Abstract

We show that Sarnak’s conjecture on Möbius disjointness holds in every uniquely ergodic model of a quasi-discrete spectrum automorphism. A consequence of this result is that, for each non constant polynomial |$P\in{\mathbb{R}}[x]$| with irrational leading coefficient and for each multiplicative function |$\boldsymbol{\nu}:{\mathbb{N}}\to{\mathbb{C}}$|⁠, |$|\boldsymbol{\nu}|\leq1$|⁠, we have
1MMm<2M1H|mn<m+He2πiP(n)ν(n)|0
as |$M\to\infty$|⁠, |$H\to\infty$|⁠, |$H/M\to 0$|⁠.

1 Introduction

The central result of this article is the following.

 
Theorem 1.
Let |$S$| be a uniquely ergodic continuous map defined on a compact metric space |$Y$|⁠. Let |$\nu$| be the unique |$S$|-invariant (ergodic) Borel probability measure on |$Y$|⁠. Assume that the measure-theoretic system |$(Y,\nu,S)$| has quasi-discrete spectrum. Then, for each multiplicative function |$\boldsymbol{\nu}:{\mathbb{N}}\to{\mathbb{C}}$|⁠, |$|\boldsymbol{\nu}|\leq 1$|⁠, each |$f\in C(Y)$| with |$\int_Y f\,\mathrm{d}\nu=0$| and each |$y\in Y$|⁠, we have
limN1NnNf(Sny)ν(n)=0.
(1)

If, moreover, we know that |$\lim_{N\to\infty}\frac1N\sum_{n\leq N} \boldsymbol{\nu}(n)=0$|⁠, then (1) holds regardless of the value of |$\int_Y f\,{\rm d}\nu$|⁠. In particular, by taking |$\boldsymbol{\nu}=\boldsymbol{\mu}$|⁠, the classical Möbius function: |$\boldsymbol{\mu}(1)=1$|⁠, |$\boldsymbol{\mu}(p_1\ldots p_k)=(-1)^k$| for different prime numbers |$p_1,\ldots,p_k$| and |$\boldsymbol{\mu}(n)=0$| for other values of |$n\in{\mathbb{N}}$|⁠, we obtain the validity of Sarnak’s conjecture [29] in the class of systems under consideration. Recall that Sarnak’s conjecture states that, if |$T$| is a continuous map of a compact metric space |$X$| with zero topological entropy, then |$\frac1N\sum_{n\leq N}f(T^nx)\boldsymbol{\mu}(n)\to 0$| for each |$f\in C(X)$| and |$x\in X$|⁠. Sarnak’s conjecture has been proved to hold in several cases [3-6, 8-11, 14, 15, 21, 24, 26-30]. Automorphisms with quasi-discrete spectrum have zero measure-theoretic entropy, hence the topological entropy will be equal to zero in each uniquely ergodic model of them.

So far, the only automorphisms for which Sarnak’s conjecture was known to hold in each of their uniquely ergodic model were those whose prime powers are disjoint in the sense of Furstenberg [12]. This assertion is already present in [5], see also [10]. However, it is easy to see that even irrational rotations do not enjoy this strong disjointness property (indeed, powers have “large” common factors). But Theorem 1 is known to hold for them [5]. On the other hand, irrational rotations have many other uniquely ergodic models. For those in which eigenfunctions are continuous, for example, symbolic models like Sturmian models [2], Sarnak’s conjecture has also been proved to hold in view of a general criterion on lifting Sarnak’s conjecture by continuous uniquely ergodic extensions, [6, 8, 30]. However, there are topological models of irrational rotations in which eigenfunctions are not continuous. Indeed, each ergodic transformation admits even a uniquely ergodic model which is topologically mixing [23].

An important tool for the proof of Theorem 1 is provided by a general criterion for an ergodic automorphism |$T$| to satisfy Sarnak’s conjecture in each of its uniquely ergodic model. The criterion is, roughly, to establish an approximative disjointness of sufficiently large relatively prime powers.

 
Definition 1.
We say that the measure-theoretic dynamical system |$(X,\mu,T)$| has Asymptotical Orthogonal Powers (AOP) if, for any |$f$| and |$g$| in |$L^2(\mu)$| with |$\int_X f\, {\rm d}\mu = \int_X g\, {\rm d}\mu =0$|⁠, we have
(2)

We would like to emphasize that this AOP property can even be satisfied for an automorphism |$T$| for which all non-zero powers are measure-theoretically isomorphic (Example 1). Further remarks on the AOP property are discussed in Section 5: we show that it implies zero entropy, but give also examples of zero-entropy automorphisms which do not have AOP.

 
Theorem 2.
Assume that |$T$| is a totally ergodic automorphism of a standard Borel probability space |$(X,{\cal B},\mu)$|⁠, and that it has AOP. Let |$S$| be a uniquely ergodic continuous map of a compact metric space |$Y$|⁠, and denote by |$\nu$| the unique |$S$|-invariant probability measure. Assume that the measure-theoretic systems |$(X,\mu,T)$| and |$(Y,\nu,S)$| are isomorphic. Then, for each multiplicative function |$\boldsymbol{\nu}:{\mathbb{N}}\to{\mathbb{C}}$|⁠, |$|\boldsymbol{\nu}|\leq 1$|⁠, each |$f\in C(Y)$| with |$\int_Y f\,{\rm d}\nu=0$| and each |$y\in Y$|⁠, we have
limN1NnNf(Sny)ν(n)=0.

In particular, Sarnak’s conjecture holds for |$S$|⁠. □

Using the fact that, in uniquely ergodic models, the average behavior of points is uniform, we will prove that uniquely ergodic models of automorphisms with AOP enjoy a stronger form of Sarnak’s conjecture, in which the sequence output by the system is allowed to switch orbit from time to time.

 
Theorem 3.
Let |$(X,T)$| be uniquely ergodic, and let |$\mu$| be the unique |$T$|-invariant probability measure. Assume that, for any uniquely ergodic model |$(Y,S)$| of |$(X,\mu,T)$|⁠, the conclusion of Theorem 2 holds. Let |$1=b_1<b_2<\cdots$| be an increasing sequence of integers with |$b_{k+1}-b_k\to\infty$|⁠. Let |$(x_k)$| be an arbitrary sequence of points in |$X$|⁠. Then, for each multiplicative function |$\boldsymbol{\nu}:{\mathbb{N}}\to{\mathbb{C}}$|⁠, |$|\boldsymbol{\nu}|\leq1$| and for each |$f\in C(X)$| with |$\int_X f\,{\rm d}\mu=0$|⁠, we have
limK1bK+11kKbkn<bk+1f(Tnxk)ν(n)=0.
(3)

Again, if we take |$\boldsymbol{\nu}=\boldsymbol{\mu}$|⁠, the conclusion holds even if |$\int_X f\,{\rm d}\mu\neq 0$|⁠.

One of the main results of the aritcle, Theorem 6, says that all quasi-discrete spectrum automorphisms have AOP. Quasi-discrete spectrum automorphisms (see Section 3.3 for a precise definition) have been first studied by Abramov [1]. They admit special uniquely ergodic models which are of the form |$Tx=Ax+b$|⁠, where |$X$| is a compact, Abelian, connected group, |$A$| is a continuous group automorphism and |$b\in X$| (with some additional assumptions on |$A$| and |$b$|⁠). It has already been proved by Liu and Sarnak in [24] that, in these algebraic models, Sarnak’s conjecture holds. As a consequence of Theorem 2, we obtain that it holds in every uniquely ergodic model of any of these systems (in fact, Theorem 2 directly gives Theorem 1).

In Section 4.2, we apply Theorem 3 in a special situation of such algebraic models which are built by taking affine cocycle extensions of irrational rotations (see [12]). This leads to the following theorem.

 
Theorem 4.
Assume that |$\boldsymbol{\nu}:{\mathbb{N}}\to{\mathbb{C}}$|⁠, |$|\boldsymbol{\nu}|\leq1$|⁠, is multiplicative. Then, for each increasing sequence |$1=b_1<b_2<\cdots$| with |$b_{k+1}-b_k\to\infty$| and for each non constant polynomial |$P\in{\mathbb{R}}[x]$| with irrational leading coefficient, we have
1bK+1kK|bkn<bk+1e2πiP(n)ν(n)|0 when K.

It has been noted to us by N. Frantzikinakis that the particular case |$P(n)=\alpha n$|⁠, which we present in Section 4.1, follows from a recent result of Matomäki et al. [25]. We can also reformulate Theorem 4 in the following way.

 
Theorem 5.
Assume that |$\boldsymbol{\nu}:{\mathbb{N}}\to{\mathbb{C}}$|⁠, |$|\boldsymbol{\nu}|\leq1$|⁠, is multiplicative. For each non constant polynomial |$P\in{\mathbb{R}}[x]$| with irrational leading coefficient, we have
1MMm<2M1H|mn<m+He2πiP(n)ν(n)|0
as |$M\to\infty$|⁠, |$H\to\infty$|⁠, |$H/M\to 0$|⁠. □

2 Main Tools

2.1 Joinings

Let |$(X,{\cal B},\mu)$| be a standard Borel probability space. We denote by |$C_2(X,\mu)$| the corresponding space of couplings: |$\rho\in C_2(X,\mu)$| if |$\rho$| is a probability on |$X\times X$| with both projections equal to |$\mu$|⁠. The space |$C_2(X,\mu)$| is endowed with a topology which coincides with the weak topology when |$X$| is a compact metric space. In this topology, convergence of a sequence of couplings |$(\rho_n)$| to some coupling |$\rho$| is characterized by
f,gL2(μ),X×XfgdρnnX×Xfgdρ.
This topology turns |$C_2(X,\mu)$| into a compact metrizable space. For example, the formula
d(ρ1,ρ2):=k,1|fkfdρ1fkfdρ2|2k+,
where |$\{f_k: k\ge 1\}$| is a linearly |$L^2$|-dense set of functions of |$L^2$|-norm 1 yields a metric compatible with the topology on |$C_2(X,\mu)$|⁠.

Now, let |$T$| be a totally ergodic automorphism of |$(X,{\cal B},\mu)$|⁠, that is, |$T^m$| is assumed to be ergodic for each |$m\neq0$|⁠. Let |$J(T^r,T^s)$| stand for the set of joinings between |$T^r$| and |$T^s$|⁠, that is, those |$\rho\in C_2(X,\mu)$| which are |$T^r\times T^s$|-invariant. By |$J^e(T^r,T^s)$| we denote the subset of ergodic joinings, which is never empty when both |$T^r$| and |$T^s$| are ergodic.

2.2 The KBSZ criterion

We will need the following version of the result of Bourgain et al. [5] (see also [18,22]).

 
Proposition 1.
Assume that |$(a_n)$| is a bounded sequence of complex numbers. Assume, moreover, that
(4)
Then, for each multiplicative function |$\boldsymbol{\nu}:{\mathbb{N}}\to{\mathbb{C}}$|⁠, |$|\boldsymbol{\nu}|\leq 1$|⁠, we have
limN1NnNanν(n)=0.
(5)

The proof of this proposition follows directly from [5] by considering a version of their result omitting a finite set of prime numbers.

 
Proof of Theorem 2
We begin by observing that the condition of having AOP is an invariant of measure-theoretic isomorphism. Therefore, if it holds for |$(X,\mu,T)$|⁠, it also holds in each of its uniquely ergodic model. Let us fix such a model |$(Y,\nu,S)$|⁠. Take |$y\in Y$| and |$f\in C(Y)$| with |$\int_Yf\,{\rm d}\nu=0$|⁠. Let |$\varepsilon>0$|⁠. In view of (2), for some |$M\geq1$|⁠, whenever |$r,s\geq M$| are different prime numbers, we have for each |$\kappa\in J^e(S^r,S^s)$|
|Y×Yff¯dκ|ε.
(6)
Now, let |$r, s\geq M$| be different prime numbers, and select a subsequence |$(N_k)$| so that
1NknNkδ(Srny,Ssny)ρ.
(7)
Then, by the unique ergodicity of |$S$|⁠, which implies the unique ergodicity of |$S^r$| and |$S^s$|⁠, it follows that |$\rho\in J(S^r,S^s)$|⁠. Let
ρ=Je(Sr,Ss)κdP(κ)
stand for the ergodic decomposition into ergodic joinings between |$S^r$| and |$S^s$| of |$\rho$| (here, we use again that the non-trivial powers of |$S$| are ergodic). Since |$f$| is continuous, in view of (7), we have
1NknNkf(Srny)f(Ssny)¯Y×Yff¯dρ.
But, in view of (6),
|Y×Yff¯dρ|=|Je(Sr,Ss)(Y×Yff¯dκ)dP(κ)|ε.
It follows that
lim supN|1NnNf(Srny)f(Ssny)¯|ε.

We have obtained that condition (4) holds for |$a_n:=f(S^ny)$|⁠, |$n\geq0$|⁠, and the result follows from Proposition 1. ■

2.3 Special uniquely ergodic models

Our idea in this section will be, given a uniquely ergodic system |$(X,T)$| (with |$\mu$| as unique invariant probability measure), to build many new uniquely ergodic models of the measure-theoretic system |$(X,\mu,T)$|⁠. Any such model will depend on the choice of an increasing sequence of integers |$1=b_1<b_2<\cdots$| with |$b_{k+1}-b_k\to\infty$|⁠, and of an arbitrary sequence of points |$(x_k)$| in |$X$|⁠.

We define a sequence |$\underline{y}=(y_n)\in X^{{\mathbb{N}}}$| by setting |$y_n:=T^n x_k$| whenever |$b_k\le n< b_{k+1}$|⁠:
Y_=(Tx1,,Tb21x1,Tb2x2,,Tb31x2,Tb3x3,).
Let |$S$| denote the shift in |$X^{{\mathbb{N}}}$|⁠, where |$X^{{\mathbb{N}}}$| is considered with the product topology, and set
Y:={Sny_:n0}¯XN.
 
Proposition 2.
The topological dynamical system |$(Y,S)$| defined in the previous paragraph is uniquely ergodic. The unique |$S$|-invariant probability measure on |$Y$| is the graph measure |$\nu$| determined by the formula
ν(A1×A2××Am×X×):=μ(A1T1A2Tm+1Am).

Moreover, the measure-theoretic dynamical system |$(Y,\nu,S)$| is isomorphic to |$(X,\mu,T)$|⁠. □

 
Proof
We first prove that |$\underline{y}$| is generic for the measure |$\nu$| defined in the statement of the proposition. Indeed, let |$\kappa$| be a probability measure on |$Y$| for which |$\underline{y}$| is quasi-generic, that is, assume that there exists a sequence of integers |$N_j\to\infty$| such that
1NjnNjδSny_κ.
For fixed |$j\ge 1$|⁠, the projection of the left-hand side on the first coordinate is
νj:=1Nj(b1n<b2δTnx1+b2n<b3δTnx2++bk1n<bkδTnxk1+bkn<bk+uδTnxk),
where |$b_k+u=N_j<b_{k+1}$|⁠. Then, the difference the pushforward of |$\nu_j$| by |$T$| and |$\nu_j$| is
Tνjνj=1Nj(δTb2x1δTx1+δTb3x2δTb2x2++δTbkxk1δTbk1xk1+δTbk+u+1xkδTbkxk),
and this measure goes weakly to zero as |$b_{k+1}-b_k\to\infty$|⁠. Hence the limit of |$\nu_j$| is |$T$|-invariant and, by unique ergodicity, the projection of |$\kappa$| on the first coordinate has to be equal to |$\mu$|⁠. Now, using again |$b_{k+1}-b_k\to\infty$|⁠, we see that the proportion of integers |$n\le N_j$| such that the second coordinate of |$S^n \underline{y}$| is not the image by |$T$| of the first coordinate of |$S^n \underline{y}$| go to zero as |$j\to\infty$|⁠, and it follows that, |$\kappa$|-almost surely, the second coordinate of |$\underline{z}\in Y$| is the image by |$T$| of the first coordinate of |$\underline{z}$|⁠. The same argument works for any coordinate and we conclude that |$\kappa=\nu$|⁠.
Now, the fact that each |$\underline{z}\in Y$| is also generic for the same measure |$\nu$| is a direct consequence of the following easy fact: If |$n_j\to\infty$| and |$\underline{z}=\lim_{j\to\infty}S^{n_j}\underline{y}$|⁠, then either there exists |$z\in X$| such that |$\underline{z}=(z,Tz,T^2z,\ldots)$|⁠, or there exist |$z_1,z_2\in X$| and |$\ell\ge0$| such that
z_=(z1,Tz1,,Tz1,z2,Tz2,T2z2,).

Indeed, we can approximate any “window” |$\underline{z}[1,M]$| by |$S^{n_j}\underline{y}[1,M]=\underline{y}[n_j+1,n_j+M]$|⁠, and when |$n_j\to\infty$|⁠, such a window has at most one point of “discontinuity”, that is, it contains at most once two consecutive coordinates which are not successive images by |$T$| of some |$x_k$|⁠.

Finally, by the construction of |$\nu$|⁠, the map |$x\mapsto (x,Tx,T^2x,\ldots)$| is clearly an isomorphism between |$(X,\mu,T)$| and |$(Y,\nu,S)$|⁠. ■

 
Proof of Theorem 3

Apply the conclusion of Theorem 2 to the specific uniquely ergodic model |$(Y,S)$| which is constructed above, starting from the point |$\underline{y}$| and taking an arbitrary continuous function of the first coordinate. ■

3 Asymptotic Orthogonality of Powers for Quasi-discrete Spectrum Automorphisms

3.1 The key lemma

We say that |$c\in \mathbb{S}^1$| is irrational if |$c^n\neq1$| for each integer |$n\neq0$|⁠. Recall that an ergodic automorphism |$T$| is totally ergodic if and only if all its eigenvalues except 1 are irrational.

 
Lemma 1.

Let |$c_1$| and |$c_2$| be irrational in |$\mathbb{S}^1$|⁠. Then there exists at most one pair |$(r,s)$| of mutually prime natural numbers such that |$c_1^r=c_2^s$|⁠. □

 
Proof

Let |$(r,s)$| and |$(r',s')$| be two pairs of mutually prime natural numbers satisfying |$c_1^r=c_2^s$|⁠, and |$c_1^{r'}= c_2^{r'}$|⁠. Let |$d\in \mathbb{S}^1$| be such that |$d^s=c_1$|⁠. We have |$(d^r)^s=c_1^r=c_2^s$|⁠. Multiplying if necessary |$d$| by an appropriate |$s$|-th root of the unity, and using the fact that |$r$| and |$s$| are mutually prime, we can assume that we also have |$d^r=c_2$|⁠. The equality |$c_1^{r'}= c_2^{r'}$| now yields |$d^{r's}=d^{s'r}$|⁠. But |$d$| is also irrational, hence |$r's=sr'$|⁠. Since |$(r,s)=1=(r',s')$|⁠, |$r=r'$| and |$s=s'$|⁠. ■

3.2 AOP for discrete spectrum

Recall that discrete spectrum automorphisms are characterized by the fact that their eigenfunctions form a linearly dense subset of |$L^2$|⁠. They fall into the more general case of quasi-discrete spectrum ones, but we nevertheless include a direct proof that they have AOP as soon as they are totally ergodic, since this will introduce the main ideas in a simpler context.

Assume that |$T$| is a totally ergodic and discrete spectrum automorphism of |$(X,{\cal B},\mu)$|⁠. Let |$c_1=1, c_2,c_3,\ldots\in \mathbb{S}^1$| be its eigenvalues, and let |$f_1, f_2, \ldots$| be corresponding eigenfunctions: |$f_i\circ T=c_i\cdot f_i$|⁠, |$|f_i|=1$|⁠. Note that |$c_i$| is irrational and |$\int f_i\,{\rm d}\mu=0$| whenever |$i\ge2$|⁠.

Let us prove that the AOP property holds for |$T$|⁠. Since the linear subspace spanned by |$f_2,f_3,\ldots$| is dense in the subspace of square integrable functions with zero integral, it is enough to show that, for any fixed |$i,j\ge 2$|⁠,
fifj¯dρ=0,ρJe(Tr,Ts), if r and s are large enough and (r,s)=1.
But we have
where graphic is an eigenfunction f or |$(T^r\times T^s,\rho)$| corresponding to the eigenvalue |$c_i^{r}$|⁠, while graphic is an eigenfunction for |$(T^r\times T^s,\rho)$| corresponding to |$c_j^s$|⁠. But, if these two eigenvalues are different, the functions graphic and graphic are orthogonal. It is then enough to observe that, by Lemma 1, the equality |$c_i^r=c_j^s$| has at most one solution with |$(r,s)=1$|⁠.
 
Example 1.

If we consider |$T$| as an ergodic rotation whose group of eigenvalues is of the form |$\{e^{2\pi iq\alpha}:\: q\in\mathbb{Q}\}$| and |$\alpha\in[0,1)$| is irrational, then by the Halmos--von Neumann theorem (e.g., [13]) it follows that all non-zero powers of |$T$| are isomorphic. This is an extremal example of an automorphism with non-disjoint powers for which the AOP property holds. □

3.3 Quasi-discrete spectrum

We denote by |$M=M(X,{\cal B},\mu)$| the multiplicative group of functions |$f\in L^2(X,{\cal B},\mu)$| satisfying |$|f|=1$|⁠. Let |$T$| be an ergodic automorphism of |$(X,{\cal B},\mu)$|⁠. We define on |$M$| the group homomorphism |$W_T$| by setting
WT(f):=fT/f.

Note that |$(W_Tf)\circ T=W_T (f\circ T)$|⁠.

Let |$E_0(T)\subset\mathbb{S}^1$| denote the group of eigenvalues of |$T$|⁠, which we also consider as a subgroup of (constant) functions in |$M$|⁠. Then, for each integer |$k \geq1$|⁠, we inductively set
Ek(T):={fM:WTfEk1(T)}.
It is easily proved by induction on |$k$| that |$E_k(T)$| is the subgroup of |$M$| constituted of all |$f\in M$| satisfying |$W_T^k f\in E_0(T)$|⁠, and that, by the ergodicity of |$T$|⁠,
fEk(T)Ek1(T)WTkfE0(T){1}.

The elements of |$\bigcup_{k\geq0}E_k(T)$| are called quasi-eigenfunctions. Clearly, |$E_k(T)\subset E_{k+1}(T)$|⁠, |$k\geq0$|⁠.

We denote by |$E_0^{\text{irr}}(T)$| the set of irrational eigenvalues of |$T$|⁠, and for each |$k\ge1$|⁠, we set
Ekirr(T):={fEk(T):WTkfE0irr(T)}.
 
Lemma 2.
Let |$T$| be an ergodic automorphism of |$(X,{\cal B},\mu)$|⁠, |$k\ge1$| and |$f\in E_k^\text{irr}(T)$|⁠. Then
Xfdμ=0.
 
Proof
We prove the result by induction on |$k$|⁠. For |$k=1$|⁠, the conclusion holds for each eigenfunction |$f$| whose eigenvalue is different from 1. Assume that the result is proved for |$k\ge1$|⁠, and consider |$f\in E_{k+1}^\text{irr}(T)$|⁠. Set |$g:=W_Tf\in E_k^\text{irr}(T)$|⁠, so that |$f\circ T=gf$|⁠. Then, for each |$n\ge1$|⁠, we have
fTn=g(n)f,
where
g(n):=ggTgTn1Ek(T).
It follows that
Xf(fTn)¯dμ=Xg(n)dμ¯.
(8)
Observe now that
WTkg(n)=WTkgWTk(gT)WTk(gTn1)=(WTkg)n=(WTk+1f)n,
which is an irrational eigenvalue. We therefore have |$g^{(n)}\in E_k^\text{irr}(T)$|⁠, and by induction hypothesis,
Xg(n)dμ=0.

Coming back to (8), we see that the spectral measure of |$f$| is the Lebesgue measure on the circle, which implies the result. ■

 
Lemma 3.
Let |$T$| be an ergodic automorphism of |$(X,{\cal B},\mu)$|⁠, |$k\ge1$|⁠, and |$f\in E_k(T)$|⁠. Then for each integer |$r\ge1$|⁠,
WTrkf=(WTkf)rk.
 
Proof
Again, we prove the result by induction on |$k$|⁠. For |$k=1$|⁠, take |$f\in E_1(T)$| and let |$c:=W_T f$| be the corresponding eigenvalue (for |$T$|⁠). Then
WTrf=fTr/f=cr=(WTf)r.
Now, assume the result is proved for |$k\ge1$|⁠, and take |$f\in E_{k+1}(T)$|⁠. Note that |$W_{T^r f}\in E_k(T)$|⁠, so that, using the induction hypothesis, we have
WTrk+1f=WTrk(WTrf)=(WTkWTrf)rk=(WTkfTr/WTkf)rk.

But |$W_T^k f$| is an eigenfunction of |$T$|⁠. Denoting by |$c:=W_T^{k+1}f$| the corresponding eigenvalue, the right-hand side of the above equality becomes |$(c^r)^{r^k}=c^{r^{k+1}}$|⁠. ■

 
Proposition 3.
Let |$T$| be a totally ergodic automorphism of |$(X,{\cal B},\mu)$|⁠, |$k\ge1$|⁠, |$f\in E_k(T)\setminus E_{k-1}(T)$| and |$g\in E_k(T)$|⁠. Then there exists at most one pair |$(r,s)$| of mutually prime natural numbers for which we can find |$\rho\in J^e(T^r,T^s)$| satisfying
X×Xfg¯dρ0.
 
Proof
Assume that |$r$|⁠, |$s$|⁠, and |$\rho$| are as in the statement of the proposition. In the ergodic dynamical system |$(X\times X, \rho, T^r\times T^s)$|⁠, we have
WTr×Tsk(fg¯)=WTrkfWTskg¯,
where both |$W_{T^r}^k f$| and |$W_{T^s}^k \overline{g}$| are eigenvalues of |$T$| by Lemma 3. This product is therefore constant, let us denote it by |$c$|⁠. It is an eigenvalue of |$T^r\times T^s$|⁠, and it is also the product of two eigenvalues of |$T$|⁠, hence it is also an eigenvalue of |$T$|⁠. This proves that |$f\otimes \overline{g}\in E_k(T^r\times T^s)$|⁠. Since we assume that |$\int f\otimes \overline g\, {\rm d}\rho \neq 0$|⁠, Lemma 2 yields that |$W^k_{T^r\times T^s} f\otimes \overline g$| is not an irrational eigenvalue of |$T^r\times T^s$|⁠. But, by the total ergodicity of |$T$|⁠, the only eigenvalue of |$T$| which is not irrational is 1. Using again Lemma 3, we get
c=1=WTr×Tsk(fg¯)=(WTkf)rk(WTkg¯)sk.

Remember that |$W_T^k f$| is irrational because |$f\in E_k(T)\setminus E_{k-1}(T)$|⁠, and |$T$| is totally ergodic. It follows that |$W_T^k g$| is also irrational, hence |$g\in E_k(T)\setminus E_{k-1}(T)$|⁠. Then |$W_T^k f$| and |$W_T^k \overline g$| are both eigenvalues of |$T$| which are different from 1, hence they are both irrational. We only need now to apply Lemma 1 to conclude the proof. ■

 
Definition 2.
|$T$| is said to have quasi-discrete spectrum if:
T is totally ergodic,
(9)
and
L2(X,B,μ)=span¯(k0Ek(T)).
(10)
 
Theorem 6.

Let |$T$| be an automorphism of |$(X,{\cal B},\mu)$| with quasi-discrete spectrum. Then it has the AOP property. □

 
Proof
By the definition of quasi-discrete spectrum, it is enough to check that, for any |$k\ge1$| and any |$f,g\in E_k(T)$|⁠, we have, for |$r,s$| mutually prime and large enough and for each |$\rho\in J^e(T^r,T^s)$|⁠,
X×Xfg¯dρ=0.

But this immediately follows from Proposition 3. ■

4 Application to Some Algebraic Models of Discrete and Quasi-discrete Spectrum Transformations

4.1 Irrational rotation

Fix some irrational number |$\alpha$|⁠, and consider the transformation |$R_\alpha: x\mapsto x+\alpha$| on |${\mathbb{T}}$|⁠. Then |$({\mathbb{T}},R_\alpha)$| is uniquely ergodic, and the corresponding measure-theoretic dynamical system is totally ergodic and has discrete spectrum. It therefore has AOP, and Theorem 3 applies in this case (we consider |$f(x):=e^{2\pi i x}$|⁠). Then, for any multiplicative function |$\boldsymbol{\nu}$|⁠, |$|\boldsymbol{\nu}|\le1$|⁠, any sequence |$(x_k)$| of points in |${\mathbb{T}}$| and any increasing sequence |$1=b_1<b_2<\cdots$| with |$b_{k+1}-b_k\to\infty$|⁠, we have
limK1bK+11kKe2πixkbkn<bk+1e2πinαν(n)=0.
Now, for each |$k\ge1$|⁠, we can choose |$x_k\in{\mathbb{T}}$| such that
e2πixkbkn<bk+1e2πinαν(n)=|bkn<bk+1e2πinαν(n)|.

This proves Theorem 4 in the case of a polynomial of degree 1.

4.2 Affine transformations of the |$d$|-dimensional torus

To prove Theorem 4 in its full form, we generalize the preceding case by considering some transformation of the |$d$|-dimensional torus of the form
T:(x1,,xd)(x1+α,x2+x1,,xd+xd1).
This transformation is an affine transformation, it can be written as |$x\mapsto Ax+b$| where |$A=[a_{ij}]_{i,j=1}^d$| is the matrix defined by |$a_1:=1, a_{i-1,i}=a_{ii}:=1$| and all other coefficients equal to zero, and |$b:=(\alpha,0,\ldots,0)$|⁠. Taking again |$\alpha$| irrational, |$({\mathbb{T}}^d,T)$| is a uniquely ergodic dynamical system, and it is totally ergodic with respect to the Haar measure on |${\mathbb{T}}^d$|⁠, which is the unique invariant measure [12]. Moreover, the corresponding measure-theoretic dynamical system has quasi-discrete spectrum [1]. Hence it has AOP, and Theorem 3 applies again. In particular, we will use this theorem with the function |$f(x_1,\ldots,x_d):=e^{2\pi i x_d}$|⁠, observing as in [7,12] that the last coordinate of |$T^n(x_1,\ldots,x_d)$| is the following polynomial in |$n$|⁠:
(nd)α+(nd1)x1++nxd1+xd.
 
Proof of Theorem 4
Consider a fixed polynomial |$P\in{\mathbb{R}}[x]$|⁠, whose leading coefficient is irrational. Then we can always choose |$\alpha, x_1,\ldots,x_d$| so that
(nd)α+(nd1)x1++nxd1+xd=P(n)for all n.
Let |$1=b_1<b_2<\cdots$| be a sequence of integers satisfying |$b_{k+1}-b_k\to\infty$|⁠, and define a sequence |$(x^k)$| of points in |${\mathbb{T}}^d$| by setting
xk:=(x1,x2,,xd1,xd+tk),
where |$(t_k)$| is a sequence of points in |${\mathbb{T}}$| to be precised later. Then, Theorem 3 applied to |$T$| gives
limK1bK+11kKe2πitkbkn<bk+1e2πiP(n)ν(n)=0.
Now, for each |$k$| we can choose |$t_k\in{\mathbb{T}}$| such that
e2πitkbkn<bk+1e2πiP(n)ν(n)=|bkn<bk+1e2πiP(n)ν(n)|,
and this concludes the proof. ■
 
Proof of Theorem 5
Assume that the result is false. Then we can find a non constant |$P\in{\mathbb{R}}[x]$| with irrational leading coefficient, and two sequences |$(M_\ell)$| and |$(H_\ell)$| with |$H_\ell\to\infty$| and |$H_\ell/M_\ell\to 0$|⁠, satisfying for some |$\varepsilon>0$|⁠:
,1MMm<2M1H|mn<m+He2πiP(n)ν(n)|>ε.
(11)
By passing to a subsequence if necessary, we can also assume that for each |$\ell$|⁠, |$M_{\ell+1}>2M_\ell+H_\ell$|⁠. Rewriting the left-hand side of the above inequality as
1H0r<H1M/HMm<2Mm=r   modH1H|mn<m+He2πiP(n)ν(n)|,
we see that for each |$\ell$| there exists |$0\le r_\ell <H_\ell$| such that
1M/HMm<2Mm=rmodH1H|mn<m+He2πiP(n)ν(n)|>ε.
(12)
Now, set
{1=b1<b2<}:={1}{m:Mm<2M+H and m=rmodH},
and let |$K_\ell$| be the largest |$k$| such that |$b_k<2M_\ell$|⁠. Observing that |$b_{K_\ell+1}/(2M_\ell)\to1$|⁠, and using inequality (12), we get
liminf1bK+1kK|bk+1n<bk+1e2πiP(n)ν(n)|liminf12MMm<2Mm=r     mod  H|mn<m+He2πiP(n)ν(n)|ε/2,
which contradicts Theorem 4. ■

4.3 Algebraic models of general quasi-discrete spectrum transformations

Assume that |$T$| is a quasi-discrete spectrum automorphism of |$(X,{\cal B},\mu)$|⁠. Then, in view of [1, 16, 17], up to measure theoretic isomorphism, we may assume that:

  • (i) |$X$| is a compact, connected, Abelian group.

  • (ii) |$Tx=Ax+b$|⁠, where |$A:X\to X$| is a continuous group automorphism, |$b\in X$|⁠, and the group generated by |$(A-I)X$| and |$b$| is dense in |$X$|⁠.

  • (iii) For each character |$\chi\in\widehat{X}$|⁠, there exists |$m\geq 0$| such that |$\chi\circ (A-I)^m=1$|⁠. (Note that |$\chi(Tx)=\chi(x)\big(\chi((A-I)x)\chi(b)\big)$|⁠, so by an easy induction, we obtain that |$\chi\in E_m(T)$| if |$m$| is the smallest such that |$\chi\circ(A-I)^m=1$|⁠.)

By (iii), |$\bigcap_{n\geq0}(A-I)^nX=\{0\}$|⁠, that is, |$T$| is unipotent, and by [19, 20], it follows that |$T$| is minimal. Then, in view of [16], |$(T,X)$| is uniquely ergodic.

 
Theorem 7.
Assume that |$T$| satisfies (i), (ii), and (iii). Let |$\boldsymbol{\nu}:{\mathbb{N}}\to {\mathbb{C}}$| be any multiplicative function, with |$|\boldsymbol{\nu}|\le1$|⁠. Then, for each |$1\neq\chi\in \widehat{X}$| and each increasing sequence |$1=b_1<b_2<\ldots$| with |$b_{k+1}-b_k\to\infty$| and each choice of |$(z_k)\in X$|⁠, we have
1bK+1kK|bk+1n<bk+1χ(Tnzk)ν(n)|0 when K.
 
Proof

We have already noticed that |$(X,T)$| is uniquely ergodic. Fix a character |$1\neq \chi\in\widehat{X}$|⁠. Using (iii), let |$m\geq 0$| be the smallest such that |$\chi\circ (A-I)^m=1$|⁠. Since |$\chi\neq1$|⁠, we have |$m\geq1$|⁠. Let |$Y_m:=(A-I)^{m-1}X$|⁠. Note that:

  • |$Y_m$| is a non-trivial subgroup of |$X$| (indeed, otherwise, |$\chi\circ(A-I)^{m-1}=1$|⁠);

  • |$Y_m$| is connected (indeed, |$(A-I)^{m-1}$| is a continuous group homomorphism and, by (i), |$Y_m$| is the image of a connected space);

  • |$\chi(Y_m)=\mathbb{S}^1$| (by the previous observation).

Note that, for each |$y\in Y_m$|⁠, we have |$\chi\circ A(y)=\chi(y)$|⁠, and since |$AY_m\subset Y_m$|⁠, we also have |$\chi\circ A^n(y)=\chi(y)$| for each |$n\geq1$|⁠. Moreover, for each |$y_k\in Y_m$|⁠, we have
χ(Tn(zk+yk))=χ(Anyk)χ(Tnzk)=χ(yk)χ(Tnzk).
(13)
It follows by the above that, for each |$k\geq1$|⁠, we can find |$y_k\in Y_m$| such that
bk+1n<bk+1χ(Tn(zk+yk))ν(n)=χ(yk)bk+1n<bk+1χ(Tnzk)ν(n)=|bk+1n<bk+1χ(Tnzk)ν(n)|
and the result follows by the AOP property of the system, and Theorem 3. ■

5 Remarks on the AOP Property

 
Remark 1.

The class of measure-theoretic dynamical systems with AOP is stable under taking factors, since any ergodic joining of the actions of |$T^r$| and |$T^s$| on a factor |$\sigma$|-algebra extends to an ergodic joining of |$T^r$| and |$T^s$|⁠. It is also close by inverse limits, since it is enough to check (2) for a dense subclass of functions |$f$| in |$L^2$|⁠. □

 
Proposition 4.

If |$T$| satisfies (2) then |$T$| is of zero entropy. □

 
Proof

Because of Remark 1, it is enough to show that there exists a Bernoulli automorphism with arbitrarily small entropy for which (2) fails.

Fix |$\delta>0$| and let |$T$| be the shift on |$X=\{0,1\}^{{\mathbb{Z}}}$| considered with Bernoulli measure |$\mu=(\delta,1-\delta)^{\otimes{\mathbb{Z}}}$|⁠. Given |$r\geq1$|⁠, consider the map |$\varphi_r:X\to X$|⁠,
φr((xi)iZ):=(xri)iZ
for each |$(x_i)_{i\in{\mathbb{Z}}}\in X$|⁠. Then, |$\varphi_r\circ T^r=T\circ\varphi_r$| and |$\varphi_r$| yields a homomorphism (factor map) from |$(X,\mu,T^r)$| on to |$(X,\mu,T)$| (because the laws of the independent processes |$(X_i)_{i\in {\mathbb{Z}}}$|⁠, |$X_i((x_j)_{j\in{\mathbb{Z}}})=x_i$| and |$(X_{ri})_{i\in{\mathbb{Z}}}$| are the same). Moreover, the extension
φr:(X,μ,Tr)(X,μ,T) is relatively Bernoulli
(14)
(see e.g., [13]) since |$\varphi_r^{-1}({\cal B})$| has the complementary independent sub-|$\sigma$|-algebra given by the smallest |$T^r$| invariant sub-|$\sigma$|-algebra making the variables |$X_1,\ldots,X_{r-1}$|-measurable. Take now |$r\neq s\geq1$| to obtain the joining |$\kappa_{r,s}$| defined as
ghdκr,s:=E(g|φr1(B))(φr1(x))E(h|φs1(B))(φs1(x))dμ(x).
(15)
In view of (14) is ergodic. Now, consider |$f\in L^2(X,{\cal B},\mu)$| defined by |$f((x_i)_{i\in Z}):=x_0$|⁠. We claim that
E(f|φr1(B))(φr1(x))=f(x)=x0.
(16)
Indeed, given |$x=(x_i)_{i\in{\mathbb{Z}}}\in X$|⁠,
φr1(x)={yX:yir=xifor eachiZ}
and on that fiber we have a relevant Bernoulli measure as the conditional measure. Note however that |$f$| is constant on the fiber, in other words, |$f$| is measurable with respect to |$\varphi_r^{-1}({\cal B})$|⁠. Since (16) holds also for |$\varphi_s$|⁠, using additionally (15), it follows that
ff¯dκr,s=fL2(X,B,μ)2>0,
which excludes the possibility of satisfying (2). ■
 
Remark 2.
There exist some zero-entropy automorphisms for which the AOP property fails. Indeed, we will show that it does not hold for horocycle flows. We have the classical relation between the geodesic and the horocycle flow: for each |$s,u\in{\mathbb{R}}$|⁠, we have
guhsgu1=he2us.

Set |$r=e^{-2u}s$|⁠. Assume now that |$r\neq s$| are coprime natural numbers, |$r,s\to \infty$| but we also assume that |$r/s\to 1$|⁠. For such pair |$(r,s)$|⁠, the number |$u$| is determined by |$e^{-2u}=r/s$|⁠, so |$u$| is close to zero. We can hence assume that |$g_u$|⁠, whose graph |$\Delta_{g_u}$| is an ergodic joining of |$h_r=(h_1)^r$| and |$h_s=(h_1)^s$|⁠, belongs to the compact set of joinings |$\{\Delta_{g_t}:t\in[0,1]\}$|⁠. Since a graph joining is clearly never equal to the product measure, there is a positive distance of |$\{\Delta_{g_t}:\:t\in[0,1]\}$| from the product measure in the relevant space of couplings. Therefore, (2) fails to hold. □

Funding

This research was supported by Narodowe Centrum Nauki grant UMO-2014/15/B/ST1/03736.

References

[1]

Abramov
L. M.
“Metric automorphisms with quasi-discrete spectrum.
Izvestiya Akademii Nauk SSSR. Seriya Matematicheskaya
26
(
1962
):
513
30
.

[2]

Arnoux
P.
“Sturmian Sequences.”
In
Substitutions in Dynamics, Arithmetics and Combinatorics. Lecture Notes in Mathematics,
edited by
Berthé
V.
Ferenczi
S.
Mauduit
C.
and
Siegel
A.
vol.
1794.
Berlin:
Springer,
2002,
pp.
143
98
.

[3]

Bourgain
J.
“Möbius-Walsh correlation bounds and an estimate of Mauduit and Rivat.
Journal d’Analyse Mathématique
119
(
2013
):
147
63
.

[4]

Bourgain
J.
“On the correlation of the Moebius function with rank-one systems.
Journal d’Analyse Mathématique
120
(
2013
):
105
30
.

[5]

Bourgain
J.
,
Sarnak
P.
and
Ziegler.
T.
“Disjointness of Moebius from Horocycle Flows.”
In
From Fourier Analysis and Number Theory to Radon Transforms and Geometry.
Developmental Mathematics,
edited by
Farkas
H. M.
,
Gunning
R. C.
,
Knopp
M. I.
and
Taylor
B. A.
, vol.
28
.
New York
:
Springer
,
2013
, pp.
67
83
.

[6]

Downarowicz
T.
, and
Kasjan.
S
“Odometers and Toeplitz systems revisited in the context of Sarnak’s conjecture.
Studia Mathematica
229
, no.
1
(
2015
):
45
72
.

[7]

Einsiedler
M.
, and
Ward.
T
“Ergodic Theory with A View Towards Number Theory.”
In
Graduate Texts in Mathematics
. vol.
259
.
London
:
Springer,
2011
.

[8]

el Abdalaoui
H.
,
Kasjan
S.
, and
Lemaúczyk
M.
.
“0–1 sequences of the Thue-Morse type and Sarnak’s conjecture.
Proceedings of the American Mathematical Society
144
, no.
1
(
2016
):
161
76
.

[9]

el Abdalaoui
H.
,
Kułaga-Przymus
J.
,
Lemańczyk
M.
, and
de la Rue
T.
The Chowla and the Sarnak conjectures from ergodic theory point of view
,
hal-01070531, preprint, to appear in Discrete and Continuous Dynamical Systems
,
2015
(accessed June 20, 2016)
.

[10]

el Abdalaoui
H.
,
Lemańczyk
M.
, and
de la Rue
T.
“On spectral disjointness of powers for rank-one transformations and Möbius orthogonality.
Journal of Functional Analysis
266
, no.
1
(
2014
):
284
317
.

[11]

Ferenczi
S.
, and
Mauduit
C.
On Sarnak’s conjecture and Veech’s question for interval exchanges, to appear in Journal d’Analyse Mathématique
.

[12]

Furstenberg
H.
“Strict ergodicity and transformation of the torus.
American Journal of Mathematics
83
(
1961
):
573
601
.

[13]

Glasner
E.
Ergodic Theory via Joinings. Mathematical Surveys and Monographs
, vol.
101
.
Providence, RI
:
American Mathematical Society
,
2003
.

[14]

Green
B.
“On (not) computing the Möbius function using bounded depth circuits.
Combinatorics, Probability and Computing
21
, no.
6
(
2012
):
942
51
.

[15]

Green
B.
, and
Tao
T
“The Möbius function is strongly orthogonal to nilsequences.
Annals of Mathematics (2)
175
, no.
2
(
2012
):
541
66
.

[16]

Hahn
F.
, and
Parry
W
“Minimal dynamical systems with quasi-discrete spectrum.
Journal of the London Mathematical Society
40
(
1965
):
309
23
.

[17]

Hahn
F.
, and
Parry
W
“Some characteristic properties of dynamical systems with quasi-discrete spectra.
Mathematical Systems Theory. An International Journal on Mathematical Computing Theory
2
(
1968
):
179
90
.

[18]

Harper
A. J.
A different proof of a finite version of Vinogradov’s bilinear sum inequality
,
preprint
(
2011
), https://www.dpmms.cam.ac.uk/~ajh228/FiniteBilinearNotes.pdf(
accessed June 20
,
2016
).

[19]

Hoare
H.
, and
Parry.
W
“Affine transformations with quasi-discrete spectrum. I.
Journal of the London Mathematical Society
41
(
1966
):
88
96
.

[20]

Hoare
H.
, and
Parry
W
“Affine transformations with quasi-discrete spectrum. II.
Journal of the London Mathematical Society
41
(
1966
):
529
30
.

[21]

Karagulyan
D.
“On Möbius orthogonality for interval maps of zero entropy and orientation-preserving circle homeomorphisms.
Arkiv för Matematik
53
, no.
2
(
2015
):
317
27
.

[22]

Kátai
I.
“A remark on a theorem of H. Daboussi.
Acta Mathematica Hungarica
47
, no.
1-2
(
1986
):
223
25
.

[23]

Lehrer
E.
“Topological mixing and uniquely ergodic systems.
Israel Journal of Mathematics
57
, no.
2
(
1987
):
239
55
.

[24]

Liu
J.
, and
Sarnak
P.
,
“The Möbius function and distal flows.
Duke Mathematical Journal
164
, no.
7
(
2015
):
1353
99
.

[25]

Matomäki
K.
,
Radziwiłl
M.
, and
Tao
T
“An averaged form of Chowla’s conjecture.
Algebra Number Theory
9
, no.
9
(
2015
):
2167
96
.

[26]

Mauduit
C.
, and
Rivat
J
“Sur un problème de Gelfond: la somme des chiffres des nombres premiers.
Annals of Mathematics (2)
171
, no.
3
(
2010
):
1591
46
.

[27]

Mauduit
C.
, and
Rivat.
J
“Prime numbers along Rudin-Shapiro sequences.
Journal of the European Mathematical Society (JEMS)
17
, no.
10
(
2015
):
2595
642
.

[28]

Peckner
R.
“Two dynamical perspectives on the randomness of the Mobius function.”
ProQuest LLC, Ann Arbor, MI. Thesis PhD
,
Princeton University
,
2015
.

[29]

Sarnak
P.
Three lectures on the Möbius function, randomness and dynamics
. http://publications.ias.edu/sarnak/
2010
(accessed June 20, 2016)
.

[30]

Veech
W. A.
“Möbius orthogonality for generalized Morse-Kakutani flows”.
preprint
.