On normal numbers and self-similar measures
Abstract.
Let be a self-similar IFS on and let be a Pisot number. We prove that if for some then for every diffeomorphism and every non-atomic self similar measure , the measure is supported on numbers that are normal in base .
1. Introduction
1.1. Background and main result
Let be an integer greater or equal to . Let be the times- map,
A number is called -normal, or normal in base , if its orbit equidistributes for the Lebesgue measure on . In 1909 Borel proved that Lebesgue almost every 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 we define the -expansion of to be the lexicographically largest sequence such that . This sequence is obtained from the orbit of under the map similarly to the integer case. It is known that admits a unique absolutely continuous invariant measure, commonly referred to as the Parry measure [20]. When is an integer then the Parry measure is just the Lebesgue measure restricted to . Thus, in analogy with the integer case, we say that is -normal if it equidistributes under for the Parry measure. Our main result will allow us to conclude that for certain fractal measures, a typical element will be -normal, where may be a non-integer.
In this paper we will work with self-similar measures on . These are defined as follows: Let be a finite set of real non-singular contracting similarity maps of a compact interval . That is, for every we can write where and , and . We will refer to as a self similar IFS (Iterated Function System). It is well known that there exists a unique compact set such that
The set is called a self-similar set, and the attractor of the IFS . We always assume that there exist such that the fixed point of is not equal to the fixed point of . This ensures that is infinite.
Next, let be a strictly positive probability vector, that is, for all and . It is well known that there exists a unique Borel probability measure such that
The measure is called a self-similar measure, and is supported on . Our assumptions that is infinite and that for every are known to imply that is non-atomic. In particular, all self-similar measures in this paper are non-atomic.
Recall that is called a Pisot number if it is an algebraic integer whose algebraic conjugates are all of modulus strictly less than . Note that every integer larger than is a Pisot number. Also, we will write if the real numbers satisfy that . Otherwise, we write , in which case are said to be multiplicatively independent. Finally, given a measure on and we say that is pointwise -normal if almost every is -normal. We are now ready to state the main result of this paper:
Theorem 1.1.
Let be a Pisot number, and let be a non-atomic self-similar measure with respect to the self-similar IFS and a strictly positive probability vector. If there is some such that then is pointwise -normal. Furthermore, is pointwise -normal for all .
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 for all and some integer , and no element in the attractor is -normal - for example, no element in the middle-thirds Cantor set is -normal. Nonetheless, at least for integer , we will soon recall some recent results showing that even if all the contractions it is still possible that every self-similar measure is pointwise -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 as long as . This was later generalized by Feldman and Smorodinsky [8] to all non-degenerate Cantor-Lebesgue measures with respect to any base . We remark that the work of Host [12] and the subsequent works of Meiri [18] and Lindenstrauss [17] about pointwise normality for -invariant measures are also related to this, though in the somewhat different context of Furstenberg’s Conjecture.
In 2015, Hochman and Shmerkin [11, Theorem 1.4] proved Theorem 1.1 under the additional assumptions that for all , and the intervals , 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 -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 at 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 such that 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 -normal numbers even though the underlying IFS satisfies for all . 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 , if is a function satisfying then the Fourier transform of 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 for all , 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 be a self similar measure as in Theorem 1.1, and let be a Pisot number such that for some . We want to show that is pointwise -normal. The first and most critical step is to express 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 . 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 , 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 to conclude, via [11, Theorem 1.2], that almost every random measure is pointwise -normal. Since in the first step we disintegrated according to these measures, Theorem 1.1 follows. The assertion made in Theorem 1.1 about pushforwards of also follows along these lines, by noting that images of typical random measures will still be pointwise -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 be a finite set of iterated function systems of similarities , . We assume that each IFS is homogeneous and uniformly contracting. That is, for every there is some such that for every we have
We allow to equal for some , i.e. to have degenerate iterated function systems.
Let . Given a sequence we define the space of words of length (possibly with ) with respect to via
Next, for every we define subsets of via
Thus, for every we have a surjective coding map defined by
(1) |
Let denote the left-shift on (or other shift spaces), and note that for every the following dynamical self-similarity relation holds:
(2) |
We say that the model satisfies the Strong Separation Condition (SSC) if the union in (2) is disjoint for all . Here we adopt the convention that the union of a single set is considered to be disjoint.
Next, for each , let be a probability vector with strictly positive entries. On each we can then define the product measure . The projection of via the coding map is a Borel probability measure supported on . For every the measure also satisfies a dynamical self similarity relation, namely,
(3) |
Finally, let be a -invariant measure on . We refer to the triple as the model under consideration. We say that the model is non-trivial if is non-atomic for -a.e. . We say that the model is ergodic if the selection measure is an ergodic measure on . We say that the model is Bernoulli if the selection measure is a Bernoulli measure on .
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 be a non-atomic self similar measure with respect to an IFS , and let . Then there exists a non-trivial Bernoulli model with strong separation such that:
-
(1)
.
-
(2)
If for some then there exist such that .
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 denote the family of Borel probability measures on a metrizable space , and define
(4) |
For and , we define the translated and renormalized measure by
Here is a normalizing constant. Note that in the definition of , the measure does not need to be supported on (but in any case ).
For , we define the scaled measure by
where is again a normalizing constant.
The scaling flow is the Borel flow acting on . The scenery of at is the orbit of under , that is, the one parameter family of measures for . Thus, the scenery of the measure at some point is what one sees as one “zooms into” the measure with focal point .
Notice that . As is standard in this context, we shall refer to elements of as distributions (and denote them by capital letters such as ), and to elements of as measures (and denote them by Greek letters such as ). A measure generates a distribution at if the scenery at equidistributes for in , i.e. if
(Any limits involving measures are understood to be in the weak topology.) We say that generates , and call a uniformly scaling measure, if it generates at almost every . If generates , then is supported on and is -invariant [10, Theorem 1.7]. Moreover, 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 is trivial if it is the distribution supported on the atom - a fixed point of . We can now state the second main step towards the proof of Theorem 1.1:
Theorem 1.3.
Let be a non-trivial Bernoulli model with strong separation. Then there exists a non-trivial -ergodic distribution such that for -a.e. the measure generates .
In fact, we will derive an explicit expression for 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 , showing that it does not contain a non-zero integer multiple of . 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 be a self similar IFS. Fixing weights , we begin by constructing a model . We then show that it meets the requirements of Theorem 1.2.
2.1. Construction of the model
Write and let be a strictly positive probability vector on . For every and write , and let
Recall that we are assuming that the attractor is infinite. It is not hard to check (see [29, Proof of Lemma 4.2]) that there exist some and such that
where denotes the convex hull of a set . In addition, notice that , the self similar measure corresponding to and , is also a self similar measure with respect to the IFS , with weights
The contraction factors of include , so the hypothesis that some is preserved by this iteration. This shows that we may assume, without loss of generality, that there exists such that
(5) |
After relabeling, we may also assume . We will use (5) to show that the model satisfies the SSC.
We now define . For every index in we can associate an IFS as follows:
-
•
For we associate the IFS .
-
•
For every we associate the degenerate IFS .
Notice that the construction above gives us a dichotomy: Each IFS in is either degenerate, or is homogeneous with strong separation in the sense of (5).
Recall that was our fixed probability vector on . We now define a probability vector on as follows:
-
•
The mass gives to is , that is, .
-
•
For every , the singleton gets mass .
Recalling our notation from Section §2.1, let be the corresponding Bernoulli measure on . This will be our (Bernoulli) selection measure.
Next, for we define the probability vector
For every we define the (degenerate) probability vector
Recall from Section §1.2 that with these probability vectors, on each we can then define the product measure , and the projection of via the coding map is a Borel probability measure supported on .
2.2. Proof of the required properties
In this section we show that the model constructed in Section §2.1 satisfies the conclusion of Theorem 1.2. First, the model is Bernoulli by definition. Secondly, the union in (2) is disjoint for all : if there is nothing to do; otherwise, this follows from (5) and the fact that all sets are contained in the original attractor . Since among the we find integer powers of all the original , if there is such that , then there is also some such that .
We proceed to prove part (1) of Theorem 1.2: Let be the self similar measure that corresponds to the weights . We will show that
Since is the unique Borel probability measure satisfying this self-similarity relation, it will then follow that
We have
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 -a.e. , the measure is non-atomic. Indeed, by the SSC for any the coding map is injective. So by our choice of weights it follows that for every and , we have
(6) |
Since is a Bernoulli measure on with strictly positive weights, -a.s. the digit occurs in infinitely times. This shows that the right hand side of (6) tends to as for -a.e. . Since any point in is covered by sets for all , the measures are -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. for all ; this case avoids certain technicalities. We then consider the case when we have at least one satisfying .
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 is a non-degenerate Bernoulli measure. Without loss of generality, we may assume that the sets are at distance from each other for all . See the discussion in §3.3 on how to treat the general case.
First, we define the random pair , where is drawn according to and is drawn according to Let denote this distribution. In fact, is the Bernoulli measure on the symbols with weights . We let be the shift map on . In particular, is -invariant and ergodic. Write for simplicity; recall that this is the translation of that centers it at .
Claim 3.1.
We have: .
Proof.
The claim is essentially due to dynamical self-similarity and the SSC. Fix , , and let be a Borel subset. It follows from (1) that
(7) |
Write , and note that for a Borel set we have
Indeed, by our assumption that the distance between the sets is greater than , at most one of them may intersect (a set of diameter ). Continuing this line of reasoning, we have
where we used (7) in the last line. ∎
Claim 3.2.
For -a.e. , the measure generates the distribution that is defined as follows: let be the suspension measure for the suspension flow with base and roof function . That is, denoting Lebesgue measure by ,
Then is the push-forward of under .
Proof.
Our goal is to show that for almost all , the measure generates at , or in other words, the scenery equidistributes for . Fix and let
By Claim 3.1,
(8) |
By the ergodic theorem, for -a.e. ,
and
Multiplying these two identities, recalling (8), and using that is bounded away from and , we see that for -almost all and -almost all , we have
Finally, by the separability of , for -almost all and -almost all the equation above holds for every . This proves the claim. ∎
Proof of Theorem 1.3 in the orientation-preserving case.
First, by Claim 3.2, for -a.e. , the measure generates the distribution as in Claim 3.2. Furthermore, is non-trivial, since the model is non-trivial. Finally, we need to establish the ergodicity of . This follows since is ergodic, being the suspension of an ergodic (in fact, Bernoulli) measure, and the map taking to 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 is a non-trivial ergodic model. In this case 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 satisfying . This case is slightly more complicated because as we zoom in on we are not necessarily going to see . Because of the presence of orientation reversing maps, it is possible that as we zoom in on we may see the image of reflected around the origin. Our analogue of the map has to take this behaviour into account, and our distribution also has to see these reflected measures.
We again assume the conditions of Theorem 1.3 are satisfied and that the sets are at distance from each other. We let be as in the orientation preserving case, and recall that its support is denoted by . We define a map as follows:
We let denote the Haar measure on and let
Lemma 3.3.
is -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 and the probability of transitioning from to is if and or if and ; and the transition is forbidden otherwise. Then
is readily checked to be a stationary probability for the Markov chain. The associated Markov measure is supported on and satisfies
Since and determine for almost surely according to the transition rule, the projection provides a natural isomorphism between and .
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 , and hence , is ergodic. ∎
To each element of we associate a measure as follows: Let be a Borel set, then
Claim 3.4.
We have: .
Proof.
The claim follows from a case analysis based upon the sign of and the value of . Note that the case where and 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 and . Let be a Borel set. Then, using the separation between the sets , and denoting , we have
∎
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 is defined as the push-forward of under the factor map , where is the suspension measure over with roof function .
3.3. Removing the distance assumption
Finally, we discuss the general case when the smallest distance between the sets is not larger than (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 is smaller than the smallest distance between the sets , to condition our measures not on but rather on 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 be a measurable (semi)flow on a space , and let be -invariant. Let . The pure point spectrum of is the set of all the for which there exists a non-zero measurable eigenfunction such that for every , on a set of full measure. If is -ergodic, then is constant for any eigenfunction ; in particular, any such measurable eigenfunction is in fact in . We also note that if is a factor of then : indeed, if is the factor map, then any eigenfunction on gives rise to the eigenfunction (with the same eigenvalue) on .
We now specialize to the scenery flow; recall Section §1.2. In this case, the existence of an eigenvalue indicates that some non-trivial feature of the measures generated by repeats periodically under magnification by . 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 is called trivial if it is the distribution supported on .
Theorem 4.1.
Let be a measure generating a non-trivial -ergodic distribution , and let be a Pisot number. If does not contain a nonzero integer multiple of , then is pointwise -normal. Furthermore, the same is true for for all .
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 be a subshift of finite type endowed with a Markov measure , let be a non-negative measurable roof function, and let denote the corresponding suspension flow. If , then there is such that
(9) |
With these ingredients, we can conclude the proof of theorem 1.1.
Proof of Theorem 1.1.
Let be a self similar measure that satisfies the condition of Theorem 1.1 with respect to a Pisot number , and let . First, apply Theorem 1.2 to obtain a disintegration of according to the corresponding model . That is,
It follows that
Therefore, to show that is pointwise -normal it is enough to show that -almost surely the measure is pointwise -normal.
By Theorem 1.2 there is some such that . Also, by Theorem 1.3, for -a.e. the measure generates the non-trivial -ergodic distribution , constructed in Section §3. Moreover, 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 in the orientation-preserving case (in this case the Markov measure is actually Bernoulli), and when at least one IFS in the model is orientation-reversing. In both cases, the roof function is where corresponds to the current state of the Markov chain. Applying (9) to the fixed point in the orientation-preserving case, or applying (9) twice to the periodic point otherwise (which is allowed since is continuous), we deduce that .
Since , we see that cannot be a non-zero integer multiple of . As we discussed at the beginning of this section, the same holds for the factor . 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. 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. -adic foliation and equidistribution. Israel J. Math., 122:29–42, 2001.
- [18] David Meiri. Entropy and uniform distribution of orbits in . 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 -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 . 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]