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

Uniform distribution in nilmanifolds along functions from a Hardy field

Florian K. Richter
Abstract

We study equidistribution properties of translations on nilmanifolds along functions of polynomial growth from a Hardy field. More precisely, if X=G/ΓX=G/\Gamma is a nilmanifold, a1,,akGa_{1},\ldots,a_{k}\in G are commuting nilrotations, and f1,,fkf_{1},\ldots,f_{k} are functions of polynomial growth from a Hardy field then we show that

  • the distribution of the sequence a1f1(n)akfk(n)Γa_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)}\Gamma is governed by its projection onto the maximal factor torus, which extends Leibman’s Equidistribution Criterion from polynomials to a much wider range of functions; and

  • the orbit closure of a1f1(n)akfk(n)Γa_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)}\Gamma is always a finite union of sub-nilmanifolds, which extends previous work of Leibman and Frantzikinakis on this topic.

1.   Introduction

The study of the distribution of orbital sequences in nilsystems is an important part of contemporary ergodic theory. Not only do nilsystems play a key role in the structure theory of measure preserving systems (cf. [HK18]), but they are also tightly connected to the theory of higher order Fourier analysis, which finds applications to combinatorics and number theory (cf. [Tao12]). The purpose of this article is to study the distribution of orbits in nilsystems along functions of polynomial growth from a Hardy field. This has applications to additive combinatorics and leads to new refinements of Szemerédi’s theorem on arithmetic progressions (see Section 1.1). Our main results in this direction are Theorems A, B, C, and D below, which expand on the work of Leibman on the uniform distribution of polynomial sequences in nilmanifolds [Leibman05a], and the work of Frantzikinakis on the uniform distribution of nil-orbits along functions of different polynomial growth from a Hardy field [Frantzikinakis09]. Our results also connect to various conjectures and problems posed by Frantzikinakis over the years, including [Frantzikinakis09, Conjecture on p. 357], [Frantzikinakis10, Problems 1 and 4], [Frantzikinakis15b, Problem 1], and [Frantzikinakis16arXiv, Problems 23 and 25].

A closed subgroup Γ\Gamma of a Lie group GG is called uniform if the quotient G/ΓG/\Gamma is compact, and it is called discrete if there exists a cover of Γ\Gamma by open subsets of the ambient group GG in which every open set contains exactly one element of Γ\Gamma. Given a (ss-step) nilpotent Lie group GG and a uniform and discrete subgroup Γ\Gamma of GG, the quotient space XG/ΓX\coloneqq G/\Gamma is called a (ss-step) nilmanifold. For any group element aGa\in G and point xXx\in X, we define the translation of xx by aa as ax(ab)Γax\coloneqq(ab)\Gamma, where bb is any element in GG such that x=bΓx=b\Gamma. This way, the Lie group GG acts continuously and transitively on the nilmanifold XX. There exists a unique Borel probability measure on XX invariant under this action by GG, called the Haar measure on XX (see [Raghunathan72]), which we denote by μX\mu_{X}. A sequence (xn)n(x_{n})_{n\in\mathbb{N}} of points in XX is then said to be uniformly distributed in XX if

limN1Nn=1NF(xn)=F𝖽μX\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}F(x_{n})=\int F\leavevmode\nobreak\ \mathsf{d}\mu_{X}

holds for every continuous function F𝖢(X)F\in\mathsf{C}(X).

When it comes to the study of uniform distribution in nilmanifolds, an important role is played by the largest toral factor of the nilmanifold, called the maximal factor torus. Let GG^{\circ} denote the identity component of a nilpotent Lie group GG, and let [G,G][G^{\circ},G^{\circ}] be the commutator subgroup generated by GG^{\circ}.

Definition 1.1 (Maximal factor torus).

Given a connected nilmanifold X=G/ΓX=G/\Gamma, the maximal factor torus of XX is the quotient [G,G]\X[G^{\circ},G^{\circ}]\backslash X. We will use ϑ:X[G,G]\X\vartheta\colon X\to[G^{\circ},G^{\circ}]\backslash X to denote the natural factor map from XX onto [G,G]\X[G^{\circ},G^{\circ}]\backslash X.

The maximal factor torus is diffeomorphic to a torus 𝕋dd/d\mathbb{T}^{d}\coloneqq\mathbb{R}^{d}/\mathbb{Z}^{d} whose dimension dd equals the dimension of the quotient group G/[G,G]G^{\circ}/[G^{\circ},G^{\circ}]. Moreover, as the name suggests, it is the torus of highest dimension that is a factor of XX.

If X=G/ΓX=G/\Gamma is connected then it is well-known that the distribution of orbits along many sequences is governed by their projection onto [G,G]\X[G^{\circ},G^{\circ}]\backslash X. A classical result in this direction is Green’s Theorem [Green61] (see also [AGH63, Parry69, Parry70, Leibman05a]) which states that a niltranslation acts ergodically on the nilmanifold if and only if it acts ergodically on the maximal factor torus. An important generalization of Green’s Theorem is Leibman’s Equidistribution Criterion for polynomial sequences. Given an element aa in a simply connected nilpotent Lie group GG, we write 𝖽𝗈𝗆(a)\mathsf{dom}(a) for the set of all tt\in\mathbb{R} for which ata^{t} is a well-defined element of the group.111For instance, a rational number r/q{r}/{q} with gcd(r,q)=1\gcd(r,q)=1 belongs to 𝖽𝗈𝗆(a)\mathsf{dom}(a) if and only if there exists bGb\in G such that bq=arb^{q}=a^{r}. Since GG is assumed to be simply connected, if such an element bb exists then it must be unique. Note that 𝖽𝗈𝗆(a)=\mathsf{dom}(a)=\mathbb{R} if and only if aGa\in G^{\circ}.

Theorem 1.2 (Leibman’s Equidistribution Criterion, [Leibman05a, Theorem C]).

Let GG be a simply connected nilpotent Lie group and Γ\Gamma a uniform and discrete subgroup of GG. Suppose

u(n)=a1p1(n)akpk(n),n,u(n)\,=\,a_{1}^{p_{1}(n)}\cdot\ldots\cdot a_{k}^{p_{k}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G are commuting and p1,,pk[t]p_{1},\ldots,p_{k}\in\mathbb{R}[t] with pi()𝖽𝗈𝗆(ai)p_{i}(\mathbb{N})\subset\mathsf{dom}(a_{i}). Then the following are equivalent:

  1. (i)

    (u(n)Γ)n(u(n)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in the nilmanifold X=G/ΓX=G/\Gamma.

  2. (ii)

    (ϑ(u(n)Γ))n(\vartheta(u(n)\Gamma))_{n\in\mathbb{N}} is uniformly distributed in the maximal factor torus [G,G]\X[G^{\circ},G^{\circ}]\backslash X.

Remark 1.3.

Leibman actually proves 1.2 without the assumption that the elements a1,,akGa_{1},\ldots,a_{k}\in G are commuting. But for the scope of this paper, we stick to the commuting case because it suffices for the applications to combinatorics that we have in mind.

Closely related to 1.2 is Leibman’s Equidistribution Theorem, which describes the orbit closure of polynomial sequences in a nilmanifold.

Theorem 1.4 (Leibman’s Equidistribution Theorem, [Leibman05a, Theorem B]).

Let GG be a simply connected nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and

u(n)a1p1(n)akpk(n),n,u(n)\,\coloneqq\,a_{1}^{p_{1}(n)}\cdot\ldots\cdot a_{k}^{p_{k}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G are commuting and p1,,pk[t]p_{1},\ldots,p_{k}\in\mathbb{R}[t] with pi()𝖽𝗈𝗆(ai)p_{i}(\mathbb{N})\subset\mathsf{dom}(a_{i}). Then there exists a closed and connected subgroup HH of GG and points x0,,xq1Xx_{0},\ldots,x_{q-1}\in X such that, for all r=0,1,,q1r=0,1,\ldots,q-1, the set YrHxrY_{r}\coloneqq Hx_{r} is a closed sub-nilmanifold of XX and the sequence (u(qn+r)Γ)n(u(qn+r)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in YrY_{r}.

It is natural to ask whether Theorems 1.2 and 1.4 remain true if one replaces the polynomials p1,,pkp_{1},\ldots,p_{k} with other sufficiently smooth and eventually monotone functions of polynomial growth. For instance, we can consider the class of logarithmico-exponential functions introduced by Hardy in [Hardy12, Hardy10]. This class, which we denote by \mathcal{LE}, consists of all real-valued functions defined on some half-line [t0,)[t_{0},\infty) that can be build from real polynomials, the logarithmic function log(t)\log(t), and the exponential function exp(t)\exp(t) using the standard arithmetical operations ++, -, \cdot, ÷\div and the operation of composition. Examples of logartihmico-exponential functions are p(t)/q(t){p(t)}/{q(t)} for p(t),q(t)[t]p(t),q(t)\in\mathbb{R}[t], tct^{c} for cc\in\mathbb{R}, t/log(t){t}/{\log(t)}, and ete^{\sqrt{t}}, as well as any products or linear combinations thereof.

\mathcal{LE} is an example of a so-called Hardy field. Although our main results apply to arbitrary Hardy fields, and are stated as such in this introduction, we delay giving the definition of a Hardy field until Section 2 (see 2.1). Instead, we ask the reader to keep \mathcal{LE} as a representative example in mind, since our results are already new and interesting for this class.

Given two functions f,g:[1,)f,g\colon[1,\infty)\to\mathbb{R} we will write f(t)g(t)f(t)\prec g(t) when g(t)/f(t){g(t)}/{f(t)}\to\infty as tt\to\infty, and f(t)g(t)f(t)\ll g(t) when there exist C>0C>0 and t01t_{0}\geqslant 1 such that f(t)Cg(t)f(t)\leqslant Cg(t) for all tt0t\geqslant t_{0}. We say f(t)f(t) has polynomial growth if it satisfies |f(t)|td|f(t)|\ll t^{d} for some dd\in\mathbb{N}; in this case the smallest such dd is called the degree of ff and denoted by deg(f)\deg(f).

In the case of tori, the uniform distribution of functions from a Hardy field has been studied extensively by Boshernitzan [Boshernitzan94]. In the case of nilmanifolds, Frantzikinakis obtained the following result.

Theorem 1.5 ([Frantzikinakis09, Theorem 1.3, part (i)]).

Let \mathcal{H} be a Hardy field, GG a connected and simply connected nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and consider the nilmanifold XkGk/ΓkX^{k}\coloneqq G^{k}/\Gamma^{k}. Suppose

v(n)=(a1f1(n),,akfk(n)),n,v(n)\,=\,\big{(}a_{1}^{f_{1}(n)},\ldots,\hskip 1.00006pta_{k}^{f_{k}(n)}\big{)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G, f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} have different growth222A finite set of functions {f1,,fk}\{f_{1},\ldots,f_{k}\} is said to have different growth if, after potentially reordering, one has f1(t)fk(t)f_{1}(t)\prec\ldots\prec f_{k}(t)., and for any f{f1,,fk}f\in\{f_{1},\ldots,f_{k}\} there exists \ell\in\mathbb{N} such that t1log(t)f(t)tt^{\ell-1}\log(t)\prec f(t)\prec t^{\ell}. Then the sequence (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in the sub-nilmanifold a1Γ¯××akΓ¯\overline{a_{1}^{\mathbb{R}}\Gamma}\times\ldots\times\overline{a_{k}^{\mathbb{R}}\Gamma}.

1.5 has significant implications to additive combinatorics, leading to analogues and refinements of Szemerédi’s theorem on arithmetic progressions (see [FW09, Frantzikinakis10, Frantzikinakis15b]). It was conjectured by Frantzikinakis that the assumption in 1.5 that the functions f1,,fkf_{1},\ldots,f_{k} have different growth can be relaxed considerably (cf. [Frantzikinakis09, Conjecture on p. 357]). This also relates to [Frantzikinakis10, Problems 1], [Frantzikinakis15b, Problem 1], and [Frantzikinakis16, Problem 23]. Our first main result addresses a special case of this conjecture. It gives a generalization of 1.2 to all finite collections f1,,fkf_{1},\ldots,f_{k} of functions from a Hardy field \mathcal{H} satisfying Property (P) below. Throughout this work, we use f(t)f^{\prime}(t), f′′(t)f^{\prime\prime}(t), and f(n)(t)f^{(n)}(t) to denote the 1st1^{\text{st}}, 2nd2^{\text{nd}}, and nthn^{\text{th}} derivative of a function f(t)f(t), respectively.

  1. Property (P):

    For all c1,,ckc_{1},\ldots,c_{k}\in\mathbb{R} and n1,,nk{0}n_{1},\ldots,n_{k}\in\mathbb{N}\cup\{0\} the function f(t)=c1f1(n1)(t)++ckfk(nk)(t)f(t)\,=\,c_{1}f_{1}^{(n_{1})}(t)+\ldots+c_{k}f_{k}^{(n_{k})}(t) has the property that for every p[t]p\in\mathbb{R}[t] either |f(t)p(t)|1|f(t)-p(t)|\ll 1 or |f(t)p(t)|log(t)|f(t)-p(t)|\succ\log(t).

Theorem A.

Let \mathcal{H} be a Hardy field, GG a simply connected nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and assume X=G/ΓX=G/\Gamma is connected. Suppose

v(n)=a1f1(n)akfk(n),n,v(n)\,=\,a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G are commuting, f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} satisfy Property (P), and fi()𝖽𝗈𝗆(ai)f_{i}(\mathbb{N})\subset\mathsf{dom}(a_{i}). Then the following are equivalent:

  1. (i)

    (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in the nilmanifold X=G/ΓX=G/\Gamma.

  2. (ii)

    (ϑ(v(n)Γ))n(\vartheta(v(n)\Gamma))_{n\in\mathbb{N}} is uniformly distributed in the maximal factor torus [G,G]\X[G^{\circ},G^{\circ}]\backslash X.

Whilst A provides us with a convenient criteria for checking whether a sequence of the form na1f1(n)akfk(n)Γn\mapsto a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)}\Gamma is uniformly distributed in the entire nilmanifold, it would also be desirable to have an analogue of 1.4 describing the orbit closure of such sequences in general. The next result provides exactly that.

Theorem B.

Let GG be a simply connected nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and \mathcal{H} a Hardy field. Let

v(n)=a1f1(n)akfk(n),n,v(n)\,=\,a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G are commuting, f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} satisfy Property (P), and fi()𝖽𝗈𝗆(ai)f_{i}(\mathbb{N})\subset\mathsf{dom}(a_{i}) for all i=1,,ki=1,\ldots,k. Then there exists a closed and connected subgroup HH of GG and points x0,x1,,xq1Xx_{0},x_{1},\ldots,x_{q-1}\in X such that, for all r=0,1,,q1r=0,1,\ldots,q-1, the set YrHxrY_{r}\coloneqq Hx_{r} is a closed sub-nilmanifold of XX and (v(qn+r)Γ)n(v(qn+r)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in YrY_{r}.

Note that if f1,,fkf_{1},\ldots,f_{k} are real polynomials then Property (P) is automatically satisfied, which is why Theorems A and B imply Theorems 1.2 and 1.4. More generally, any finite subset of {c1tr1++cmtrm:m,c1,,cm,r1,,rm}\{c_{1}t^{r_{1}}+\ldots+c_{m}t^{r_{m}}:m\in\mathbb{N},c_{1},\ldots,c_{m},r_{1},\ldots,r_{m}\in\mathbb{R}\} satisfies Property (P). This shows that, in contrast to 1.5, Theorems A and B apply to collections of functions that don’t necessarily have different growth. But Theorems A and B do not imply all cases of 1.5, because not every collection of functions f1,,fkf_{1},\ldots,f_{k} satisfying the hypothesis of 1.5 also satisfy Property (P). For instance, fi(t)=tilog(t)f_{i}(t)=t^{i}\log(t) for i{1,,k}i\in\{1,\ldots,k\} is such an example.

We also prove analogues of Theorems A and B that apply to an arbitrary collection of functions f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} of polynomial growth, even when Property (P) is not satisfied. However, in the absence of Property (P) we need to replace Cesàro averages with other, weaker, methods of summation.

Definition 1.6.

Let W:(0,)W\colon\mathbb{N}\to(0,\infty) be a non-decreasing sequence, let w(n)ΔW(n)=W(n+1)W(n)w(n)\coloneqq\Delta W(n)=W(n+1)-W(n) be its discrete derivative, and assume W(n)W(n)\to\infty as nn\to\infty and w(n)1w(n)\ll 1. A sequence (xn)n(x_{n})_{n\in\mathbb{N}} of points in a nilmanifold X=G/ΓX=G/\Gamma is said to be uniformly distributed with respect to WW-averages in XX if

limN1W(N)n=1Nw(n)F(xn)=F𝖽μX\lim_{N\to\infty}\frac{1}{W(N)}\sum_{n=1}^{N}w(n)F(x_{n})=\int F\leavevmode\nobreak\ \mathsf{d}\mu_{X}

holds for every continuous function F𝖢(X)F\in\mathsf{C}(X).

When dealing with functions of “slow growth” from a Hardy field, it turns out to be necessary to switch form Cesàro averages to WW-averages to adequately capture the way in which orbits along such functions distribute. For example, if f(t)f(t) satisfies loglog(t)f(t)log(t)\log\log(t)\prec f(t)\ll\log(t) then the sequence (f(n)mod1)n(f(n)\bmod 1)_{n\in\mathbb{N}} is not uniformly distributed with respect to Cesàro averages in the unit interval [0,1)[0,1), but it is uniformly distributed with respect to logarithmic averages, i.e., WW-averages where W(n)=log(n)W(n)=\log(n) for all nn\in\mathbb{N}. Likewise, if logloglog(t)f(t)loglog(t)\log\log\log(t)\prec f(t)\ll\log\log(t) then the sequence (f(n)mod1)n(f(n)\bmod 1)_{n\in\mathbb{N}} is neither uniformly distributed with respect to Cesàro averages nor uniformly distributed with respect to logarithmic averages, but it is uniformly distributed with respect to double-logarithmic averages, i.e., WW-averages where W(n)=loglog(n)W(n)=\log\log(n) for all nn\in\mathbb{N} (cf. 5.1 below).

Theorems C and D below are generalizations of Theorems A and B that apply to all finite collections of functions f1,,fkf_{1},\ldots,f_{k} of polynomial growth from a Hardy field. However, the increased generality comes at the cost of having to change the method of summation from Cesàro averages to WW-averages.

Theorem C.

Let \mathcal{H} be a Hardy field, GG a simply connected nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and assume X=G/ΓX=G/\Gamma is connected. Suppose

v(n)=a1f1(n)akfk(n),n,v(n)\,=\,a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G are commuting and f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} have polynomial growth and satisfy fi()𝖽𝗈𝗆(ai)f_{i}(\mathbb{N})\subset\mathsf{dom}(a_{i}). Then there exists WW\in\mathcal{H} with 1W(t)t1\prec W(t)\ll t such that the following are equivalent:

  1. (i)

    The sequence (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}} is uniformly distributed with respect to WW-averages in the nilmanifold X=G/ΓX=G/\Gamma.

  2. (ii)

    The sequence (ϑ(v(n)Γ))n(\vartheta(v(n)\Gamma))_{n\in\mathbb{N}} is uniformly distributed with respect to WW-averages in the maximal factor torus [G,G]\X[G^{\circ},G^{\circ}]\backslash X.

Aside from this change in the averaging scheme, C implies 1.5 as well as Frantzinakis’ aforementioned conjecture [Frantzikinakis09, Conjecture on p. 357]. Moreover, C provides the following aesthetic corollary.

Corollary 1.7.

Let \mathcal{H} be a Hardy field, GG a simply connected nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and assume X=G/ΓX=G/\Gamma is connected. Suppose

v(n)=a1f1(n)akfk(n),n,v(n)\,=\,a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G are commuting and f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} have polynomial growth and satisfy fi()𝖽𝗈𝗆(ai)f_{i}(\mathbb{N})\subset\mathsf{dom}(a_{i}). Then the following are equivalent:

  1. (i)

    The sequence (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}} is dense in the nilmanifold X=G/ΓX=G/\Gamma.

  2. (ii)

    The sequence (ϑ(v(n)Γ))n(\vartheta(v(n)\Gamma))_{n\in\mathbb{N}} is dense in the maximal factor torus [G,G]\X[G^{\circ},G^{\circ}]\backslash X.

Finally, let us state our last main result.

Theorem D.

Let GG be a simply connected nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and \mathcal{H} a Hardy field. Suppose

v(n)=a1f1(n)akfk(n),n,v(n)\,=\,a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G are commuting and f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} have polynomial growth and satisfy fi()𝖽𝗈𝗆(ai)f_{i}(\mathbb{N})\subset\mathsf{dom}(a_{i}). Then there exists WW\in\mathcal{H} with 1W(t)t1\prec W(t)\ll t, a closed and connected subgroup HH of GG, and points x0,x1,,xq1Xx_{0},x_{1},\ldots,x_{q-1}\in X such that YrHxrY_{r}\coloneqq Hx_{r} is a closed sub-nilmanifold of XX and (v(qn+r)Γ)n(v(qn+r)\Gamma)_{n\in\mathbb{N}} is uniformly distributed with respect to WW-averages in YrY_{r} for all r=0,1,,q1r=0,1,\ldots,q-1.

The connection between WW and f1,,fkf_{1},\ldots,f_{k} in Theorems C and D is given by a variant of Property (P):

  1. Property (PW):

    For all c1,,ckc_{1},\ldots,c_{k}\in\mathbb{R} and n1,,nk{0}n_{1},\ldots,n_{k}\in\mathbb{N}\cup\{0\} the function f(t)=c1f1(n1)(t)++ckfk(nk)(t)f(t)\,=\,c_{1}f_{1}^{(n_{1})}(t)+\ldots+c_{k}f_{k}^{(n_{k})}(t) has the property that for every p[t]p\in\mathbb{R}[t] either |f(t)p(t)|1|f(t)-p(t)|\ll 1 or |f(t)p(t)|log(W(t))|f(t)-p(t)|\succ\log(W(t)).

As we will show below (see A.5), for an arbitrary finite collection of functions from a Hardy field f1,,fkf_{1},\ldots,f_{k} of polynomial growth there exists some WW with 1W(t)t1\prec W(t)\ll t such that Property (PW) holds. Moreover, it will be clear from the proofs of Theorems C and D that if f1,,fkf_{1},\ldots,f_{k} satisfy Property (PW) for some WW then we can take this WW to be the same as the one appearing in the statements of Theorems C and D.

1.1.   Applications to Combinatorics

The motivation for obtaining Theorems A, B, C, and D is their connection to additive combinatorics. Indeed, these results play a crucial role in a forthcoming paper [BMR20draft], where a far-reaching generalization of Szemerédi’s theorem on arithmetic progressions is explored. To motivate our combinatorial results in this direction, let us first recall the statement of Szemerédi’s Theorem. The upper density of a set EE\subset\mathbb{N} is defined as d¯(E)=lim supN|E{1,,N}|/N\overline{d}(E)=\limsup_{N\to\infty}|E\cap\{1,\ldots,N\}|/N.

Theorem 1.8 (Szemerédi’s Theorem).

For any set EE\subset\mathbb{N} of positive upper density and any kk\in\mathbb{N} there exist a,na,n\in\mathbb{N} such that {a,a+n,,a+(k1)n}E\{a,a+n,\ldots,a+(k-1)n\}\subset E.

Szemerédi’s Theorem has been generalized numerous times and in many different directions. One of the most noteworhty extensions is due to Bergelson and Leibman in [BL96], where a polynomial version was obtained. The following theorem pertains to the one-dimensional case of their result.

Theorem 1.9 (Polynomial Szemerédi Theorem).

For any set EE\subset\mathbb{N} of positive upper density and any polynomials p1,,pk[t]p_{1},\ldots,p_{k}\in\mathbb{Z}[t] satisfying p1(0)==pk(0)=0p_{1}(0)=\ldots=p_{k}(0)=0 there exist a,na,n\in\mathbb{N} such that {a,a+p1(n),,a+pk(n)}E\{a,\,a+p_{1}(n),\ldots,a+p_{k}(n)\}\subset E.

1.9 was later improved in [BLL08] to include an “if and only if” condition.

Theorem 1.10.

Given p1,,pk[t]p_{1},\ldots,p_{k}\in\mathbb{Z}[t] the following are equivalent:

  1. (i)

    The polynomials p1,,pkp_{1},\ldots,p_{k} are jointly intersective, i.e., for any mm\in\mathbb{N} there exists nn\in\mathbb{N} such that pi(n)0modmp_{i}(n)\equiv{0}\bmod{m} for all i{1,,k}i\in\{1,\ldots,k\}.

  2. (ii)

    For any set EE\subset\mathbb{N} of positive upper density there exist a,na,n\in\mathbb{N} such that {a,a+p1(n),,a+pk(n)}E\{a,\,a+p_{1}(n),\ldots,a+p_{k}(n)\}\subset E.

Via the Host-Kra structure theory (see [HK05a, HK18]), Szemerédi’s Theorem and its generalizations are intimately connected to questions about uniform distribution in nilmanifolds. This connection played an important role in the proof of 1.10 in [BLL08], but was also used by Frantzikinakis in [Frantzikinakis15b] (see also [FW09, Frantzikinakis10]) to derive from 1.5 the following combinatorial theorem. Let .:\lfloor.\rfloor\colon\mathbb{R}\to\mathbb{Z} denote the floor function.

Theorem 1.11.

Let f1,,fkf_{1},\ldots,f_{k} be functions from a Hardy field of different growth and with the property that for every f{f1,,fk}f\in\{f_{1},\ldots,f_{k}\} there is \ell\in\mathbb{N} such that t1logtf(t)tt^{\ell-1}\log t\prec f(t)\prec t^{\ell}. Then for any set EE\subset\mathbb{N} of positive upper density there exist a,na,n\in\mathbb{N} such that {a,a+f1(n),,a+fk(n)}E\{a,\,a+\lfloor f_{1}(n)\rfloor,\ldots,a+\lfloor f_{k}(n)\rfloor\}\subset E.

Given a finite collection of functions f1,,fkf_{1},\ldots,f_{k} we denote by span(f1,,fk)\mathrm{span}_{\mathbb{R}}(f_{1},\ldots,f_{k}) the set of all functions of the form f(t)=c1f1(t)++ckfk(t)f(t)=c_{1}f_{1}(t)+\ldots+c_{k}f_{k}(t) for (c1,,ck)k(c_{1},\ldots,c_{k})\in\mathbb{R}^{k}, and by span(f1,,fk)\mathrm{span}_{\mathbb{R}}^{*}(f_{1},\ldots,f_{k}) the set of all functions of the form f(t)=c1f1(t)++ckfk(t)f(t)=c_{1}f_{1}(t)+\ldots+c_{k}f_{k}(t) for (c1,,ck)k\{0}(c_{1},\ldots,c_{k})\in\mathbb{R}^{k}\backslash\{0\}. The following open conjecture is an extension of 1.11 and was posed by Frantzikinakis on multiple occasions.

Conjecture 1.12 (see [Frantzikinakis10, Problems 4 and 4’] and [Frantzikinakis16, Problem 25]).

Let f1,,fkf_{1},\ldots,f_{k} be functions of polynomial growth from a Hardy field such that

|f(t)p(t)||f(t)-p(t)|\to\infty

for every p[t]p\in\mathbb{Z}[t], and every fspan(f1,,fk)f\in\mathrm{span}_{\mathbb{R}}^{*}(f_{1},\ldots,f_{k}). Then for any set EE\subset\mathbb{N} of positive upper density there exist a,na,n\in\mathbb{N} such that {a,a+f1(n),,a+fk(n)}E\{a,a+\lfloor f_{1}(n)\rfloor,\ldots,a+\lfloor f_{k}(n)\rfloor\}\subset E.

Finally, another variant of Szemerédi’s Theorem, which also involves functions from a Hardy field, was obtained in [BMR17arXiv].

Theorem 1.13.

Let ff be a function from a Hardy field and assume there is \ell\in\mathbb{N} such that t1f(t)tt^{\ell-1}\prec f(t)\prec t^{\ell}. Then for any set EE\subset\mathbb{N} of positive upper density there exist a,na,n\in\mathbb{N} such that {a,a+f(n),a+f(n+1),,a+f(n+k)}E\{a,\,a+\lfloor f(n)\rfloor,a+\lfloor f(n+1)\rfloor,\ldots,a+\lfloor f(n+k)\rfloor\}\subset E.

Similar to the proofs of 1.10 and 1.11, the proof of 1.13 also hinges on uniform distribution results in nilmanifolds.

With the help of Theorems C and D, we prove in [BMR20draft] a theorem which not only unifies Theorems 1.81.91.101.111.13, but also confirms 1.12.

Theorem E ([BMR20draft]).

Let [.]:[.]\colon\mathbb{R}\to\mathbb{Z} be the rounding to the closest integer function, let f1,,fkf_{1},\ldots,f_{k} be functions from a Hardy field with polynomial growth, and assume at least one of the following two conditions holds:

  1. (1)

    For all q[t]q\in\mathbb{Z}[t] and fspan(f1,,fk)f\in\mathrm{span}_{\mathbb{R}}^{*}(f_{1},\ldots,f_{k}) we have limt|f(t)q(t)|=\lim_{t\to\infty}|f(t)-q(t)|=\infty.

  2. (2)

    There is jointly intersective collection of polynomials p1,,p[t]p_{1},\ldots,p_{\ell}\in\mathbb{Z}[t] such that any real polynomial “appearing” in span(f1,,fk)\mathrm{span}_{\mathbb{R}}(f_{1},\ldots,f_{k}) also appears in span(p1,,p)\mathrm{span}_{\mathbb{R}}(p_{1},\ldots,p_{\ell}), where we say a polynomial p[t]p\in\mathbb{R}[t] “appears” in span(f1,,fk)\mathrm{span}_{\mathbb{R}}(f_{1},\ldots,f_{k}) if there is fspan(f1,,fk)f\in\mathrm{span}_{\mathbb{R}}(f_{1},\ldots,f_{k}) such that limt|f(t)p(t)|=0\lim_{t\to\infty}|f(t)-p(t)|=0.

Then for any set EE\subset\mathbb{N} of positive upper density there exist a,na,n\in\mathbb{N} such that {a,a+[f1(n)],,a+[fk(n)]}E\{a,\,a+[f_{1}(n)],\ldots,a+[f_{k}(n)]\}\subset E.

Acknowledgements.

The author thanks Vitaly Bergelson, Nikos Frantzikinakis, and Joel Moreira, and the anonymous referee for providing useful comments. The author is supported by the National Science Foundation under grant number DMS 1901453.

2.   Preliminaries

In the proofs of our main theorems we utilize numerous well-known facts and results regarding Hardy fields, nilpotent Lie groups, and nilmanifolds. For convenience, we collect them here in this preparatory section.

2.1.   Preliminaries on Hardy fields

A germ at \infty is any equivalence class of real-valued functions in one real variable under the equivalence relationship (fg)(t0>0such thatf(t)=g(t)for allt[t0,))(f\sim g)\Leftrightarrow\big{(}\exists t_{0}>0\leavevmode\nobreak\ \text{such that}\leavevmode\nobreak\ f(t)=g(t)\leavevmode\nobreak\ \text{for all}\leavevmode\nobreak\ t\in[t_{0},\infty)\big{)}. Let 𝖡\mathsf{B} denote the set of all germs at \infty of real valued functions defined on some half-line [s,)[s,\infty) for some ss\in\mathbb{R}. Note that 𝖡\mathsf{B} forms a ring under pointwise addition and multiplication, which we denote by (𝖡,+,)(\mathsf{B},+,\cdot).

Definition 2.1 (see [Boshernitzan94, Definition 1.2]).

Any subfield of the ring (𝖡,+,)(\mathsf{B},+,\cdot) that is closed under differentiation is called a Hardy field.

By abuse of language, we say that a function f:[s,)f\colon[s,\infty)\to\mathbb{R} belongs to some Hardy field \mathcal{H} (and write ff\in\mathcal{H}) if its germ at \infty belongs to \mathcal{H}.

Functions from a Hardy field have a number of convenient properties. For instance, it was shown in [Boshernitzan81, Proposition 2.1] that for any function ff belonging to a Hardy field \mathcal{H} and any cc\in\mathbb{R} the function f(t)cf(t)-c is either eventually non-negative or eventually non-positive. Combined with the fact that \mathcal{H} is a field and closed under differentiation, this implies that:

  • the limit limtf(t)\lim_{t\to\infty}f(t) always exists as an element in {,}\mathbb{R}\cup\{-\infty,\infty\};

  • ff is either eventually increasing, eventually decreasing, or eventually constant;

  • for any f,gf,g\in\mathcal{H} either f(t)g(t)f(t)\prec g(t), or g(t)f(t)g(t)\prec f(t), or limtf(t)/g(t)\lim_{t\to\infty}f(t)/g(t) is a non-zero real number.

The following lemma was proved in [Frantzikinakis09, Subsection 2.1] using L’Hôpital’s rule.

Lemma 2.2.

Let \mathcal{H} be a Hardy field.

  1. 1.

    If ff\in\mathcal{H} satisfies tkf(t)tkt^{-k}\prec f(t)\prec t^{k} for some kk\in\mathbb{N} and f(t)f(t) is not asymptotically equal to a constant then

    f(t)tlog2(t)f(t)f(t)t.\frac{f(t)}{t\log^{2}(t)}\prec f^{\prime}(t)\ll\frac{f(t)}{t}.
  2. 2.

    If ff\in\mathcal{H} satisfies t1/kf(t)tkt^{1/k}\prec f(t)\prec t^{k} for some kk\in\mathbb{N} then limttf(t)/f(t)\lim_{t\to\infty}tf(t)/f^{\prime}(t) is a non-zero constant.

For more information on Hardy fields we refer the reader to [Boshernitzan81, Boshernitzan82, Boshernitzan84a, Boshernitzan84b, Boshernitzan94, Frantzikinakis09].

2.2.   Preliminaries on nilpotent Lie groups

Let GG be a ss-step nilpotent Lie group with identity element 1G1_{G}. The lower central series of GG, which we denote by C{C1,C2,,Cs,Cs+1}C_{\bullet}\coloneqq\{C_{1},C_{2},\ldots,C_{s},C_{s+1}\}, is a decreasing nested sequence of normal subgroups,

G=C1C2CsCs+1={1G},G=C_{1}\trianglerighteq C_{2}\trianglerighteq\ldots\trianglerighteq C_{s}\trianglerighteq C_{s+1}=\{1_{G}\},

where Ci+1:=[Ci,G]C_{i+1}:=[C_{i},G] is the subgroup of GG generated by all the commutators aba1b1aba^{-1}b^{-1} with aCia\in C_{i} and bGb\in G. Note that Cs+1={1G}C_{s+1}=\{1_{G}\} because GG is ss-step nilpotent. Also, each CiC_{i} is a closed subgroup of GG (cf. [Leibman05a, Section 2.11]).

The upper central series of GG, denoted by Z{Z0,Z1,Z2,,Zs}Z_{\bullet}\coloneqq\{Z_{0},Z_{1},Z_{2},\ldots,Z_{s}\}, is an increasing nested sequence of normal subgroups,

{1G}=Z0Z1Zs1Zs=G,\{1_{G}\}=Z_{0}\trianglelefteq Z_{1}\trianglelefteq\ldots\trianglelefteq Z_{s-1}\trianglelefteq Z_{s}=G,

where the Z0,Z1,,ZsZ_{0},Z_{1},\ldots,Z_{s} are defined inductively by Z0={1G}Z_{0}=\{1_{G}\} and Zi+1={aG:[a,b]Zi for all bG}Z_{i+1}=\{a\in G:[a,b]\in Z_{i}\text{ for all }b\in G\}. Note that Z1Z_{1} is equal to the center Z(G)Z(G) of GG and Zs=GZ_{s}=G because GG is ss-step nilpotent.

Given a uniform and discrete subgroup Γ\Gamma of a nilpotent Lie group GG, an element gGg\in G with the property that gnΓg^{n}\in\Gamma for some nn\in\mathbb{N} is called rational (or rational with respect to Γ\Gamma). A closed subgroup HH of GG is then called rational (or rational with respect to Γ\Gamma) if rational elements are dense in HH. For example, the subgroups C1,,Cs,Cs+1C_{1},\ldots,C_{s},C_{s+1} in the lower central series of GG, as well as Z0,Z1,,ZsZ_{0},Z_{1},\ldots,Z_{s} in the upper central series of GG, are rational with respect to any uniform and discrete subgroup Γ\Gamma of GG (cf. [Raghunathan72, Corollary 1 of Theorem 2.1] for a proof of this fact for connected GG and [Leibman05a, Section 2.11] for the general case).

Rational subgroups play a key role in the description of sub-nilmanifolds. If X=G/ΓX=G/\Gamma is a nilmanifold, then a sub-nilmanifold YY of XX is any closed set of the form Y=HxY=Hx, where xXx\in X and HH is a closed subgroup of GG. It is not true that for every closed subgroup HH of GG and every element x=gΓx=g\Gamma in X=G/ΓX=G/\Gamma the set HxHx is a sub-nilmanifold of XX, because HxHx need not be closed. In fact, it is shown in [Leibman06] that HxHx is closed in XX (and hence a sub-nilmanifold) if and only if the subgroup g1Hgg^{-1}Hg is rational with respect to Γ\Gamma.

For more information on rational elements and rational subgroups see [Leibman06].

2.3.   Preliminaries on the center and central characters

Throughout the paper we use Z(G)Z(G) to denote the center of a group GG.

Lemma 2.3.

Let GG be a nilpotent group and LL a non-trivial normal subgroup of GG. Then LZ(G){1G}L\cap Z(G)\neq\{1_{G}\}.

Proof.

We will make use of the upper central series {1G}=Z0Z1Zs=G,\{1_{G}\}=Z_{0}\trianglelefteq Z_{1}\trianglelefteq\ldots\trianglelefteq Z_{s}=G, which was defined in the previous subsection. Consider the intersection LjLZjL_{j}\coloneqq L\cap Z_{j} for j=0,1,,sj=0,1,\ldots,s. Note that L0={1G}L_{0}=\{1_{G}\} and Ls=LL_{s}=L. Let J={1js:Lj{1G}}J=\{1\leqslant j\leqslant s:L_{j}\neq\{1_{G}\}\} and note that JJ is non-empty because Ls{1G}L_{s}\neq\{1_{G}\}. Let jminj_{\min{}} denote the minimum of JJ. By definition, we have [Zjmin,G]Zjmin1[Z_{j_{\min{}}},G]\subset Z_{j_{\min{}}-1}, which implies [Ljmin,G]Zjmin1[L_{j_{\min{}}},G]\subset Z_{j_{\min{}}-1}. Moreover, LjminL_{j_{\min{}}} is a normal subgroup of GG, because it is the intersection of two normal subgroups of GG, and hence [Ljmin,G]Ljmin[L_{j_{\min{}}},G]\subset L_{j_{\min{}}}. We conclude that

[Ljmin,G]LjminZjmin1=Ljmin1.[L_{j_{\min{}}},G]\subset L_{j_{\min{}}}\cap Z_{j_{\min{}}-1}=L_{j_{\min{}}-1}.

In view of the minimality assumption on jminj_{\min{}} we have Ljmin1={1G}L_{j_{\min{}}-1}=\{1_{G}\}, from which it follows that [Ljmin,G]={1G}[L_{j_{\min{}}},G]=\{1_{G}\}. This proves that LjminL_{j_{\min{}}} is a subset of Z(G)Z(G). Thus LZ(G)=L1=Ljmin{1G}L\cap Z(G)=L_{1}=L_{j_{\min{}}}\neq\{1_{G}\} as desired. ∎

Corollary 2.4.

Let GG be a simply-connected nilpotent Lie group and LL a non-trivial, connected, and normal subgroup of GG. Then LZ(G){1G}L\cap Z(G)^{\circ}\neq\{1_{G}\}.

Proof.

According to 2.3 there exists an element a1Ga\neq 1_{G} in the intersection LZ(G)L\cap Z(G). Since LL is connected and aLa\in L, the element aa belongs to the identity component GG^{\circ} of GG. Moreover, since GG is simply-connected, the 11-parameter subgroup a={at:t}a^{\mathbb{R}}=\{a^{t}:t\in\mathbb{R}\} is well defined. It follows from [Malcev49, Lemma 3] (and, alternatively, also from the Baker-Campbell-Hausdorff formula) that if aa commutes with an element bGb\in G then the entire 11-parameter subgroup aa^{\mathbb{R}} commutes with bb. This implies that if aa belongs to the center of GG, then so does aa^{\mathbb{R}}. In particular, aZ(G)a^{\mathbb{R}}\subset Z(G), which proves aZ(G)a\in Z(G)^{\circ} and hence LZ(G){1G}L\cap Z(G)^{\circ}\neq\{1_{G}\}. ∎

Definition 2.5 (Central characters).

Let GG be a nilpotent Lie group and Γ\Gamma a uniform and discrete subgroup of GG. A central character of (G,Γ)(G,\Gamma) is any continuous map φ:X\varphi\colon X\to\mathbb{C} with the property that there exists a continuous group homomorphism χ:Z(G){z:|z|=1}\chi\colon Z(G)\to\{z\in\mathbb{C}:|z|=1\} such that

φ(sx)=χ(s)φ(x),sZ(G),xX.\varphi(sx)=\chi(s)\varphi(x),\qquad\forall s\in Z(G),\leavevmode\nobreak\ \forall x\in X. (2.1)
Remark 2.6.

Since the set of all central characters is closed under conjugation and separates points333We claim that for any two distinct points x,yX=G/Γx,y\in X=G/\Gamma there exists a central character φ\varphi such that φ(x)φ(y)\varphi(x)\neq\varphi(y). To verify this claim, we distinguish two cases, the case y{sx:sZ(G)}y\notin\{sx:s\in Z(G)\} and the case y{sx:sZ(G)}y\in\{sx:s\in Z(G)\}. If we are in the first case then Z(G)xZ(G)yZ(G)x\neq Z(G)y. This implies there exists a continuous function φ\varphi^{\prime} on the quotient space Z(G)\XZ(G)\backslash X with φ(Z(G)x)φ(Z(G)y)\varphi^{\prime}(Z(G)x)\neq\varphi^{\prime}(Z(G)y), which we can lift to a continuous and Z(G)Z(G)-invariant function φ\varphi on XX satisfying φ(x)φ(y)\varphi(x)\neq\varphi(y). If we are in the second case then there is s0Z(G)s_{0}\in Z(G) such that y=s0xy=s_{0}x. Note that s0s_{0} cannot be an element of Γ\Gamma because xyx\neq y. Let χ0\chi_{0} be any group character of Z(G)Z(G) with the property that Z(G)ΓkerχZ(G)\cap\Gamma\subset\ker\chi and χ0(s0)1\chi_{0}(s_{0})\neq 1. Define TZ(G)/(ΓZ(G))T\coloneqq Z(G)/(\Gamma\cap Z(G)) and observe that TT is a compact abelian group. Let UU be a small neighborhood of xx, and let ρ:X[0,1]\rho\colon X\to[0,1] be a continuous function with the property that ρ(x)=1\rho(x)=1 and ρ(z)=0\rho(z)=0 for all zUz\notin U. Now define φ(z)ρ(sz)χ0(s)𝖽μT(s)\varphi(z)\coloneqq\int\rho(sz)\chi_{0}(s)\leavevmode\nobreak\ \mathsf{d}\mu_{T}(s), where μT\mu_{T} is the normalized Haar measure on TT. It is straightforward to check that φ\varphi is a non-zero and continuous function on XX satisfying φ(sz)=χ0(s)φ(z)\varphi(sz)=\chi_{0}(s)\varphi(z) for all sZ(G)s\in Z(G) and zXz\in X. In particular φ(y)=φ(s0x)=χ0(s0)φ(x)\varphi(y)=\varphi(s_{0}x)=\chi_{0}(s_{0})\varphi(x) and so φ(y)φ(x)\varphi(y)\neq\varphi(x). in XX, it follows from the Stone-Weierstrass Theorem that their linear span is uniformly dense in 𝖢(X)\mathsf{C}(X). We will make use of this fact multiple times in the upcoming sections.

Remark 2.7.

Let φ\varphi be a central character of (G,Γ)(G,\Gamma) and let χ:Z(G){z:|z|=1}\chi\colon Z(G)\to\{z\in\mathbb{C}:|z|=1\} be the corresponding continuous group homomorphism such that (2.1) is satisfied. We claim that if χ\chi is non-trivial (meaning that there exists sZ(G)s\in Z(G) such that χ(s)1\chi(s)\neq 1) then the integral φ𝖽μX\int\varphi\leavevmode\nobreak\ \mathsf{d}\mu_{X} equals 0. To verify this claim, note that the measure μX\mu_{X} is invariant under left-multiplication by ss. Therefore,

φ(x)𝖽μX(x)=φ(sx)𝖽μX(x)=χ(s)φ(x)𝖽μX(x),\int\varphi(x)\leavevmode\nobreak\ \mathsf{d}\mu_{X}(x)\,=\,\int\varphi(sx)\leavevmode\nobreak\ \mathsf{d}\mu_{X}(x)\,=\,\chi(s)\int\varphi(x)\leavevmode\nobreak\ \mathsf{d}\mu_{X}(x),

which can only hold if φ𝖽μX=0\int\varphi\leavevmode\nobreak\ \mathsf{d}\mu_{X}=0.

2.4.   Preliminaries on relatively independent self-products

One of the key ideas featured in the proofs of Theorems A, B, C, and D is the utilization of a special type of “relative product group”. Similar product groups played an important role in the inductive procedure employed by Green and Tao in [GT12a].

Definition 2.8 (Relatively independent product).

Let GG be a group and LL a normal subgroup of GG. We define the relatively independent self-product of GG over LL as

G×LG{(a1,a2)G×G:a1a21L}.G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G\coloneqq\{(a_{1},a_{2})\in G{\mkern-1.0mu\times\mkern-0.5mu}G:a_{1}a_{2}^{-1}\in L\}.

If GG is a nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and LL a normal and rational subgroup of GG then the group

Γ×LΓ{(γ1,γ2)Γ×Γ:γ1γ21L}=(Γ×Γ)(G×LG)\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma\coloneqq\{(\gamma_{1},\gamma_{2})\in\Gamma{\mkern-1.0mu\times\mkern-0.5mu}\Gamma:\gamma_{1}\gamma_{2}^{-1}\in L\}=(\Gamma{\mkern-1.0mu\times\mkern-0.5mu}\Gamma)\cap(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)

is a uniform and discrete subgroup of G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G. This gives rise to the nilmanifold

X×LX(G×LG)/(Γ×LΓ),X{\mkern-1.0mu\times\mkern-1.5mu}_{L}X\coloneqq(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)/(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma),

which we call the relatively independent self-product of XX over LL.

Remark 2.9.

In ergodic theory, the notion of a relatively independent self-joining of a system XX over one of its factors YY is an important notion and finds many applications (see [EW11, Definition 6.15] for the definition). If X=G/ΓX=G/\Gamma is a nilmanifold and LL is a normal and rational subgroup of GG then the quotient space Y=L\XY=L\backslash X is a factor of XX. It turns out that that the relatively independent self-product X×LXX{\mkern-1.0mu\times\mkern-1.5mu}_{L}X is exactly the same as the relatively independent self-joining of XX over the factor YY.

Next, let us state and prove a few results regarding relatively independent self-products that will be useful in the later sections.

Lemma 2.10.

We have [G×LG,G×LG]=[G,G]×[G,L][G,G]\big{[}G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G\big{]}=[G,G]{\mkern-1.0mu\times\mkern-1.5mu}_{[G,L]}[G,G].

Proof.

Note that [G,L][G,L] is a normal subgroup of GG and a subset of LL, because LL is a normal subgroup of GG. It follows that [G,L]×[G,L][G,L]{\mkern-1.0mu\times\mkern-0.5mu}[G,L] is a normal subgroup of G×GG{\mkern-1.0mu\times\mkern-0.5mu}G. Moreover, [G,L]×[G,L][G,L]{\mkern-1.0mu\times\mkern-0.5mu}[G,L] is a subset of [G×LG,G×LG]\big{[}G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G\big{]} because [G,L]L[G,L]\subset L, and [G,L]×[G,L][G,L]{\mkern-1.0mu\times\mkern-0.5mu}[G,L] is a subset of [G,G]×[G,L][G,G][G,G]{\mkern-1.0mu\times\mkern-1.5mu}_{[G,L]}[G,G] because [G,L][G,G][G,L]\subset[G,G]. Thus, to show

[G×LG,G×LG]=[G,G]×[G,L][G,G],\big{[}G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G\big{]}=[G,G]{\mkern-1.0mu\times\mkern-1.5mu}_{[G,L]}[G,G],

it suffices to show

[G×LG,G×LG]mod[G,L]×[G,L]=[G,G]×[G,L][G,G]mod[G,L]×[G,L].\big{[}G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G\big{]}\bmod[G,L]{\mkern-1.0mu\times\mkern-0.5mu}[G,L]=[G,G]{\mkern-1.0mu\times\mkern-1.5mu}_{[G,L]}[G,G]\bmod[G,L]{\mkern-1.0mu\times\mkern-0.5mu}[G,L]. (2.2)

Note that [G×LG,G×LG]\big{[}G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G\big{]} is generated by elements of the form ([g1,g2],[g1l1,g2,l2])([g_{1},g_{2}],[g_{1}l_{1},g_{2},l_{2}]) for g1,g2Gg_{1},g_{2}\in G and l1,l2Ll_{1},l_{2}\in L. Since elements in GG commute with elements in LL modulo [G,L][G,L], we have ([g1,g2],[g1l1,g2,l2])([g1,g2],[g1,g2])mod[G,L]×[G,L]([g_{1},g_{2}],[g_{1}l_{1},g_{2},l_{2}])\equiv([g_{1},g_{2}],[g_{1},g_{2}])\bmod[G,L]{\mkern-1.0mu\times\mkern-0.5mu}[G,L]. In other words,

[G×LG,G×LG]mod[G,L]×[G,L]={(g,g):g[G,G]}mod[G,L]×[G,L].\big{[}G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G\big{]}\bmod[G,L]{\mkern-1.0mu\times\mkern-0.5mu}[G,L]=\{(g,g):g\in[G,G]\}\bmod[G,L]{\mkern-1.0mu\times\mkern-0.5mu}[G,L].

Similarly, since [G,G]×[G,L][G,G]={(g,gh):g[G,G],h[G,L]}[G,G]{\mkern-1.0mu\times\mkern-1.5mu}_{[G,L]}[G,G]=\{(g,gh):g\in[G,G],\,h\in[G,L]\}, we have

[G,G]×[G,L][G,G]={(g,g):g[G,G]}mod[G,L]×[G,L].[G,G]{\mkern-1.0mu\times\mkern-1.5mu}_{[G,L]}[G,G]=\{(g,g):g\in[G,G]\}\bmod[G,L]{\mkern-1.0mu\times\mkern-0.5mu}[G,L].

This finishes the proof of (2.2). ∎

From 2.10 we can derive the following corollary.

Corollary 2.11.

G/[G,G]×L/[G,L]G/[G,G]{\mkern-1.0mu\times\mkern-0.5mu}L/[G,L] and G×LG/[G×LG,G×LG]G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G/[G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G] are isomorphic as nilpotent Lie groups.

Proof.

Consider the map Φ:G×LG×LG/[G×LG,G×LG]\Phi\colon G{\mkern-1.0mu\times\mkern-0.5mu}L\to G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G/[G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G] defined as

Φ(a,b)=(a,ba)[G×LG,G×LG],aG,bL.\Phi(a,b)=(a,ba)[G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G],\qquad\forall a\in G,\leavevmode\nobreak\ \forall b\in L.

Clearly, Φ\Phi is well defined, smooth, surjective, and a homomorphism. Moreover, it follows from 2.10 that the kernel of Φ\Phi equals [G,G]×[G,L][G,G]{\mkern-1.0mu\times\mkern-0.5mu}[G,L]. Indeed, (a,b)(a,b) belongs to the kernel of Φ\Phi if and only if (a,ab)(a,ab) belongs to [G×LG,G×LG][G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G]. According to 2.10, this happens exactly when (a,ab)(a,ab) belongs to [G,G]×[G,L][G,G][G,G]{\mkern-1.0mu\times\mkern-1.5mu}_{[G,L]}[G,G]. Using the definition of relative independent product groups, we see that (a,ab)[G,G]×[G,L][G,G](a,ab)\in[G,G]{\mkern-1.0mu\times\mkern-1.5mu}_{[G,L]}[G,G] if and only if a[G,G]a\in[G,G] and b[G,L]b\in[G,L] as claimed. ∎

2.11 helps us better understand the maximal factor torus of the relatively independent self-product X×LXX{\mkern-1.0mu\times\mkern-1.5mu}_{L}X, which will turn out to be an important aspect in the proofs of our main results. First, let us introduce the notion of a horizontal character.

Definition 2.12 (Horizontal characters, cf. [GT12a]).

Let GG be a nilpotent Lie group and Γ\Gamma a uniform and discrete subgroup of GG. A horizontal character of (G,Γ)(G,\Gamma) is any continuous map η:X\{0}\eta\colon X\to\mathbb{C}\backslash\{0\} satisfying

η(abΓ)=η(aΓ)η(bΓ),a,bG.\eta(ab\Gamma)\,=\,\eta(a\Gamma)\eta(b\Gamma),\qquad\forall a,b\in G. (2.3)

We say η\eta is non-trivial if η\eta is not constant equal to 11.

Remark 2.13.

With the help of 2.11 it is easy to describe the horizontal characters of the relatively independent product (G×LG,Γ×LΓ)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma) in terms of the horizontal characters of (G,Γ)(G,\Gamma) and (L,ΓL)(L,\Gamma_{L}), where ΓLΓL\Gamma_{L}\coloneqq\Gamma\cap L. Indeed, for any horizontal character η\eta of (G×LG,Γ×LΓ)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma) there exists a horizontal character η1\eta_{1} of (G,Γ)(G,\Gamma) and a horizontal character η2\eta_{2} of (L,ΓL)(L,\Gamma_{L}) with [G,L]kerη2[G,L]\subset\ker\eta_{2} such that

η((a1,a2)Γ×LΓ)=η1(a2Γ)η2(a1a21ΓL),(a1,a2)G×LG.\eta\big{(}(a_{1},a_{2})\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma\big{)}=\eta_{1}\big{(}a_{2}\Gamma\big{)}\eta_{2}\big{(}a_{1}a_{2}^{-1}\Gamma_{L}\big{)},\qquad\forall(a_{1},a_{2})\in G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G.

Observe that if GG is connected then horizontal characters descend to the maximal factor torus, where they generate an algebra that is uniformly dense due to the Stone-Weierstrass Theorem. Therefore, 2.13 helps us understand the maximal factor torus of the relatively independent self-product X×LXX{\mkern-1.0mu\times\mkern-1.5mu}_{L}X in the case when GG is connected. However, we also need to better understand the maximal factor torus if GG is not connected. First, let us characterize the identity component of G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G.

Lemma 2.14.

We have (G×LG)=G×LG(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)^{\circ}=G^{\circ}{\mkern-1.0mu\times\mkern-1.5mu}_{L^{\circ}}G^{\circ}.

Proof.

If (Gi)i(G_{i})_{i\in\mathcal{I}} are the connected components of GG and (Lj)j𝒥(L_{j})_{j\in\mathcal{J}} are the connected components of LL then the connected components of G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G are (Gi×LjGk)i,k,j𝒥(G_{i}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{j}}G_{k})_{i,k\in\mathcal{I},j\in\mathcal{J}}, where

Gi×LjGk{(a1,a2)Gi×Gk:a1a21Lj}.G_{i}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{j}}G_{k}\coloneqq\{(a_{1},a_{2})\in G_{i}\times G_{k}:a_{1}a_{2}^{-1}\in L_{j}\}.

This is because Gi×LjGkG_{i}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{j}}G_{k} are open and connected subsets of G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G, and if (i,k,j)(i,k,j)(i,k,j)\neq(i^{\prime},k^{\prime},j^{\prime}) then Gi×LjGkG_{i}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{j}}G_{k} and Gi×LjGkG_{i^{\prime}}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{j^{\prime}}}G_{k^{\prime}} are disjoint. It follows that the connected component of G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G that contains the identity is G×LGG^{\circ}{\mkern-1.0mu\times\mkern-1.5mu}_{L^{\circ}}G^{\circ}. ∎

To study the maximal factor torus for non-connected GG, we utilize a variant of the notion of a horizontal character.

Definition 2.15 (Pseudo-horizontal characters).

Let GG be a nilpotent Lie group and Γ\Gamma a uniform and discrete subgroup of GG. A pseudo-horizontal character of (G,Γ)(G,\Gamma) is any continuous map η:X\{0}\eta\colon X\to\mathbb{C}\backslash\{0\} satisfying

η(abΓ)=η(aΓ)η(bΓ),aG,bG.\eta(ab\Gamma)\,=\,\eta(a\Gamma)\eta(b\Gamma),\qquad\forall a\in G^{\circ},\leavevmode\nobreak\ \forall b\in G. (2.4)

Note that if GG is connected then pseudo-horizontal characters are the same as horizontal characters. Also, pseudo-horizontal characters descend to the maximal factor torus, where they generated a dense algebra, provided that G/ΓG/\Gamma is connected. Indeed, if G/ΓG/\Gamma is connected then GG^{\circ} acts transitively on G/ΓG/\Gamma and hence G/[G,G]G^{\circ}/[G^{\circ},G^{\circ}] acts transitively on the maximal factor torus. This implies that the maximal factor torus is isomorphic to the compact abelian group G/[G,G](ΓG)G^{\circ}/[G^{\circ},G^{\circ}](\Gamma\cap G^{\circ}) and the pseudo-horizontal characters descend to the group characters of this group.

Remark 2.16.

By combining 2.11 with 2.14 we are now able to describe the pseudo-horizontal characters of (G×LG,Γ×LΓ)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma) in the case when (G×LG)/(Γ×LΓ)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)/(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma) is connected, similar to the way we described the horizontal characters of (G×LG,Γ×LΓ)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma) in 2.13 above. If (G×LG)/(Γ×LΓ)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)/(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma) is connected then it is isomorphic to

(G×LG)/((G×LG)(Γ×LΓ))(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)^{\circ}/((G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)^{\circ}\cap(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma))

and hence the pseudo-horizontal characters of G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G can be identified with the horizontal characters of (G×LG)=G×LG(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)^{\circ}=G^{\circ}{\mkern-1.0mu\times\mkern-1.5mu}_{L^{\circ}}G^{\circ}. It follows that for any pseudo-horizontal character η\eta of (G×LG,Γ×LΓ)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma) there exists a pseudo-horizontal character η1\eta_{1} of (G,Γ)(G,\Gamma) and a pseudo-horizontal character η2\eta_{2} of (L,ΓL)(L,\Gamma_{L}) with [G,L]kerη2[G^{\circ},L^{\circ}]\subset\ker\eta_{2} such that

η((a1,a2)Γ×LΓ)=η1(a2Γ)η2(a1a21ΓL),(a1,a2)G×LG.\eta\big{(}(a_{1},a_{2})\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma\big{)}=\eta_{1}\big{(}a_{2}\Gamma\big{)}\eta_{2}\big{(}a_{1}a_{2}^{-1}\Gamma_{L}\big{)},\qquad\forall(a_{1},a_{2})\in G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G.

Finally, here is another lemma that we will invoke numerous times in the later sections.

Lemma 2.17.

Let GG be a nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and LL a normal and rational subgroup of GG. Let π:GX\pi\colon G\to X denote the natural projection of GG onto the nilmanifold XG/ΓX\coloneqq G/\Gamma. If both XX and the sub-nilmanifold π(L)\pi(L) are connected then the relatively independent self-product X×LXX{\mkern-1.0mu\times\mkern-1.5mu}_{L}X is also connected.

Proof.

Note that XX is connected if and only if G=GΓG=G^{\circ}\Gamma. Likewise, π(L)\pi(L) is connected if and only if LΓ=LΓ=ΓLL^{\circ}\Gamma=L\Gamma=\Gamma L. In particular, any aGa\in G can be written as bγb\gamma with bGb\in G^{\circ} and γG\gamma\in G, and any sLs\in L can be written as κt\kappa t with tLt\in L^{\circ} and κLΓ\kappa\in L\cap\Gamma. It follows that

G×LG\displaystyle G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G ={(a,as):aG,sL}\displaystyle=\{(a,as):a\in G,\leavevmode\nobreak\ s\in L\}
={(bγ,bγκt):bG,tL,γΓ,κΓL}.\displaystyle=\{(b\gamma,b\gamma\kappa t):b\in G^{\circ},\leavevmode\nobreak\ t\in L^{\circ},\leavevmode\nobreak\ \gamma\in\Gamma,\leavevmode\nobreak\ \kappa\in\Gamma\cap L\}.

Since LL^{\circ} is normal, we have bγκt=t~bγκb\gamma\kappa t=\tilde{t}b\gamma\kappa where t~=(bγκ)t(bγκ)1L\tilde{t}=(b\gamma\kappa)t(b\gamma\kappa)^{-1}\in L^{\circ}. Hence

G×LG\displaystyle G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G ={(bγ,t~bγκ):bG,t~L,γΓ,κΓL}\displaystyle=\{(b\gamma,\tilde{t}b\gamma\kappa):b\in G^{\circ},\leavevmode\nobreak\ \tilde{t}\in L^{\circ},\leavevmode\nobreak\ \gamma\in\Gamma,\leavevmode\nobreak\ \kappa\in\Gamma\cap L\}
={(b,t~b):bG,t~L}{(γ,γκ):γΓ,κΓL}\displaystyle=\{(b,\tilde{t}b):b\in G^{\circ},\leavevmode\nobreak\ \tilde{t}\in L^{\circ}\}\leavevmode\nobreak\ \{(\gamma,\gamma\kappa):\gamma\in\Gamma,\leavevmode\nobreak\ \kappa\in\Gamma\cap L\}
=(G×LG)(Γ×LΓ),\displaystyle=(G^{\circ}{\mkern-1.0mu\times\mkern-1.5mu}_{L^{\circ}}G^{\circ})(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma),

which implies X×LXX{\mkern-1.0mu\times\mkern-1.5mu}_{L}X is connected. ∎

3.   Reducing Theorems A and B to F

Our first step in proving Theorems A and B is to reduce them to the following result.

Theorem F.

Let GG be a simply connected nilpotent Lie group and Γ\Gamma a uniform and discrete subgroup of GG. Assume v:Gv\colon\mathbb{N}\to G is a mapping of the form

v(n)=a1f1(n)akfk(n)b1p1(n)bmpm(n),n,v(n)\,=\,a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)}b_{1}^{p_{1}(n)}\cdot\ldots\cdot b_{m}^{p_{m}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G^{\circ}, b1,,bmGb_{1},\ldots,b_{m}\in G, the elements a1,,ak,b1,,bma_{1},\ldots,a_{k},b_{1},\ldots,b_{m} are pairwise commuting, b1Γ¯,,bmΓ¯\overline{b_{1}^{\mathbb{Z}}\Gamma},\ldots,\overline{b_{m}^{\mathbb{Z}}\Gamma} are connected sub-nilmanifolds of X=G/ΓX=G/\Gamma, p1,,pm[t]p_{1},\ldots,p_{m}\in\mathbb{R}[t] are polynomials satisfying

  1.  (F1)

    pj()p_{j}(\mathbb{Z})\subset\mathbb{Z}, for all j=1,,mj=1,\ldots,m,

  2.  (F2)

    deg(pj)=j\deg(p_{j})=j for all j=1,,mj=1,\ldots,m,

and f1,,fkf_{1},\ldots,f_{k} are functions belonging to some Hardy field \mathcal{H} satisfying

  1.  (F3)

    f1(t)fk(t)f_{1}(t)\prec\ldots\prec f_{k}(t),

  2.  (F4)

    for all f{f1,,fk}f\in\{f_{1},\ldots,f_{k}\} there exists \ell\in\mathbb{N} such that t1log(t)f(t)tt^{\ell-1}\log(t)\prec f(t)\prec t^{\ell},

  3.  (F5)

    for all f{f1,,fk}f\in\{f_{1},\ldots,f_{k}\} with deg(f)2\deg(f)\geqslant 2 we have f{f1,,fk}f^{\prime}\in\{f_{1},\ldots,f_{k}\}.

Then (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in the sub-nilmanifold a1akb1bmΓ¯\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma}.

We remark that the set a1akb1bmΓ¯\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma} appearing in the formulation of F above is indeed a sub-nilmanifold of XX, which follows from [BMR17arXiv, Lemma A.6].

Remark 3.1.

Using the “change of base-point” trick (cf. [Frantzikinakis09, p. 368]), it is straightforward to see that if the sequence a1f1(n)akfk(n)b1p1(n)bmpm(n)Γa_{1}^{f_{1}(n)}\cdots a_{k}^{f_{k}(n)}b_{1}^{p_{1}(n)}\cdots b_{m}^{p_{m}(n)}\Gamma is uniformly distributed in a1akb1bmΓ¯\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma} then for every cGc\in G the sequence a1f1(n)akfk(n)b1p1(n)bmpm(n)cΓa_{1}^{f_{1}(n)}\cdots a_{k}^{f_{k}(n)}b_{1}^{p_{1}(n)}\cdots b_{m}^{p_{m}(n)}c\Gamma is uniformly distributed in a1akb1bmcΓ¯\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}c\Gamma}.

The proof of F is given in Section 4. The remainder of this section is dedicated to showing that F implies Theorems A and B. For this, we will make use of A.2 and A.6, which are formulated and proved in the appendix.

Proof that F implies Theorems A and B.

Let GG be a simply connected nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and XX the nilmanifold G/ΓG/\Gamma. Let a1,,akGa_{1},\ldots,a_{k}\in G be pairwise commuting, f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} satisfy Property (P), suppose fi()𝖽𝗈𝗆(ai)f_{i}(\mathbb{N})\subset\mathsf{dom}(a_{i}) for all i=1,,ki=1,\ldots,k, and consider the sequence

v(n)a1f1(n)akfk(n),n.v(n)\coloneqq a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)},\qquad\forall n\in\mathbb{N}.

To begin with, we divide the set {1,,k}\{1,\ldots,k\} into two pieces. The first piece, which we denote by \mathcal{I}, consists of all i{1,,k}i\in\{1,\ldots,k\} for which aia_{i} belongs to the identity component GG^{\circ}. The second piece c\mathcal{I}^{c} is defined as {1,,k}\\{1,\ldots,k\}\backslash\mathcal{I}. According to A.2, for any ici\in\mathcal{I}^{c} there exist mim_{i}\in\mathbb{N} and pi[t]p_{i}\in\mathbb{R}[t] such that pi()p_{i}(\mathbb{Z})\subset\mathbb{Z}, 1/mi𝖽𝗈𝗆(ai){1}/{m_{i}}\in\mathsf{dom}(a_{i}), and

fi(n)=pi(n)/mi,n.f_{i}(n)=p_{i}(n)/m_{i},\qquad\forall n\in\mathbb{N}.

Then, by A.6 applied with V(t)=tV(t)=t to the family of functions {fi:i}\{f_{i}:i\in\mathcal{I}\}, we can find mm\in\mathbb{N}, functions g1,,gmg_{1},\ldots,g_{m}\in\mathcal{H}, a set of polynomials {pi:i}[t]\{p_{i}:i\in\mathcal{I}\}\subset\mathbb{R}[t], and coefficients {λi,1,,λi,m:i}\{\lambda_{i,1},\ldots,\lambda_{i,m}:i\in\mathcal{I}\}\subset\mathbb{R} such that the following properties hold:

  1. (I)

    g1(t)gm(t)g_{1}(t)\prec\ldots\prec g_{m}(t);

  2. (II)

    for all g{g1,,gm}g\in\{g_{1},\ldots,g_{m}\} either g=0g=0 or there exists \ell\in\mathbb{N} such that t1log(t)g(t)tt^{\ell-1}\log(t)\prec g(t)\prec t^{\ell};

  3. (III)

    for all g{g1,,gm}g\in\{g_{1},\ldots,g_{m}\} with deg(g)2\deg(g)\geqslant 2 we have g{g1,,gm}g^{\prime}\in\{g_{1},\ldots,g_{m}\};

  4. (IV)

    for all ii\in\mathcal{I},

    limt|fi(t)l=1mλi,lgl(t)pi(t)|=0.\lim_{t\to\infty}\Bigg{|}f_{i}(t)-\sum_{l=1}^{m}\lambda_{i,l}g_{l}(t)-p_{i}(t)\Bigg{|}=0.

For ii\in\mathcal{I} the polynomial pip_{i} can be written as

pi(n)=ci,0(n0)+ci,1(n1)+ci,2(n2)++ci,M(nM)p_{i}(n)\,=\,c_{i,0}\binom{n}{0}+c_{i,1}\binom{n}{1}+c_{i,2}\binom{n}{2}+\ldots+c_{i,M}\binom{n}{M}

for some real coefficients ci,0,,ci,Mc_{i,0},\ldots,c_{i,M}. For ii\notin\mathcal{I} the polynomial pip_{i} can also be written as

pi(n)=ci,0(n0)+ci,1(n1)+ci,2(n2)++ci,M(nM)p_{i}(n)\,=\,c_{i,0}\binom{n}{0}+c_{i,1}\binom{n}{1}+c_{i,2}\binom{n}{2}+\ldots+c_{i,M}\binom{n}{M}

but with an additional feature. It is a standard fact from algebra that any polynomial which takes integer values on the integers can we expressed as an integer linear combination of binomial coefficients. Therefore, if ici\in\mathcal{I}^{c} then the coefficients ci,0,,ci,Mc_{i,0},\ldots,c_{i,M} are actually integers.

Next, for l={1,,m}l=\{1,\ldots,m\}, define

uliaiλi,lu_{l}\,\coloneqq\,\prod_{i\in\mathcal{I}}a_{i}^{\lambda_{i,l}}

and for j{0,1,,M}j\in\{0,1,\ldots,M\} define

ej(iaici,j)(icaici,j/mi).e_{j}\,\coloneqq\,\left(\prod_{i\in\mathcal{I}}a_{i}^{c_{i,j}}\right)\cdot\left(\prod_{i\in\mathcal{I}^{c}}a_{i}^{c_{i,j}/m_{i}}\right).

Note that for all ii\in\mathcal{I} the elements aiλi,ja_{i}^{\lambda_{i,j}} and aici,ja_{i}^{c_{i,j}} are well defined because aiGa_{i}\in G^{\circ}, and for all ici\in\mathcal{I}^{c} the elements aici,j/mia_{i}^{c_{i,j}/m_{i}} are well defined because 1/mi𝖽𝗈𝗆(ai){1}/{m_{i}}\in\mathsf{dom}(a_{i}) and ci,jc_{i,j}\in\mathbb{Z}. It follows that the elements u1,,umu_{1},\ldots,u_{m} belong to GG^{\circ} and u1,,um,e0,e1,,eMu_{1},\ldots,u_{m},e_{0},e_{1},\ldots,e_{M} are pairwise commuting.

Let dGd_{G} be any right-invariant metric on the nilpotent Lie group GG. Using property (IV), it is straightforward to check that

limndG(a1f1(n)akfk(n),u1g1(n)umgm(n)e0(n0)e1(n1)eM(nM))= 0.\lim_{n\to\infty}\leavevmode\nobreak\ d_{G}\left(a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)},\leavevmode\nobreak\ u_{1}^{g_{1}(n)}\cdot\ldots\cdot u_{m}^{g_{m}(n)}e_{0}^{\binom{n}{0}}e_{1}^{\binom{n}{1}}\cdot\ldots\cdot e_{M}^{\binom{n}{M}}\right)\,=\,0.

Therefore, instead of showing the conclusions of Theorems A and B for (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}}, it suffices to show the conclusions of Theorems A and B for the sequence (w(n)Γ)n(w(n)\Gamma)_{n\in\mathbb{N}}, where

w(n)u1g1(n)umgm(n)e1(n1)eM(nM),n.w(n)\,\coloneqq\,u_{1}^{g_{1}(n)}\cdot\ldots\cdot u_{m}^{g_{m}(n)}e_{1}^{\binom{n}{1}}\cdot\ldots\cdot e_{M}^{\binom{n}{M}},\qquad n\in\mathbb{N}.

Note that e0(n0)e_{0}^{\binom{n}{0}} was omitted because it is a constant (cf. 3.1).

It is a consequence of 1.4 that for every j{1,,M}j\in\{1,\ldots,M\} there exists qjq_{j}\in\mathbb{N} such that for all r{0,1,,qj1}r\in\{0,1,\ldots,q_{j}-1\} the set

ejqj+rΓ¯={ejqjn+rΓ:n}¯\overline{e_{j}^{q_{j}\mathbb{Z}+r}\Gamma}=\overline{\{e_{j}^{q_{j}n+r}\Gamma:n\in\mathbb{Z}\}}

is a connected sub-nilmanifold of XX. If we take q=lcm(q1,,qM)q=\mathrm{lcm}(q_{1},\ldots,q_{M}) then

ejq+rΓ¯\overline{e_{j}^{q\mathbb{Z}+r}\Gamma}

is also connected for all r{0,1,,q1}r\in\{0,1,\ldots,q-1\} and all j{1,,M}j\in\{1,\ldots,M\}. Let pj,r(n)p_{j,r}(n) be the polynomial defined as

pj,r(n)1q((qn+rj)(rj)),p_{j,r}(n)\coloneqq\frac{1}{q}\left(\binom{qn+r}{j}-\binom{r}{j}\right),

let sjejqs_{j}\coloneqq e_{j}^{q} and cre1(r1)eM(rM)c_{r}\,\coloneqq\,e_{1}^{\binom{r}{1}}\cdot\ldots\cdot e_{M}^{\binom{r}{M}}, let hi,r(t)qdeg(gi)gi(qt+r)h_{i,r}(t)\coloneqq q^{-\deg(g_{i})}g_{i}(qt+r) and ziuiqdeg(gi)z_{i}\coloneqq u_{i}^{q^{\deg(g_{i})}}, and define

wr(n)z1h1,r(n)zmhm,r(n)s1p1,r(n)sMpM,r(n).w_{r}(n)\coloneqq z_{1}^{h_{1,r}(n)}\cdot\ldots\cdot z_{m}^{h_{m,r}(n)}s_{1}^{p_{1,r}(n)}\cdot\ldots\cdot s_{M}^{p_{M,r}(n)}.

A straightforward calculation shows that

w(qn+r)=wr(n)cr,r{0,1,,q1}.w(qn+r)\,=\,w_{r}(n)c_{r},\qquad\forall r\in\{0,1,\ldots,q-1\}.

Also note that the elements z1,,zmz_{1},\ldots,z_{m} belong to GG^{\circ} and z1,,zm,s1,,sMz_{1},\ldots,z_{m},s_{1},\ldots,s_{M} are pairwise commuting. Additionally, the polynomials p1,r,,pM,rp_{1,r},\ldots,p_{M,r} satisfy conditions (F1) and (F2) from F, and the functions h1,r,,hm,rh_{1,r},\ldots,h_{m,r} belong to \mathcal{H} and satisfy conditions (F3) and (F4). Condition (F5) is also satisfied, because

hi,r(t)=qdeg(gi)+1gi(qt+r)=qdeg(gi)gi(qt+r)h_{i,r}^{\prime}(t)\,=\,q^{-\deg(g_{i})+1}g_{i}^{\prime}(qt+r)\,=\,q^{-\deg(g_{i}^{\prime})}g_{i}^{\prime}(qt+r)

and g1,,gmg_{1},\ldots,g_{m} satisfy (III). In conclusion, all the conditions of F are met, which means that for every r{0,1,,q1}r\in\{0,1,\ldots,q-1\} the sequence (wr(n)crΓ)n=(w(qn+r)Γ)n(w_{r}(n)c_{r}\Gamma)_{n\in\mathbb{N}}=(w(qn+r)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in the sub-nilmanifold z1zms1sMcrΓ¯\overline{z_{1}^{\mathbb{R}}\cdots z_{m}^{\mathbb{R}}s_{1}^{\mathbb{Z}}\cdots s_{M}^{\mathbb{Z}}c_{r}\Gamma}. Define

Yr=z1zms1sMcrΓ¯Y_{r}=\overline{z_{1}^{\mathbb{R}}\cdots z_{m}^{\mathbb{R}}s_{1}^{\mathbb{Z}}\cdots s_{M}^{\mathbb{Z}}c_{r}\Gamma}

and note that YrY_{r} is connected because ejq+rΓ¯\overline{e_{j}^{q\mathbb{Z}+r}\Gamma} is connected for all j{1,,M}j\in\{1,\ldots,M\}. Moreover, using that the map gcrgg\mapsto c_{r}g from GGG\to G is homeomorphic, we have

Yr\displaystyle Y_{r} =z1zms1sMcrΓ¯\displaystyle=\overline{z_{1}^{\mathbb{R}}\cdots z_{m}^{\mathbb{R}}s_{1}^{\mathbb{Z}}\cdots s_{M}^{\mathbb{Z}}c_{r}\Gamma}
=crz1zms1sMΓ¯\displaystyle=\overline{c_{r}z_{1}^{\mathbb{R}}\cdots z_{m}^{\mathbb{R}}s_{1}^{\mathbb{Z}}\cdots s_{M}^{\mathbb{Z}}\Gamma}
=crz1zms1sMΓ¯\displaystyle=c_{r}\,\overline{z_{1}^{\mathbb{R}}\cdots z_{m}^{\mathbb{R}}s_{1}^{\mathbb{Z}}\cdots s_{M}^{\mathbb{Z}}\Gamma}
=crY0.\displaystyle=c_{r}Y_{0}.

Let Z=z1zme1eMΓ¯Z=\overline{z_{1}^{\mathbb{R}}\cdots z_{m}^{\mathbb{R}}e_{1}^{\mathbb{Z}}\cdots e_{M}^{\mathbb{Z}}\Gamma} and observe that Y0,Y1,,Yq1Y_{0},Y_{1},\ldots,Y_{q-1} are sub-nilmanifolds of ZZ, because sj=ejqs_{j}=e_{j}^{q} and cr=e1(r1)eM(rM)c_{r}=e_{1}^{\binom{r}{1}}\cdot\ldots\cdot e_{M}^{\binom{r}{M}}. Also, finitely many translates of Y0Y_{0} cover ZZ, because

0α1,,αMq1e1α1eMαMY0=Z.\bigcup_{0\leqslant\alpha_{1},\ldots,\alpha_{M}\leqslant q-1}e_{1}^{\alpha_{1}}\cdot\ldots\cdot e_{M}^{\alpha_{M}}Y_{0}\leavevmode\nobreak\ =\leavevmode\nobreak\ Z.

Since Y0Y_{0} is connected and finitely many translates of it cover ZZ, we conclude that Y0Y_{0} is the identity component of ZZ, i.e., Y0Y_{0} is the connect component of ZZ containing Γ\Gamma. Let H~\tilde{H} be the stabilizer of ZZ,

H~={gG:gZ=Z},\tilde{H}=\{g\in G:gZ=Z\},

and define HH~H\coloneqq\tilde{H}^{\circ}. Then Z=H~ΓZ=\tilde{H}\Gamma and Y0=HΓY_{0}=H\Gamma. Moreover, since crH~c_{r}\in\tilde{H} and HH is a normal subgroup of H~\tilde{H}, it follows that

Yr\displaystyle Y_{r} =crY0\displaystyle=c_{r}Y_{0}
=crHΓ\displaystyle=c_{r}H\Gamma
=HcrΓ.\displaystyle=Hc_{r}\Gamma.

Thus, choosing xr=crΓx_{r}=c_{r}\Gamma, we have shown that Yr=HxrY_{r}=Hx_{r} for a connected and closed subgroup HH of GG, which finishes the proof of B. Also, since z1zms1sMcrΓ¯\overline{z_{1}^{\mathbb{R}}\cdots z_{m}^{\mathbb{R}}s_{1}^{\mathbb{Z}}\cdots s_{M}^{\mathbb{Z}}c_{r}\Gamma} is dense in XX if and only if its projection onto the maximal factor torus is dense there444Note that any nilmanifold possesses a niltranslation that acts ergodically on it. Therefore, given any sub-nilmanifold YY of X=G/ΓX=G/\Gamma there exists some aGa\in G that acts ergodically on YY. If the projection of YY onto the maximal factor torus is surjective, then aa also acts ergodically on the maximal factor torus, which in view of 1.2 implies that aa acts ergodically on XX. This shows that a sub-nilmanifold of XX coincides with XX if and only if its projection onto the maximal factor torus equals the the maximal factor torus., it follows that (w(qn+r)Γ)n(w(qn+r)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in XX if and only if its projection (ϑ(w(qn+r)Γ))n(\vartheta(w(qn+r)\Gamma))_{n\in\mathbb{N}} onto [G,G]\X[G^{\circ},G^{\circ}]\backslash X is uniformly distributed there. In particular, (w(n)Γ)n(w(n)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in XX if and only if (ϑ(w(n)Γ))n(\vartheta(w(n)\Gamma))_{n\in\mathbb{N}} is uniformly distributed in [G,G]\X[G^{\circ},G^{\circ}]\backslash X, which proves the conclusion of A. ∎

4.   Proof of F

For the proof of F we distinguish three cases:

The first case that we consider is when GG is abelian. Although this case essentially follows from the work of Boshernitzan [Boshernitzan94], for completeness we state and prove it in Section 4.1 below. It serves as the base case for the induction used in the proofs of the two subsequent cases.

The second case of F that we consider is when fk(t)tf_{k}(t)\prec t and b1==bm=1Gb_{1}=\ldots=b_{m}=1_{G}. We will refer to this as the “sub-linear” case of F, and it is proved in Section 4.2 below using induction on the nilpotency step of the Lie group GG. The reason why we consider this case separately is because its proof requires the use of a special type of van der Corput Lemma that is specifically designed to handle functions of sub-linear growth from a Hardy field (see 4.4).

Finally, in Section 4.3, we prove the general case of F. The proof of the general case bears many similarities to the proof of the “sub-linear” case, but relies on the standard van der Corput Lemma and instead of induction on the nilpotency step of GG uses induction on the so-called degree of the sequence v(n)v(n), see 4.7.

4.1.   The abelian case

The following corresponds to the special case of F where GG is abelian.

Theorem 4.1 (The abelian case of F).

Let d=d1+d2d=d_{1}+d_{2} and consider a map v:d1×d2v\colon\mathbb{N}\to\mathbb{R}^{d_{1}}\times\mathbb{Z}^{d_{2}} of the form

v(n)=f1(n)α1++fk(n)αk+p1(n)β1++pm(n)βm,v(n)=f_{1}(n)\alpha_{1}+\ldots+f_{k}(n)\alpha_{k}+p_{1}(n)\beta_{1}+\ldots+p_{m}(n)\beta_{m},

where α1,,αkd×{(0,,0)}\alpha_{1},\ldots,\alpha_{k}\in\mathbb{R}^{d}\times\{(0,\ldots,0)\}, β1,,βmd1×d2\beta_{1},\ldots,\beta_{m}\in\mathbb{R}^{d_{1}}\times\mathbb{Z}^{d_{2}}, p1,,pm[t]p_{1},\ldots,p_{m}\in\mathbb{R}[t] satisfy properties (F1) and (F2), and f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} satisfy properties (F3), (F4), and (F5). Moreover, let Λ\Lambda be a subgroup of d2\mathbb{Z}^{d_{2}} of finite index, define Δd2/Λ\Delta\coloneqq\mathbb{Z}^{d_{2}}/\Lambda, and assume that for every j=1,,mj=1,\ldots,m the set {nβjmod(d1×Λ):n}¯\overline{\{n\beta_{j}\bmod(\mathbb{Z}^{d_{1}}\times\Lambda):n\in\mathbb{Z}\}} is a connected subgroup of (d1×d2)/(d1×Λ)=𝕋d1×Δ(\mathbb{R}^{d_{1}}\times\mathbb{Z}^{d_{2}})/(\mathbb{Z}^{d_{1}}\times\Lambda)=\mathbb{T}^{d_{1}}\times\Delta. Then the sequence (v(n)mod(d1×Λ))n(v(n)\bmod(\mathbb{Z}^{d_{1}}\times\Lambda))_{n\in\mathbb{N}} is uniformly distributed in the subgroup

Tα1++αk+β1++βmmod(d1×Λ)¯T\coloneqq\overline{\mathbb{R}\alpha_{1}+\ldots+\mathbb{R}\alpha_{k}+\mathbb{Z}\beta_{1}+\ldots+\mathbb{Z}\beta_{m}\bmod(\mathbb{Z}^{d_{1}}\times\Lambda)}

of 𝕋d×Δ\mathbb{T}^{d}\times\Delta.

Proof.

Since 𝕋d1×Δ\mathbb{T}^{d_{1}}\times\Delta is a compact abelian group, the algebra generated by continuous group characters η:𝕋d1×ΔS1={z:|z|=1}\eta\colon\mathbb{T}^{d_{1}}\times\Delta\to S^{1}=\{z\in\mathbb{C}:|z|=1\} is uniformly dense C(𝕋d1×Δ)C(\mathbb{T}^{d_{1}}\times\Delta). Therefore, to prove that (v(n)mod(d1×Λ))n(v(n)\bmod(\mathbb{Z}^{d_{1}}\times\Lambda))_{n\in\mathbb{N}} is uniformly distributed in TT it suffices to show that for every continuous group characters η\eta that is non-trivial when restricted to TT we have

limN1Nn=1Nη(v(n)mod(d1×Λ))= 0.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\eta(v(n)\bmod(\mathbb{Z}^{d_{1}}\times\Lambda))\,=\,0. (4.1)

Let e(x)e(x) be shorthand for e2πixe^{2\pi ix} and choose ξ1,,ξk\xi_{1},\ldots,\xi_{k}\in\mathbb{R} and ζ1,,ζm[0,1)\zeta_{1},\ldots,\zeta_{m}\in[0,1) such that η(tαimod(d1×Λ))=e(tξi)\eta(t\alpha_{i}\bmod(\mathbb{Z}^{d_{1}}\times\Lambda))=e(t\xi_{i}) for all tt\in\mathbb{R} as well as η(nβjmod(d1×Λ))=e(nζj)\eta(n\beta_{j}\bmod(\mathbb{Z}^{d_{1}}\times\Lambda))=e(n\zeta_{j}) for all nn\in\mathbb{N}. This allows us to rewrite (4.1) as

limN1Nn=1Ne(f1(n)ξ1++fk(n)ξk+p1(n)ζ1++pm(n)ζm)= 0.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}e\big{(}f_{1}(n)\xi_{1}+\ldots+f_{k}(n)\xi_{k}+p_{1}(n)\zeta_{1}+\ldots+p_{m}(n)\zeta_{m}\big{)}\,=\,0. (4.2)

Since f1,,fkf_{1},\ldots,f_{k} have different growth and satisfy property (F4), if at least one of the numbers ξ1,,ξk\xi_{1},\ldots,\xi_{k} is non-zero then it follows from Boshernitzan’s Equidistribution Theorem ([Boshernitzan94, Theorem 1.8]) that (4.2) holds. If all of the ξ1,,ξk\xi_{1},\ldots,\xi_{k} are zero and at least one of the numbers ζ1,,ζk\zeta_{1},\ldots,\zeta_{k} is non-zero then (4.2) holds too, because p1,,pkp_{1},\ldots,p_{k} have different degree and, since {nβjmod(d1×Λ):n}¯\overline{\{n\beta_{j}\bmod(\mathbb{Z}^{d_{1}}\times\Lambda):n\in\mathbb{Z}\}} is connected for all j{1,,M}j\in\{1,\ldots,M\}, any non-zero ζj\zeta_{j} must be an irrational number and polynomials with irrational coefficients are uniformly distributed mod 11 by Weyl’s Equidistribution Theorem [Weyl16, Satz 14]. To finish the proof, note that not all the numbers ξ1,,ξk,ζ1,,ζm\xi_{1},\ldots,\xi_{k},\zeta_{1},\ldots,\zeta_{m} can be zero since η\eta was assumed to be non-trivial when restricted to TT. ∎

4.2.   The sub-linear case

The purpose of this subsection is to prove the special case of F when b1,,bm=1Gb_{1},\ldots,b_{m}=1_{G} and fk(t)tf_{k}(t)\prec t, which we dubbed the “sub-linear case”. For the convenience of the reader, let us state it as a separate theorem here.

Theorem 4.2 (The sub-linear case of F).

Let GG be a simply connected nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and \mathcal{H} a Hardy field. For any mapping v:Gv\colon\mathbb{N}\to G of the form

v(n)=a1f1(n)akfk(n),n,v(n)=a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G^{\circ} are commuting and log(t)f1fkt\log(t)\prec f_{1}\prec\ldots\prec f_{k}\prec t\in\mathcal{H}, the sequence (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in the sub-nilmanifold a1akΓ¯\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}\Gamma}.

Given a sub-nilmanifold YY of a nilmanifold X=G/ΓX=G/\Gamma and a function F𝖢(X)F\in\mathsf{C}(X) we will write F(Y)F(Y) to denote the quantity F𝖽μY\int F\leavevmode\nobreak\ \mathsf{d}\mu_{Y}. A sequence of sub-nilmanifolds (Yn)n(Y_{n})_{n\in\mathbb{N}} is said to be uniformly distributed in XX if

limN1Nn=1NF(Yn)=F𝖽μX,F𝖢(X).\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}F(Y_{n})\leavevmode\nobreak\ =\leavevmode\nobreak\ \int F\leavevmode\nobreak\ \mathsf{d}\mu_{X},\qquad\forall F\in\mathsf{C}(X).

The following lemma will be instrumental in our proof of 4.2.

Lemma 4.3.

Let GG be a connected simply connected nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG and consider the nilmanifold XG/ΓX\coloneqq G/\Gamma. Let bGb\in G be arbitrary and let LL denote the smallest connected, normal, rational, and closed subgroup of GG containing b={bt:t}b^{\mathbb{R}}=\{b^{t}:t\in\mathbb{R}\}. Let XX^{\triangle} denote the diagonal {(x,x):xX}\{(x,x):x\in X\}. Then for all but countably many ξ\xi\in\mathbb{R}, the sequence ((bξn,1G)X)n\big{(}(b^{\xi n},1_{G})X^{\triangle}\big{)}_{n\in\mathbb{N}} is uniformly distributed in the relatively independent product X×LXX{\mkern-1.0mu\times\mkern-1.5mu}_{L}X.555Since {(g,g):gG}\{(g,g):g\in G\} is a rational and closed subgroup of G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G, we can identify (gΓ,gΓ)(g\Gamma,g\Gamma) with (g,g)Γ×LΓ(g,g)\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma. This allows us to view the diagonal X{(x,x):xX}X^{\triangle}\coloneqq\{(x,x):x\in X\} as a sub-nilmanifold of the relatively independent product X×LXX{\mkern-1.0mu\times\mkern-1.5mu}_{L}X.

Proof.

It follows from [Leibman05b, Corollary 1.9] that ((bξn,1G)X)n\big{(}(b^{\xi n},1_{G})X^{\triangle}\big{)}_{n\in\mathbb{N}} is uniformly distributed in X×LXX{\mkern-1.0mu\times\mkern-1.5mu}_{L}X if and only if

limN1Nn=1Nη((bξn,1G)X)=0\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\eta\big{(}(b^{\xi n},1_{G})X^{\triangle}\big{)}=0 (4.3)

for every non-trivial horizontal character η\eta of (G×LG,Γ×LΓ)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma). As was mentioned in 2.13, for any horizontal character η\eta of (G×LG,Γ×LΓ)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma) there exists a horizontal character η1\eta_{1} of (G,Γ)(G,\Gamma) and a horizontal character η2\eta_{2} of (L,ΓL)(L,\Gamma_{L}) with [G,L]kerη2[G,L]\subset\ker\eta_{2} and such that

η((a1,a2)Γ×LΓ)=η1(a2Γ)η2(a1a21ΓL),(a1,a2)G×LG.\eta\big{(}(a_{1},a_{2})\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma\big{)}=\eta_{1}\big{(}a_{2}\Gamma\big{)}\eta_{2}\big{(}a_{1}a_{2}^{-1}\Gamma_{L}\big{)},\qquad\forall(a_{1},a_{2})\in G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G.

Therefore we get

limN1Nn=1Nη((bξn,1G)X)=(η1𝖽μX)(limN1Nn=1Nη2(bξnΓL)).\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\eta\big{(}(b^{\xi n},1_{G})X^{\triangle}\big{)}=\left(\int\eta_{1}\leavevmode\nobreak\ \mathsf{d}\mu_{X}\right)\left(\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\eta_{2}(b^{\xi n}\Gamma_{L})\right).

Since η\eta is non-trivial, either η1\eta_{1} or η2\eta_{2} is non-trivial. If η1\eta_{1} is non trivial then η1𝖽μX=0\int\eta_{1}\leavevmode\nobreak\ \mathsf{d}\mu_{X}=0 and hence (4.3) is satisfied. It remains to deal with the case when η2\eta_{2} is non-trivial.

Note that η2(btΓL)=λt\eta_{2}(b^{t}\Gamma_{L})=\lambda^{t} for some λ\lambda\in\mathbb{C} with |λ|=1|\lambda|=1. We claim that λ1\lambda\neq 1. Before we verify this claim, let us show how it allows us finish the proof of (4.3). Indeed, if λ1\lambda\neq 1 then λξ1\lambda^{\xi}\neq 1 for all but countably many ξ\xi. Since there are only countably many horizontal characters, by excluding countably many “bad” ξ\xis for each horizontal character, there exists a co-countable set of “good” ξ\xis independent of the choice of η2\eta_{2}. Since the Cesàro average of λξn\lambda^{\xi n} is 0 whenever λξ1\lambda^{\xi}\neq 1, it follows that for any such “good” ξ\xi we have

limN1Nn=1Nη2(bξnΓL)=0,\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\eta_{2}(b^{\xi n}\Gamma_{L})=0,

which proves that (4.3) holds.

Let us now prove the claim that λ1\lambda\neq 1 whenever η2\eta_{2} is non-trivial. By way of contradiction, assume λ=1\lambda=1. Therefore η2(btΓL)=1\eta_{2}(b^{t}\Gamma_{L})=1 for all tt\in\mathbb{R}. Consider the set

K{gL:η2(gΓL)=1}.K\coloneqq\{g\in L:\eta_{2}(g\Gamma_{L})=1\}.

Clearly bKb^{\mathbb{R}}\subset K. Since η2(g1g2ΓL)=η2(g1ΓL)η2(g2ΓL)\eta_{2}(g_{1}g_{2}\Gamma_{L})=\eta_{2}(g_{1}\Gamma_{L})\eta_{2}(g_{2}\Gamma_{L}) for all g1,g2Lg_{1},g_{2}\in L, we see that KK is a subgroup of LL, and hence also a subgroup of GG. Moreover, [L,G]kerη2[L,G]\subset\ker\eta_{2} implies [L,G]K[L,G]\subset K, from which we conclude that KK is a normal subgroup of GG. Since ggΓLg\mapsto g\Gamma_{L} and η2:L/ΓL\eta_{2}\colon L/\Gamma_{L}\to\mathbb{C} are continuous maps, KK is a closed set. Also, since KΓL=η1({1})K\Gamma_{L}=\eta^{-1}(\{1\}) is a closed subset of L/ΓLL/\Gamma_{L}, KK is rational. In summary, KK is a normal, rational, and closed subgroup of GG. Let KK^{\circ} be the identity component of KK. Then KK^{\circ} is also normal, rational, and closed. On top of that, KK^{\circ} is connected and contains bb^{\mathbb{R}}. By the minimality assumption on LL, we must have K=LK=L. However, we also have KLK\neq L because η2\eta_{2} was assumed to be non-trivial. This is a contradiction. ∎

Besides 4.3, another important ingredient in our proof of 4.2 is the following variant of van der Corput’s Lemma.

Proposition 4.4 (van der Corput’s Lemma for sub-linear functions).

Assume f1,,fkf_{1},\ldots,f_{k} are functions from a Hardy field \mathcal{H} satisfying log(t)f1(t)fkt\log(t)\prec f_{1}(t)\prec\ldots\prec f_{k}\prec t. Let Ψ:k\Psi\colon\mathbb{R}^{k}\to\mathbb{C} be a bounded and uniformly continuous function and suppose for all ss\in\mathbb{R} the limit

A(s)limN1Nn=1NΨ(f1(n),,fk1(n),fk(n)+s)Ψ(f1(n),,fk(n))¯A(s)\leavevmode\nobreak\ \coloneqq\leavevmode\nobreak\ \lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\Psi(f_{1}(n),\ldots,f_{k-1}(n),f_{k}(n)+s)\overline{\Psi(f_{1}(n),\ldots,f_{k}(n))}

exists. If for every ε>0\varepsilon>0 there exists ξ(0,ε)\xi\in(0,\varepsilon) such that

limH1Hh=1HA(ξh)= 0\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}A(\xi h)\leavevmode\nobreak\ =\leavevmode\nobreak\ 0

then necessarily

limN1Nn=1NΨ(f1(n),,fk(n))=0.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\Psi(f_{1}(n),\ldots,f_{k}(n))=0. (4.4)

4.4 is a special case of 6.1, which is stated and proved in Section 6. Let us now turn to the proof of 4.2.

Proof of 4.2.

We proceed by induction on the nilpotency step of GG. The case of 4.2 where GG is abelian (i.e., where the nilpotency step of GG equals 11) has already been taken care of by 4.1. Let us therefore assume that GG is a dd-step nilpotent Lie group with d2d\geqslant 2 and that 4.2 has already been proven for all cases where the nilpotency step of the Lie group is smaller than dd.

By replacing XX with the sub-nilmanifold a1akΓ¯\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}\Gamma} (and GG with a rational and closed subgroup of itself) if necessary, we can assume without loss of generality that

X=a1akΓ¯.X\,=\,\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}\Gamma}. (4.5)

In particular, XX is connected and therefore GΓ=GG^{\circ}\Gamma=G. On top of that, a1,,akGa_{1},\ldots,a_{k}\in G^{\circ}. This means we can replace GG by GG^{\circ} if needed, which allows us to also assume that GG is connected.

To show that (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in XX we must verify

limN1Nn=1NF(v(n)Γ)=F𝖽μX\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}F\big{(}v(n)\Gamma\big{)}=\int F\leavevmode\nobreak\ \mathsf{d}\mu_{X} (4.6)

for all continuous functions F𝖢(X)F\in\mathsf{C}(X). However, in light 2.6 it is actually not necessary to check (4.6) for all continuous functions. Indeed, since the linear span of central characters is uniformly dense in 𝖢(X)\mathsf{C}(X), instead of (4.6) it suffices to show

limN1Nn=1Nφ(v(n)Γ)=φ𝖽μX\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\varphi\big{(}v(n)\Gamma\big{)}=\int\varphi\leavevmode\nobreak\ \mathsf{d}\mu_{X} (4.7)

for central characters φ\varphi only. Let us therefore fix a central character φ\varphi and let χ\chi be the character of Z(G)Z(G) corresponding to φ\varphi, i.e., the continuous group homomorphism from Z(G)Z(G) to the unit circle {z:|z|=1}\{z\in\mathbb{C}:|z|=1\} such that (2.1) holds.

Write LL for the smallest connected, normal, rational and closed subgroup of GG containing the element aka_{k}. Let VV be the intersection of LL with Z(G)Z(G) and note that by 2.3, VV is a non-trivial subgroup of Z(G)Z(G). Moreover, since LL is connected and Z(G)Z(G) is connected (note that Z(G)Z(G) is connected because G=GG=G^{\circ}), VV is connected too.666It follows from [Malcev49, Lemma 3] that if aa belongs to a connected subgroup of a simply connected nilpotent Lie group GG then so does ata^{t} for all tt\in\mathbb{R}. This property implies that the intersection of connected subgroups of GG is connected.

We now claim that we can assume χ\chi is non-trivial when restricted to VV (by which we mean that there exists sVs\in V such that χ(s)1\chi(s)\neq 1). Indeed, if χ\chi is trivial on VV then φ\varphi is invariant under the action of VV. In this case, φ\varphi descends to a continuous function on the quotient space X/VX/V. We can identify X/VX/V with the nilmanifold (G/V)/(Γ/V)(G/V)/(\Gamma/V) and, since VV is connected and non-trivial, the dimension of X/VX/V is smaller than the dimension of XX. This allows us to reduce (4.7) to an analogous question on a nilmanifold of strictly smaller dimension. By induction on the dimension, we can thus assume that χ\chi is non-trivial when restricted to VV.

Since χ\chi is non-trivial when restricted to VV, it is in particular a non-trivial central character. Non-trivial central character have zero mean (see 2.7), and hence (4.7) becomes

limN1Nn=1Nφ(v(n)Γ)=0.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\varphi\big{(}v(n)\Gamma\big{)}=0. (4.8)

In order to prove (4.8) we use the “van der Corput Lemma for sub-linear functions”, i.e., 4.4. Define

A(s)limN1Nn=1Nφ(v(n)aksΓ)φ¯(v(n)Γ).A(s)\coloneqq\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\varphi\big{(}v(n)a_{k}^{s}\Gamma\big{)}\overline{\varphi}\big{(}v(n)\Gamma\big{)}. (4.9)

According to 4.4, if we can show that A(s)A(s) is well defined for all ss\in\mathbb{R} (meaning that the limit on the right hand side of (4.9) exists for all ss\in\mathbb{R}) and for every ε>0\varepsilon>0 there exists ξ(0,ε)\xi\in(0,\varepsilon) such that

limH1Hh=1HA(ξh)= 0,\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}A(\xi h)\leavevmode\nobreak\ =\leavevmode\nobreak\ 0, (4.10)

then (4.8) holds. Define

v(n)(v(n),v(n)),n,v^{\triangle}(n)\coloneqq\big{(}v(n),v(n)\big{)},\qquad\forall n\in\mathbb{N},

and let bakb\coloneqq a_{k}. Clearly, (bs,1G)v(n)(b^{s},1_{G})v^{\triangle}(n) takes values in G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G for all ss\in\mathbb{R} and nn\in\mathbb{N}. This will allow us to express A(s)A(s) in terms of the relatively independent product X×LXX{\mkern-1.0mu\times\mkern-1.5mu}_{L}X instead of the cartesian product X×XX{\mkern-1.0mu\times\mkern-0.5mu}X, which, as we will see, turns out to be a big advantage. Define the map Φ:X×LX\Phi\colon X{\mkern-1.0mu\times\mkern-1.5mu}_{L}X\to\mathbb{C} as Φ((g1,g2)Γ×LΓ)=φ(g1Γ)φ¯(g2Γ)\Phi\big{(}(g_{1},g_{2})\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma\big{)}=\varphi(g_{1}\Gamma)\overline{\varphi}(g_{2}\Gamma) for all (g1,g2)G×LG(g_{1},g_{2})\in G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G. Note that Φ\Phi is well defined and continuous. We can now rewrite (4.9) as

A(s)=limN1Nn=1NΦ((bs,1G)v(n)(Γ×LΓ)).A(s)=\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\Phi\big{(}(b^{s},1_{G})v^{\triangle}(n)(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma)\big{)}.

We make two claims:

Claim 1.

The integral Φ𝖽μX×LX\int\Phi\leavevmode\nobreak\ \mathsf{d}\mu_{X{\mkern-1.0mu\times\mkern-1.5mu}_{L}X} equals 0.

Claim 2.

For all ss\in\mathbb{R},

limN1Nn=1NΦ((bs,1G)v(n)(Γ×LΓ))=Φ((bs,1G)X).\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\Phi\big{(}(b^{s},1_{G})v^{\triangle}(n)(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma)\big{)}\,=\,\Phi\big{(}(b^{s},1_{G})\,X^{\triangle}\big{)}. (4.11)

Once Claim 1 and Claim 2 have been verified, we can finish the proof of (4.10) rather quickly. Indeed, Claim 2 implies that the limit in A(s)A(s) exists for all ss\in\mathbb{R} and

limH1Hh=1HA(ξh)=limH1Hh=1HΦ((bξh,1G)X).\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}A(\xi h)=\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}\Phi\big{(}(b^{\xi h},1_{G})X^{\triangle}\big{)}.

In view of 4.3, we thus have for a co-countable set of ξ\xi that

limH1Hh=1HΦ((bξh,1G)X)=Φ𝖽μX×LX,\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}\Phi\big{(}(b^{\xi h},1_{G})X^{\triangle}\big{)}=\int\Phi\leavevmode\nobreak\ \mathsf{d}\mu_{X{\mkern-1.0mu\times\mkern-1.5mu}_{L}X},

which together with Claim 1 implies (4.10).

It remains to verify Claims 1 and 2.

Proof of Claim 1.

Recall that there exists sVs\in V such that χ(s)1\chi(s)\neq 1. Using the definition of Φ\Phi it is straightforward to check that

Φ((s,1G)(g1,g2)(Γ×LΓ))=χ(s)Φ((g1,g2)(Γ×LΓ)),(g1,g2)G×LG.\Phi\big{(}(s,1_{G})(g_{1},g_{2})(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma)\big{)}=\chi(s)\Phi\big{(}(g_{1},g_{2})(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma)\big{)},\qquad\forall(g_{1},g_{2})\in G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G.

Note that (s,1G)(s,1_{G}) is an element of G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G because sVs\in V and VLV\subset L. Therefore μX×LX\mu_{X{\mkern-1.0mu\times\mkern-1.5mu}_{L}X} is invariant under left-multiplication by (s,1G)(s,1_{G}), which gives

Φ𝖽μX×LX=(s,1G)Φ𝖽μX×LX=χ(s)Φ𝖽μX×LX.\int\Phi\leavevmode\nobreak\ \mathsf{d}\mu_{X{\mkern-1.0mu\times\mkern-1.5mu}_{L}X}\leavevmode\nobreak\ =\leavevmode\nobreak\ \int(s,1_{G})\cdot\Phi\leavevmode\nobreak\ \mathsf{d}\mu_{X{\mkern-1.0mu\times\mkern-1.5mu}_{L}X}\leavevmode\nobreak\ =\leavevmode\nobreak\ \chi(s)\int\Phi\leavevmode\nobreak\ \mathsf{d}\mu_{X{\mkern-1.0mu\times\mkern-1.5mu}_{L}X}.

Since χ(s)1\chi(s)\neq 1, we obtain Φ𝖽μX×LX=0\int\Phi\leavevmode\nobreak\ \mathsf{d}\mu_{X{\mkern-1.0mu\times\mkern-1.5mu}_{L}X}=0 as claimed. ∎

Proof of Claim 2.

Define a new function Φs:X\Phi_{s}\colon X\to\mathbb{C} via

Φs(gΓ)Φ((bs,1G)(g,g)(Γ×LΓ)),gG.\Phi_{s}\big{(}g\Gamma\big{)}\coloneqq\Phi\big{(}(b^{s},1_{G})(g,g)(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma)\big{)},\qquad\forall g\in G.

It is straightforward to check that (4.11) is equivalent to

limN1Nn=1NΦs(v(n)Γ)=Φs𝖽μX.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\Phi_{s}\big{(}v(n)\Gamma\big{)}=\int\Phi_{s}\leavevmode\nobreak\ \mathsf{d}\mu_{X}. (4.12)

Write σ:GG/Z(G)\sigma\colon G\to G/Z(G) for the natural projection of GG onto G/Z(G)G/Z(G) and set

G^\displaystyle\widehat{G} \displaystyle\coloneqq G/Z(G);\displaystyle G/Z(G);
Γ^\displaystyle\widehat{\Gamma} \displaystyle\coloneqq σ(Γ);\displaystyle\sigma(\Gamma);
X^\displaystyle\widehat{X} \displaystyle\coloneqq G^/Γ^;\displaystyle\widehat{G}/\widehat{\Gamma};
v^(n)\displaystyle\widehat{v}(n) \displaystyle\coloneqq σ(v(n)).\displaystyle\sigma(v(n)).

It follows from (2.1) that Φ\Phi is invariant under the action of Z(G)={(g,g):gZ(G)}Z(G)^{\triangle}=\{(g,g):g\in Z(G)\} and so Φs\Phi_{s} is invariant under the action of Z(G)Z(G). Therefore, Φs\Phi_{s} descends to a continuous function Φ^s\widehat{\Phi}_{s} on X^\widehat{X}, which makes (4.12) equivalent to

limN1Nn=1NΦ^s(v^(n)Γ^)=Φ^s𝖽μX^.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\widehat{\Phi}_{s}\big{(}\widehat{v}(n)\widehat{\Gamma}\big{)}=\int\widehat{\Phi}_{s}\leavevmode\nobreak\ \mathsf{d}\mu_{\widehat{X}}. (4.13)

Since G^\widehat{G} has nilpotency step d1d-1, we can invoke the induction hypothesis and conclude that the sequence (v^(n)Γ^)n(\widehat{v}(n)\widehat{\Gamma})_{n\in\mathbb{N}} is uniformly distributed on the sub-nilmanifold a^1a^kΓ^¯,\overline{\widehat{a}_{1}^{\mathbb{R}}\cdot\ldots\cdot\widehat{a}_{k}^{\mathbb{R}}\widehat{\Gamma}}, where a^iσ(ai)\widehat{a}_{i}\coloneqq\sigma(a_{i}), i=0,1,,ki=0,1,\ldots,k. However, (4.5) implies

X^=a^1a^kΓ^¯,\widehat{X}\,=\,\overline{\widehat{a}_{1}^{\mathbb{R}}\cdot\ldots\cdot\widehat{a}_{k}^{\mathbb{R}}\widehat{\Gamma}},

which proves (4.13). ∎

This finishes the proofs of Claims 1 and 2, which in turn completes the proof of 4.2. ∎

4.3.   The general case

In this subsection we deal with the general case of F. In the proof of the “sub-linear” case in the previous subsection we got away with using induction on the nilpotency step of the Lie group GG. Unfortunately, the proof of the general case requires a more complicated inductive procedure. This inductive scheme bears similarities to the ones used in [GT12a] and [BMR17arXiv, Section 5] and relies on the notion of the “degree” associated to a mapping v:Gv\colon\mathbb{N}\to G. For the definition of this degree, the notion of a filtration is needed.

Definition 4.5.

Let GG be a nilpotent Lie group and Γ\Gamma a uniform and discrete subgroup of GG. Let dd\in\mathbb{N} and let G1GdGd+1G_{1}\supset\ldots\supset G_{d}\supset G_{d+1} be subgroups of GG that are normal, rational and closed. We call G{G1,,Gd,Gd+1}G_{\bullet}\coloneqq\{G_{1},\ldots,G_{d},G_{d+1}\} a dd-step filtration of GG if G1=GG_{1}=G, Gd+1={1G}G_{d+1}=\{1_{G}\}, and

[Gi,Gj]Gi+j,i,j{1,,d}withi+jd+1.[G_{i},G_{j}]\subset G_{i+j},\qquad\forall i,j\in\{1,\ldots,d\}\leavevmode\nobreak\ \text{with}\leavevmode\nobreak\ i+j\leqslant d+1.

The next lemma shows how one can turn a dd-step filtration of GG into a dd-step filtration of the relatively independent self-product G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G.

Lemma 4.6 (cf. [GT12a, Proposition 7.2]).

Let G={G1,,Gd,Gd+1}G_{\bullet}=\{G_{1},\ldots,G_{d},G_{d+1}\} be a dd-step filtration of GG and suppose LL is a normal subgroup of GG with LG2L\subset G_{2}. Define LiLGi+1L_{i}\coloneqq L\cap G_{i+1} for i=0,1,,di=0,1,\ldots,d, and set Ld+1{1G}L_{d+1}\coloneqq\{1_{G}\}. Then

(G×LG)={G1×L1G1,,Gd×LdGd,Gd+1×Ld+1Gd+1}(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)_{\bullet}=\big{\{}G_{1}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{1}}G_{1},\ \ldots,\ G_{d}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{d}}G_{d},\ G_{d+1}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{d+1}}G_{d+1}\big{\}}

is a dd-step filtration of G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G.

Proof.

Suppose (a,b)Gi×LiGi(a,b)\in G_{i}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{i}}G_{i} and (c,d)Gj×LjGj(c,d)\in G_{j}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{j}}G_{j}, where ii and jj belong to {1,,d}\{1,\ldots,d\} and satisfy i+jd+1i+j\leqslant d+1. To complete the proof we must show that ([a,c],[b,d])Gi+j×Li+jGi+j([a,c],[b,d])\in G_{i+j}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{i+j}}G_{i+j}. Clearly, [a,c]Gi+j[a,c]\in G_{i+j} and [b,d]Gi+j[b,d]\in G_{i+j}. It remains to prove that

[a,c][b,d]1Li+j.[a,c][b,d]^{-1}\in L_{i+j}. (4.14)

Since Li+j=LGi+j+1L_{i+j}=L\cap G_{i+j+1}, we will establish (4.14) in two steps: First we will show that [a,c][b,d]1L[a,c][b,d]^{-1}\in L, and thereafter we will show that [a,c][b,d]1Gi+j+1[a,c][b,d]^{-1}\in G_{i+j+1}.

Write sa1bs\coloneqq a^{-1}b and tc1dt\coloneqq c^{-1}d. Then sLis\in L_{i} and (a,b)=(a,as)(a,b)=(a,as). Likewise, tLjt\in L_{j} and (c,d)=(c,ct)(c,d)=(c,ct). So [a,c][b,d]1L[a,c][b,d]^{-1}\in L can be written as

[a,c][b,d]1=a1c1act1c1s1a1ctasL.[a,c][b,d]^{-1}=a^{-1}c^{-1}act^{-1}c^{-1}s^{-1}a^{-1}ctas\in L. (4.15)

Since sLs\in L, (4.15) is equivalent to

a1c1act1c1s1a1ctaL.a^{-1}c^{-1}act^{-1}c^{-1}s^{-1}a^{-1}cta\in L. (4.16)

Next, using aLa1=LaLa^{-1}=L and tLt\in L, we see that (4.16) is equivalent to

c1act1c1s1a1cL.c^{-1}act^{-1}c^{-1}s^{-1}a^{-1}c\in L. (4.17)

Using normality of LL again, (4.17) reduces to

ct1c1s1L.ct^{-1}c^{-1}s^{-1}\in L.

Finally, since sLs\in L and tLt\in L, it follows that ct1c1s1Lct^{-1}c^{-1}s^{-1}\in L, which finishes the proof of (4.15).

It remains to show that

[a,c][b,d]1=a1c1act1c1s1a1ctasGi+j+1.[a,c][b,d]^{-1}=a^{-1}c^{-1}act^{-1}c^{-1}s^{-1}a^{-1}ctas\in G_{i+j+1}. (4.18)

We can assume without loss of generality that iji\geqslant j. Since sGi+1s\in G_{i+1}, we have that [s,c]Gi+j+1[s,c]\in G_{i+j+1}, [s,t]Gi+j+2Gi+j+1[s,t]\in G_{i+j+2}\subset G_{i+j+1} and [s,a]G2i+1Gi+j+1[s,a]\in G_{2i+1}\subset G_{i+j+1}. In particular, modulo Gi+j+1G_{i+j+1} the element ss commutes with aa, cc, and tt. Hence

[a,c][b,d]1modGi+j+1=a1c1act1c1a1cta.[a,c][b,d]^{-1}\bmod G_{i+j+1}=a^{-1}c^{-1}act^{-1}c^{-1}a^{-1}cta.

Next, observe that [t,a]Gi+j+1[t,a]\in G_{i+j+1} and hence tt commutes with aa modulo Gi+j+1G_{i+j+1}. Therefore

[a,c][b,d]1modGi+j+1\displaystyle[a,c][b,d]^{-1}\bmod G_{i+j+1} =\displaystyle= a1c1act1c1a1ctamodGi+j+1\displaystyle a^{-1}c^{-1}act^{-1}c^{-1}a^{-1}cta\bmod G_{i+j+1}
=\displaystyle= a1c1act1c1a1catmodGi+j+1\displaystyle a^{-1}c^{-1}act^{-1}c^{-1}a^{-1}cat\bmod G_{i+j+1}
=\displaystyle= [[c,a],t]modGi+j+1.\displaystyle[[c,a],t]\bmod G_{i+j+1}.

Finally, since [c,a]Gi+j[c,a]\in G_{i+j} we have [[c,a],t]Gi+j+1[[c,a],t]\in G_{i+j+1} and conclude that [a,c][b,d]1Gi+j+1[a,c][b,d]^{-1}\in G_{i+j+1}. ∎

Given a nilpotent Lie group GG, let dG:G×G[0,)d_{G}\colon G{\mkern-1.0mu\times\mkern-0.5mu}G\to[0,\infty) be a right-invariant metric on GG. For any uniform and discrete subgroup Γ\Gamma the metric dGd_{G} descends to a metric dXd_{X} on the nilmanifold X=G/ΓX=G/\Gamma in the following way:

dX(xΓ,yΓ)inf{dG(xγ,yγ):γ,γΓ}.d_{X}(x\Gamma,y\Gamma)\coloneqq\inf\{d_{G}(x\gamma,y\gamma^{\prime}):\gamma,\gamma^{\prime}\in\Gamma\}. (4.19)

Given a subset SGS\subset G and a point gGg\in G we denote by d(g,S)infsSdG(g,s)d(g,S)\coloneqq\inf_{s\in S}d_{G}(g,s) the distance between SS and gg.

Definition 4.7 (cf. [GT12a, Definition 1.8] and [BMR17arXiv, Definition 5.9]).

Let GG be a simply connected nilpotent Lie group, \mathcal{H} a Hardy field, and v:Gv\colon\mathbb{N}\to G a mapping of the form

v(n)a1f1(n)akfk(n)b1p1(n)bmpm(n),n,v(n)\,\coloneqq\,a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)}b_{1}^{p_{1}(n)}\cdot\ldots\cdot b_{m}^{p_{m}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G^{\circ}, b1,,bmGb_{1},\ldots,b_{m}\in G, the elements a1,,ak,b1,,bma_{1},\ldots,a_{k},b_{1},\ldots,b_{m} are commuting, f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} have polynomial growth, and p1,,pm[t]p_{1},\ldots,p_{m}\in\mathbb{R}[t] with pj()p_{j}(\mathbb{Z})\subset\mathbb{Z}. We define the degree of vv to be the smallest number dd\in\mathbb{N} such that there exists a dd-step filtration G={G1,G2,,Gd,Gd+1}G_{\bullet}=\{G_{1},G_{2},\ldots,G_{d},G_{d+1}\} with the property that bjGdeg(pj)+1b_{j}\in G_{\deg(p_{j})+1} for all j=1,,mj=1,\ldots,m and aiGdeg(fi)a_{i}\in G_{\deg(f_{i})}^{\circ} for all i=1,,ki=1,\ldots,k. If GG_{\bullet} is such a minimal filtration then we say GG_{\bullet} realizes the degree of vv. If there exists no such filtration, then we say that vv has infinite degree.

Lemma 4.8.

Let GG be a simply connected nilpotent Lie group GG and \mathcal{H} a Hardy field. Assume v:Gv\colon\mathbb{N}\to G is a mapping of the form

v(n)a1f1(n)akfk(n)b1p1(n)bmpm(n),n,v(n)\,\coloneqq\,a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)}b_{1}^{p_{1}(n)}\cdot\ldots\cdot b_{m}^{p_{m}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G^{\circ}, b1,,bmGb_{1},\ldots,b_{m}\in G, the elements a1,,ak,b1,,bma_{1},\ldots,a_{k},b_{1},\ldots,b_{m} are commuting, f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} have polynomial growth, and p1,,pm[t]p_{1},\ldots,p_{m}\in\mathbb{R}[t] with pj()p_{j}(\mathbb{Z})\subset\mathbb{Z}. Then vv has finite degree.

Proof.

Let MM\in\mathbb{N} be any number such that deg(fi)M\deg(f_{i})\leqslant M for all i=1,,ki=1,\ldots,k and deg(pj)+1M\deg(p_{j})+1\leqslant M for all j=1,,mj=1,\ldots,m. Let C{C1,C2,,Cs,Cs+1}C_{\bullet}\coloneqq\{C_{1},C_{2},\ldots,C_{s},C_{s+1}\} denote the lower central series of GG. Set r(s+1)(M+1)r\coloneqq(s+1)(M+1) and define a filtration

G={G1,G2,,Gr,Gr+1}G_{\bullet}=\{G_{1},G_{2},\ldots,G_{r},G_{r+1}\}

by setting G(j1)(M+1)+iCjG_{(j-1)(M+1)+i}\coloneqq C_{j} for all j{1,,s+1}j\in\{1,\ldots,s+1\} and i{1,,M+1}i\in\{1,\ldots,M+1\} and Gr+1={1G}G_{r+1}=\{1_{G}\}. It is straightforward to check that GG_{\bullet} is a filtration. Also, since Gi=GG_{i}=G for all i=1,,M+1i=1,\ldots,M+1, we certainly have that bjGdeg(pj)+1b_{j}\in G_{\deg(p_{j})+1} for all j=1,,mj=1,\ldots,m and aiGdeg(fi)a_{i}\in G_{\deg(f_{i})}^{\circ} for all i=1,,ki=1,\ldots,k. ∎

Remark 4.9.

Note that the the filtration G={G1,G2,,Gr,Gr+1}G_{\bullet}=\{G_{1},G_{2},\ldots,G_{r},G_{r+1}\} constructed in the above proof is not necessarily a filtration that realizes the degree of vv. Nonetheless, its existence proves that the degree of vv does not exceed r=(s+1)(M+1)r=(s+1)(M+1).

Proof of F, the general case.

Let v:Gv\colon\mathbb{N}\to G be as in the statement of F. We use induction on the degree dd of vv, which is finite due to 4.8. The base case of this induction, which is when d=1d=1, is covered by 4.1, because if d=1d=1 then GG must be abelian. Therefore, we only have to deal with the inductive step. Assume d>1d>1 and F has already been proven for all mappings v^:G^\widehat{v}\colon\mathbb{N}\to\widehat{G} satisfying the hypothesis of F and whose degree is strictly smaller than dd.

By replacing XX with a1akb1bmΓ¯\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma} if necessary777Let GG^{\prime} be the the smallest rational and closed subgroup of GG containing a1akb1bma_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}. Then Γ=GΓ\Gamma^{\prime}=G^{\prime}\cap\Gamma is a uniform and discrete subgroup of GG^{\prime} and the nilmanifold XG/ΓX^{\prime}\coloneqq G^{\prime}/\Gamma^{\prime} can be identified with the sub-nilmanifold a1akb1bmΓ¯\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma} of XX. Moreover, (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in a1akb1bmΓ¯\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma} if and only if (v(n)Γ)n(v(n)\Gamma^{\prime})_{n\in\mathbb{N}} is uniformly distributed in XX^{\prime}. Thus, by replacing GG with GG^{\prime}, Γ\Gamma with Γ\Gamma^{\prime}, and XX with XX^{\prime}, we can assume without loss of generality that X=a1akb1bmΓ¯X=\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma}., we will assume that

X=a1akb1bmΓ¯.X\,=\,\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma}. (4.20)

Since b1Γ¯,,bmΓ¯\overline{b_{1}^{\mathbb{Z}}\Gamma},\ldots,\overline{b_{m}^{\mathbb{Z}}\Gamma} are assumed to be connected, the sub-nilmanifold b1bmΓ¯\overline{b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma} is also connected. It follows that a1ak(b1bmΓ¯)a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}(\overline{b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma}) is connected, which in turn implies that

a1ak(b1bmΓ¯)¯=a1akb1bmΓ¯=X\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}(\overline{b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma})}\,=\,\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma}\,=\,X

is connected. For technical reasons, it will be convenient to assume that for every j=1,,mj=1,\ldots,m, if the sub-nilmanifold bjΓ¯\overline{b_{j}^{\mathbb{Z}}\Gamma} is a point then bj=1Gb_{j}=1_{G}. This assumption can be made without loss of generality, because if bjΓ¯\overline{b_{j}^{\mathbb{Z}}\Gamma} is a point then bjb_{j} must belong to Γ\Gamma, in which case we can simply replace bjb_{j} with 1G1_{G} and the sequence (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}} remains unchanged.

Our goal is to show that (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}} is uniformly distributed in XX, or equivalently,

limN1Nn=1NF(v(n)Γ)=F𝖽μX\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}F\big{(}v(n)\Gamma\big{)}=\int F\leavevmode\nobreak\ \mathsf{d}\mu_{X} (4.21)

for all F𝖢(X)F\in\mathsf{C}(X). Repeating the same argument as was already used in the proof of 4.2, we see that instead of (4.21) it suffices to show

limN1Nn=1Nφ(v(n)Γ)=φ𝖽μX\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\varphi\big{(}v(n)\Gamma\big{)}=\int\varphi\leavevmode\nobreak\ \mathsf{d}\mu_{X} (4.22)

for all central characters φ\varphi. Let \mathcal{I} denote all the numbers i{1,,k}i\in\{1,\ldots,k\} for which deg(fi)2\deg(f_{i})\geqslant 2, and let 𝒥\mathcal{J} be all the numbers j{1,,m}j\in\{1,\ldots,m\} for which bj1Gb_{j}\neq 1_{G}. Note that if \mathcal{I} and 𝒥\mathcal{J} are both the empty set then fk(t)tf_{k}(t)\prec t and b1==bm=1Gb_{1}=\ldots=b_{m}=1_{G}, and so we find ourselves in the “sub-linear case” of F. Since this case has already been taken care of by 4.2, we can assume that either \mathcal{I} or 𝒥\mathcal{J} is non-empty.

Let G={G1,G2,,Gd,Gd+1}G_{\bullet}=\{G_{1},G_{2},\ldots,G_{d},G_{d+1}\} be a dd-step filtration that realizes the degree of vv (cf. 4.7). Among other things, this means bjG2b_{j}^{\mathbb{Z}}\subset G_{2} for all j=1,,mj=1,\ldots,m and aiG2a_{i}\subset G_{2}^{\circ} for all ii\in\mathcal{I}. Note that aiG2a_{i}\subset G_{2}^{\circ} implies aiG2a_{i}^{\mathbb{R}}\subset G_{2}^{\circ} (cf. Footnote 6). Let LL denote the smallest closed rational and normal subgroup of GG containing aia_{i}^{\mathbb{R}} for all ii\in\mathcal{I} and bjb_{j}^{\mathbb{Z}} for all j=1,,mj=1,\ldots,m. Then, if π:GX\pi\colon G\to X denotes the natural projection of GG onto XX, the set π(L)\pi(L) is a sub-nilmanifold of XX containing the sub-nilmanifold iai1jmbjΓ¯\overline{\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}\prod_{1\leqslant j\leqslant m}b_{j}^{\mathbb{Z}}\Gamma}. This sub-nilmanifold is what Leibman calls the normal closure, and it is shown in [Leibman10a, p. 844] that the normal closure of a connected sub-nilmanifold is connected. In particular, since iai1jmbjΓ¯\overline{\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}\prod_{1\leqslant j\leqslant m}b_{j}^{\mathbb{Z}}\Gamma} is connected, π(L)\pi(L) is connected too.

Next, we claim that the identity component LL^{\circ} of LL is non-trivial. To verify this claim, we are going to distinguish two cases. The first case is when \mathcal{I} is non-empty. In this case, LL^{\circ} contains a one-parameter subgroup aia_{i}^{\mathbb{R}} for ii\in\mathcal{I} and is therefore non-trivial (we assume without loss of generality that ai1Ga_{i}\neq 1_{G} for all i=1,,ki=1,\ldots,k). The second case is when 𝒥\mathcal{J} is non-empty. Note that if we are not in the first case, then we must be in the second, since either \mathcal{I} or 𝒥\mathcal{J} is non-empty. If 𝒥\mathcal{J} is non-empty then π(L)\pi(L^{\circ}) contains bjΓ¯\overline{b_{j}^{\mathbb{Z}}\Gamma} for some j𝒥j\in\mathcal{J}, and since bjΓ¯\overline{b_{j}^{\mathbb{Z}}\Gamma} is connected and not a point for every j𝒥j\in\mathcal{J}, it follows that LL^{\circ} is non-trivial.

Let VV be the intersection of LL^{\circ} with Z(G)Z(G)^{\circ}. By 2.4, VV is a non-trivial subgroup of Z(G)Z(G). Also, as an intersection of connected subgroups, VV is connected (cf. Footnote 6). We can now use the same argument as in the proof of 4.2, which involved induction on the dimension of the nilmanifold XX, to show that it suffices to prove (4.22) for the case when the central character φ\varphi has a “central frequency” χ\chi that is non-trivial when restricted to VV, i.e., there exists sVs\in V such that χ(s)0\chi(s)\neq 0. This also implies that φ𝖽μX=0\int\varphi\leavevmode\nobreak\ \mathsf{d}\mu_{X}=0, and hence (4.22) can be written as

limN1Nn=1Nφ(v(n)Γ)=0.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\varphi\big{(}v(n)\Gamma\big{)}=0. (4.23)

To prove (4.23) we use van der Corput’s trick. Define

A(h)limN1Nn=1Nφ(v(n+h)Γ)φ¯(v(n)Γ)A(h)\coloneqq\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\varphi\big{(}v(n+h)\Gamma\big{)}\overline{\varphi}\big{(}v(n)\Gamma\big{)} (4.24)

whenever this limit exists. In light of A.8 (applied with pn=1p_{n}=1 for all nn\in\mathbb{N}), (4.23) holds if we can show that the limit on the right hand side of (4.24) exists for all hh\in\mathbb{N} and

limH1Hh=1HA(h)= 0.\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}A(h)\leavevmode\nobreak\ =\leavevmode\nobreak\ 0. (4.25)

We can interpret φ(v(n+h)Γ)φ(v(n)Γ)¯\varphi(v(n+h)\Gamma)\overline{\varphi(v(n)\Gamma)} as φφ¯((v(n+h),v(n))(Γ×Γ))\varphi{\mkern 1.8mu\otimes\mkern 1.6mu}\overline{\varphi}((v(n+h),v(n))(\Gamma{\mkern-1.0mu\times\mkern-0.5mu}\Gamma)), where φφ¯\varphi{\mkern 1.8mu\otimes\mkern 1.6mu}\overline{\varphi} is a continuous function on the product nilmanifold X×X=(G×G)/(Γ×Γ)X{\mkern-1.0mu\times\mkern-0.5mu}X=(G{\mkern-1.0mu\times\mkern-0.5mu}G)/(\Gamma{\mkern-1.0mu\times\mkern-0.5mu}\Gamma) and (v(n+h),v(n))(v(n+h),v(n)) is an element in G×GG{\mkern-1.0mu\times\mkern-0.5mu}G. Note that the sequence (v(n+h),v(n))(v(n+h),v(n)) can be rewritten as

(v(n+h),v(n))=(Δhw(n),1G)v(n),(v(n+h),v(n))\leavevmode\nobreak\ =\leavevmode\nobreak\ (\Delta_{h}w(n),1_{G})\,v^{\triangle}(n), (4.26)

where v(n)=(v(n),v(n))v^{\triangle}(n)=(v(n),v(n)) and

Δhw(n)=a1Δhf1(n)akΔhfk(n)b1Δhp1(n)bmΔhpm(n),n,h.\Delta_{h}w(n)\leavevmode\nobreak\ =\leavevmode\nobreak\ a_{1}^{\Delta_{h}f_{1}(n)}\cdot\ldots\cdot a_{k}^{\Delta_{h}f_{k}(n)}b_{1}^{\Delta_{h}p_{1}(n)}\cdot\ldots\cdot b_{m}^{\Delta_{h}p_{m}(n)},\qquad\forall n,h\in\mathbb{N}. (4.27)

For every ii\notin\mathcal{I} the function fif_{i} has degree 11, which means its discrete derivative Δhfi(n)\Delta_{h}f_{i}(n) is negligibly small for large nn. This implies that for every ii\notin\mathcal{I} the element aiΔhfi(n)a_{i}^{\Delta_{h}f_{i}(n)} converges to the identity 1G1_{G} and can therefore be ignored. More precisely, using the right-invariance of the metric dGd_{G}, we have

limndG(Δhw(n),wh(n))= 0,\lim_{n\to\infty}d_{G}\left(\Delta_{h}w(n),\,w_{h}(n)\right)\,=\,0,

where

wh(n)iaiΔhfi(n)1jmbjΔhpj(n).w_{h}(n)\coloneqq\prod_{i\in\mathcal{I}}a_{i}^{\Delta_{h}f_{i}(n)}\cdot\prod_{1\leqslant j\leqslant m}b_{j}^{\Delta_{h}p_{j}(n)}.

It follows that if we set

vh(n)(wh(n),1G)v(n)v^{\square}_{h}(n)\,\coloneqq\,(w_{h}(n),1_{G})\,v^{\triangle}(n) (4.28)

then, in view of (4.26), the difference between (v(n+h),v(n))(v(n+h),v(n)) and vh(n)v^{\square}_{h}(n) goes to zero as nn\to\infty. Hence A(h)A(h) equals

A(h)=limN1Nn=1N(φφ¯)(vh(n)(Γ×Γ)).A(h)\,=\,\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}(\varphi{\mkern 1.8mu\otimes\mkern 1.6mu}\overline{\varphi})\big{(}v_{h}^{\square}(n)(\Gamma{\mkern-1.0mu\times\mkern-0.5mu}\Gamma)\big{)}. (4.29)

The advantage of using (4.29) instead of (4.24) is that vh(n)G×LGv^{\square}_{h}(n)\in G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G for all nn\in\mathbb{N}. Define the map Φ:X×LX\Phi\colon X{\mkern-1.0mu\times\mkern-1.5mu}_{L}X\to\mathbb{C} as Φ((g1,g2)Γ×LΓ)=φ(g1Γ)φ(g2Γ)¯\Phi\big{(}(g_{1},g_{2})\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma\big{)}=\varphi(g_{1}\Gamma)\overline{\varphi(g_{2}\Gamma)} for all (g1,g2)G×LG(g_{1},g_{2})\in G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G. Note that Φ\Phi is well defined and continuous. This allows us to rewrite (4.29) as

A(h)=limN1Nn=1NΦ(vh(n)(Γ×LΓ)).A(h)\,=\,\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\Phi\big{(}v^{\square}_{h}(n)(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma)\big{)}. (4.30)

Let Z(G){(g,g):gZ(G)}Z(G)^{\triangle}\coloneqq\{(g,g):g\in Z(G)\} and denote by σ:G×LG(G×LG)/Z(G)\sigma\colon G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G\to(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)/Z(G)^{\triangle} the natural projection of G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G onto (G×LG)/Z(G)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)/Z(G)^{\triangle}. Define

G^\displaystyle\widehat{G} \displaystyle\coloneqq σ(G×LG),\displaystyle\sigma(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G),
Γ^\displaystyle\widehat{\Gamma} \displaystyle\coloneqq σ(Γ×LΓ),\displaystyle\sigma\big{(}\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma\big{)},
X^\displaystyle\widehat{X} \displaystyle\coloneqq G^/Γ^,\displaystyle\widehat{G}/\widehat{\Gamma},
v^h(n)\displaystyle\widehat{v}_{h}(n) \displaystyle\coloneqq σ(vh(n)).\displaystyle\sigma\big{(}v^{\square}_{h}(n)\big{)}.

It follows from (2.1) that Φ\Phi is invariant under the action of Z(G)Z(G)^{\triangle}. Therefore, Φ\Phi descends to a continuous function on X^\widehat{X}, meaning there exists Φ^𝖢(X^)\widehat{\Phi}\in\mathsf{C}(\widehat{X}) such that

Φ((g1,g2)(Γ×LΓ))=Φ^(σ(g1,g2)Γ^),(g1,g2)G×LG.\Phi\big{(}(g_{1},g_{2})(\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma)\big{)}=\widehat{\Phi}\big{(}\sigma(g_{1},g_{2})\widehat{\Gamma}\big{)},\qquad\forall(g_{1},g_{2})\in G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G.

It thus follows form (4.30) that

A(h)=limN1Nn=1NΦ^(v^h(n)Γ^).A(h)\,=\,\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}\widehat{\Phi}\big{(}\widehat{v}_{h}(n)\widehat{\Gamma}\big{)}. (4.31)

We now make four claims.

Claim 1.

The integral Φ^𝖽μX^\int\widehat{\Phi}\leavevmode\nobreak\ \mathsf{d}\mu_{\widehat{X}} equals zero.

Claim 2.

For all non-trivial pseudo-horizontal characters η^\widehat{\eta} of (G^,Γ^)(\widehat{G},\widehat{\Gamma}) (see 2.15) we have

limHlimN1Hh=1H1Nn=1Nη^(v^h(n)Γ^)=0.\lim_{H\to\infty}\lim_{N\to\infty}\frac{1}{H}\sum_{h=1}^{H}\frac{1}{N}\sum_{n=1}^{N}\widehat{\eta}\big{(}\widehat{v}_{h}(n)\,\widehat{\Gamma}\big{)}=0. (4.32)

In the following, we call any mapping u:Hu\colon\mathbb{N}\to H of the from u(n)=g1p1(n)gp(n)u(n)=g_{1}^{p_{1}(n)}\cdot\ldots\cdot g_{\ell}^{p_{\ell}(n)}, where g1,,gng_{1},\ldots,g_{n} are elements in a nilpotent Lie group HH and p1,,pp_{1},\ldots,p_{\ell} are polynomials with pi()p_{i}(\mathbb{Z})\subset\mathbb{Z}, a polynomial mapping (cf. [Leibman05a, Subsection 1.3]).

Claim 3.

There exist polynomials q1,,qm[t]q_{1},\ldots,q_{m}\in\mathbb{R}[t] with qj()q_{j}(\mathbb{Z})\subset\mathbb{Z} and 1deg(q1)<<deg(qm)m1\leqslant\deg(q_{1})<\ldots<\deg(q_{m})\leqslant m, polynomial mappings c,e1,,em:G^c,e_{1},\ldots,e_{m}\colon\mathbb{N}\to\widehat{G}, and polynomial mappings u1,,uk:G^u_{1},\ldots,u_{k}\colon\mathbb{N}\to\widehat{G}^{\circ} such that the elements c(h),e1(h),,em(h),u1(h),,uk(h)c(h),e_{1}(h),\ldots,e_{m}(h),u_{1}(h),\ldots,u_{k}(h) are pairwise commuting for every hh\in\mathbb{N}, and

dG^(v^h(n),u1(h)f1(n)uk(h)fk(n)e1(h)q1(n)em(h)qm(n)c(h))=on(1)d_{\widehat{G}}\Big{(}\widehat{v}_{h}(n),\ u_{1}(h)^{f_{1}(n)}\cdot\ldots\cdot u_{k}(h)^{f_{k}(n)}e_{1}(h)^{q_{1}(n)}\cdot\ldots\cdot e_{m}(h)^{q_{m}(n)}c(h)\Big{)}={\rm o}_{n\to\infty}(1)

for every hh\in\mathbb{N}.

Claim 4.

For every hh the degree of v^h\widehat{v}_{h} is smaller than dd.

Before we provide the proofs of Claims 1, 2, 3, and 4, let us see how they can be used to prove that the limit in A(h)A(h) exists for all hh\in\mathbb{N} and (4.25) holds. Claims 3 and 4 allow us to invoke the induction hypothesis and deduce that for every hh\in\mathbb{N} the sequence (v^h(n)Γ^)n(\widehat{v}_{h}(n)\widehat{\Gamma})_{n\in\mathbb{N}} is uniformly distributed in the sub-nilmanifold

Y^hu1(h)uk(h)e1(h)em(h)c(h)Γ^¯.\widehat{Y}_{h}\,\coloneqq\,\overline{u_{1}(h)^{\mathbb{R}}\cdot\ldots\cdot u_{k}(h)^{\mathbb{R}}e_{1}(h)^{\mathbb{Z}}\cdot\ldots\cdot e_{m}(h)^{\mathbb{Z}}c(h)\widehat{\Gamma}}.

(As was explained in 3.1, it is not a problem that the “base point” of the sequence (v^h(n)Γ^)n(\widehat{v}_{h}(n)\widehat{\Gamma})_{n\in\mathbb{N}} is c(h)Γ^c(h)\widehat{\Gamma} instead of Γ^\widehat{\Gamma}.) As a consequence we have A(h)=Φ^(Y^h)A(h)=\widehat{\Phi}(\widehat{Y}_{h}), which in particular proves that the limit in A(h)A(h) exists for all hh\in\mathbb{N}. Moreover, (4.25) will follow if we can show that

limH1Hh=1HΦ^(Y^h)= 0.\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}\widehat{\Phi}(\widehat{Y}_{h})\,=\,0. (4.33)

Given a vector ξ=(ξ1,,ξk)k\xi=(\xi_{1},\ldots,\xi_{k})\in\mathbb{R}^{k}, consider the multi-parameter polynomial sequence wξ:k+m+1G^w_{\xi}\colon\mathbb{Z}^{k+m+1}\to\widehat{G} defined as

wξ(h,n1,,nk,1,,m)=u1(h)ξ1n1uk(h)ξknke1(h)1em(h)mc(h).w_{\xi}(h,n_{1},\ldots,n_{k},\ell_{1},\ldots,\ell_{m})=u_{1}(h)^{\xi_{1}n_{1}}\cdot\ldots\cdot u_{k}(h)^{\xi_{k}n_{k}}e_{1}(h)^{\ell_{1}}\cdot\ldots\cdot e_{m}(h)^{\ell_{m}}c(h).

Arguing as in the proof of [BMR17arXiv, Lemma A.7] we can find for every hh\in\mathbb{N} a co-null set Ξhk\Xi_{h}\subset\mathbb{R}^{k} such that for all ξ=(ξ1,,ξk)Ξh\xi=(\xi_{1},\ldots,\xi_{k})\in\Xi_{h} we have

u1(h)uk(h)e1(h)em(h)c(h)Γ^¯=u1(h)ξ1uk(h)ξke1(h)em(h)c(h)Γ^¯.\begin{split}&\overline{u_{1}(h)^{\mathbb{R}}\cdot\ldots\cdot u_{k}(h)^{\mathbb{R}}e_{1}(h)^{\mathbb{Z}}\cdot\ldots\cdot e_{m}(h)^{\mathbb{Z}}c(h)\widehat{\Gamma}}\\ &\qquad\qquad\leavevmode\nobreak\ =\leavevmode\nobreak\ \overline{u_{1}(h)^{\xi_{1}\mathbb{Z}}\cdot\ldots\cdot u_{k}(h)^{\xi_{k}\mathbb{Z}}e_{1}(h)^{\mathbb{Z}}\cdot\ldots\cdot e_{m}(h)^{\mathbb{Z}}c(h)\widehat{\Gamma}}.\end{split} (4.34)

It follows that for every hh\in\mathbb{N} and every ξΞh\xi\in\Xi_{h} the sequence

(wξ(h,n1,,nk,1,,m,h)Γ^)(n1,,nk,1,,m)k+m(w_{\xi}(h,n_{1},\ldots,n_{k},\ell_{1},\ldots,\ell_{m},h)\widehat{\Gamma})_{(n_{1},\ldots,n_{k},\ell_{1},\ldots,\ell_{m})\in\mathbb{N}^{k+m}}

is dense in Y^h\widehat{Y}_{h}. By invoking [Leibman05b, Theorem A, p. 216], we have that since this sequence is dense in Y^h\widehat{Y}_{h}, it is also uniformly distributed in Y^h\widehat{Y}_{h}. This means that

limN1Nm+kn1,,m=1NF^(wξ(h,n1,,nk,1,,m,h)Γ^)=F^𝖽μY^h\lim_{N\to\infty}\frac{1}{N^{m+k}}\sum_{n_{1},\ldots,\ell_{m}=1}^{N}\widehat{F}\big{(}w_{\xi}(h,n_{1},\ldots,n_{k},\ell_{1},\ldots,\ell_{m},h)\widehat{\Gamma}\big{)}=\int\widehat{F}\leavevmode\nobreak\ \mathsf{d}\mu_{\widehat{Y}_{h}}

for all continuous functions F^:X^\widehat{F}\colon\widehat{X}\to\mathbb{C}. Henceforth, let ξ\xi be any number in hΞh\bigcap_{h\in\mathbb{N}}\Xi_{h}. Note that Claim 2 implies

limH1Hh=1Hη^(Y^h)= 0\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}\widehat{\eta}(\widehat{Y}_{h})\,=\,0

for all non-trivial pseudo-horizontal characters η^\widehat{\eta} of (G^,Γ^)(\widehat{G},\widehat{\Gamma}). It follows that

limHlimN1Hh=1H1Nm+kn1,,m=1Nη^(wξ(n1,,nk,1,,m,h)Γ^)= 0.\lim_{H\to\infty}\lim_{N\to\infty}\frac{1}{H}\sum_{h=1}^{H}\frac{1}{N^{m+k}}\sum_{n_{1},\ldots,\ell_{m}=1}^{N}\widehat{\eta}\big{(}w_{\xi}(n_{1},\ldots,n_{k},\ell_{1},\ldots,\ell_{m},h)\widehat{\Gamma}\big{)}\,=\,0.

Note also that X×LXX{\mkern-1.0mu\times\mkern-1.5mu}_{L}X is connected, due to 2.17 and the fact that both XX and π(L)\pi(L) are connected. Thus, it follows from the work of Leibman (see [Leibman05b, Theorems A and B]) that the sequence

(wξ(h,n1,,nk,1,,m,h)Γ^)(h,n1,,nk,1,,m)k+m+1(w_{\xi}(h,n_{1},\ldots,n_{k},\ell_{1},\ldots,\ell_{m},h)\widehat{\Gamma})_{(h,n_{1},\ldots,n_{k},\ell_{1},\ldots,\ell_{m})\in\mathbb{N}^{k+m+1}}

is well distributed888A sequence (xn1,,nk)(n1,,nk)k(x_{n_{1},\ldots,n_{k}})_{(n_{1},\ldots,n_{k})\in\mathbb{N}^{k}} of points in a nilmanifold XX is said to be well distributed in XX if for all F𝖢(X)F\in\mathsf{C}(X) and all ε>0\varepsilon>0 there exists KK\in\mathbb{N} such that for all M1,N1,M2,N2,,Mk,NkM_{1},N_{1},M_{2},N_{2},\ldots,M_{k},N_{k}\in\mathbb{N} with NiMiKN_{i}-M_{i}\geqslant K for all i=1,,ki=1,\ldots,k we have |1(N1M1)(NkMk)(n1,,nk)[M1,N1)××[Mk,Nk)F(xn1,,nk)F𝖽μX|ε.\Big{|}\frac{1}{(N_{1}-M_{1})\cdot\ldots\cdot(N_{k}-M_{k})}\sum_{(n_{1},\ldots,n_{k})\in[M_{1},N_{1}){\mkern-1.0mu\times\mkern-0.5mu}\ldots{\mkern-1.0mu\times\mkern-0.5mu}[M_{k},N_{k})}F(x_{n_{1},\ldots,n_{k}})-\int F\leavevmode\nobreak\ \mathsf{d}\mu_{X}\Big{|}\leavevmode\nobreak\ \leqslant\leavevmode\nobreak\ \varepsilon. in X^\widehat{X}. We conclude that

limH1H\displaystyle\lim_{H\to\infty}\frac{1}{H} h=1HF^(Y^h)\displaystyle\sum_{h=1}^{H}\widehat{F}\big{(}\widehat{Y}_{h}\big{)}
=limHlimN1Hh=1H1Nm+kn1,,m=1NF^(wξ(n1,,nk,1,,m,h)Γ^)\displaystyle=\lim_{H\to\infty}\lim_{N\to\infty}\frac{1}{H}\sum_{h=1}^{H}\frac{1}{N^{m+k}}\sum_{n_{1},\ldots,\ell_{m}=1}^{N}\widehat{F}\big{(}w_{\xi}(n_{1},\ldots,n_{k},\ell_{1},\ldots,\ell_{m},h)\widehat{\Gamma}\big{)}
=F^𝖽μX^\displaystyle=\int\widehat{F}\leavevmode\nobreak\ \mathsf{d}\mu_{\widehat{X}}

for all continuous functions F^:X^\widehat{F}\colon\widehat{X}\to\mathbb{C}. In particular,

limH1Hh=1HΦ^(Y^h)=Φ^𝖽μX^.\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}\widehat{\Phi}(\widehat{Y}_{h})\,=\,\int\widehat{\Phi}\leavevmode\nobreak\ \mathsf{d}\mu_{\widehat{X}}.

Now we can simply invoke Claim 1 to conclude that (4.33) holds.

Let us now turn to the proofs of Claims 1, 2, 3, and 4.

Proof of Claim 1.

The following argument is very similar to the proof of Claim 1 which appeared in the proof of 4.2 in Section 4.2 above. Recall that χ\chi is non-trivial when restricted to VV, meaning that there exists sVs\in V such that χ(s)1\chi(s)\neq 1. Let s^σ(s,1G)\widehat{s}\coloneqq\sigma(s,1_{G}), where 1G1_{G} denotes the identity element of GG. Using the definition of Φ^\widehat{\Phi} it is straightforward to check that

Φ^(s^x^)=χ(s)Φ^(x^),x^X^.\widehat{\Phi}(\widehat{s}\widehat{x})\,=\,\chi(s)\widehat{\Phi}(\widehat{x}),\qquad\forall\widehat{x}\in\widehat{X}.

Since μX^\mu_{\widehat{X}} is invariant under s^\widehat{s}, we have that

Φ^(x^)𝖽μX^(x^)=Φ^(s^x^)𝖽μX^(x^)=χ(s)Φ^(x^)𝖽μX^(x^),\int\widehat{\Phi}(\widehat{x})\leavevmode\nobreak\ \mathsf{d}\mu_{\widehat{X}}(\widehat{x})\,=\,\int\widehat{\Phi}(\widehat{s}\widehat{x})\leavevmode\nobreak\ \mathsf{d}\mu_{\widehat{X}}(\widehat{x})\,=\,\chi(s)\int\widehat{\Phi}(\widehat{x})\leavevmode\nobreak\ \mathsf{d}\mu_{\widehat{X}}(\widehat{x}),

and hence Φ^𝖽μX^=0\int\widehat{\Phi}\leavevmode\nobreak\ \mathsf{d}\mu_{\widehat{X}}=0 as claimed. ∎

Proof of Claim 2.

For any pseudo-horizontal character η^\widehat{\eta} of (G^,Γ^)(\widehat{G},\widehat{\Gamma}) there exists a pseudo-horizontal character η\eta of (G×LG,Γ×LΓ)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma) such that η^σ=η\widehat{\eta}\circ\sigma=\eta. Thus, instead of (4.32), it suffices to show that for all non-trivial pseudo-horizontal characters η\eta of (G×LG,Γ×LΓ)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma) we have

limHlimN1Hh=1H1Nn=1Nη(vh(n)Γ×LΓ)=0.\lim_{H\to\infty}\lim_{N\to\infty}\frac{1}{H}\sum_{h=1}^{H}\frac{1}{N}\sum_{n=1}^{N}\eta\big{(}v_{h}^{\square}(n)\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma\big{)}=0. (4.35)

According to 2.16, there exist a pseudo-horizontal character η1\eta_{1} of (G,Γ)(G,\Gamma) and a pseudo-horizontal character η2\eta_{2} of (L,(LΓ))(L,(L\cap\Gamma)) with [G,L]kerη2[G^{\circ},L^{\circ}]\subset\ker\eta_{2} such that

η((a,b)Γ×LΓ)=η1(bΓ)η2(ab1ΓL),(a,b)G×LG,\eta\big{(}(a,b)\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma\big{)}=\eta_{1}(b\Gamma)\eta_{2}(ab^{-1}\Gamma_{L}),\qquad\forall(a,b)\in G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G,

where ΓLΓL\Gamma_{L}\coloneqq\Gamma\cap L. Thus, by (4.28),

η(vh(n)Γ×LΓ)=η1(v(n)Γ)η2(wh(n)ΓL).\eta\big{(}v_{h}^{\square}(n)\Gamma{\mkern-1.0mu\times\mkern-1.5mu}_{L}\Gamma\big{)}=\eta_{1}\big{(}v(n)\Gamma\big{)}\eta_{2}\big{(}w_{h}(n)\Gamma_{L}\big{)}.

Although we have b1,,bmLb_{1},\ldots,b_{m}\in L, we don’t necessarily have b1,,bmLb_{1},\ldots,b_{m}\in L^{\circ}. This makes it more difficult to study the expressions η1(v(n)Γ)\eta_{1}(v(n)\Gamma) and η2(wh(n)ΓL)\eta_{2}(w_{h}(n)\Gamma_{L}), because η1\eta_{1} and η2\eta_{2} are only pseudo-horizontal characters and not horizontal characters. However, we can circumvent these difficulties in the following way. Since π(L)\pi(L) is connected, we have

LΓL=L.L^{\circ}\Gamma_{L}=L. (4.36)

As is explained in [Leibman05a, Subsections 2.6 and 2.7 on p. 204], under these conditions there exist a polynomial sequence g1:Lg_{1}\colon\mathbb{N}\to L^{\circ} such that

g1(n)ΓL=b1p1(n)bmpm(n)ΓL,n.g_{1}(n)\Gamma_{L}\,=\,b_{1}^{p_{1}(n)}\cdot\ldots\cdot b_{m}^{p_{m}(n)}\Gamma_{L},\qquad\forall n\in\mathbb{N}. (4.37)

The advantage of using g1(n)g_{1}(n) instead of b1p1(n)bmpm(n)b_{1}^{p_{1}(n)}\cdot\ldots\cdot b_{m}^{p_{m}(n)} is that g1(n)g_{1}(n) takes values in LGL^{\circ}\subset G^{\circ} and hence the image of g1(n)g_{1}(n) under η1\eta_{1} and η2\eta_{2} is easier to understand. A downside of making this trade-off is that, unlike b1,,bmb_{1},\ldots,b_{m}, the values of g1(n)g_{1}(n) do not necessarily commute with a1,,aka_{1},\ldots,a_{k}. But for the current proof (meaning the proof of Claim 2) this commutativity is not needed. Similarly, we can find a polynomial sequence in two variables g2:2Lg_{2}\colon\mathbb{N}^{2}\to L^{\circ} such that

g2(n,h)ΓL=b1Δhp1(n)bmΔhpm(n)ΓL,n,h.g_{2}(n,h)\Gamma_{L}\,=\,b_{1}^{\Delta_{h}p_{1}(n)}\cdot\ldots\cdot b_{m}^{\Delta_{h}p_{m}(n)}\Gamma_{L},\qquad\forall n,h\in\mathbb{N}.

Note that even though

b1Δhp1(n)bmΔhpm(n)=(b1p1(n+h)bmpm(n+h))(b1p1(n)bmpm(n))1,b_{1}^{\Delta_{h}p_{1}(n)}\cdot\ldots\cdot b_{m}^{\Delta_{h}p_{m}(n)}\leavevmode\nobreak\ =\leavevmode\nobreak\ \left(b_{1}^{p_{1}(n+h)}\cdot\ldots\cdot b_{m}^{p_{m}(n+h)}\right)\left(b_{1}^{p_{1}(n)}\cdot\ldots\cdot b_{m}^{p_{m}(n)}\right)^{-1},

we do not necessarily have g2(n,h)=g1(n+h)g1(n)1g_{2}(n,h)=g_{1}(n+h)g_{1}(n)^{-1}. But we do have

g2(n,h)[L,L]ΓL=g1(n+h)g1(n)1[L,L]ΓL,g_{2}(n,h)[L^{\circ},L^{\circ}]\Gamma_{L}\leavevmode\nobreak\ =\leavevmode\nobreak\ g_{1}(n+h)g_{1}(n)^{-1}[L^{\circ},L^{\circ}]\Gamma_{L}, (4.38)

which we will make use of later.

It will be convenient to pick α1,,αk\alpha_{1},\ldots,\alpha_{k} and (ζi)i(\zeta_{i})_{i\in\mathcal{I}} such that

η1(aitΓ)\displaystyle\eta_{1}(a_{i}^{t}\Gamma) =\displaystyle= e(tαi),i=1,,k,t,\displaystyle e(t\alpha_{i}),\quad i=1,\ldots,k,\leavevmode\nobreak\ \forall t\in\mathbb{R},
η2(aitΓL)\displaystyle\eta_{2}(a_{i}^{t}\Gamma_{L}) =\displaystyle= e(tζi),i,t,\displaystyle e(t\zeta_{i}),\quad i\in\mathcal{I},\leavevmode\nobreak\ \forall t\in\mathbb{R},

where e(x)e(x) is shorthand for e2πixe^{2\pi ix}. From (2.4) and the fact that a1,,akGa_{1},\ldots,a_{k}\in G^{\circ} as well as aiLa_{i}\in L^{\circ} for all ii\in\mathcal{I}, it follows that

η1(v(n)Γ)η2(wh(n)ΓL)==η1(1ikaifi(n)1jmbjpj(n)Γ)η2(iaiΔhfi(n)1jmbjΔhpj(n)ΓL)=e(i=1kfi(n)αi+iΔhfi(n)ζi)η1(1jmbjpj(n)Γ)η2(1jmbjΔhpj(n)ΓL)=e(i=1kfi(n)αi+iΔhfi(n)ζi)η1(g1(n)Γ)η2(g2(n,h)ΓL).\begin{split}\eta_{1}\big{(}v(n)&\Gamma\big{)}\eta_{2}\big{(}w_{h}(n)\Gamma_{L}\big{)}=\\ =&\leavevmode\nobreak\ \eta_{1}\left(\prod_{1\leqslant i\leqslant k}a_{i}^{f_{i}(n)}\cdot\prod_{1\leqslant j\leqslant m}b_{j}^{p_{j}(n)}\Gamma\right)\eta_{2}\left(\prod_{i\in\mathcal{I}}a_{i}^{\Delta_{h}f_{i}(n)}\cdot\prod_{1\leqslant j\leqslant m}b_{j}^{\Delta_{h}p_{j}(n)}\Gamma_{L}\right)\\ =&\leavevmode\nobreak\ e\left(\sum_{i=1}^{k}f_{i}(n)\alpha_{i}\,+\sum_{i\in\mathcal{I}}\Delta_{h}f_{i}(n)\zeta_{i}\right)\eta_{1}\left(\prod_{1\leqslant j\leqslant m}b_{j}^{p_{j}(n)}\Gamma\right)\eta_{2}\left(\prod_{1\leqslant j\leqslant m}b_{j}^{\Delta_{h}p_{j}(n)}\Gamma_{L}\right)\\ =&\leavevmode\nobreak\ e\left(\sum_{i=1}^{k}f_{i}(n)\alpha_{i}\,+\sum_{i\in\mathcal{I}}\Delta_{h}f_{i}(n)\zeta_{i}\right)\eta_{1}\left(g_{1}(n)\Gamma\right)\eta_{2}\left(g_{2}(n,h)\Gamma_{L}\right).\end{split} (4.39)

Note that [L,L]ΓL[L^{\circ},L^{\circ}]\Gamma_{L} belongs to the kernel of η2\eta_{2} and so it follows from (4.38) that

η2(g2(n,h)ΓL)=η2(g1(n+h)g1(n)1ΓL).\eta_{2}\left(g_{2}(n,h)\Gamma_{L}\right)\leavevmode\nobreak\ =\leavevmode\nobreak\ \eta_{2}\left(g_{1}(n+h)g_{1}(n)^{-1}\Gamma_{L}\right).

Let r1,r2[t]r_{1},r_{2}\in\mathbb{R}[t] be polynomials such that

η1(g1(n))=e(r1(n))\eta_{1}\left(g_{1}(n)\right)\leavevmode\nobreak\ =\leavevmode\nobreak\ e\big{(}r_{1}(n)\big{)}

as well as

η2(g1(n)ΓL)=e(r2(n)).\eta_{2}\left(g_{1}(n)\Gamma_{L}\right)\leavevmode\nobreak\ =\leavevmode\nobreak\ e\big{(}r_{2}(n)\big{)}.

Then (4.39) implies

η1(v(n)Γ)η2(wh(n)ΓL)=e(i=1kfi(n)αi+iΔhfi(n)ζi+r1(n)+r2(n+h)r2(n))\eta_{1}\big{(}v(n)\Gamma\big{)}\eta_{2}\big{(}w_{h}(n)\Gamma_{L}\big{)}=e\left(\sum_{i=1}^{k}f_{i}(n)\alpha_{i}\,+\sum_{i\in\mathcal{I}}\Delta_{h}f_{i}(n)\zeta_{i}+r_{1}(n)+r_{2}(n+h)-r_{2}(n)\right)

and so (4.35) becomes

limHlimN1Hh=1H1Nn=1Ne(i=1kfi(n)αi+iΔhfi(n)ζi+r1(n)+r2(n+h)r2(n))=0.\lim_{H\to\infty}\lim_{N\to\infty}\frac{1}{H}\sum_{h=1}^{H}\frac{1}{N}\sum_{n=1}^{N}e\left(\sum_{i=1}^{k}f_{i}(n)\alpha_{i}+\sum_{i\in\mathcal{I}}\Delta_{h}f_{i}(n)\zeta_{i}+r_{1}(n)+r_{2}(n+h)-r_{2}(n)\right)=0. (4.40)

Since f1,,fkf_{1},\ldots,f_{k} have different growth (see property (F3)) and behave independently from polynomials (due to property (F4)), it follows that if at least one of the αi\alpha_{i} is non-zero or at least one of the ζi\zeta_{i} is non-zero, then (4.40) is satisfied and we are done. Let us therefore assume αi=0\alpha_{i}=0 for all i=1,,ki=1,\ldots,k and ζi=0\zeta_{i}=0 for all ii\in\mathcal{I}. In this case, (4.40) is equivalent to

limHlimN1Hh=1H1Nn=1Ne(r1(n)+r2(n+h)r2(n))=0.\lim_{H\to\infty}\lim_{N\to\infty}\frac{1}{H}\sum_{h=1}^{H}\frac{1}{N}\sum_{n=1}^{N}e\left(r_{1}(n)+r_{2}(n+h)-r_{2}(n)\right)=0. (4.41)

Averages of polynomial sequences are known to behave very regularly. In particular, the order of limits in (4.41) can be interchanged freely, which means that (4.41) is equivalent to

limNlimH1Nn=1N1Hh=1He(r1(n)+r2(n+h)r2(n))=0,\lim_{N\to\infty}\lim_{H\to\infty}\frac{1}{N}\sum_{n=1}^{N}\frac{1}{H}\sum_{h=1}^{H}e\left(r_{1}(n)+r_{2}(n+h)-r_{2}(n)\right)=0,

which is the same as

(limN1Nn=1Ne(r1(n)r2(n)))(limH1Hh=1He(r2(h)))=0.\left(\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}e\left(r_{1}(n)-r_{2}(n)\right)\right)\left(\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}e\left(r_{2}(h)\right)\right)=0. (4.42)

Recall that biΓ¯\overline{b_{i}^{\mathbb{Z}}\Gamma} is connected for all i=1,,mi=1,\ldots,m. This implies that biΓL¯\overline{b_{i}^{\mathbb{Z}}\Gamma_{L}} is also connected for all i=1,,mi=1,\ldots,m and therefore

{b1p1(n)bmpm(n):n}¯=b1bmΓ¯\overline{\big{\{}b_{1}^{p_{1}(n)}\cdot\ldots\cdot b_{m}^{p_{m}(n)}:n\in\mathbb{Z}\big{\}}}\leavevmode\nobreak\ =\leavevmode\nobreak\ \overline{b_{1}^{\mathbb{Z}}\cdot\ldots\cdot b_{m}^{\mathbb{Z}}\Gamma}

is connected. It now follows from (4.37) that g1()Γ¯\overline{g_{1}(\mathbb{Z})\Gamma} and g1()ΓL¯\overline{g_{1}(\mathbb{Z})\Gamma_{L}} are connected. But if g1()ΓL¯\overline{g_{1}(\mathbb{Z})\Gamma_{L}} is connected then, because e(r2(n))=η2(g1(n)ΓL)e(r_{2}(n))=\eta_{2}(g_{1}(n)\Gamma_{L}), as soon as the function ne(r2(n))n\mapsto e(r_{2}(n)) is non-constant, the average

limH1Hh=1He(r2(h))\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}e\left(r_{2}(h)\right)

must equal 0. If this average equals 0 then (4.42) holds, which implies that (4.41) holds, and once again we are done. Let us therefore assume that ne(r2(n))n\mapsto e(r_{2}(n)) is constant. Since e(r2(0))=1e(r_{2}(0))=1, if ne(r2(n))n\mapsto e(r_{2}(n)) is constant then we must have e(r2(n))=1e(r_{2}(n))=1 for all nn\in\mathbb{N}. Therefore g1(n)g_{1}(n) belongs to the kernel of η2\eta_{2} and (4.41) becomes

limN1Nn=1Ne(r1(n))=0.\lim_{N\to\infty}\frac{1}{N}\sum_{n=1}^{N}e\left(r_{1}(n)\right)=0. (4.43)

Since g1()Γ¯\overline{g_{1}(\mathbb{Z})\Gamma} is connected, we once again only have two possibilities: either e(r1(n))e(r_{1}(n)) is constant equal to 11 or (4.43) is satisfied. Since we are done if (4.43) is satisfied, the proof of Claim 2 is completed if we can show that e(r1(n))e(r_{1}(n)) cannot be constant equal to 11 under the current assumptions.

By way of contradiction, assume e(r1(n))=1e(r_{1}(n))=1 for all nn\in\mathbb{N}. This implies that g1(n)g_{1}(n) belongs to the kernel of η1\eta_{1}. But we also have that a1,,aka_{1},\ldots,a_{k} belong to the kernel of η1\eta_{1} and aia_{i} for ii\in\mathcal{I} as well as g1(n)g_{1}(n) for all nn\in\mathbb{N} belong to the kernel of η2\eta_{2}. We claim that having all those elements belong to the kernels of η1\eta_{1} and η2\eta_{2} contradicts the hypothesis that either η1\eta_{1} or η2\eta_{2} are non-trivial.

To verify this claim, we first need to make a simplifying assumption. Note that the group generated by GG^{\circ} and b1,,bmb_{1},\ldots,b_{m} is closed and rational999 Since XX is connected, be have GΓ=GG^{\circ}\Gamma=G. Therefore, for every i=1,,mi=1,\ldots,m there exists γiΓ\gamma_{i}\in\Gamma such that biγiGb_{i}\gamma_{i}\in G^{\circ}. This means that the group generated by GG^{\circ} and b1,,bmb_{1},\ldots,b_{m} equals GΓG^{\circ}\Gamma^{\prime}, where Γ\Gamma^{\prime} is the subgroup of Γ\Gamma generated by γ1,,γm\gamma_{1},\ldots,\gamma_{m}. This proves that the group generated by GG^{\circ} and b1,,bmb_{1},\ldots,b_{m} is both closed and rational., and it contains a1,,aka_{1},\ldots,a_{k} as well as b1,,bmb_{1},\ldots,b_{m}. Therefore we can replace GG by G,b1,,bm\langle G^{\circ},b_{1},\ldots,b_{m}\rangle if necessary, and will henceforth assume that G=G,b1,,bmG=\langle G^{\circ},b_{1},\ldots,b_{m}\rangle.

Next, define

Niaib1bm[G,L]ΓL,N\,\coloneqq\,\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}\cdot b_{1}^{\mathbb{Z}}\cdot\ldots\cdot b_{m}^{\mathbb{Z}}[G^{\circ},L^{\circ}]\Gamma_{L^{\circ}},

where ΓL=ΓL\Gamma_{L^{\circ}}=\Gamma\cap L^{\circ}. We claim that NN is a normal subgroup of GG with a dense subset of rational elements. Once verified, this claim will imply that the closure of NN, which we denote by N¯\overline{N}, is a closed, rational, and normal subgroup of GG.

To show that NN is a group, define

N0iai[G,L]ΓLandHb1bmN_{0}\,\coloneqq\,\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}[G^{\circ},L^{\circ}]\Gamma_{L^{\circ}}\qquad\text{and}\qquad H\,\coloneqq\,b_{1}^{\mathbb{Z}}\cdot\ldots\cdot b_{m}^{\mathbb{Z}}

and note that N=HN0N=HN_{0}. Certainly, HH is a subgroup of GG. If we can show that N0N_{0} is a normal subgroup of GG then it will follow that NN is a group.

Since G=G,b1,,bmG=\langle G^{\circ},b_{1},\ldots,b_{m}\rangle, to prove that N0N_{0} is normal it suffices to show that g1N0g=N0g^{-1}N_{0}g=N_{0} for all gGg\in G^{\circ}, and bj1N0bj=N0b_{j}^{-1}N_{0}b_{j}=N_{0} for all j=1,,mj=1,\ldots,m. It is easy to see that that g1N0g=N0g^{-1}N_{0}g=N_{0} holds for all gGg\in G^{\circ}, because N0N_{0} contains [G,L][G^{\circ},L^{\circ}]. To show that bj1N0bj=N0b_{j}^{-1}N_{0}b_{j}=N_{0} for all j=1,,mj=1,\ldots,m fix some jj between 11 and mm. Since bjb_{j} commutes with aia_{i}, we have

bj1N0bj=bj1(iai[G,L]ΓL)bj=iaibj1[G,L]ΓLbj.b_{j}^{-1}N_{0}b_{j}\,=\,b_{j}^{-1}\left(\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}[G^{\circ},L^{\circ}]\Gamma_{L^{\circ}}\right)b_{j}\,=\,\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}b_{j}^{-1}[G^{\circ},L^{\circ}]\Gamma_{L^{\circ}}b_{j}.

Since both GG^{\circ} and LL^{\circ} are normal subgroups of GG, the commutator [G,L][G^{\circ},L^{\circ}] is normal and hence

iaibj1[G,L]ΓLbj=iai[G,L]bj1ΓLbj.\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}b_{j}^{-1}[G^{\circ},L^{\circ}]\Gamma_{L^{\circ}}b_{j}\,=\,\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}[G^{\circ},L^{\circ}]b_{j}^{-1}\Gamma_{L^{\circ}}b_{j}.

Since LΓL=ΓLL=LL^{\circ}\Gamma_{L}=\Gamma_{L}L^{\circ}=L (cf. (4.36)), there exists γΓL\gamma\in\Gamma_{L} and cLc\in L^{\circ} such that γc=bj\gamma c=b_{j}. Hence

iai[G,L]bj1ΓLbj=iai[G,L]c1γ1ΓLγc.\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}[G^{\circ},L^{\circ}]b_{j}^{-1}\Gamma_{L^{\circ}}b_{j}\,=\,\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}[G^{\circ},L^{\circ}]c^{-1}\gamma^{-1}\Gamma_{L^{\circ}}\gamma c.

Since ΓL\Gamma_{L^{\circ}} is a normal subgroup of Γ\Gamma, we have γ1ΓLγ=ΓL\gamma^{-1}\Gamma_{L^{\circ}}\gamma=\Gamma_{L^{\circ}}. So

iai[G,L]c1γ1ΓLγc=iai[G,L]c1ΓLc.\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}[G^{\circ},L^{\circ}]c^{-1}\gamma^{-1}\Gamma_{L^{\circ}}\gamma c\,=\,\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}[G^{\circ},L^{\circ}]c^{-1}\Gamma_{L^{\circ}}c.

Finally, observe that c1ΓLc=ΓLmod[G,L]c^{-1}\Gamma_{L^{\circ}}c=\Gamma_{L^{\circ}}\bmod[G^{\circ},L^{\circ}], which gives

iai[G,L]c1ΓLc=iai[G,L]ΓL=N0.\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}[G^{\circ},L^{\circ}]c^{-1}\Gamma_{L^{\circ}}c\,=\,\prod_{i\in\mathcal{I}}a_{i}^{\mathbb{R}}[G^{\circ},L^{\circ}]\Gamma_{L^{\circ}}\,=\,N_{0}.

This proves that N0N_{0} is a normal subgroup of GG and hence NN is a subgroup of GG.

Next, let us show that NN is normal too. When proving that N0N_{0} is normal, we used that GG is generated by GG^{\circ} and b1,,bmb_{1},\ldots,b_{m}. For the proof that NN is normal, this does not seem to be particularly helpful. Instead, we shall use that the set a1akb1bmΓa_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma is dense in GG (which follows from (4.20)). Therefore, to show that NN is normal, it suffices to prove that ai1Nai=Na_{i}^{-1}Na_{i}=N for all i=1,,ki=1,\ldots,k, bj1Nbj=Nb_{j}^{-1}Nb_{j}=N for all j=1,,mj=1,\ldots,m, and γ1Nγ=N\gamma^{-1}N\gamma=N for all γΓ\gamma\in\Gamma. To verify the first assertion, namely that ai1Nai=Na_{i}^{-1}Na_{i}=N, simply note that N=HN0N=HN_{0}, where N0N_{0} is normal and HH is a group every element of which commutes with aia_{i}. A similar argument shows that bj1Nbj=Nb_{j}^{-1}Nb_{j}=N. To see why γ1Nγ=N\gamma^{-1}N\gamma=N holds for all γΓ\gamma\in\Gamma, simply note that N=HN0=ΓLN0N=HN_{0}=\Gamma_{L}N_{0} and ΓL\Gamma_{L} is a normal subgroup of Γ\Gamma.

Finally let us show that rational elements are dense in NN, or equivalently, that N¯\overline{N} is rational. It is well known (see [Leibman05a, Subsection 2.2, p. 203–204]) that a closed subgroup of GG is rational if and only if its intersection with Γ\Gamma is a uniform subgroup of that group. Hence, to prove that N¯\overline{N} is rational, it suffices to show that ΓN¯\Gamma\cap\overline{N} is a uniform subgroup of N¯\overline{N}. However, since ΓLN\Gamma_{L}\subset N and N¯L\overline{N}\subset L, it follows that ΓN¯=ΓL\Gamma\cap\overline{N}=\Gamma_{L}. Since ΓL\Gamma_{L} is a uniform subgroup of LL and N¯\overline{N} is a subgroup of LL, it follows that ΓL\Gamma_{L} is a uniform subgroup of N¯\overline{N} and we are done.

In conclusion, N¯\overline{N} is a closed and rational subgroup of LL that contains aia_{i} for all ii\in\mathcal{I} and bjb_{j} for all j=1,,mj=1,\ldots,m. Moreover, N¯\overline{N} is a normal subgroup of GG. Since, by definition, LL is the smallest subgroup of GG with all these properties, we must have

L=N¯.L=\overline{N}.

Recall that a1,,aka_{1},\ldots,a_{k} and g1(n)g_{1}(n) for all nn\in\mathbb{N} belong to the kernel of η1\eta_{1} and aia_{i} for ii\in\mathcal{I} and g1(n)g_{1}(n) for all nn\in\mathbb{N} belong to the kernel of η2\eta_{2}. From this it follows that b1,,bmb_{1},\ldots,b_{m} also belong to the kernel of both η1\eta_{1} and η2\eta_{2}. Recall also that [G,L]kerη2[G^{\circ},L^{\circ}]\subset\ker\eta_{2}. In other words a1akb1bmΓa_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma is a subset of ker(η1)\ker(\eta_{1}) and NN is a subset of ker(η2)\ker(\eta_{2}). Since a1akb1bmΓa_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma is dense in GG it follows that η1\eta_{1} is trivial, and since NN is dense in LL, η2\eta_{2} is also trivial. This contradicts the fact that either η1\eta_{1} or η2\eta_{2} is non-trivial and finishes the proof of Claim 2. ∎

Proof of Claim 3.

Recall that v^h(n)=σ(vh(n))\widehat{v}_{h}(n)=\sigma(v^{\square}_{h}(n)), where vh(n)=(wh(n),1G)v(n)v^{\square}_{h}(n)=(w_{h}(n),1_{G})\,v^{\triangle}(n),

v(n)=a1f1(n)akfk(n)b1p1(n)bmpm(n),n,v(n)=a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)}b_{1}^{p_{1}(n)}\cdot\ldots\cdot b_{m}^{p_{m}(n)},\qquad\forall n\in\mathbb{N},

and

wh(n)=iaiΔhfi(n)1jmbjΔhpj(n).w_{h}(n)\,=\,\prod_{i\in\mathcal{I}}a_{i}^{\Delta_{h}f_{i}(n)}\cdot\prod_{1\leqslant j\leqslant m}b_{j}^{\Delta_{h}p_{j}(n)}.

Let νi\nu_{i} denote the degree of fif_{i}. Using Taylor’s Theorem, we can approximate Δhfi(n)\Delta_{h}f_{i}(n) by

Δhfi(n)=hfi(n)++h(νi1)(νi1)!fi(νi1)(n)+O(f(νi)(n)).\Delta_{h}f_{i}(n)\,=\,hf_{i}^{\prime}(n)+\ldots+\frac{h^{(\nu_{i}-1)}}{(\nu_{i}-1)!}f_{i}^{(\nu_{i}-1)}(n)+{\rm O}\Big{(}f^{(\nu_{i})}(n)\Big{)}.

In view of 2.2, we have O(f(νi)(n))=on(1){\rm O}(f^{(\nu_{i})}(n))={\rm o}_{n\to\infty}(1). Thus,

dG(wh(n),iaihfi(n)++h(νi1)(νi1)!fi(νi1)(n)1jmbjΔhpj(n))=on(1).d_{G}\Big{(}w_{h}(n),\,\prod_{i\in\mathcal{I}}a_{i}^{hf_{i}^{\prime}(n)+\ldots+\frac{h^{(\nu_{i}-1)}}{(\nu_{i}-1)!}f_{i}^{(\nu_{i}-1)}(n)}\cdot\prod_{1\leqslant j\leqslant m}b_{j}^{\Delta_{h}p_{j}(n)}\Big{)}\,=\,{\rm o}_{n\to\infty}(1).

If ii\in\mathcal{I} then νi2\nu_{i}\geqslant 2. Also, according to the hypothesis of F, for every j{1,,νi1}j\in\{1,\ldots,\nu_{i}-1\} there exists z(i,j){1,,k}z(i,j)\in\{1,\ldots,k\} such that fi(j)=fz(i,j)f_{i}^{(j)}=f_{z(i,j)}. For every l{1,,k}l\in\{1,\ldots,k\} define Ql{(i,j):z(i,j)=l}Q_{l}\coloneqq\{(i,j):z(i,j)=l\} and set

u~l(h)(i,j)Qlaihjj!.\tilde{u}_{l}(h)\,\coloneqq\,\prod_{(i,j)\in Q_{l}}a_{i}^{\frac{h^{j}}{j!}}.

Now define, for every i{1,,k}i\in\{1,\ldots,k\}, the polynomial mapping ui:G^u_{i}\colon\mathbb{N}\to\widehat{G}^{\circ} as

ui(h){σ(ai,ai),ifi,σ(aiu~i(h),ai),ifi.u_{i}(h)\,\coloneqq\,\begin{cases}\sigma(a_{i},a_{i}),&\text{if}\leavevmode\nobreak\ i\notin\mathcal{I},\\ \sigma(a_{i}\tilde{u}_{i}(h),a_{i}),&\text{if}\leavevmode\nobreak\ i\in\mathcal{I}.\end{cases}

In a similar way, one can find c,e1,,em:G^c,e_{1},\ldots,e_{m}\colon\mathbb{N}\to\widehat{G}. ∎

Proof of Claim 4.

Fix hh\in\mathbb{N}. Since G={G1,G2,,Gd,Gd+1}G_{\bullet}=\{G_{1},G_{2},\ldots,G_{d},G_{d+1}\} is filtration that realizes the degree of vv, we have bjGdeg(pj)+1b_{j}\in G_{\deg(p_{j})+1} for all j=1,,mj=1,\ldots,m and aiGdeg(fi)a_{i}\in G_{\deg(f_{i})}^{\circ} for all i=1,,ki=1,\ldots,k. Define LiLGi+1L_{i}\coloneqq L\cap G_{i+1}, i=0,1,,di=0,1,\ldots,d, Ld+1{1G}L_{d+1}\coloneqq\{1_{G}\}, and

(G×LG)={G1×L1G1,,Gd×LdGd,Gd+1×Ld+1Gd+1}.(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)_{\bullet}=\big{\{}G_{1}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{1}}G_{1},\ \ldots,\ G_{d}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{d}}G_{d},\ G_{d+1}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{d+1}}G_{d+1}\big{\}}.

According to 4.6, (G×LG)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)_{\bullet} is a dd-step filtration of G×LGG{\mkern-1.0mu\times\mkern-1.5mu}_{L}G. We claim that (G×LG)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)_{\bullet} is a filtration that realizes the degree of vh(n)v_{h}^{\square}(n). Recall that

vh(n)=(wh(n),1G)v(n),v_{h}^{\square}(n)\,=\,(w_{h}(n),1_{G})\,v^{\triangle}(n),

where

v(n)=1ik(ai,ai)fi(n)1jm(bj,bj)pj(n)v^{\triangle}(n)\,=\,\prod_{1\leqslant i\leqslant k}(a_{i},a_{i})^{f_{i}(n)}\cdot\prod_{1\leqslant j\leqslant m}(b_{j},b_{j})^{p_{j}(n)}

and

(wh(n),1G)=i(ai,1G)Δhfi(n)1jm(bj,1G)Δhpj(n).(w_{h}(n),1_{G})\,=\,\prod_{i\in\mathcal{I}}(a_{i},1_{G})^{\Delta_{h}f_{i}(n)}\cdot\prod_{1\leqslant j\leqslant m}(b_{j},1_{G})^{\Delta_{h}p_{j}(n)}.

To show that (G×LG)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)_{\bullet} is a filtration realizing the degree of vh(n)v_{h}^{\square}(n), we must prove that

  1. (i)

    (bj,bj)Gdeg(pj)+1×Ldeg(pj)+1Gdeg(pj)+1(b_{j},b_{j})\in G_{\deg(p_{j})+1}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{\deg(p_{j})+1}}G_{\deg(p_{j})+1} for all j=1,,mj=1,\ldots,m;

  2. (ii)

    (ai,ai)(Gdeg(fi)×Ldeg(fi)Gdeg(fi))(a_{i},a_{i})\in(G_{\deg(f_{i})}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{\deg(f_{i})}}G_{\deg(f_{i})})^{\circ} for all i=1,,ki=1,\ldots,k;

  3. (iii)

    (bj,1G)Gdeg(Δhpj)+1×Ldeg(Δhpj)+1Gdeg(Δhpj)+1(b_{j},1_{G})\in G_{\deg(\Delta_{h}p_{j})+1}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{\deg(\Delta_{h}p_{j})+1}}G_{\deg(\Delta_{h}p_{j})+1} for all j=1,,mj=1,\ldots,m;

  4. (iv)

    (ai,1G)(Gdeg(Δhfi)×Ldeg(Δhfi)Gdeg(Δhfi))(a_{i},1_{G})\in(G_{\deg(\Delta_{h}f_{i})}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{\deg(\Delta_{h}f_{i})}}G_{\deg(\Delta_{h}f_{i})})^{\circ} for all ii\in\mathcal{I};

Parts (i) and (ii) follow from the fact that GiGi×LiGiG_{i}^{\triangle}\subset G_{i}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{i}}G_{i} for all i=1,,di=1,\ldots,d and that bjGdeg(pj)+1b_{j}\in G_{\deg(p_{j})+1} for all j=1,,mj=1,\ldots,m and aiGdeg(fi)a_{i}\in G_{\deg(f_{i})}^{\circ} for all i=1,,ki=1,\ldots,k. Parts (iii) and (iv) follow from the fact that deg(Δhfi)=deg(fi)1\deg(\Delta_{h}f_{i})=\deg(f_{i})-1 and deg(Δhpj)=deg(pj)1\deg(\Delta_{h}p_{j})=\deg(p_{j})-1 and that (LGi+1)×{1G}Gi×LiGi(L\cap G_{i+1})\times\{1_{G}\}\subset G_{i}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{i}}G_{i}.

To complete the proof of Claim 4, note that if (G×LG)(G{\mkern-1.0mu\times\mkern-1.5mu}_{L}G)_{\bullet} is a filtration realizing the degree of vh(n)v_{h}^{\square}(n), then the filtration G^={G^1,G^2,,G^d}\widehat{G}_{\bullet}=\{\widehat{G}_{1},\widehat{G}_{2},\ldots,\widehat{G}_{d}\}, defined as

G^iσ(Gi×LiGi),i=1,,d,\widehat{G}_{i}\coloneqq\sigma\big{(}G_{i}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{i}}G_{i}\big{)},\qquad i=1,\ldots,d,

is a filtration realizing the degree of v^h(n)\widehat{v}_{h}(n). Moreover, Gd×LdGdG_{d}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{d}}G_{d} is equal to GdG_{d}^{\triangle} because Ld={1G}L_{d}=\{1_{G}\}, and hence Gd×LdGdG_{d}{\mkern-1.0mu\times\mkern-1.5mu}_{L_{d}}G_{d} belongs to the kernel of σ\sigma. This shows that G^={G^1,G^2,,G^d}\widehat{G}_{\bullet}=\{\widehat{G}_{1},\widehat{G}_{2},\ldots,\widehat{G}_{d}\} is a (d1)(d-1)-step filtration and hence v^h(n)\widehat{v}_{h}(n) has degree d1d-1. ∎

This finishes the proofs of Claims 1, 2, 3, and 4, which in turn completes the proof of F. ∎

5.   Theorems C and D

For proving Theorems C and D we use essentially the same ideas as were used in the proofs of Theorems A and B, only that all Cesàro averages get replaced with WW-averages. Similar to what we did in Section 3, the first step is to reduce Theorems C and D to the following analogue of F.

Theorem G.

Let GG be a simply connected nilpotent Lie group, Γ\Gamma a uniform and discrete subgroup of GG, and \mathcal{H} a Hardy field. Assume v:Gv\colon\mathbb{N}\to G is a mapping of the form

v(n)=a1f1(n)akfk(n)b1p1(n)bmpm(n),n,v(n)=a_{1}^{f_{1}(n)}\cdot\ldots\cdot a_{k}^{f_{k}(n)}b_{1}^{p_{1}(n)}\cdot\ldots\cdot b_{m}^{p_{m}(n)},\qquad\forall n\in\mathbb{N},

where a1,,akGa_{1},\ldots,a_{k}\in G^{\circ}, b1,,bmGb_{1},\ldots,b_{m}\in G, the elements a1,,ak,b1,,bma_{1},\ldots,a_{k},b_{1},\ldots,b_{m} are pairwise commuting, b1Γ¯,,bmΓ¯\overline{b_{1}^{\mathbb{Z}}\Gamma},\ldots,\overline{b_{m}^{\mathbb{Z}}\Gamma} are connected sub-nilmanifolds of X=G/ΓX=G/\Gamma, p1,,pm[t]p_{1},\ldots,p_{m}\in\mathbb{R}[t] are polynomials satisfying

  1.  (G1)

    pj()p_{j}(\mathbb{Z})\subset\mathbb{Z}, for all j=1,,mj=1,\ldots,m,

  2.  (G2)

    deg(pj)=j\deg(p_{j})=j, for all j=1,,mj=1,\ldots,m,

and f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} satisfy

  1.  (G3)

    f1(t)fk(t)f_{1}(t)\prec\ldots\prec f_{k}(t),

  2.  (G4)

    for all f{f1,,fk}f\in\{f_{1},\ldots,f_{k}\} there exists \ell\in\mathbb{N} such that t1log(W(t))f(t)tt^{\ell-1}\log(W(t))\prec f(t)\prec t^{\ell},

  3.  (G5)

    for all f{f1,,fk}f\in\{f_{1},\ldots,f_{k}\} with deg(f)2\deg(f)\geqslant 2 we have f{f1,,fk}f^{\prime}\in\{f_{1},\ldots,f_{k}\},

where WW\in\mathcal{H} has degree 11. Then (v(n)Γ)n(v(n)\Gamma)_{n\in\mathbb{N}} is uniformly distributed with respect to WW-averages in the sub-nilmanifold a1akb1bmΓ¯\overline{a_{1}^{\mathbb{R}}\cdots a_{k}^{\mathbb{R}}b_{1}^{\mathbb{Z}}\cdots b_{m}^{\mathbb{Z}}\Gamma}.

The proof that G implies Theorems D and C is almost identical to the proof that F implies Theorems A and B given in Section 3. The only difference is that instead of applying A.6 with V(t)=tV(t)=t, we apply A.6 with V(t)=W(t)V(t)=W(t) where WW\in\mathcal{H} is chosen (using A.5) such that f1,,fkf_{1},\ldots,f_{k} satisfy Property (PW). Since these proofs are so similar, we omit the details.

The proof of G is, just like the proof of F, split into three cases: the abelian case, the sub-linear case, and the general case. The proof of the sub-linear case of G is the same as the proof of the sub-linear case of F, except that all Cesàro averages are replaced with WW-averages and instead of utilizing 4.4 one uses 6.1, which is precisely the analogue of 4.4 for WW-averages. Therefore, we omit the details of this part of the proof of G too.

Similarly, the arguments used in the proof of the general case of G are almost identical to the ones used in the proof of the general case of F in Section 4.3 if one replaces all Cesàro averages with WW-averages and instead of applying A.8 with pn=1p_{n}=1 one applies A.8 with pn=w(n)p_{n}=w(n). We omit the details of this part as well.

This leaves only the abelian case of G to be verified. For the proof of this case we can also copy the proof of the abelian case of F given in Section 4.1. The only missing ingredient is a variant of Boshernitzan’s Equidistribution Theorem ([Boshernitzan94, Theorem 1.8]) for WW-averages. Let us formulate and prove such a variant now.

Theorem 5.1.

Let \mathcal{H} be a Hardy field, let W,fW,f\in\mathcal{H}, and assume 1W(t)t1\prec W(t)\ll t and t1log(W(t))f(t)tt^{\ell-1}\log(W(t))\prec f(t)\prec t^{\ell} for some \ell\in\mathbb{N}. Then

limN1W(n)n=1Nw(n)e(f(n))= 0,\lim_{N\to\infty}\frac{1}{W(n)}\sum_{n=1}^{N}w(n)\,e(f(n))\,=\,0, (5.1)

where w=ΔWw=\Delta W.

Although 5.1 for W(t)=tW(t)=t does not imply Boshernitzan’s Equidistribution Theorem in full generality, it is good enough for the proof of the abelian case of G.

Proof of 5.1.

For the proof we use induction on \ell. For the base case of the induction, when =1\ell=1, we use 6.1. In light of 6.1, instead of (5.1) it suffices to show that for every ε>0\varepsilon>0 there exists ξ(0,ε)\xi\in(0,\varepsilon) such that

limHlimN1Hh=1H1W(N)n=1Nw(n)e(f(n)+ξh)e(f(n))¯= 0.\lim_{H\to\infty}\lim_{N\to\infty}\frac{1}{H}\sum_{h=1}^{H}\frac{1}{W(N)}\sum_{n=1}^{N}w(n)e(f(n)+\xi h)\overline{e(f(n))}\leavevmode\nobreak\ =\leavevmode\nobreak\ 0. (5.2)

We can simplify the left hand side of (5.2) to

limHlimN1Hh=1H1W(N)n=1Nw(n)e(f(n)+ξh)e(f(n))¯=limHlimN1Hh=1H1W(N)n=1Nw(n)e(ξh)=limH1Hh=1He(ξh),\begin{split}\lim_{H\to\infty}\lim_{N\to\infty}\frac{1}{H}\sum_{h=1}^{H}\frac{1}{W(N)}&\sum_{n=1}^{N}w(n)e(f(n)+\xi h)\overline{e(f(n))}\\ =&\lim_{H\to\infty}\lim_{N\to\infty}\frac{1}{H}\sum_{h=1}^{H}\frac{1}{W(N)}\sum_{n=1}^{N}w(n)e(\xi h)\\ =&\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}e(\xi h),\end{split} (5.3)

which equals 0 for all ξ\xi\notin\mathbb{Z}.

For the proof of the inductive step, we use A.8. In view of A.8, applied with PN=W(N)P_{N}=W(N) and pn=w(n)p_{n}=w(n), we see that (5.1) holds if we can show

limHlimN1Hh=1H1W(n)n=1Nw(n)e(f(n+h)f(n))= 0.\lim_{H\to\infty}\lim_{N\to\infty}\frac{1}{H}\sum_{h=1}^{H}\frac{1}{W(n)}\sum_{n=1}^{N}w(n)\,e(f(n+h)-f(n))\leavevmode\nobreak\ =\leavevmode\nobreak\ 0. (5.4)

However, since ff satisfied t1log(W(t))f(t)tt^{\ell-1}\log(W(t))\prec f(t)\prec t^{\ell} and 2\ell\geqslant 2, the function Δhf(t)f(t+h)f(t)\Delta_{h}f(t)\coloneqq f(t+h)-f(t) satisfies t2log(W(t))Δhf(t)t1t^{\ell-2}\log(W(t))\prec\Delta_{h}f(t)\prec t^{\ell-1}, which is a consequence of 2.2. Hence (5.4) follows form the induction hypothesis. ∎

6.   A variant of van der Corput’s Lemma

The purpose of this section is to prove the following proposition which was used in the proofs of 4.2 in Section 4.2 and G in Section 5.

Proposition 6.1.

Assume WW and f1,,fkf_{1},\ldots,f_{k} are functions from a Hardy field \mathcal{H} satisfying 1W(t)t1\prec W(t)\ll t and log(W(t))f1(t)fkt\log(W(t))\prec f_{1}(t)\prec\ldots\prec f_{k}\prec t. Let Ψ:k\Psi\colon\mathbb{R}^{k}\to\mathbb{C} be a bounded and uniformly continuous function and suppose for all ss\in\mathbb{R} the limit

A(s)limN1W(N)n=1Nw(n)Ψ(f1(n),,fk1(n),fk(n)+s)Ψ(f1(n),,fk(n))¯A(s)\leavevmode\nobreak\ \coloneqq\leavevmode\nobreak\ \lim_{N\to\infty}\frac{1}{W(N)}\sum_{n=1}^{N}w(n)\Psi(f_{1}(n),\ldots,f_{k-1}(n),f_{k}(n)+s)\overline{\Psi(f_{1}(n),\ldots,f_{k}(n))} (6.1)

exists, where w=ΔWw=\Delta W. If for every ε>0\varepsilon>0 there exists ξ(0,ε)\xi\in(0,\varepsilon) such that

limH1Hh=1HA(ξh)= 0\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}A(\xi h)\leavevmode\nobreak\ =\leavevmode\nobreak\ 0

then necessarily

limN1W(N)n=1Nw(n)Ψ(f1(n),,fk(n))=0.\lim_{N\to\infty}\frac{1}{W(N)}\sum_{n=1}^{N}w(n)\Psi(f_{1}(n),\ldots,f_{k}(n))=0. (6.2)

Note that 4.4 follows from 6.1 by choosing W(t)=tW(t)=t.

The next lemma will be useful for the proof of 6.1. We say a function f:[1,)f\colon[1,\infty)\to\mathbb{R} has sub-exponential growth if |f(t)|ct|f(t)|\prec c^{t} for all c>1c>1.

Lemma 6.2.

Let ff\in\mathcal{H} with log(t)f(t)\log(t)\prec f(t). Then f1f^{-1} has sub-exponential growth.

Proof.

Note that f1(t)f^{-1}(t) has sub-exponential growth if and only if

limtf1(t+1)f1(t)=1.\lim_{t\to\infty}\frac{f^{-1}(t+1)}{f^{-1}(t)}=1.

Let us therefore consider the number c0limnf1(n+1)/f1(n)c_{0}\coloneqq\lim_{n\to\infty}{f^{-1}(n+1)}/{f^{-1}(n)}. Note that this limit exists because if ff belongs to some Hardy Field, then so does f1f^{-1}.

Since f1f^{-1} is eventually increasing, we have c01c_{0}\geqslant 1. It remains to show that c01c_{0}\leqslant 1, which we will do by showing that c1c\leqslant 1 for all c<c0c<c_{0}. Thus, fix any cc with 0<c<c00<c<c_{0}. There exists M>0M>0 such that for all but finitely many nn we have

f1(n)Mcnf^{-1}(n)\geqslant Mc^{n}

and hence, using f(f1(n))=nf(f^{-1}(n))=n, we obtain

f(Mcn)n.f(Mc^{n})\leqslant n.

Since log(t)f(t)\log(t)\prec f(t), we conclude that log(Mcn)n\log(Mc^{n})\prec n and hence c1c\leqslant 1. ∎

Remark 6.3.

It follows from 6.2 that if W,fW,f\in\mathcal{H} with 1W(t)1\prec W(t) and log(W(t))f(t)\log(W(t))\prec f(t) then Wf1W\circ f^{-1} has sub-exponential growth.

Lemma 6.4.

Let W,fW,f\in\mathcal{H} with 1W(t)11\prec W(t)\ll 1 and log(W(t))f(t)t\log(W(t))\prec f(t)\prec t, and define wΔWw\coloneqq\Delta W. For every nn\in\mathbb{N} and ξ(0,1]\xi\in(0,1] define

Kn{j:f(j)(ξn,ξ(n+1)]}.K_{n}\coloneqq\big{\{}j\in\mathbb{N}:f(j)\in\left(\xi n,\xi(n+1)\right]\big{\}}.

Define g(t)ξ1f(t)g(t)\coloneqq\xi^{-1}f(t), pniKnw(i)p_{n}\coloneqq\sum_{i\in K_{n}}w(i), and PNn=1NpnP_{N}\coloneqq\sum_{n=1}^{N}p_{n}. Then the following hold:

  1. (i)

    limnW(g1(n+1))W(g1(n))pn=1\lim_{n\to\infty}\frac{W(g^{-1}(n+1))-W(g^{-1}(n))}{p_{n}}=1.

  2. (ii)

    limnW(g1(n))Pn=1\lim_{n\to\infty}\frac{W(g^{-1}(n))}{P_{n}}=1;

  3. (iii)

    limnPn=\lim_{n\to\infty}P_{n}=\infty;

  4. (iv)

    limnpnPn=0\lim_{n\to\infty}\frac{p_{n}}{P_{n}}=0.

Proof.

Since 1f(t)t1\prec f(t)\prec t and ff is eventually monotone increasing, for sufficiently large nn the set KnK_{n} is an interval of the form [an,bn][a_{n},b_{n}]\subset\mathbb{N}, where limnbnan=\lim_{n\to\infty}b_{n}-a_{n}=\infty. For all such nn we thus also have pn=W(bn+1)W(an)=W(bn)W(an)+on(1)p_{n}=W(b_{n}+1)-W(a_{n})=W(b_{n})-W(a_{n})+{\rm o}_{n\to\infty}(1). Let snmin{t:f(t)ξn}s_{n}\coloneqq\min\{t\in\mathbb{R}:f(t)\geqslant\xi n\}. Then for all but finitely many nn we have that sn=g1(n)s_{n}=g^{-1}(n). Note that the difference between ana_{n} and sns_{n} can be bounded from above by Δf(n)\Delta f(n), and since Δf(n)0\Delta f(n)\to 0 as nn\to\infty, we have that limnansn=0\lim_{n\to\infty}a_{n}-s_{n}=0. Similarly, we can show that limnbnsn+1=0\lim_{n\to\infty}b_{n}-s_{n+1}=0. Therefore limnW(an)W(g1(n))=limnW(bn)W(g1(n+1))=0\lim_{n\to\infty}W(a_{n})-W(g^{-1}(n))=\lim_{n\to\infty}W(b_{n})-W(g^{-1}(n+1))=0 and hence

limnW(g1(n+1))W(g1(n))pn=limnW(g1(n+1))W(g1(n))W(bn)W(an)=1.\lim_{n\to\infty}\frac{W(g^{-1}(n+1))-W(g^{-1}(n))}{p_{n}}=\lim_{n\to\infty}\frac{W(g^{-1}(n+1))-W(g^{-1}(n))}{W(b_{n})-W(a_{n})}=1.

Part (ii) follows straight away from (i) and part (iii) follows immediately from part (ii) and the fact that W(g1(t))W(g^{-1}(t))\to\infty.

For the proof of part (iv) note that

limnpn/Pn=limnW(g1(n+1))W(g1(n))W(g1(n))\lim_{n\to\infty}p_{n}/P_{n}=\lim_{n\to\infty}\frac{W(g^{-1}(n+1))-W(g^{-1}(n))}{W(g^{-1}(n))}

because of parts (i) and (ii). However, limn(g1(n+1)g1(n))/g1(n)=0\lim_{n\to\infty}(g^{-1}(n+1)-g^{-1}(n))/g^{-1}(n)=0 because W(g1(t))W(g^{-1}(t)) has sub-exponential growth due to 6.3 and the fact that log(W(t))g(t)\log(W(t))\prec g(t). ∎

Proof of 6.1.

Fix ξ(0,1]\xi\in(0,1] and define Kj{n:fk(n)(ξ(j1),ξj]}K_{j}\coloneqq\left\{n\in\mathbb{N}:f_{k}(n)\in\left(\xi(j-1),\xi j\right]\right\} and g(t)ξ1fk(t)g(t)\coloneqq\xi^{-1}f_{k}(t). Since

1W(N)n=1Nw(n)Ψ(f1(n),,fk(n))=1W(N)j=1g(N)nKjw(n)Ψ(f1(n),,fk(n))+oN(1),\begin{split}\frac{1}{W(N)}\sum_{n=1}^{N}&w(n)\Psi(f_{1}(n),\ldots,f_{k}(n))\\ &=\frac{1}{W(N)}\sum_{j=1}^{\lfloor g(N)\rfloor}\sum_{n\in K_{j}}w(n)\Psi(f_{1}(n),\ldots,f_{k}(n))\,+\,{\rm o}_{N\to\infty}(1),\end{split}

instead of (6.2) is suffices to show that

limN1W(N)j=1g(N)nKjw(n)Ψ(f1(n),,fk(n))= 0.\lim_{N\to\infty}\frac{1}{W(N)}\sum_{j=1}^{\lfloor g(N)\rfloor}\sum_{n\in K_{j}}w(n)\Psi(f_{1}(n),\ldots,f_{k}(n))\leavevmode\nobreak\ =\leavevmode\nobreak\ 0. (6.3)

Set pjnKjw(n)p_{j}\coloneqq\sum_{n\in K_{j}}w(n) and PJj=1JpjP_{J}\coloneqq\sum_{j=1}^{J}p_{j}. According to 6.4, part (ii), we have

limNW(N)Pg(N)= 1.\lim_{N\to\infty}\frac{W(N)}{P_{\lfloor g(N)\rfloor}}\leavevmode\nobreak\ =\leavevmode\nobreak\ 1.

Therefore, (6.3) is equivalent to

limN1Pg(N)j=1g(N)nKjw(n)Ψ(f1(n),,fk(n))= 0.\lim_{N\to\infty}\frac{1}{P_{\lfloor g(N)\rfloor}}\sum_{j=1}^{\lfloor g(N)\rfloor}\sum_{n\in K_{j}}w(n)\Psi(f_{1}(n),\ldots,f_{k}(n))\leavevmode\nobreak\ =\leavevmode\nobreak\ 0.

which we can write as

limJ1PJj=1JnKjw(n)Ψ(f1(n),,fk(n))= 0.\lim_{J\to\infty}\frac{1}{P_{J}}\sum_{j=1}^{J}\sum_{n\in K_{j}}w(n)\Psi(f_{1}(n),\ldots,f_{k}(n))\leavevmode\nobreak\ =\leavevmode\nobreak\ 0. (6.4)

Define gi(n)fi(g1(n))g_{i}(n)\coloneqq f_{i}(g^{-1}(n)) for i=1,,ki=1,\ldots,k and note that gk(n)=ξng_{k}(n)=\xi n. Then supnKj|gk(j)fk(n)|=O(ξ)\sup_{n\in K_{j}}|g_{k}(j)-f_{k}(n)|={\rm O}(\xi) and, for all i<ki<k, we have supnKj|gi(j)fi(n)|=on(1)\sup_{n\in K_{j}}|g_{i}(j)-f_{i}(n)|={\rm o}_{n\to\infty}(1). Therefore, using the uniform continuity of Ψ\Psi, we have that

nKjw(n)Ψ(f1(n),,fk(n))=pjΨ(g1(j),,gk(j))+oj,ξ0(pj).\sum_{n\in K_{j}}w(n)\Psi(f_{1}(n),\ldots,f_{k}(n))=p_{j}\Psi\big{(}g_{1}(j),\ldots,g_{k}(j)\big{)}+{\rm o}_{j\to\infty,\xi\to 0}(p_{j}).

It follows that

1PJj=1JnKjw(n)Ψ(f1(n),,fk(n))=1PJj=1JpjΨ(g1(j),,gk(j))+oJ,ξ0(1).\frac{1}{P_{J}}\sum_{j=1}^{J}\,\sum_{n\in K_{j}}w(n)\Psi(f_{1}(n),\ldots,f_{k}(n))=\frac{1}{P_{J}}\sum_{j=1}^{J}p_{j}\Psi(g_{1}(j),\ldots,g_{k}(j))+{\rm o}_{J\to\infty,\xi\to 0}(1).

In conclusion, (6.4), and therefore also (6.2), are equivalent to

limJ1PJj=1JpjΨ(g1(j),,gk(j))=oξ0(1).\lim_{J\to\infty}\frac{1}{P_{J}}\sum_{j=1}^{J}p_{j}\Psi(g_{1}(j),\ldots,g_{k}(j))\leavevmode\nobreak\ =\leavevmode\nobreak\ {\rm o}_{\xi\to 0}(1). (6.5)

Using essentially the same argument one can also show that A(s)A(s), which was defined in (6.1), is given by

A(s)=limJ1PJj=1JpjΨ(g1(j),,gk1(j),gk(j)+s)Ψ(g1(j),,gk(j))¯.A(s)\,=\,\lim_{J\to\infty}\frac{1}{P_{J}}\sum_{j=1}^{J}p_{j}\Psi(g_{1}(j),\ldots,g_{k-1}(j),g_{k}(j)+s)\overline{\Psi(g_{1}(j),\ldots,g_{k}(j))}. (6.6)

According to A.8, instead of (6.5), it is enough to prove that

limH1Hh=1H(limJ1PJj=1JpjΨ(g1(j+h),,gk(j+h))Ψ(g1(j),,gk(j))¯)=oξ0(1).\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}\left(\lim_{J\to\infty}\frac{1}{P_{J}}\sum_{j=1}^{J}p_{j}\Psi(g_{1}(j+h),\ldots,g_{k}(j+h))\overline{\Psi(g_{1}(j),\ldots,g_{k}(j))}\right)={\rm o}_{\xi\to 0}(1). (6.7)

For i<ki<k we have gi(t)tg_{i}(t)\prec t and hence gi(j+h)=gi(j)+oj(1)g_{i}(j+h)=g_{i}(j)+{\rm o}_{j\to\infty}(1). For i=ki=k we have g(j+h)=g(j)+ξg(j+h)=g(j)+\xi. Therefore the left hand side of (6.7) can be replaced with

limH1Hh=1H(limJ1PJj=1JpjΨ(g1(j),,gk1(j),gk(j)+ξh)Ψ(g1(j),,gk(j))¯),\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}\left(\lim_{J\to\infty}\frac{1}{P_{J}}\sum_{j=1}^{J}p_{j}\Psi(g_{1}(j),\ldots,g_{k-1}(j),g_{k}(j)+\xi h)\overline{\Psi(g_{1}(j),\ldots,g_{k}(j))}\right),

which in combination with (6.6) shows that (6.7) is equivalent to

limH1Hh=1HA(ξh)=oξ0(1).\lim_{H\to\infty}\frac{1}{H}\sum_{h=1}^{H}A(\xi h)={\rm o}_{\xi\to 0}(1).

This finishes the proof. ∎

Appendix A Appendix

A.1.   Some basic results regarding functions form a Hardy field

Lemma A.1.

Let \mathcal{H} be a Hardy field and let ff\in\mathcal{H} be of polynomial growth. If f()f(\mathbb{N})\subset\mathbb{Z} then f[t]f\in\mathbb{R}[t].

Proof.

For kk\in\mathbb{N}, let Δkf\Delta^{k}f denote the kk-fold finite difference of ff, that is, Δ0f(n)=f(n)\Delta^{0}f(n)=f(n), Δ1f(n)=Δf(n)=f(n+1)f(n)\Delta^{1}f(n)=\Delta f(n)=f(n+1)-f(n), Δ2f(n)=Δf(n+1)Δf(n)=f(n+2)2f(n+1)+f(n)\Delta^{2}f(n)=\Delta f(n+1)-\Delta f(n)=f(n+2)-2f(n+1)+f(n), and so on. If dd is the degree of ff then the function Δdf\Delta^{d}f has degree 0. Moreover, since f()f(\mathbb{N})\subset\mathbb{Z}, we have Δdf(n)\Delta^{d}f(n)\in\mathbb{Z} for all nn\in\mathbb{N}. However, the only function from a Hardy field that has degree 0 and for which all its values belong to \mathbb{Z} is a constant function. That means that Δdf(n)=c\Delta^{d}f(n)=c for some cc\in\mathbb{Z}, which implies that ff is a polynomial of degree dd satisfying f()f(\mathbb{N})\subset\mathbb{Z}. ∎

Lemma A.2.

Let GG be a simply connected nilpotent Lie group, aGa\in G, \mathcal{H} a Hardy field, and ff\in\mathcal{H} of polynomial growth. If f()𝖽𝗈𝗆(a)f(\mathbb{N})\subset\mathsf{dom}(a) then one of the following two cases holds:

  1. (i)

    either aGa\in G^{\circ};

  2. (ii)

    or there exist mm\in\mathbb{N} and p[t]p\in\mathbb{R}[t] with p()p(\mathbb{Z})\subset\mathbb{Z} such that 1m𝖽𝗈𝗆(a)\frac{1}{m}\in\mathsf{dom}(a) and f(n)=p(n)/mf(n)=p(n)/m for all nn\in\mathbb{N}.

Proof.

Suppose aa is not an element of GG^{\circ}. This means there exists mm\in\mathbb{N} such that 𝖽𝗈𝗆(a)=1m\mathsf{dom}(a)=\frac{1}{m}\mathbb{Z}. In particular, f()1mf(\mathbb{N})\subset\frac{1}{m}\mathbb{Z}. By A.1 the function p(n)mf(n)p(n)\coloneqq mf(n) is polynomial with p()p(\mathbb{N})\subset\mathbb{Z}. This finishes the proof. ∎

Define 𝒮(f1,,fk)={λ1f1++λkfk+p:λ1,,λk,p[t]}\mathcal{S}(f_{1},\ldots,f_{k})=\{\lambda_{1}f_{1}+\ldots+\lambda_{k}f_{k}+p:\lambda_{1},\ldots,\lambda_{k}\in\mathbb{R},\leavevmode\nobreak\ p\in\mathbb{R}[t]\}.

Lemma A.3.

Let \mathcal{H} be a Hardy field and assume f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} have polynomial growth. Then there exist mm\in\mathbb{N}, g1,,gm𝒮(f1,,fk)g_{1},\ldots,g_{m}\in\mathcal{S}(f_{1},\ldots,f_{k}), p1,,pk[t]p_{1},\ldots,p_{k}\in\mathbb{R}[t], and λ1,1,,λk,m\lambda_{1,1},\ldots,\lambda_{k,m}\in\mathbb{R} with the following properties:

  1. (1)

    g1(t)gm(t)g_{1}(t)\prec\ldots\prec g_{m}(t);

  2. (2)

    for all g{g1,,gm}g\in\{g_{1},\ldots,g_{m}\} either g=0g=0 or there exists \ell\in\mathbb{N} such that t1g(t)tt^{\ell-1}\prec g(t)\prec t^{\ell};

  3. (3)

    for all i{1,,k}i\in\{1,\ldots,k\},

    limt|fi(t)j=1mλi,jgj(t)pi(t)|=0.\lim_{t\to\infty}\Bigg{|}f_{i}(t)-\sum_{j=1}^{m}\lambda_{i,j}g_{j}(t)-p_{i}(t)\Bigg{|}=0.
Proof.

Let us associate to every finite set of functions h1,,hrh_{1},\ldots,h_{r}\in\mathcal{H} of polynomial growth a pair (d,e)({0})×(d,e)\in(\mathbb{N}\cup\{0\})\times\mathbb{N}, which we will call the characteristic pair associated to {h1,,hr}\{h_{1},\ldots,h_{r}\}, in the following way: The number dd is the maximal degree among degrees of functions in {h1,,hr}\{h_{1},\ldots,h_{r}\}, i.e.,

d=max{deg(hi):1ir},d\,=\,\max\big{\{}\deg(h_{i}):1\leqslant i\leqslant r\big{\}},

and the number ee equals the number of functions in {h1,,hr}\{h_{1},\ldots,h_{r}\} whose degree is dd, i.e.,

e=|{i{1,,r}:deg(hi)=d}|.e\,=\,\big{|}\big{\{}i\in\{1,\ldots,r\}:\deg(h_{i})=d\big{\}}\big{|}.

Using this notion of a characteristic pair, we can define a partial ordering on the set of finite subsets of functions in \mathcal{H} of polynomial growth: Given h1,,hr,h1,,hrh_{1},\ldots,h_{r},{h}^{*}_{1},\ldots,{h}^{*}_{{r}^{*}}\in\mathcal{H} of polynomial growth we write {h1,,hr}{h1,,hr}\{h_{1},\ldots,h_{r}\}\prec\{{h}^{*}_{1},\ldots,{h}^{*}_{{r}^{*}}\} if

  • either d<dd<{d}^{*},

  • or d=dd={d}^{*} and e<ee<{e}^{*},

where (d,e)(d,e), (d,e)({d}^{*},{e}^{*}) are the characteristic pairs associated to {h1,,hr}\{h_{1},\ldots,h_{r}\} and {h1,,hr}\{{h}^{*}_{1},\ldots,{h}^{*}_{{r}^{*}}\} respectively.

Recall, our goal is to show for any f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} of polynomial growth there exist mm\in\mathbb{N}, g1,,gm𝒮(f1,,fk)g_{1},\ldots,g_{m}\in\mathcal{S}(f_{1},\ldots,f_{k}), p1,,pk[t]p_{1},\ldots,p_{k}\in\mathbb{R}[t], and λ1,1,,λk,m\lambda_{1,1},\ldots,\lambda_{k,m}\in\mathbb{R} such that properties (1), (2), and (3) are satisfied. To accomplish this goal, we will use induction on the just defined partial ordering.

The base case of this induction corresponds to (d,e)=(0,e)(d,e)=(0,e) for some ee\in\mathbb{N}. In this case we have k=ek=e. Let m1m\coloneqq 1, cilimtfi(t)c_{i}\coloneqq\lim_{t\to\infty}f_{i}(t), g1(t)=0g_{1}(t)=0, pi(t)cip_{i}(t)\coloneqq c_{i}, and λi,10\lambda_{i,1}\coloneqq 0. With this choice, (1), (2), and (3) are satisfied, and we are done.

Next, suppose we are in the case when the characteristic pair associated to {f1,,fk}\{f_{1},\ldots,f_{k}\} is of the form (d,e)(d,e) and d1d\geqslant 1. Define σilimtfi(t)/td\sigma_{i}\coloneqq\lim_{t\to\infty}f_{i}(t)/t^{d} and set

hi(t)fi(t)σitd,i=1,,k.h_{i}(t)\coloneqq f_{i}(t)-\sigma_{i}t^{d},\qquad i=1,\ldots,k.

This yields a new collection of functions {h1,,hk}\{h_{1},\ldots,h_{k}\} with the property that hi(t)tdh_{i}(t)\prec t^{d} for all i=1,,ki=1,\ldots,k. We now distinguish two cases, the case when the characteristic pair of {h1,,hk}\{h_{1},\ldots,h_{k}\} is the same as the characteristic pair of {f1,,fk}\{f_{1},\ldots,f_{k}\}, and the case when {h1,,hk}{f1,,fk}\{h_{1},\ldots,h_{k}\}\prec\{f_{1},\ldots,f_{k}\}.

If we are in the first case then there exists some function in {h1,,hk}\{h_{1},\ldots,h_{k}\} of degree dd. By relabeling h1,,hkh_{1},\ldots,h_{k} if necessary, we can assume without loss of generality that h1hkh_{1}\ll\ldots\ll h_{k}. Then hkh_{k} has degree dd. Define ηilimthi(t)/hk(t)\eta_{i}\coloneqq\lim_{t\to\infty}h_{i}(t)/h_{k}(t) and set hihiηihk{h}^{*}_{i}\coloneqq h_{i}-\eta_{i}h_{k} for all i=1,,k1i=1,\ldots,k-1. It is straightforward to check that h1,,hk1{h}^{*}_{1},\ldots,{h}^{*}_{k-1}\in\mathcal{H} satisfies {h1,,hk1}{h1,,hk}\{{h}^{*}_{1},\ldots,{h}^{*}_{k-1}\}\prec\{h_{1},\ldots,h_{k}\}. Therefore, by the induction hypothesis, we can find m{m}^{*}\in\mathbb{N}, g1,,gm𝒮(h1,,hk1){g}^{*}_{1},\ldots,{g}^{*}_{{m}^{*}}\in\mathcal{S}({h}^{*}_{1},\ldots,{h}^{*}_{k-1}), p1,,pk1[t]{p}^{*}_{1},\ldots,{p}^{*}_{k-1}\in\mathbb{R}[t], and λ1,1,,λk1,m{\lambda}^{*}_{1,1},\ldots,{\lambda}^{*}_{k-1,{m}^{*}}\in\mathbb{R} such that g1,,gm{g}^{*}_{1},\ldots,{g}^{*}_{{m}^{*}} satisfy properties (1) and (2), and for all i{1,,k1}i\in\{1,\ldots,k-1\} we have

limt|hi(t)j=1mλi,jgj(t)pi(t)|=0.\lim_{t\to\infty}\Bigg{|}{h}^{*}_{i}(t)-\sum_{j=1}^{{m}^{*}}{\lambda}^{*}_{i,j}{g}^{*}_{j}(t)-{p}^{*}_{i}(t)\Bigg{|}=0.

Define m=m+1m={m}^{*}+1, let pipi+σitdp_{i}\coloneqq{p}^{*}_{i}+\sigma_{i}t^{d}, and set

λi,j={λi,j,ifi<kandj<mηi,ifi<kandj=m0,ifi=kandj<m1,ifi=kandj=m,andgj={gj,ifj<mhk,ifj=m.\lambda_{i,j}=\begin{cases}{\lambda}^{*}_{i,j},&\text{if}\leavevmode\nobreak\ i<k\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ j<m\\ \eta_{i},&\text{if}\leavevmode\nobreak\ i<k\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ j=m\\ 0,&\text{if}\leavevmode\nobreak\ i=k\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ j<m\\ 1,&\text{if}\leavevmode\nobreak\ i=k\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ j=m\end{cases},\quad\text{and}\quad g_{j}=\begin{cases}{g}^{*}_{j},&\text{if}\leavevmode\nobreak\ j<m\\ h_{k},&\text{if}\leavevmode\nobreak\ j=m\end{cases}.

Then, for i{1,,k1}i\in\{1,\ldots,k-1\} we have

fi(t)\displaystyle f_{i}(t) =\displaystyle= hi(t)+ηihk(t)+σitd+ot(1)\displaystyle{h}^{*}_{i}(t)+\eta_{i}h_{k}(t)+\sigma_{i}t^{d}+{\rm o}_{t\to\infty}(1)
=\displaystyle= j=1mλi,jgj(t)+pi(t)+ηihk(t)+σitd+ot(1)\displaystyle\sum_{j=1}^{{m}^{*}}{\lambda}^{*}_{i,j}{g}^{*}_{j}(t)+{p}^{*}_{i}(t)+\eta_{i}h_{k}(t)+\sigma_{i}t^{d}+{\rm o}_{t\to\infty}(1)
=\displaystyle= j=1m1λi,jgj(t)+λi,mgm(t)+pi(t)+ot(1)\displaystyle\sum_{j=1}^{m-1}\lambda_{i,j}g_{j}(t)+\lambda_{i,m}g_{m}(t)+p_{i}(t)+{\rm o}_{t\to\infty}(1)
=\displaystyle= j=1mλi,jgj(t)+pi(t)+ot(1).\displaystyle\sum_{j=1}^{m}\lambda_{i,j}g_{j}(t)+p_{i}(t)+{\rm o}_{t\to\infty}(1).

For i=ki=k we have

fk(t)=hk(t)+σitd+ot(1)=j=1mλk,jgj(t)+pk(t)+ot(1).f_{k}(t)\,=\,h_{k}(t)+\sigma_{i}t^{d}+{\rm o}_{t\to\infty}(1)\,=\,\sum_{j=1}^{m}\lambda_{k,j}g_{j}(t)+p_{k}(t)+{\rm o}_{t\to\infty}(1).

This shows that property (3) is satisfied. Property (2) holds by construction (and the fact that gm=hkg_{m}=h_{k} has degree dd but satisfies also gm(t)tdg_{m}(t)\prec t^{d}) and property (1) holds because hkh_{k} grows faster than any hi{h}^{*}_{i} for i=1,,k1i=1,\ldots,k-1, which implies that gmg_{m} also grows faster than any gjg_{j}, 1jm11\leqslant j\leqslant m-1.

It remains to deal with the second case, i.e., the case when {h1,,hk}{f1,,fk}\{h_{1},\ldots,h_{k}\}\prec\{f_{1},\ldots,f_{k}\}. By the induction hypothesis, we can find mm\in\mathbb{N}, g1,,gm𝒮(h1,,hk)g_{1},\ldots,g_{m}\in\mathcal{S}(h_{1},\ldots,h_{k}), p1,,pk[t]p_{1}^{*},\ldots,p_{k}^{*}\in\mathbb{R}[t], and λ1,1,,λk,m\lambda_{1,1},\ldots,\lambda_{k,m}\in\mathbb{R} such that g1,,gmg_{1},\ldots,g_{m} satisfy properties (1), (2), and for all i{1,,k}i\in\{1,\ldots,k\},

limt|hi(t)j=1mλi,jgj(t)pi(t)|=0.\lim_{t\to\infty}\Bigg{|}h_{i}(t)-\sum_{j=1}^{m}\lambda_{i,j}g_{j}(t)-p_{i}^{*}(t)\Bigg{|}=0.

Note that 𝒮(h1,,hk)=𝒮(f1,,fk)\mathcal{S}(h_{1},\ldots,h_{k})=\mathcal{S}(f_{1},\ldots,f_{k}) and hence g1,,gm𝒮(f1,,fk)g_{1},\ldots,g_{m}\in\mathcal{S}(f_{1},\ldots,f_{k}). Moreover, if we take pi(t)pi(t)+σitdp_{i}(t)\coloneqq p_{i}^{*}(t)+\sigma_{i}t^{d} then we get

limt|fi(t)j=1mλi,jgj(t)pi(t)|=0\lim_{t\to\infty}\Bigg{|}f_{i}(t)-\sum_{j=1}^{m}\lambda_{i,j}g_{j}(t)-p_{i}(t)\Bigg{|}=0

as desired. ∎

Define

𝒮(f1,,fk)={λ1f1(n1)++λkfk(nk)+p:λ1,,λk,p[t],n1,,nk{0}}.\mathcal{S}^{*}(f_{1},\ldots,f_{k})=\{\lambda_{1}f_{1}^{(n_{1})}+\ldots+\lambda_{k}f_{k}^{(n_{k})}+p:\lambda_{1},\ldots,\lambda_{k}\in\mathbb{R},\leavevmode\nobreak\ p\in\mathbb{R}[t],\leavevmode\nobreak\ n_{1},\ldots,n_{k}\in\mathbb{N}\cup\{0\}\}.
Lemma A.4.

Let \mathcal{H} be a Hardy field and assume f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} have polynomial growth. Then there exists mm\in\mathbb{N}, g1,,gm𝒮(f1,,fk)g_{1},\ldots,g_{m}\in\mathcal{S}^{*}(f_{1},\ldots,f_{k}), p1,,pk[t]p_{1},\ldots,p_{k}\in\mathbb{R}[t], and λ1,1,,λk,m\lambda_{1,1},\ldots,\lambda_{k,m}\in\mathbb{R} with the following properties:

  1. (1)

    g1(t)gm(t)g_{1}(t)\prec\ldots\prec g_{m}(t);

  2. (2)

    for all g{g1,,gm}g\in\{g_{1},\ldots,g_{m}\} either g=0g=0 or there exists \ell\in\mathbb{N} such that t1g(t)tt^{\ell-1}\prec g(t)\prec t^{\ell};

  3. (3)

    For all g{g1,,gm}g\in\{g_{1},\ldots,g_{m}\} with 2deg(g)deg(gm)2\leqslant\deg(g)\leqslant\deg(g_{m}) we have g{g1,,gm}g^{\prime}\in\{g_{1},\ldots,g_{m}\} and for all g{g1,,gm}g\in\{g_{1},\ldots,g_{m}\} with 1deg(g)deg(gm)11\leqslant\deg(g)\leqslant\deg(g_{m})-1 there exists an antiderivative of gg in {g1,,gm}\{g_{1},\ldots,g_{m}\};

  4. (4)

    for all i{1,,k}i\in\{1,\ldots,k\},

    limt|fi(t)j=1mλi,jgj(t)pi(t)|=0.\lim_{t\to\infty}\Bigg{|}f_{i}(t)-\sum_{j=1}^{m}\lambda_{i,j}g_{j}(t)-p_{i}(t)\Bigg{|}=0.

In the proof of A.3 we associated to every finite set of functions h1,,hrh_{1},\ldots,h_{r}\in\mathcal{H} of polynomial growth a pair (d,e)({0})×(d,e)\in(\mathbb{N}\cup\{0\})\times\mathbb{N}, called the characteristic pair, which gave rise to a partial ordering \prec on the set of finite subsets of functions from \mathcal{H} of polynomial growth. For the proof of A.4 we shall use inductions on the same partial ordering.

Proof of A.4.

If the characteristic pair associated to {f1,,fk}\{f_{1},\ldots,f_{k}\} is either of the from (0,e)(0,e) or (1,e)(1,e) for some ee\in\mathbb{N} then the conclusion of A.4 follows from A.3. Let us therefore assume that the characteristic pair associated to {f1,,fk}\{f_{1},\ldots,f_{k}\} is (d,e)(d,e) with d2d\geqslant 2 and that A.4 has already been proven for all f1,,fk{f}^{*}_{1},\ldots,{f}^{*}_{{k}^{*}}\in\mathcal{H} satisfying {f1,,fk}{f1,,fk}\{{f}^{*}_{1},\ldots,{f}^{*}_{{k}^{*}}\}\prec\{f_{1},\ldots,f_{k}\}.

By replacing fif_{i} with fi-f_{i} if necessary, we can assume without loss of generality that all functions in f1,,fkf_{1},\ldots,f_{k} are eventually non-negative. Also, since functions from a Hardy field can always be reordered according to their growth, we can relabel f1,,fkf_{1},\ldots,f_{k} such that f1f2fkf_{1}\ll f_{2}\ll\ldots\ll f_{k}. Define ηilimtfi(t)/fk(t)\eta_{i}\coloneqq\lim_{t\to\infty}f_{i}(t)/f_{k}(t). Set kk{k}^{*}\coloneqq k, define fifiηifk{f}^{*}_{i}\coloneqq f_{i}-\eta_{i}f_{k} for all i=1,,k1i=1,\ldots,{k}^{*}-1 and fkfk{f}^{*}_{{k}^{*}}\coloneqq f_{k}^{\prime}. It is straightforward to check that f1,,fk{f}^{*}_{1},\ldots,{f}^{*}_{{k}^{*}}\in\mathcal{H} satisfies {f1,,fk}{f1,,fk}\{{f}^{*}_{1},\ldots,{f}^{*}_{{k}^{*}}\}\prec\{f_{1},\ldots,f_{k}\}. By the induction hypothesis, we can find m{m}^{*}\in\mathbb{N}, g1,,gm{g}^{*}_{1},\ldots,{g}^{*}_{{m}^{*}}\in\mathcal{H}, p1,,pk[t]{p}^{*}_{1},\ldots,{p}^{*}_{{k}^{*}}\in\mathbb{R}[t], and λ1,1,,λk,m{\lambda}^{*}_{1,1},\ldots,{\lambda}^{*}_{{k}^{*},{m}^{*}}\in\mathbb{R} such that g1,,gm{g}^{*}_{1},\ldots,{g}^{*}_{{m}^{*}} satisfy properties (1), (2), and (3), and for all i{1,,k}i\in\{1,\ldots,{k}^{*}\} we have

limt|fi(t)j=1mλi,jgj(t)pi(t)|=0.\lim_{t\to\infty}\Bigg{|}{f}^{*}_{i}(t)-\sum_{j=1}^{{m}^{*}}{\lambda}^{*}_{i,j}{g}^{*}_{j}(t)-{p}^{*}_{i}(t)\Bigg{|}=0.

Next, set gj(t)gj(t)g_{j}(t)\coloneqq{g}^{*}_{j}(t) for all j{1,,m}j\in\{1,\ldots,{m}^{*}\} and pi=pip_{i}={p_{i}}^{*} for all i{1,,k1}={1,,k1}i\in\{1,\ldots,{k}^{*}-1\}=\{1,\ldots,k-1\}. Let j0j_{0} be the number in {1,,m}\{1,\ldots,{m}^{*}\} uniquely determined by the property that gjg_{j} has an antiderivative in {g1,,gm}\{g_{1},\ldots,g_{{m}^{*}}\} if jj0j\leqslant j_{0}, and has no antiderivative in {g1,,gm}\{g_{1},\ldots,g_{{m}^{*}}\} if j>j0j>j_{0}. In other words, j0j_{0} is the largest number in {1,,m}\{1,\ldots,{m}^{*}\} for which deg(gj0)=deg(gm)1\deg(g_{j_{0}})=\deg(g_{{m}^{*}})-1. Then, for every jj0j\leqslant j_{0}, let β(j)\beta(j) be the number in {j+1,j+2,,m}\{j+1,j+2,\ldots,{m}^{*}\} such that the derivative of gβ(j)g_{\beta(j)} equals gjg_{j}. Define m2mj0m\coloneqq 2{m}^{*}-j_{0} and, for every j{m+1,,m}j\in\{{m}^{*}+1,\ldots,m\}, let gjg_{j} be any antiderivative of gj0+jmg_{j_{0}+j-{m}^{*}}. Let 𝒥{β(j):1jj0}\mathcal{J}\coloneqq\{\beta(j):1\leqslant j\leqslant j_{0}\} and take

λi,j={λk,β1(j),ifi=kandj𝒥0,ifi=kandj{1,,m}\𝒥,λk,j+j0m,ifi=kandm<jmλi,j+ηiλk,j,ifi<kandjmηiλk,j,ifi<kandm<jm.\lambda_{i,j}=\begin{cases}{\lambda}^{*}_{k,\beta^{-1}(j)},&\text{if}\leavevmode\nobreak\ i=k\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ j\in\mathcal{J}\\ 0,&\text{if}\leavevmode\nobreak\ i=k\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ j\in\{1,\ldots,{m}^{*}\}\backslash\mathcal{J},\\ {\lambda}^{*}_{k,j+j_{0}-{m}^{*}},&\text{if}\leavevmode\nobreak\ i=k\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ {m}^{*}<j\leqslant m\\ {\lambda}^{*}_{i,j}+\eta_{i}\lambda_{k,j},&\text{if}\leavevmode\nobreak\ i<k\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ j\leqslant{m}^{*}\\ \eta_{i}\lambda_{k,j},&\text{if}\leavevmode\nobreak\ i<k\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ {m}^{*}<j\leqslant m\end{cases}.

Also, let pkp_{k} be any polynomial with the property that it is an antiderivative of pk{p}^{*}_{k}. Since fk(t)=fk(t)f_{k}^{\prime}(t)={f}^{*}_{{k}^{*}}(t) and k=k{k}^{*}=k, we can write

fk(t)=j=1mλk,jgj(t)+pk(t)+ot(1)=j=1mλk,jgj(t)+pk(t)+ot(1).f_{k}^{\prime}(t)\,=\,\sum_{j=1}^{{m}^{*}}{\lambda}^{*}_{{k}^{*},j}\,{g}^{*}_{j}(t)+{p}^{*}_{{k}^{*}}(t)+{\rm o}_{t\to\infty}(1)\,=\,\sum_{j=1}^{{m}^{*}}{\lambda}^{*}_{k,j}\,g_{j}(t)+{p}^{*}_{k}(t)+{\rm o}_{t\to\infty}(1).

By integrating we get

fk(t)=j=1j0λk,jgβ(j)(t)+j=j0+1mλk,jgm+jj0(t)+c+pk(t)+ot(1),f_{k}(t)\,=\,\sum_{j=1}^{j_{0}}{\lambda}^{*}_{k,j}\,g_{\beta(j)}(t)+\sum_{j=j_{0}+1}^{{m}^{*}}{\lambda}^{*}_{k,j}\,g_{{m}^{*}+j-j_{0}}(t)+c+p_{k}(t)+{\rm o}_{t\to\infty}(1), (A.1)

where cc is some real constant. We can absorb cc into pkp_{k}, since pkp_{k} was chosen to be an arbitrary antiderivative of pk{p}^{*}_{k}. Thus, (A.1) becomes

fk(t)\displaystyle f_{k}(t) =\displaystyle= j=1j0λk,jgβ(j)(t)+j=j0+1mλk,jgm+jj0(t)+pk(t)+ot(1)\displaystyle\sum_{j=1}^{j_{0}}{\lambda}^{*}_{k,j}\,g_{\beta(j)}(t)+\sum_{j=j_{0}+1}^{{m}^{*}}{\lambda}^{*}_{k,j}\,g_{{m}^{*}+j-j_{0}}(t)+p_{k}(t)+{\rm o}_{t\to\infty}(1)
=\displaystyle= j=1j0λk,jgβ(j)(t)+j=m+1mλk,j+j0mgj(t)+pk(t)+ot(1)\displaystyle\sum_{j=1}^{j_{0}}{\lambda}^{*}_{k,j}\,g_{\beta(j)}(t)+\sum_{j={m}^{*}+1}^{m}{\lambda}^{*}_{k,j+j_{0}-{m}^{*}}\,g_{j}(t)+p_{k}(t)+{\rm o}_{t\to\infty}(1)
=\displaystyle= j𝒥λk,β1(j)gj(t)+j=m+1mλk,jgj(t)+pk(t)+ot(1)\displaystyle\sum_{j\in\mathcal{J}}{\lambda}^{*}_{k,\beta^{-1}(j)}\,g_{j}(t)+\sum_{j={m}^{*}+1}^{m}\lambda_{k,j}\,g_{j}(t)+p_{k}(t)+{\rm o}_{t\to\infty}(1)
=\displaystyle= j=1mλk,jgj(t)+pk(t)+ot(1).\displaystyle\sum_{j=1}^{m}\lambda_{k,j}\,g_{j}(t)+p_{k}(t)+{\rm o}_{t\to\infty}(1).

For i{1,,k1}i\in\{1,\ldots,k-1\} we have

fi(t)\displaystyle f_{i}(t) =\displaystyle= fi(t)+ηifk(t)\displaystyle{f}^{*}_{i}(t)+\eta_{i}f_{k}(t)
=\displaystyle= j=1mλi,jgj(t)+pi(t)+ηifk(t)+ot(1)\displaystyle\sum_{j=1}^{{m}^{*}}{\lambda}^{*}_{i,j}{g}^{*}_{j}(t)+{p}^{*}_{i}(t)+\eta_{i}f_{k}(t)+{\rm o}_{t\to\infty}(1)
=\displaystyle= j=1mλi,jgj(t)+ηij=1mλk,jgj(t)+pi(t)+ot(1)\displaystyle\sum_{j=1}^{{m}^{*}}{\lambda}^{*}_{i,j}g_{j}(t)+\eta_{i}\sum_{j=1}^{m}\lambda_{k,j}\,g_{j}(t)+p_{i}(t)+{\rm o}_{t\to\infty}(1)
=\displaystyle= j=1m(λi,j+ηiλk,j)gj(t)+j=m+1mηiλk,jgj(t)+pi(t)+ot(1)\displaystyle\sum_{j=1}^{{m}^{*}}({\lambda}^{*}_{i,j}+\eta_{i}\lambda_{k,j})g_{j}(t)+\sum_{j={m}^{*}+1}^{m}\eta_{i}\lambda_{k,j}\,g_{j}(t)+p_{i}(t)+{\rm o}_{t\to\infty}(1)
=\displaystyle= j=1mλi,jgj(t)+pi(t)+ot(1).\displaystyle\sum_{j=1}^{m}\lambda_{i,j}\,g_{j}(t)+p_{i}(t)+{\rm o}_{t\to\infty}(1).

This shows that property (4) holds. Since g1,,gm{g}^{*}_{1},\ldots,{g}^{*}_{{m}^{*}} satisfy properties (1) and gj=gjg_{j}={g}^{*}_{j} for all j{1,,m}j\in\{1,\ldots,{m}^{*}\}, we have g1(t)gm(t)g_{1}(t)\prec\ldots\prec g_{{m}^{*}}(t). For j1<j2{m+1,,m}j_{1}<j_{2}\in\{{m}^{*}+1,\ldots,m\}, it follows from L’Hôpital’s rule and gj1(t)gj2(t)g_{j_{1}}^{\prime}(t)\prec g_{j_{2}}^{\prime}(t) that gj1(t)gj2(t)g_{j_{1}}(t)\prec g_{j_{2}}(t). Also, gm(t)gm+1(t)g_{{m}^{*}}(t)\prec g_{{m}^{*}+1}(t) because deg(gm+1)=deg(gm)+1\deg(g_{{m}^{*}+1})=\deg(g_{{m}^{*}})+1. This shows that Property (1) holds. Properties (2), and (3) are straightforward to derive from the definition of g1,,gmg_{1},\ldots,g_{m}. ∎

Corollary A.5.

Let \mathcal{H} be a Hardy field and assume f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} have polynomial growth. Then there exists WW\in\mathcal{H} with 1W(t)t1\prec W(t)\ll t such that f1,,fkf_{1},\ldots,f_{k} satisfy Property (PW).

Proof.

Let mm\in\mathbb{N}, g1,,gm𝒮(f1,,fk)g_{1},\ldots,g_{m}\in\mathcal{S}^{*}(f_{1},\ldots,f_{k}), p1,,pk[t]p_{1},\ldots,p_{k}\in\mathbb{R}[t], and λ1,1,,λk,m\lambda_{1,1},\ldots,\lambda_{k,m}\in\mathbb{R} be as guaranteed by A.4. According to (2), for all g{g1,,gm}g\in\{g_{1},\ldots,g_{m}\} either g=0g=0 or there exists \ell\in\mathbb{N} such that t1g(t)tt^{\ell-1}\prec g(t)\prec t^{\ell}. If g=0g=0 for some g{g1,,gm}g\in\{g_{1},\ldots,g_{m}\} then we must have g=g1g=g_{1}, since g1gmg_{1}\prec\ldots\prec g_{m}. By discarding g1g_{1} if it is the zero function, we can assume that for all j{1,,m}j\in\{1,\ldots,m\} there exists j\ell_{j}\in\mathbb{N} such that tj1gj(t)tjt^{\ell_{j}-1}\prec g_{j}(t)\prec t^{\ell_{j}}. Consider the functions gj(t)/tj1g_{j}(t)/t^{\ell_{j}-1}, j=1,,mj=1,\ldots,m, and pick any function WW\in\mathcal{H} with the property that 1W(t)t1\prec W(t)\ll t and log(W(t))gj(t)/tj1\log(W(t))\prec g_{j}(t)/t^{\ell_{j}-1} for all j=1,,mj=1,\ldots,m. Any WW with these properties is as desired. ∎

Corollary A.6.

Let \mathcal{H} be a Hardy field, VV\in\mathcal{H} with 1V(t)t1\prec V(t)\ll t, and assume f1,,fkf_{1},\ldots,f_{k}\in\mathcal{H} have the property that

  1. (0)

    for all c1,,ckc_{1},\ldots,c_{k}\in\mathbb{R}, n1,,nk{0}n_{1},\ldots,n_{k}\in\mathbb{N}\cup\{0\}, and p[t]p\in\mathbb{R}[t] with p(0)=0p(0)=0 the function f=c1f1(n1)++ckfk(nk)+pf=c_{1}f_{1}^{(n_{1})}+\ldots+c_{k}f_{k}^{(n_{k})}+p satisfies either |f(t)|1|f(t)|\ll 1 or t1log(V(t))|f(t)|tt^{\ell-1}\log(V(t))\prec|f(t)|\ll t^{\ell} for some \ell\in\mathbb{N}.

Then there exists mm\in\mathbb{N}, g1,,gmg_{1},\ldots,g_{m}\in\mathcal{H}, p1,,pk[t]p_{1},\ldots,p_{k}\in\mathbb{R}[t], and λ1,1,,λk,m\lambda_{1,1},\ldots,\lambda_{k,m}\in\mathbb{R} with the following properties:

  1. (1)

    g1(t)gm(t)g_{1}(t)\prec\ldots\prec g_{m}(t);

  2. (2)

    for all g{g1,,gm}g\in\{g_{1},\ldots,g_{m}\} there exists \ell\in\mathbb{N} such that t1log(V(t))g(t)tt^{\ell-1}\log(V(t))\prec g(t)\prec t^{\ell};

  3. (3)

    for all g{g1,,gm}g\in\{g_{1},\ldots,g_{m}\} with deg(g)2\deg(g)\geqslant 2 we have g{g1,,gm}g^{\prime}\in\{g_{1},\ldots,g_{m}\};

  4. (4)

    for all i{1,,k}i\in\{1,\ldots,k\},

    limt|fi(t)j=1mλi,jgj(t)pi(t)|=0.\lim_{t\to\infty}\Bigg{|}f_{i}(t)-\sum_{j=1}^{m}\lambda_{i,j}g_{j}(t)-p_{i}(t)\Bigg{|}=0.
Proof.

This follows straightaway from A.4. ∎

Remark A.7.

It follows from 2.2 that t1log(V(t))|f(t)|tt^{\ell-1}\log(V(t))\prec|f(t)|\ll t^{\ell} if and only if log(V(t))|f(1)(t)|\log(V(t))\prec|f^{(\ell-1)}(t)|. Therefore, if V(t)=tV(t)=t then Property (0) from A.6 is the same as Property (P) from Section 1.

A.2.   Another variant of van der Corput’s Lemma

Theorem A.8.

Let p1,p2,p3,p_{1},p_{2},p_{3},\ldots be a sequence of positive real numbers that is either non-decreasing or non-increasing. Let PNn=1NpnP_{N}\coloneqq\sum_{n=1}^{N}p_{n} and assume

limNPN=andlimNpNPN= 0.\lim_{N\to\infty}P_{N}\,=\,\infty\qquad\text{and}\qquad\lim_{N\to\infty}\frac{p_{N}}{P_{N}}\,=\,0.

Then for every ε>0\varepsilon>0 there exists δ>0\delta>0 such that for every arithmetic function f:f\colon\mathbb{N}\to\mathbb{C} bounded in modulus by 11 and with the property that for every hh\in\mathbb{N} the limit

A(h)limN1PNn=1Npnf(n+h)f(n)¯A(h)\leavevmode\nobreak\ \coloneqq\leavevmode\nobreak\ \lim_{N\to\infty}\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}f(n+h)\overline{f(n)}

exists, we have

lim supH|1Hh=1HA(h)|δlimN|1PNn=1Npnf(n)|ε.\limsup_{H\to\infty}\left|\frac{1}{H}\sum_{h=1}^{H}A(h)\right|\leqslant\delta\leavevmode\nobreak\ \leavevmode\nobreak\ \implies\leavevmode\nobreak\ \leavevmode\nobreak\ \lim_{N\to\infty}\left|\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}f(n)\right|\leqslant\varepsilon. (A.2)
Proof.

Assume p1,p2,p_{1},p_{2},\ldots is non-decreasing, i.e., pnpn1p_{n}\geqslant p_{n-1} for all nn\in\mathbb{N}. For the case when p1,p2,p_{1},p_{2},\ldots is non-increasing similar arguments apply. We claim that for any bounded function f:f\colon\mathbb{N}\to\mathbb{C} we have

limN|1PNn=1Npn(f(n)f(n+1))|= 0.\lim_{N\to\infty}\left|\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}(f(n)-f(n+1))\right|\leavevmode\nobreak\ =\leavevmode\nobreak\ 0. (A.3)

Let f:f\colon\mathbb{N}\to\mathbb{C} be bounded. For the proof of (A.3) we can assume without loss of generality that supn|f(n)|1\sup_{n\in\mathbb{N}}|f(n)|\leqslant 1. After an index-shift we obtain

|1PNn=1Npn(f(n)f(n+1))|\displaystyle\left|\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}(f(n)-f(n+1))\right| \displaystyle\leqslant p1PN+pNPN+|1PNn=2N(pnpn1)f(n)|\displaystyle\frac{p_{1}}{P_{N}}+\frac{p_{N}}{P_{N}}+\left|\frac{1}{P_{N}}\sum_{n=2}^{N}(p_{n}-p_{n-1})f(n)\right|

Using pnpn1p_{n}\geqslant p_{n-1} we can estimate

|1PNn=2N(pnpn1)f(n)|1PNn=1N1(pnpn1).\left|\frac{1}{P_{N}}\sum_{n=2}^{N}(p_{n}-p_{n-1})f(n)\right|\leavevmode\nobreak\ \leqslant\leavevmode\nobreak\ \frac{1}{P_{N}}\sum_{n=1}^{N-1}(p_{n}-p_{n-1}).

The sum n=2N(pnpn1)\sum_{n=2}^{N}(p_{n}-p_{n-1}) is telescoping and equals pNp1p_{N}-p_{1}. We are left with

|1PNn=1Npn(f(n)f(n+1))|p1PN+pNPN+pNPNp1PN=2pNPN.\left|\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}(f(n)-f(n+1))\right|\leavevmode\nobreak\ \leqslant\leavevmode\nobreak\ \frac{p_{1}}{P_{N}}+\frac{p_{N}}{P_{N}}+\frac{p_{N}}{P_{N}}-\frac{p_{1}}{P_{N}}=\frac{2p_{N}}{P_{N}}.

The Claim now follows from the assumption pN/PN0p_{N}/P_{N}\to 0 as NN\to\infty.

Next, fix any f:f\colon\mathbb{N}\to\mathbb{C} bounded by 11. Using (A.3), we get for all HH\in\mathbb{N} that

|1PNn=1Npnf(n)|2=|1Hh=1H1PNn=1Npnf(n+h)|2+oN(1).\left|\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}f(n)\right|^{2}=\left|\frac{1}{H}\sum_{h=1}^{H}\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}f(n+h)\right|^{2}\,+\,{\rm o}_{N\to\infty}(1).

By Jensen’s inequality we have

|1Hh=1H1PNn=1Npnf(n+h)|2\displaystyle\left|\frac{1}{H}\sum_{h=1}^{H}\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}f(n+h)\right|^{2} \displaystyle\leqslant 1PNn=1Npn|1Hh=1Hf(n+h)|2\displaystyle\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}\left|\frac{1}{H}\sum_{h=1}^{H}f(n+h)\right|^{2}
=\displaystyle= 1PNn=1Npn1H2h1,h2=1Hf(n+h1)f(n+h2)¯\displaystyle\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}\frac{1}{H^{2}}\sum_{h_{1},h_{2}=1}^{H}f(n+h_{1})\overline{f(n+h_{2})}
=\displaystyle= 1PNn=1Npn1H2h1,h2=1Hf(n+h1h2)f(n)¯+oN(1)\displaystyle\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}\frac{1}{H^{2}}\sum_{h_{1},h_{2}=1}^{H}f(n+h_{1}-h_{2})\overline{f(n)}\,+\,{\rm o}_{N\to\infty}(1)
=\displaystyle= 1PNn=1Npn1Hh=HHH|h|Hf(n+h)f(n)¯+oN(1).\displaystyle\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}\frac{1}{H}\sum_{h=-H}^{H}\frac{H-|h|}{H}f(n+h)\overline{f(n)}\,+\,{\rm o}_{N\to\infty}(1).

Moreover, we can write

1PNn=1Npn1Hh=HHH|h|Hf(n+h)f(n)¯\displaystyle\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}\frac{1}{H}\sum_{h=-H}^{H}\frac{H-|h|}{H}f(n+h)\overline{f(n)} =\displaystyle= 1Hh=HHH|h|HA(h)+oN(1).\displaystyle\frac{1}{H}\sum_{h=-H}^{H}\frac{H-|h|}{H}A(h)\,+\,{\rm o}_{N\to\infty}(1).

In summary, we have shown that

|1PNn=1Npnf(n)|21Hh=HHH|h|HA(h)+oN(1).\left|\frac{1}{P_{N}}\sum_{n=1}^{N}p_{n}f(n)\right|^{2}\leavevmode\nobreak\ \leqslant\leavevmode\nobreak\ \frac{1}{H}\sum_{h=-H}^{H}\frac{H-|h|}{H}A(h)\,+\,{\rm o}_{N\to\infty}(1).

It is now not hard to show that for every ε>0\varepsilon>0 there exists δ>0\delta>0 such that

lim supH|1Hh=1HA(h)|δlim supH|1Hh=HHH|h|HA(h)|ε,\limsup_{H\to\infty}\left|\frac{1}{H}\sum_{h=1}^{H}A(h)\right|\leqslant\delta\leavevmode\nobreak\ \leavevmode\nobreak\ \implies\leavevmode\nobreak\ \leavevmode\nobreak\ \limsup_{H\to\infty}\left|\frac{1}{H}\sum_{h=-H}^{H}\frac{H-|h|}{H}A(h)\right|\leqslant\varepsilon, (A.4)

from which (A.2) follows. Indeed, lim supH|1Hh=1HA(h)|δ\limsup_{H\to\infty}\left|\frac{1}{H}\sum_{h=1}^{H}A(h)\right|\leqslant\delta implies for all KK\in\mathbb{N} and r{0,1,,K1}r\in\{0,1,\ldots,K-1\} that

lim supH|1HrHK|h|(r+1)HKA(h)|\displaystyle\limsup_{H\to\infty}\left|\frac{1}{H}\sum_{\frac{rH}{K}\leqslant|h|\leqslant\frac{(r+1)H}{K}}A(h)\right| =lim supH|1H|h|(r+1)HKA(h)1H|h|rHKA(h)|\displaystyle=\limsup_{H\to\infty}\left|\frac{1}{H}\sum_{|h|\leqslant\frac{(r+1)H}{K}}A(h)\leavevmode\nobreak\ -\leavevmode\nobreak\ \frac{1}{H}\sum_{|h|\leqslant\frac{rH}{K}}A(h)\right|
lim supH|1H|h|(r+1)HKA(h)|+lim supH|1H|h|rHKA(h)|\displaystyle\leqslant\limsup_{H\to\infty}\left|\frac{1}{H}\sum_{|h|\leqslant\frac{(r+1)H}{K}}A(h)\right|+\limsup_{H\to\infty}\left|\frac{1}{H}\sum_{|h|\leqslant\frac{rH}{K}}A(h)\right|
lim supH|2H1h(r+1)HKA(h)|+lim supH|2H1hrHKA(h)|\displaystyle\leqslant\limsup_{H\to\infty}\left|\frac{2}{H}\sum_{1\leqslant h\leqslant\frac{(r+1)H}{K}}A(h)\right|+\limsup_{H\to\infty}\left|\frac{2}{H}\sum_{1\leqslant h\leqslant\frac{rH}{K}}A(h)\right|
4δ.\displaystyle\leqslant 4\delta.

Hence

lim supH|1Hh=HHH|h|HA(h)|\displaystyle\limsup_{H\to\infty}\left|\frac{1}{H}\sum_{h=-H}^{H}\frac{H-|h|}{H}A(h)\right| K(max0rK|1HrHKh(r+1)HKH|h|HA(h)|)\displaystyle\leqslant K\left(\max_{0\leqslant r\leqslant K}\left|\frac{1}{H}\sum_{\frac{rH}{K}\leqslant h\leqslant\frac{(r+1)H}{K}}\frac{H-|h|}{H}A(h)\right|\right)
=K(max0rK|1HrHKh(r+1)HK(1r/K)A(h)|+1/K2)\displaystyle=K\left(\max_{0\leqslant r\leqslant K}\left|\frac{1}{H}\sum_{\frac{rH}{K}\leqslant h\leqslant\frac{(r+1)H}{K}}\big{(}1-r/K\big{)}A(h)\right|+1/K^{2}\right)
4Kδ+1/K.\displaystyle\leqslant 4K\delta+1/K.

Choosing K=δ1/2K=\lfloor\delta^{-1/2}\rfloor we obtain equation (A.4) with ε=10δ1/2\varepsilon=10\delta^{1/2}. ∎

References


Florian K. Richter
École Polytechnique Fédérale de Lausanne