Polynomial Fourier decay and a cocycle version of Dolgopyat’s method for self conformal measures
Abstract
We show that every self conformal measure with respect to a IFS has polynomial Fourier decay under some mild and natural non-linearity conditions. In particular, every such measure has polynomial decay if is and contains a non-affine map.
A key ingredient in our argument is a cocycle version of Dolgopyat’s method, that does not require the cylinder covering of the attractor to be a Markov partition. It is used to obtain spectral gap-type estimates for the transfer operator, which in turn imply a renewal theorem with an exponential error term in the spirit of Li (2022).
1 Introduction
1.1 Background and main results
Let be a Borel probability measure on . For every the Fourier transform of at is defined by
The measure is called a Rajchman measure if as . By the Riemann-Lebesgue Lemma, if is absolutely continuous then it is Rajchman. On the other hand, by Wiener’s Lemma if has an atom then it is not Rajchman. For measures that are both continuous (no atoms) and singular, determining whether or not is a Rajchman measure may be a challenging problem even for well structured measures. The Rajchman property has various geometric consequences on the measure and its support, e.g. regarding the uniqueness problem [36].
Further information about the rate of decay of has even stronger geometric consequences. For example, by a classical Theorem of Davenport-Erdős-LeVeque [20], if decays at a logarithmic rate then -a.e. point is normal to all integer bases (see also [40]). Wide ranging geometric information can be derived if decays at a polynomial rate, that is, if there exists some such that
For example, by a result of Shmerkin [44], it implies that the convolution of with any measure of dimension is absolutely continuous. We refer to Mattila’s recent book [37] for various further applications of Fourier decay in geometric measure theory and related fields.
The goal of this paper is to prove that every measure in a fundamental class of fractal measures has polynomial Fourier decay, as long as it satisfies some very mild non-linearity conditions; Furthermore, these conditions can be easily verified in concrete examples. To define this class of measures, let be a finite set of strict contractions of a compact interval (an IFS - Iterated Function System), such that every is differentiable. We say that is smooth if every is at least smooth for some . It is well known that there exists a unique compact set such that
(1) |
The set is called the attractor of the IFS . We always assume that there exist such that the fixed point of does not equal the fixed point of . This ensures that is infinite. We call uniformly contracting if
Next, writing , for every and let
Fix . Then we have a surjective coding map given by
(2) |
which is well defined because of uniform contraction (see e.g. [13, Section 2.1]).
Let be a strictly positive probability vector, that is, for all and , and let be the corresponding Bernoulli measure on . We call the measure on the self conformal measure corresponding to , and note that our assumptions are known to imply that it is non-atomic. Equivalently, is the unique Borel probability measure on such that
When all the maps in are affine we call a self-similar IFS and a self-similar measure.
Next, we say that a IFS is linear if for every and . This notion was introduced in our previous work [3], though it implicitly appeared in the literature prior to that, notably in the work of Hochman-Shmerkin [28]. It is clear that if is and linear then it must be self-similar. While we believe such IFSs should exist, we are not aware of any known example of a linear smooth IFS that is not self-similar for . A major source of examples of non-linear IFSs arise in smooth dynamics from certain homoclinic intersections, see e.g. [21, Section I].
We can now state the main result of this paper. We say that a IFS is conjugate to an IFS if there is a diffeomorphism such that .
Theorem 1.1.
Let be a uniformly contracting IFS where . If is not conjugate to a linear IFS then every non-atomic self-conformal measure admits some such that
Moreover, our argument yields easily verifiable conditions when Theorem 1.1 may be applied (without directly checking whether the IFS is conjugate to linear); As we discuss later, it suffices for a IFS to satisfy conditions (10) and (11) below for the conclusion of Theorem 1.1 to hold true.
A great deal of attention has been given to the case when IFS in question is real analytic. This is mainly because such IFSs arise naturally in number theory, e.g. as (finitely many) inverse branches of the Gauss map [29], as Furstenberg measures for some cocycles [52, 6], and are closely related to Patterson-Sullivan measures on limit sets of some Schottky groups [35, 39, 14], among others. Combining Theorem 1.1 with a new method from a paper of Algom et al. [2], we can derive the following Corollary regarding Fourier decay in this setting:
Corollary 1.2.
Let be a IFS. If contains a non-affine map then every non-atomic self-conformal measure admits some such that
We emphasize that no separation conditions are imposed on the IFS in both Theorem 1.1 and Corollary 1.2. Thus, as we discuss below, these are essentially the first instances where polynomial decay is obtained for such measures without an underlying Markov partition, or a more specialized algebraic or dynamical setup. We remark that, simultaneously and independently of our work, Baker and Sahlsten [9] obtained similar results but with a different method. See Remark 1.3 for more details. Also, we note that Baker and Banaji [8] independently obtained a proof of Corollary 1.2, with a method that differs considerably from that of [2]. We refer to [2, 8] for more discussion and comparisons between the two techniques.
Relying on his own previous work [33] and on the work of Bourgain-Dyatlov [14], Li [34] proved polynomial Fourier decay for Furstenberg measures for cocycles under mild assumptions. Corollary 1.2 gives a new proof of these results when the measure is self-conformal (see [52, 6] for conditions that ensure this happens). We remark that, as we discuss in Section 1.2 below, the renewal theoretic parts of our argument are closely related to Li [33, 34]. Sahlsten-Stevens [43, Theorem 1.1] proved polynomial Fourier decay for a class of stationary measures that includes self-conformal measures, with respect to totally non-linear IFSs with strong separation (i.e. the union (1) is disjoint). Their methods also apply when the IFS is only under the additional assumptions that and some further separation conditions on . For self-conformal measures, Theorem 1.1 improves these results since we do not require any separation conditions on the IFS, or that the attractor is an interval. In fact, Corollary 1.2 removes virtually all the assumptions made in the corresponding result of Sahlsten-Stevens [43, Theorem 1.1], except for the existence of an analytic non-affine map in the IFS. Theorem 1.1 also improves our previous result [3, Theorem 1.1] by upgrading the rate of decay in the non-conjugate to linear setting from logarithmic to polynomial. Finally, we point out that when the IFS in question is conjugate to self-similar via a non-linear map, Kaufman [32] (for some Bernoulli convolutions) and later Mosquera-Shmerkin [38] (for homogeneous self-similar measures) proved polynomial Fourier decay for all self-conformal measures. The only concrete examples of polynomial Fourier decay in the fully self-similar setup were given by Dai-Feng-Wang [19] and Streck [50] for some homogeneous IFSs that enjoy nice number theoretic properties. It is known, though, that very few self-similar IFSs should fail this property; See Solomyak [46].
Finally, combining Corollary 1.2 with the recent works of Brémont [16] and Li-Sahlsten [36] we obtain the following complete characterization of the Rajchman property for IFSs: Let be a IFS. If admits a non-atomic self-conformal measure that is not Rajchman, then is self-similar and there exists a Pisot number such that for some . Furthermore, is affinely conjugated to an IFS that has all of its translations in . For more recent results on Fourier decay for self similar measures we refer to [46, 41, 51, 4, 38, 17, 19, 18, 50] and references therein.
1.2 Outline of proof: A cocycle version of Dolgopyat’s method
We proceed to discuss the method of proof of Theorem 1.1. There is no loss of generality in assuming our IFS is . Fix a self-conformal measure , and assume the conditions as in Theorem 1.1 are met. We aim to show that has polynomial Fourier decay. Our proof consists of four steps:
The first and most involved step is arguing that the transfer operator corresponding to the derivative cocycle and (Definition 2.3) satisfies a spectral gap-type estimate (Theorem 2.8). Both our result and method of proof are strongly related to the works of Dolgopyat [22, 24], Naud [39], and Stoyanov [49, 48] (see also [5, 7, 10]) . These papers utilize a proof technique that originates from the work of Dolgopyat [22], commonly known as Dolgopyat’s method. It is used to obtain spectral gap-type estimates by, roughly speaking, a reduction to an contraction estimate (in the spirit of Proposition 2.9), and then a proof of this estimate via the construction of so-called Dolgopyat operators (in the spirit of Lemma 2.12), that control cancellations of the transfer operator.
There are, among others, three critical properties of and that are used in Naud’s version [39] of Dolgopyat’s method, which is the one we ultimately rely on here: That the union (1) corresponds to a Markov partition (i.e. for , and may intersect only at their endpoints), that the measure satisfies the Federer property (see e.g. Theorem 2.4 Part (6)), and that satisfies the uniform non-integrability condition (UNI in short - see Claim 2.2). In our setting, since we assume no separation conditions (in particular, the cylinder covering is not a Markov partition), we cannot assume has the Federer property. As for (UNI), in our previous work [3, proof of Claim 2.13] we showed that if is not conjugate to linear then (UNI) does hold true (regardless of any separation assumptions). In fact, our non-linearity assumption is only used to obtain the (UNI) condition.
One of the main innovations of this paper is the introduction a cocycle version of Dolgopyat’s method, that we use get around these issues: First, we express as an integral over a certain family of random measures, and express the transfer operator as a corresponding convex combination of smaller ”pieces” of itself. This is based upon a technique that first appeared in [26], and was subsequently applied in several other papers. e.g. [30, 42, 47]. In our variant, that is closest to the construction in [1], these random measures typically satisfy a certain dynamical self-conformality relation with strong separation, and satisfy the Federer property. Critically, we are able to preserve the (UNI) condition into all the different pieces of the transfer operator corresponding to this decomposition. An important preliminary step is the construction of a special covering of our IFS by (possibly overlapping) sub-IFSs (Claim 2.1), followed by the construction of this disintegration in Theorem 2.4. We then proceed to state and prove our spectral gap-type estimate (Theorem 2.8) by first reducing to a certain randomized contraction estimate (Proposition 2.9), and then prove this estimate via the construction of corresponding randomized Dolgopyat operators (Lemma 2.12). All of this discussion takes place in Section 2.
In the second step of our argument, we deduce from Theorem 2.8 (our spectral gap estimate) a renewal theorem with an exponential error term. This is Proposition 3.2 in Section 3. Such strong renewal Theorems were first proved by Li [34] for random walks arising from certain random matrix products. Our proof technique is based on that of Li [33, 34], and Boyer [15], relating the error term in the renewal operator to the resolvent of the transfer operator, on which we have good control thanks to Theorem 2.8.
In the third step of our proof, we deduce from Proposition 3.2 (our renewal Theorem) an effective equidistribution result for certain random walks driven by the derivative cocycle. This is Theorem 4.1, that critically holds with an exponential error term. The exponent we obtain is related to the size of the strip where we have spectral gap (the from Theorem 2.8). Finally, in Section 5 we obtain the desired Fourier decay bound on by combining Theorem 4.1 with our previous proof scheme from [4, Section 4.2], that relies on delicate linerization arguments and estimation of certain oscillatory integrals as in [27].
As for the proof of Corollary 1.2, our main ingredient is that the following dichotomy holds for any given IFS: Either it is not conjugate to linear, or it is conjugate to a self-similar IFS. This is Claim 6.1, that critically relies on the Poincaré-Siegel Theorem [31, Theorem 2.8.2]. Corollary 1.2 then follows from Theorem 1.1 (if the IFS is not conjugate to linear), or from a result of Algom et al. [2] (if the IFS is analytically conjugate to self-similar).
Remark 1.3.
Let us now revisit the parallel project of Baker and Sahlsten [9] where they obtain similar results to ours. Roughly speaking, they extend the arguments of Sahlsten-Stevens [43] that rely on the additive combinatorial approach of Bourgain and Dyatlov [14]. This is a significant difference between the two papers, as we make no use of additive combinatorics in our work. Baker and Sahlsten [9] also require a spectral gap type estimate for the transfer operator without an underlying Markov partition. To this end, they utilize a disintegration technique as in [1], and combine it with the proof outline of Naud [39]. This has some similarities to our corresponding argument. However, even in this part there are some essential differences; We invite the interested reader to compare our Section 2 with [9, Section 4].
2 From non-linearity to spectral gap
The main goal of this Section is to establish the spectral gap-type estimate Theorem 2.8. To do this, we construct our random model in Theorem 2.4 in Section 2.3, based on Claim 2.1 in Section 2.1. This construction requires moving to an induced IFS of higher generation, but standard considerations show that this is allowed in our setting. We then proceed to state Theorem 2.8, and then reduce it to a randomized -contraction estimate Proposition 2.9, in Section 2.4. We then reduce Proposition 2.9 to the key Lemma 2.12 in Section 2.5, and spend the rest of the Section proving it. The proof of Lemma 2.12 can be seen as a randomized version of Naud’s arguments in [39, Sections 5, 6, and 7].
2.1 An induced IFS
Fix a IFS and write . We assume without the loss of generality that for every the map is a self map of , and that
(3) |
Indeed, this follows since we may assume the IFS acts on an open interval, and since Fourier decay estimates like in Theorem 1.1 are invariant under conjugation of the IFS by a non-singular affine map. Let
and let
Let us recall the bounded distortion property [4, Lemma 2.1]: There is some such that for every we have
(4) |
Next, note that
(5) |
So, there is a uniform constant , such that, assuming as we may (see the next paragraph) that ,
(6) |
For all let be the corresponding induced IFS of generation , that is,
Since is finite, we may order its functions ( and so on) at our will. Note that the same constants from (4) and from (6) will work for the induced IFS for all .
Our first step is, assuming is not conjugate to linear, to construct an induced IFS that has useful separation and non-linearity properties:
Claim 2.1.
Let be a IFS that is not conjugate to linear. Then there exists such that admits satisfying:
-
1.
For every there exists such that
is a disjoint union. In particular, the union is disjoint for .
-
2.
There exists such that: For every , for both ,
-
3.
We have
We first remark that the existence of an upper bound as in Part (2) holds for all and every ; This is a standard fact, see e.g. [39, Section 4] or (6) for similar estimates. We thus focus on the lower bound, which is much harder to obtain, in our proof of Claim 2.1.
We require the following Claim:
Claim 2.2.
There exist some such that for all there exist such that for every ,
Proof.
By [3, proof of Claim 2.13], if for all -periodic111Observe that for the proof of [3, Claim 2.13] to follow through it is enough to consider only -periodic elements. Furthermore, for such elements the limits in (7) exist by (6). We remark that similarly, in [3, Claim 2.12] and its proof it is enough to consider only -periodic elements. and we have
(7) |
then is conjugate to a linear IFS. Thus, since is not conjugate to linear, there exist some , and -periodic such that for infinitely many ,
Recalling our coding map (2), let be such that . Using -periodicity, by the uniform convergence ([3, Claim 2.12] or (6)) as of
there is some and some such that for every
Let , and let be as above. Then for all ,
So, for , the words and satisfy the claimed bound from below for all . ∎
Proof of Claim 2.1, Case 1: In this case we assume that, with the notations of Claim 2.2, that there exist arbitrarily large such that the as in Claim 2.2 satisfy
Note that there exist some and such that
For example, can be chosen to be the -prefixes of codes of the endpoints of the interval . Clearly every large enough will work. We thus assume that is large enough so that as in Claim 2.2, the first displayed equation holds, and
(8) |
.
We now define and our maps in via
Let us show that all parts of the Claim hold: First, for part (2), for all ,
Now, by Claim 2.2 since ,
On the other hand, arguing similarly to (6),
Since we may assume is also large depending on (that are both known a-priori), we conclude that
The same calculation holds for . This shows Part (2) holds true for our functions with
Part (3) essentially follows from the same argument since, assuming is sufficiently large depending on , we can get that also
For Part (1), for every and there is some with
This shows that
Next, by the choice of and it is clear that
Now, fix any with . If
is a disjoint union then we are done. Otherwise, suppose
Then, since and by (8),
So, since the unions and have already been shown to be disjoint, the union
is disjoint. The other cases are similar. The proof of Claim 2.1 in Case 1 is complete.
Proof of Claim 2.1, Case 2: Assume now that for all , for as in Claim 2.2,
Let us find some such that for all
(9) |
Such must exist since is infinite.
Thus, and are words that can be made arbitrarily long, that satisfy that for some and for all
Also, by (9),
Thus, we are back in Case 1 with these (that we recall can be made arbitrarily long) and .
Thus, we have completely reduced Case 2 to Case 1, completing the proof of Claim 2.1.
2.2 The Euclidean derivative cocycle and associated transfer operator
Fix a IFS as in Section 2.1, and let us retain the other assumptions and notations from that Section. Let be a strictly positive probability vector on , and let be the corresponding product measure on . Let be the corresponding self-conformal measure.
Note that for every the induced IFS has the same attractor as . That is, . Furthermore, is also a self-conformal measure with respect and the induced probability vector on .
Thus, by working with the induced IFS as in Claim 2.1, we may assume without the loss of generality that our (not conjugate to linear) original IFS already admits that satisfy:
-
1.
For every there exists such that
(10) is a disjoint union. In particular, the union is disjoint for .
-
2.
There exists such that: For every , for both ,
(11)
Note that for the second part of (11) we use Claim 2.1 Part (3) and that, with the notation as in Section 2.1,
Let to be the free semigroup generated by the family , which acts on by composing the corresponding ’s. We define the derivative cocycle via
(12) |
Slightly adjusting our notation from Section 2.1, let
Also, we note that by uniform contraction
(13) |
A major part in our argument is played by the transfer operator:
Definition 2.3.
For every such that is small enough, let denote the transfer operator defined by, for and ,
Note that is the unique stationary measure corresponding to the measure on . So, the detailed discussion about this operator as in [12, Section 11.5] applies in the setting we are considering.
2.3 Disintegration of and the transfer operator
We can now construct the random measures we discussed in Section 1.2. We begin by recalling the notion of a model (as in e.g. [45]): Let be a finite set of iterated function systems . Let . For and let
Define a coding map via
Next, for each let be a probability vector with strictly positive entries. On each we define the product measure
We can now define
Letting be the left shift, we have the following dynamical self-conformality relation:
(14) |
We also define
and note that here we have
Next, for , and we define an operator by
(15) |
Note that we only need to know in order for to be well defined. Iterating (14) we have the following equivariance relations, whose proof is left to the reader: First, for every , , and ,
(16) |
Furthermore, for all integers we have that
(17) |
Finally, let be a -invariant measure on . The triplet is called a model. We say the model is non-trivial if is non-atomic for -a.e. . We say it is Bernoulli if is a Bernoulli measure. We are now ready to state the main result of this Section:
Theorem 2.4.
Let be a IFS satisfying properties (10) and (11), and let be a self-conformal measure. Then there exists a non-trivial Bernoulli model such that:
-
1.
(Disintegration of measure) .
-
2.
(Disintegration of operator) For every and we have
(18) -
3.
(Non-trivial branching) For every we have or .
-
4.
(Separation) For every the union
is disjoint.
-
5.
(UNI in all parts) There exist and such that for all , for every there exist such that
.
-
6.
(Federer property) For every there exists such that:
For every , for every and ,
Recall that our IFS meets the conditions of Theorem 2.4 courtesy of Claim 2.1. The proof of Theorem 2.4 is given in the next three subsections.
2.3.1 Construction of the model
Recall that we are assuming conditions (10) and (11) hold for . For we define the IFSs
For every define the IFS
as in (10).
We now define
Thus, for every we associate the IFS .
Note that certain appear in multiple IFS’s in . Write, for ,
Recall that where . We now define a probability vector on as follows:
Let . This will be our Bernoulli selection measure on .
Next, for every we define the probability vector
Our model is now fully defined.
2.3.2 Proof of Parts (1)-(5) of Theorem 2.4
Proof of Part (1) We first argue that
The proof is almost entirely the same as [1, Section 2.2], so we omit the details. Part (1) now follows since is the unique measure satisfying this identity.
Proof of Parts (2) and (3) These are straightforward given our construction. Indeed, for part (2), when it suffices to note that for every ,
For general similar considerations apply.
Proof of Part (4) This follows directly from our construction and from Claim 2.1 part (1).
Proof of Part (5) Let and let . Choose any . By (11) and our construction, there exist such that for some with
Let , and put . Then for all ,
By arguing similarly to (6),
By the choice of we conclude that
where the last inequality is due to the second part of (11).
2.3.3 Proof of Part (6) of Theorem 2.4
This part of the proof is modelled after Naud’s work in [39, Section 6]. Similarly to Naud, we require the following Lemmas and definitions, that will also be used elsewhere in this note. However, unlike Naud [39], the Federer property we establish is for the random measures in our model. Thus, we make sure that all our estimates are uniform in (i.e. depend only on the model and not the measure under consideration).
Definition 2.5.
Fix and . The cylinder that corresponds to is the set
From now on we fix the in question and suppress it in our notation. Also, note that by the definition of , (14), and Theorem 2.4 Part (4),
(19) |
For a cylinder set let denote its diameter. Recall that we are fixing some , and considering cylinders with respect to .
Lemma 2.6.
There exist constants and uniform in such that for all cylinders
Proof.
Write where . Then, omitting superscripts, for some
and for some we have
Letting be as in (4), recalling that is a composition of maps from ,
as claimed. ∎
Lemma 2.7.
For every cylinder we have
where . Moreover, there is some uniform in such that for any two distinct cylinders with and we have
Proof.
Proof of Theorem 2.4 Part (6) Fix , and . In general, we aim to show that there exist cylinders such that depends only on , and
If this holds then by (19):
Note that the latter bound depends on , and in particular is uniform in .
Set and . We first assume for some 1st generation cylinder of . Set
Then . Let be such that . Let with . By definition of , so there are two options:
-
1.
If then by Lemma 2.6 we have
-
2.
Otherwise, there exists a cylinder such that such that and are consecutive and . Indeed, is covered by such cylinders of generation and none of them are included in . Consider now a bigger cylinder such that , and assume that
The maximality of implies that and where . Since the gap between and is included in , by Lemma 2.7 we find that
We have just shown that there exists a cylinder such that and , where is independent of .
Finally, there exists a decreasing sequence of cylinders such that
By Lemma 2.6 and the estimate on we have
So, whenever we have . This is the same as asking that
Hence there is a cylinder as required, such that and only depends on .
Finally, we discuss the case when is not included in a first generation cylinder. Note that for some . If then and following the same ideas we find and with only depending on . So,
where only depends on (in particular, does not depend on ). If then by a similar gap argument to the one previously used,
Noting that the latter constant can be bounded above uniformly in terms of the model, the proof is concluded in the same way.
2.4 Spectral gap and reduction to an contraction estimate
We equip with the norm
(20) |
Following Dolgopyat [23, Section 6] and Naud [39, the discussion prior to Lemma 5.2], for every we define yet another norm on via
(21) |
These two norms on are clearly equivalent.
Recall the notations and definitions from Section 2.2. The main goal of this (entire) Section is to prove the following spectral gap-type estimate:
Theorem 2.8.
Recall that by Claim 2.1 every not conjugate to linear IFS admits an induced IFS satisfying the conditions of Theorem 2.8.
We will work with the model constructed in Theorem 2.4. Also, recall the definition of the operators from (15). We first reduce Theorem 2.8 to the following statement about the -contraction of all parts of the transfer operator.
Proposition 2.9.
Assume the conditions of Theorem 2.8, and let be the model from Theorem 2.4. Then there is some and such that for with small enough and large enough:
For every ,
for all with .
This is a randomized analogue of [39, Proposition 5.3] in Naud’s work.
In the reminder of this Section, we reduce Theorem 2.8 to Proposition 2.9. Our argument is roughly based on Naud’s corresponding argument [39, Section 5], with some significant variations due to our model construction. To this end, we require the following two Lemmas. For let denote the Lipschitz constant of .
Lemma 2.10.
[13, Proof of Theorem 2.1.1 part (ii)] There exists a constant such that for all , for every and ,
Proof.
The operator
contracts the space with the dual Lipschitz metric by a factor of . Indeed, the IFS satisfies that . The Lemma now follows by noting that , taking to be the diameter of . ∎
Recall that is as in (13).
Lemma 2.11.
(A-priori bounds) There exists such that for all small enough and large enough, for all , writing ,
In particular, there exists a constant such that for all and
Proof.
Proof that Proposition 2.9 implies Theorem 2.8 Fix and as in Proposition 2.9. Set
For all and with and all we have, by (18),
(23) |
Now, fix . Then, by (17),
Set . Note that
Then by Cauchy-Schwartz,
Via Lemma 2.10 applied with , (16), and the previous calculation we have
Applying Proposition 2.9,
where are positive constants. Recalling that is already assumed to be small, we can now choose with and even closer to so that if is small enough and is large enough,
for some . Using that the same bound works for every term in the convex combination (23), we obtain
Next, by (18)
(24) |
So, since (as ) we have, as in Lemma 2.11, for some
Thus,
So, using similar ideas, for small and large and possibly a large ,
for some and .
Finally, given we write
Let be such that for all small enough . Applying Lemma 2.11,
for some . This is true for all with small, and since
Theorem 2.3 follows.
2.5 The key Lemma
Given let
Note: If and then
We will prove the following variant of Naud’s result [39, Lemma 5.4], that is the key to the proof of Proposition 2.9:
Lemma 2.12.
There exist and such that for all with small and large:
For every there exist a finite set of bounded operators on such that:
-
1.
The cone is stable under for all .
-
2.
For all and ,
-
3.
Let and be such that and . Then for every there exists such that
2.6 The triple intersections property
The following Proposition allows for the construction, for every , of a special partition of that has the triple intersections property on : This means that whenever a cell intersects , two other nearby cells must also intersect . It is based upon [39, Proposition 5.6], but as usual there are significant variation due to our model setting. Thus, while the partitions themselves depend on , certain metric features of them, e.g. the size of the cells and the distance of the endpoints from , are uniform across our model .
Proposition 2.13.
There exist constants such that for all and every small enough:
There exists a finite collection of closed intervals ordered along such that:
-
1.
, and .
-
2.
For all we have .
-
3.
For all such that , either and , or and , or and .
-
4.
For all such that , we have
It is critical to our argument is that the constants may be chosen uniformly across the model.
The proof of Proposition 2.13, that we discuss now, is roughly modelled after Naud’s arguments in [39, Section 7]. Fix and recall Definition 2.5 (cylinders of ). We require the following Lemmas:
Lemma 2.14.
There exists a constant such that for all and all , there exists a cylinder such that
Proof.
Consider a cylinder with , and assume is minimal. Since such a cylinder exists. Let with . By minimality of we cannot have included in , and thus . Via Lemma 2.6 we obtain
as claimed. ∎
Lemma 2.15.
Let be a cylinder. Then there a finite set of at least words such that
where and .
Proof.
By definition, there is some such that
Furthermore, by Theorem 2.4 Part (4)
Now, by the definition of the model,
and by Theorem 2.4 Part (3). Note that we are also using Theorem 2.4 Part (4) to see that does not have exact overlaps (maps with different coding are not equal). Thus,
which is a union of cylinders that are contained in . As required. ∎
Remark 2.16.
Proof of Proposition 2.13 Let . Set . First, we divide into closed intervals with disjoint interiors such that
For every write . We may assume and .
We first deal with part (4): For , if we set . Otherwise, let . Applying Lemma 2.14, we find a cylinder with . By Lemma 2.15, there are consecutive cylinders such that
Set
In both cases
As for the boundary points , by (3)
So, upon taking , we may put and .
For all set . Then, writing we have
Furthermore, the intervals still have disjoint interiors. Thus, this collection of intervals satisfies parts (1),(2), and (4) of the Proposition. Let us call these intervals .
2.7 Proof of Lemma 2.12
Fix and let . We begin by constructing the Dolgopyat operators as in Lemma 2.12. Let be sufficiently large in the sense of Theorem 2.4 part (5), and in other ways that will be specified soon, and let be the length words satisfying the conclusion of Theorem 2.4 Part (5). Let us fix uniformly in , that are small enough (to be determined later). Let be a triadic partition as in Proposition 2.13 of of modulus . For all set
By Proposition 2.13 part (4)
Then there exists a cut off function such that on , on , and outside of . Then there exists that depends only on the previous (uniform in ) constants such that
Define
Fix to be determined later. Let . Define a function by
Note that is well defined by the separation property in Theorem 2.4 Part (4), and since the ’s intersect potentially only at their endpoints by Proposition 2.13, where all the vanish.
We can now define the Dolgopyat operators on by:
We proceed to prove the three assertions of Lemma 2.12.
2.7.1 Part 1: construction of an invariant cone
We follow the same notations of the construction carried out in the previous section, and prove Lemma 2.12 Part (1):
Lemma 2.17.
There exist , and such that for with sufficiently small and sufficiently large, for every ,
-
1.
The cone is stable under every .
-
2.
If and satisfy
then
-
3.
If then .
Our proof is roughly based on [39, proof of equation (4)]:
Proof.
Fix where is yet to be determined. Then for all ,
Now, for every , by separation (Theorem 2.4 Part (4)) we have
Also, we can find a constant uniform in , and such that if is small enough then
Therefore,
Using that we obtain
assuming is large enough, is small enough, and
Note that the above calculation works for any . Now, if and satisfy
then
where is independent of and is large enough. Note that when and so in this case . So, under this assumption
as long as and .
We thus fix and take large enough so that and fix . A similar argument shows that this choice of parameters also yields Part (3).
∎
2.7.2 Part 2: contraction of the cones
We proceed to prove that the operators contract these cones in the norms. We require the following Definition:
Definition 2.18.
A subset is called dense if for every such that there exists such that:
For a dense subset we write
We need the following key Lemma, a variant of [39, Lemma 5.7]:
Lemma 2.19.
Let be dense and fix . Then there exists independent of and such that
Proof.
Let
and note that . For every , by density of , there exists some index such that for some such that . We thus get a function such that for every the set contains at most elements.
Let . For every fix some . By Proposition 2.13,
Moreover, by Proposition 2.13 part (4), for and some carefully chosen we have
Note that
So, by the Federer property of proved in Theorem 2.4,
where we note that does not depend on .
Now, let . Then, as long as is large enough,
Since the constant are uniform (in particular, in and ), the proof is complete. ∎
Definition 2.20.
We define
We can now prove the required contraction property in Lemma 2.12 Part (2).
Proposition 2.21.
There exists uniform in such that for all with small and large, for all and all we have
This is a randomized analogue of [39, Proposition 5.9].
Proof.
Let . For every , by the Cauchy-Schwartz inequality and the definition of ,
is bounded above by the product of
(25) |
and
For all there exists such that
Recall that for every we have by (13)
Therefore, recalling the notations from Section 2.3, if then the sum in (25) is bounded by
Now, by Lemma 2.17 since . Therefore, applying Lemma 2.19 to this function we obtain
Recall that
So, if is small enough there is some such that
Therefore
as claimed.
∎
2.7.3 Part 3: Domination of the Dolgopyat operators
We now turn to Part (3) of Lemma 2.12. First we need the following key Lemma:
Lemma 2.22.
Let be such that
For every define functions via
Then for and small enough, and for all small enough and large enough, for all such that there exist such that and and some such that: For every ,
This is a model version of [39, Lemma 5.10]. For the proof, we require the following basic Lemmas from [39]:
Lemma 2.23.
In the following arguments, for let be the unique real number such that .
Lemma 2.24.
Proof of Lemma 2.22 First, we choose small enough so that the conclusion of Lemma 2.23 is true for all , and one checks that this does not change and (see the end of the proof of Lemma 2.17, and recall that are contractions). It is clear that . We add the assumption that
Let be a triad of intervals such that each of them intersects . Write
If for some we have for all then for all , and we are done.
Otherwise, by Lemma 2.23 we have for all and every ,
We aim to make use of Lemma 2.24. For every set
Let
We claim that for ,
Indeed,
If for some we have then for all we obtain
and so . If for all then
We next control the relative variations in the arguments of . Since
there exist two functions such that
For one possible construction see [39, Proof of Lemma 5.10]. For all set
Then for
Now, by our assumptions on
and so by Theorem 2.4 Part (5) for all
Let . By the mean value Theorem
So, if is large enough then there are constants such that, independently of and ,
We now pick so that
and put .
Suppose, towards a contradiction, that there are and such that both
Since we cannot have
with . Indeed, it will imply
a contradiction. But then
which is also a contradiction.
We conclude that there exists such that for all ,
Since , the conditions of Lemma 2.24 are met. Thus, either for every
or for every
depending on whether
Choosing , we have for all and some .
Proof of Lemma 2.12 Part (3) Fix constants so that parts (1) and (2) of Lemma 2.12, and so that Lemma 2.22, all hold true. Let be such that
We aim to show that there exists a dense subset such that
Since the latter statement holds true for all by Lemma 2.17, we focus on the first one.
Now, suppose for some .
Case 1 If but then by (26) for all : Indeed, (26) implies that since . Since we find that
The case but is similar.
Case 2 Suppose now both and . Then both . Taking half of the sum of these inequalities, we deduce that
Since for every with we have , it follows that, similarly to case 1,
The proof is complete.
3 From spectral gap to a renewal theorem with an exponential error term
We keep our notations and assumptions from Section 2. In particular, we are working with the induced IFS from Claim 2.1, so that Theorem 2.8 holds as stated. However, from this point forward we will no longer require the model constructed in Theorem 2.4. Recall also that is , and that is a self-conformal measure such that is a strictly positive probability vector. Recall that is our Bernoulli measure on .
First, we record the following consequence of Theorem 2.8; More precisely, only the third part follows from Theorem 2.8, the rest are already well known in our setting. For every let
Theorem 3.1.
Note that in [33, Section 4] the transfer operator has an additional minus in the exponent, which explains the left hand side of the equation in Part (2) in our setting. Part (3) was treated by Li in [34] in a related setting to ours (cocycles that arise from actions of algebraic groups on certain projective spaces). It follows from the analyticity of when for the as in Theorem 2.8, and otherwise from a direct application of Theorem 2.8 via the identity
We refer to the discussion in [34, Proposition 4.26] for more details.
We now define a renewal operator as follows: For a non-negative bounded function on and , define
Since is positive, this sum is well defined. For every let be the function
We also define, for the Fourier transform
The following Proposition is the main result of this Section:
Proposition 3.2.
Let be as in Theorem 2.8, and let be a compact set. Fix a non-negative bounded and continuous function on such that for every , and for every fixed . Assume
Then for every and
where is the supremum of the absolute value of the elements of .
We remark that Proposition 3.2 is an IFS-type analogue of the renewal Theorem of Li [34, Theorem 1.1 and Proposition 4.27]. Similarly to Li, we derive it from our spectral gap result Theorem 2.8 via Theorem 3.1.
Proof.
Indeed, for write
and put . Note that
By the monotone convergence Theorem, since and we have
Thus,
(28) |
Recall that for every . So, using the inverse Fourier transform, we have
(29) |
We now argue that the sum in (29) is absolutely convergent: Since is compactly supported for every , . Note that for every , , , and ,
So, plugging in and ,
(30) |
Thus, as long as , putting ,
It follows that the sum in (29) is absolutely convergent. Thus, we can use Fubini’s Theorem to change the order of integration. Since for every , we can apply Theorem 3.1 to obtain
Since for , and since for all by (30), we have
When , since is integrable with respect to , this converges to
by monotone convergence. Taking tally of our computations, (27) is proved.
Furthermore, using the bound on the norm of from Theorem 3.1 and (30), by dominated convergence
Define
Then is a tempered distribution with an analytic continuation to by Theorem 2.8 and Theorem 3.1, such that for any we have . Furthermore, by Theorem 3.1 and (30), for any , (see the computation below). Applying [34, Lemma 4.28] we have, for all
Finally, by (30) and Theorem 3.1,
where in the last equality we used that . Via (27) and the preceding paragraph, the proof is complete.
∎
4 From the renewal Theorem to equidistribution
We keep our notations and assumptions from Section 2. In particular, we are working with the induced IFS from Claim 2.1 so that Theorem 2.8, and therefore Proposition 3.2, hold as stated. Recall also that is , and that is a self-conformal measure such that is a strictly positive probability vector. Recall that is our Bernoulli measure on .
Recall the definition of the semigroup from Section 2.2. In this Section we discuss an effective equidistribution result for a certain random walk, driven by a symbolic version the derivative cocycle : It is defined as
(31) |
We can now define a random walk as follows: Recalling that is the left shift, for every we define a function on via
(32) |
Let be the random variable
(33) |
For every integer we define
Let be the law of the random variable . Then for every , . By uniform contraction there exists as in (13), so . In particular, the support of is bounded away from . It is easy to see that for every and we have
Thus, in this sense is a random walk.
Next, we define a function on that resembles a stopping time: For let
We emphasize that is allowed to take non-integer values. We also recall that is the corresponding Lyapunov exponent. Recalling (13), it is clear that for every and we have
We also define a local norm on by, for ,
(34) |
The following Theorem is an effective equidistribution result for the random variables :
Theorem 4.1.
Let be as in Proposition 3.2. Then for every , , and ,
Theorem 4.1 is a version of results of Li and Sahlsten [36, Propositions 2.1 and 2.2], that has better (exponential) error terms due to our non-linear setting. Indeed, Li-Sahlsten work with self-similar IFS’s, where necessarily has slower equidistribution rates (no faster than polynomial is possible). This also makes our treatment more complicated since the random walk is not IID as in their work. Another point of difference is that our renewal operator from Section 3, that is used critically for the proof of Theorem 4.1 via Proposition 3.2, is defined for functions on , rather than just on . Nonetheless, we will follow along the scheme of proof as in [36, Section 4], and deduce Theorem 4.1 from Proposition 3.2.
4.1 Regularity properties of renewal measures
Let be the even smooth bump function given by
Here is chosen so that is also a probability density function. Next, for write
The following Proposition is an analogue of [36, Proposition 4.5]:
Proposition 4.2.
There exists some such that, for our from Theorem 2.8:
For all , , , and such that ,
where the function is constant in (the coordinate).
Note that our error term is upgraded compared with [36, Proposition 4.5], due to the exponential error term in Proposition 3.2.
Proof.
If then contains either or . Therefore, as is even
So, as functions on ,
We proceed to bound , where we note that is a compactly supported function in . By Proposition 3.2, since our function is constant in the -coordinate,
Up to global multiplicative constants, the first and third terms are less than . For the second term, recall that
and so
Finally, is supported on . This yields the desired bound. ∎
We will also require the following Lemma from [36], that follows directly from Proposition 4.2 in our case:
Lemma 4.3.
[36, Lemma 4.6] There is some such that for all , , and we have
where the function is constant in the coordinate.
4.2 Residue process
Let be a non-negative bounded Borel function on . For and define the residue operator by
Recall that for every we let be the function
The following Proposition is an analogue of [36, Proposition 4.7]:
Proposition 4.4.
(Residue process) Let be a compact set. Fix a non-negative bounded, compactly supported, and continuous function on such that for every , and for every . Assume
Then for every and ,
Proof.
For and define a non-negative bounded Borel function
Then
Since is a IFS, the map for all , and therefore for all and . So, as our assumptions on imply that meets its conditions, the result is now a direct application of Proposition 3.2. ∎
4.3 Residue process with cut-off
Define a cutoff operator on real non-negative bounded Borel functions on via, for ,
We have the following analogue of [36, Lemma 4.9], which follows here from Lemma 4.3 (or Proposition 4.2):
Lemma 4.5.
There exists such that for all and we have
Proof.
By Lemma 4.5 the operator is well defined for bounded Borel functions. For a function on we denote, for all ,
We have the following analogue of [36, Proposition 4.10]:
Proposition 4.6.
Let be a compact set, and fix a bounded and continuous function on such that contains , for every , and for every . Assume
Then for and every we have
Remark 4.7.
By a standard decomposition into real and imaginary parts, and then each one into positive and negative parts, it suffices to prove Proposition 4.6 under the assumption that is non-negative (with the same parameters assumed in the original Proposition).
The following Lemma relates the operators and :
Fix small enough so that . We use to regularize these functions; Let
The following Lemma is an upgraded version of [36, Lemma 4.13], as it has an exponential error term:
Lemma 4.9.
Under the assumptions of Proposition 4.6,
Proof.
We first wish to apply Proposition 4.4 to the function . To this end, notice that by our assumptions for every the function , and satisfies the required integrability conditions. Also, since for every , it is clear that for every , the function
belongs to .
So, applying Proposition 4.4,
Also, for every
Now, since and then
It follows that for every we have
So,
which implies the Lemma. ∎
The following Lemma is based on [36, Lemma 4.15]. It is upgraded due to the fact that our bump function vanishes outside of .
Lemma 4.10.
Let be a function with and . Let
Then is bounded by:
-
•
if .
-
•
if or .
-
•
if or .
Proof.
If then, since is supported on
If then . Since for we obtain
In the second case, we use the trivial bound
In the third case, if then , and so
This gives the Lemma. ∎
Since , applying Lemma 4.10, is bounded by the sum of the following three terms (we sometimes omit the variable in the following computation):
-
•
if .
-
•
if or
-
•
if or .
By the definition of , the first term is smaller than
The third term is equal to
This gives us
By Lemma 4.3 the first term is dominated by
For the second term, since is a cocycle,
So, the second term is smaller than . By proposition 4.2, this is dominated by
The third term, by a change of the order of integration (as in [36, Proof of Proposition 4.10]), is smaller than
By Lemma 4.5, this is smaller than
This latter term is since is supported on .
Conclusion of proof Assuming is small enough, we have shown that
with the sum following error terms: From Lemma 4.9 we have
and from the previous argument above,
Note that, for every ,
Indeed, the second equality holds since exists except for at most two points, the third inequality is Young’s inequality, and the last one is trivial. A similar calculation shows that
Next, put
so that
and the error becomes
This completes the proof.
4.4 Proof of Theorem 4.1
First, let use relate our previous discussion to the symbolic setting outlined prior to Theorem 4.1: Recall that for and we defined a cutoff operator
We can define a symbolic analogue via, for ,
Then, by the definition of our cocycles (12) and (31)
Thus, the conclusion of Proposition 4.6 applies to with the same error terms.
Now, let . Let be a smooth cutoff function such that , and such that it becomes outside of . Let
Then when . By definition,
The function satisfies the conditions of Proposition 4.6, and so we can also apply this proposition to . The conclusion of Theorem 4.1 follows, with an error term of
Since , for every the function is supported on . So, we can take , and for all
Similarly, for all
Combining the previous three displayed equation, the Theorem is proved.
5 From equidistribution to Fourier decay
We keep our assumptions and notations from Section 2. In particular, we are working with the induced IFS from Claim 2.1 so that Theorem 2.8 holds. Therefore, Proposition 3.2 holds, and so we have Theorem 4.1 at our disposal. This Theorem will be the key to our arguments in this Section. Recall also that is , and that is a self-conformal measure such that is a strictly positive probability vector. Recall that is our Bernoulli measure on .
In this section we show that:
We will prove this under the additional (minor) assumption that is orientation preserving; The general case is similar, but notationally heavier.
Let be large and let to be chosen later. Let be as in Theorem 4.1. We define a stopping time by
We require the following Theorem from [4], that relates the random variables , (defined in Section 4), and . For any let denote the scaling map .
Theorem 5.1.
(Linearization) For any ,
Also, there exists a global constant such that for every
(35) |
Proof.
This is a combination of [4, Lemma 4.3, Lemma 4.4, and Claim 4.5]. ∎
Theorem 5.1 is the only place in the proof where we use our additional assumption that is orientation preserving. See [4, Corollary 4.6 and Remark 4.7] on how to remove this assumption.
Next, for every we define a measurable partition of via the relation
Writing for the expectation with respect to the conditional measure of on a cell corresponding to a -typical , it follows from Theorem 5.1 and the law of total probability that: For any ,
(36) |
And, for every ,
Now, for every fixed and , define the function
Then by the definition of the local norm on as in (34), assuming as ,
We emphasize that the bound above holds uniformly across . Applying Theorem 4.1 we obtain for a -typical
Finally, we use the following Lemma (originally due to Hochman [27]) to deal with the oscillatory integral above:
Lemma 5.2.
[4, Lemma 2.6] (Oscillatory integral) For every , , and
Note that we use (35) to get uniformity in the first term on the right hand side.
Conclusion of proof By the argument above we can bound by the sum of the following terms. Every term is bounded with implicit dependence on and the underlying IFS. For simplicity, we ignore global multiplicative constants so we omit the big- notation:
Linearization: For any prefixed ,
Equidistribution:
Oscillatory integral: For every
Choice of parameters For we choose that satisfies
We also choose and . Then we get:
Linearization:
Equidistribution:
Oscillatory integral: There is some such that
Here we made use of [25, Proposition 2.2], where it is shown that there is some such that for every small enough .
Finally, summing these error terms, we see that for some we have . Since as we have for some uniform , our claim follows.
6 On the proof of Corollary 1.2
In this Section we prove Corollary 1.2: All self-conformal measures with respect to a IFS that contains a non-affine map have polynomial Fourier decay: First, we show that any given IFS is either not conjugate to linear, or it is conjugate to a self-similar IFS (Claim 6.1). By this dichotomy, Corollary 1.2 follows from Theorem 1.1 and from a Fourier decay result about smooth images of self-similar measures from a paper of of Algom et al. [2].
6.1 On conjugate to linear real analytic IFSs
Recall that , a IFS, is called linear if the following holds:
In particular, if is and linear then it is self-similar. Recall that an IFS is called conjugate to if there is a diffeomorphsim between neighbourhoods of the corresponding attractors such that
The geometric properties of linear non self-similar smooth IFSs when are not well understand. In fact, even showing the existence of such IFSs is a highly non-trivial question (no such example is known). However, in the analytic category our understanding is much better:
Claim 6.1.
Let be a IFS. Then there is a dichotomy:
-
1.
is not conjugate to linear, or
-
2.
is conjugate to a linear IFS via an analytic map , that is a diffeomorphism on . In particular, is conjugate to a self-similar IFS.
Several variants of Claim 6.1 exist in the literature under various assumptions (see e.g. [11]). We also note that a version of the Claim holds if is an interval (via a closely related argument).
First, we require the following Proposition, a special case of the Poincaré-Siegel Theorem [31, Theorem 2.8.2]:
Proposition 6.2.
[31, Proposition 2.1.3] Let be a contracting map. Then there exists some non-trivial interval and a diffeomorphism such that is affine.
We also require the following Lemma:
Lemma 6.3.
Let be a IFS that is conjugate to a linear IFS. If there exists such that on its attractor , then is already linear; That is, for every , on .
Proof.
Let be such that the IFS is linear. Let be such that on . We first show that this implies that vanishes on : By our assumption, for all we have
Writing , we compute:
Combining the two previous displayed equations, it follow that
By our assumption since . We conclude that
(37) |
Finally, let . Since for all and for every we have for the -fold composition , then for all , . A similar argument shows that for every the IFS is linear. It follows that in (37) we can substitute for . As the IFS is uniformly contracting, this shows that the LHS of (37) can be made arbitrarily small, but the RHS remains fixed. This is only possible if . We conclude that vanishes on .
Finally, let . Then for all we have, for
Since then and so by the previous paragraph and . As on by assumption, this is only possible if . It follows that vanishes on , as claimed. ∎
Proof of Claim 6.1 Suppose is conjugate to linear. Let be any map. By Proposition 6.2, there is some non-trivial interval and a map such that is affine. Then the IFS is and contains an affine map. Since is conjugate to linear, so is . Since this IFS contains an affine map, by Lemma 6.3 it is already linear. So, it is linear and analytic, hence it must be self-similar. Thus, the second alternative of Claim 6.1 holds true.
6.2 Proof of Corollary 1.2
We now prove Corollary 1.2. Let be a IFS, and assume contains a non-affine map. Recall that we are always assuming is infinite. By Claim 6.1 there are two cases to consider:
The first alternative is that is not conjugate to linear. Then, by Theorem 1.1, every self-conformal measure admits some such that
The second alternative is that is conjugate to a self-similar IFS . Let denote the conjugating map. We have the following easy Lemma:
Lemma 6.4.
The analytic map is not affine.
Proof.
Suppose towards a contradiction that is affine. Recall that is a conjugating map between , a IFS, and a self-similar IFS. So, both IFS’s in question are in fact self-similar. However, our standing assumption is that contains contains a non affine map. This is a contradiction. ∎
So, in the second alternative, is conjugate to a self-similar IFS via a map that is not affine. In particular,
Since every self-conformal measure with respect to can be written as where is a self-similar measure with respect to , the Fourier decay bound in the second alternative case is a direct consequence of the following Theorem of Algom et al. [2]:
Theorem 6.5.
[2, Corollary 1.3] Let be a non-atomic self-similar measure with respect to , and let be such that , and such that except for possibly finitely many points in . Then there exists some such that
The proof of Corollary 1.2 is complete.
7 Acknowledgements
We thank Simon Baker, Tuomas Sahlsten, Meng Wu, and Osama Khalil, for useful discussions and for their remarks on this project. We also thank Joey Veltri for pointing out some bugs in a previous version of this manuscript. This research was supported by Grant No. 2022034 from the United States - Israel Binational Science Foundation (BSF), Jerusalem, Israel.
References
- [1] Amir Algom, Simon Baker, and Pablo Shmerkin. On normal numbers and self-similar measures. Adv. Math., 399:Paper No. 108276, 17, 2022.
- [2] Amir Algom, Yuanyang Chang, Meng Wu, and Yu-Liang Wu. Van der corput and metric theorems for geometric progressions for self-similar measures. arXiv preprint arXiv:2401.01120, 2024.
- [3] Amir Algom, Federico Rodriguez Hertz, and Zhiren Wang. Logarithmic Fourier decay for self conformal measures. J. Lond. Math. Soc. (2), 106(2):1628–1661, 2022.
- [4] Amir Algom, Federico Rodriguez Hertz, and Zhiren Wang. Pointwise normality and Fourier decay for self-conformal measures. Adv. Math., 393:Paper No. 108096, 72, 2021.
- [5] Vitor Araújo and Ian Melbourne. Exponential decay of correlations for nonuniformly hyperbolic flows with a stable foliation, including the classical Lorenz attractor. Ann. Henri Poincaré, 17(11):2975–3004, 2016.
- [6] Artur Avila, Jairo Bochi, and Jean-Christophe Yoccoz. Uniformly hyperbolic finite-valued -cocycles. Comment. Math. Helv., 85(4):813–884, 2010.
- [7] Artur Avila, Sébastien Gouëzel, and Jean-Christophe Yoccoz. Exponential mixing for the Teichmüller flow. Publ. Math. Inst. Hautes Études Sci., (104):143–211, 2006.
- [8] Simon Baker and Amlan Banaji. Polynomial Fourier decay for fractal measures and their pushforwards. arXiv preprint arXiv:2401.01241, 2024.
- [9] Simon Baker and Tuomas Sahlsten. Spectral gaps and Fourier dimension for self-conformal sets with overlaps. arXiv preprint arXiv:2306.01389, 2023.
- [10] Viviane Baladi and Brigitte Vallée. Exponential decay of correlations for surface semi-flows without finite Markov partitions. Proc. Amer. Math. Soc., 133(3):865–874, 2005.
- [11] Tim Bedford and Albert M. Fisher. Ratio geometry, rigidity and the scenery process for hyperbolic Cantor sets. Ergodic Theory Dynam. Systems, 17(3):531–564, 1997.
- [12] Yves Benoist and Jean-François Quint. Random walks on reductive groups, volume 62 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer, Cham, 2016.
- [13] Christopher J. Bishop and Yuval Peres. Fractals in probability and analysis, volume 162 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2017.
- [14] Jean Bourgain and Semyon Dyatlov. Fourier dimension and spectral gaps for hyperbolic surfaces. Geom. Funct. Anal., 27(4):744–771, 2017.
- [15] Jean-Baptiste Boyer. The rate of convergence for the renewal theorem in . arXiv preprint arXiv:1603.07214, 2016.
- [16] Julien Brémont. Self-similar measures and the Rajchman property. Ann. H. Lebesgue, 4:973–1004, 2021.
- [17] Alexander I. Bufetov and Boris Solomyak. On the modulus of continuity for spectral measures in substitution dynamics. Adv. Math., 260:84–129, 2014.
- [18] Xin-Rong Dai. When does a Bernoulli convolution admit a spectrum? Adv. Math., 231(3-4):1681–1693, 2012.
- [19] Xin-Rong Dai, De-Jun Feng, and Yang Wang. Refinable functions with non-integer dilations. J. Funct. Anal., 250(1):1–20, 2007.
- [20] H. Davenport, P. Erdős, and W. J. LeVeque. On Weyl’s criterion for uniform distribution. Michigan Math. J., 10:311–314, 1963.
- [21] Carlos Gustavo T. de A. Moreira. Stable intersections of Cantor sets and homoclinic bifurcations. Ann. Inst. H. Poincaré C Anal. Non Linéaire, 13(6):741–781, 1996.
- [22] Dmitry Dolgopyat. On decay of correlations in Anosov flows. Ann. of Math. (2), 147(2):357–390, 1998.
- [23] Dmitry Dolgopyat. Prevalence of rapid mixing in hyperbolic flows. Ergodic Theory Dynam. Systems, 18(5):1097–1114, 1998.
- [24] Dmitry Dolgopyat. Prevalence of rapid mixing. II. Topological prevalence. Ergodic Theory Dynam. Systems, 20(4):1045–1059, 2000.
- [25] De-Jun Feng and Ka-Sing Lau. Multifractal formalism for self-similar measures with weak separation condition. J. Math. Pures Appl. (9), 92(4):407–428, 2009.
- [26] Daniel Galicer, Santiago Saglietti, Pablo Shmerkin, and Alexia Yavicoli. dimensions and projections of random measures. Nonlinearity, 29(9):2609–2640, 2016.
- [27] Michael Hochman. A short proof of Host’s equidistribution theorem. Israel J. Math., 251(2):527–539, 2022.
- [28] Michael Hochman and Pablo Shmerkin. Equidistribution from fractal measures. Inventiones mathematicae, 202(1):427–479, 2015.
- [29] Thomas Jordan and Tuomas Sahlsten. Fourier transforms of Gibbs measures for the Gauss map. Math. Ann., 364(3-4):983–1023, 2016.
- [30] Antti Käenmäki and Tuomas Orponen. Absolute continuity in families of parametrised non-homogeneous self-similar measures. To appear in J. Fractal Geom.
- [31] Anatole Katok and Boris Hasselblatt. Introduction to the modern theory of dynamical systems, volume 54 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1995. With a supplementary chapter by Katok and Leonardo Mendoza.
- [32] Robert Kaufman. On Bernoulli convolutions. In Conference in modern analysis and probability (New Haven, Conn., 1982), volume 26 of Contemp. Math., pages 217–222. Amer. Math. Soc., Providence, RI, 1984.
- [33] Jialun Li. Decrease of Fourier coefficients of stationary measures. Math. Ann., 372(3-4):1189–1238, 2018.
- [34] Jialun Li. Fourier decay, renewal theorem and spectral gaps for random walks on split semisimple Lie groups. Ann. Sci. Éc. Norm. Supér. (4), 55(6):1613–1686, 2022.
- [35] Jialun Li, Frédéric Naud, and Wenyu Pan. Kleinian Schottky groups, Patterson-Sullivan measures, and Fourier decay. Duke Math. J., 170(4):775–825, 2021. With an appendix by Li.
- [36] Jialun Li and Tuomas Sahlsten. Trigonometric series and self-similar sets. J. Eur. Math. Soc. (JEMS), 24(1):341–368, 2022.
- [37] Pertti Mattila. Fourier analysis and Hausdorff dimension, volume 150 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2015.
- [38] 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.
- [39] Frédéric Naud. Expanding maps on Cantor sets and analytic continuation of zeta functions. Ann. Sci. École Norm. Sup. (4), 38(1):116–153, 2005.
- [40] Andrew D. Pollington, Sanju Velani, Agamemnon Zafeiropoulos, and Evgeniy Zorin. Inhomogeneous Diophantine approximation on -sets with restricted denominators. Int. Math. Res. Not. IMRN, (11):8571–8643, 2022.
- [41] Ariel Rapaport. On the Rajchman property for self-similar measures on . Adv. Math., 403:Paper No. 108375, 53, 2022.
- [42] Santiago Saglietti, Pablo Shmerkin, and Boris Solomyak. Absolute continuity of non-homogeneous self-similar measures. Adv. Math., 335:60–110, 2018.
- [43] Tuomas Sahlsten and Connor Stevens. Fourier transform and expanding maps on cantor sets. To appear in Amer. J. Math.
- [44] Pablo Shmerkin. On the exceptional set for absolute continuity of Bernoulli convolutions. Geom. Funct. Anal., 24(3):946–958, 2014.
- [45] Pablo Shmerkin and Boris Solomyak. Absolute continuity of self-similar measures, their projections and convolutions. Trans. Amer. Math. Soc., 368(7):5125–5151, 2016.
- [46] Boris Solomyak. Fourier decay for self-similar measures. Proc. Amer. Math. Soc., 149(8):3277–3291, 2021.
- [47] Boris Solomyak and Adam Śpiewak. Absolute continuity of self-similar measures on the plane. arXiv preprint arXiv:2301.10620, 2023.
- [48] Luchezar Stoyanov. Spectrum of the Ruelle operator and exponential decay of correlations for open billiard flows. Amer. J. Math., 123(4):715–759, 2001.
- [49] Luchezar Stoyanov. Spectra of Ruelle transfer operators for axiom A flows. Nonlinearity, 24(4):1089–1120, 2011.
- [50] Lauritz Streck. On absolute continuity and maximal garsia entropy for self-similar measures with algebraic contraction ratio. arXiv preprint arXiv:2303.07785, 2023.
- [51] Péter P. Varjú and Han Yu. Fourier decay of self-similar measures and self-similar sets of uniqueness. Anal. PDE, 15(3):843–858, 2022.
- [52] Jean-Christophe Yoccoz. Some questions and remarks about cocycles. In Modern dynamical systems and applications, pages 447–458. Cambridge Univ. Press, Cambridge, 2004.
Department of Mathematics, The University of Haifa at Oranim, Tivon 36006, Israel
E-mail address [email protected]
Department of Mathematics, the Pennsylvania State University, University Park, PA 16802, USA
E-mail address [email protected]
E-mail address [email protected]