Hostname: page-component-76fb5796d-dfsvx Total loading time: 0 Render date: 2024-04-27T12:34:53.407Z Has data issue: false hasContentIssue false

Bohr chaoticity of principal algebraic actions and Riesz product measures

Published online by Cambridge University Press:  06 March 2024

AI HUA FAN
Affiliation:
Wuhan Institute for Math & AI, Wuhan University, Wuhan 430072, PR China LAMFA, UMR 7352 CNRS, University of Picardie, 33 rue Saint Leu, 80039 Amiens, France (e-mail: ai-hua.fan@u-picardie.fr)
KLAUS SCHMIDT
Affiliation:
Mathematics Institute, University of Vienna, Oskar-Morgenstern-Platz 1, 1090 Vienna, Austria (e-mail: klaus.schmidt@univie.ac.at)
EVGENY VERBITSKIY*
Affiliation:
Mathematical Institute, Leiden University, P.O. Box 9512, 2300 RA Leiden, The Netherlands Korteweg-de Vries Institute for Mathematics, University of Amsterdam, Postbus 94248, 1090 GE Amsterdam, The Netherlands
Rights & Permissions [Opens in a new window]

Abstract

For a continuous $\mathbb {N}^d$ or $\mathbb {Z}^d$ action on a compact space, we introduce the notion of Bohr chaoticity, which is an invariant of topological conjugacy and which is proved stronger than having positive entropy. We prove that all principal algebraic $\mathbb {Z}$ actions of positive entropy are Bohr chaotic. The same is proved for principal algebraic actions of $\mathbb {Z}^d$ with positive entropy under the condition of existence of summable homoclinic points.

Type
Original Article
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - SA
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike licence (https://creativecommons.org/licenses/by-nc-sa/4.0), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the same Creative Commons licence is used to distribute the re-used or adapted article and the original article is properly cited. The written permission of Cambridge University Press must be obtained prior to any commercial use.
Copyright
© The Author(s), 2024. Published by Cambridge University Press

1 Introduction

Bohr chaoticity is a topological invariant introduced in [Reference Fan, Fan, Ryzhikov and Shen9] for topological dynamical systems. For defining this invariant, we recall that a sequence $\boldsymbol w=(w_n)_{n\ge 0}\in \ell ^{\infty } (\mathbb {N},\mathbb {C})$ is a non-trivial weight sequence if

(1.1) $$ \begin{align} \limsup_{N\to \infty} \frac{1}{N} \sum_{n=0}^{N-1}|w_n|>0. \end{align} $$

A non-trivial weight sequence $\boldsymbol {w}$ is orthogonal to a topological dynamical system $(X,T)$  if

(1.2) $$ \begin{align} \lim_{N\to\infty} \frac 1N \sum_{n=0}^{N-1} w_n f(T^nx) =0 \end{align} $$

for every continuous function $f\in C(X)$ and every $x\in X$ .

Definition 1.1. [Reference Fan, Fan, Ryzhikov and Shen9]

A topological dynamical system $(X, T)$ is Bohr chaotic if it is non-orthogonal to every non-trivial weight sequence $\boldsymbol {w}\in \ell ^{\infty }(\mathbb {N},\mathbb {C})$ . In other words, $(X,T)$ is Bohr chaotic if we can find, for every non-trivial weight sequence $\boldsymbol {w}\in \ell ^{\infty }(\mathbb {N},\mathbb {C})$ , a continuous function $g\in C(X)$ and a point $x \in X$ such that

(1.3) $$ \begin{align} \limsup_{N\to \infty} \frac{1}{N} \bigg|\!\sum_{n=0}^{N-1}w_n g(T^n x)\bigg|>0. \end{align} $$

Bohr chaotic systems must have positive entropy: for example, almost all $(\tfrac 12,\tfrac 12)$ Bernoulli sequences taking values $-1$ and $1$ are orthogonal to every topological dynamical system $(X,T)$ with zero entropy (see [Reference El Abdalaoui, Kułaga-Przymus, Lemańczyk and de la Rue5]). We list some further basic results on Bohr chaoticity, taken from [Reference Fan, Fan, Ryzhikov and Shen9]:

  • any extension of a Bohr chaotic topological dynamical system is Bohr chaotic;

  • if a topological dynamical system $(X,T)$ has a non-empty, closed, T-invariant subset $Y\subset X$ such that $(Y,T|_Y)$ is Bohr chaotic, then $(X,T)$ is Bohr chaotic;

  • no uniquely ergodic dynamical system is Bohr chaotic (this is generalized by Tal [Reference Tal31] to systems having at most countably many ergodic measures);

  • all affine toral endomorphisms of positive entropy are Bohr chaotic;

  • all systems having an m-order horseshoe, $m\ge 1$ , are Bohr chaotic. By an m-order horseshoe K of a system $(X, T)$ , we mean a $T^m$ -invariant closed non-empty set $K\subset X$ such that the subsystem $(K, T^m)$ is conjugate either to the one-sided shift $(\{0, 1\}^{\mathbb {N}}, \sigma )$ or to the two-sided shift $(\{0,1\}^{\mathbb {Z}}, \sigma )$ ;

  • all subshifts of finite type with positive entropy are Bohr chaotic;

  • all piecewise monotone $C^1$ interval maps of positive entropy are Bohr chaotic. For example, the $\beta $ -shifts with $\beta>1$ ;

  • every $C^{1+\delta }$ ( $\delta>0$ ) diffeomorphism of a compact smooth manifold admitting an ergodic non-atomic Borel probability invariant measure with non-zero Lyapunov exponents is Bohr chaotic.

The reason for the last two classes is that any such system admits a subsystem which is conjugate to a subshift of finite type of positive entropy [Reference Katok15, Reference Young32].

It is interesting to note that for the examples of Bohr chaotic systems constructed in [Reference Fan, Fan, Ryzhikov and Shen9], the sets of points $x\in X$ satisfying equation (1.3) are large in the sense that they are of full Hausdorff dimension. Actually, weighted ergodic averages on typical dynamical systems would be multifractal and a study on symbolic spaces is carried out in [Reference Fan7].

In the present paper, we extend the notion of Bohr chaoticity from ${\mathbb Z}$ - to ${\mathbb Z}^d$ -actions and prove that a large class of algebraic dynamical systems—the so-called principal algebraic actions—are Bohr chaotic, provided they have positive entropy.

By analogy with equation (1.1), we say that a complex sequence $\boldsymbol w=(w_{\boldsymbol n})_{\boldsymbol n \in \mathbb {N}^d}\in \ell ^{\infty }(\mathbb {N}^d,\mathbb C)$ is a non-trivial weight if

$$ \begin{align*} \limsup_{N\to \infty} \frac{1}{N^d} \sum_{{\boldsymbol n} \in [0, N-1]^d} |w_{\boldsymbol n}|>0. \end{align*} $$

Consider a continuous $\mathbb {N}^d$ - or $\mathbb {Z}^d$ -action $\alpha \colon \boldsymbol {n}\mapsto \alpha ^{\boldsymbol {n}}$ on some compact space X. As in equation (1.2), say that a (non-trivial) weight $(w_{\boldsymbol n})_{\boldsymbol n \in \mathbb {N}^d}$ is orthogonal to the dynamical system $(X, \alpha )$ if

(1.4) $$ \begin{align} \lim_{N\to \infty} \frac{1}{N^d} \sum_{{\boldsymbol n} \in [0, N-1]^d} w_{\boldsymbol n} g (\alpha ^{\boldsymbol n} x) =0 \end{align} $$

for every continuous function $g \in C(X)$ and every point $x\in X$ .

Definition 1.2. If $\alpha $ is a continuous $\mathbb {N}^d$ - or ${\mathbb Z}^d$ -action on a compact space X, we call $(X, \alpha )$ Bohr chaotic if it is not orthogonal to any non-trivial weight, that is to say, if for any non-trivial weight $\boldsymbol w=(w_{\boldsymbol n})_{\boldsymbol n \in \mathbb {N}^d}$ , there exist $g \in C(X)$ and $x\in X$ such that

(1.5) $$ \begin{align} \limsup_{N\to \infty} \frac{1}{N^d} \bigg|\!\sum_{{\boldsymbol n} \in [0, N-1]^d} w_{\boldsymbol n} g(\alpha ^{\boldsymbol n} x)\bigg|>0. \end{align} $$

Note that if $\alpha $ is a continuous $\mathbb {N}^d$ -action on X and if $(\bar {X},\bar {\alpha })$ is the natural extension of $(X,\alpha )$ to a continuous $\mathbb {Z}^d$ -action $\bar {\alpha }$ on a compact space $\bar {X}$ , then $(\bar {X},\bar {\alpha })$ is Bohr chaotic if and only if the same is true for $(X,\alpha )$ . Conversely, if a continuous $\mathbb {Z}^d$ -action is Bohr chaotic, it is obviously also Bohr chaotic as an $\mathbb {N}^d$ -action. In view of this last property, we focus our attention in much of this paper on Bohr chaoticity of $\mathbb {Z}^d$ -actions, referring to $\mathbb {N}^d$ -actions only where necessary (like in Proposition 3.1 or Example 3.3).

As in the one-dimensional case, one can easily verify the following properties of continuous $\mathbb {Z}^d$ -actions $(X,\alpha )$ :

  1. (i) if X has a closed, $\alpha $ -invariant subset Y such that $(Y,\alpha |_Y)$ is Bohr chaotic, then $(X,\alpha )$ is Bohr chaotic;

  2. (ii) if $(X,\alpha )$ has a Bohr chaotic factor $(Y,\beta )$ (that is, if $(Y,\beta )$ is a Bohr chaotic $\mathbb {Z}^d$ -action and there exists a continuous, surjective, equivariant map $\phi \colon X\to Y$ ), then $(X,\alpha )$ is Bohr chaotic.

In particular, Bohr chaoticity is an invariant of topological conjugacy.

Our main results will be proved by using Riesz product measures borrowed from harmonic analysis and the main technical tool is the notion of m-goodness.

The paper is organized as follows. In §2, we present algebraic ${\mathbb Z}^d$ -actions and their basic properties, state our main results on Bohr chaoticity of principal algebraic $\mathbb {Z}^d$ -actions (Theorems 2.1 and 2.3), and prove that Bohr chaotic algebraic $\mathbb {Z}^d$ -actions have to have completely positive entropy (Example 3.2). In §3, we show that zero-entropy $\mathbb {Z}^d$ -actions are not Bohr chaotic. Our main tool, Riesz products, is presented in §4 where lacunarity of polynomials is discussed. In §5, we prove that any principal algebraic ${\mathbb Z}^d$ -action defined by a so-called m-good polynomial is Bohr chaotic (Theorem 5.1). Section 6 is devoted to the proof of m-goodness for all irreducible polynomials $f\in R_1$ with positive Mahler measure, and Theorem 2.1 (for $d=1$ ) is proved there. Theorem 2.3 (for $d \ge 2$ ) is proved in §7, where we prove a gap theorem (Theorem 7.3) for irreducible polynomials which admit summable homoclinic points. This gap theorem is of independent interest. In §8, we speculate briefly on the necessity of our atorality assumptions for our main results and give some examples of Bohr chaotic principal actions arising from toral polynomials.

2 Algebraic ${\mathbb Z}^d$ -actions

In this section, we present the principal algebraic ${\mathbb Z}^d$ -actions which are our main objects of study and then state our main results (Theorems 2.1 and 2.3).

An algebraic $\mathbb {Z}^d$ -action is an action of $\mathbb {Z}^d$ by (continuous) automorphisms of a compact metrizable abelian group. Algebraic $\mathbb {Z}^d$ -actions provide a useful source of examples of continuous $\mathbb {Z}^d$ -actions with a wide range of properties, both with zero and with positive entropy, and with or without Bohr chaoticity.

We are interested in a particular family of algebraic $\mathbb {Z}^d$ -actions, the so-called cyclic actions. Denote by $\sigma $ the shift-action of $\mathbb {Z}^d$ on $\mathbb {T}^{\mathbb {Z}^d}$ given by

(2.1) $$ \begin{align} \sigma ^{\boldsymbol{m}}(x)_{\boldsymbol{n}} = x_{\boldsymbol{n}+\boldsymbol{m}} \end{align} $$

for every $x=(x_{\boldsymbol {n}})_{\boldsymbol {n}\in \mathbb {Z}^d}\in \mathbb {T}^{\mathbb {Z}^d}$ . A cyclic algebraic $\mathbb {Z}^d$ -action is a pair $(X,\alpha _X)$ , where $X\subset \mathbb {T}^{\mathbb {Z}^d}$ is a closed, shift-invariant subgroup and $\alpha _X=\sigma |_X$ is the restriction to X of the shift-action $\sigma $ in equation (2.1).

To describe these actions in more detail, we denote by $R_d = \mathbb {Z}[z_1^{\pm 1},\ldots ,z_d^{\pm 1}]$ the ring of Laurent polynomials in the variables $z_1,\ldots ,z_d$ with coefficients in $\mathbb {Z}$ . Every $f\in R_d$ will be written as $f=\sum _{{\boldsymbol n}\in \mathbb {Z}^d}f_{\boldsymbol n} {{\boldsymbol z}}^{\boldsymbol n}$ with $f_{\boldsymbol n}\in \mathbb {Z}$ and $\boldsymbol {z}^{\boldsymbol n}=z_1^{n_1}\cdots z_d^{n_d}$ for every ${\boldsymbol n}=(n_1,\ldots ,n_d)\in \mathbb {Z}^d$ . The set $\textsf {supp}(f) = \{{\boldsymbol n}\in \mathbb {Z}^d\mid f_{\boldsymbol n} \ne 0\}$ will be called the support of f, and we set $\|f\|_1=\sum _{{\boldsymbol n}\in \mathbb {Z}^d}|f_{\boldsymbol n}|$ and $\|f\|_\infty =\max _{{\boldsymbol n}\in \mathbb {Z}^d}|f_{\boldsymbol n}|$ . Following standard terminology, we call a non-zero element $f\in R_d$ primitive if the greatest common divisor $\gcd (\{f_{\boldsymbol n}\mid {\boldsymbol n}\in \mathbb {Z}^d\})$ of its coefficients is equal to $1$ , and irreducible if it is not a product of two non-units in $R_d$ . For example, $2{\boldsymbol z}^{\boldsymbol n}$ is irreducible, but not primitive, while $2(1+{\boldsymbol z}^{\boldsymbol n})$ is neither primitive nor irreducible, for every ${\boldsymbol n}\in \mathbb {Z}^d$ .

Every non-zero $f=\sum _{{\boldsymbol n}\in \mathbb {Z}^d}f_{\boldsymbol n} {{\boldsymbol z}}^{\boldsymbol n}\in R_d$ defines a surjective group homomorphism $f(\sigma )=\sum _{\boldsymbol {m}\in \mathbb {Z}^d}f_{\boldsymbol {m}}\sigma ^{\boldsymbol {m}}\colon \mathbb {T}^{\mathbb {Z}^d} \to \mathbb {T}^{\mathbb {Z}^d}$ . Consider the closed, shift-invariant subgroup

(2.2) $$ \begin{align} X_f =\bigg\{ x\in{\mathbb T}^{\mathbb Z^d}\bigg| \sum_{{\boldsymbol m}\in{\mathbb Z}^d} x_{{\boldsymbol n}+{\boldsymbol m}} f_{{\boldsymbol m}} =0 \enspace (\text{mod}\;1)\enspace \text{for all}\enspace {\boldsymbol n}\in{\mathbb Z}^d\bigg\} = \ker(f(\sigma ))\subset \mathbb{T}^{\mathbb{Z}^d}, \end{align} $$

and denote by

(2.3) $$ \begin{align} \alpha _f = \sigma |_{X_f} \end{align} $$

the restriction to $X_f$ of the shift-action $\sigma $ on $\mathbb {T}^{\mathbb {Z}^d}$ . The dynamical system $(X_f,\alpha _f)$ is called the principal algebraic action corresponding to $f\in R_d$ .

Formally we can extend this definition of a principal action $(X_f,\alpha _f)$ to include the case $f=0$ , the zero polynomial in $R_d$ . In this case, the definitions in equations (2.2) and (2.3) reduce to $X_f=\mathbb {T}^{\mathbb {Z}^d}$ and $\alpha _f = \sigma $ , that is, $(X_f,\alpha _f)$ is simply the shift action of $\mathbb {Z}^d$ on $\mathbb {T}^{\mathbb {Z}^d}$ .

For every cyclic algebraic action $(X,\alpha _X)$ with $X\subsetneq \mathbb {T}^{\mathbb {Z}^d}$ , the set

(2.4) $$ \begin{align} I_X=\{f\in R_d\mid X\subset X_f\} \end{align} $$

is an ideal in $R_d$ (which is, of course, finitely generated since the ring $R_d$ is Noetherian) and $X= \bigcap _{f \in I_X} X_f$ . Conversely, if $I\subset R_d$ is an ideal, generated by $\{f^{(1)},\ldots ,f^{(r)}\}$ , say, we denote by $(X_I,\alpha _I)$ the cyclic $\mathbb {Z}^d$ -action defined by

(2.5) $$ \begin{align} X_I=\bigcap_{f\in I} X_f = \bigcap_{i=1}^r X_{f^{(i)}}\subset \mathbb{T}^{\mathbb{Z}^d}\quad \text{and}\quad \alpha _I=\sigma |_{X_I}, \end{align} $$

and write $\unicode{x3bb} _I$ for the normalized Haar measure of $X_I$ . If the ideal $I\subset R_d$ is principal, $I=(f)$ , say, we write $(X_f,\alpha _f)$ instead of $(X_{(f)},\alpha _{(f)})$ and denote by $\unicode{x3bb} _f$ the normalized Haar measure on $X_f$ .

2.1 Mahler measure

The topological entropy $h_{\text {top}}(\alpha _f)$ of a principal algebraic action $(X_f,\alpha _f)$ , $f\in R_d\smallsetminus \{0\}$ , coincides with its measure-theoretic entropy $h_{\unicode{x3bb} _f}(\alpha _f)$ and is given by the (logarithmic) Mahler measure of its defining polynomial f:

(2.6)

For polynomials in a single variable (that is, for $f\in R_1$ ), Mahler measure can be computed using Jensen’s formula: let $f(z)=f_0+f_1z +\cdots +f_k z^k$ with $f_0f_k\ne 0$ and (complex) roots $\unicode{x3bb} _1,\ldots , \unicode{x3bb} _k$ . Then

(2.7) $$ \begin{align} \mathsf{m}(f)=\log|f_k| + \sum_{j:\ |\unicode{x3bb}_j|>1} \log|\unicode{x3bb}_j| \end{align} $$

(cf. [Reference Lind, Schmidt and Ward21, p. 597], [Reference Schmidt29, equation (16.2)], or [Reference Young33]). The Kronecker lemma [Reference Kronecker17] states that if a polynomial $f\in R_1$ is irreducible, monic, and all its roots have absolute value at most $1$ , then f is cyclotomic, that is, for some integer n,

(2.8) $$ \begin{align} f(z)=\Phi_n(z) := \prod_{\substack{1\le \ell\le n \\ \textsf{gcd}(\ell,n)=1}} (z-e^{2 \pi i\ell/n}). \end{align} $$

Since Mahler measure is additive in the sense that $\mathsf {m}(f\cdot g)=\mathsf {m}(f)+\mathsf {m}(g)$ for all ${f,g\in R_d\smallsetminus \{0\}}$ , any $f\in R_1\smallsetminus \{0\}$ with $\mathsf {m}(f)=0$ must be a product of cyclotomic polynomials.

A similar statement is true for multivariate polynomials as well: if $f\in R_d\smallsetminus \{0\}$ , then $\mathsf {m}(f)=0$ if and only if f is a product of so-called generalized cyclotomic polynomials

$$ \begin{align*} f({\boldsymbol z})=\pm {{\boldsymbol z}^{{\boldsymbol n}_0}}\Phi_{m_1}({\boldsymbol z}^{{\boldsymbol n}_1} ) \cdots \Phi_{m_r}({\boldsymbol z}^{{\boldsymbol n}_r} ) \end{align*} $$

for some integers $m_1,\ldots , m_r$ and ${\boldsymbol n}_0,{\boldsymbol n}_1,\ldots , {\boldsymbol n}_r\in {\mathbb Z}^d$ ([Reference Boyd2], [Reference Schmidt29, Theorem 19.5], [Reference Smyth30]).

We recall the following properties of cyclic algebraic $\mathbb {Z}^d$ -action $(X_I,\alpha _I)$ (cf. [Reference Schmidt29, Ch. 6]).

  • The normalized Haar measure $\unicode{x3bb} _{X_I}$ of $X_I$ is shift-invariant.

  • If $I\subset R_d$ is non-zero and principal, $I=(f)$ , say, the topological entropy of $(X_f,\alpha _f)$ is given by the Mahler measure in equation (2.6) of f; if $I\subset R_d$ contains at least two non-zero elements $f,g$ which are relatively prime to each other (that is, without a non-trivial common factor), then $h_{\text {top}}(X_I,\alpha _I)=0$ .

  • If $d>1$ , every principal $\mathbb {Z}^d$ -action $(X_f,\alpha _f)$ is ergodic (with respect to $\unicode{x3bb} _f$ ); if $d=1$ , a principal $\mathbb {Z}$ -action $(X_f,\alpha _f)$ is ergodic if and only if f has no cyclotomic divisor.

  • For every non-zero and irreducible $f\in R_d$ , the following conditions are equivalent:

    1. $\unicode{x3bb} _f$ is mixing under $(X_f,\alpha _f)$ ;

    2. $h_{\text {top}}(X_f,\alpha _f)>0$ .

2.2 Main results

Our main results are the following theorems which will be proved in §§6 and 7.

Theorem 2.1. Suppose $f\in R_1\smallsetminus \{0\}$ with $\mathsf {m}(f)>0$ . Then the principal algebraic ${\mathbb Z}$ -action $(X_f,\alpha _f)$ is Bohr chaotic.

For the higher dimensional case, we need an extra condition.

Definition 2.2. [Reference Lind, Schmidt and Verbitskiy20]

A non-zero Laurent polynomial $f\in R_d$ is atoral if it is not a unit in $R_d$ and its unitary variety

$$ \begin{align*} \mathsf U(f)=\{(t_1,\ldots ,t_d)\in \mathbb{T}^d\mid f(e^{2\pi it_1},\ldots ,e^{2\pi it_d})=0\} \end{align*} $$

of f has dimension $\le d-2$ . This includes the possibility that $\mathsf U(f)=\varnothing $ , which is equivalent to expansivity of the $\mathbb {Z}^d$ -action $\alpha _f$ . If $\mathsf U(f)$ has dimension $d-1$ , f is called toral.

With this definition, the following is true.

Theorem 2.3. Suppose that $d\ge 2$ and that $f\in R_d$ is atoral. Then $h_{\text {top}}(X_f,\alpha _f)>0$ and $(X_f,\alpha _f)$ is Bohr chaotic.

We illustrate these definitions with a few examples.

Example 2.4. (Toral automorphisms)

We start with a special case: let $f=f_0+\cdots +f_kz^k \in R_1$ with $k\ge 1$ and $f_k=|f_0|=1$ . Then the principal $\mathbb {Z}$ -action $(X_f,\alpha _f)$ is conjugate to the toral automorphism $(\mathbb {T}^k,A_f)$ , where

(2.9)

is the companion matrix of f. The map $\phi \colon X_f \to \mathbb {T}^k$ , defined by

$$ \begin{align*} \phi (x)= \left( \begin{smallmatrix} x_0 \\ \vdots \\ x_{k-1} \end{smallmatrix}\right) \end{align*} $$

for every $x=(x_n)_{n\in \mathbb {Z}}$ , implements this conjugacy. We conclude that $(X_f,\alpha _f)$ and thus $(\mathbb {T}^k,A_f)$ is Bohr chaotic if and only if $\mathsf {m}(f)>0$ (cf. equation (2.6)).

Consider now an irreducible toral automorphism $T_A\colon {\mathbb T}^k\to {\mathbb T}^k$ defined by a matrix $A\in GL_k({\mathbb Z})$ . Then the characteristic polynomial $f(z)$ of A is irreducible, and the principal algebraic $\mathbb {Z}$ -action $({\mathbb T}^k,A_f) \cong (X_f,\alpha _f)$ in equation (2.9) is a finite-to-one factor of $({\mathbb T}^d,T_A)$ . Hence, if $({\mathbb T}^k, A_f)$ is Bohr chaotic (which is the case if and only if $\mathsf (m)(f)>0$ ), then $({\mathbb T}^k,T_A)$ is also Bohr chaotic (as an extension).

If a toral automorphism $T_A\colon {\mathbb T}^k\to {\mathbb T}^k$ with $A\in GL_k({\mathbb Z})$ is reducible (that is, has a proper invariant subtorus $V\subsetneq \mathbb {T}^k$ ), then the characteristic polynomial g of $T_A|_V$ will be a proper factor of f. If $g'$ is one of the irreducible factors of g (and hence of f), then the system $(\mathbb {T}^{\text {deg}(g')},T_{A_{g'}})$ will be Bohr chaotic if and only if $\mathsf {m}(g')>0$ , in which case both $(V,T_A|_V)$ and $(\mathbb {T}^d,T_A)$ will be Bohr chaotic. By varying V over the A-invariant irreducible subtori of $\mathbb {T}^k$ , we see that $(\mathbb {T}^k,T_A)$ is Bohr chaotic if and only if $h_{\text {top}}(T_A) = \mathsf {m}(f)>0$ .

Example 2.5. (Constant polynomials)

Suppose that $f=p\in \mathbb {N}$ , $p>1$ , viewed as a constant polynomial in $R_d$ . Then the principal algebraic action $(X_f,\alpha _f)$ arising from this polynomial is the shift-action in equation (2.1) on , which is certainly Bohr chaotic. If $p=1$ (or, more generally, if f is a unit in $R_d$ ), then $X_f$ reduces to a single point and the $\mathbb {Z}^d$ -action $\alpha _f$ becomes trivial. By default, $(X_f,\alpha _f)$ is not Bohr chaotic.

Example 2.6. (The zero polynomial)

So far, we have always assumed that the polynomial $f\in R_d$ defining a principal algebraic action $(X_f,\alpha _f)$ is non-zero. If we deviate from this assumption and set $f=0$ (the zero polynomial in $R_d$ ), then equation (2.2) reduces to $X_f=\mathbb {T}^{\mathbb {Z}^d}$ , and $\alpha _f$ becomes the shift action $\sigma $ of $\mathbb {Z}^d$ on $\mathbb {T}^{\mathbb {Z}^d}$ . For every integer $p>1$ , $\mathbb {T}^{\mathbb {Z}^d}$ contains the closed, shift invariant subset in Example 2.5. Since $(X_p,\sigma )$ is Bohr periodic, the same is true for $(X_f,\alpha _f) = (\mathbb {T}^{\mathbb {Z}^d},\sigma )$ .

Remark 2.7. To prove our Theorems 2.1 and 2.3, we may assume without loss of generality that the polynomial f in either of these theorems is primitive and irreducible.

Indeed, if f is not primitive, then $f = pg$ for some primitive polynomial $g\in R_d$ , and $(X_f,\alpha _f)$ has the subsystem $(X_p,\alpha _p= \sigma |_{X_p})$ appearing in Example 2.5. Since $(X_p,\alpha _p)$ is Bohr chaotic, the same holds for $(X_f,\alpha _f)$ .

Similarly, if $f\in R_d$ is reducible and $\mathsf {m}(f)>0$ , then at least one of the irreducible factors g of f has positive Mahler measure $\mathsf {m}(g)>0$ . If $(X_g,\alpha _g)$ is Bohr chaotic, the Bohr chaoticity of $(X_f,\alpha _f)$ follows immediately from Bohr chaoticity of the subsystem $(X_g,\alpha _g)$ . For $d\ge 2$ in Theorem 2.3, we also note that atorality of a polynomial $f\in R_d$ is inherited by all its irreducible factors.

3 ${\mathbb Z}^d$ actions of zero entropy

Before going on to study our Bohr chaotic principal algebraic $\mathbb {Z}^d$ -actions, we would like to justify that any continuous ${\mathbb N}^d$ - or ${\mathbb Z}^d$ -action with zero topological entropy is not Bohr chaotic (Proposition 3.1). This is an immediate consequence of the disjointness completely positive entropy systems and zero entropy systems. Also, we would like to point out that principal ${\mathbb Z}$ -actions with zero topological entropy are disjoint from the Möbius function (Proposition 3.5).

3.1 Zero entropy $\mathbb {Z}^d$ -actions are not Bohr chaotic

Consider a measure-preserving $\mathbb {N}^d$ - or ${\mathbb Z}^d$ -action $\gamma $ on a Lebesgue space $(\Omega , \mu )$ , where $\Omega $ is a compact space equipped with its Borel field. We say that the measure-theoretic system $(\Omega , \mu , \gamma )$ has completely positive entropy if any non-trivial factor of $(\Omega , \mu , \gamma )$ has positive entropy. Bernoulli systems have complete positive entropy. For $d=1$ , the following result is folklore; for $d\ge 1$ , we include a proof for completeness, based on a disjointness theorem due to Glasner, Thouvenot, and Weiss [Reference Glasner, Thouvenot and Weiss12, Theorem 1].

Proposition 3.1. Suppose that $(\Omega , \mu , \gamma )$ has completely positive entropy, $\omega \in \Omega $ is a $\mu $ -generic point, and $\phi \in C(\Omega )$ is a continuous function having zero mean. Then $(\phi (\gamma ^{{\boldsymbol n}} \omega ))_{{\boldsymbol n} \in \mathbb {N}^d}$ is orthogonal to every zero entropy $\mathbb {N}^d$ - or $\mathbb {Z}^d$ -action $(X, \alpha )$ . That is to say, for every $f\in C(X)$ and every $x\in X$ , we have

(3.1) $$ \begin{align} \lim_{N\to\infty}\frac{1}{N^d}\sum_{{\boldsymbol n} \in [0, N-1]^d} \phi(\gamma ^{{\boldsymbol n}}\omega ) f(\alpha ^{\boldsymbol n} x)=0. \end{align} $$

In particular, continuous ${\mathbb N}^d$ - or ${\mathbb Z}^d$ -actions with zero topological entropy are not Bohr chaotic.

Proof. Suppose that for some f and some x, there exists a sequence $(N_j)$ tending to infinity such that

We can assume that along this sequence $(N_j)$ , the following weak limits of measures exist:

where $\delta _\omega $ and $\delta _x$ denote the point masses at the points $\omega $ and x, respectively. Clearly, the measure $\unicode{x3bb} $ is $\gamma \negthinspace \times \negthinspace \alpha $ -invariant, and the projection of $\unicode{x3bb} $ on X is equal to $\nu $ . Since $\omega $ is $\mu $ -generic, the projection of $\unicode{x3bb} $ onto $\Omega $ is equal to $\mu $ . In other words, $\unicode{x3bb} $ is a joining of $\mu $ and $\nu $ , where $\nu $ has zero entropy. Since systems of completely positive entropy are disjoint from systems of zero entropy by [Reference Glasner, Thouvenot and Weiss12, Theorem 1], we obtain that $\unicode{x3bb} =\mu \negthinspace \times \negthinspace \nu $ . Thus, by the definition of $\unicode{x3bb} $ and the hypothesis that $\mathbb {E}_\mu \phi =0$ , we get that

$$ \begin{align*} \ell = \mathbb{E}_\unicode{x3bb} (\phi \otimes f) = \mathbb{E}_\mu \phi \cdot \mathbb{E}_\nu f =0, \end{align*} $$

which is a contradiction.

Corollary 3.2. Let $(X,\alpha )$ be an algebraic $\mathbb {Z}^d$ -action which does not have completely positive entropy (with respect to the Haar measure $\unicode{x3bb} _X$ ). Then $(X,\alpha )$ is not Bohr chaotic.

Proof. If $(X,\alpha )$ does not have completely positive entropy, then [Reference Schmidt29, Theorem 20.8] implies that there exists a non-trivial closed, $\alpha $ -invariant subgroup $Y\subset X$ such that the $\mathbb {Z}^d$ -action $\alpha _{X/Y}$ induced by $\alpha $ on $X/Y$ has zero entropy. Condition (ii) at the beginning of §3, combined with Proposition 3.1, shows that $(X,\alpha )$ cannot be Bohr chaotic.

Example 3.3. (Furstenberg’s example)

Let $d=2$ and let $I = (2-z_1, 3-z_2)\subset R_2$ . Then $X_I=\{x\in \mathbb {T}^{\mathbb {Z}^2}\mid \sigma ^{(1,0)}x=2x,\,\sigma ^{(0,1)}x=3x\}$ , so that $x_{k,l}=2^k3^lx_{(0,0)}$ for every $x\in X_I$ and $(k,l)\in \mathbb {Z}^2$ . Since $f^{(2)}=2-z_1$ and $f^{(3)}=3-z_2$ are irreducible and relatively prime to each other, I is a prime ideal, and hence $h_{\text {top}}(X_I,\alpha _I)=0$ [Reference Schmidt29, Proposition 17.5].

If $\gamma $ is a continuous $\mathbb {Z}^2$ -action on a compact space $\Omega $ , $\mu $ is a probability measure on $\Omega $ with completely positive entropy under $\gamma $ , $\omega \in \Omega $ is a $\mu $ -generic point, and $\phi \in C(\Omega )$ has mean zero, Proposition 3.1 shows that

$$ \begin{align*} \lim_{N\to\infty } \frac{1}{N^2} \sum_{(m,n)\in [0,N-1]^2} \phi (\gamma ^{(m,n)}\omega )h(2^m3^nt)=0 \end{align*} $$

for every $h\in C(\mathbb {T})$ and $t\in \mathbb {T}$ .

In [Reference Furstenberg11], Furstenberg’s example was defined as the $\mathbb {N}^2$ -action $\alpha $ on $X=\mathbb {T}$ given by

$$ \begin{align*} \alpha ^{(m,n)}t=2^m3^nt\ (\text{mod 1}) \end{align*} $$

for every $(m,n)\in \mathbb {N}^2$ and $t\in \mathbb {T}$ .

We set $\Omega =\mathbb {T}^{\mathbb {N}^d}$ , write the coordinates of every $\omega =(\omega _{\boldsymbol n})_{{\boldsymbol n}\in \mathbb {N}^d}\in \Omega $ in the form $\omega _{\boldsymbol n} = (\omega _{\boldsymbol n}^{(1)},\ldots ,\omega _{\boldsymbol n}^{(d)})$ , and denote by $\gamma $ the one-sided shift-action of $\mathbb {N}^d$ on $\Omega $ (cf. equation (2.1)). According to Franklin [Reference Franklin10], for Lebesgue-almost everywhere (a.e.) $(\beta _1, \ldots , \beta _d)$ with $\beta _1>1, \ldots , \beta _d>1$ , the point $\beta =(\beta _{\boldsymbol n})_{{\boldsymbol n}\in \mathbb {N}^d} \in \Omega $ with ${\beta _{\boldsymbol n} = (\beta _1^{n_1}\;(\text {mod}\,1), \ldots , \beta _d^{n_d}\;(\text {mod}\,1))}$ for every ${\boldsymbol n}\in \mathbb {N}^d$ is Lebesgue-generic for $\gamma $ on $\Omega $ . If $\phi \colon \Omega \to \mathbb {C}$ is the map defined by

$$ \begin{align*} \phi (\omega )= e^{2\pi i(\omega_{\boldsymbol{0}}^{(1)} + \cdots + \omega_{\boldsymbol{0}}^{(d)})}, \end{align*} $$

then

$$ \begin{align*} \phi(\gamma ^{{\boldsymbol n}}\beta ) =e^{2\pi i(\beta _1^{n_1}+ \cdots +\beta_d^{n_d})} \end{align*} $$

for every ${\boldsymbol n} = (n_1,\ldots ,n_d)\in \mathbb {N}^d$ . By Proposition 3.1, the sequence $(\phi (\gamma ^{{\boldsymbol n}}\beta ))_{{\boldsymbol n}\in \mathbb {N}^d}$ is almost surely orthogonal to all systems of zero entropy. Since Furstenberg’s example $(\mathbb {T},\alpha )$ described in the preceding paragraph has zero entropy, we obtain the following corollary of Proposition 3.1.

Corollary 3.4. For almost all $(\beta _1, \beta _2)$ with $\beta _1>1$ and $\beta _2>1$ ,

$$ \begin{align*} \lim_{N\to\infty}\frac{1}{N^2}\sum_{0\le m, n<N} e^{2\pi i (\beta_1^m + \beta_2^n)} f(2^m 3^n t)=0 \end{align*} $$

for every continuous function $f \in C(\mathbb {T})$ and every $t\in \mathbb {T}$ .

3.2 Möbius disjointness and principal actions

We have just shown that zero entropy ${\mathbb Z}^d$ -actions are not Bohr chaotic. In fact, for principal actions, the result can be strengthened. We recall that a topological dynamical system $(X,T)$ is Möbius disjoint if

(3.2) $$ \begin{align} \lim_{n\to\infty} \frac 1n \sum_{k=1}^n \mu (k) f(T^kx) =0\quad \text{for every } f\in C(X) \text{ and every } x\in X. \end{align} $$

Proposition 3.5. A zero entropy principal ${\mathbb Z}$ -algebraic action $(X_f,\alpha _f)$ , $f\in R_1$ , is Möbius disjoint.

Proof. Since $(X_f,\alpha _f)$ has zero entropy, that is, $\mathsf {m}(f)=0$ , the Kronecker lemma implies that f has the form

(3.3) $$ \begin{align} f(z)= \pm z^{m_0} \Phi_{n_1}(z^{m_1})\cdots \Phi_{n_k}(z^{m_k}), \end{align} $$

where $m_0\in {\mathbb Z}$ , $n_j,m_j\in {\mathbb N}$ , $j=1,\ldots ,k$ , and $\Phi _n$ is the nth cyclotomic polynomial defined in equation (2.8). One immediately concludes from equation (3.3) that

$$ \begin{align*} f(z)=a_0+a_1z+\cdots+a_{N}z^N\quad \text{with } |a_0|=|a_N|=1, \end{align*} $$

and hence $(X_f,\alpha _f)$ is topologically conjugate to the toral automorphism $({\mathbb T}^{N},T_A)$ , where $T_A:{\mathbb T}^{N}\to {\mathbb T}^{N}$ is a linear automorphism with the matrix $A=A_{f}$ —the companion matrix of f, see Example 2.4. However, toral automorphisms with zero entropy are known to be Möbius disjoint [Reference Liu and Sarnak23, Theorem 1.1]. In fact, toral automorphisms, and more generally affine maps of compact abelian groups, are the primary examples motivating Sarnak’s (still unproven) conjecture that all topological dynamical systems with zero entropy are Möbius disjoint [Reference Sarnak28].

4 Riesz product measures on $X_f$

In this section, we start on the proofs of Theorems 2.1 and 2.3. As explained in Remark 2.7, we assume from now on—and without loss in generality—that the polynomial $f\in R_d$ defining our principal action $(X_f,\alpha _f)$ is primitive, irreducible, and has positive Mahler measure.

For the proofs of these theorems, we shall use a class of measures called Riesz products. First, we will recall the general construction of Riesz product measures on arbitrary compact abelian groups. Second, we will construct Riesz products on $X_f$ based on lacunary polynomials in the dual group $\widehat {X}_f \subset R_d$ .

4.1 Riesz product measures

Let X be a compact abelian group with dual group $\widehat X$ .

Definition 4.1. [Reference Hewitt and Zuckerman13]

An infinite sequence of distinct characters $\Lambda =(\gamma _n)_{n\in \mathbb {N}}=\{\gamma _0, \gamma _1,\ldots \,\}\subset \widehat X$ is said to be dissociate if for every $k\ge 1$ and every k-tuple $(n_1,n_2,\ldots , n_k)\in \mathbb {N}^k$ of distinct non-negative integers, the equality

$$ \begin{align*} \gamma _{n_1}^{\varepsilon _1}\gamma _{n_2}^{\varepsilon _2}\cdots \gamma _{n_k}^{\varepsilon _k}=1 \end{align*} $$

with $\varepsilon _j\in \{-2,-1,0,1,2\}$ for every $j=1,\ldots ,k$ , implies that

$$ \begin{align*} \gamma _{n_1}^{\varepsilon _1}=\gamma _{n_2}^{\varepsilon _2}=\cdots =\gamma _{n_k}^{\varepsilon _k}=1. \end{align*} $$

Equivalently, $\Lambda $ is dissociate if any character in $\widehat X$ can be represented in at most one way as a finite product $ \gamma _{n_1}^{\varepsilon _1}\gamma _{n_2}^{\varepsilon _2}\cdots \gamma _{n_k}^{\varepsilon _k} $ of elements of $\Lambda $ , where all $n_j$ are distinct and ${\varepsilon _j\in \{-1, 0,1\}}$ .

Using dissociate sequences of characters, Hewitt and Zuckermann [Reference Hewitt and Zuckerman13] proposed a construction of interesting probability measures—the so-called Riesz products, generalizing Riesz products on ${\mathbb T}$ constructed by Riesz [Reference Riesz27] in 1918. More precisely, denote by $\unicode{x3bb} _X$ the Haar measure on X. Suppose that:

  1. (i) $\Lambda = (\gamma _n)_{n\ge 0}$ is a dissociate sequence of characters in $\widehat X$ ;

  2. (ii) $a = (a_n)_{n\ge 0}$ is a sequence of complex numbers such that $ |a_n|\le 1$ for all n.

For any $N\ge 0$ , denote by $\mu _a^{(N)}$ the measure on X which is absolutely continuous with respect to $\unicode{x3bb} _X$ with density

$$ \begin{align*} \frac {d\mu _a^{(N)}}{d\unicode{x3bb} _X}(x) = \prod_{n=0}^N(1 +{\mathsf{Re\, }} a_n\gamma _n(x)). \end{align*} $$

It is not very difficult to show that the sequence of measures $(\mu _{a}^{(N)})_{N\ge 0}$ converges weakly; the limiting measure $\mu _a=\lim _N \mu _a^{(N)}$ is called the Riesz product, and we denote it as

(4.1) $$ \begin{align} \mu _a=\prod_{n=0}^{\infty} (1 +{\mathsf{Re\, }} a_n\gamma _n(x)). \end{align} $$

The Riesz product $\mu _a$ is absolutely continuous with respect to the Haar measure $\unicode{x3bb} _X$ if and only if $\sum _{n} |a_n|^2<\infty $ , and it is singular to the Haar measure $\unicode{x3bb} _X$ if and only if ${\sum _{n} |a_n|^2=\infty }$ (see [Reference Peyrière25, Reference Zygmund34]). We will omit dependence of $\mu _a$ on the sequence $\Lambda $ , since $\Lambda $ will usually be fixed.

Since

$$ \begin{align*} 1 +{\mathsf{Re\, }} a_n\gamma _n(x) =1 + \frac {a_n}2\gamma _n(x)+ \frac {\overline a_n}2\gamma _n^{-1}(x), \end{align*} $$

the Riesz product $\mu _a$ , associated to the sequences $\Lambda $ and a, can be characterized by the Fourier coefficients $ \widehat {\mu }_a(\gamma )=\int \overline {\gamma }(x)\,d\mu _a(x)$ , $\gamma \in \widehat {X}$ , as follows.

  1. (a) For any finite set of distinct characters $\{\gamma _{n_1},\gamma _{n_2},\ldots , \gamma _{n_k}\}\subset \Lambda $ and any $(\varepsilon _1, \varepsilon _2, \ldots ,\varepsilon _k)\in \{-1,0,1\}^k$ ,

    (4.2) $$ \begin{align} {\widehat\mu }_a(\gamma _{n_1}^{\varepsilon _1} \gamma _{n_2}^{\varepsilon _2}\cdots \gamma _{n_k}^{\varepsilon _k}) = a_{n_1}^{(\varepsilon _1)} a_{n_2}^{(\varepsilon _2)} \cdots a_{n_k}^{(\varepsilon _k)}, \end{align} $$
    where $a_n^{(\varepsilon )}= ({a_n}/{2}), 0,$ or ${\overline a_n}/{2}$ , depending on whether $\varepsilon =1,0$ , or $-1$ .
  2. (b) For any character $\gamma \in \widehat {X}$ not of the form $\gamma _{n_1}^{\varepsilon _1}\gamma _{n_2}^{\varepsilon _2}\cdots \gamma _{n_k}^{\varepsilon _k}$ with $\varepsilon _1,\varepsilon _2,\ldots ,\varepsilon _k\in \{-1,0,1\}$ as in case (a) above, one has

    (4.3) $$ \begin{align} \widehat{\mu }_a(\gamma )=0. \end{align} $$

For any two Riesz products $\mu _a$ and $\mu _b$ , it is proved in [Reference Peyrière25] that $\mu _a$ and $\mu _b$ are mutually singular if $\sum |a_n-b_n|^2=\infty $ , and mutually equivalent if $\sum |a_n-b_n|^2<\infty $ and $\sup _n|a_n|<1$ . For any Riesz product $\mu _a$ , it is proved in [Reference Fan6] that the orthogonal series $\sum c_n (\gamma _n(x)- a_n/2)$ (with $c_n\in \mathbb {C}$ ) converges $\mu _a$ -a.e. if and only if $\sum |c_n|^2<\infty $ . Such convergence results will be useful to us in the proofs of Theorems 2.1 and 2.3. Riesz products on $\mathbb {T}$ and some generalized Riesz products appear as spectral measures of some dynamical systems (see [Reference Bourgain1, Reference Ledrappier18, Reference Queffélec26]). Riesz products are tools in harmonic analysis (see [Reference Kahane14, Reference Katznelson16, Reference Zygmund34]).

4.2 The dual group $\widehat {X}_f$

Before constructing Riesz products on $X_f$ , let us describe the dual group of $X_f$ (cf. [Reference Lind and Schmidt19, Reference Schmidt29]). Every Laurent polynomial with integer coefficients

$$ \begin{align*} h({\boldsymbol z}) =\sum_{{\boldsymbol m} \in \mathbb{Z}^d} h_{\boldsymbol m} {\boldsymbol z}^{\boldsymbol m} \in R_d \end{align*} $$

defines a character $\gamma ^{(h)} \in \widehat {\mathbb {T}^{\mathbb {Z}^d}}$ , given by

where

$$ \begin{align*} \langle h,x\rangle=\sum_{{\boldsymbol m}\in{\mathbb Z}} h_{\boldsymbol m} x_{\boldsymbol m} \end{align*} $$

for every $x\in \mathbb {T}^{\mathbb {Z}^d}$ . Conversely, every character of $\mathbb {T}^{\mathbb {Z}^d}$ is of the form $\gamma =\gamma ^{(h)}$ for some $h\in R_d$ , so that we may identify $\widehat {\mathbb {T}^{\mathbb {Z}^d}}$ with $R_d$ . Note, however, that the group operation in $R_d$ is addition, whereas in $\widehat {\mathbb {T}^{\mathbb {Z}^d}}$ , it is multiplication:

$$ \begin{align*} \gamma ^{(h+h')} = \gamma ^{(h)}\gamma ^{(h')} \end{align*} $$

for all $h,h'\in R_d$ .

Since $X_f$ is a subgroup of $\mathbb {T}^{\mathbb {Z}^d}$ , every character $\gamma ^{(h)}\in \widehat {\mathbb {T}^{\mathbb {Z}^d}}$ , $h\in R_d$ , restricts to a character $\tilde {\gamma }^{(h)}\in \widehat {X}_f$ . From the definition of $X_f$ in equation (2.2), it is clear that, for any two polynomials $h,h'\in R_d$ , $\tilde {\gamma }^{(h)} = \tilde {\gamma }^{(h')}$ if and only if $h-h'$ is a multiple of f. This allows us to identify the dual group $\widehat {X}_f$ with $R_d/(f)$ , where $(f)=R_d\cdot f$ is the principal ideal in $R_d$ generated by f:

$$ \begin{align*} \widehat{X}_f = R_d/(f). \end{align*} $$

More generally, if $I\subset R_d$ is an ideal and $X_I$ is given by equation (2.5), then

$$ \begin{align*} \widehat{X}_I= R_d/I. \end{align*} $$

4.3 Lacunary polynomials

For the construction of Riesz product measures on $X_f$ , we have to take a closer look at dissociate families $\Lambda \subset \widehat {X}_f$ in the sense of Definition 4.1.

Definition 4.2. Given an integer $m\in \mathbb N$ , we say that a primitive irreducible polynomial $f\in R_d$ is m-good if the following conditions hold.

  1. (C1) The collection of characters

    $$ \begin{align*} \{ \tilde{\gamma }^{(\boldsymbol{z}^{m{\boldsymbol n}})} \mid {\boldsymbol n}\in\mathbb{N}^d\}\subset \widehat{X}_f \end{align*} $$
    is dissociate. Explicitly, this means that any non-zero polynomial of the form $g({\boldsymbol z}^m)$ where
    $$ \begin{align*} g({\boldsymbol z})=\sum_{{\boldsymbol n} \in \mathbb{Z}^d} \varepsilon _{\boldsymbol n} {\boldsymbol z}^{{\boldsymbol n}} \end{align*} $$
    with $\varepsilon _{\boldsymbol n} \in \{-2,-1,0,1,2\}$ is not divisible by f.
  2. (C2) For any ${\boldsymbol k} \in \mathbb {N}^d/ m\mathbb {N}^d$ , any two points ${\boldsymbol n} \ne {\boldsymbol n}'$ in $\mathbb {Z}^d$ , and any non-zero polynomial of the form with $\varepsilon _{\boldsymbol n} \in \{-1,0,1\}$ , the polynomial

    $$ \begin{align*} {\boldsymbol z}^{m{\boldsymbol n}+{\boldsymbol k}}-{\boldsymbol z}^{m{\boldsymbol n}'+{\boldsymbol k}} + g({\boldsymbol z}^{m}) \end{align*} $$
    is not divisible by f.

For a given principal algebraic action $(X_f,\alpha _f)$ , where f is m-good, Riesz product measures $\mu _a$ can be constructed using the countable dissociate collection of characters $\Lambda =\{ \tilde {\gamma }^{({\boldsymbol z}^{m{\boldsymbol n}})}\mid \ {\boldsymbol n}\in \mathbb {N}^d\}$ (cf. condition (C1)). Condition (C2) ensures that any shifted family of characters $\Lambda _{{\boldsymbol k}}=\{ \tilde {\gamma }^{({\boldsymbol z}^{m{\boldsymbol n}+{\boldsymbol k}})}\mid \ {\boldsymbol n}\in \mathbb {N}^d \}$ (with ${\boldsymbol k}\in [0,m-1]^d\smallsetminus \{\boldsymbol {0}\}$ being fixed) is a $\mu _a$ -orthogonal system, as a direct consequence of equation (4.2) applied with $k=2$ —a useful property which will help us control the behavior of weighted ergodic averages. As we will see, the coefficient sequence a will be chosen depending on the non-trivial weight sequence $\boldsymbol w$ .

5 $(X_f, \alpha _f)$ is Bohr chaotic when f is m-good

The following theorem will allow us to reduce the proof of Bohr chaoticity of $(X_f, \alpha _f)$ to checking the m-goodness of the polynomial f.

Theorem 5.1. If a primitive irreducible polynomial $f\in R_d$ with positive Mahler measure is m-good, that is, if the conditions (C1) and (C2) hold for some positive integer m, then $(X_f,\alpha _f)$ is Bohr chaotic.

We begin with a simple auxiliary lemma.

Lemma 5.2. Let $\alpha $ be a continuous ${\mathbb Z}^d$ -action on a compact metrizable space X, and let ${\boldsymbol w} = (w_{\boldsymbol n})_{{\boldsymbol n} \in \mathbb {N}^d}$ be a non-trivial weight. Then $(X,\alpha )$ is not disjoint from ${\boldsymbol w}=(w_{\boldsymbol n})$ if and only if for any ${\boldsymbol k}\in \mathbb {N}^d$ , $(X,\alpha )$ is not disjoint from the weight $\widetilde {{\boldsymbol w}}=(\widetilde {w}_{\boldsymbol n})$ defined by $\widetilde {w}_{\boldsymbol n}=w_{{\boldsymbol n}+{\boldsymbol k}}$ for all ${\boldsymbol n}\in \mathbb {N}^d$ .

Proof. Introduce the following notation: for a continuous function $\phi $ on X, let

$$ \begin{align*} S_N^{{\boldsymbol w}}\phi(x)=\sum_{{\boldsymbol n} \in [0, N-1]^d} w_{\boldsymbol n} \phi(\alpha ^{\boldsymbol n} x). \end{align*} $$

For any ${\boldsymbol k}\in \mathbb {N}^d$ and for any $x\in X$ , one has

$$ \begin{align*}\begin{aligned} |S_{N+\|{\boldsymbol k}\|_\infty}^{{\boldsymbol w}}\phi(x)-S^{\widetilde{{\boldsymbol w}}}_N\phi(\alpha ^{{\boldsymbol k}}x)|&=\bigg|\!\sum_{{\boldsymbol n} \in [0, N+\|{\boldsymbol k}\|_\infty-1]^d}\! w_{{\boldsymbol n}}\phi(\alpha ^{{\boldsymbol n}}x)-\!\!\sum_{{\boldsymbol n} \in [0, N-1]^d}\! {w}_{{\boldsymbol n}+{\boldsymbol k}}\phi(\alpha ^{{\boldsymbol n}+{\boldsymbol k}}x)\bigg|\\ &\le \|\phi\|_{\infty}\cdot | [0,N+\|{\boldsymbol k}\|_\infty-1]^d\triangle ({\boldsymbol k}+[0,N-1]^d)|. \end{aligned} \end{align*} $$

To finish the proof, it suffices to notice that the cardinality of the symmetric difference is of order $O(N^{d-1})$ .

Proof of Theorem 5.1

Fix $m\in {\mathbb N}$ such that conditions (C1) and (C2) hold. Assume that ${\boldsymbol w}$ is a non-trivial weight (cf. equation (1.5)). Then for some ${\boldsymbol k}\in [0,\ldots ,m-1]^{d}$ , one has

(5.1) $$ \begin{align} \limsup_{N\to \infty} \frac{1}{N^d} \sum_{{{\boldsymbol n}\mid m{\boldsymbol n}+{\boldsymbol k}} \in [0, N-1]^d} |w_{m{\boldsymbol n}+{\boldsymbol k}}|>0. \end{align} $$

Without loss of generality, we can assume ${\boldsymbol k}=\boldsymbol 0$ . Otherwise, consider the shifted weight $\widetilde {{\boldsymbol w}}=(\widetilde {w}_{\boldsymbol n})$ with $\widetilde {w}_{\boldsymbol n}=w_{{\boldsymbol n}+{\boldsymbol k}}$ . By Lemma 5.2, $(X_f,\alpha _f)$ is not disjoint from ${\boldsymbol w}$ if and only if $(X_f,\alpha _f)$ is not disjoint from $\widetilde {\boldsymbol w}$ . Thus, it is sufficient to consider the weight $\widetilde {{\boldsymbol w}}$ for which we can assume that equation (5.1) holds with ${\boldsymbol k}=0$ . In the following, we consider an arbitrary such weight.

Step 1. Choice of the function $\phi $ and the point x. We are going to show that equation (1.5) holds for and for almost all $x\in X_f$ with respect to an appropriately chosen Riesz product measure. Note that for all ${\boldsymbol n}\in \mathbb {N}^d$ ,

$$ \begin{align*} \phi(\alpha_f ^{{\boldsymbol n}}x) = e^{2\pi i x_{{\boldsymbol n}}}= e^{2\pi i \langle {\boldsymbol z}^{{\boldsymbol n}}, x\rangle}=\gamma ^{({\boldsymbol z}^{{\boldsymbol n}})} (x). \end{align*} $$

Step 2. Choice of the measure. By condition (C1), the collection of characters

is dissociate. Consider now the following collection of coefficients:

Since $|a_{{\boldsymbol n}}|=1$ for all ${\boldsymbol n}$ , the Riesz product $\mu _a$ in equation (4.1) is well defined.

Step 3. Orthonormality. For each ${\boldsymbol k}\in [0,m-1]^d\smallsetminus \{\boldsymbol 0\}$ , consider the following collection of functions:

We claim that for each ${\boldsymbol k}\in [0,m-1]^d\smallsetminus \{\boldsymbol 0\}$ , $\mathcal F_{{\boldsymbol k}}$ is orthonormal in $L^2(X_f,\mu _a)$ . Indeed, for each ${\boldsymbol n}\ne {\boldsymbol n}'$ , condition (C2) means that the character corresponding to the polynomial ${\boldsymbol z}^{m{\boldsymbol n}+{\boldsymbol k}}-{\boldsymbol z}^{m{\boldsymbol n}'+{\boldsymbol k}}$ :

$$ \begin{align*} \gamma ^{({\boldsymbol z}^{m{\boldsymbol n}+{\boldsymbol k}}-{\boldsymbol z}^{m{\boldsymbol n}'+{\boldsymbol k}})}(x)=\gamma ^{({\boldsymbol z}^{m{\boldsymbol n}+{\boldsymbol k}})}(x) \overline{\gamma ^{({\boldsymbol z}^{m{\boldsymbol n}'+{\boldsymbol k}})}(x)}, \end{align*} $$

cannot be expressed as a product of characters in $\Lambda $ , and hence using equation (4.3) for the Fourier coefficients of Riesz products, one gets that

$$ \begin{align*} \int_{X_f} \gamma ^{({\boldsymbol z}^{m{\boldsymbol n}+{\boldsymbol k}})}(x) \overline{\gamma ^{({\boldsymbol z}^{m{\boldsymbol n}'+{\boldsymbol k}})}(x)}\,d\mu _a(x) = \widehat{\mu _a}(\overline{\gamma ^{({\boldsymbol z}^{m{\boldsymbol n}+{\boldsymbol k}}-{\boldsymbol z}^{m{\boldsymbol n}'+{\boldsymbol k}})}})=0. \end{align*} $$

Since $|\gamma ^{({\boldsymbol z}^{m{\boldsymbol n}+{\boldsymbol k}})}(x)|^2=1$ for all x, the orthonormality of $\mathcal F_{{\boldsymbol k}}$ is thus proved.

For ${\boldsymbol k}=0$ , we set

Direct application of equations (4.2) and (4.3) immediately gives that the collection of functions $\mathcal F_{\boldsymbol 0}$ is orthogonal in $L^2(X_f,\mu _a)$ , and that

$$ \begin{align*} \int_{X_f} |\gamma ^{({\boldsymbol z}^{m{\boldsymbol n}})}(x)|^2\,d\mu _a(x)=1-\frac{|a_{{\boldsymbol n}}|^2}{4}=\frac {3}{4}\quad \text{for all } {\boldsymbol n}\in\mathbb{N}^d. \end{align*} $$

Step 4. Almost everywhere convergence. Write

$$ \begin{align*} S_N^{{\boldsymbol w}}\phi(x)= \sum_{{\boldsymbol n}\in [0,N-1]^d} w_{\boldsymbol n} \phi(\alpha _f^{\boldsymbol n} x) = \sum_{{\boldsymbol k}\in [0,m-1]^d}S_{N,{\boldsymbol k}}^{{\boldsymbol w}}\phi(x), \end{align*} $$

where

We claim that for any ${\boldsymbol k}\in [0,m-1]^d\smallsetminus \{\boldsymbol 0\}$ , one has

(5.2) $$ \begin{align} \frac 1{N^d} S_{N,{\boldsymbol k}}^{{\boldsymbol w}}\phi(x)\to 0\quad\mu _a\!-\!a.e., \end{align} $$

and for ${\boldsymbol k}=0$ , one has

(5.3) $$ \begin{align} \frac 1{N^d}\bigg( S_{N,{\boldsymbol 0}}^{{\boldsymbol w}}\phi(x)-\frac 12\sum_{\{{\boldsymbol n}\mid m{\boldsymbol n}\in[0,N-1]^d\}}|w_{\boldsymbol n}|\bigg)\to0 \quad\mu _a\!-\!a.e. \end{align} $$

Now we write

$$ \begin{align*} \frac {1}{N^d}S_N^{{\boldsymbol w}}\phi(x)=\frac {1}{N^d}\bigg(S_N^{{\boldsymbol w}}\phi(x) -\frac 12\sum_{\{{\boldsymbol n}\mid m{\boldsymbol n}\in[0,N-1]^d\}}|w_{\boldsymbol n}|\bigg)+\frac {1}{2N^d}\sum_{\{{\boldsymbol n}\mid m{\boldsymbol n}\in[0,N-1]^d\}}|w_{\boldsymbol n}|. \end{align*} $$

If equations (5.2) and (5.3) are indeed true, the first term in the brackets on the right-hand side converges to $0$ for $\mu _a$ -almost all $x\in X_f$ , and the second term does not converge to $0$ by equation (5.1). Hence, we will be able to conclude that

$$ \begin{align*} \limsup_{N\to\infty} \frac 1{N^d} | S_N^{{\boldsymbol w}}\phi(x)|>0, \quad \mu _a\mbox{-a.e.}, \end{align*} $$

and thus, that $(X_f,\alpha _f)$ is Bohr chaotic.

Finally, to establish equations (5.2) and (5.3), we will use the following multivariate generalization of the result of Davenport, Erdös, and LeVeque [Reference Davenport, Erdős and LeVeque3] due to Fan, Fan, and Qiu [Reference Fan, Fan and Qiu8, Theorem 6.1]: suppose that $\{\xi _{\boldsymbol \ell }\mid {\boldsymbol \ell }\in {\mathbb N}^d\}$ is a collection of measurable complex valued uniformly bounded functions on a probability space $(\Omega ,\mathbb {P})$ such that

(5.4) $$ \begin{align} \sum_{N=1}^{\infty} \frac 1{N}\int_{\Omega } |Z_{N}|^2\, d\mathbb{P}<\infty, \end{align} $$

where

$$ \begin{align*} Z_N=\frac {1}{N^d}\sum_{{\boldsymbol \ell}\in[0,N-1]^d} \xi_{{\boldsymbol \ell}} \quad (N\ge 1). \end{align*} $$

Then $Z_N\to 0$ as $N\to \infty \ \mathbb {P}$ -a.e. on $\Omega $ .

In particular, if $\{\xi _{\boldsymbol \ell }\mid {\boldsymbol \ell }\in {\mathbb N}^d\}$ are uniformly bounded and orthogonal in $L^2(\Omega ,\mathbb {P})$ , then

$$ \begin{align*} \frac 1{N}\int_{\Omega } |Z_{N}|^2\,d\mathbb{P} =\frac 1{N^{2d+1}}\sum_{{\boldsymbol \ell}\in[0,N-1]^d} \int_{X} |\xi_{{\boldsymbol \ell}}|^2\,d\mathbb{P} \le \frac{C}{N^{d+1}}, \end{align*} $$

and hence equation (5.4) holds for any $d\ge 1$ .

If we now apply this result to the orthogonal families of bounded functions

$$ \begin{align*} \mathcal F_{{\boldsymbol k}}^{\boldsymbol w}=\{ w_{m{\boldsymbol n}+{\boldsymbol k}}\psi\circ\alpha ^{m{\boldsymbol n}+{\boldsymbol k}}(x)\mid {\boldsymbol n}\in{\mathbb Z}^d_+ \},\quad {\boldsymbol k}\in[0,m-1]^d\smallsetminus\{\boldsymbol 0\}, \end{align*} $$

and

$$ \begin{align*} \mathcal F_{\boldsymbol 0}^{\boldsymbol w}=\bigg\{ w_{m{\boldsymbol n}}\bigg(\psi\circ\alpha ^{m{\boldsymbol n}}(x)-\frac {a_{{\boldsymbol n}}}{2}\bigg) \mid {\boldsymbol n}\in{\mathbb N}^d \bigg\}, \end{align*} $$

we obtain equations (5.2) and (5.3), and hence, we complete the proof.

6 Bohr chaoticity of $(X_f, \alpha _f)$ : the case of $d=1$

In this section, we complete the proof of Theorem 2.1 in the case where $d=1$ : every principal algebraic ${\mathbb Z}$ -action $(X_f,\alpha _f)$ with positive entropy is Bohr chaotic. Theorem 2.1 will follow from Remark 2.7, Theorem 5.1, and the following result.

Theorem 6.1. Every primitive irreducible polynomial $f\in R_1$ with $\mathsf {m}(f)>0$ is m-good for some positive integer m.

The proof of Theorem 6.1 consists of the following three lemmas.

Lemma 6.2. (Preliminary lemma)

Let $f(z) = f_0 + f_1 z + \cdots + f_r z^r \in \mathbb {Z}[z]$ be an irreducible polynomial with $r\ge 1$ and $f_0f_r\ne 0$ . If $\mathsf {m}(f)>0$ , then at least one of the following statements is true:

  1. (1) there exists a root of f in $\mathbb {C}$ which is not on the unit circle;

  2. (2) there exists a prime $p\ge 2$ such that f admits a root $\unicode{x3bb} $ in the algebraic closure $\overline {\mathbb {Q}}_p$ of the field of p-adic numbers $\mathbb {Q}_p$ such that $|\unicode{x3bb} |_p>1$ .

Proof. Suppose $r=1$ , that is, $f(z)=f_0+f_1z$ . If the only root of f lies on the unit circle, then necessarily $|f_0|=|f_1|$ . The irreducibility of f implies that $|f_0|=|f_1|=1$ , and thus $\mathsf {m}(f)=0$ by equation (2.7). Therefore, assuming $\mathsf {m}(f)>0$ for irreducible f with $\deg (f)=1$ , we conclude that $|f_0|\ne |f_1|$ , and hence condition (1) must hold.

Suppose $r\ge 2$ and the roots of f are all on the unit circle, that is, suppose that condition (1) does not hold. If $|f_r|=1$ , then Kronecker’s theorem [Reference Kronecker17] implies that all roots of f are roots of unity, and equation (2.7) shows that $\mathsf {m}(f)=0$ , in contradiction to our hypothesis. However, if $|f_r|>1$ , then

$$ \begin{align*} f(z) =f_r \bigg( z^r +\frac{f_{r-1}}{f_r} z^{r-1}+\cdots +\frac{f_0}{f_r}\bigg), \end{align*} $$

and Vieta’s formula implies that $|{f_0}/{f_r}| = |\prod _{\zeta \in \mathbb {C}: f(\zeta )=0}\zeta | =1$ , that is, that $|f_0|=|f_r|$ . Since f is irreducible, $f_j/f_r$ is not integer for some $1\le j\le r-1$ . Then there exists a rational prime p such that $|f_j/f_r|_p>1$ . Let $\unicode{x3bb} _i,\; 1\le i \le r$ , be the roots of f in the algebraic closure $\overline {\mathbb {Q}}_p$ of $\mathbb {Q}_p$ , the p-adic rationals. By considering the jth elementary symmetric polynomials of the roots and once again applying Vieta’s formulas, we get that

$$ \begin{align*} 1<\bigg|\frac{f_j}{f_r}\bigg|_p = \bigg|\!\sum_{1\le i_1<i_2<\cdots <i_j \le r}\unicode{x3bb} _{i_1}\cdots \unicode{x3bb} _{i_j}\bigg|_p\le (\max_{1\le i\le r} |\unicode{x3bb} _i|_p)^j. \end{align*} $$

Thus one has $|\unicode{x3bb} _i|_p>1$ for some $i\in \{1,\ldots ,r\}$ .

The following key lemma will be used to show that for sufficiently large m, the sequence of polynomials $\{z^{nm}\}_{n\ge 0}$ gives a dissociate sequence of characters of $X_f$ .

Lemma 6.3. (Condition (C1))

Suppose that $f=f_0+f_1z+\cdots +f_rz^r\in {\mathbb Z}[z]$ has a root in $\mathbb {C}$ or in $\overline {\mathbb {Q}}_p$ (for some p) of absolute value larger than $1$ . Then for any sufficiently large m and any $D\ge 0$ , the polynomials

(6.1) $$ \begin{align} P(z) =\sum_{j=0}^D \varepsilon _j z^{mj} \quad \text{with }\varepsilon _0,\varepsilon _1,\ldots,\varepsilon _D\in \{-2,-1,0,1,2\}, \end{align} $$

are not divisible by f unless $\epsilon _0=\epsilon _1=\cdots = \epsilon _D=0$ .

Proof. First we consider the case that f has a root in $\mathbb {C}$ of modulus larger than $1$ . For any polynomial g, we introduce the notation

$$ \begin{align*} \rho_g=\max\{|z|\mid g(z)=0\}. \end{align*} $$

Without loss of generality, we may assume that $\varepsilon _D\ne 0$ and consider the reduced polynomial

$$ \begin{align*} \widetilde{P}(z) =\sum_{j=0}^D \varepsilon _j z^{j}, \end{align*} $$

such that $ P(z)= \widetilde {P}(z^m)$ . Clearly, $\rho _{\widetilde {P}} = \rho _P^m$ .

However, using the Cauchy bound on the roots of polynomials, one gets that

$$ \begin{align*} \rho_{\widetilde{P}}\le 1+\max_{j=0,\ldots,D-1} \bigg|\frac {\varepsilon _j}{\varepsilon _D}\bigg|\le 3, \end{align*} $$

and hence $ \rho _P \le 3^{{1}/{m}}. $ Choose an integer $M\ge 1$ large enough such that $3^{{1}/{M}} <\rho _f$ (this is possible because $\rho _f>1$ ). Thus for all $m\ge M$ , we have $\rho _P <\rho _f$ . However, if $P(z)$ was divisible by f, we would have $\rho _f \le \rho _P$ , thus arriving to a contradiction.

If f has a root in $\overline {\mathbb {Q}}_p$ (for some prime p) of absolute value larger than $1$ , the same argument works with $|\cdot |$ replaced by $|\cdot |_p$ . Indeed, suppose $\zeta $ is a root of f with $|\zeta |_p>1$ . If $f|P$ , then $\zeta $ is also a root of P, so one has

$$ \begin{align*} |\zeta|_{p}^{mD}\!=\!|\zeta^{mD}|_p\!=\!\bigg|\! \sum_{j=0}^{D-1} \frac{\varepsilon _j}{\varepsilon _D} \zeta^{mj}\bigg|_p \le \max_{j=0,\ldots, D-1}\bigg| \frac{\varepsilon _j}{\varepsilon _D} \zeta^{mj}\bigg|_p \le \bigg(\max_{j=0,\ldots, D-1} \bigg| \frac{\varepsilon _j}{\varepsilon _D}\bigg|_p\bigg) |\zeta|_{p}^{m(D-1)}. \end{align*} $$

Thus arriving at a contradiction.

Lemma 6.4. (Condition (C2))

Suppose that $f(z)= f_rz^r+\cdots + f_1z+f_0 \in \mathbb {Z}[z]$ has a root in $\mathbb {C}$ or in $\overline {\mathbb {Q}}_p$ (for some prime p) of absolute value larger than $1$ . Then for all sufficiently large integers m, any integer k with $1\le k <m$ , every $D\ge 0$ , and all arbitrary $(D+1)$ -tuples $\varepsilon =(\varepsilon _0,\ldots ,\varepsilon _D)$ and $\delta =(\delta _0,\ldots ,\delta _D)$ in $\{-1,0,1\}^{D+1}$ , the polynomial

(6.2) $$ \begin{align} Q(z) = \sum_{j=0}^D \varepsilon _j z^{mj}-\sum_{j=0}^D \delta_j z^{mj+k} \end{align} $$

is not divisible by f unless $Q(z)\equiv 0$ , that is, unless all $\epsilon _j$ and $\delta _j$ are equal to zero.

Proof. Assume Q is divisible by f. We treat the complex case first. Namely, assume $\zeta \in \mathbb {C}$ is such that $f(\zeta )=0$ and . Without loss of generality, we may assume that $|\varepsilon _D|+|\delta _D|>0$ . We distinguish two cases.

Case I. $\delta _D\ne 0$ . If the polynomial $Q(z)$ , defined by equation (6.2), is divisible by f, then $Q(\zeta )=0$ , in other words,

(6.3) $$ \begin{align} \smash[b]{\delta_D \zeta^{mD+k} =\sum_{j=0}^D \varepsilon _j \zeta^{mj}-\sum_{j=0}^{D-1} \delta_j \zeta^{mj+k}.} \end{align} $$

It follows that

$$ \begin{align*} \smash{R^{mD+k} \le \sum_{j=0}^D R^{mj}+\sum_{j=0}^{D-1} R^{mj+k}= \frac{R^{(D+1)m}-1}{R^m-1}+ \frac {R^{Dm}-1}{R^m-1}\cdot R^k,} \end{align*} $$

and hence

$$ \begin{align*} \smash[t]{R^{k} < \frac{R^{m} }{R^m-1}+ \frac {R^k}{R^m-1}}. \end{align*} $$

As $m\to \infty $ , the right-hand side of this inequality converges to $1$ , but the left-hand side remains equal to $R^k>1$ . If m is large enough, our assumption that Q is divisible by f leads to a contradiction.

Case II. $\delta _D=0$ but $\varepsilon _D\ne 0$ . In this case, we have

(6.4) $$ \begin{align} \varepsilon _D \zeta^{mD} =-\sum_{j=0}^{D-1}\varepsilon _j \zeta^{jm}+\sum_{j=0}^{D-1}\delta_j \zeta^{jm+k}. \end{align} $$

It follows that

$$ \begin{align*} 1 < \frac{1}{R^m -1} +\frac{R^k}{R^m -1}=\frac {R^k+1}{R^m-1}\le \frac{R^{m-1}+1}{R^m-1}. \end{align*} $$

Since $R>1$ , the last inequality is violated for all sufficiently large m, and we again arrive at a contradiction with our assumption that Q is divisible by f.

In the p-adic case, the argument is simpler because of the non-archimedean triangle inequality $|\zeta +\xi |_p\le \max (|\zeta |_p,|\xi |_p)$ . Indeed, from equation (6.3), we get that $|\zeta |_p^{mD+k}\le |\zeta |_p^{mD}$ (impossible), and from equation (6.4), we get that $|\zeta |_p^{mD}\le |\zeta |_p^{m(D-1)+k}$ (equally impossible).

Proof of Theorem 2.1

By Remark 2.7, we may assume that the polynomial $f(z) = f_0+\cdots +f_rz^r\in \mathbb {Z}[z]$ is primitive, irreducible, and has positive Mahler measure (cf. equation (2.7)). If f has a root in $\mathbb {C}$ of absolute value larger than 1, Lemmas 6.3 and 6.4 show that f is m-good for some $m\in {\mathbb N}$ (in fact, for all sufficiently large m). An application of Theorem 5.1 completes the proof of Theorem 2.1 in this case.

If all complex roots of f have absolute value $\le 1$ , and if $\zeta $ is a root of f, we denote by $\textbf {k}=\mathbb {Q}(\zeta )$ the algebraic number field generated by $\zeta $ and write $P_{\textbf {k}}$ for the set of places (or valuations) of $\textbf {k}$ . If $|\zeta |_v$ is the absolute value of $\zeta $ at a place $v\in P_{\textbf {k}}$ , then our last assumption implies that $|\zeta |_v\le 1$ for every infinite place $v\in P_{\textbf {k}}$ . Since the product formula for algebraic number fields shows that $\prod _{v\in P_{\textbf {k}}}|\zeta |_v=1$ , we either have that $|\zeta |_v=1$ for every $v\in P_{\textbf {k}}$ , or that there exists a finite place $v\in P_{\textbf {k}}$ with $|\zeta |_v>1$ . In the first case, $\zeta $ is a root of unity, which makes f cyclotomic and $\mathsf {m}(f)=0$ , contrary to our hypothesis. In the second case, we can once again use the Lemmas 6.3 and 6.4 to show that f is m-good for all sufficiently large m, and Theorem 5.1 completes the proof of Theorem 2.1 as above.

7 Bohr chaoticity of $(X_f, \alpha _f)$ : the case of $d\ge 2$

This section is devoted to the proof of Theorem 2.3 for $d\ge 2$ , which will again be based on Theorem 5.1. As before, we assume that the polynomial $f\in R_d$ is primitive and irreducible.

7.1 Homoclinic points of atoral polynomials in $R_d$ and the gap theorem

For every ${t\in \mathbb {T}}$ , we set

$$ \begin{align*} \lVert t\rVert = \min_{q\in \mathbb{Z}}|t-q|. \end{align*} $$

Definition 7.1. A point $x\in X_f$ is homoclinic (or, more precisely, homoclinic to $0$ ) if $\lim _{{\boldsymbol n}\to \infty }\lVert x_{\boldsymbol n}\rVert =0$ . A homoclinic point $x\in X_f$ is summable if $\sum _{{\boldsymbol n}\in \mathbb {Z}^d}\lVert x_{\boldsymbol n}\rVert <\infty $ .

The existence of non-zero summable homoclinic points of $(X_f,\alpha _f)$ is equivalent to atorality of the polynomial f.

Theorem 7.2. [Reference Lind, Schmidt and Verbitskiy20]

If $0\ne f\in R_d$ , the following conditions are equivalent:

  1. (1) the principal algebraic action $(X_f,\alpha _f)$ has a non-zero summable homoclinic point;

  2. (2) the Laurent polynomial f is atoral in the sense of Definition 2.2.

For a principal algebraic $\mathbb {Z}^d$ -action $(X_f,\alpha _f)$ , the existence of summable homoclinic points has a number of important consequences (cf. [Reference Lind and Schmidt19]): it implies positivity of entropy and very strong specification properties of the action, and it guarantees the coincidence of entropy with the logarithmic growth rate of the number of periodic points of $\alpha _f$ (that is, of points in $X_f$ with finite orbits under $\alpha _f$ —cf. [Reference Lind and Schmidt19, Reference Lind, Schmidt and Verbitskiy20]). Somewhat surprisingly, it also plays a role in the gap theorem stated below, which will imply conditions (C1) and (C2) in Definition 4.2.

We remark in passing that some of these consequences of atorality also hold for toral polynomials, but with considerably harder proofs and/or weaker conclusions—cf. e.g., [Reference Dimitrov4] or [Reference Lind, Schmidt and Verbitskiy20]). However, it is not known if specification or gap properties hold in the toral case.

To state the gap theorem referred to above, we consider, for any non-empty subset ${{\mathcal S}\subset {\mathbb Z}^d}$ and any integer $H\ge 1$ , the set $\mathcal P({\mathcal S},H)\subset R_d$ of all Laurent polynomials with support in ${\mathcal S}$ and coefficients bounded in absolute value by H:

$$ \begin{align*} \mathcal P({\mathcal S},H)=\{ v\in R_d\mid \textsf{supp}(v)\subseteq {\mathcal S}\quad\text{and}\quad \|v\|_\infty \le H\}. \end{align*} $$

For every ${\boldsymbol n}=(n_1,\ldots ,n_d)\in \mathbb {Z}^d$ , we set $\|{\boldsymbol n}\|=\max \{|n_1|,\ldots ,|n_d|\}$ . Then the following is true.

Theorem 7.3. (Gap theorem)

Suppose that $g\in R_d$ is primitive, irreducible, and atoral. For every $H\ge 1$ , there exists an integer $m\ge 1$ with the following property: for every pair of sets ${\mathcal S},{\mathcal S}'\subset \mathbb {Z}^d$ with distance

and for every $v=\sum _{{\boldsymbol n}\in {\mathcal S}\cup {\mathcal S}'} v_{{\boldsymbol n}}{\boldsymbol z}^{{\boldsymbol n}} \in \mathcal {P}({\mathcal S}\cup {\mathcal S}',H)$ which is divisible by g, the restriction of v to ${\mathcal S}$ ,

(7.1) $$ \begin{align} v_{\mathcal S}=\sum_{{\boldsymbol n}\in {\mathcal S}} v_{{\boldsymbol n}}{\boldsymbol z}^{{\boldsymbol n}}, \end{align} $$

is also divisible by g.

For the proof of Theorem 7.3, we consider the algebra $\ell ^1(\mathbb {Z}^d,\mathbb {R})$ of all functions $v\colon {\boldsymbol n} \mapsto v_{\boldsymbol n}$ from $\mathbb {Z}^d$ to $\mathbb {R}$ with $\|v\|_1=\sum _{{\boldsymbol n}\in \mathbb {Z}^d}|v_{\boldsymbol n}|<\infty $ , furnished with its usual multiplication (or convolution) $(v,w)\mapsto v\cdot w$ and involution $w\mapsto w^{*}$ , given by

(7.2) $$ \begin{align} \smash[b]{(v\cdot w)_{\boldsymbol n} = \sum_{{\boldsymbol m}\in \mathbb{Z}^d}v_{{\boldsymbol m}}w_{{\boldsymbol n} - {\boldsymbol m}} = \sum_{{\boldsymbol m}\in \mathbb{Z}^d}v_{{\boldsymbol n}-{\boldsymbol m}}w_{\boldsymbol m},} \end{align} $$

and

(7.3) $$ \begin{align} w^{*}_{\boldsymbol m} = w_{-{\boldsymbol m}} \end{align} $$

for every $v,w\in \ell ^1(\mathbb {Z}^d,\mathbb {R})$ and ${\boldsymbol m},{\boldsymbol n}\in \mathbb {Z}^d$ . If we denote by $\ell ^1(\mathbb {Z}^d,\mathbb {Z})\subset \ell ^1(\mathbb {Z}^d,\mathbb {R})$ the set of all integer-valued elements of $\ell ^1(\mathbb {Z}^d,\mathbb {R})$ and identify every $h = \sum _{{\boldsymbol n}\in \mathbb {Z}^d}h_{\boldsymbol n}\boldsymbol {z}^{\boldsymbol n}\in R_d$ with the element $(h_{\boldsymbol n})_{{\boldsymbol n}\in \mathbb {Z}^d}\in \ell ^1(\mathbb {Z}^d,\mathbb {Z})$ , we obtain an embedding

$$ \begin{align*} R_d = \ell ^1(\mathbb{Z}^d,\mathbb{Z}) \subset \ell ^1(\mathbb{Z}^d,\mathbb{R}) \end{align*} $$

in which the multiplication $(h,h')\mapsto h\cdot h'$ of Laurent polynomials extends to the composition in equation (7.2) in $\ell ^1(\mathbb {Z}^d,\mathbb {R})$ . In fact, the multiplication $(v,w)\mapsto v\cdot w$ in equation (7.2) is also well defined for $w\in \ell ^1 (\mathbb {Z}^d,\mathbb {R})$ and $v\in \ell ^{\infty } (\mathbb {Z}^d,\mathbb {R})$ , the space of all bounded sequences $(v_{\boldsymbol n})_{{\boldsymbol n}\in \mathbb {Z}^d}$ in the supremum norm $\|v\|_\infty =\sup _{{\boldsymbol n}\in \mathbb {Z}^d}|v_{\boldsymbol n}|$ , and

$$ \begin{align*} \|v\cdot w\|_\infty \le\|v\|_\infty \|w\|_1 \end{align*} $$

for all $w\in \ell ^1 (\mathbb {Z}^d,\mathbb {R})$ and $v\in \ell ^{\infty } (\mathbb {Z}^d,\mathbb {R})$ .

The shift action $\bar \sigma $ of $\mathbb {Z}^d$ on $\ell ^{\infty } (\mathbb {Z}^d,\mathbb {R})$ , defined exactly as in equation (2.1) by

(7.4) $$ \begin{align} (\bar\sigma ^{\boldsymbol m} v)_{\boldsymbol n} = v_{{\boldsymbol m}+{\boldsymbol n}} \end{align} $$

for every ${\boldsymbol m}\in \mathbb {Z}^d$ and $v\in \ell ^{\infty } (\mathbb {Z}^d,\mathbb {R})$ , extends to an action $w\mapsto w(\bar \sigma )$ of $\ell ^1(\mathbb {Z}^d,\mathbb {R})$ on $\ell ^{\infty } (\mathbb {Z}^d,\mathbb {R})$ by bounded linear operators with

$$ \begin{align*} w(\bar\sigma )= \sum_{{\boldsymbol m}\in \mathbb{Z}^d}w_{\boldsymbol m}\bar\sigma ^{\boldsymbol m}\colon \ell^{\infty} (\mathbb{Z}^d,\mathbb{R})\to \ell^{\infty} (\mathbb{Z}^d,\mathbb{R}) \end{align*} $$

for every $w\in \ell ^1(\mathbb {Z}^d,\mathbb {R})$ . Equation (7.4) implies that

$$ \begin{align*} (w(\bar\sigma )v)_{\boldsymbol n}=\sum_{{\boldsymbol m}\in \mathbb{Z}^d}w_{\textbf{m}}(\bar\sigma ^{\boldsymbol m} v)_{\boldsymbol n} = \sum_{{\boldsymbol m}\in \mathbb{Z}^d} w_{\boldsymbol m} v_{{\boldsymbol m}+{\boldsymbol n}} = (w^{*}\cdot v)_{\boldsymbol n}, \end{align*} $$

so that

(7.5) $$ \begin{align} w(\bar\sigma )v = w^{*}\cdot v \end{align} $$

for every $w\in \ell ^1(\mathbb {Z}^d,\mathbb {R})$ and $v\in \ell ^{\infty } (\mathbb {Z}^d,\mathbb {R})$ (cf. equation (7.3)).

We define a surjective group homomorphism $\eta \colon \ell ^{\infty } (\mathbb {Z}^d,\mathbb {R}) \to \mathbb {T}^{\mathbb {Z}^d}$ by setting

(7.6) $$ \begin{align} \eta (v)_{\boldsymbol n} = v_{\boldsymbol n} \;(\text{mod}\;1) \end{align} $$

for every $v=(v_{\boldsymbol n})_{{\boldsymbol n}\in \mathbb {Z}^d}$ and ${\boldsymbol n}\in \mathbb {Z}^d$ . Note that $\eta $ is shift-equivariant in the sense that

$$ \begin{align*} \eta \circ \bar\sigma ^{\boldsymbol n} = \sigma ^{\boldsymbol n} \circ \eta \end{align*} $$

for every ${\boldsymbol n}\in \mathbb {Z}^d$ ; more generally, if $w\in R_d=\ell ^1(\mathbb {Z}^d,\mathbb {Z})$ , then

(7.7) $$ \begin{align} \eta \circ w(\bar\sigma ) = w(\sigma )\eta. \end{align} $$

For every $x\in \mathbb {T}^{\mathbb {Z}^d}$ , there exists a unique point $x^{\#} \in (-\tfrac 12,\tfrac 12]^{\mathbb {Z}^d}\subset \ell ^{\infty } (\mathbb {Z}^d,\mathbb {R})$ , called the lift of x, such that

(7.8) $$ \begin{align} \eta (x^{\#}) = x. \end{align} $$

Let $g=\sum _{{\boldsymbol n}\in \mathbb {Z}^d} g_{\boldsymbol n} \boldsymbol {z}^{\boldsymbol n} \in R_d$ be the Laurent polynomial appearing in the statement of Theorem 7.3 and set

$$ \begin{align*} f=g^{*} = \sum_{{\boldsymbol n}\in \mathbb{Z}^d} g_{\boldsymbol n} \boldsymbol{z}^{-{\boldsymbol n} }. \end{align*} $$

Since $U(f)= U(g)$ , f is again atoral and has non-trivial summable homoclinic points by Theorem 7.2.

Lemma 7.4. For every $x\in \mathbb {T}^{\mathbb {Z}^d}$ , the following is true:

  1. (1) $x\in X_f$ if and only if $f(\bar {\sigma })x^{\#} \in \ell ^{\infty } (\mathbb {Z}^d,\mathbb {Z}),\ that\ is, \ x^{\#}\cdot f^{*} \in \ell ^{\infty } (\mathbb {Z}^d,\mathbb {Z}) $ (cf. equation (7.5));

  2. (2) x is a non-trivial summable homoclinic point of $\alpha _f$ if and only if $x^{\#}\in \ell ^1(\mathbb {Z}^d,\mathbb {R})$ , , and h is not divisible by $f^{*}$ in $R_d$ .

Proof. (1) Suppose that $x\in {\mathbb T}^{{\mathbb Z}^d}$ . By equation (7.7), we have

$$ \begin{align*} \eta (f(\bar{\sigma })x^{\#})=f(\sigma )\eta (x^{\#})=f(\sigma )x. \end{align*} $$

So, $x\in X_f$ , that is, $f(\sigma )x=0$ if and only if $f(\bar {\sigma })x^{\#} \in \ell ^{\infty } (\mathbb {Z}^d,\mathbb {Z})$ .

(2) If x is a non-trivial summable homoclinic point of $\alpha _f$ , then $x^\# \in \ell ^1(\mathbb{Z}^d,\mathbb{R})$ , and part (1) of this proof implies that $h=x^\#\cdot f^*\in \ell^1(\mathbb{Z}^d,\mathbb{Z}) = R_d$ . If h were divisible by $f^*$ , that is, if $x^\#\cdot f^* = h\cdot f^*$ for some $h\in R_d$ , then $(x^\#-h)\cdot f^*=0$ and [Reference Linnell and Puls22, Theorem 2.1] would imply that $x^\#=h$ and $x=\eta (x^\#)=\eta (h)=0$ . This violates our conditions on x. The converse is obvious.

Proof of Theorem 7.3

Since $f=g^{*}$ is atoral, there exists a non-trivial summable homoclinic point $x\in X_f$ . Let $x^{\#}\in \ell ^1(\mathbb {Z}^d,\mathbb {R})$ be the lift of x (cf. equation (7.8)), and let $h=f(\bar {\sigma })x^{\#}=x^{\#}\cdot f^{*} \in R_d$ (cf. Lemma 7.4). Since $x^{\#}\in \ell ^1(\mathbb {Z}^d,\mathbb {R})$ , there exists an integer $R=R(x,f,H)$ such that

$$ \begin{align*} \smash[t]{\sum_{\|{\boldsymbol n}\|\ge R} |x^{\#}_{{\boldsymbol n}}| <\frac 1{2H\|f\|_1}.} \end{align*} $$

For every non-empty subset ${\mathcal S}\subset \mathbb {Z}^d$ , we set

$$ \begin{align*} B_R({\mathcal S})=\{{\boldsymbol n}\in \mathbb{Z}^d\mid d({\boldsymbol n},{\mathcal S})=\min_{{\boldsymbol n}'\in {\mathcal S}}\|{\boldsymbol n}-{\boldsymbol n}'\|\le R\}. \end{align*} $$

Let ${\mathcal S},{\mathcal S}'\subset \mathbb {Z}^d$ be two subsets of $\mathbb {Z}^d$ with distance $d({\mathcal S},{\mathcal S}')\ge 3R$ . Suppose that a Laurent polynomial $v\in \mathcal P({\mathcal S}\cup {\mathcal S}', H)$ is divisible by $f^{*}$ , that is, that $v=\phi \cdot f^{*}$ for some $\phi \in R_d$ . Then:

  1. (i) $v\cdot x^{\#}\in R_d$ ;

  2. (ii) $\textsf {supp}(v\cdot x^{\#})\subset B_R({\mathcal S})\cup B_R({\mathcal S}')$ .

Indeed, item (i) follows from Lemma 7.4(2):

$$ \begin{align*} v\cdot x^{\#} = (\phi \cdot f^{*}) \cdot x^{\#}= \phi\cdot (f^{*}\cdot x^{\#})=\phi\cdot h\in R_d, \end{align*} $$

because both $\phi $ and h belong to $R_d$ . For item (ii), we note that every ${\boldsymbol n} \notin B_R({\mathcal S})\cup B_R({\mathcal S}')$ satisfies that $d({\boldsymbol n}, {\mathcal S}\cup {\mathcal S}')> R$ . Then $v_{{\boldsymbol n}-{\boldsymbol m}}=0$ for all ${\boldsymbol m}$ with $\|{\boldsymbol m}\|\le R$ , and hence

$$ \begin{align*} |(v\cdot x^{\#})_{{\boldsymbol n}}|=\bigg|\!\sum_{{\boldsymbol m}\in{\mathbb Z}^d} x^{\#}_{{\boldsymbol m}}v_{{\boldsymbol n}-{\boldsymbol m}}\bigg|\le \|v\|_\infty \sum_{\|{\boldsymbol m}\|>R}| x^{\#}_{{\boldsymbol m}}| < H\cdot \frac {1}{2H\|f\|_1}\le \frac 12. \end{align*} $$

Since $(v\cdot x^{\#})_{{\boldsymbol n}} \in \mathbb {Z}$ by item (i), it follows that $(v\cdot x^{\#})_{{\boldsymbol n}}=0$ .

Let $\psi $ be the restriction of $v \cdot x^{\#}$ to $B_R({\mathcal S})$ , and let $v_{\mathcal S}$ and $v_{{\mathcal S}'}$ be the restrictions of v to ${\mathcal S}$ and ${\mathcal S}'$ , respectively. Then $\psi \in R_d$ by item (i), and we claim that

(7.9) $$ \begin{align} \|\psi - v_{\mathcal S}\cdot x^{\#}\|_{\infty} < \frac{1}{2\|f\|_1}, \end{align} $$

that is, that

(7.10) $$ \begin{align} |\psi_{\boldsymbol n}- (v_{\mathcal S}\cdot x^{\#})_{\boldsymbol n}| < \frac{1}{2\|f\|_1} \quad \text{for every}\enspace {\boldsymbol n} \in {\mathbb Z}^d. \end{align} $$

Indeed, if ${\boldsymbol n} \in B_R({\mathcal S})$ , then $d({\boldsymbol n},B_R({\mathcal S}')) \ge R$ , and hence

(7.11) $$ \begin{align} |(v_{{\mathcal S}'}\cdot x^{\#})_{\boldsymbol n}| = \bigg|\!\sum_{{\boldsymbol m}\in {\mathcal S}'} v_{\boldsymbol m} x^{\#}_{{\boldsymbol n} -{\boldsymbol m}}\bigg|\le \|v\|_\infty \sum_{\|{\boldsymbol \ell}\|\ge R} |x^{\#}_{{\boldsymbol \ell}}|\le H\cdot \frac{1}{2 H \|f\|_1} = \frac{1}{2\|f\|_1}. \end{align} $$

Since $\psi _{\boldsymbol n} = (v_{{\mathcal S}}\cdot x^{\#})_{\boldsymbol n} + (v_{{\mathcal S}'}\cdot x^{\#})_{\boldsymbol n}$ , it follows that

$$ \begin{align*} |\psi _{\boldsymbol n} - (v_{{\mathcal S}}\cdot x^{\#})_{\boldsymbol n}| = |(v_{{\mathcal S}'}\cdot x^{\#})_{\boldsymbol n}| < \frac{1}{2\|f\|_1} \end{align*} $$

by equation (7.11). However, if ${\boldsymbol n}\notin B_R({\mathcal S})$ , then

$$ \begin{align*} |\psi _{\boldsymbol n} - (v_{\mathcal S} \cdot x^{\#})_{\boldsymbol n} |=|(v_{\mathcal S} \cdot x^{\#})_{\boldsymbol n} | = \bigg|\!\sum_{{\boldsymbol m}\in {\mathcal S}}(v_{\boldsymbol m} x^{\#}_{{\boldsymbol n}-{\boldsymbol m}})\bigg| \le H\cdot \frac{1}{2 H \|f\|_1} = \frac{1}{2\|f\|_1}. \end{align*} $$

This proves equation (7.10) for every ${\boldsymbol n}\in \mathbb {Z}^d$ .

Since both $v_{\mathcal S}\cdot x^{\#}\cdot f^{*} = v_{\mathcal S}\cdot h$ and $\psi $ lie in $R_d$ , we have that $(\psi -v_{\mathcal S}\cdot x^{\#})\cdot f^{*} \in R_d$ , but the smallness of the coordinates of $\psi -v_{\mathcal S}\cdot x^{\#}$ in equation (7.10) implies that $(\psi -v_{\mathcal S}\cdot x^{\#})\cdot f^{*}=0$ . Thus, we have proved that $\psi \cdot f^{*}=v_{\mathcal S}\cdot x^{\#}\cdot f^{*}=v_{\mathcal S}\cdot h$ , where h is not divisible by $f^{*}$ (cf. Lemma 7.4(2)). As $g=f^{*}$ is irreducible, we have proved that $v_{\mathcal S}$ is divisible by g, as claimed in the statement of this theorem.

This completes the proof of Theorem 7.3 with $m\ge 3R$ .

7.2 The conditions (C1) and (C2): divisibility by f of lacunary polynomials

According to Theorem 5.1, to prove Theorem 2.3, it suffices to prove that any irreducible and atoral polynomial $f\in R_d$ is m-good for a sufficiently large $m\in {\mathbb N}$ . Now we are going to prove this and finish the proof of Theorem 2.3.

Theorem 7.3 has an immediate corollary which implies that any atoral polynomial is m-good for sufficiently large m.

Corollary 7.5. Suppose that $f\in R_d$ is irreducible and atoral, and that $|\mathrm {supp}(f)|> 1$ . Then there exists, for every $H\ge 1$ , an integer $m\ge 1$ with the following property: for any set ${\mathcal S}\in {\mathbb Z}^d$ which is m-separated in the sense that

$$ \begin{align*} \|{\boldsymbol k}-{\boldsymbol n}\|\ge m\quad \text{for any pair } {\boldsymbol k}, {\boldsymbol n}\in{\mathcal S},\ {\boldsymbol k}\ne{\boldsymbol n}, \end{align*} $$

no non-zero polynomial $g\in \mathcal P({\mathcal S},H)$ is divisible by f.

Proof. For $H\ge 1$ and f fixed, choose m as in the statement of Theorem 7.3 (that is, $m\ge 3R$ in the proof of that theorem). Consider an arbitrary m-separated set ${\mathcal S}$ and any non-trivial polynomial $v = \sum _{{\boldsymbol n}\in {\mathcal S}}v_{\boldsymbol n}\boldsymbol {z}^{\boldsymbol n}\in \mathcal P({\mathcal S},H)$ .

If $|\textsf {supp}(v)|=1$ , then v cannot be divisible by f, since $|\textsf {supp}(f)|>1$ by assumption. Assume therefore that $|\textsf {supp}(v)|\ge 2$ , and that v is divisible by f. Since for any ${\boldsymbol n}\in \textsf {supp}(v)$ , the sets

$$ \begin{align*} \mathcal{T}=\{{\boldsymbol n}\},\quad \mathcal{T}'=\textsf{supp}(v)\smallsetminus \{{\boldsymbol n}\} \end{align*} $$

have distance at least m and hence, by Theorem 7.3, the restriction of v to $\mathcal {T}$ , that is, $v_{\mathcal {T}}=v_{{\boldsymbol n}}{\boldsymbol z}^{{\boldsymbol n}}$ must be divisible by f, which is impossible, v is not divisible by f.

The condition that $|\textsf {supp}(f)|>1$ in Corollary 7.5 is obviously necessary: the polynomial $f=2$ is obviously irreducible and atoral, and divides $2g$ for every $g\in R_d$ (irrespective of whether g is m-separated or not).

Corollary 7.6. Suppose that $f\in R_d$ is irreducible and atoral, and that $|\mathrm {supp}(f)|>~1$ . For all sufficiently large $m\ge 1$ and every $\boldsymbol {k}\in [0,m-1]^d\smallsetminus \{{\boldsymbol 0}\}$ , no $v\in \mathcal {P}(m\mathbb {Z}^d \cup (m\mathbb {Z}^d+{\boldsymbol k}),1)$ with $v \ne 0$ is divisible by f.

Proof. Put $H=1$ and let $m\ge 6R$ , where R is the number appearing in the proof of Theorem 7.3. Suppose $v\in \mathcal {P}(m\mathbb {Z}^d \cup (m\mathbb {Z}^d+\boldsymbol {k}),1)$ is a non-trivial polynomial divisible by f. Consider the decomposition $\textsf {supp}(v)={\mathcal S}_0\sqcup {\mathcal S}_1$ where

$$ \begin{align*} {\mathcal S}_0=\textsf{supp}(v)\cap m{\mathbb Z}^d,\quad {\mathcal S}_1=\textsf{supp}(v)\cap (m{\mathbb Z}^d+{\boldsymbol k}). \end{align*} $$

Both sets ${\mathcal S}_0,{\mathcal S}_1$ are m-separated, as subsets of $m{\mathbb Z}^d$ and $m{\mathbb Z}^d+{\boldsymbol k}$ , respectively.

We claim that for any ${\boldsymbol n}\in {\mathcal S}_0$ , there exists ${\boldsymbol n}'={\boldsymbol n}'({\boldsymbol n})\in {\mathcal S}_1$ such that $d({\boldsymbol n},{\boldsymbol n}')<3R$ . Otherwise, there exists ${\boldsymbol n}\in {\mathcal S}_0$ such that $d({\boldsymbol n}, {\mathcal S}_1)\ge 3R$ so that $d({\boldsymbol n}, \textsf {supp}(v)\smallsetminus \{{\boldsymbol n}\})\ge 3R$ . Then, by Theorem 7.3, the restriction of v to $\{{\boldsymbol n}\}$ , that is, $\pm {\boldsymbol z}^{{\boldsymbol n}}$ , is divisible by f, which is impossible. Similarly, for any ${\boldsymbol n}'\in {\mathcal S}_1$ , there exists ${\boldsymbol n}\in S_0$ such that $d({\boldsymbol n},{\boldsymbol n}')<3R$ . Thus, the support of v is a union of distinct pairs:

$$ \begin{align*} \textsf{supp}(v)=\bigcup_{{\boldsymbol n}\in{\mathcal S}_0}\{{\boldsymbol n},{\boldsymbol n}'\}, \end{align*} $$

where the distance within each pair is at most $3R$ .

Given a pair $\{{\boldsymbol n},{\boldsymbol n}'\}$ , consider the decomposition of $\textsf {supp}(v)$ :

$$ \begin{align*} {\mathcal S}=\{{\boldsymbol n},{\boldsymbol n}'\}, \quad {\mathcal S}'=\textsf{supp}(v)\smallsetminus {\mathcal S}. \end{align*} $$

The fact that $m\ge 6R$ implies $d({\mathcal S}, {\mathcal S}')\ge 3R$ . Indeed, $d({\boldsymbol n}, {\mathcal S}')= d({\boldsymbol n}, {\boldsymbol n}^{*})$ for some ${{\boldsymbol n}^{*}\in {\mathcal S}'}$ and

$$ \begin{align*} \begin{cases} d({\boldsymbol n}, {\boldsymbol n}^{*}) \ge m &\mbox{if}\ {\boldsymbol n}^{*}\in {\mathcal S}_0; \\ d({\boldsymbol n}, {\boldsymbol n}^{*})\ge d({\boldsymbol n}', {\boldsymbol n}^{*}) - d({\boldsymbol n}', {\boldsymbol n}) \ge m -3R &\mbox{if}\ {\boldsymbol n}^{*}\in {\mathcal S}_1. \end{cases} \end{align*} $$

It follows that $d({\boldsymbol n}, {\mathcal S}')>3R$ . Similarly, $d({\boldsymbol n}', {\mathcal S}')>3R$ .

Applying Theorem 7.3 to ${\mathcal S}$ and ${\mathcal S}'$ , we conclude that the restriction of v to ${\mathcal S}=\{{\boldsymbol n},{\boldsymbol n}'\}$ , that is,

$$ \begin{align*} v_{{\mathcal S}} =v_{{\boldsymbol n}} {\boldsymbol z}^{{\boldsymbol n}}+v_{{\boldsymbol n}'} {\boldsymbol z}^{{\boldsymbol n}'}, \quad v_{{\boldsymbol n}}, v_{{\boldsymbol n}'}\in\{-1,1\}, \end{align*} $$

must be divisible by f, which is impossible, since $v_{\mathcal S}$ is of the form

$$ \begin{align*} \pm {\boldsymbol z}^{{\boldsymbol m}}(1\pm {\boldsymbol z}^{{\boldsymbol \ell}}),\quad {\boldsymbol m}\in{\mathbb Z}^d, {\boldsymbol \ell}\in\mathbb{N}^d, \end{align*} $$

and hence is a product of a unit ( $\pm {\boldsymbol z}^{{\boldsymbol m}}$ ) and a generalized cyclotomic polynomial ${(1\pm {\boldsymbol z}^{{\boldsymbol \ell }})}$ , and thus must have zero Mahler measure $m(v_S)=0$ . This implies that $\mathsf {m}(f) =0$ , in violation of Theorem 7.2.

Proof of Theorem 2.3

Since Remark 2.7 allows us to assume without loss in generality that the polynomial $f\in R_d$ is primitive and irreducible, the proof of Bohr chaoticity under the additional assumption of atorality of f is now complete.

Indeed, if $|\textsf {supp}(f)|=1$ , atorality implies that we are in the situation of Example 2.5 with $p>1$ , so that $(X_f,\alpha _f)$ is Bohr chaotic. If $|\textsf {supp}(f)|\ge 2$ , Corollary 7.5 for ${H=2}$ and Corollary 7.6 show that conditions (C1) and (C2) are satisfied. Therefore, Bohr chaoticity of $(X_f,\alpha _f)$ for irreducible atoral polynomials $f\in R_d$ follows from Theorem 5.1.

8 Concluding remarks: toral polynomials

We have shown that a principal ${\mathbb Z}$ -action is Bohr chaotic if and only if it has positive entropy. A principal ${\mathbb Z}^d$ -action, $d>1$ , is shown to be Bohr chaotic if it has positive entropy and is atoral. We believe that the atorality assumption (equivalently, existence of a non-trivial summable homoclinic point) can be removed.

Irreducible toral polynomials come in two flavors: those for which $X_f$ has no non-trivial homoclinic points (cf. e.g., [Reference Lind and Schmidt19, Example 7.1]), and those for which $X_f$ has no summable homoclinic points, but uncountably many non-zero homoclinic points $v\in X_f$ with the property that $v^{\#} \cdot f^{*} \in f^{*} R_d$ (cf. e.g., [Reference Lind and Schmidt19, Example 7.3]; for notation, we refer to equation (7.8)). Unfortunately, none of these latter homoclinic points can be used in our proof of the gap theorem (Theorem 7.3), since the key Lemma 7.4 is not valid in this case.

Remarkably, for toral examples of the kind illustrated in [Reference Lind and Schmidt19, Example 7.1], we can still prove Bohr chaoticity by using Theorems 5.1 and 6.1.

Theorem 8.1. Let $g\in R_1$ be an irreducible polynomial with positive Mahler measure and with all roots of absolute value $1$ , and define $f\in R_d$ by $f(z_1,\ldots ,z_d)=g(z_1)$ . Then $(X_f,\alpha _f)$ is Bohr chaotic.

Remark 8.2. There exist infinitely many distinct irreducible polynomials $g\in R_1$ with the properties required in Theorem 8.1. To see this, we follow a short note from math.stackexchange.com [24].

Let $\vartheta $ be a totally real algebraic number all of whose conjugates $\vartheta _1=\vartheta ,\vartheta _2,\ldots ,\vartheta _r$ have absolute values strictly less than 2 (to find such a $\vartheta $ , take any totally real algebraic number and divide by a big integer). Assume also that $\vartheta $ is not itself an algebraic integer. Let $\beta $ be a solution to the equation

$$ \begin{align*} \beta +1/\beta =\vartheta. \end{align*} $$

All the conjugates of $\vartheta $ are, by assumption, real numbers in the interval $(-2,2)$ . This forces all the conjugates of $\beta $ to be complex numbers of absolute value one—which is what we want. Moreover, $\beta $ will not be a root of unity, since otherwise, $\vartheta $ would be an algebraic integer. Then $\beta $ is a root of the polynomial

$$ \begin{align*} h(x)=\prod _{i=1}^r (x^2 - \vartheta _ix + 1). \end{align*} $$

Clearing denominators in h, one gets the desired polynomial g. It will be irreducible, because, by looking at infinite places, $[\mathbb {Q}(\beta ):\mathbb {Q}(\vartheta )]=2$ .

For example, $\vartheta =1/2$ yields $g=2z^2-1+2$ , $\vartheta = -6/5$ yields $g=5z^2-6z+5$ , $\vartheta = 1/\sqrt 2$ yields $g=2z^4+3z^2+2$ , etc.

Proof of Theorem 8.1

Since g with $\mathsf m(g)>0$ is m-good for some sufficiently large m by Theorem 6.1, $f(z_1,\ldots ,z_d)=g(z_1)$ is also m-good, but now viewed as a polynomial in d-variables. Indeed, if $f({\boldsymbol z})=g(z_1)$ divides a non-trivial polynomial h of the form

$$ \begin{align*} h({\boldsymbol z}) &=\sum_{{\boldsymbol n} \in \mathbb{Z}^d} \varepsilon _{\boldsymbol n} {\boldsymbol z}^{m{\boldsymbol n}} \end{align*} $$

with $\varepsilon _{\boldsymbol n}=\varepsilon _{(n_1,n_2,\ldots ,n_d)} \in \{-2,-1,0,1,2\}$ , then by rewriting h as

$$ \begin{align*} h({\boldsymbol z}) &= \sum_{(n_2,\ldots,n_{d})\in {\mathbb Z}^{d-1}} \bigg( \sum_{n_1\in{\mathbb Z}} \varepsilon_{(n_1,n_2,\ldots,n_d)} z_1^{m n_1}\bigg) z_2^{m n_2}\cdots z_{d}^{m n_d} \\ &=: \sum_{(n_2,\ldots,n_{d})\in {\mathbb Z}^{d-1}} h_{n_2,\ldots,n_d}(z_1^m) z_2^{mn_2}\cdots z_{d}^{mn_d}, \end{align*} $$

we conclude that $g(z_1)$ must divide all polynomials $h_{n_2,\ldots ,n_d}(z_1^m)$ , some of which are non-zero. Since $g(z_1)$ is m-good, the resulting contradiction proves that condition (C1) is valid for f. For the proof of condition (C2), we can proceed similarly.

Theorem 8.1 now follows from Theorem 5.1.

Theorem 8.1 obviously applies only to an extremely restricted class of toral polynomials. A more interesting example of an irreducible toral polynomial with positive entropy in [Reference Lind and Schmidt19, Example 7.3] is given by the polynomial

$$ \begin{align*} f(z_1,z_2)= 3-z_1-\frac 1{z_1}-z_2-\frac 1{z_2}. \end{align*} $$

The unitary variety of f is a smooth real-analytic curve

$$ \begin{align*} \mathsf U(f)=\bigg\{(e^{2\pi i s},e^{2\pi i t}): t=\pm \frac{1}{2 \pi} \cos ^{-1}\bigg(\frac{3}{2}-\cos 2 \pi s\bigg), - \frac{1}{6} \leq s \leq \frac{1}{6} \bigg\}. \end{align*} $$

Moreover, $\mathsf U(f)$ is connected and has curvature bounded away from zero. As explained in [Reference Lind and Schmidt19, Example 7.3], one can use this fact to prove the existence of probability measures supported on $\mathsf U(f)$ whose Fourier transforms (as functions on $\widehat {\mathbb {S}^d}=\mathbb {Z}^d$ ) vanish at infinity. By translating this information back to the system $(X_f,\alpha _f)$ , one obtains uncountably many homoclinic points $x\in X_f$ satisfying

$$ \begin{align*} |x_{{\boldsymbol n}}| \le \frac {C}{1+\|{\boldsymbol n}\|^{\tfrac 12}}\quad \text{for all } {\boldsymbol n}\in {\mathbb Z}^2. \end{align*} $$

Unfortunately, none of these homoclinic points can be used in our proof of the gap theorem (Theorem 7.3), since they are not summable.

It would be interesting to see whether one can prove the gap theorem for $f=3-z_1- ({1}/{z_1})-z_2- ({1}/{z_2})$ directly, using some elementary methods, or establish Bohr chaoticity of $X_f$ by some other means.

Acknowledgements

The authors would like to thank B. Weiss for valuable discussions and D. Lind for alerting them to an error in the submitted manuscript. K.S. and E.V. are grateful to Central China Normal University for their hospitality, where part of the work was done. A.H.F. was partly supported by NSF of China (Grant Nos. 11971192 and 12231013).

References

Bourgain, J.. On the spectral type of Ornstein’s class one transformations. Israel J. Math. 84(1–2) (1993), 5363.CrossRefGoogle Scholar
Boyd, D. W.. Kronecker’s theorem and Lehmer’s problem for polynomials in several variables. J. Number Theory 13(1) (1981), 116121.CrossRefGoogle Scholar
Davenport, H., Erdős, P. and LeVeque, W. J.. On Weyl’s criterion for uniform distribution. Michigan Math. J. 10 (1963), 311314.Google Scholar
Dimitrov, V.. Convergence to the Mahler measure and the distribution of periodic points for algebraic Noetherian ${\mathrm{Z}}^{\mathrm{d}}$ -actions. Preprint, 2017, arXiv:1611.04664v2 [math.DS].Google Scholar
El Abdalaoui, E. H., Kułaga-Przymus, J., Lemańczyk, M. and de la Rue, T.. The Chowla and the Sarnak conjectures from ergodic theory point of view. Discrete Contin. Dyn. Syst. 37(6) (2017), 28992944.CrossRefGoogle Scholar
Fan, A.. Quelques propriétés des produits de Riesz. Bull. Sci. Math. 117(4) (1993), 421439.Google Scholar
Fan, A.. Multifractal analysis of weighted ergodic averages. Adv. Math. 377 (2021), 107488.CrossRefGoogle Scholar
Fan, A., Fan, S. and Qiu, Y.. Some properties of stationary determinantal point processes on $\mathbb{Z}$ . J. Lond. Math. Soc. (2) 98(3) (2018), 517535.CrossRefGoogle Scholar
Fan, A., Fan, S., Ryzhikov, V. and Shen, W.. Bohr-chaoticity of topological dynamical systems. Math. Z. 302 (2022), 11271154.CrossRefGoogle Scholar
Franklin, J. N.. Deterministic simulation of random processes. Math. Comp. 17 (1963), 2859.CrossRefGoogle Scholar
Furstenberg, H.. Disjointness in ergodic theory, minimal sets, and a problem in Diophantine approximation. Math. Systems Theory 1 (1967), 149.CrossRefGoogle Scholar
Glasner, E., Thouvenot, J.-P. and Weiss, B.. Entropy theory without a past . Ergod. Th. & Dynam. Sys. 20(5) (2000), 13551370.CrossRefGoogle Scholar
Hewitt, E. and Zuckerman, H. S.. Singular measures with absolutely continuous convolution squares. Math. Proc. Cambridge Philos. Soc. 62(3) (1966), 399420.CrossRefGoogle Scholar
Kahane, J.-P.. Séries de Fourier absolument convergentes (Ergebnisse der Mathematik und ihrer Grenzgebiete, 50). Springer-Verlag, Berlin, 1970.CrossRefGoogle Scholar
Katok, A.. Lyapunov exponents, entropy and periodic orbits for diffeomorphisms. Publ. Math. Inst. Hautes Études Sci. 51 (1980), 137173.CrossRefGoogle Scholar
Katznelson, Y.. An Introduction to Harmonic Analysis (Cambridge Mathematical Library), 3rd edn. Cambridge University Press, Cambridge, 2004.CrossRefGoogle Scholar
Kronecker, L.. Zwei Sätze über Gleichungen mit ganzzahligen Coefficienten. J. Reine Angew. Math. 53 (1857), 173175 (in German).Google Scholar
Ledrappier, F.. Des produits de Riesz comme mesures spectrales. Ann. Inst. H. Poincaré Sect. B (N.S.) 6 (1970), 335344.Google Scholar
Lind, D. and Schmidt, K.. Homoclinic points of algebraic Zd-actions. J. Amer. Math. Soc. 12(4) (1999), 953980.CrossRefGoogle Scholar
Lind, D., Schmidt, K. and Verbitskiy, E.. Homoclinic points, atoral polynomials, and periodic points of algebraic ${\mathbb{Z}}^{\mathrm{d}}$ -actions. Ergod. Th. & Dynam. Sys. 33(4) (2013), 10601081.CrossRefGoogle Scholar
Lind, D., Schmidt, K. and Ward, T.. Mahler measure and entropy for commuting automorphisms of compact groups. Invent. Math. 101(3) (1990), 593629.CrossRefGoogle Scholar
Linnell, P. A. and Puls, M. J.. Zero divisors and $L^p(G)$ . II. New York J. Math. 7 (2001), 4958.Google Scholar
Liu, J. and Sarnak, P.. The Möbius function and distal flows. Duke Math. J. 164(7) (2015), 13531399.CrossRefGoogle Scholar
math.stackexchange. Non-monic polynomial with roots on the unit circle (version: 2017-04-13). Available at https://math.stackexchange.com/q/1275978.Google Scholar
Peyrière, J.. Étude de quelques propriétés des produits de Riesz. Ann. Inst. Fourier (Grenoble) 25(2) (1975), 127169.CrossRefGoogle Scholar
Queffélec, M.. Substitution Dynamical Systems—Spectral Analysis (Lecture Notes in Mathematics, 1294). Springer-Verlag, Berlin, 1987.CrossRefGoogle Scholar
Riesz, F.. Über die Fourierkoeffizienten einer stetigen Funktion von beschränkter Schwankung. Math. Z. 2(3–4) (1918), 312315 (in German).CrossRefGoogle Scholar
Sarnak, P.. Three lectures on the Möbius function randomness and dynamics (2011). Available at http://publications.ias.edu/sites/default/files/MobiusFunctionsLectures(2).pdf.Google Scholar
Schmidt, K.. Dynamical Systems of Algebraic Origin. Birkhäuser Verlag, Basel, 1995.Google Scholar
Smyth, C. J.. A Kronecker-type theorem for complex polynomials in several variables. Canad. Math. Bull. 24(4) (1981), 447452; Addenda and errata, Canad. Math. Bull. 24(4) (1981), 504.CrossRefGoogle Scholar
Tal, M.. Some remarks on the notion of Bohr chaos and invariant measures. Studia Math. 271 (2023), 347359.CrossRefGoogle Scholar
Young, L. S.. On the prevalence of horseshoes. Trans. Amer. Math. Soc. 263(1) (1981), 7588.CrossRefGoogle Scholar
Young, R. M.. Notes: On Jensen’s Formula and ${\int}_0^{2\pi}\log |1-{e}^{i\theta}| d\theta$ . Amer. Math. Monthly 93(1) (1986), 4445.Google Scholar
Zygmund, A.. Trigonometric Series. Vol. I, II (Cambridge Mathematical Library), 3rd edn. Cambridge University Press, Cambridge, 2002; with a foreword by R. A. Fefferman.Google Scholar