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

On normal numbers and self-similar measures

Amir Algom, Simon Baker, and Pablo Shmerkin
Abstract.

Let {fi(x)=six+ti}\{f_{i}(x)=s_{i}\cdot x+t_{i}\} be a self-similar IFS on \mathbb{R} and let β>1\beta>1 be a Pisot number. We prove that if log|si|logβ\frac{\log|s_{i}|}{\log\beta}\notin\mathbb{Q} for some ii then for every C1C^{1} diffeomorphism gg and every non-atomic self similar measure μ\mu, the measure gμg\mu is supported on numbers that are normal in base β\beta.

P.S. was supported by an NSERC discovery grant

1. Introduction

1.1. Background and main result

Let bb be an integer greater or equal to 22. Let TbT_{b} be the times-bb map,

Tb(x)=bxmod1,x.T_{b}(x)=b\cdot x\mod 1,\,x\in\mathbb{R}.

A number xx\in\mathbb{R} is called bb-normal, or normal in base bb, if its orbit {Tbk(x)}k\{T_{b}^{k}(x)\}_{k\in\mathbb{N}} equidistributes for the Lebesgue measure on [0,1][0,1]. In 1909 Borel proved that Lebesgue almost every xx is absolutely normal, that is, normal in all bases. In the absence of obvious obstructions, it is believed that this phenomenon should continue to hold true for typical elements of well structured sets with respect to appropriate measures. The purpose of this paper is to make a contribution in this direction for a well studied class of fractal measures.

Before stating our main result, we recall that one may define digit expansions of numbers in non-integer bases as well: Following Rényi [25], for β>1\beta>1 we define the β\beta-expansion of x(0,1)x\in(0,1) to be the lexicographically largest sequence xn{0,,[β]}x_{n}\in\{0,...,[\beta]\} such that x=n=1xnβnx=\sum_{n=1}^{\infty}x_{n}\beta^{-n}. This sequence is obtained from the orbit of xx under the map Tβ(x)=βxmod1T_{\beta}(x)=\beta\cdot x\mod 1 similarly to the integer case. It is known that TβT_{\beta} admits a unique absolutely continuous invariant measure, commonly referred to as the Parry measure [20]. When β\beta is an integer then the Parry measure is just the Lebesgue measure restricted to [0,1)[0,1). Thus, in analogy with the integer case, we say that xx is β\beta-normal if it equidistributes under TβT_{\beta} for the Parry measure. Our main result will allow us to conclude that for certain fractal measures, a typical element will be β\beta-normal, where β>1\beta>1 may be a non-integer.

In this paper we will work with self-similar measures on \mathbb{R}. These are defined as follows: Let Φ={f1,,fn}\Phi=\{f_{1},...,f_{n}\} be a finite set of real non-singular contracting similarity maps of a compact interval JJ\subset\mathbb{R}. That is, for every ii we can write fi(x)=six+tif_{i}(x)=s_{i}\cdot x+t_{i} where si(1,1){0}s_{i}\in(-1,1)\setminus\{0\} and tit_{i}\in\mathbb{R}, and fi(J)Jf_{i}(J)\subseteq J. We will refer to Φ\Phi as a self similar IFS (Iterated Function System). It is well known that there exists a unique compact set K=KΦJ\emptyset\neq K=K_{\Phi}\subseteq J such that

K=i=1nfi(K).K=\bigcup_{i=1}^{n}f_{i}(K).

The set KK is called a self-similar set, and the attractor of the IFS Φ\Phi. We always assume that there exist iji\neq j such that the fixed point of fif_{i} is not equal to the fixed point of fjf_{j}. This ensures that KK is infinite.

Next, let p=(p1,,pn)\textbf{p}=(p_{1},...,p_{n}) be a strictly positive probability vector, that is, pi>0p_{i}>0 for all ii and ipi=1\sum_{i}p_{i}=1. It is well known that there exists a unique Borel probability measure μ\mu such that

μ=i=1npifiμ, where fiμ is the push-forward of μ via fi.\mu=\sum_{i=1}^{n}p_{i}\cdot f_{i}\mu,\quad\text{ where }f_{i}\mu\text{ is the push-forward of }\mu\text{ via }f_{i}.

The measure μ\mu is called a self-similar measure, and is supported on KK. Our assumptions that KK is infinite and that pi>0p_{i}>0 for every ii are known to imply that μ\mu is non-atomic. In particular, all self-similar measures in this paper are non-atomic.

Recall that β>1\beta>1 is called a Pisot number if it is an algebraic integer whose algebraic conjugates are all of modulus strictly less than 11. Note that every integer larger than 11 is a Pisot number. Also, we will write aba\sim b if the real numbers a,b0a,b\neq 0 satisfy that log|a|log|b|\frac{\log|a|}{\log|b|}\in\mathbb{Q}. Otherwise, we write a≁ba\not\sim b, in which case a,ba,b are said to be multiplicatively independent. Finally, given a measure μ\mu on \mathbb{R} and β>1,\beta>1, we say that μ\mu is pointwise β\beta-normal if μ\mu almost every xx is β\beta-normal. We are now ready to state the main result of this paper:

Theorem 1.1.

Let β>1\beta>1 be a Pisot number, and let μ\mu be a non-atomic self-similar measure with respect to the self-similar IFS {fi(x)=six+ti}\{f_{i}(x)=s_{i}\cdot x+t_{i}\} and a strictly positive probability vector. If there is some ii such that si≁βs_{i}\not\sim\beta then μ\mu is pointwise β\beta-normal. Furthermore, gμg\mu is pointwise β\beta-normal for all gdiff1()g\in\text{diff}^{1}(\mathbb{R}).

We emphasize that no separation condition is imposed on the underlying IFS. Theorem 1.1 is sharp in the sense that there are many IFSs such that si=ns_{i}=n for all ii and some integer nn, and no element in the attractor is nn-normal - for example, no element in the middle-thirds Cantor set is 33-normal. Nonetheless, at least for integer β\beta, we will soon recall some recent results showing that even if all the contractions siβs_{i}\sim\beta it is still possible that every self-similar measure is pointwise β\beta-normal. We also note that there are other ways self-similar measures can lead to equidistribution - see e.g. [3]

Theorem 1.1 extends a long line of research on pointwise normality for dynamically defined measures, and provides a unified proof for many recent results regarding self similar measures. The first instances of a special case of Theorem 1.1 were obtained by Cassels [5] and Schmidt [28] around 1960. They considered the Cantor-Lebesgue measure on the middle thirds Cantor set, proving that almost every point is normal to base nn as long as n≁3n\not\sim 3. This was later generalized by Feldman and Smorodinsky [8] to all non-degenerate Cantor-Lebesgue measures with respect to any base bb. We remark that the work of Host [12] and the subsequent works of Meiri [18] and Lindenstrauss [17] about pointwise normality for TnT_{n}-invariant measures are also related to this, though in the somewhat different context of Furstenberg’s ×2,×3\times 2,\times 3 Conjecture.

In 2015, Hochman and Shmerkin [11, Theorem 1.4] proved Theorem 1.1 under the additional assumptions that si>0s_{i}>0 for all ii, and the intervals fi(J)f_{i}(J), iJi\in J are disjoint except potentially at their endpoints (a condition stronger than the open set condition). In fact, [11, Theorem 1.2] provides a general criterion for a uniformly scaling measure (that is, a measure such that at almost every point its scenery flow equidistributes for the same distribution) to be pointwise β\beta-normal. We refer the reader to Section §1.2 for the relevant definitions about the scenery flow. The main reason they imposed this separation condition was to ensure that the corresponding self-similar measures are indeed uniformly scaling. This fact had previously been established by Hochman [10, Section 4.3]. Recently, Pyörälä [23] proved that it suffices to assume that the IFS satisfies the weak separation condition [15, 32] in order for self similar measures to be uniformly scaling. Utilizing this result and the methods from [11], Pyörälä obtained a version of Theorem 1.1 under the additional assumption that the IFS satisfies the weak separation condition [23, Corollary 5.4]. We note that there are many self-similar sets which satisfy neither the open set nor the weak separation condition; indeed, this is often the generic behaviour in parametrized families of self-similar sets: see [22].

Very recently, Algom, Rodriguez Hertz, and Wang, proved that if the Fourier transform of a self similar measure decays to 0 at \infty then almost every point is absolutely normal [2, Theorem 1.4]. Combining this with recent progress on the Fourier decay problem for self similar measures leads to many special cases of Theorem 1.1: For example, by a result of Li and Sahlsten [16], if there are i,ji,j such that si≁sjs_{i}\not\sim s_{j} then all self similar measures have this property, and consequently are supported on absolutely normal numbers. Furthermore, via [2, Theorem 1.4] and the recent work of Brémont [4] one obtains many instances of self similar measures that are supported on nn-normal numbers even though the underlying IFS satisfies sins_{i}\sim n for all ii. Examples of this form were also recently obtained by Dayan, Ganguly, and Weiss [7].

The results of [2] are related to a classical theorem of Davenport-Erdős-LeVeque [6]. This result states that if the Fourier transform of a Borel measure decays to zero sufficiently quickly, then a typical number with respect to this measure will be normal (for integer bases). Recall that by [2, Theorem 1.4], this property holds for self-similar measures with a decaying Fourier transform, regardless of the rate of decay. In [19] it was shown that for any homogeneous IFS with non-atomic self-similar measure μ\mu, if gg is a C2C^{2} function satisfying g′′>0g^{\prime\prime}>0 then the Fourier transform of gμg\mu decays to zero sufficiently quickly so that the Davenport-Erdős-LeVeque theorem applies. For dynamically defined measures there are many recent results that establish a sufficiently fast rate of decay for this result to apply: [1, 13, 27, 16, 30, 31, 24] to name just a few examples (see e.g. [2] for many more references).

Let us compare Theorem 1.1 with the results of [2, 11, 23]: First, we fully relax the separation conditions on the IFS from [11, 23]. Second, while no separation conditions are needed for [2, Theorem 1.4], it does not cover two important cases where Theorem 1.1 does apply: The general case when sisjs_{i}\sim s_{j} for all i,ji,j, and pointwise normality for non-integer Pisot numbers. Third, our work gives a simple and unified approach to this problem that is relatively self-contained. In particular, we do not need to invoke any results on the Fourier transform of self-similar measures as in [2]. On the other hand, we note that the methods and results of both [11] and [2] apply for more general self-conformal IFS’s (i.e. they allow for non-affine smooth maps in the IFS), and that most of the results of [11] work for a broader class of measures on the fractal.

We conclude this section with a brief informal discussion of our method. It consists of three steps: Let μ\mu be a self similar measure as in Theorem 1.1, and let β\beta be a Pisot number such that β≁si\beta\not\sim s_{i} for some ii. We want to show that μ\mu is pointwise β\beta-normal. The first and most critical step is to express μ\mu as an integral over a certain family of random measures. This is based upon a technique that first appeared in [9], and was subsequently applied in [14] and [26], to study the dimension of μ\mu. What was important for these authors was that these random measures could be expressed as an infinite convolution. The crucial new feature in our construction is to ensure that these random measures typically satisfy a certain dynamical self-similarity relation with strong separation. We are also able to preserve the arithmetic condition β≁si\beta\not\sim s_{i}, in some sense, into these random measures.

In the second step of the proof, we show that the scenery flow of typical random measures satisfying this dynamical self-similarity relation with strong separation equidistributes for the same non-trivial ergodic distribution, which we construct explicitly.

The third and final step of the proof is to use spectral analysis of this distribution together with independence from β\beta to conclude, via [11, Theorem 1.2], that almost every random measure is pointwise β\beta-normal. Since in the first step we disintegrated μ\mu according to these measures, Theorem 1.1 follows. The assertion made in Theorem 1.1 about C1C^{1} pushforwards of μ\mu also follows along these lines, by noting that C1C^{1} images of typical random measures will still be pointwise β\beta-normal by [11, Theorem 1.2].

In the next section, after recalling some definitions, we make this sketch precise and formally state the main steps in the proof.

1.2. Statement of the main steps in the proof

We begin by recalling the notion of a model (as in e.g. [26]): Let II be a finite set of iterated function systems of similarities Φ(i)=(f1(i),,fki(i))\Phi^{(i)}=(f_{1}^{(i)},...,f_{k_{i}}^{(i)}), iIi\in I. We assume that each IFS is homogeneous and uniformly contracting. That is, for every iIi\in I there is some ri(1,0)(0,1)r_{i}\in(-1,0)\cup(0,1) such that for every 1jki1\leq j\leq k_{i} we have

fj(i)(x)=rix+tj(i), where tj(i).f^{(i)}_{j}(x)=r_{i}\cdot x+t^{(i)}_{j},\,\text{ where }t^{(i)}_{j}\in\mathbb{R}.

We allow kik_{i} to equal 11 for some ii, i.e. to have degenerate iterated function systems.

Let Ω=I\Omega=I^{\mathbb{N}}. Given a sequence ω=(ωn)nΩ\omega=(\omega_{n})_{n\in\mathbb{N}}\in\Omega we define the space of words of length nn (possibly with n=n=\infty) with respect to ω\omega via

Xn(ω)=j=1n{1,,kωj}.X_{n}^{(\omega)}=\prod_{j=1}^{n}\{1,...,k_{\omega_{j}}\}.

Next, for every ω\omega we define subsets of \mathbb{R} via

Y(ω)={n=1(j=1n1rωj)tun(ωn):uX(ω)}.Y^{(\omega)}=\left\{\sum_{n=1}^{\infty}\left(\prod_{j=1}^{n-1}r_{\omega_{j}}\right)\cdot t^{(\omega_{n})}_{u_{n}}:\,u\in X_{\infty}^{(\omega)}\right\}.

Thus, for every ω\omega we have a surjective coding map Πω:X(ω)Y(ω)\Pi_{\omega}:X^{(\omega)}_{\infty}\rightarrow Y^{(\omega)} defined by

(1) Πω(u)=n=1(j=1n1rωj)tun(ωn),uX(ω).\Pi_{\omega}(u)=\sum_{n=1}^{\infty}\left(\prod_{j=1}^{n-1}r_{\omega_{j}}\right)\cdot t^{(\omega_{n})}_{u_{n}},\quad u\in X^{(\omega)}_{\infty}.

Let σ\sigma denote the left-shift on Ω\Omega (or other shift spaces), and note that for every ωΩ\omega\in\Omega the following dynamical self-similarity relation holds:

(2) Y(ω)=uX1(ω)fu(ω1)(Yσ(ω)).Y^{(\omega)}=\bigcup_{u\in X_{1}^{(\omega)}}f_{u}^{(\omega_{1})}\left(Y^{\sigma(\omega)}\right).

We say that the model satisfies the Strong Separation Condition (SSC) if the union in (2) is disjoint for all ωΩ\omega\in\Omega. Here we adopt the convention that the union of a single set is considered to be disjoint.

Next, for each iIi\in I, let 𝐩i=(p1(i),,pki(i))\mathbf{p}_{i}=(p_{1}^{(i)},...,p_{k_{i}}^{(i)}) be a probability vector with strictly positive entries. On each X(ω)X_{\infty}^{(\omega)} we can then define the product measure η¯(ω):=n=1𝐩ωn\bar{\eta}^{(\omega)}:=\prod_{n=1}^{\infty}\mathbf{p}_{\omega_{n}}. The projection of η¯(ω)\bar{\eta}^{(\omega)} via the coding map is a Borel probability measure η(ω)\eta^{(\omega)} supported on Y(ω)Y^{(\omega)}. For every ω\omega the measure η(ω)\eta^{(\omega)} also satisfies a dynamical self similarity relation, namely,

(3) η(ω)=uX1(ω)pu(ω1)fu(ω)η(σ(ω)).\eta^{(\omega)}=\sum_{u\in X_{1}^{(\omega)}}p_{u}^{(\omega_{1})}f_{u}^{(\omega)}\eta^{(\sigma(\omega))}.

Finally, let \mathbb{P} be a σ\sigma-invariant measure on Ω\Omega. We refer to the triple Σ=(ΦiI(i),(𝐩i)iI,)\Sigma=(\Phi^{(i)}_{i\in I},(\mathbf{p}_{i})_{i\in I},\mathbb{P}) as the model under consideration. We say that the model is non-trivial if η(ω)\eta^{(\omega)} is non-atomic for \mathbb{P}-a.e. ω\omega. We say that the model is ergodic if the selection measure \mathbb{P} is an ergodic measure on Ω\Omega. We say that the model is Bernoulli if the selection measure \mathbb{P} is a Bernoulli measure on Ω\Omega.

As indicated earlier, the proof of Theorem 1.1 consists of three main steps. The first step is to prove that every self similar measure admits a disintegration over the typical measures of a model. Furthermore, this model is Bernoulli, and the arithmetic properties of the original IFS are preserved in an appropriate sense:

Theorem 1.2.

Let μ\mu be a non-atomic self similar measure with respect to an IFS {φi(x)=six+ti}\{\varphi_{i}(x)=s_{i}\cdot x+t_{i}\}, and let β>1\beta>1. Then there exists a non-trivial Bernoulli model Σ=Σ(μ)\Sigma=\Sigma(\mu) with strong separation such that:

  1. (1)

    μ=η(ω)𝑑(ω)\mu=\int\eta^{(\omega)}d\mathbb{P}(\omega).

  2. (2)

    If sj≁βs_{j}\not\sim\beta for some jj then there exist iIi\in I such that ri≁βr_{i}\not\sim\beta.

This theorem is proved in Section §2.

The second step in the proof of Theorem 1.1 is a general result about random model measures being uniformly scaling. Before stating this result, we recall the definition of the scaling scenery of a measure, and related notions. We follow the notations as in [11].

Let 𝒫(X)\mathcal{P}(X) denote the family of Borel probability measures on a metrizable space XX, and define

(4) ={μ𝒫([1,1]): 0supp(μ)}.\mathcal{M}^{\square}=\{\mu\in\mathcal{P}([-1,1]):\,0\in\text{supp}(\mu)\}.

For μ𝒫()\mu\in\mathcal{P}(\mathbb{R}) and xsupp(μ)x\in\text{supp}(\mu), we define the translated and renormalized measure μx\mu_{x}\in\mathcal{M}^{\square} by

μx(E)=cμ(E+x),for E[1,1] a Borel set.\mu_{x}(E)=c\cdot\mu(E+x),\quad\text{for }E\subset[-1,1]\text{ a Borel set.}

Here c=μ([x1,x+1])1c=\mu([x-1,x+1])^{-1} is a normalizing constant. Note that in the definition of μx\mu_{x}, the measure μ\mu does not need to be supported on [1,1][-1,1] (but in any case μx\mu_{x}\in\mathcal{M}^{\square}).

For μ\mu\in\mathcal{M}^{\square}, we define the scaled measure StμS_{t}\mu\in\mathcal{M}^{\square} by

Stμ(E)=cμ(etE),for E[1,1] a Borel set,S_{t}\mu(E)=c^{\prime}\cdot\mu(e^{-t}E),\quad\text{for }E\subset[-1,1]\text{ a Borel set},

where c=μ([et,et])1c^{\prime}=\mu([-e^{-t},e^{-t}])^{-1} is again a normalizing constant.

The scaling flow is the Borel +\mathbb{R}^{+} flow S=(St)t0S=(S_{t})_{t\geq 0} acting on \mathcal{M}^{\square}. The scenery of μ\mu at xsupp(μ)x\in\text{supp}(\mu) is the orbit of μx\mu_{x} under SS, that is, the one parameter family of measures μx,t:=St(μx)\mu_{x,t}:=S_{t}(\mu_{x}) for t0t\geq 0. Thus, the scenery of the measure at some point xx is what one sees as one “zooms into” the measure with focal point xx.

Notice that 𝒫()𝒫(𝒫([1,1]))\mathcal{P}(\mathcal{M}^{\square})\subseteq\mathcal{P}(\mathcal{P}([-1,1])). As is standard in this context, we shall refer to elements of 𝒫(𝒫([1,1]))\mathcal{P}(\mathcal{P}([-1,1])) as distributions (and denote them by capital letters such as QQ), and to elements of 𝒫()\mathcal{P}(\mathbb{R}) as measures (and denote them by Greek letters such as μ\mu). A measure μ𝒫()\mu\in\mathcal{P}(\mathbb{R}) generates a distribution P𝒫(𝒫([1,1]))P\in\mathcal{P}(\mathcal{P}([-1,1])) at xsupp(μ)x\in\text{supp}(\mu) if the scenery at xx equidistributes for PP in 𝒫(𝒫([1,1]))\mathcal{P}(\mathcal{P}([-1,1])), i.e. if

limT1T0Tf(μx,t)𝑑t=f(ν)𝑑P(ν), for all fC(𝒫([1,1])).\lim_{T\rightarrow\infty}\frac{1}{T}\int_{0}^{T}f(\mu_{x,t})dt=\int f(\nu)dP(\nu),\quad\text{ for all }f\in C(\mathcal{P}([-1,1])).

(Any limits involving measures are understood to be in the weak topology.) We say that μ\mu generates PP, and call μ\mu a uniformly scaling measure, if it generates PP at μ\mu almost every xx. If μ\mu generates PP, then PP is supported on \mathcal{M}^{\square} and is SS-invariant [10, Theorem 1.7]. Moreover, PP satisfies a sort of translation invariance known as the quasi-Palm condition (see [10]); we will not use this fact directly but it plays a key role in the proof of [11, Theorem 1.2]) which is one of the main tools in the proof of Theorem 1.1 (and is recalled as Theorem 4.1 below).

We say that PP is trivial if it is the distribution supported on the atom δ0\delta_{0}\in\mathcal{M}^{\square} - a fixed point of SS. We can now state the second main step towards the proof of Theorem 1.1:

Theorem 1.3.

Let Σ\Sigma be a non-trivial Bernoulli model with strong separation. Then there exists a non-trivial SS-ergodic distribution QQ such that for \mathbb{P}-a.e. ω\omega the measure η(ω)\eta^{(\omega)} generates QQ.

In fact, we will derive an explicit expression for QQ as a factor of a suspension flow over a Markov measure. This theorem is proved in Section §3.

The third and final step is to deduce Theorem 1.1 from Theorem 1.2 and Theorem 1.3. To do this, we study the pure point spectrum of the corresponding distribution QQ, showing that it does not contain a non-zero integer multiple of 1logβ\frac{1}{\log\beta}. Once this is established, we apply Theorem 4.1 to obtain the desired pointwise normality result. See Section §4 for more details.

2. Proof of Theorem 1.2

Let Φ={φi(x)=six+ti}i=1n\Phi=\{\varphi_{i}(x)=s_{i}\cdot x+t_{i}\}_{i=1}^{n} be a self similar IFS. Fixing weights 𝐩\mathbf{p}, we begin by constructing a model Σ=Σ(Φ,𝐩)\Sigma=\Sigma(\Phi,\mathbf{p}). We then show that it meets the requirements of Theorem 1.2.

2.1. Construction of the model

Write 𝒜:={1,,n}\mathcal{A}:=\{1,...,n\} and let 𝐩\mathbf{p} be a strictly positive probability vector on 𝒜\mathcal{A}. For every MM\in\mathbb{N} and I=(i1,,im)𝒜MI=(i_{1},...,i_{m})\in\mathcal{A}^{M} write φI=φi1φiM\varphi_{I}=\varphi_{i_{1}}\circ\cdots\circ\varphi_{i_{M}}, and let

ΦM:={φI:I𝒜M}.\Phi^{M}:=\{\varphi_{I}:\,I\in\mathcal{A}^{M}\}.

Recall that we are assuming that the attractor KΦK_{\Phi} is infinite. It is not hard to check (see [29, Proof of Lemma 4.2]) that there exist some MM\in\mathbb{N} and I,J𝒜MI,J\in\mathcal{A}^{M} such that

Conv(φI(KΦ))Conv(φJ(KΦ))= and φI=φJ,\text{Conv}\left(\varphi_{I}(K_{\Phi})\right)\bigcap\text{Conv}\left(\varphi_{J}(K_{\Phi})\right)=\emptyset\,\text{ and }\varphi_{I}^{\prime}=\varphi_{J}^{\prime},

where Conv(X)\text{Conv}(X) denotes the convex hull of a set XX. In addition, notice that μ\mu, the self similar measure corresponding to Φ\Phi and 𝐩\mathbf{p}, is also a self similar measure with respect to the IFS Φm\Phi^{m}, with weights

𝐩M={pi1piM:ij𝒜 for all j}.\mathbf{p}^{M}=\{p_{i_{1}}\cdot\cdot\cdot p_{i_{M}}:\,i_{j}\in\mathcal{A}\text{ for all }j\}.

The contraction factors of Φm\Phi^{m} include sims_{i}^{m}, so the hypothesis that some si≁βs_{i}\not\sim\beta is preserved by this iteration. This shows that we may assume, without loss of generality, that there exists i,j𝒜i,j\in\mathcal{A} such that

(5) Conv(φi(KΦ))Conv(φj(KΦ))= and φi=φj.\text{Conv}\left(\varphi_{i}(K_{\Phi})\right)\bigcap\text{Conv}\left(\varphi_{j}(K_{\Phi})\right)=\emptyset\,\text{ and }\varphi_{i}^{\prime}=\varphi_{j}^{\prime}.

After relabeling, we may also assume i=1,j=2i=1,j=2. We will use (5) to show that the model satisfies the SSC.

We now define I:={j:j=0 or j𝒜{1,2}}I:=\{j:\,j=0\text{ or }j\in\mathcal{A}\setminus\{1,2\}\}. For every index in II we can associate an IFS as follows:

  • For j=0j=0 we associate the IFS {φ1,φ2}\{\varphi_{1},\varphi_{2}\}.

  • For every jI{0},j\in I\setminus\{0\}, we associate the degenerate IFS {φj}\{\varphi_{j}\}.

Notice that the construction above gives us a dichotomy: Each IFS in II is either degenerate, or is homogeneous with strong separation in the sense of (5).

Recall that 𝐩\mathbf{p} was our fixed probability vector on 𝒜\mathcal{A}. We now define a probability vector 𝐪\mathbf{q} on II as follows:

  • The mass 𝐪\mathbf{q} gives to 0 is p1+p2p_{1}+p_{2}, that is, q0=p1+p2q_{0}=p_{1}+p_{2}.

  • For every jI{0}j\in I\setminus\{0\}, the singleton jj gets mass qj=pjq_{j}=p_{j}.

Recalling our notation from Section §2.1, let \mathbb{P} be the corresponding Bernoulli measure 𝐪\mathbf{q}^{\mathbb{N}} on I:=ΩI^{\mathbb{N}}:=\Omega. This \mathbb{P} will be our (Bernoulli) selection measure.

Next, for 0I0\in I we define the probability vector

𝐩~0=(p1p1+p2,p2p1+p2)=(p1q0,p2q0).\tilde{\mathbf{p}}_{0}=\left(\frac{p_{1}}{p_{1}+p_{2}},\,\frac{p_{2}}{p_{1}+p_{2}}\right)=\left(\frac{p_{1}}{q_{0}},\,\frac{p_{2}}{q_{0}}\right).

For every jI{0}j\in I\setminus\{0\} we define the (degenerate) probability vector

𝐩~j=(1).\tilde{\mathbf{p}}_{j}=(1).

Recall from Section §1.2 that with these probability vectors, on each X(ω)X_{\infty}^{(\omega)} we can then define the product measure η¯(ω)\bar{\eta}^{(\omega)}, and the projection of η¯(ω)\bar{\eta}^{(\omega)} via the coding map is a Borel probability measure η(ω)\eta^{(\omega)} supported on Y(ω)Y^{(\omega)}.

2.2. Proof of the required properties

In this section we show that the model Σ\Sigma constructed in Section §2.1 satisfies the conclusion of Theorem 1.2. First, the model Σ\Sigma is Bernoulli by definition. Secondly, the union in (2) is disjoint for all ω\omega: if ω10\omega_{1}\neq 0 there is nothing to do; otherwise, this follows from (5) and the fact that all sets Y(ω)Y^{(\omega^{\prime})} are contained in the original attractor KK. Since among the |ri||r_{i}| we find integer powers of all the original |sj||s_{j}|, if there is jj such that sj≁βs_{j}\not\sim\beta, then there is also some iIi\in I such that ri≁βr_{i}\not\sim\beta.

We proceed to prove part (1) of Theorem 1.2: Let μ\mu be the self similar measure that corresponds to the weights 𝐩\mathbf{p}. We will show that

η(ω)𝑑(ω)=i𝒜piφiη(ω)𝑑(ω).\int\eta^{(\omega)}d\mathbb{P}(\omega)=\sum_{i\in\mathcal{A}}p_{i}\varphi_{i}\int\eta^{(\omega)}d\mathbb{P}(\omega).

Since μ\mu is the unique Borel probability measure satisfying this self-similarity relation, it will then follow that

μ=η(ω)𝑑(ω).\mu=\int\eta^{(\omega)}d\mathbb{P}(\omega).

We have

η(ω)𝑑(ω)\displaystyle\int\eta^{(\omega)}d\mathbb{P}(\omega) =\displaystyle= (uX1(ω)p~u(ω1)fu(ω)η(σ(ω)))𝑑(ω)\displaystyle\int\left(\sum_{u\in X_{1}^{(\omega)}}\tilde{p}_{u}^{(\omega_{1})}f_{u}^{(\omega)}\eta^{(\sigma(\omega))}\right)d\mathbb{P}(\omega)
=\displaystyle= iI[i](uX1(ω)p~u(ω1)fu(ω)η(σ(ω)))𝑑(ω)\displaystyle\sum_{i\in I}\int_{[i]}\left(\sum_{u\in X_{1}^{(\omega)}}\tilde{p}_{u}^{(\omega_{1})}f_{u}^{(\omega)}\eta^{(\sigma(\omega))}\right)d\mathbb{P}(\omega)
=\displaystyle= jI,j0pj(φjη(σ(jω)))𝑑(ω)\displaystyle\sum_{j\in I,j\neq 0}p_{j}\int\left(\varphi_{j}\eta^{(\sigma(j*\omega))}\right)d\mathbb{P}(\omega)
+\displaystyle+ (p1+p2)(p1p1+p2φ1η(σ(1ω))+p2p1+p2φ2η(σ(2ω)))𝑑(ω)\displaystyle(p_{1}+p_{2})\cdot\int\left(\frac{p_{1}}{p_{1}+p_{2}}\varphi_{1}\eta^{(\sigma(1*\omega))}+\frac{p_{2}}{p_{1}+p_{2}}\varphi_{2}\eta^{(\sigma(2*\omega))}\right)d\mathbb{P}(\omega)
=\displaystyle= i𝒜piφiη(ω)𝑑(ω).\displaystyle\sum_{i\in\mathcal{A}}p_{i}\varphi_{i}\int\eta^{(\omega)}d\mathbb{P}(\omega).

Which is what we claimed.

So far we have established all the claims in Theorem 1.2, except for the non-degeneracy of the model: We claim that for \mathbb{P}-a.e. ω\omega, the measure η(ω)\eta^{(\omega)} is non-atomic. Indeed, by the SSC for any ω\omega the coding map Πω\Pi_{\omega} is injective. So by our choice of weights it follows that for every nn\in\mathbb{N} and InXn(ω)I_{n}\in X_{n}^{(\omega)}, we have

(6) η(ω)(Πω(In))=η¯(ω)({In})(max{p1p1+p2,p2p1+p2})|{1kn:ωk=0}|.\eta^{(\omega)}\left(\Pi_{\omega}\left(I_{n}\right)\right)=\bar{\eta}^{(\omega)}\left(\{I_{n}\}\right)\leq\left(\max\left\{\frac{p_{1}}{p_{1}+p_{2}},\,\frac{p_{2}}{p_{1}+p_{2}}\right\}\right)^{\left|\{1\leq k\leq n:\,\omega_{k}=0\}\right|}.

Since \mathbb{P} is a Bernoulli measure on II^{\mathbb{N}} with strictly positive weights, \mathbb{P}-a.s. the digit 0 occurs in ω\omega infinitely times. This shows that the right hand side of (6) tends to 0 as nn\rightarrow\infty for \mathbb{P}-a.e. ω\omega. Since any point in X(ω)X_{\infty}^{(\omega)} is covered by sets Πω(In)\Pi_{\omega}\left(I_{n}\right) for all nn, the measures η(ω)\eta^{(\omega)} are \mathbb{P}-a.s. non-atomic, as claimed. This concludes the proof of Theorem 1.2.

3. Proof of Theorem 1.3

We split out proof of Theorem 1.3 into two cases. We first detail the case when each IFS in our model is orientation preserving, i.e. ri(0,1)r_{i}\in(0,1) for all iIi\in I; this case avoids certain technicalities. We then consider the case when we have at least one ii satisfying ri(1,0)r_{i}\in(-1,0).

3.1. The orientation preserving case

Assume the conditions of Theorem 1.3 hold true. In particular, recall that we are assuming the SSC (2), and that the selection measure \mathbb{P} is a non-degenerate Bernoulli measure. Without loss of generality, we may assume that the sets (fu(ω1)(Y(σω)))uX1(ω)(f_{u}^{(\omega_{1})}(Y^{(\sigma\omega)}))_{u\in X_{1}^{(\omega)}} are at distance >2>2 from each other for all ωΩ\omega\in\Omega. See the discussion in §3.3 on how to treat the general case.

First, we define the random pair (ω,u)(\omega,u), where ω\omega is drawn according to \mathbb{P} and uu is drawn according to η¯(ω).\overline{\eta}^{(\omega)}. Let \mathbb{P}^{\prime} denote this distribution. In fact, \mathbb{P}^{\prime} is the Bernoulli measure on the symbols {(i,u):iI,1uki}\{(i,u):i\in I,1\leq u\leq k_{i}\} with weights qi,u=([i])pu(i)q_{i,u}=\mathbb{P}([i])\cdot p_{u}^{(i)}. We let MM be the shift map on Ω=supp()\Omega^{\prime}=\operatorname{supp}(\mathbb{P}^{\prime}). In particular, \mathbb{P}^{\prime} is MM-invariant and ergodic. Write ηω,u=ηΠω(u)(ω)\eta_{\omega,u}=\eta^{(\omega)}_{\Pi_{\omega}(u)} for simplicity; recall that this is the translation of η(ω)\eta^{(\omega)} that centers it at Πω(u)\Pi_{\omega}(u).

Claim 3.1.

We have: Slog(rω1)ηω,u=ηM(ω,u)S_{-\log(r_{\omega_{1}})}\eta_{\omega,u}=\eta_{M(\omega,u)}.

Proof.

The claim is essentially due to dynamical self-similarity and the SSC. Fix ωΩ\omega\in\Omega, uX(ω)u\in X_{\infty}^{(\omega)}, and let E[1,1]E\subset[-1,1] be a Borel subset. It follows from (1) that

(7) Πω(u)=fu(ω1)(Πσω(σu)).\Pi_{\omega}(u)=f_{u}^{(\omega_{1})}(\Pi_{\sigma\omega}(\sigma u)).

Write x=Πω(u)x=\Pi_{\omega}(u), and note that for a Borel set E[1,1]E\subset[-1,1] we have

Slog(rω1)ηω,u(E)=cη(ω)(rω1E+x)=cfuω1η(σω)(rω1E+x).S_{-\log(r_{\omega_{1}})}\eta_{\omega,u}(E)=c\cdot\eta^{(\omega)}(r_{\omega_{1}}\cdot E+x)=c\cdot f_{u}^{\omega_{1}}\eta^{(\sigma\omega)}(r_{\omega_{1}}\cdot E+x).

Indeed, by our assumption that the distance between the sets fv(ω1)(Y(σω))f_{v}^{(\omega_{1})}(Y^{(\sigma\omega)}) is greater than 22, at most one of them may intersect rω1E+xr_{\omega_{1}}\cdot E+x (a set of diameter <2<2). Continuing this line of reasoning, we have

cfu(ω1)η(σω)(rω1E+x)\displaystyle c\cdot f_{u}^{(\omega_{1})}\eta^{(\sigma\omega)}(r_{\omega_{1}}\cdot E+x) =cη(σω)(rω1E+xtuω1rω1)\displaystyle=c\cdot\eta^{(\sigma\omega)}\left(\frac{r_{\omega_{1}}\cdot E+x-t_{u}^{\omega_{1}}}{r_{\omega_{1}}}\right)
=cη(σω)(E+(fu(ω1))1(x))\displaystyle=c\cdot\eta^{(\sigma\omega)}(E+(f_{u}^{(\omega_{1})})^{-1}(x))
=ηM(ω,u),\displaystyle=\eta_{M(\omega,u)},

where we used (7) in the last line. ∎

Claim 3.2.

For \mathbb{P}-a.e. ω\omega, the measure η(ω)\eta^{(\omega)} generates the distribution QQ that is defined as follows: let Q¯\overline{Q} be the suspension measure for the suspension flow with base (Ω,)(\Omega^{\prime},\mathbb{P}^{\prime}) and roof function ρ(ω,u)=log(rω1)\rho(\omega,u)=-\log(r_{\omega_{1}}). That is, denoting Lebesgue measure by λ\lambda,

Q¯=1log(rω1)d(ω)(×λ)|{(ω,u,t):0t<log(rω1)}.\overline{Q}=\frac{1}{\int-\log(r_{\omega_{1}})\,d\mathbb{P}(\omega)}(\mathbb{P}^{\prime}\times\lambda)|_{\{(\omega,u,t):0\leq t<-\log(r_{\omega_{1}})\}}.

Then QQ is the push-forward of Q¯\overline{Q} under (ω,u,t)Stηω,u(\omega,u,t)\mapsto S_{t}\eta_{\omega,u}.

Proof.

Our goal is to show that for \mathbb{P}^{\prime} almost all (ω,u)(\omega,u), the measure η(ω)\eta^{(\omega)} generates QQ at Πω(u)\Pi_{\omega}(u), or in other words, the scenery (Stηω,u)t=0(S_{t}\eta_{\omega,u})_{t=0}^{\infty} equidistributes for QQ. Fix fC(𝒫([1,1]))f\in C(\mathcal{P}([-1,1])) and let

F(ω,u):=0logrω1f(Stηω,u)𝑑t.F(\omega,u):=\int_{0}^{-\log r_{\omega_{1}}}f(S_{t}\eta_{\omega,u})\,dt.

By Claim 3.1,

(8) j=0N1F(Mj(ω,u))=0j=1Nlog(rωj)f(Stηω,u)𝑑t.\sum_{j=0}^{N-1}F(M^{j}(\omega,u))=\int_{0}^{\sum_{j=1}^{N}-\log(r_{\omega_{j}})}f(S_{t}\eta_{\omega,u})\,dt.

By the ergodic theorem, for \mathbb{P}^{\prime}-a.e. (ω,u)(\omega,u),

limN1Nj=0N1F(Mj(ω,u))=ΩX(ω)0logrω1f(Stηω,u)𝑑t𝑑η¯(ω)(u)𝑑(ω)\lim_{N\to\infty}\frac{1}{N}\sum_{j=0}^{N-1}F(M^{j}(\omega,u))=\int_{\Omega}\int_{X_{\infty}^{(\omega)}}\int_{0}^{-\log r_{\omega_{1}}}f(S_{t}\eta_{\omega,u})\,dt\,d\overline{\eta}^{(\omega)}(u)\,d\mathbb{P}(\omega)

and

limNNj=1Nlog(rωj)=1i([i])log(1/ri).\lim_{N\to\infty}\frac{N}{\sum_{j=1}^{N}-\log(r_{\omega_{j}})}=\frac{1}{\sum_{i}\mathbb{P}([i])\log(1/r_{i})}.

Multiplying these two identities, recalling (8), and using that rir_{i} is bounded away from 0 and 11, we see that for \mathbb{P}-almost all ω\omega and η¯(ω)\overline{\eta}^{(\omega)}-almost all uu, we have

limT1T0Tf(Stηω,u)𝑑t=f(ν)𝑑Q(ν).\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}f(S_{t}\eta_{\omega,u})\,dt=\int f(\nu)\,dQ(\nu).

Finally, by the separability of C(𝒫([1,1]))C(\mathcal{P}([-1,1])), for \mathbb{P}-almost all ω\omega and η¯(ω)\overline{\eta}^{(\omega)}-almost all u,u, the equation above holds for every fC(𝒫([1,1]))f\in C(\mathcal{P}([-1,1])). This proves the claim. ∎

Proof of Theorem 1.3 in the orientation-preserving case.

First, by Claim 3.2, for \mathbb{P}-a.e. ω\omega, the measure η(ω)\eta^{(\omega)} generates the distribution QQ as in Claim 3.2. Furthermore, QQ is non-trivial, since the model is non-trivial. Finally, we need to establish the ergodicity of QQ. This follows since Q¯\overline{Q} is ergodic, being the suspension of an ergodic (in fact, Bernoulli) measure, and the map (ω,u,t)Stηω,u(\omega,u,t)\mapsto S_{t}\eta_{\omega,u} taking Q¯\overline{Q} to QQ is a factor map, as can be seen from Claim 3.1. ∎

We remark that our proof of Theorem 1.3 in the orientation preserving case also works under the weaker assumption that Σ\Sigma is a non-trivial ergodic model. In this case \mathbb{P}^{\prime} is no longer Bernoulli but can still be seen to be ergodic. Our proof of this theorem in the orientation reversing case relies upon the ergodicity of certain group extensions. The ergodicity of these group extensions does not always hold if our model is just assumed to be ergodic. Fortunately however, ergodicity can be shown to hold if our model is Bernoulli.

3.2. The case with some orientation-reversing map

Throughout this section we will assume that our model is such that there exists at least one ii satisfying ri<0r_{i}<0. This case is slightly more complicated because as we zoom in on ηω,x,\eta_{\omega,x}, we are not necessarily going to see ηM(ω,x)\eta_{M(\omega,x)}. Because of the presence of orientation reversing maps, it is possible that as we zoom in on ηω,x,\eta_{\omega,x}, we may see the image of ηM(ω,x)\eta_{M(\omega,x)} reflected around the origin. Our analogue of the map MM has to take this behaviour into account, and our distribution QQ also has to see these reflected measures.

We again assume the conditions of Theorem 1.3 are satisfied and that the sets (fu(ω1)(Y(σω)))uX1(ω)(f_{u}^{(\omega_{1})}(Y^{(\sigma\omega)}))_{u\in X_{1}^{(\omega)}} are at distance >2>2 from each other. We let \mathbb{P}^{\prime} be as in the orientation preserving case, and recall that its support is denoted by Ω\Omega^{\prime}. We define a map M:Ω×2Ω×2M^{\prime}:\Omega^{\prime}\times\mathbb{Z}_{2}\to\Omega^{\prime}\times\mathbb{Z}_{2} as follows:

M(ω,u,a)={(M(ω,u),a)if rω1>0;(M(ω,u),a+1)if rω1<0.M^{\prime}(\omega,u,a)=\left\{\begin{array}[]{ll}\left(M(\omega,u),a\right)&\mbox{if $r_{\omega_{1}}>0$};\\ \left(M(\omega,u),a+1\right)&\mbox{if $r_{\omega_{1}}<0$}.\end{array}\right.

We let 𝔪\mathfrak{m} denote the Haar measure on 2\mathbb{Z}_{2} and let ′′=×𝔪.\mathbb{P}^{\prime\prime}=\mathbb{P}^{\prime}\times\mathfrak{m}.

Lemma 3.3.

′′\mathbb{P}^{\prime\prime} is MM^{\prime}-invariant and ergodic; moreover, it is isomorphic to an ergodic Markov measure on an irreducible Markov chain.

Proof.

The state space of the Markov chain is 𝒮={(i,u,a):iI,1uki,a2}\mathcal{S}=\{(i,u,a):i\in I,1\leq u\leq k_{i},a\in\mathbb{Z}_{2}\} and the probability of transitioning from (i,u,a)(i,u,a) to (i,u,a)(i^{\prime},u^{\prime},a^{\prime}) is ([i])pu(i)\mathbb{P}([i^{\prime}])\cdot p_{u^{\prime}}^{(i^{\prime})} if a=aa=a^{\prime} and ri>0r_{i}>0 or if aaa\neq a^{\prime} and ri<0r_{i}<0; and the transition is forbidden otherwise. Then

(12([i])pu(i))(i,u,a)𝒮\left(\frac{1}{2}\mathbb{P}([i])\cdot p_{u}^{(i)}\right)_{(i,u,a)\in\mathcal{S}}

is readily checked to be a stationary probability for the Markov chain. The associated Markov measure ~′′\widetilde{\mathbb{P}}^{\prime\prime} is supported on Ω×2\Omega^{\prime}\times\mathbb{Z}_{2}^{\mathbb{N}} and satisfies

~′′([(ω1,u1,a1)(ωn,un,an)])=12([(ω1,u1)(ωn,un)]).\widetilde{\mathbb{P}}^{\prime\prime}([(\omega_{1},u_{1},a_{1})\cdots(\omega_{n},u_{n},a_{n})])=\frac{1}{2}\mathbb{P}^{\prime}([(\omega_{1},u_{1})\cdots(\omega_{n},u_{n})]).

Since ω\omega and a1a_{1} determine aja_{j} for j2j\geq 2 almost surely according to the transition rule, the projection (ω,u,a)(ω,u,a1)(\omega,u,a)\mapsto(\omega,u,a_{1}) provides a natural isomorphism between (Ω×2,~′′,σ)(\Omega^{\prime}\times\mathbb{Z}_{2}^{\mathbb{N}},\widetilde{\mathbb{P}}^{\prime\prime},\sigma) and (Ω×2,′′,M)(\Omega^{\prime}\times\mathbb{Z}_{2},\mathbb{P}^{\prime\prime},M^{\prime}).

Finally, the subshift of finite type underlying the Markov chain is irreducible (in fact, one can get from one state to any other in at most two steps), so ~′′\widetilde{\mathbb{P}}^{\prime\prime}, and hence ′′\mathbb{P}^{\prime\prime}, is ergodic. ∎

To each element of Ω×2\Omega^{\prime}\times\mathbb{Z}_{2} we associate a measure as follows: Let E[1,1]E\subset[-1,1] be a Borel set, then

ηω,u,a(E)={ηω,u(E)if a=0;ηω,u(E)if a=1.\eta_{\omega,u,a}(E)=\left\{\begin{array}[]{ll}\eta_{\omega,u}(E)&\mbox{if $a=0$};\\ \eta_{\omega,u}(-E)&\mbox{if $a=1$}.\end{array}\right.
Claim 3.4.

We have: Slog(|rω1|)ηω,u,a=ηM(ω,u,a)S_{-\log(|r_{\omega_{1}}|)}\eta_{\omega,u,a}=\eta_{M^{\prime}(\omega,u,a)}.

Proof.

The claim follows from a case analysis based upon the sign of rω1r_{\omega_{1}} and the value of aa. Note that the case where rω1>0r_{\omega_{1}}>0 and a=0a=0 is essentially Claim 3.1. For the purpose of brevity we do not cover each case here. We instead content ourselves with the case where rω1<0r_{\omega_{1}}<0 and a=0a=0. Let E[1,1]E\subset[-1,1] be a Borel set. Then, using the separation between the sets (fv(ω1)(Y(σω)))vX1(ω)(f_{v}^{(\omega_{1})}(Y^{(\sigma\omega)}))_{v\in X_{1}^{(\omega)}}, and denoting x=Πω(u)x=\Pi_{\omega}(u), we have

Slog(|rω1|)ηω,u,0(E)\displaystyle S_{-\log(|r_{\omega_{1}}|)}\eta_{\omega,u,0}(E) =cη(ω)(|rω1|E+x)\displaystyle=c\cdot\eta^{(\omega)}(|r_{\omega_{1}}|\cdot E+x)
=cfuω1η(σω)(|rω1|E+x)\displaystyle=c\cdot f_{u}^{\omega_{1}}\eta^{(\sigma\omega)}(|r_{\omega_{1}}|\cdot E+x)
=cη(σω)(|rω1|E+xtuω1rω1)\displaystyle=c\cdot\eta^{(\sigma\omega)}\left(\frac{|r_{\omega_{1}}|\cdot E+x-t_{u}^{\omega_{1}}}{r_{\omega_{1}}}\right)
=cη(σω)(E+(fuω1)1(x))\displaystyle=c\cdot\eta^{(\sigma\omega)}\left(-E+(f_{u}^{\omega_{1}})^{-1}(x)\right)
=ηM(ω,u)(E)=ηM(ω,u),1(E)=ηM(ω,u,0)(E).\displaystyle=\eta_{M(\omega,u)}(-E)=\eta_{M(\omega,u),1}(E)=\eta_{M^{\prime}(\omega,u,0)}(E).

What remains of the proof of Theorem 1.3 in the orientation reversing case is an almost identical argument to that presented in the orientation preserving case after Claim 3.1 is established. For the purpose of completion, we mention that in this case the distribution QQ is defined as the push-forward of Q¯\overline{Q} under the factor map (ω,x,a,t)Stηω,x,a(\omega,x,a,t)\mapsto S_{t}\eta_{\omega,x,a}, where Q¯\overline{Q} is the suspension measure over ′′\mathbb{P}^{\prime\prime} with roof function log|rω1|-\log|r_{\omega_{1}}|.

3.3. Removing the distance >2>2 assumption

Finally, we discuss the general case when the smallest distance between the sets (fu(ω1)(Y(σω)))uX1(ω)(f_{u}^{(\omega_{1})}(Y^{(\sigma\omega)}))_{u\in X_{1}^{(\omega)}} is not larger than 22 (note that by compactness, under the SSC there is indeed a smallest positive such distance). There are (at least) two ways to get around this. The first is to note that this can be achieved by rescaling all the translations of the IFS’s in the model by a common factor. The second is, assuming 2t0>02t_{0}>0 is smaller than the smallest distance between the sets (fu(ω1)(Y(σω)))uX1(ω)(f_{u}^{(\omega_{1})}(Y^{(\sigma\omega)}))_{u\in X_{1}^{(\omega)}}, to condition our measures not on [1,1][-1,1] but rather on [t0,t0][-t_{0},t_{0}] in the magnification process. By the discussion in [10, Section 1 and Section 3.1], the first way does not affect the uniformly scaling nature of the measures or the generated distributions, and the second has a corresponding very mild such effect.

4. Proof of Theorem 1.1

Let S=(St)t=0S=(S_{t})_{t=0}^{\infty} be a measurable (semi)flow on a space XX, and let PP be SS-invariant. Let e(s)=exp(2πis)e(s)=\exp(2\pi is). The pure point spectrum Σ(P,S)\Sigma(P,S) of PP is the set of all the α\alpha\in\mathbb{R} for which there exists a non-zero measurable eigenfunction φ:X\varphi:X\rightarrow\mathbb{C} such that φSt=e(αt)φ\varphi\circ S_{t}=e(\alpha t)\varphi for every t0t\geq 0, on a set of full PP measure. If PP is SS-ergodic, then |φ||\varphi| is constant for any eigenfunction φ\varphi; in particular, any such measurable eigenfunction is in fact in L2(P)L^{2}(P). We also note that if (Y,Q,S)(Y,Q,S) is a factor of (X,P,S)(X,P,S^{\prime}) then Σ(Q,S)Σ(P,S)\Sigma(Q,S)\subset\Sigma(P,S^{\prime}): indeed, if Π\Pi is the factor map, then any eigenfunction φ\varphi on YY gives rise to the eigenfunction φΠ\varphi\circ\Pi (with the same eigenvalue) on XX.

We now specialize to the scenery flow; recall Section §1.2. In this case, the existence of an eigenvalue α\alpha indicates that some non-trivial feature of the measures generated by PP repeats periodically under magnification by e1/αe^{1/\alpha}. The following theorem of Hochman and Shmerkin [11, Theorems 1.1 and 1.2] relates the pure point spectrum of a distribution generated by a measure with equidistribution. Recall that a distribution PP is called trivial if it is the distribution supported on δ0\delta_{0}\in\mathcal{M}^{\square}.

Theorem 4.1.

Let μ\mu be a measure generating a non-trivial SS-ergodic distribution PP, and let β\beta be a Pisot number. If Σ(P,S)\Sigma(P,S) does not contain a nonzero integer multiple of 1logβ\frac{1}{\log\beta}, then μ\mu is pointwise β\beta-normal. Furthermore, the same is true for gμg\mu for all gdiff1()g\in\text{diff}^{1}(\mathbb{R}).

Finally, we recall a classical fact regarding eigenfunctions of suspensions over Markov measures; see [21, Proposition 6.2] for the proof (of a more general statement).

Theorem 4.2.

Let (X,,σ)(X,\mathbb{P},\sigma) be a subshift of finite type XX endowed with a Markov measure \mathbb{P}, let ρL1()\rho\in L^{1}(\mathbb{P}) be a non-negative measurable roof function, and let (Xρ,ρ,S)(X_{\rho},\mathbb{P}_{\rho},S) denote the corresponding suspension flow. If αΣ(ρ,S)\alpha\in\Sigma(\mathbb{P}_{\rho},S), then there is fC(X)f\in C(X) such that

(9) f(σx)=e(αρ(x))f(x),xX.f(\sigma x)=e(\alpha\rho(x))f(x),\quad x\in X.

With these ingredients, we can conclude the proof of theorem 1.1.

Proof of Theorem 1.1.

Let μ\mu be a self similar measure that satisfies the condition of Theorem 1.1 with respect to a Pisot number β>1\beta>1, and let gdiff1()g\in\text{diff}^{1}(\mathbb{R}). First, apply Theorem 1.2 to obtain a disintegration of μ\mu according to the corresponding model Σ=(ΦiI(i),(𝐩i)iI,)\Sigma=(\Phi^{(i)}_{i\in I},(\mathbf{p}_{i})_{i\in I},\mathbb{P}). That is,

μ=η(ω)𝑑(ω).\mu=\int\eta^{(\omega)}d\mathbb{P}(\omega).

It follows that

gμ=gη(ω)𝑑(ω).g\mu=\int g\eta^{(\omega)}d\mathbb{P}(\omega).

Therefore, to show that gμg\mu is pointwise β\beta-normal it is enough to show that \mathbb{P}-almost surely the measure gη(ω)g\eta^{(\omega)} is pointwise β\beta-normal.

By Theorem 1.2 there is some jIj\in I such that |rj|≁β|r_{j}|\not\sim\beta. Also, by Theorem 1.3, for \mathbb{P}-a.e. ω\omega the measure η(ω)\eta^{(\omega)} generates the non-trivial SS-ergodic distribution QQ, constructed in Section §3. Moreover, (,Q,S)(\mathcal{M}^{\square},Q,S) arises as a factor of a suspension flow with an ergodic Markov measure on the base - recall Claim 3.2, Lemma 3.3 and the discussion at the end of §3.2.

The state space of the corresponding Markov chain is either {(i,u):iI,u{1,,ki}}\{(i,u):i\in I,u\in\{1,\ldots,k_{i}\}\} in the orientation-preserving case (in this case the Markov measure is actually Bernoulli), and {(i,u,a):iI,u{1,,ki},a2}\{(i,u,a):i\in I,u\in\{1,\ldots,k_{i}\},a\in\mathbb{Z}_{2}\} when at least one IFS in the model is orientation-reversing. In both cases, the roof function is log(|ri|)-\log(|r_{i}|) where ii corresponds to the current state of the Markov chain. Applying (9) to the fixed point (j,1)(j,1)^{\infty} in the orientation-preserving case, or applying (9) twice to the periodic point ((j,1,0)(j,1,1))((j,1,0)(j,1,1))^{\infty} otherwise (which is allowed since ff is continuous), we deduce that 2αlog(|rj|)2\alpha\log(|r_{j}|)\in\mathbb{Z}.

Since |rj|≁β|r_{j}|\not\sim\beta, we see that α\alpha cannot be a non-zero integer multiple of 1/log(β)1/\log(\beta). As we discussed at the beginning of this section, the same holds for the factor (,Q,S)(\mathcal{M}^{\square},Q,S). Now the desired conclusion follows from Theorem 4.1.

References

  • [1] Amir Algom, Federico Rodriguez Hertz, and Zhiren Wang. Logarithmic Fourier decay for self conformal measures. arXiv preprint arXiv: 2109.13017, 2021.
  • [2] Amir Algom, Federico Rodriguez Hertz, and Zhiren Wang. Pointwise normality and Fourier decay for self-conformal measures. to appear in Advances in Mathematics, 2021.
  • [3] Simon Baker. Equidistribution results for self-similar measures. To appear in International Mathematics Research Notices, 2021.
  • [4] Julien Brémont. On the rajchman property for self-similar measures. To appear in Annales H. Lebesgue, 2021.
  • [5] John William Scott Cassels. On a problem of Steinhaus about normal numbers. Colloq. Math., 7:95–101, 1959.
  • [6] Harold Davenport, Paul Erdős, and William J. LeVeque. On Weyl’s criterion for uniform distribution. Michigan Math. J., 10:311–314, 1963.
  • [7] Yiftach Dayan, Arijit Ganguly, and Barak Weiss. Random walks on tori and normal numbers in self similar sets. arXiv preprint arXiv:2002.00455, 2020.
  • [8] Jacob Feldman and Meir Smorodinsky. Normal numbers from independent processes. Ergodic Theory Dynam. Systems, 12(4):707–712, 1992.
  • [9] Daniel Galicer, Santiago Saglietti, Pablo Shmerkin, and Alexia Yavicoli. LqL^{q} dimensions and projections of random measures. Nonlinearity, 29(9):2609–2640, 2016.
  • [10] Michael Hochman. Dynamics on fractals and fractal distributions. arXiv preprint arXiv:1008.3731, 2010.
  • [11] Michael Hochman and Pablo Shmerkin. Equidistribution from fractal measures. Inventiones mathematicae, 202(1):427–479, 2015.
  • [12] Bernard Host. Nombres normaux, entropie, translations. Israel J. Math., 91(1-3):419–428, 1995.
  • [13] Thomas Jordan and Tuomas Sahlsten. Fourier transforms of Gibbs measures for the Gauss map. Math. Ann., 364(3-4):983–1023, 2016.
  • [14] Antti Käenmäki and Tuomas Orponen. Absolute continuity in families of parametrised non-homogeneous self-similar measures. arXiv preprint arXiv:1812.05006, 2018.
  • [15] Ka-Sing Lau and Sze-Man Ngai. Multifractal measures and a weak separation condition. Adv. Math., 141(1):45–96, 1999.
  • [16] Jialun Li and Tuomas Sahlsten. Trigonometric series and self-similar sets. To appear in Journal of the European Mathematical Society, 2021.
  • [17] Elon Lindenstrauss. pp-adic foliation and equidistribution. Israel J. Math., 122:29–42, 2001.
  • [18] David Meiri. Entropy and uniform distribution of orbits in 𝐓d{\bf T}^{d}. Israel J. Math., 105:155–183, 1998.
  • [19] Carolina A. Mosquera and Pablo S. Shmerkin. Self-similar measures: asymptotic bounds for the dimension and Fourier decay of smooth images. Ann. Acad. Sci. Fenn. Math., 43(2):823–834, 2018.
  • [20] William Parry. On the β\beta-expansions of real numbers. Acta Math. Acad. Sci. Hungar., 11:401–416, 1960.
  • [21] William Parry and Mark Pollicott. Zeta functions and the periodic orbit structure of hyperbolic dynamics. Astérisque, (187-188):268, 1990.
  • [22] Yuval Peres, Károly Simon, and Boris Solomyak. Self-similar sets of zero Hausdorff measure and positive packing measure. Israel J. Math., 117:353–379, 2000.
  • [23] Aleksi Pyörälä. The scenery flow of self-similar measures with weak separation condition. Ergodic Theory and Dynamical Systems, page 1–24, 2021.
  • [24] Ariel Rapaport. On the Rajchman property for self-similar measures on d\mathbb{R}^{d}. arXiv preprint arXiv:2104.03955, 2021.
  • [25] Alfréd Rényi. Representations for real numbers and their ergodic properties. Acta Math. Acad. Sci. Hungar., 8:477–493, 1957.
  • [26] Santiago Saglietti, Pablo Shmerkin, and Boris Solomyak. Absolute continuity of non-homogeneous self-similar measures. Adv. Math., 335:60–110, 2018.
  • [27] Tuomas Sahlsten and Connor Stevens. Fourier transform and expanding maps on cantor sets. arXiv preprint arXiv:2009.01703, 2020.
  • [28] Wolfgang M. Schmidt. On normal numbers. Pacific J. Math., 10:661–672, 1960.
  • [29] Pablo Shmerkin. Projections of Self-Similar and Related Fractals: A Survey of Recent Developments, pages 53–74. Springer International Publishing, Cham, 2015.
  • [30] Boris Solomyak. Fourier decay for self-similar measures. Proc. Amer. Math. Soc., 149(8):3277–3291, 2021.
  • [31] Péter P Varjú and Han Yu. Fourier decay of self-similar measures and self-similar sets of uniqueness. To appear in Analysis and PDE.
  • [32] Martin P. W. Zerner. Weak separation properties for self-similar sets. Proc. Amer. Math. Soc., 124(11):3529–3539, 1996.

Department of Mathematics, the Pennsylvania State University, University Park, PA 16802, USA

E-mail address [email protected]


School of Mathematics, University of Birmingham, Birmingham, B15 2TT, UK

E-mail address [email protected]


Department of Mathematics, the University of British Columbia, 1984 Mathematics Road Vancouver, BC, V6T 1Z2, Canada

E-mail address [email protected]