Mqh,k(ai,qh,k,ai,qh,k+h] and I2,k=i=1Mqh,k(ai,qh,k+h,ai,qh,k+2h].I_{1,k}=\bigcup_{i=1}^{M_{q_{h,k}}}(a_{i,q_{h,k}},a_{i,q_{h,k}}+h]\text{ and }I_{2,k}=\bigcup_{i=1}^{M_{q_{h,k}}}(a_{i,q_{h,k}}+h,a_{i,q_{h,k}}+2h]. (52)

Firstly, we see that D|I1,k|((xn+hxn)nI1,k)=1D_{|I_{1,k}|}((x_{n+h}-x_{n})_{n\in I_{1,k}})=1. Secondly, we see that

(xn+hxn)nI2,k=j=h2h((Nqh,k1h+4hj+2hn)2)n=1Mqh,k.(x_{n+h}-x_{n})_{n\in I_{2,k}}=\bigcup_{j=h}^{2h}((N_{q_{h,k}-1}h+4hj+2hn^{\prime})\sqrt{2})_{n^{\prime}=1}^{M_{q_{h,k}}}. (53)

Since 2\sqrt{2} is an algebraic irrational, we may use Theorem 2.2 to pick a constant C=C(2,12)C=C(\sqrt{2},\frac{1}{2}) such that for any M<Mqh,kM<M_{q_{h,k}} we have

DM((Nqh,k1h+4hj+2hn)2)n=1MCM12D_{M}((N_{q_{h,k}-1}h+4hj+2hn^{\prime})\sqrt{2})_{n^{\prime}=1}^{M}\leq CM^{-\frac{1}{2}} (54)

After recalling that the discrepancy function is subadditive, we see that

D|I2,k|((xn+hxn)nI2,k)hCMqh,k12, hence\displaystyle D_{|I_{2,k}|}((x_{n+h}-x_{n})_{n\in I_{2,k}})\leq hCM_{q_{h,k}}^{-\frac{1}{2}}\text{, hence}
DNqh,kNqh,k1((xn+hxn)n=Nqh,k1+1Nqh,k)(12hCMqh,k12,12+hCMqh,k12).\displaystyle D_{N_{q_{h,k}}-N_{q_{h,k}-1}}\left((x_{n+h}-x_{n})_{n=N_{q_{h,k}-1}+1}^{N_{q_{h,k}}}\right)\in\left(\frac{1}{2}-hCM_{q_{h,k}}^{-\frac{1}{2}},\frac{1}{2}+hCM_{q_{h,k}}^{-\frac{1}{2}}\right).

Since Nq>>Nq1N_{q}>>N_{q-1}, we see that

limkDNqh,k(xn+hxn)n=1Nqh,k=limkDNqh,kNqh,k1(xn+hxn)n=Nqh,k1+1Nqh,k=12.\lim_{k\rightarrow\infty}D_{N_{q_{h,k}}}(x_{n+h}-x_{n})_{n=1}^{N_{q_{h,k}}}=\lim_{k\rightarrow\infty}D_{N_{q_{h,k}}-N_{q_{h,k}-1}}(x_{n+h}-x_{n})_{n=N_{q_{h,k}-1}+1}^{N_{q_{h,k}}}=\frac{1}{2}.

The calculations thusfar show that D¯((xn+hxn)n=1)12\overline{D}((x_{n+h}-x_{n})_{n=1}^{\infty})\geq\frac{1}{2} for all hh\in\mathbb{N}. For the reverse inequality, it suffices to show that

limqqf11(h)DNq((xn+hxn)n=1)=0.\lim_{\underset{q\notin f_{1}^{-1}(h)}{q\rightarrow\infty}}D_{N_{q}}((x_{n+h}-x_{n})_{n=1}^{\infty})=0. (55)

To this end, let h1h2h_{1}\neq h_{2}\in\mathbb{N} be arbitrary, and observe that for
n(Nqh1,k1,Nqh1,kh2]n\in(N_{q_{h_{1},k}-1},N_{q_{h_{1},k}}-h_{2}] we have

xn+h2xn=((n+f(n))2n2)2=2f(n)n2+f(n)22,x_{n+h_{2}}-x_{n}=((n+f(n))^{2}-n^{2})\sqrt{2}=2f(n)n\sqrt{2}+f(n)^{2}\sqrt{2}, (56)

where f:{0}f:\mathbb{N}\rightarrow\mathbb{Z}\setminus\{0\} is some periodic function with period pp dividing 2h12h_{1}. Using the same bound as before for the discrepancy of (n2)n=1(n\sqrt{2})_{n=1}^{\infty}, we see that for MMqh1,kh2M\leq M_{q_{h_{1},k}}-h_{2} and 1j2h11\leq j\leq 2h_{1} we have

DM((xNqh1,k1+2h1n+j+h2xNqh1,k1+2h1n+j)n=1MC(h1,h2,j)M12, hence\displaystyle D_{M}((x_{N_{q_{h_{1},k}-1}+2h_{1}n^{\prime}+j+h_{2}}-x_{N_{q_{h_{1},k}-1}+2h_{1}n^{\prime}+j})_{n^{\prime}=1}^{M}\leq C(h_{1},h_{2},j)M^{-\frac{1}{2}}\text{, hence}
D2h1M((xNqh1,k1+n+h2xNqh1,k1+n)n=12h1M)C(h1,h2)M12,\displaystyle D_{2h_{1}M}((x_{N_{q_{h_{1},k}-1}+n+h_{2}}-x_{N_{q_{h_{1},k}-1}+n})_{n=1}^{2h_{1}M})\leq C(h_{1},h_{2})M^{-\frac{1}{2}},

from which the desired result follows.

It remains to show that (xn)n=1(x_{n})_{n=1}^{\infty} is a sL-sequence. To this end, let \ell\in\mathbb{N} be arbitrary and let ((e(xn))n=1,(e(xn))n=1,(Nm)m=1)((e(\ell x_{n}))_{n=1}^{\infty},(e(\ell x_{n}))_{n=1}^{\infty},(N_{m})_{m=1}^{\infty}) be a permissible triple. It suffices to show that

limm1Nqn=1Nme((xn+hxn))=0,\lim_{m\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{m}}e(\ell(x_{n+h}-x_{n}))=0, (57)

for all but at most 33 values of hh\in\mathbb{N}. If Equation (57) held for all hh\in\mathbb{N}, then we would be done, so let h0h_{0}\in\mathbb{N} be such that

limm1Nqn=1Nme((xn+h0xn))=c0.\lim_{m\rightarrow\infty}\frac{1}{N_{q}}\sum_{n=1}^{N_{m}}e(\ell(x_{n+h_{0}}-x_{n}))=c\neq 0. (58)

Theorem 2.3 tells us that

12|c|4πDNm((xn+h0xn)n=1Nm),\frac{1}{2}\cdot\frac{|c|}{4\pi\ell}\leq D_{N_{m}}\left((x_{n+h_{0}}-x_{n})_{n=1}^{N_{m}}\right), (59)

for all mm0m\geq m_{0}. Since Nm>>Nm1N_{m}>>N_{m-1}, we must have that Nqh0,kmNm<Nqh0,km+1N_{q_{h_{0},k_{m}}}\leq N_{m}<N_{q_{h_{0},k_{m}}+1} for all mm0m\geq m_{0}. Since f1(qh0,km+1){h01,h0,h0+1}f_{1}(q_{h_{0},k_{m}}+1)\in\{h_{0}-1,h_{0},h_{0}+1\}, we see that Equation (57) holds for all h{h01,h0,h0+1}h\notin\{h_{0}-1,h_{0},h_{0}+1\}. ∎

The sequence constructed in the previous Lemma is an example of an sL-sequence that is not an o-sequence, as any o-sequence (xn)n=1(x_{n})_{n=1}^{\infty} will have (xn+hxn)n=1(x_{n+h}-x_{n})_{n=1}^{\infty} be uniformly distributed for all hh\in\mathbb{N}. Our next result gives an another example of a sL-sequence that is not an o-sequence.

Lemma 4.3 ([5, Theorem 2.4.28]).

The sequence (xn)n=1[0,1](x_{n})_{n=1}^{\infty}\subseteq[0,1] given by

xn={n2α if n is even2(n1)2α if n is oddx_{n}=\begin{cases}n^{2}\alpha&\text{ if }n\text{ is even}\\ 2(n-1)^{2}\alpha\text{ if }n\text{ is odd}\end{cases} (60)

is a sL-sequence but not an o-sequence. In particular, (xn+hxn)n=1(x_{n+h}-x_{n})_{n=1}^{\infty} is uniformly distributed for all hh\in\mathbb{N} and (xn,xn+h)n=1(x_{n},x_{n+h})_{n=1}^{\infty} is uniformly distributed if and only if h>1h>1.

Acknowledgement.

I would like to thank Mariusz Lemańczyk for helpful discussions regarding spectral theory. I would also like to thank Teturo Kamae for reading an earlier draft of this paper and making some useful remarks. I acknowledge being supported by grant 2019/34/E/ST1/00082 for the project “Set theoretic methods in dynamics and number theory,” NCN (The National Science Centre of Poland).

5. REFERENCES

References

  • [1] \byBergelson, V. Weakly mixing PET. Ergodic Theory Dynam. Systems, 7(3):337-349,1987.
  • [2] \byBergelson, V., and McCutcheon, R. Idempotent ultrafilters, multiple weak mix- ing and Szemerédi’s theorem for generalized polynomials. In: J. Anal. Math. 111 (2010), pp. 77–130. issn: 0021-7670.
  • [3] \byBergelson, V. and Moreira, J. Van der Corput’s difference theorem: some mod- ern developments. Indag. Math. (N.S.) 27.2 (2016), pp. 437–479. issn: 0019-3577.
  • [4] \byCoquet, J. and Kamae, T. and Mendès France, M. Sur la mesure spectrale de certaines suites arithmétiques. Bull. Soc. Math. France 105 (1977), pp. 369-384.
  • [5] \byFarhangi, S. Topics in ergodic theory and ramsey theory, PhD Dissertation, the Ohio State University, 2022.
  • [6] \byFarhangi, S. A generalization of van der Corput’s difference theorem with applications to recurrence and multiple ergodic averages. Dynamical Systems, DOI: 10.1080/14689367.2023.2230160 2023.
  • [7] \byEisner, T. and Farkas, B. and Haase, M. and Nagel, R. Operator theoretic aspects of ergodic theory. Graduate Texts in Mathematics, 272 Springer, Cham, (2015) pp. xviii+628.
  • [8] \byFrantzikinakis, N. and Host, B. Furstenberg systems of bounded multiplicative functions and applications. Int. Math. Res. Not. IMRN 8, (2021), pp. 6077–6107.
  • [9] \byKakutani, S. Ergodic theory of shift transformations. Proc. Fifth Berkeley Sympos.on Math. Statist. and Probability, California Univ. 1967.
  • [10] \byKamae, T. and Mendès France, M. Van der Corput’s difference theorem. Israel J. Math. 31.3-4, (1978) pp. 335-342.
  • [11] \byKatznelson, Y. An introduction to harmonic analysis. Cambridge Mathematical Library. Cambridge University Press, Cambridge, third edition, 2004.
  • [12] \byKuipers, L. and Niederreiter, H. Uniform distribution of sequences. Pure and Applied Mathematics. Wiley-Interscience [John Wiley & Sons], New York-London-Sydney, 1974.
  • [13] \byMahler, K. On the translation properties of a simple class of arithmetical functions. Journal of Mat. Phys. (1927)
  • [14] \byMoreira, J. and Richter, F. K. and Robertson, D. A proof of a sumset conjecture of Erdős. Ann. of Math. (2) 189.2 (2019), pp. 605–652. issn: 0003-486X.
  • [15] \byTserunyan, A. A Ramsey theorem on semigroups and a general van der Corput lemma. J. Symb. Log. (2) 81 (2016) pp. 718–741.
  • [16] \byvan der Corput, J. G. Diophantische Ungleichungen. I. Zur Gleichverteilung Modulo Eins. Acta Math., 56(1):373–456, 1931.