Uniform distribution in nilmanifolds along functions from a Hardy field
Abstract
We study equidistribution properties of translations on nilmanifolds along functions of polynomial growth from a Hardy field. More precisely, if is a nilmanifold, are commuting nilrotations, and are functions of polynomial growth from a Hardy field then we show that
-
•
the distribution of the sequence 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 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 of a Lie group is called uniform if the quotient is compact, and it is called discrete if there exists a cover of by open subsets of the ambient group in which every open set contains exactly one element of . Given a (-step) nilpotent Lie group and a uniform and discrete subgroup of , the quotient space is called a (-step) nilmanifold. For any group element and point , we define the translation of by as , where is any element in such that . This way, the Lie group acts continuously and transitively on the nilmanifold . There exists a unique Borel probability measure on invariant under this action by , called the Haar measure on (see [Raghunathan72]), which we denote by . A sequence of points in is then said to be uniformly distributed in if
holds for every continuous function .
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 denote the identity component of a nilpotent Lie group , and let be the commutator subgroup generated by .
Definition 1.1 (Maximal factor torus).
Given a connected nilmanifold , the maximal factor torus of is the quotient . We will use to denote the natural factor map from onto .
The maximal factor torus is diffeomorphic to a torus whose dimension equals the dimension of the quotient group . Moreover, as the name suggests, it is the torus of highest dimension that is a factor of .
If is connected then it is well-known that the distribution of orbits along many sequences is governed by their projection onto . 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 in a simply connected nilpotent Lie group , we write for the set of all for which is a well-defined element of the group.111For instance, a rational number with belongs to if and only if there exists such that . Since is assumed to be simply connected, if such an element exists then it must be unique. Note that if and only if .
Theorem 1.2 (Leibman’s Equidistribution Criterion, [Leibman05a, Theorem C]).
Let be a simply connected nilpotent Lie group and a uniform and discrete subgroup of . Suppose
where are commuting and with . Then the following are equivalent:
-
(i)
is uniformly distributed in the nilmanifold .
-
(ii)
is uniformly distributed in the maximal factor torus .
Remark 1.3.
Leibman actually proves 1.2 without the assumption that the elements 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 be a simply connected nilpotent Lie group, a uniform and discrete subgroup of , and
where are commuting and with . Then there exists a closed and connected subgroup of and points such that, for all , the set is a closed sub-nilmanifold of and the sequence is uniformly distributed in .
It is natural to ask whether Theorems 1.2 and 1.4 remain true if one replaces the polynomials 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 , consists of all real-valued functions defined on some half-line that can be build from real polynomials, the logarithmic function , and the exponential function using the standard arithmetical operations , , , and the operation of composition. Examples of logartihmico-exponential functions are for , for , , and , as well as any products or linear combinations thereof.
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 as a representative example in mind, since our results are already new and interesting for this class.
Given two functions we will write when as , and when there exist and such that for all . We say has polynomial growth if it satisfies for some ; in this case the smallest such is called the degree of and denoted by .
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 be a Hardy field, a connected and simply connected nilpotent Lie group, a uniform and discrete subgroup of , and consider the nilmanifold . Suppose
where , have different growth222A finite set of functions is said to have different growth if, after potentially reordering, one has ., and for any there exists such that . Then the sequence is uniformly distributed in the sub-nilmanifold .
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 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 of functions from a Hardy field satisfying Property (P) below. Throughout this work, we use , , and to denote the , , and derivative of a function , respectively.
-
Property (P):
For all and the function has the property that for every either or .
Theorem A.
Let be a Hardy field, a simply connected nilpotent Lie group, a uniform and discrete subgroup of , and assume is connected. Suppose
where are commuting, satisfy Property (P), and . Then the following are equivalent:
-
(i)
is uniformly distributed in the nilmanifold .
-
(ii)
is uniformly distributed in the maximal factor torus .
Whilst A provides us with a convenient criteria for checking whether a sequence of the form 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 be a simply connected nilpotent Lie group, a uniform and discrete subgroup of , and a Hardy field. Let
where are commuting, satisfy Property (P), and for all . Then there exists a closed and connected subgroup of and points such that, for all , the set is a closed sub-nilmanifold of and is uniformly distributed in .
Note that if 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 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 satisfying the hypothesis of 1.5 also satisfy Property (P). For instance, for is such an example.
We also prove analogues of Theorems A and B that apply to an arbitrary collection of functions 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 be a non-decreasing sequence, let be its discrete derivative, and assume as and . A sequence of points in a nilmanifold is said to be uniformly distributed with respect to -averages in if
holds for every continuous function .
When dealing with functions of “slow growth” from a Hardy field, it turns out to be necessary to switch form Cesàro averages to -averages to adequately capture the way in which orbits along such functions distribute. For example, if satisfies then the sequence is not uniformly distributed with respect to Cesàro averages in the unit interval , but it is uniformly distributed with respect to logarithmic averages, i.e., -averages where for all . Likewise, if then the sequence 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., -averages where for all (cf. 5.1 below).
Theorems C and D below are generalizations of Theorems A and B that apply to all finite collections of functions 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 -averages.
Theorem C.
Let be a Hardy field, a simply connected nilpotent Lie group, a uniform and discrete subgroup of , and assume is connected. Suppose
where are commuting and have polynomial growth and satisfy . Then there exists with such that the following are equivalent:
-
(i)
The sequence is uniformly distributed with respect to -averages in the nilmanifold .
-
(ii)
The sequence is uniformly distributed with respect to -averages in the maximal factor torus .
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 be a Hardy field, a simply connected nilpotent Lie group, a uniform and discrete subgroup of , and assume is connected. Suppose
where are commuting and have polynomial growth and satisfy . Then the following are equivalent:
-
(i)
The sequence is dense in the nilmanifold .
-
(ii)
The sequence is dense in the maximal factor torus .
Finally, let us state our last main result.
Theorem D.
Let be a simply connected nilpotent Lie group, a uniform and discrete subgroup of , and a Hardy field. Suppose
where are commuting and have polynomial growth and satisfy . Then there exists with , a closed and connected subgroup of , and points such that is a closed sub-nilmanifold of and is uniformly distributed with respect to -averages in for all .
The connection between and in Theorems C and D is given by a variant of Property (P):
-
Property (PW):
For all and the function has the property that for every either or .
As we will show below (see A.5), for an arbitrary finite collection of functions from a Hardy field of polynomial growth there exists some with such that Property (PW) holds. Moreover, it will be clear from the proofs of Theorems C and D that if satisfy Property (PW) for some then we can take this 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 is defined as .
Theorem 1.8 (Szemerédi’s Theorem).
For any set of positive upper density and any there exist such that .
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 of positive upper density and any polynomials satisfying there exist such that .
1.9 was later improved in [BLL08] to include an “if and only if” condition.
Theorem 1.10.
Given the following are equivalent:
-
(i)
The polynomials are jointly intersective, i.e., for any there exists such that for all .
-
(ii)
For any set of positive upper density there exist such that .
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 denote the floor function.
Theorem 1.11.
Let be functions from a Hardy field of different growth and with the property that for every there is such that . Then for any set of positive upper density there exist such that .
Given a finite collection of functions we denote by the set of all functions of the form for , and by the set of all functions of the form for . 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 be functions of polynomial growth from a Hardy field such that
for every , and every . Then for any set of positive upper density there exist such that .
Finally, another variant of Szemerédi’s Theorem, which also involves functions from a Hardy field, was obtained in [BMR17arXiv].
Theorem 1.13.
Let be a function from a Hardy field and assume there is such that . Then for any set of positive upper density there exist such that .
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 1.8, 1.9, 1.10, 1.11, 1.13, but also confirms 1.12.
Theorem E ([BMR20draft]).
Let be the rounding to the closest integer function, let be functions from a Hardy field with polynomial growth, and assume at least one of the following two conditions holds:
-
(1)
For all and we have .
-
(2)
There is jointly intersective collection of polynomials such that any real polynomial “appearing” in also appears in , where we say a polynomial “appears” in if there is such that .
Then for any set of positive upper density there exist such that .
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 is any equivalence class of real-valued functions in one real variable under the equivalence relationship . Let denote the set of all germs at of real valued functions defined on some half-line for some . Note that forms a ring under pointwise addition and multiplication, which we denote by .
Definition 2.1 (see [Boshernitzan94, Definition 1.2]).
Any subfield of the ring that is closed under differentiation is called a Hardy field.
By abuse of language, we say that a function belongs to some Hardy field (and write ) if its germ at belongs to .
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 belonging to a Hardy field and any the function is either eventually non-negative or eventually non-positive. Combined with the fact that is a field and closed under differentiation, this implies that:
-
•
the limit always exists as an element in ;
-
•
is either eventually increasing, eventually decreasing, or eventually constant;
-
•
for any either , or , or 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 be a Hardy field.
-
1.
If satisfies for some and is not asymptotically equal to a constant then
-
2.
If satisfies for some then 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 be a -step nilpotent Lie group with identity element . The lower central series of , which we denote by , is a decreasing nested sequence of normal subgroups,
where is the subgroup of generated by all the commutators with and . Note that because is -step nilpotent. Also, each is a closed subgroup of (cf. [Leibman05a, Section 2.11]).
The upper central series of , denoted by , is an increasing nested sequence of normal subgroups,
where the are defined inductively by and . Note that is equal to the center of and because is -step nilpotent.
Given a uniform and discrete subgroup of a nilpotent Lie group , an element with the property that for some is called rational (or rational with respect to ). A closed subgroup of is then called rational (or rational with respect to ) if rational elements are dense in . For example, the subgroups in the lower central series of , as well as in the upper central series of , are rational with respect to any uniform and discrete subgroup of (cf. [Raghunathan72, Corollary 1 of Theorem 2.1] for a proof of this fact for connected and [Leibman05a, Section 2.11] for the general case).
Rational subgroups play a key role in the description of sub-nilmanifolds. If is a nilmanifold, then a sub-nilmanifold of is any closed set of the form , where and is a closed subgroup of . It is not true that for every closed subgroup of and every element in the set is a sub-nilmanifold of , because need not be closed. In fact, it is shown in [Leibman06] that is closed in (and hence a sub-nilmanifold) if and only if the subgroup is rational with respect to .
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 to denote the center of a group .
Lemma 2.3.
Let be a nilpotent group and a non-trivial normal subgroup of . Then .
Proof.
We will make use of the upper central series which was defined in the previous subsection. Consider the intersection for . Note that and . Let and note that is non-empty because . Let denote the minimum of . By definition, we have , which implies . Moreover, is a normal subgroup of , because it is the intersection of two normal subgroups of , and hence . We conclude that
In view of the minimality assumption on we have , from which it follows that . This proves that is a subset of . Thus as desired. ∎
Corollary 2.4.
Let be a simply-connected nilpotent Lie group and a non-trivial, connected, and normal subgroup of . Then .
Proof.
According to 2.3 there exists an element in the intersection . Since is connected and , the element belongs to the identity component of . Moreover, since is simply-connected, the -parameter subgroup is well defined. It follows from [Malcev49, Lemma 3] (and, alternatively, also from the Baker-Campbell-Hausdorff formula) that if commutes with an element then the entire -parameter subgroup commutes with . This implies that if belongs to the center of , then so does . In particular, , which proves and hence . ∎
Definition 2.5 (Central characters).
Let be a nilpotent Lie group and a uniform and discrete subgroup of . A central character of is any continuous map with the property that there exists a continuous group homomorphism such that
(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 there exists a central character such that . To verify this claim, we distinguish two cases, the case and the case . If we are in the first case then . This implies there exists a continuous function on the quotient space with , which we can lift to a continuous and -invariant function on satisfying . If we are in the second case then there is such that . Note that cannot be an element of because . Let be any group character of with the property that and . Define and observe that is a compact abelian group. Let be a small neighborhood of , and let be a continuous function with the property that and for all . Now define , where is the normalized Haar measure on . It is straightforward to check that is a non-zero and continuous function on satisfying for all and . In particular and so . in , it follows from the Stone-Weierstrass Theorem that their linear span is uniformly dense in . We will make use of this fact multiple times in the upcoming sections.
Remark 2.7.
Let be a central character of and let be the corresponding continuous group homomorphism such that (2.1) is satisfied. We claim that if is non-trivial (meaning that there exists such that ) then the integral equals . To verify this claim, note that the measure is invariant under left-multiplication by . Therefore,
which can only hold if .
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 be a group and a normal subgroup of . We define the relatively independent self-product of over as
If is a nilpotent Lie group, a uniform and discrete subgroup of , and a normal and rational subgroup of then the group
is a uniform and discrete subgroup of . This gives rise to the nilmanifold
which we call the relatively independent self-product of over .
Remark 2.9.
In ergodic theory, the notion of a relatively independent self-joining of a system over one of its factors is an important notion and finds many applications (see [EW11, Definition 6.15] for the definition). If is a nilmanifold and is a normal and rational subgroup of then the quotient space is a factor of . It turns out that that the relatively independent self-product is exactly the same as the relatively independent self-joining of over the factor .
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 .
Proof.
Note that is a normal subgroup of and a subset of , because is a normal subgroup of . It follows that is a normal subgroup of . Moreover, is a subset of because , and is a subset of because . Thus, to show
it suffices to show
(2.2) |
Note that is generated by elements of the form for and . Since elements in commute with elements in modulo , we have . In other words,
Similarly, since , we have
This finishes the proof of (2.2). ∎
From 2.10 we can derive the following corollary.
Corollary 2.11.
and are isomorphic as nilpotent Lie groups.
Proof.
Consider the map defined as
Clearly, is well defined, smooth, surjective, and a homomorphism. Moreover, it follows from 2.10 that the kernel of equals . Indeed, belongs to the kernel of if and only if belongs to . According to 2.10, this happens exactly when belongs to . Using the definition of relative independent product groups, we see that if and only if and as claimed. ∎
2.11 helps us better understand the maximal factor torus of the relatively independent self-product , 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 be a nilpotent Lie group and a uniform and discrete subgroup of . A horizontal character of is any continuous map satisfying
(2.3) |
We say is non-trivial if is not constant equal to .
Remark 2.13.
With the help of 2.11 it is easy to describe the horizontal characters of the relatively independent product in terms of the horizontal characters of and , where . Indeed, for any horizontal character of there exists a horizontal character of and a horizontal character of with such that
Observe that if 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 in the case when is connected. However, we also need to better understand the maximal factor torus if is not connected. First, let us characterize the identity component of .
Lemma 2.14.
We have .
Proof.
If are the connected components of and are the connected components of then the connected components of are , where
This is because are open and connected subsets of , and if then and are disjoint. It follows that the connected component of that contains the identity is . ∎
To study the maximal factor torus for non-connected , we utilize a variant of the notion of a horizontal character.
Definition 2.15 (Pseudo-horizontal characters).
Let be a nilpotent Lie group and a uniform and discrete subgroup of . A pseudo-horizontal character of is any continuous map satisfying
(2.4) |
Note that if 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 is connected. Indeed, if is connected then acts transitively on and hence acts transitively on the maximal factor torus. This implies that the maximal factor torus is isomorphic to the compact abelian group 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 in the case when is connected, similar to the way we described the horizontal characters of in 2.13 above. If is connected then it is isomorphic to
and hence the pseudo-horizontal characters of can be identified with the horizontal characters of . It follows that for any pseudo-horizontal character of there exists a pseudo-horizontal character of and a pseudo-horizontal character of with such that
Finally, here is another lemma that we will invoke numerous times in the later sections.
Lemma 2.17.
Let be a nilpotent Lie group, a uniform and discrete subgroup of , and a normal and rational subgroup of . Let denote the natural projection of onto the nilmanifold . If both and the sub-nilmanifold are connected then the relatively independent self-product is also connected.
Proof.
Note that is connected if and only if . Likewise, is connected if and only if . In particular, any can be written as with and , and any can be written as with and . It follows that
Since is normal, we have where . Hence
which implies is connected. ∎
3. Reducing Theorems A and B to F
Theorem F.
Let be a simply connected nilpotent Lie group and a uniform and discrete subgroup of . Assume is a mapping of the form
where , , the elements are pairwise commuting, are connected sub-nilmanifolds of , are polynomials satisfying
and are functions belonging to some Hardy field satisfying
Then is uniformly distributed in the sub-nilmanifold .
We remark that the set appearing in the formulation of F above is indeed a sub-nilmanifold of , 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 is uniformly distributed in then for every the sequence is uniformly distributed in .
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 be a simply connected nilpotent Lie group, a uniform and discrete subgroup of , and the nilmanifold . Let be pairwise commuting, satisfy Property (P), suppose for all , and consider the sequence
To begin with, we divide the set into two pieces. The first piece, which we denote by , consists of all for which belongs to the identity component . The second piece is defined as . According to A.2, for any there exist and such that , , and
Then, by A.6 applied with to the family of functions , we can find , functions , a set of polynomials , and coefficients such that the following properties hold:
-
(I)
;
-
(II)
for all either or there exists such that ;
-
(III)
for all with we have ;
-
(IV)
for all ,
For the polynomial can be written as
for some real coefficients . For the polynomial can also be written as
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 then the coefficients are actually integers.
Next, for , define
and for define
Note that for all the elements and are well defined because , and for all the elements are well defined because and . It follows that the elements belong to and are pairwise commuting.
Let be any right-invariant metric on the nilpotent Lie group . Using property (IV), it is straightforward to check that
Therefore, instead of showing the conclusions of Theorems A and B for , it suffices to show the conclusions of Theorems A and B for the sequence , where
Note that was omitted because it is a constant (cf. 3.1).
It is a consequence of 1.4 that for every there exists such that for all the set
is a connected sub-nilmanifold of . If we take then
is also connected for all and all . Let be the polynomial defined as
let and , let and , and define
A straightforward calculation shows that
Also note that the elements belong to and are pairwise commuting. Additionally, the polynomials satisfy conditions (F1) and (F2) from F, and the functions belong to and satisfy conditions (F3) and (F4). Condition (F5) is also satisfied, because
and satisfy (III). In conclusion, all the conditions of F are met, which means that for every the sequence is uniformly distributed in the sub-nilmanifold . Define
and note that is connected because is connected for all . Moreover, using that the map from is homeomorphic, we have
Let and observe that are sub-nilmanifolds of , because and . Also, finitely many translates of cover , because
Since is connected and finitely many translates of it cover , we conclude that is the identity component of , i.e., is the connect component of containing . Let be the stabilizer of ,
and define . Then and . Moreover, since and is a normal subgroup of , it follows that
Thus, choosing , we have shown that for a connected and closed subgroup of , which finishes the proof of B. Also, since is dense in 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 of there exists some that acts ergodically on . If the projection of onto the maximal factor torus is surjective, then also acts ergodically on the maximal factor torus, which in view of 1.2 implies that acts ergodically on . This shows that a sub-nilmanifold of coincides with if and only if its projection onto the maximal factor torus equals the the maximal factor torus., it follows that is uniformly distributed in if and only if its projection onto is uniformly distributed there. In particular, is uniformly distributed in if and only if is uniformly distributed in , 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 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 and . 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 . 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 uses induction on the so-called degree of the sequence , see 4.7.
4.1. The abelian case
The following corresponds to the special case of F where is abelian.
Theorem 4.1 (The abelian case of F).
Let and consider a map of the form
where , , satisfy properties (F1) and (F2), and satisfy properties (F3), (F4), and (F5). Moreover, let be a subgroup of of finite index, define , and assume that for every the set is a connected subgroup of . Then the sequence is uniformly distributed in the subgroup
of .
Proof.
Since is a compact abelian group, the algebra generated by continuous group characters is uniformly dense . Therefore, to prove that is uniformly distributed in it suffices to show that for every continuous group characters that is non-trivial when restricted to we have
(4.1) |
Let be shorthand for and choose and such that for all as well as for all . This allows us to rewrite (4.1) as
(4.2) |
Since have different growth and satisfy property (F4), if at least one of the numbers is non-zero then it follows from Boshernitzan’s Equidistribution Theorem ([Boshernitzan94, Theorem 1.8]) that (4.2) holds. If all of the are zero and at least one of the numbers is non-zero then (4.2) holds too, because have different degree and, since is connected for all , any non-zero must be an irrational number and polynomials with irrational coefficients are uniformly distributed mod by Weyl’s Equidistribution Theorem [Weyl16, Satz 14]. To finish the proof, note that not all the numbers can be zero since was assumed to be non-trivial when restricted to . ∎
4.2. The sub-linear case
The purpose of this subsection is to prove the special case of F when and , 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 be a simply connected nilpotent Lie group, a uniform and discrete subgroup of , and a Hardy field. For any mapping of the form
where are commuting and , the sequence is uniformly distributed in the sub-nilmanifold .
Given a sub-nilmanifold of a nilmanifold and a function we will write to denote the quantity . A sequence of sub-nilmanifolds is said to be uniformly distributed in if
The following lemma will be instrumental in our proof of 4.2.
Lemma 4.3.
Let be a connected simply connected nilpotent Lie group, a uniform and discrete subgroup of and consider the nilmanifold . Let be arbitrary and let denote the smallest connected, normal, rational, and closed subgroup of containing . Let denote the diagonal . Then for all but countably many , the sequence is uniformly distributed in the relatively independent product .555Since is a rational and closed subgroup of , we can identify with . This allows us to view the diagonal as a sub-nilmanifold of the relatively independent product .
Proof.
It follows from [Leibman05b, Corollary 1.9] that is uniformly distributed in if and only if
(4.3) |
for every non-trivial horizontal character of . As was mentioned in 2.13, for any horizontal character of there exists a horizontal character of and a horizontal character of with and such that
Therefore we get
Since is non-trivial, either or is non-trivial. If is non trivial then and hence (4.3) is satisfied. It remains to deal with the case when is non-trivial.
Note that for some with . We claim that . Before we verify this claim, let us show how it allows us finish the proof of (4.3). Indeed, if then for all but countably many . Since there are only countably many horizontal characters, by excluding countably many “bad” s for each horizontal character, there exists a co-countable set of “good” s independent of the choice of . Since the Cesàro average of is whenever , it follows that for any such “good” we have
which proves that (4.3) holds.
Let us now prove the claim that whenever is non-trivial. By way of contradiction, assume . Therefore for all . Consider the set
Clearly . Since for all , we see that is a subgroup of , and hence also a subgroup of . Moreover, implies , from which we conclude that is a normal subgroup of . Since and are continuous maps, is a closed set. Also, since is a closed subset of , is rational. In summary, is a normal, rational, and closed subgroup of . Let be the identity component of . Then is also normal, rational, and closed. On top of that, is connected and contains . By the minimality assumption on , we must have . However, we also have because 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 are functions from a Hardy field satisfying . Let be a bounded and uniformly continuous function and suppose for all the limit
exists. If for every there exists such that
then necessarily
(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 . The case of 4.2 where is abelian (i.e., where the nilpotency step of equals ) has already been taken care of by 4.1. Let us therefore assume that is a -step nilpotent Lie group with and that 4.2 has already been proven for all cases where the nilpotency step of the Lie group is smaller than .
By replacing with the sub-nilmanifold (and with a rational and closed subgroup of itself) if necessary, we can assume without loss of generality that
(4.5) |
In particular, is connected and therefore . On top of that, . This means we can replace by if needed, which allows us to also assume that is connected.
To show that is uniformly distributed in we must verify
(4.6) |
for all continuous functions . 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 , instead of (4.6) it suffices to show
(4.7) |
for central characters only. Let us therefore fix a central character and let be the character of corresponding to , i.e., the continuous group homomorphism from to the unit circle such that (2.1) holds.
Write for the smallest connected, normal, rational and closed subgroup of containing the element . Let be the intersection of with and note that by 2.3, is a non-trivial subgroup of . Moreover, since is connected and is connected (note that is connected because ), is connected too.666It follows from [Malcev49, Lemma 3] that if belongs to a connected subgroup of a simply connected nilpotent Lie group then so does for all . This property implies that the intersection of connected subgroups of is connected.
We now claim that we can assume is non-trivial when restricted to (by which we mean that there exists such that ). Indeed, if is trivial on then is invariant under the action of . In this case, descends to a continuous function on the quotient space . We can identify with the nilmanifold and, since is connected and non-trivial, the dimension of is smaller than the dimension of . 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 is non-trivial when restricted to .
Since is non-trivial when restricted to , it is in particular a non-trivial central character. Non-trivial central character have zero mean (see 2.7), and hence (4.7) becomes
(4.8) |
In order to prove (4.8) we use the “van der Corput Lemma for sub-linear functions”, i.e., 4.4. Define
(4.9) |
According to 4.4, if we can show that is well defined for all (meaning that the limit on the right hand side of (4.9) exists for all ) and for every there exists such that
(4.10) |
then (4.8) holds. Define
and let . Clearly, takes values in for all and . This will allow us to express in terms of the relatively independent product instead of the cartesian product , which, as we will see, turns out to be a big advantage. Define the map as for all . Note that is well defined and continuous. We can now rewrite (4.9) as
We make two claims:
Claim 1.
The integral equals .
Claim 2.
For all ,
(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 exists for all and
In view of 4.3, we thus have for a co-countable set of that
It remains to verify Claims 1 and 2.
Proof of Claim 1.
Recall that there exists such that . Using the definition of it is straightforward to check that
Note that is an element of because and . Therefore is invariant under left-multiplication by , which gives
Since , we obtain as claimed. ∎
Proof of Claim 2.
Define a new function via
It is straightforward to check that (4.11) is equivalent to
(4.12) |
Write for the natural projection of onto and set
It follows from (2.1) that is invariant under the action of and so is invariant under the action of . Therefore, descends to a continuous function on , which makes (4.12) equivalent to
(4.13) |
Since has nilpotency step , we can invoke the induction hypothesis and conclude that the sequence is uniformly distributed on the sub-nilmanifold where , . However, (4.5) implies
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 . 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 . For the definition of this degree, the notion of a filtration is needed.
Definition 4.5.
Let be a nilpotent Lie group and a uniform and discrete subgroup of . Let and let be subgroups of that are normal, rational and closed. We call a -step filtration of if , , and
The next lemma shows how one can turn a -step filtration of into a -step filtration of the relatively independent self-product .
Lemma 4.6 (cf. [GT12a, Proposition 7.2]).
Let be a -step filtration of and suppose is a normal subgroup of with . Define for , and set . Then
is a -step filtration of .
Proof.
Suppose and , where and belong to and satisfy . To complete the proof we must show that . Clearly, and . It remains to prove that
(4.14) |
Since , we will establish (4.14) in two steps: First we will show that , and thereafter we will show that .
Write and . Then and . Likewise, and . So can be written as
(4.15) |
Since , (4.15) is equivalent to
(4.16) |
Next, using and , we see that (4.16) is equivalent to
(4.17) |
Using normality of again, (4.17) reduces to
Finally, since and , it follows that , which finishes the proof of (4.15).
It remains to show that
(4.18) |
We can assume without loss of generality that . Since , we have that , and . In particular, modulo the element commutes with , , and . Hence
Next, observe that and hence commutes with modulo . Therefore
Finally, since we have and conclude that . ∎
Given a nilpotent Lie group , let be a right-invariant metric on . For any uniform and discrete subgroup the metric descends to a metric on the nilmanifold in the following way:
(4.19) |
Given a subset and a point we denote by the distance between and .
Definition 4.7 (cf. [GT12a, Definition 1.8] and [BMR17arXiv, Definition 5.9]).
Let be a simply connected nilpotent Lie group, a Hardy field, and a mapping of the form
where , , the elements are commuting, have polynomial growth, and with . We define the degree of to be the smallest number such that there exists a -step filtration with the property that for all and for all . If is such a minimal filtration then we say realizes the degree of . If there exists no such filtration, then we say that has infinite degree.
Lemma 4.8.
Let be a simply connected nilpotent Lie group and a Hardy field. Assume is a mapping of the form
where , , the elements are commuting, have polynomial growth, and with . Then has finite degree.
Proof.
Let be any number such that for all and for all . Let denote the lower central series of . Set and define a filtration
by setting for all and and . It is straightforward to check that is a filtration. Also, since for all , we certainly have that for all and for all . ∎
Remark 4.9.
Note that the the filtration constructed in the above proof is not necessarily a filtration that realizes the degree of . Nonetheless, its existence proves that the degree of does not exceed .
Proof of F, the general case.
Let be as in the statement of F. We use induction on the degree of , which is finite due to 4.8. The base case of this induction, which is when , is covered by 4.1, because if then must be abelian. Therefore, we only have to deal with the inductive step. Assume and F has already been proven for all mappings satisfying the hypothesis of F and whose degree is strictly smaller than .
By replacing with if necessary777Let be the the smallest rational and closed subgroup of containing . Then is a uniform and discrete subgroup of and the nilmanifold can be identified with the sub-nilmanifold of . Moreover, is uniformly distributed in if and only if is uniformly distributed in . Thus, by replacing with , with , and with , we can assume without loss of generality that ., we will assume that
(4.20) |
Since are assumed to be connected, the sub-nilmanifold is also connected. It follows that is connected, which in turn implies that
is connected. For technical reasons, it will be convenient to assume that for every , if the sub-nilmanifold is a point then . This assumption can be made without loss of generality, because if is a point then must belong to , in which case we can simply replace with and the sequence remains unchanged.
Our goal is to show that is uniformly distributed in , or equivalently,
(4.21) |
for all . 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
(4.22) |
for all central characters . Let denote all the numbers for which , and let be all the numbers for which . Note that if and are both the empty set then and , 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 or is non-empty.
Let be a -step filtration that realizes the degree of (cf. 4.7). Among other things, this means for all and for all . Note that implies (cf. Footnote 6). Let denote the smallest closed rational and normal subgroup of containing for all and for all . Then, if denotes the natural projection of onto , the set is a sub-nilmanifold of containing the sub-nilmanifold . 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 is connected, is connected too.
Next, we claim that the identity component of is non-trivial. To verify this claim, we are going to distinguish two cases. The first case is when is non-empty. In this case, contains a one-parameter subgroup for and is therefore non-trivial (we assume without loss of generality that for all ). The second case is when is non-empty. Note that if we are not in the first case, then we must be in the second, since either or is non-empty. If is non-empty then contains for some , and since is connected and not a point for every , it follows that is non-trivial.
Let be the intersection of with . By 2.4, is a non-trivial subgroup of . Also, as an intersection of connected subgroups, 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 , to show that it suffices to prove (4.22) for the case when the central character has a “central frequency” that is non-trivial when restricted to , i.e., there exists such that . This also implies that , and hence (4.22) can be written as
(4.23) |
To prove (4.23) we use van der Corput’s trick. Define
(4.24) |
whenever this limit exists. In light of A.8 (applied with for all ), (4.23) holds if we can show that the limit on the right hand side of (4.24) exists for all and
(4.25) |
We can interpret as , where is a continuous function on the product nilmanifold and is an element in . Note that the sequence can be rewritten as
(4.26) |
where and
(4.27) |
For every the function has degree , which means its discrete derivative is negligibly small for large . This implies that for every the element converges to the identity and can therefore be ignored. More precisely, using the right-invariance of the metric , we have
where
It follows that if we set
(4.28) |
then, in view of (4.26), the difference between and goes to zero as . Hence equals
(4.29) |
The advantage of using (4.29) instead of (4.24) is that for all . Define the map as for all . Note that is well defined and continuous. This allows us to rewrite (4.29) as
(4.30) |
Let and denote by the natural projection of onto . Define
It follows from (2.1) that is invariant under the action of . Therefore, descends to a continuous function on , meaning there exists such that
It thus follows form (4.30) that
(4.31) |
We now make four claims.
Claim 1.
The integral equals zero.
Claim 2.
For all non-trivial pseudo-horizontal characters of (see 2.15) we have
(4.32) |
In the following, we call any mapping of the from , where are elements in a nilpotent Lie group and are polynomials with , a polynomial mapping (cf. [Leibman05a, Subsection 1.3]).
Claim 3.
There exist polynomials with and , polynomial mappings , and polynomial mappings such that the elements are pairwise commuting for every , and
for every .
Claim 4.
For every the degree of is smaller than .
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 exists for all and (4.25) holds. Claims 3 and 4 allow us to invoke the induction hypothesis and deduce that for every the sequence is uniformly distributed in the sub-nilmanifold
(As was explained in 3.1, it is not a problem that the “base point” of the sequence is instead of .) As a consequence we have , which in particular proves that the limit in exists for all . Moreover, (4.25) will follow if we can show that
(4.33) |
Given a vector , consider the multi-parameter polynomial sequence defined as
Arguing as in the proof of [BMR17arXiv, Lemma A.7] we can find for every a co-null set such that for all we have
(4.34) |
It follows that for every and every the sequence
is dense in . By invoking [Leibman05b, Theorem A, p. 216], we have that since this sequence is dense in , it is also uniformly distributed in . This means that
for all continuous functions . Henceforth, let be any number in . Note that Claim 2 implies
for all non-trivial pseudo-horizontal characters of . It follows that
Note also that is connected, due to 2.17 and the fact that both and are connected. Thus, it follows from the work of Leibman (see [Leibman05b, Theorems A and B]) that the sequence
is well distributed888A sequence of points in a nilmanifold is said to be well distributed in if for all and all there exists such that for all with for all we have in . We conclude that
for all continuous functions . In particular,
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 is non-trivial when restricted to , meaning that there exists such that . Let , where denotes the identity element of . Using the definition of it is straightforward to check that
Since is invariant under , we have that
and hence as claimed. ∎
Proof of Claim 2.
For any pseudo-horizontal character of there exists a pseudo-horizontal character of such that . Thus, instead of (4.32), it suffices to show that for all non-trivial pseudo-horizontal characters of we have
(4.35) |
According to 2.16, there exist a pseudo-horizontal character of and a pseudo-horizontal character of with such that
where . Thus, by (4.28),
Although we have , we don’t necessarily have . This makes it more difficult to study the expressions and , because and are only pseudo-horizontal characters and not horizontal characters. However, we can circumvent these difficulties in the following way. Since is connected, we have
(4.36) |
As is explained in [Leibman05a, Subsections 2.6 and 2.7 on p. 204], under these conditions there exist a polynomial sequence such that
(4.37) |
The advantage of using instead of is that takes values in and hence the image of under and is easier to understand. A downside of making this trade-off is that, unlike , the values of do not necessarily commute with . 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 such that
Note that even though
we do not necessarily have . But we do have
(4.38) |
which we will make use of later.
It will be convenient to pick and such that
where is shorthand for . From (2.4) and the fact that as well as for all , it follows that
(4.39) |
Note that belongs to the kernel of and so it follows from (4.38) that
Let be polynomials such that
as well as
Then (4.39) implies
and so (4.35) becomes
(4.40) |
Since have different growth (see property (F3)) and behave independently from polynomials (due to property (F4)), it follows that if at least one of the is non-zero or at least one of the is non-zero, then (4.40) is satisfied and we are done. Let us therefore assume for all and for all . In this case, (4.40) is equivalent to
(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
which is the same as
(4.42) |
Recall that is connected for all . This implies that is also connected for all and therefore
is connected. It now follows from (4.37) that and are connected. But if is connected then, because , as soon as the function is non-constant, the average
must equal . If this average equals then (4.42) holds, which implies that (4.41) holds, and once again we are done. Let us therefore assume that is constant. Since , if is constant then we must have for all . Therefore belongs to the kernel of and (4.41) becomes
(4.43) |
Since is connected, we once again only have two possibilities: either is constant equal to 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 cannot be constant equal to under the current assumptions.
By way of contradiction, assume for all . This implies that belongs to the kernel of . But we also have that belong to the kernel of and for as well as for all belong to the kernel of . We claim that having all those elements belong to the kernels of and contradicts the hypothesis that either or are non-trivial.
To verify this claim, we first need to make a simplifying assumption. Note that the group generated by and is closed and rational999 Since is connected, be have . Therefore, for every there exists such that . This means that the group generated by and equals , where is the subgroup of generated by . This proves that the group generated by and is both closed and rational., and it contains as well as . Therefore we can replace by if necessary, and will henceforth assume that .
Next, define
where . We claim that is a normal subgroup of with a dense subset of rational elements. Once verified, this claim will imply that the closure of , which we denote by , is a closed, rational, and normal subgroup of .
To show that is a group, define
and note that . Certainly, is a subgroup of . If we can show that is a normal subgroup of then it will follow that is a group.
Since , to prove that is normal it suffices to show that for all , and for all . It is easy to see that that holds for all , because contains . To show that for all fix some between and . Since commutes with , we have
Since both and are normal subgroups of , the commutator is normal and hence
Since (cf. (4.36)), there exists and such that . Hence
Since is a normal subgroup of , we have . So
Finally, observe that , which gives
This proves that is a normal subgroup of and hence is a subgroup of .
Next, let us show that is normal too. When proving that is normal, we used that is generated by and . For the proof that is normal, this does not seem to be particularly helpful. Instead, we shall use that the set is dense in (which follows from (4.20)). Therefore, to show that is normal, it suffices to prove that for all , for all , and for all . To verify the first assertion, namely that , simply note that , where is normal and is a group every element of which commutes with . A similar argument shows that . To see why holds for all , simply note that and is a normal subgroup of .
Finally let us show that rational elements are dense in , or equivalently, that is rational. It is well known (see [Leibman05a, Subsection 2.2, p. 203–204]) that a closed subgroup of is rational if and only if its intersection with is a uniform subgroup of that group. Hence, to prove that is rational, it suffices to show that is a uniform subgroup of . However, since and , it follows that . Since is a uniform subgroup of and is a subgroup of , it follows that is a uniform subgroup of and we are done.
In conclusion, is a closed and rational subgroup of that contains for all and for all . Moreover, is a normal subgroup of . Since, by definition, is the smallest subgroup of with all these properties, we must have
Recall that and for all belong to the kernel of and for and for all belong to the kernel of . From this it follows that also belong to the kernel of both and . Recall also that . In other words is a subset of and is a subset of . Since is dense in it follows that is trivial, and since is dense in , is also trivial. This contradicts the fact that either or is non-trivial and finishes the proof of Claim 2. ∎
Proof of Claim 3.
Recall that , where ,
and
Let denote the degree of . Using Taylor’s Theorem, we can approximate by
In view of 2.2, we have . Thus,
If then . Also, according to the hypothesis of F, for every there exists such that . For every define and set
Now define, for every , the polynomial mapping as
In a similar way, one can find . ∎
Proof of Claim 4.
Fix . Since is filtration that realizes the degree of , we have for all and for all . Define , , , and
According to 4.6, is a -step filtration of . We claim that is a filtration that realizes the degree of . Recall that
where
and
To show that is a filtration realizing the degree of , we must prove that
-
(i)
for all ;
-
(ii)
for all ;
-
(iii)
for all ;
-
(iv)
for all ;
Parts (i) and (ii) follow from the fact that for all and that for all and for all . Parts (iii) and (iv) follow from the fact that and and that .
To complete the proof of Claim 4, note that if is a filtration realizing the degree of , then the filtration , defined as
is a filtration realizing the degree of . Moreover, is equal to because , and hence belongs to the kernel of . This shows that is a -step filtration and hence has degree . ∎
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 -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 be a simply connected nilpotent Lie group, a uniform and discrete subgroup of , and a Hardy field. Assume is a mapping of the form
where , , the elements are pairwise commuting, are connected sub-nilmanifolds of , are polynomials satisfying
and satisfy
where has degree . Then is uniformly distributed with respect to -averages in the sub-nilmanifold .
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 , we apply A.6 with where is chosen (using A.5) such that 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 -averages and instead of utilizing 4.4 one uses 6.1, which is precisely the analogue of 4.4 for -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 -averages and instead of applying A.8 with one applies A.8 with . 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 -averages. Let us formulate and prove such a variant now.
Theorem 5.1.
Let be a Hardy field, let , and assume and for some . Then
(5.1) |
where .
Although 5.1 for does not imply Boshernitzan’s Equidistribution Theorem in full generality, it is good enough for the proof of the abelian case of G.
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 and are functions from a Hardy field satisfying and . Let be a bounded and uniformly continuous function and suppose for all the limit
(6.1) |
exists, where . If for every there exists such that
then necessarily
(6.2) |
The next lemma will be useful for the proof of 6.1. We say a function has sub-exponential growth if for all .
Lemma 6.2.
Let with . Then has sub-exponential growth.
Proof.
Note that has sub-exponential growth if and only if
Let us therefore consider the number . Note that this limit exists because if belongs to some Hardy Field, then so does .
Since is eventually increasing, we have . It remains to show that , which we will do by showing that for all . Thus, fix any with . There exists such that for all but finitely many we have
and hence, using , we obtain
Since , we conclude that and hence . ∎
Remark 6.3.
It follows from 6.2 that if with and then has sub-exponential growth.
Lemma 6.4.
Let with and , and define . For every and define
Define , , and . Then the following hold:
-
(i)
.
-
(ii)
;
-
(iii)
;
-
(iv)
.
Proof.
Since and is eventually monotone increasing, for sufficiently large the set is an interval of the form , where . For all such we thus also have . Let . Then for all but finitely many we have that . Note that the difference between and can be bounded from above by , and since as , we have that . Similarly, we can show that . Therefore and hence
Proof of 6.1.
Fix and define and . Since
instead of (6.2) is suffices to show that
(6.3) |
Set and . According to 6.4, part (ii), we have
Therefore, (6.3) is equivalent to
which we can write as
(6.4) |
Define for and note that . Then and, for all , we have . Therefore, using the uniform continuity of , we have that
It follows that
In conclusion, (6.4), and therefore also (6.2), are equivalent to
(6.5) |
Using essentially the same argument one can also show that , which was defined in (6.1), is given by
(6.6) |
According to A.8, instead of (6.5), it is enough to prove that
(6.7) |
For we have and hence . For we have . Therefore the left hand side of (6.7) can be replaced with
which in combination with (6.6) shows that (6.7) is equivalent to
This finishes the proof. ∎
Appendix A Appendix
A.1. Some basic results regarding functions form a Hardy field
Lemma A.1.
Let be a Hardy field and let be of polynomial growth. If then .
Proof.
For , let denote the -fold finite difference of , that is, , , , and so on. If is the degree of then the function has degree . Moreover, since , we have for all . However, the only function from a Hardy field that has degree and for which all its values belong to is a constant function. That means that for some , which implies that is a polynomial of degree satisfying . ∎
Lemma A.2.
Let be a simply connected nilpotent Lie group, , a Hardy field, and of polynomial growth. If then one of the following two cases holds:
-
(i)
either ;
-
(ii)
or there exist and with such that and for all .
Proof.
Suppose is not an element of . This means there exists such that . In particular, . By A.1 the function is polynomial with . This finishes the proof. ∎
Define .
Lemma A.3.
Let be a Hardy field and assume have polynomial growth. Then there exist , , , and with the following properties:
-
(1)
;
-
(2)
for all either or there exists such that ;
-
(3)
for all ,
Proof.
Let us associate to every finite set of functions of polynomial growth a pair , which we will call the characteristic pair associated to , in the following way: The number is the maximal degree among degrees of functions in , i.e.,
and the number equals the number of functions in whose degree is , i.e.,
Using this notion of a characteristic pair, we can define a partial ordering on the set of finite subsets of functions in of polynomial growth: Given of polynomial growth we write if
-
•
either ,
-
•
or and ,
where , are the characteristic pairs associated to and respectively.
Recall, our goal is to show for any of polynomial growth there exist , , , and 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 for some . In this case we have . Let , , , , and . 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 is of the form and . Define and set
This yields a new collection of functions with the property that for all . We now distinguish two cases, the case when the characteristic pair of is the same as the characteristic pair of , and the case when .
If we are in the first case then there exists some function in of degree . By relabeling if necessary, we can assume without loss of generality that . Then has degree . Define and set for all . It is straightforward to check that satisfies . Therefore, by the induction hypothesis, we can find , , , and such that satisfy properties (1) and (2), and for all we have
Define , let , and set
Then, for we have
For we have
This shows that property (3) is satisfied. Property (2) holds by construction (and the fact that has degree but satisfies also ) and property (1) holds because grows faster than any for , which implies that also grows faster than any , .
Define
Lemma A.4.
Let be a Hardy field and assume have polynomial growth. Then there exists , , , and with the following properties:
-
(1)
;
-
(2)
for all either or there exists such that ;
-
(3)
For all with we have and for all with there exists an antiderivative of in ;
-
(4)
for all ,
In the proof of A.3 we associated to every finite set of functions of polynomial growth a pair , called the characteristic pair, which gave rise to a partial ordering on the set of finite subsets of functions from 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 is either of the from or for some then the conclusion of A.4 follows from A.3. Let us therefore assume that the characteristic pair associated to is with and that A.4 has already been proven for all satisfying .
By replacing with if necessary, we can assume without loss of generality that all functions in are eventually non-negative. Also, since functions from a Hardy field can always be reordered according to their growth, we can relabel such that . Define . Set , define for all and . It is straightforward to check that satisfies . By the induction hypothesis, we can find , , , and such that satisfy properties (1), (2), and (3), and for all we have
Next, set for all and for all . Let be the number in uniquely determined by the property that has an antiderivative in if , and has no antiderivative in if . In other words, is the largest number in for which . Then, for every , let be the number in such that the derivative of equals . Define and, for every , let be any antiderivative of . Let and take
Also, let be any polynomial with the property that it is an antiderivative of . Since and , we can write
By integrating we get
(A.1) |
where is some real constant. We can absorb into , since was chosen to be an arbitrary antiderivative of . Thus, (A.1) becomes
For we have
This shows that property (4) holds. Since satisfy properties (1) and for all , we have . For , it follows from L’Hôpital’s rule and that . Also, because . This shows that Property (1) holds. Properties (2), and (3) are straightforward to derive from the definition of . ∎
Corollary A.5.
Let be a Hardy field and assume have polynomial growth. Then there exists with such that satisfy Property (PW).
Proof.
Let , , , and be as guaranteed by A.4. According to (2), for all either or there exists such that . If for some then we must have , since . By discarding if it is the zero function, we can assume that for all there exists such that . Consider the functions , , and pick any function with the property that and for all . Any with these properties is as desired. ∎
Corollary A.6.
Let be a Hardy field, with , and assume have the property that
-
(0)
for all , , and with the function satisfies either or for some .
Then there exists , , , and with the following properties:
-
(1)
;
-
(2)
for all there exists such that ;
-
(3)
for all with we have ;
-
(4)
for all ,
Proof.
This follows straightaway from A.4. ∎
Remark A.7.
It follows from 2.2 that if and only if . Therefore, if 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 be a sequence of positive real numbers that is either non-decreasing or non-increasing. Let and assume
Then for every there exists such that for every arithmetic function bounded in modulus by and with the property that for every the limit
exists, we have
(A.2) |
Proof.
Assume is non-decreasing, i.e., for all . For the case when is non-increasing similar arguments apply. We claim that for any bounded function we have
(A.3) |
Let be bounded. For the proof of (A.3) we can assume without loss of generality that . After an index-shift we obtain
Using we can estimate
The sum is telescoping and equals . We are left with
The Claim now follows from the assumption as .
Next, fix any bounded by . Using (A.3), we get for all that
By Jensen’s inequality we have
Moreover, we can write
In summary, we have shown that
It is now not hard to show that for every there exists such that
(A.4) |
from which (A.2) follows. Indeed, implies for all and that
Hence
Choosing we obtain equation (A.4) with . ∎
References
Florian K. Richter
École Polytechnique Fédérale de Lausanne