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

Polynomial Fourier decay and a cocycle version of Dolgopyat’s method for self conformal measures

Amir Algom, Federico Rodriguez Hertz, and Zhiren Wang
Abstract

We show that every self conformal measure with respect to a C2()C^{2}(\mathbb{R}) IFS Φ\Phi has polynomial Fourier decay under some mild and natural non-linearity conditions. In particular, every such measure has polynomial decay if Φ\Phi is Cω()C^{\omega}(\mathbb{R}) 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 ν\nu be a Borel probability measure on \mathbb{R}. For every qq\in\mathbb{R} the Fourier transform of ν\nu at qq is defined by

q(ν):=exp(2πiqx)𝑑ν(x).\mathcal{F}_{q}(\nu):=\int\exp(2\pi iqx)d\nu(x).

The measure ν\nu is called a Rajchman measure if q(ν)=o(1)\mathcal{F}_{q}(\nu)=o(1) as |q||q|\rightarrow\infty. By the Riemann-Lebesgue Lemma, if ν\nu is absolutely continuous then it is Rajchman. On the other hand, by Wiener’s Lemma if ν\nu has an atom then it is not Rajchman. For measures that are both continuous (no atoms) and singular, determining whether or not ν\nu is a Rajchman measure may be a challenging problem even for well structured measures. The Rajchman property has various geometric consequences on the measure ν\nu and its support, e.g. regarding the uniqueness problem [36].

Further information about the rate of decay of q(ν)\mathcal{F}_{q}(\nu) has even stronger geometric consequences. For example, by a classical Theorem of Davenport-Erdős-LeVeque [20], if q(ν)\mathcal{F}_{q}(\nu) decays at a logarithmic rate then ν\nu-a.e. point is normal to all integer bases (see also [40]). Wide ranging geometric information can be derived if q(ν)\mathcal{F}_{q}(\nu) decays at a polynomial rate, that is, if there exists some α>0\alpha>0 such that

q(ν)=O(1|q|α).\mathcal{F}_{q}(\nu)=O\left(\frac{1}{|q|^{\alpha}}\right).

For example, by a result of Shmerkin [44], it implies that the convolution of ν\nu with any measure of dimension 11 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 Φ={f1,,fn}\Phi=\{f_{1},...,f_{n}\} be a finite set of strict contractions of a compact interval II\subseteq\mathbb{R} (an IFS - Iterated Function System), such that every fif_{i} is differentiable. We say that Φ\Phi is CαC^{\alpha} smooth if every fif_{i} is at least CαC^{\alpha} smooth for some α1\alpha\geq 1. It is well known that there exists a unique compact set K=KΦI\emptyset\neq K=K_{\Phi}\subseteq I such that

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

The set KK is called the attractor of the IFS {f1,,fn}\{f_{1},...,f_{n}\}. We always assume that there exist i,ji,j such that the fixed point of fif_{i} does not equal the fixed point of fjf_{j}. This ensures that KK is infinite. We call Φ\Phi uniformly contracting if

0<inf{|f(x)|:fΦ,xI}sup{|f(x)|:fΦ,xI}<1.0<\inf\{|f^{\prime}(x)|:\,f\in\Phi,x\in I\}\leq\sup\{|f^{\prime}(x)|:\,f\in\Phi,x\in I\}<1.

Next, writing 𝒜={1,,n}\mathcal{A}=\{1,...,n\}, for every ω𝒜\omega\in\mathcal{A}^{\mathbb{N}} and mm\in\mathbb{N} let

fω|m:=fω1fωm.f_{\omega|_{m}}:=f_{\omega_{1}}\circ\circ\circ f_{\omega_{m}}.

Fix x0Ix_{0}\in I. Then we have a surjective coding map π:𝒜K\pi:\mathcal{A}^{\mathbb{N}}\rightarrow K given by

ω𝒜xω:=limmfω|m(x0),\omega\in\mathcal{A}^{\mathbb{N}}\mapsto x_{\omega}:=\lim_{m\rightarrow\infty}f_{\omega|_{m}}(x_{0}), (2)

which is well defined because of uniform contraction (see e.g. [13, Section 2.1]).

Let p=(p1,,pn)\textbf{p}=(p_{1},...,p_{n}) be a strictly positive probability vector, that is, pi>0p_{i}>0 for all ii and ipi=1\sum_{i}p_{i}=1, and let =𝐩\mathbb{P}=\mathbf{p}^{\mathbb{N}} be the corresponding Bernoulli measure on 𝒜\mathcal{A}^{\mathbb{N}}. We call the measure ν=ν𝐩=π\nu=\nu_{\mathbf{p}}=\pi\mathbb{P} on KK the self conformal measure corresponding to 𝐩\mathbf{p}, and note that our assumptions are known to imply that it is non-atomic. Equivalently, ν𝐩\nu_{\mathbf{p}} is the unique Borel probability measure on KK such that

ν=i=1npifiν, where fiν is the push-forward of ν via fi.\nu=\sum_{i=1}^{n}p_{i}\cdot f_{i}\nu,\quad\text{ where }f_{i}\nu\text{ is the push-forward of }\nu\text{ via }f_{i}.

When all the maps in Φ\Phi are affine we call Φ\Phi a self-similar IFS and ν\nu a self-similar measure.

Next, we say that a C2()C^{2}(\mathbb{R}) IFS Ψ\Psi is linear if g′′(x)=0g^{\prime\prime}(x)=0 for every xKΨx\in K_{\Psi} and gΨg\in\Psi. 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 Ψ\Psi is Cω()C^{\omega}(\mathbb{R}) 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 Cr()C^{r}(\mathbb{R}) smooth IFS that is not self-similar for r1r\geq 1. 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 CrC^{r} IFS Φ\Phi is conjugate to an IFS Ψ\Psi if there is a CrC^{r} diffeomorphism hh such that Φ={hgh1}gΨ\Phi=\{h\circ g\circ h^{-1}\}_{g\in\Psi}.

Theorem 1.1.

Let Φ\Phi be a uniformly contracting Cr()C^{r}(\mathbb{R}) IFS where r2r\geq 2. If Φ\Phi is not conjugate to a linear IFS then every non-atomic self-conformal measure ν\nu admits some α=α(ν)>0\alpha=\alpha(\nu)>0 such that

|q(gν)|=O(1|q|α).\left|\mathcal{F}_{q}\left(g\nu\right)\right|=O\left(\frac{1}{|q|^{\alpha}}\right).

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 C2()C^{2}(\mathbb{R}) 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 SL(2,)\text{SL}(2,\mathbb{R}) 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 Φ\Phi be a Cω()C^{\omega}(\mathbb{R}) IFS. If Φ\Phi contains a non-affine map then every non-atomic self-conformal measure ν\nu admits some α=α(ν)>0\alpha=\alpha(\nu)>0 such that

|q(ν)|=O(1|q|α).\left|\mathcal{F}_{q}\left(\nu\right)\right|=O\left(\frac{1}{|q|^{\alpha}}\right).

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 SL(2,)\text{SL}(2,\mathbb{R}) 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 CωC^{\omega} IFSs with strong separation (i.e. the union (1) is disjoint). Their methods also apply when the IFS is only C2()C^{2}(\mathbb{R}) under the additional assumptions that KΦ=[0,1]K_{\Phi}=[0,1] and some further separation conditions on Φ\Phi. 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 C2C^{2} 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 Cω()C^{\omega}(\mathbb{R}) IFSs: Let Φ\Phi be a Cω()C^{\omega}(\mathbb{R}) IFS. If Φ\Phi admits a non-atomic self-conformal measure that is not Rajchman, then Φ={r1x+t1,,rnx+tn}\Phi=\{r_{1}\cdot x+t_{1},...,r_{n}\cdot x+t_{n}\} is self-similar and there exists a Pisot number r1r^{-1} such that ri=rir_{i}=r^{\ell_{i}} for some i+\ell_{i}\in\mathbb{Z}_{+}. Furthermore, Φ\Phi is affinely conjugated to an IFS that has all of its translations in (r)\mathbb{Q}(r). 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 C2([0,1])C^{2}([0,1]). Fix a self-conformal measure ν=ν𝐩\nu=\nu_{\mathbf{p}}, and assume the conditions as in Theorem 1.1 are met. We aim to show that ν\nu 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 𝐩\mathbf{p} (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 L2L^{2} 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 ν\nu and Φ\Phi 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 iji\neq j, fi([0,1])f_{i}([0,1]) and fj([0,1])f_{j}([0,1]) may intersect only at their endpoints), that the measure ν\nu satisfies the Federer property (see e.g. Theorem 2.4 Part (6)), and that Φ\Phi 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 ν\nu has the Federer property. As for (UNI), in our previous work [3, proof of Claim 2.13] we showed that if Φ\Phi 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 ν\nu 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 L2L^{2} 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 ϵ\epsilon from Theorem 2.8). Finally, in Section 5 we obtain the desired Fourier decay bound on q(ν)\mathcal{F}_{q}(\nu) 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 Cω()C^{\omega}(\mathbb{R}) IFS: Either it is not C2C^{2} conjugate to linear, or it is CωC^{\omega} 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 L2L^{2}-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 C2()C^{2}(\mathbb{R}) IFS Φ={f1,,fn}\Phi=\{f_{1},...,f_{n}\} and write 𝒜={1,,n}\mathcal{A}=\{1,...,n\}. We assume without the loss of generality that for every 1an1\leq a\leq n the map faf_{a} is a self map of [0,1][0,1], and that

K=KΦ does not contain the points 0,1.K=K_{\Phi}\text{ does not contain the points }0,1. (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

ρ~:=supfΦf(0,1),\tilde{\rho}:=\sup_{f\in\Phi}||f^{\prime}||_{\infty}\in(0,1),

and let

ρmin:=mini𝒜,x[0,1]|fi(x)|.\rho_{\min}:=\min_{i\in\mathcal{A},x\in[0,1]}|f_{i}^{\prime}(x)|.

Let us recall the bounded distortion property [4, Lemma 2.1]: There is some L=L(Φ)L=L(\Phi) such that for every η𝒜\eta\in\mathcal{A}^{*} we have

L1|fη(x)||fη(y)|L for all x,y[0,1].L^{-1}\leq\frac{\left|f_{\eta}^{\prime}(x)\right|}{\left|f_{\eta}^{\prime}(y)\right|}\leq L\text{ for all }x,y\in[0,1]. (4)

Next, note that

supx[0,1],ξ𝒜,n|ddx(logfξ|n)(x)|=supx[0,1],ξ𝒜,n|k=1nfξk′′fσk(ξ)|nk(x)fσk(ξ)|nk(x)fξk(x)fσk(ξ)|nk(x)|\sup_{x\in[0,1],\xi\in\mathcal{A}^{\mathbb{N}},n\in\mathbb{N}}\left|\frac{d}{dx}\left(\log f^{\prime}_{\xi|_{n}}\right)(x)\right|=\sup_{x\in[0,1],\xi\in\mathcal{A}^{\mathbb{N}},n\in\mathbb{N}}\left|\sum_{k=1}^{n}\frac{f_{\xi_{k}}^{\prime\prime}\circ f_{\sigma^{k}(\xi)|_{n-k}}(x)\cdot f_{\sigma^{k}(\xi)|_{n-k}}^{\prime}(x)}{f_{\xi_{k}}^{\prime}(x)\circ f_{\sigma^{k}(\xi)|_{n-k}}(x)}\right|
supx[0,1],fΦ|f′′(x)f(x)|11ρ~.\leq\sup_{x\in[0,1],f\in\Phi}\left|\frac{f^{\prime\prime}(x)}{f^{\prime}(x)}\right|\cdot\frac{1}{1-\tilde{\rho}}. (5)

So, there is a uniform constant C~:=2supx[0,1],fΦ|f′′(x)f(x)|\tilde{C}:=2\cdot\sup_{x\in[0,1],f\in\Phi}\left|\frac{f^{\prime\prime}(x)}{f^{\prime}(x)}\right|, such that, assuming as we may (see the next paragraph) that ρ~<12\tilde{\rho}<\frac{1}{2},

supx[0,1],ξ𝒜,n|ddx(logfξ|n)(x)|C~.\sup_{x\in[0,1],\xi\in\mathcal{A}^{\mathbb{N}},n\in\mathbb{N}}\left|\frac{d}{dx}\left(\log f^{\prime}_{\xi|_{n}}\right)(x)\right|\leq\tilde{C}. (6)

For all NN let ΦN\Phi^{N} be the corresponding induced IFS of generation NN, that is,

ΦN={fη:η𝒜N}.\Phi^{N}=\{f_{\eta}:\,\eta\in\mathcal{A}^{N}\}.

Since ΦN\Phi^{N} is finite, we may order its functions (f1,f2f_{1},f_{2} and so on) at our will. Note that the same constants LL from (4) and C~\tilde{C} from (6) will work for the induced IFS ΦN\Phi^{N} for all NN.

Our first step is, assuming Φ\Phi is not conjugate to linear, to construct an induced IFS that has useful separation and non-linearity properties:

Claim 2.1.

Let Φ\Phi be a C2([0,1])C^{2}([0,1]) IFS that is not conjugate to linear. Then there exists NN\in\mathbb{N} such that ΦN\Phi^{N} admits f1,f2,f3,f4f_{1},f_{2},f_{3},f_{4} satisfying:

  1. 1.

    For every k5k\geq 5 there exists i{1,3}i\in\{1,3\} such that

    fi([0,1])fi+1([0,1])fk([0,1])f_{i}([0,1])\cup f_{i+1}([0,1])\cup f_{k}([0,1])

    is a disjoint union. In particular, the union fi([0,1])fi+1([0,1])f_{i}([0,1])\cup f_{i+1}([0,1]) is disjoint for i=1,3i=1,3.

  2. 2.

    There exists m,m>0m^{\prime},m>0 such that: For every x[0,1]x\in[0,1], for both i=1,3i=1,3,

    m|ddx(logfilogfi+1)(x)|m.m\leq\left|\frac{d}{dx}\left(\log f_{i}^{\prime}-\log f_{i+1}^{\prime}\right)\left(x\right)\right|\leq m^{\prime}.
  3. 3.

    We have

    m2C~ρ~N>0.m-2\cdot\tilde{C}\cdot\tilde{\rho}^{N}>0.

We first remark that the existence of an upper bound mm^{\prime} as in Part (2) holds for all NN and every fi,fjΦNf_{i},f_{j}\in\Phi^{N}; 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 c,m,N0>0c,m^{\prime},N_{0}>0 such that for all n>N0n>N_{0} there exist ξ,ζ𝒜n\xi,\zeta\in\mathcal{A}^{n} such that for every x[0,1]x\in[0,1],

c<|ddx(logfξlogfζ)(x)|m.c<\left|\frac{d}{dx}\left(\log f_{\xi}^{\prime}-\log f_{\zeta}^{\prime}\right)\left(x\right)\right|\leq m^{\prime}.
Proof.

By [3, proof of Claim 2.13], if for all σ\sigma-periodic111Observe that for the proof of [3, Claim 2.13] to follow through it is enough to consider only σ\sigma-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 σ\sigma-periodic elements. ξ,ζ𝒜\xi,\zeta\in\mathcal{A}^{\mathbb{N}} and xKx\in K we have

limnddxlogfξ|n(x)=limnddxlogfζ|n(x),\lim_{n}\frac{d}{dx}\log f_{\xi|_{n}}^{\prime}(x)=\lim_{n}\frac{d}{dx}\log f_{\zeta|_{n}}^{\prime}(x), (7)

then Φ\Phi is C2C^{2} conjugate to a linear IFS. Thus, since Φ\Phi is not conjugate to linear, there exist some c>0c^{\prime}>0, x0Kx_{0}\in K and σ\sigma-periodic ξ,ζ𝒜\xi,\zeta\in\mathcal{A}^{\mathbb{N}} such that for infinitely many nn,

c<|ddx(logfξ|nlogfζ|n)(x0)|.c^{\prime}<\left|\frac{d}{dx}\left(\log f_{\xi|_{n}}^{\prime}-\log f_{\zeta|_{n}}^{\prime}\right)\left(x_{0}\right)\right|.

Recalling our coding map (2), let ω𝒜\omega\in\mathcal{A}^{\mathbb{N}} be such that xω=x0x_{\omega}=x_{0}. Using σ\sigma-periodicity, by the uniform convergence ([3, Claim 2.12] or (6)) as nn\rightarrow\infty of

|ddx(logfξ|nlogfζ|n)()|\left|\frac{d}{dx}\left(\log f_{\xi|_{n}}^{\prime}-\log f_{\zeta|_{n}}^{\prime}\right)\left(\cdot\right)\right|

there is some N1N_{1} and some k=k(N1,c)k=k(N_{1},c^{\prime}) such that for every n>N1n>N_{1}

|ddx(logfξ|nlogfζ|n)(x0)|c2|ddx(logfξ|nlogfζ|n)(fω|k(x))|\left|\frac{d}{dx}\left(\log f_{\xi|_{n}}^{\prime}-\log f_{\zeta|_{n}}^{\prime}\right)\left(x_{0}\right)\right|-\frac{c^{\prime}}{2}\leq\left|\frac{d}{dx}\left(\log f_{\xi|_{n}}^{\prime}-\log f_{\zeta|_{n}}^{\prime}\right)\left(f_{\omega|_{k}}(x)\right)\right|

Let n>max{N1,N0}n>\max\{N_{1},N_{0}\}, and let k=k(N1,c)k=k(N_{1},c^{\prime}) be as above. Then for all x[0,1]x\in[0,1],

c2ρmink|ddx(logfξ|nlogfζ|n)(fω|k(x))||fω|k(x)|\frac{c^{\prime}}{2}\cdot\rho_{\min}^{k}\leq\left|\frac{d}{dx}\left(\log f_{\xi|_{n}}^{\prime}-\log f_{\zeta|_{n}}^{\prime}\right)\left(f_{\omega|_{k}}(x)\right)\right|\cdot\left|f_{\omega|_{k}}^{\prime}(x)\right|
=|ddx(log(fξ|nfω|k)log(fζ|nfω|k))(x)|.=\left|\frac{d}{dx}\left(\log\left(f_{\xi|_{n}}\circ f_{\omega|_{k}}\right)^{\prime}-\log\left(f_{\zeta|_{n}}\circ f_{\omega|_{k}}\right)^{\prime}\right)\left(x\right)\right|.

So, for c=c2ρminkc=\frac{c^{\prime}}{2}\cdot\rho_{\min}^{k}, the words ξ=ξ|nω|k\xi^{\prime}=\xi|_{n}*\omega|_{k} and ζ=ζ|nω|k\zeta^{\prime}=\zeta|_{n}*\omega|_{k} satisfy the claimed bound from below for all n>max{N1,N0}n>\max\{N_{1},N_{0}\}. ∎

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 nn such that the ξ,ζ𝒜n\xi,\zeta\in\mathcal{A}^{n} as in Claim 2.2 satisfy

dist(fξ([0,1]),fζ([0,1]))>3(ρ~)n.\text{dist}\left(f_{\xi}([0,1]),\,f_{\zeta}([0,1])\right)>3(\tilde{\rho})^{n}.

Note that there exist some kk and η1,η2𝒜k\eta_{1},\eta_{2}\in\mathcal{A}^{k} such that

dist(fη1([0,1]),fη2([0,1]))diam(K)2.\text{dist}\left(f_{\eta_{1}}([0,1]),\,f_{\eta_{2}}([0,1])\right)\geq\frac{\text{diam}(K)}{2}.

For example, η1,η2\eta_{1},\eta_{2} can be chosen to be the kk-prefixes of codes of the endpoints of the interval conv(K)¯\overline{\text{conv}(K)}. Clearly every large enough kk will work. We thus assume that nn is large enough so that n>N0n>N_{0} as in Claim 2.2, the first displayed equation holds, and

2ρ~n<diam(K)/2.2\tilde{\rho}^{n}<\text{diam}(K)/2. (8)

.

We now define N=n+kN=n+k and our maps in ΦN\Phi^{N} via

f1:=fη1fξ,f2:=fη1fζ,f3:=fη2fξ,f4:=fη2fζ.f_{1}:=f_{\eta_{1}}\circ f_{\xi},\quad f_{2}:=f_{\eta_{1}}\circ f_{\zeta},\quad f_{3}:=f_{\eta_{2}}\circ f_{\xi},\quad f_{4}:=f_{\eta_{2}}\circ f_{\zeta}.

Let us show that all parts of the Claim hold: First, for part (2), for all x[0,1]x\in[0,1],

|ddx(logf1logf2)(x)|=|ddx(log(fη1fξ)log(fη1fζ))(x)|\left|\frac{d}{dx}\left(\log f_{1}^{\prime}-\log f_{2}^{\prime}\right)\left(x\right)\right|=\left|\frac{d}{dx}\left(\log\left(f_{\eta_{1}}\circ f_{\xi}\right)^{\prime}-\log\left(f_{\eta_{1}}\circ f_{\zeta}\right)^{\prime}\right)\left(x\right)\right|
=|ddx(logfξlogfζ+log(fη1fξ)log(fη1fζ))(x)|=\left|\frac{d}{dx}\left(\log f_{\xi}^{\prime}-\log f_{\zeta}^{\prime}+\log\left(f_{\eta_{1}}^{\prime}\circ f_{\xi}\right)-\log\left(f_{\eta_{1}}^{\prime}\circ f_{\zeta}\right)\right)\left(x\right)\right|
|ddx(logfξlogfζ)(x)||ddx(log(fη1fξ)log(fη1fζ))(x)|.\geq\left|\frac{d}{dx}\left(\log f_{\xi}^{\prime}-\log f_{\zeta}^{\prime}\right)\left(x\right)\right|-\left|\frac{d}{dx}\left(\log\left(f_{\eta_{1}}^{\prime}\circ f_{\xi}\right)-\log\left(f_{\eta_{1}}^{\prime}\circ f_{\zeta}\right)\right)\left(x\right)\right|.

Now, by Claim 2.2 since n>N0n>N_{0},

|ddx(logfξlogfζ)(x)|c.\left|\frac{d}{dx}\left(\log f_{\xi}^{\prime}-\log f_{\zeta}^{\prime}\right)\left(x\right)\right|\geq c.

On the other hand, arguing similarly to (6),

|ddxlog(fη1fξ)(x)|C~|fξ(x)|C~ρ~n.\left|\frac{d}{dx}\log\left(f_{\eta_{1}}^{\prime}\circ f_{\xi}\right)(x)\right|\leq\tilde{C}\cdot\left|f_{\xi}^{\prime}(x)\right|\leq\tilde{C}\cdot\tilde{\rho}^{n}.

Since we may assume nn is also large depending on C~,c\tilde{C},c (that are both known a-priori), we conclude that

|ddx(logf1logf2)(x)|c2C~ρ~n>0.\left|\frac{d}{dx}\left(\log f_{1}^{\prime}-\log f_{2}^{\prime}\right)\left(x\right)\right|\geq c-2\cdot\tilde{C}\cdot\tilde{\rho}^{n}>0.

The same calculation holds for f3,f4f_{3},f_{4}. This shows Part (2) holds true for our functions with

m:=c2C~ρ~n.m:=c-2\cdot\tilde{C}\cdot\tilde{\rho}^{n}.

Part (3) essentially follows from the same argument since, assuming nn is sufficiently large depending on c,C~c,\tilde{C}, we can get that also

c2C~ρ~n2C~ρ~n+k>0.c-2\cdot\tilde{C}\cdot\tilde{\rho}^{n}-2\cdot\tilde{C}\cdot\tilde{\rho}^{n+k}>0.

For Part (1), for every xf1([0,1])x\in f_{1}([0,1]) and yf2([0,1])y\in f_{2}([0,1]) there is some zz with

|f1(x)f2(y)|=|fη1(z)||fξ(x)fζ(y)|ρminkdist(fξ([0,1]),fζ([0,1]))|f_{1}(x)-f_{2}(y)|=|f_{\eta_{1}}^{\prime}(z)|\cdot|f_{\xi}(x)-f_{\zeta}(y)|\geq\rho_{\min}^{k}\cdot\text{dist}\left(f_{\xi}([0,1]),\,f_{\zeta}([0,1])\right)
ρmink3ρ~n>0.\geq\rho_{\min}^{k}\cdot 3\cdot\tilde{\rho}^{n}>0.

This shows that

dist(f1([0,1]),f2([0,1])),dist(f3([0,1]),f4([0,1]))>ρmink3ρ~n.\text{dist}\left(f_{1}([0,1]),\,f_{2}([0,1])\right),\quad\text{dist}\left(f_{3}([0,1]),\,f_{4}([0,1])\right)>\rho_{\min}^{k}\cdot 3\cdot\tilde{\rho}^{n}.

Next, by the choice of η1,η2\eta_{1},\eta_{2} and f1,f2,f3,f4f_{1},f_{2},f_{3},f_{4} it is clear that

dist(f1([0,1]),f3([0,1])),dist(f1([0,1]),f4([0,1])),dist(f2([0,1]),f3([0,1]))dist(f2([0,1]),f4([0,1]))\text{dist}\left(f_{1}([0,1]),\,f_{3}([0,1])\right),\,\text{dist}\left(f_{1}([0,1]),\,f_{4}([0,1])\right),\,\text{dist}\left(f_{2}([0,1]),\,f_{3}([0,1])\right)\,\text{dist}\left(f_{2}([0,1]),\,f_{4}([0,1])\right)
diam(K)2.\geq\frac{\text{diam}(K)}{2}.

Now, fix any ff_{\ell} with 5\ell\geq 5. If

f1([0,1])f2([0,1])f([0,1])f_{1}([0,1])\cup f_{2}([0,1])\cup f_{\ell}([0,1])

is a disjoint union then we are done. Otherwise, suppose

f1([0,1])f([0,1]).f_{1}([0,1])\cap f_{\ell}([0,1])\neq\emptyset.

Then, since diamf([0,1])(ρ~)N\text{diam}f_{\ell}([0,1])\leq(\tilde{\rho})^{N} and by (8),

dist(f([0,1]),f3([0,1])),dist(f([0,1]),f4([0,1]))diam(K)22(ρ~)N>0.\text{dist}\left(f_{\ell}([0,1]),\,f_{3}([0,1])\right),\quad\text{dist}\left(f_{\ell}([0,1]),\,f_{4}([0,1])\right)\geq\frac{\text{diam}(K)}{2}-2(\tilde{\rho})^{N}>0.

So, since the unions f1([0,1])f2([0,1])f_{1}([0,1])\cup f_{2}([0,1]) and f3([0,1])f4([0,1])f_{3}([0,1])\cup f_{4}([0,1]) have already been shown to be disjoint, the union

f3([0,1])f4([0,1])f([0,1])f_{3}([0,1])\cup f_{4}([0,1])\cup f_{\ell}([0,1])

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 n>N0n>N_{0}, for ξ,ζ𝒜n\xi,\zeta\in\mathcal{A}^{n} as in Claim 2.2,

dist(fξ([0,1]),fζ([0,1]))3(ρ~)n.\text{dist}\left(f_{\xi}([0,1]),\,f_{\zeta}([0,1])\right)\leq 3(\tilde{\rho})^{n}.

Let us find some ω1,ω2𝒜\omega_{1},\omega_{2}\in\mathcal{A}^{\mathbb{N}} such that for all kk

dist(fω1|k([0,1]),fω2|k([0,1]))>3(ρ~)k.\text{dist}\left(f_{\omega_{1}|_{k}}([0,1]),\,f_{\omega_{2}|_{k}}([0,1])\right)>3(\tilde{\rho})^{k}. (9)

Such ω1,ω2\omega_{1},\omega_{2} must exist since KΦK_{\Phi} is infinite.

Then for all large n>N0n^{\prime}>N_{0} and ξ,ζ𝒜n\xi,\zeta\in\mathcal{A}^{n^{\prime}} as in Claim 2.2, for all large nn for all x[0,1]x\in[0,1], we have

|ddx(log(fω1|nfξ)log(fω2|nfζ))(x)|\left|\frac{d}{dx}\left(\log\left(f_{\omega_{1}|_{n}}\circ f_{\xi}\right)^{\prime}-\log\left(f_{\omega_{2}|_{n}}\circ f_{\zeta}\right)^{\prime}\right)\left(x\right)\right|
=|ddx(logfω1|nfξlogfω2|nfζ)(x)ddx(logfζlogfξ)(x)|=\left|\frac{d}{dx}\left(\log f_{\omega_{1}|_{n}}^{\prime}\circ f_{\xi}-\log f_{\omega_{2}|_{n}}^{\prime}\circ f_{\zeta}\right)\left(x\right)-\frac{d}{dx}\left(\log f_{\zeta}^{\prime}-\log f_{\xi}^{\prime}\right)\left(x\right)\right|
|ddx(logfζlogfξ)(x)||ddx(logfω1|nfξlogfω2|nfζ)(x)|\geq\left|\frac{d}{dx}\left(\log f_{\zeta}^{\prime}-\log f_{\xi}^{\prime}\right)\left(x\right)\right|-\left|\frac{d}{dx}\left(\log f_{\omega_{1}|_{n}}^{\prime}\circ f_{\xi}-\log f_{\omega_{2}|_{n}}^{\prime}\circ f_{\zeta}\right)\left(x\right)\right|
c2C~ρ~n=co(1).\geq c-2\tilde{C}\cdot\tilde{\rho}^{n^{\prime}}=c-o(1).

Thus, ξ′′=ω1|nξ\xi^{\prime\prime}=\omega_{1}|_{n}*\xi and ζ′′=ω2|nζ\zeta^{\prime\prime}=\omega_{2}|_{n}*\zeta are words that can be made arbitrarily long, that satisfy that for some c>0c^{\prime}>0 and for all xx

|ddx(logfξ′′logfζ′′)(x)|>c.\left|\frac{d}{dx}\left(\log f_{\xi^{\prime\prime}}^{\prime}-\log f_{\zeta^{\prime\prime}}^{\prime}\right)\left(x\right)\right|>c^{\prime}.

Also, by (9),

dist(fξ′′([0,1]),fζ′′([0,1]))>3(ρ~)n>3(ρ~)n+n.\text{dist}\left(f_{\xi^{\prime\prime}}([0,1]),\,f_{\zeta^{\prime\prime}}([0,1])\right)>3(\tilde{\rho})^{n}>3(\tilde{\rho})^{n+n^{\prime}}.

Thus, we are back in Case 1 with these ξ′′,ζ′′\xi^{\prime\prime},\zeta^{\prime\prime} (that we recall can be made arbitrarily long) and cc^{\prime}.

Thus, we have completely reduced Case 2 to Case 1, completing the proof of Claim 2.1. \Box

2.2 The Euclidean derivative cocycle and associated transfer operator

Fix a C2C^{2} IFS Φ\Phi as in Section 2.1, and let us retain the other assumptions and notations from that Section. Let p=(p1,,pn)\textbf{p}=(p_{1},...,p_{n}) be a strictly positive probability vector on 𝒜\mathcal{A}, and let =𝐩\mathbb{P}=\mathbf{p}^{\mathbb{N}} be the corresponding product measure on 𝒜\mathcal{A}^{\mathbb{N}}. Let ν=ν𝐩\nu=\nu_{\mathbf{p}} be the corresponding self-conformal measure.

Note that for every PP the induced IFS ΦP\Phi^{P} has the same attractor KK as Φ\Phi. That is, KΦP=KΦK_{\Phi^{P}}=K_{\Phi}. Furthermore, ν\nu is also a self-conformal measure with respect ΦP\Phi^{P} and the induced probability vector 𝐩P\mathbf{p}^{P} on 𝒜P\mathcal{A}^{P}.

Thus, by working with the induced IFS ΦN\Phi^{N} as in Claim 2.1, we may assume without the loss of generality that our (not conjugate to linear) original IFS Φ\Phi already admits f1,f2,f3,f4Φf_{1},f_{2},f_{3},f_{4}\in\Phi that satisfy:

  1. 1.

    For every k1k\geq 1 there exists i{1,3}i\in\{1,3\} such that

    fi([0,1])fi+1([0,1])fk([0,1]).f_{i}([0,1])\cup f_{i+1}([0,1])\cup f_{k}([0,1]). (10)

    is a disjoint union. In particular, the union fi([0,1])fi+1([0,1])f_{i}([0,1])\cup f_{i+1}([0,1]) is disjoint for i=1,3i=1,3.

  2. 2.

    There exists m,m>0m^{\prime},m>0 such that: For every x[0,1]x\in[0,1], for both i=1,3i=1,3,

    m|ddx(logfilogfi+1)(x)|m, and m2C~supfΦf>0.m\leq\left|\frac{d}{dx}\left(\log f_{i}^{\prime}-\log f_{i+1}^{\prime}\right)\left(x\right)\right|\leq m^{\prime},\text{ and }m-2\cdot\tilde{C}\cdot\sup_{f\in\Phi}||f^{\prime}||_{\infty}>0. (11)

Note that for the second part of (11) we use Claim 2.1 Part (3) and that, with the notation ρ~\tilde{\rho} as in Section 2.1,

supfΦNfρ~N.\sup_{f\in\Phi^{N}}||f^{\prime}||_{\infty}\leq\tilde{\rho}^{N}.

Let GG to be the free semigroup generated by the family {fa:1an}\{f_{a}:1\leq a\leq n\}, which acts on [0,1][0,1] by composing the corresponding faf_{a}’s. We define the derivative cocycle c:G×[0,1]c:G\times[0,1]\rightarrow\mathbb{R} via

c(I,x)=log|fI(x)|.c(I,x)=-\log\left|f^{\prime}_{I}(x)\right|. (12)

Slightly adjusting our notation from Section 2.1, let

ρ:=supfΦf(0,1).\rho:=\sup_{f\in\Phi}||f^{\prime}||_{\infty}\in(0,1).

Also, we note that by uniform contraction

0<D:=min{log|f(x)|:fΦ,xI},D:=max{log|f(x)|:fΦ,xI}<.0<D:=\min\{-\log|f^{\prime}(x)|:f\in\Phi,x\in I\},\quad D^{\prime}:=\max\{-\log|f^{\prime}(x)|:f\in\Phi,x\in I\}<\infty. (13)

A major part in our argument is played by the transfer operator:

Definition 2.3.

For every ss\in\mathbb{C} such that |(s)||\Re(s)| is small enough, let Ps:C1([0,1])C1([0,1])P_{s}:C^{1}([0,1])\rightarrow C^{1}([0,1]) denote the transfer operator defined by, for gC1g\in C^{1} and x[0,1]x\in[0,1],

Ps(g)(x)=e2πsc(a,x)gfa(x)𝑑𝐩(a).P_{s}(g)(x)=\int e^{2\pi\cdot s\cdot c(a,x)}g\circ f_{a}(x)\,d\mathbf{p}(a).

Note that ν\nu is the unique stationary measure corresponding to the measure a𝒜paδ{fa}\sum_{a\in\mathcal{A}}p_{a}\cdot\delta_{\{f_{a}\}} on GG. So, the detailed discussion about this operator as in [12, Section 11.5] applies in the setting we are considering.

2.3 Disintegration of ν\nu 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 II be a finite set of C2()C^{2}(\mathbb{R}) iterated function systems {f1(i),,fki(i)},iI\{f_{1}^{(i)},...,f_{k_{i}}^{(i)}\},i\in I. Let Ω=I\Omega=I^{\mathbb{N}}. For ωΩ\omega\in\Omega and n{}n\in\mathbb{N}\cup\{\infty\} let

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

Define a coding map Πω:X(ω)\Pi_{\omega}:X_{\infty}^{(\omega)}\rightarrow\mathbb{R} via

Πω(u)=limnfu1(ω1)fun(ωn)(0).\Pi_{\omega}(u)=\lim_{n\rightarrow\infty}f^{(\omega_{1})}_{u_{1}}\circ\circ\circ f^{(\omega_{n})}_{u_{n}}(0).

Next, for each iIi\in I let 𝐩i=(p1(i),,pki(i))\mathbf{p}_{i}=(p_{1}^{(i)},...,p_{k_{i}}^{(i)}) be a probability vector with strictly positive entries. On each X(ω)X_{\infty}^{(\omega)} we define the product measure

η(ω)=n=1𝐩(ωn).\eta^{(\omega)}=\prod_{n=1}^{\infty}\mathbf{p}^{(\omega_{n})}.

We can now define

μω:=Πω(η(ω)).\mu_{\omega}:=\Pi_{\omega}\left(\eta^{(\omega)}\right).

Letting σ:ΩΩ\sigma:\Omega\rightarrow\Omega be the left shift, we have the following dynamical self-conformality relation:

μω=uX1(ω)pu(ω1)fu(ω1)μσ(ω).\mu_{\omega}=\sum_{u\in X_{1}^{(\omega)}}p_{u}^{(\omega_{1})}f_{u}^{(\omega_{1})}\mu_{\sigma(\omega)}. (14)

We also define

Kω:=supp(μω)=Πω(X(ω)),K_{\omega}:=\text{supp}(\mu_{\omega})=\Pi_{\omega}\left(X_{\infty}^{(\omega)}\right),

and note that here we have

Kω=uX1(ω)fu(ω1)Kσ(ω).K_{\omega}=\bigcup_{u\in X_{1}^{(\omega)}}f_{u}^{(\omega_{1})}K_{\sigma(\omega)}.

Next, for ωΩ,N\omega\in\Omega,N\in\mathbb{N}, and ss\in\mathbb{C} we define an operator Ps,ω,N:C1([0,1])C1([0,1])P_{s,\omega,N}:C^{1}\left([0,1]\right)\rightarrow C^{1}\left([0,1]\right) by

Ps,ω,N(g)(x):=IXN(ω)η(ω)(I)e2πsc(I,x)gfI(x), where η(ω)(I):=n=1|I|pIn(ωn).P_{s,\omega,N}\left(g\right)(x):=\sum_{I\in X_{N}^{(\omega)}}\eta^{(\omega)}(I)e^{2\pi sc(I,x)}g\circ f_{I}(x),\quad\text{ where }\eta^{(\omega)}(I):=\prod_{n=1}^{|I|}\textbf{p}^{(\omega_{n})}_{I_{n}}. (15)

Note that we only need to know ω|N\omega|_{N} in order for Ps,ω,NP_{s,\omega,N} to be well defined. Iterating (14) we have the following equivariance relations, whose proof is left to the reader: First, for every ωΩ\omega\in\Omega, NN\in\mathbb{N}, and gC1([0,1])g\in C^{1}\left([0,1]\right),

g(x)𝑑μω(x)=P0,ω,N(g)(x)𝑑μσNω(x).\int g(x)\,d\mu_{\omega}(x)=\int P_{0,\omega,N}\left(g\right)(x)\,d\mu_{\sigma^{N}\omega}(x). (16)

Furthermore, for all integers 0n~n,N,0\leq\tilde{n}\leq n,N, we have that

Ps,ω,nN(g)(x)=Ps,σn~Nω,NnNn~(Ps,ω,n~N(g))(x).P_{s,\omega,nN}\left(g\right)(x)=P_{s,\sigma^{\tilde{n}N}\omega,Nn-N\tilde{n}}\left(P_{s,\omega,\tilde{n}N}\left(g\right)\right)\left(x\right). (17)

Finally, let \mathbb{Q} be a σ\sigma-invariant measure on Ω\Omega. The triplet Σ=(ΦiI(i),(𝐩i)iI,)\Sigma=(\Phi_{i\in I}^{(i)},(\mathbf{p}_{i})_{i\in I},\mathbb{Q}) is called a model. We say the model is non-trivial if μω\mu_{\omega} is non-atomic for \mathbb{Q}-a.e. ω\omega. We say it is Bernoulli if \mathbb{Q} is a Bernoulli measure. We are now ready to state the main result of this Section:

Theorem 2.4.

Let Φ\Phi be a C2([0,1])C^{2}([0,1]) IFS satisfying properties (10) and (11), and let ν=ν𝐩\nu=\nu_{\mathbf{p}} be a self-conformal measure. Then there exists a non-trivial Bernoulli model Σ=Σ(Φ,ν)\Sigma=\Sigma(\Phi,\nu) such that:

  1. 1.

    (Disintegration of measure) ν=μ(ω)𝑑(ω)\nu=\int\mu_{(\omega)}d\mathbb{Q}(\omega).

  2. 2.

    (Disintegration of operator) For every N,s,fC1([0,1]),N\in\mathbb{N},s\in\mathbb{C},f\in C^{1}([0,1]), and x[0,1]x\in[0,1] we have

    PsN(f)(x)=ωΩN([ω])Ps,ω,N(f)(x).P_{s}^{N}\left(f\right)(x)=\sum_{\omega\in\Omega^{N}}\mathbb{Q}\left([\omega]\right)P_{s,\omega,N}\left(f\right)(x). (18)
  3. 3.

    (Non-trivial branching) For every iIi\in I we have ki=2k_{i}=2 or ki=3k_{i}=3.

  4. 4.

    (Separation) For every ωΩ\omega\in\Omega the union

    uX1(ω)fu(ω1)([0,1])\bigcup_{u\in X_{1}^{(\omega)}}f_{u}^{(\omega_{1})}([0,1])

    is disjoint.

  5. 5.

    (UNI in all parts) There exist m,m>0m^{\prime},m>0 and N00N_{0}\geq 0 such that for all NN0N\geq N_{0}, for every ωΩ\omega\in\Omega there exist α1N,α2NXN(ω)\alpha_{1}^{N},\alpha_{2}^{N}\in X_{N}^{(\omega)} such that

    m|ddx(logfα1Nlogfα2N)(x)|m, for all x[0,1]m\leq\left|\frac{d}{dx}\left(\log f_{\alpha_{1}^{N}}^{\prime}-\log f_{\alpha_{2}^{N}}^{\prime}\right)\left(x\right)\right|\leq m^{\prime},\quad\text{ for all }x\in[0,1].

  6. 6.

    (Federer property) For every D>1D>1 there exists CD=CD(Σ)>0C_{D}=C_{D}(\Sigma)>0 such that:

    For every ωΩ\omega\in\Omega, for every xsupp(μω)x\in\text{supp}(\mu_{\omega}) and r>0r>0,

    μ(ω)(B(x,Dr))Cdμ(ω)(B(x,r)).\mu_{(\omega)}\left(B(x,Dr)\right)\leq C_{d}\mu_{(\omega)}\left(B(x,r)\right).

Recall that our IFS Φ\Phi 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 Φ\Phi. For i=1,2,3,4i=1,2,3,4 we define the IFSs

Φ1=Φ2={f1,f2}, and Φ3=Φ4={f3,f4}.\Phi_{1}=\Phi_{2}=\{f_{1},f_{2}\},\text{ and }\Phi_{3}=\Phi_{4}=\{f_{3},f_{4}\}.

For every k5k\geq 5 define the IFS

Φk={fi,fi+1,fk}\Phi_{k}=\{f_{i},f_{i+1},f_{k}\}

as in (10).

We now define

I={Φi:i𝒜}.I=\{\Phi_{i}:\,i\in\mathcal{A}\}.

Thus, for every iIi\in I we associate the IFS Φi\Phi_{i}.

Note that certain fiΦf_{i}\in\Phi appear in multiple IFS’s in II. Write, for i𝒜i\in\mathcal{A},

ni=|{Φj:fiΦj}|.n_{i}=\left|\{\Phi_{j}:\,f_{i}\in\Phi_{j}\}\right|.

Recall that ν=ν𝐩\nu=\nu_{\mathbf{p}} where 𝐩𝒫(𝒜)\mathbf{p}\in\mathcal{P}(\mathcal{A}). We now define a probability vector 𝐪\mathbf{q} on II as follows:

𝐪1=𝐪2=p1n1+p2n2,𝐪3=𝐪4=p3n3+p4n4,𝐪k=pini+pi+1ni+1+pknk for k5.{\mathbf{q}}_{1}={\mathbf{q}}_{2}=\frac{p_{1}}{n_{1}}+\frac{p_{2}}{n_{2}},\quad{\mathbf{q}}_{3}={\mathbf{q}}_{4}=\frac{p_{3}}{n_{3}}+\frac{p_{4}}{n_{4}},\quad{\mathbf{q}}_{k}=\frac{p_{i}}{n_{i}}+\frac{p_{i+1}}{n_{i+1}}+\frac{p_{k}}{n_{k}}\text{ for }k\geq 5.

Let =𝐪\mathbb{Q}=\mathbf{q}^{\mathbb{N}}. This will be our Bernoulli selection measure on Ω\Omega^{\mathbb{N}}.

Next, for every iIi\in I we define the probability vector

𝐩~1=𝐩~2=(p1n1p1n1+p2n2,p2n2p1n1+p2n2),𝐩~3=𝐩~4=(p3n3p3n3+p4n4,p4n4p3n3+p4n4),\tilde{\mathbf{p}}_{1}=\tilde{\mathbf{p}}_{2}=\left(\frac{\frac{p_{1}}{n_{1}}}{\frac{p_{1}}{n_{1}}+\frac{p_{2}}{n_{2}}},\,\frac{\frac{p_{2}}{n_{2}}}{\frac{p_{1}}{n_{1}}+\frac{p_{2}}{n_{2}}}\right),\quad\tilde{\mathbf{p}}_{3}=\tilde{\mathbf{p}}_{4}=\left(\frac{\frac{p_{3}}{n_{3}}}{\frac{p_{3}}{n_{3}}+\frac{p_{4}}{n_{4}}},\,\frac{\frac{p_{4}}{n_{4}}}{\frac{p_{3}}{n_{3}}+\frac{p_{4}}{n_{4}}}\right),
𝐩~k=(pinipini+pi+1ni+1+pknk,pi+1ni+1pini+pi+1ni+1+pknk,pknkpini+pi+1ni+1+pknk) for k5.\tilde{\mathbf{p}}_{k}=\left(\frac{\frac{p_{i}}{n_{i}}}{\frac{p_{i}}{n_{i}}+\frac{p_{i+1}}{n_{i+1}}+\frac{p_{k}}{n_{k}}},\,\frac{\frac{p_{i+1}}{n_{i+1}}}{\frac{p_{i}}{n_{i}}+\frac{p_{i+1}}{n_{i+1}}+\frac{p_{k}}{n_{k}}},\,\frac{\frac{p_{k}}{n_{k}}}{\frac{p_{i}}{n_{i}}+\frac{p_{i+1}}{n_{i+1}}+\frac{p_{k}}{n_{k}}}\right)\text{ for }k\geq 5.

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

i𝒜fipiμ(ω)𝑑(ω)=μ(ω)𝑑(ω).\sum_{i\in\mathcal{A}}f_{i}p_{i}\cdot\int\mu_{(\omega)}\,d\mathbb{Q}(\omega)=\int\mu_{(\omega)}\,d\mathbb{Q}(\omega).

The proof is almost entirely the same as [1, Section 2.2], so we omit the details. Part (1) now follows since ν\nu is the unique measure satisfying this identity.

Proof of Parts (2) and (3) These are straightforward given our construction. Indeed, for part (2), when N=1N=1 it suffices to note that for every i𝒜i\in\mathcal{A},

j:fiΦj𝐪j𝐩j(i)=nipini=pi.\sum_{j:\,f_{i}\in\Phi_{j}}{\mathbf{q}}_{j}\cdot\mathbf{p}_{j}(i)=n_{i}\cdot\frac{p_{i}}{n_{i}}=p_{i}.

For general NN 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 ωΩ\omega\in\Omega and let N0=1N_{0}=1. Choose any N>N0N>N_{0}. By (11) and our construction, there exist m,m>0m^{\prime},m>0 such that for some i,jΦωNi,j\in\Phi_{\omega_{N}} with

m|ddx(logfilogfj)(x)|m, for every x[0,1].m\leq\left|\frac{d}{dx}\left(\log f_{i}^{\prime}-\log f_{j}^{\prime}\right)\left(x\right)\right|\leq m^{\prime},\text{ for every }x\in[0,1].

Let ξXN1(ω)\xi\in X^{(\omega)}_{N-1}, and put α1:=ξi,α2:=ξjXN(ω)\alpha_{1}:=\xi*i,\alpha_{2}:=\xi*j\in X^{(\omega)}_{N}. Then for all x[0,1]x\in[0,1],

|ddx(logfα1logfα2)(x)|=|ddx(log(fξfi)log(fξfj))(x)|\left|\frac{d}{dx}\left(\log f_{\alpha_{1}}^{\prime}-\log f_{\alpha_{2}}^{\prime}\right)\left(x\right)\right|=\left|\frac{d}{dx}\left(\log\left(f_{\xi}\circ f_{i}\right)^{\prime}-\log\left(f_{\xi}\circ f_{j}\right)^{\prime}\right)\left(x\right)\right|
=|ddx(logfilogfj+log(fξfi)log(fξfj))(x)|=\left|\frac{d}{dx}\left(\log f_{i}^{\prime}-\log f_{j}^{\prime}+\log\left(f_{\xi}^{\prime}\circ f_{i}\right)-\log\left(f_{\xi}^{\prime}\circ f_{j}\right)\right)\left(x\right)\right|
|ddx(logfilogfj)(x)||ddx(log(fξfi)log(fξfj))(x)|.\geq\left|\frac{d}{dx}\left(\log f_{i}^{\prime}-\log f_{j}^{\prime}\right)\left(x\right)\right|-\left|\frac{d}{dx}\left(\log\left(f_{\xi}^{\prime}\circ f_{i}\right)-\log\left(f_{\xi}^{\prime}\circ f_{j}\right)\right)\left(x\right)\right|.

By arguing similarly to (6),

|ddxlog(fξfi)(x)|C~|fi(x)|C~ρ.\left|\frac{d}{dx}\log\left(f_{\xi}^{\prime}\circ f_{i}\right)(x)\right|\leq\tilde{C}\cdot\left|f_{i}^{\prime}(x)\right|\leq\tilde{C}\cdot\rho.

By the choice of i,ji,j we conclude that

|ddx(logfα1logfα2)(x)|m2C~ρ>0,\left|\frac{d}{dx}\left(\log f_{\alpha_{1}}^{\prime}-\log f_{\alpha_{2}}^{\prime}\right)\left(x\right)\right|\geq m-2\cdot\tilde{C}\cdot\rho>0,

where the last inequality is due to the second part of (11). \Box

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 ω\omega (i.e. depend only on the model and not the measure under consideration).

Definition 2.5.

Fix ωΩ\omega\in\Omega and nn\in\mathbb{N}. The cylinder that corresponds to uXn(ω)u\in X_{n}^{(\omega)} is the set

Cu:=fu1(ω1)fun(ωn)([0,1])[0,1].C_{u}:=f^{(\omega_{1})}_{u_{1}}\circ\circ\circ f^{(\omega_{n})}_{u_{n}}([0,1])\subseteq[0,1].

From now on we fix the ω\omega in question and suppress it in our notation. Also, note that by the definition of μω\mu_{\omega}, (14), and Theorem 2.4 Part (4),

μω(Cu)=μω(fu1(ω1)fun(ωn)([0,1]))=η(ω)([u1,,un])=i=1n𝐩ui(ωi).\mu_{\omega}(C_{u})=\mu_{\omega}(f^{(\omega_{1})}_{u_{1}}\circ\circ\circ f^{(\omega_{n})}_{u_{n}}([0,1]))=\eta^{(\omega)}([u_{1},...,u_{n}])=\prod_{i=1}^{n}\mathbf{p}^{(\omega_{i})}_{u_{i}}. (19)

For a cylinder set CαC_{\alpha} let |Cα||C_{\alpha}| denote its diameter. Recall that we are fixing some ωΩ\omega\in\Omega, and considering cylinders with respect to KωK_{\omega}.

Lemma 2.6.

There exist constants C>0C>0 and 0<δ1,δ2<10<\delta_{1},\delta_{2}<1 uniform in ω\omega such that for all cylinders CαCβC_{\alpha}\subseteq C_{\beta}

C1δ1|α||β||Cα||Cβ|Cδ2|α||β|.C^{-1}\delta_{1}^{|\alpha|-|\beta|}\leq\frac{|C_{\alpha}|}{|C_{\beta}|}\leq C\cdot\delta_{2}^{|\alpha|-|\beta|}.
Proof.

Write α=βu\alpha=\beta\cdot u where |u|=|α||β||u|=|\alpha|-|\beta|. Then, omitting superscripts, for some x0x_{0}

|Cα|=|fβfu([0,1])|=|fβfu(x0)fu(x0)||C_{\alpha}|=|f_{\beta}\circ f_{u}([0,1])|=\left|f_{\beta}^{\prime}\circ f_{u}(x_{0})\cdot f_{u}^{\prime}(x_{0})\right|

and for some y0y_{0} we have

|Cβ|=|fβ([0,1])|=|fβ(y0)|.|C_{\beta}|=|f_{\beta}([0,1])|=\left|f_{\beta}^{\prime}(y_{0})\right|.

Letting LL be as in (4), recalling that fβf_{\beta} is a composition of maps from Φ\Phi,

L1(miniI,fΦ(i)minx[0,1]|fi(x)|)|u|L1|fu(x0)||Cα||Cβ|L|fu(x0)|Lρ|u|,L^{-1}\left(\min_{i\in I,f\in\Phi^{(i)}}\min_{x\in[0,1]}\left|f_{i}^{\prime}(x)\right|\right)^{|u|}\leq L^{-1}\cdot\left|f_{u}^{\prime}(x_{0})\right|\leq\frac{|C_{\alpha}|}{|C_{\beta}|}\leq L\cdot\left|f_{u}^{\prime}(x_{0})\right|\leq L\cdot\rho^{|u|},

as claimed. ∎

Lemma 2.7.

For every cylinder CβC_{\beta} we have

CβKωCαCβCα,C_{\beta}\cap K_{\omega}\subseteq\bigcup_{C_{\alpha}\subseteq C_{\beta}}C_{\alpha},

where |α|=|β|+1|\alpha|=|\beta|+1. Moreover, there is some λ>0\lambda>0 uniform in ω\omega such that for any two distinct cylinders CαiCβC_{\alpha_{i}}\subseteq C_{\beta} with i=1,2i=1,2 and |αi|=|β|+1|\alpha_{i}|=|\beta|+1 we have

dist(Cα1,Cα2)λ|Cβ|.\text{dist}\left(C_{\alpha_{1}},\,C_{\alpha_{2}}\right)\geq\lambda\cdot|C_{\beta}|.
Proof.

For the first assertion, for every nn\in\mathbb{N} we have by definition

Kω=uXn(ω)fu(ω)Kσn(ω),KωK[0,1] for every ω, and Kσn(ω)=iX1(σnω)fi(ωn+1)(Kσn+1(ω)).K_{\omega}=\bigcup_{u\in X_{n}^{(\omega)}}f_{u}^{(\omega)}K_{\sigma^{n}(\omega)},\quad K_{\omega}\subseteq K\subseteq[0,1]\text{ for every }\omega,\text{ and }\quad K_{\sigma^{n}(\omega)}=\bigcup_{i\in X^{(\sigma^{n}\omega)}_{1}}f^{(\omega_{n+1})}_{i}(K_{\sigma^{n+1}(\omega)}).

By Theorem 2.4 part (4),

CβKω=fβ(ω)([0,1])fβ(ω)(Kσn(ω))=fβ(ω)(Kσn(ω)).C_{\beta}\cap K_{\omega}=f_{\beta}^{(\omega)}([0,1])\cap f_{\beta}^{(\omega)}(K_{\sigma^{n}(\omega)})=f_{\beta}^{(\omega)}(K_{\sigma^{n}(\omega)}).

So,

CβKωiX1(σnω)fβ(ω)fi(ωn+1)(Kσn+1(ω))iX1(σnω)fβ(ω)fi(ωn+1)([0,1])=CαCβCα.C_{\beta}\cap K_{\omega}\subseteq\bigcup_{i\in X^{(\sigma^{n}\omega)}_{1}}f_{\beta}^{(\omega)}\circ f^{(\omega_{n+1})}_{i}\circ(K_{\sigma^{n+1}(\omega)})\subseteq\bigcup_{i\in X^{(\sigma^{n}\omega)}_{1}}f_{\beta}^{(\omega)}\circ f^{(\omega_{n+1})}_{i}([0,1])=\bigcup_{C_{\alpha}\subseteq C_{\beta}}C_{\alpha}.

As for the second assertion, write α1=βi\alpha_{1}=\beta\cdot i and α2=βj\alpha_{2}=\beta\cdot j, where i,jΦω|β|+1i,j\in\Phi_{\omega_{|\beta|+1}}. Then for any x,y[0,1]x,y\in[0,1] there is some z[0,1]z\in[0,1] such that

fα1(x)fα2(y)=fβ(fi(x))fβ(fj(y))=fβ(z)(fi(x)fj(y)).f_{\alpha_{1}}(x)-f_{\alpha_{2}}(y)=f_{\beta}(f_{i}(x))-f_{\beta}(f_{j}(y))=f_{\beta}^{\prime}(z)\cdot(f_{i}(x)-f_{j}(y)).

So, by bounded distortion (4),

|fα1(x)fα2(y)|L|Cβ|dist(Ci,Cj)L|Cβ|minsI,f,gΦsdist(f([0,1]),g([0,1])),\left|f_{\alpha_{1}}(x)-f_{\alpha_{2}}(y)\right|\geq L\cdot|C_{\beta}|\cdot\text{dist}\left(C_{i},\,C_{j}\right)\geq L\cdot|C_{\beta}|\cdot\min_{s\in I,f,g\in\Phi_{s}}\text{dist}\left(f([0,1]),\,g([0,1])\right),

and by Theorem 2.4 part (4)

minsI,f,gΦsdist(f([0,1]),g([0,1]))>0.\min_{s\in I,f,g\in\Phi_{s}}\text{dist}\left(f([0,1]),\,g([0,1])\right)>0.

As required. ∎

Proof of Theorem 2.4 Part (6) Fix r>0r>0, D>1D>1 and xKωx\in K_{\omega}. In general, we aim to show that there exist cylinders CαCβC_{\alpha}\subseteq C_{\beta} such that |α||β||\alpha|-|\beta| depends only on DD, and

CαB(x,r) and B(x,Dr)KωCβ.C_{\alpha}\subseteq B(x,r)\text{ and }B(x,Dr)\cap K_{\omega}\subseteq C_{\beta}.

If this holds then by (19):

μω(B(x,Dr)μω(B(x,Dr)μω(Cβ)μω(Cα)(minjImini{1,,kj}pi(j))|β||α|.\frac{\mu_{\omega}(B(x,Dr)}{\mu_{\omega}(B(x,Dr)}\leq\frac{\mu_{\omega}(C_{\beta})}{\mu_{\omega}(C_{\alpha})}\leq(\min_{j\in I}\min_{i\in\{1,...,k_{j}\}}{p}^{(j)}_{i})^{|\beta|-|\alpha|}.

Note that the latter bound depends on DD, and in particular is uniform in ω\omega.

Set J=B(x,r)J=B(x,r) and J=B(x,Dr)J^{\prime}=B(x,Dr). We first assume JKωCiJ^{\prime}\cap K_{\omega}\subsetneq C_{i} for some 1st generation cylinder of KωK_{\omega}. Set

n=min{j1:CαJ,|α|=j}.n=\min\{j\geq 1:\,\exists C_{\alpha}\subset J^{\prime},\,|\alpha|=j\}.

Then n2n\geq 2. Let CαJC_{\alpha}\subset J^{\prime} be such that |α|=n|\alpha|=n. Let CαCαC_{\alpha}\subset C_{\alpha^{\prime}} with |α|=n1|\alpha^{\prime}|=n-1. By definition of nn CαJC_{\alpha^{\prime}}\not\subset J^{\prime}, so there are two options:

  1. 1.

    If JKωCαJ^{\prime}\cap K_{\omega}\subset C_{\alpha^{\prime}} then by Lemma 2.6 we have

    C1δ1|Cα||Cα||J|.C^{-1}\delta_{1}\left|C_{\alpha^{\prime}}\right|\leq\left|C_{\alpha}\right|\leq\left|J^{\prime}\right|.
  2. 2.

    Otherwise, there exists a cylinder CβC_{\beta^{\prime}} such that |β|=n1|\beta^{\prime}|=n-1 such that CαC_{\alpha^{\prime}} and CβC_{\beta^{\prime}} are consecutive and JKωCαCβJ^{\prime}\cap K_{\omega}\subseteq C_{\alpha^{\prime}}\cup C_{\beta^{\prime}}. Indeed, KωJK_{\omega}\cap J^{\prime} is covered by such cylinders of generation n1n-1 and none of them are included in JJ^{\prime}. Consider now a bigger cylinder CγC_{\gamma} such that CαCβCγC_{\alpha^{\prime}}\cup C_{\beta^{\prime}}\subset C_{\gamma}, and assume that

    |γ|=max{j0:CβCαCβ,|β|=j}.|\gamma|=\max\{j\geq 0:\,\exists C_{\beta}\supset C_{\alpha^{\prime}}\cup C_{\beta^{\prime}},\,|\beta|=j\}.

    The maximality of |γ||\gamma| implies that CαCα′′C_{\alpha^{\prime}}\subset C_{\alpha^{\prime\prime}} and CβCβ′′C_{\beta^{\prime}}\subset C_{\beta^{\prime\prime}} where |α′′|=|β′′|=|γ|+1|\alpha^{\prime\prime}|=|\beta^{\prime\prime}|=|\gamma|+1. Since the gap between CαC_{\alpha^{\prime}} and CβC_{\beta^{\prime}} is included in JJ^{\prime}, by Lemma 2.7 we find that

    |J|dist(Cα,Cβ)dist(Cα′′,Cβ′′)λ|Cγ|.|J^{\prime}|\geq\text{dist}(C_{\alpha^{\prime}},C_{\beta^{\prime}})\geq\text{dist}(C_{\alpha^{\prime\prime}},C_{\beta^{\prime\prime}})\geq\lambda|C_{\gamma}|.

We have just shown that there exists a cylinder CβC_{\beta} such that JKωCβJ^{\prime}\cap K_{\omega}\subseteq C_{\beta} and |Cβ|C|J||C_{\beta}|\leq C^{\prime}\cdot|J^{\prime}|, where CC^{\prime} is independent of r,x,Dr,x,D.

Finally, there exists a decreasing sequence of cylinders CγiC_{\gamma_{i}} such that

xCγiCγ1Cγ0Cβ.x\in C_{\gamma_{i}}\subsetneq...\subsetneq C_{\gamma_{1}}\subsetneq C_{\gamma_{0}}\subsetneq C_{\beta}.

By Lemma 2.6 and the estimate on |Cβ||C_{\beta}| we have

|Cγi|CCδ2|γi||β|Dr.|C_{\gamma_{i}}|\leq C\cdot C^{\prime}\cdot\delta_{2}^{|\gamma_{i}|-|\beta|}\cdot Dr.

So, whenever CCδ2|γi||β|D<1C\cdot C^{\prime}\cdot\delta_{2}^{|\gamma_{i}|-|\beta|}\cdot D<1 we have CγiJC_{\gamma_{i}}\subseteq J. This is the same as asking that

|γi||β|>logDCClogδ2.|\gamma_{i}|-|\beta|>\frac{\log DCC^{\prime}}{\log\delta_{2}}.

Hence there is a cylinder Cγ=CγiC_{\gamma}=C_{\gamma_{i}} as required, such that CαJC_{\alpha}\subset J and |α||β||\alpha|-|\beta| only depends on DD.

Finally, we discuss the case when JKωJ^{\prime}\cap K_{\omega} is not included in a first generation cylinder. Note that xCix\in C_{i} for some i{1,,kω1}i\in\{1,...,k_{\omega_{1}}\}. If CiJC_{i}\subseteq J^{\prime} then |Ci||J||C_{i}|\leq|J^{\prime}| and following the same ideas we find CαJC_{\alpha}\subset J and CαCiC_{\alpha}\subset C_{i} with |α||\alpha| only depending on DD. So,

μω(J)μω(J)1μω(Cα)CD,\frac{\mu_{\omega}(J^{\prime})}{\mu_{\omega}(J)}\leq\frac{1}{\mu_{\omega}(C_{\alpha})}\leq C_{D}^{\prime},

where CDC_{D}^{\prime} only depends on DD (in particular, does not depend on ω\omega). If CiJC_{i}\not\subset J^{\prime} then by a similar gap argument to the one previously used,

|Ci|maxj{1,,kω1}|Cj|minij{1,,kω1}dist(Ci,Cj)|J|.|C_{i}|\leq\frac{\max_{j\in\{1,...,k_{\omega_{1}}\}}|C_{j}|}{\min_{i\neq j\in\{1,...,k_{\omega_{1}}\}}\text{dist}(C_{i},C_{j})}|J^{\prime}|.

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 L2L^{2} contraction estimate

We equip C1([0,1])C^{1}([0,1]) with the norm

φC1=φ+φ.||\varphi||_{C^{1}}=||\varphi||_{\infty}+||\varphi^{\prime}||_{\infty}. (20)

Following Dolgopyat [23, Section 6] and Naud [39, the discussion prior to Lemma 5.2], for every b0b\neq 0 we define yet another norm on C1([0,1])C^{1}([0,1]) via

φ(b)=φ+φ|b|.||\varphi||_{(b)}=||\varphi||_{\infty}+\frac{||\varphi^{\prime}||_{\infty}}{|b|}. (21)

These two norms on C1([0,1])C^{1}([0,1]) 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.

Let Φ\Phi be a C2([0,1])C^{2}([0,1]) IFS satisfying properties (10) and (11), and let ν=ν𝐩\nu=\nu_{\mathbf{p}} be a self-conformal measure. Then there exist C,γ,ϵ,R>0C,\gamma,\epsilon,R>0 and some 0<α<10<\alpha<1 such that for all |b|>R|b|>R, aa\in\mathbb{R} with |a|<ϵ|a|<\epsilon, and nn\in\mathbb{N}

Pa+ibnC1C|b|1+γαn.||P_{a+ib}^{n}||_{C^{1}}\leq C\cdot|b|^{1+\gamma}\cdot\alpha^{n}. (22)

Recall that by Claim 2.1 every not conjugate to linear C2C^{2} 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 Ps,ω,kP_{s,\omega,k} from (15). We first reduce Theorem 2.8 to the following statement about the L2L^{2}-contraction of all parts of the transfer operator.

Proposition 2.9.

Assume the conditions of Theorem 2.8, and let Σ\Sigma be the model from Theorem 2.4. Then there is some NN\in\mathbb{N} and 0<α<10<\alpha<1 such that for s=a+ibs=a+ib with |a||a| small enough and |b||b| large enough:

For every ωΩ\omega\in\Omega,

|Ps,ω,nNW|2𝑑μσnNωαn\int\left|P_{s,\omega,nN}W\right|^{2}\,d\mu_{\sigma^{nN}\omega}\leq\alpha^{n}

for all WC1([0,1])W\in C^{1}([0,1]) with W(b)1||W||_{(b)}\leq 1.

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 fC1([0,1])f\in C^{1}([0,1]) let fL([0,1]||f||_{L([0,1]} denote the Lipschitz constant of ff.

Lemma 2.10.

[13, Proof of Theorem 2.1.1 part (ii)] There exists a constant C2>0C_{2}>0 such that for all fC1([0,1])f\in C^{1}([0,1]), for every ωΩ,x[0,1]\omega\in\Omega,x\in[0,1] and κ𝒫([0,1])\kappa\in\mathcal{P}([0,1]),

|P0,ω,n(f)(x)P0,ω,n(f)(y)𝑑κ(y)|C2ρnfL([0,1]).\left|P_{0,\omega,n}\left(f\right)(x)-\int P_{0,\omega,n}\left(f\right)(y)d\kappa(y)\right|\leq C_{2}\rho^{n}||f||_{L([0,1])}.
Proof.

The operator

μP0,ω,n(f)(y)𝑑μ(y)\mu\mapsto\int P_{0,\omega,n}\left(f\right)(y)d\mu(y)

contracts the space 𝒫([0,1])\mathcal{P}\left([0,1]\right) with the dual Lipschitz metric by a factor of ρn\rho^{n}. Indeed, the IFS {fβ(ω)}βXn(ω)\{f^{(\omega)}_{\beta}\}_{\beta\in X_{n}^{(\omega)}} satisfies that maxβXn(ω)(fβ(ω))ρn\max_{\beta\in X_{n}^{(\omega)}}||\left(f^{(\omega)}_{\beta}\right)^{\prime}||_{\infty}\leq\rho^{n}. The Lemma now follows by noting that P0,ω,n(f)(x)=P0,ω,n(f)(y)𝑑δx(y)P_{0,\omega,n}\left(f\right)(x)=\int P_{0,\omega,n}\left(f\right)(y)\,d\delta_{x}(y), taking C2C_{2} to be the diameter of 𝒫([0,1])\mathcal{P}\left([0,1]\right). ∎

Recall that DD^{\prime} is as in (13).

Lemma 2.11.

(A-priori bounds) There exists C1>0C_{1}>0 such that for all |a||a| small enough and |b||b| large enough, for all fC1([0,1])f\in C^{1}([0,1]), writing s=a+ibs=a+ib,

(Psnf)C1|b|Panf+ρnPan(|f|), and ||\left(P_{s}^{n}f\right)^{\prime}||_{\infty}\leq C_{1}|b|\cdot||P_{a}^{n}f||_{\infty}+\rho^{n}||P_{a}^{n}\left(|f^{\prime}|\right)||_{\infty},\text{ and }
(Ps,ω,nf)C1|b|Pa,ω,nf+ρnPa,ω,n(|f|), for all ω.||\left(P_{s,\omega,n}f\right)^{\prime}||_{\infty}\leq C_{1}|b|\cdot||P_{a,\omega,n}f||_{\infty}+\rho^{n}||P_{a,\omega,n}\left(|f^{\prime}|\right)||_{\infty},\,\text{ for all }\omega.

In particular, there exists a constant C6>0C_{6}>0 such that for all n0n\geq 0 and s=a+ibs=a+ib

Psn(b)C6enD|a|.||P_{s}^{n}||_{(b)}\leq C_{6}\cdot e^{nD^{\prime}|a|}.
Proof.

By Definition 2.3 (Psnf)(x)\left(P_{s}^{n}f\right)^{\prime}(x) equals

|I|=ne(a+ib)2πc(I,x)(a+ib)2π(logfI)(x)𝐩(I)ffI(x)+e(a+ib)2πc(I,x)𝐩(I)(ffI)(x).\sum_{|I|=n}e^{(a+ib)\cdot 2\pi\cdot c(I,x)}(a+ib)2\pi\cdot\left(\log f_{I}^{\prime}\right)^{\prime}(x)\mathbf{p}(I)f\circ f_{I}(x)+e^{(a+ib)\cdot 2\pi\cdot c(I,x)}\mathbf{p}(I)\left(f\circ f_{I}\right)^{\prime}(x).

So, the first and second assertions follows by noting that, as in e.g. [3, proof of Claim 2.12]

supx[0,1],I𝒜|ddx(logfI)(x)|=C1<.\sup_{x\in[0,1],I\in\mathcal{A}^{*}}\left|\frac{d}{dx}\left(\log f_{I}^{\prime}\right)(x)\right|=C_{1}<\infty.

For the last one, if f(b)1||f||_{(b)}\leq 1 then f1||f||_{\infty}\leq 1 and so f|b|||f^{\prime}||_{\infty}\leq|b|. Thus, by the first assertion, and since PsnfenD|a|||P_{s}^{n}f||_{\infty}\leq e^{nD^{\prime}|a|} as f1||f||_{\infty}\leq 1, we have

Psnf(b)=Psnf+(Psnf)|b|enD|a|+C1enD|a|+enD|a|=C6enD|a|.||P_{s}^{n}f||_{(b)}=||P_{s}^{n}f||_{\infty}+\frac{||\left(P_{s}^{n}f\right)^{\prime}||_{\infty}}{|b|}\leq e^{nD^{\prime}|a|}+C_{1}e^{nD^{\prime}|a|}+e^{nD^{\prime}|a|}=C_{6}e^{nD^{\prime}|a|}.

Proof that Proposition 2.9 implies Theorem 2.8 Fix s=a+ibs=a+ib and N>0N>0 as in Proposition 2.9. Set

n=2[CNlog|b|],n~=[CNlog|b|], with C>0 to be chosen later.n=2[\frac{C}{N}\log|b|],\quad\tilde{n}=[\frac{C}{N}\log|b|],\quad\text{ with }C>0\text{ to be chosen later.}

For all ss\in\mathbb{C} and fC1([0,1])f\in C^{1}([0,1]) with f(b)1||f||_{(b)}\leq 1 and all x[0,1]x\in[0,1] we have, by (18),

|PsnN(f)|(x)ωΩnN([ω])|Ps,ω,nN(f)|(x).\left|P_{s}^{nN}\left(f\right)\right|(x)\leq\sum_{\omega\in\Omega^{nN}}\mathbb{Q}\left([\omega]\right)\left|P_{s,\omega,nN}\left(f\right)\right|(x). (23)

Now, fix ωΩ\omega\in\Omega. Then, by (17),

|Ps,ω,nN(f)|(x)=|Ps,σn~Nω,nNn~N(Ps,ω,n~N(f))(x)|Pa,σn~Nω,nNn~N(|Ps,ω,n~N(f)|)(x).\left|P_{s,\omega,nN}\left(f\right)\right|(x)=\left|P_{s,\sigma^{\tilde{n}N}\omega,nN-\tilde{n}N}\left(P_{s,\omega,\tilde{n}N}\left(f\right)\right)\left(x\right)\right|\leq P_{a,\sigma^{\tilde{n}N}\omega,nN-\tilde{n}N}\left(\left|P_{s,\omega,\tilde{n}N}\left(f\right)\right|\right)(x).

Set m=(nn~)Nm=(n-\tilde{n})N. Note that

Pa,σn~Nω,m(|Ps,ω,n~N(f)|)(x)=IXm(σn~Nω)e2πac(I,x)η(σn~Nω)(I)(η(σn~Nω)(I)|Ps,ω,n~N(f)|fI(x)).P_{a,\sigma^{\tilde{n}N}\omega,m}\left(\left|P_{s,\omega,\tilde{n}N}\left(f\right)\right|\right)(x)=\sum_{I\in X^{(\sigma^{\tilde{n}N}\omega)}_{m}}e^{2\pi ac(I,x)}\sqrt{\eta^{(\sigma^{\tilde{n}N}\omega)}(I)}\cdot\left(\sqrt{\eta^{(\sigma^{\tilde{n}N}\omega)}(I)}\left|P_{s,\omega,\tilde{n}N}\left(f\right)\right|\circ f_{I}(x)\right).

Then by Cauchy-Schwartz,

(Pa,σn~Nω,m(|Ps,ω,n~N(f)|)(x))2e2m|a|DP0,σn~Nω,m(|Ps,ω,n~N(f)|2)(x).\left(P_{a,\sigma^{\tilde{n}N}\omega,m}\left(\left|P_{s,\omega,\tilde{n}N}\left(f\right)\right|\right)(x)\right)^{2}\leq e^{2m|a|\cdot D^{\prime}}\cdot P_{0,\sigma^{\tilde{n}N}\omega,m}\left(\left|P_{s,\omega,\tilde{n}N}\left(f\right)\right|^{2}\right)(x).

We wish to invoke Lemma 2.10, so we must estimate |Ps,ω,n~N(f)|2L([0,1])||\left|P_{s,\omega,\tilde{n}N}\left(f\right)\right|^{2}||_{L([0,1])}: By Lemma 2.11, using that f1||f||_{\infty}\leq 1 and f|b|||f^{\prime}||_{\infty}\leq|b|,

|Ps,ω,n~N(f)|2L([0,1])2Ps,ω,n~N(f)(Ps,ω,n~N(f))||\left|P_{s,\omega,\tilde{n}N}\left(f\right)\right|^{2}||_{L([0,1])}\leq 2\cdot||P_{s,\omega,\tilde{n}N}\left(f\right)||_{\infty}\cdot||\left(P_{s,\omega,\tilde{n}N}\left(f\right)\right)^{\prime}||_{\infty}
2en~N|a|D(C1|b|Pa,ω,n~Nf+ρn~NPa,ω,n~N(|f|))\leq 2e^{\tilde{n}N|a|D^{\prime}}\cdot\left(C_{1}|b|\cdot||P_{a,\omega,\tilde{n}N}f||_{\infty}+\rho^{\tilde{n}N}||P_{a,\omega,\tilde{n}N}\left(|f^{\prime}|\right)||_{\infty}\right)
2en~NaD(C1|b|en~N|a|D+ρn~N|b|en~N|a|D)\leq 2e^{\tilde{n}NaD^{\prime}}\cdot\left(C_{1}|b|e^{\tilde{n}N|a|D^{\prime}}+\rho^{\tilde{n}N}\cdot|b|\cdot e^{\tilde{n}N|a|D^{\prime}}\right)
C3|b|en~N2|a|D\leq C_{3}|b|e^{\tilde{n}N2|a|D^{\prime}}

Via Lemma 2.10 applied with κ=μσnN(ω)\kappa=\mu_{\sigma^{nN}(\omega)}, (16), and the previous calculation we have

Ps,ω,nN(f)2e2m|a|D(P0,σn~Nω,m(|Ps,ω,n~N(f)|2)𝑑μσnN(ω)+C2ρm|Ps,ω,n~N(f)|2L([0,1]))\|P_{s,\omega,nN}\left(f\right)\|^{2}_{\infty}\leq e^{2m|a|\cdot D^{\prime}}\left(\int P_{0,\sigma^{\tilde{n}N}\omega,m}\left(\left|P_{s,\omega,\tilde{n}N}\left(f\right)\right|^{2}\right)\,d\mu_{\sigma^{nN}(\omega)}+C_{2}\rho^{m}||\left|P_{s,\omega,\tilde{n}N}\left(f\right)\right|^{2}||_{L([0,1])}\right)
|b|C|a|2D(|Ps,ω,n~N(f)|2𝑑μσn~N(ω)+C3C2ρm|b|en~N2|a|D)\leq|b|^{C|a|2D^{\prime}}\left(\int\left|P_{s,\omega,\tilde{n}N}\left(f\right)\right|^{2}\,d\mu_{\sigma^{\tilde{n}N}(\omega)}+C_{3}C_{2}\rho^{m}|b|e^{\tilde{n}N2|a|D^{\prime}}\right)
|b|C|a|2D(|Ps,ω,n~N(f)|2𝑑μσn~N(ω)+C3C2ρm|b|1+2|a|DC).\leq|b|^{C|a|2D^{\prime}}\left(\int\left|P_{s,\omega,\tilde{n}N}\left(f\right)\right|^{2}\,d\mu_{\sigma^{\tilde{n}N}(\omega)}+C_{3}C_{2}\rho^{m}|b|^{1+2|a|D^{\prime}C}\right).

Applying Proposition 2.9,

Ps,ω,nN(f)2|b|C|a|2D(C4|b|CN|logα|+C5|b|C|logρ|12|a|DC),||P_{s,\omega,nN}\left(f\right)||_{\infty}^{2}\leq|b|^{C|a|2D^{\prime}}\left(\frac{C_{4}}{|b|^{\frac{C}{N}|\log\alpha|}}+\frac{C_{5}}{|b|^{C|\log\rho|-1-2|a|D^{\prime}C}}\right),

where C4,C5C_{4},C_{5} are positive constants. Recalling that |a||a| is already assumed to be small, we can now choose CC with C|logρ|>1+2|a|DCC|\log\rho|>1+2|a|D^{\prime}C and |a||a| even closer to 0 so that if |a||a| is small enough and |b||b| is large enough,

Ps,ω,nN(f)1|b|β||P_{s,\omega,nN}\left(f\right)||_{\infty}\leq\frac{1}{|b|^{\beta}}

for some β>0\beta>0. Using that the same bound works for every term in the convex combination (23), we obtain

PsnNf1|b|β.||P_{s}^{nN}f||_{\infty}\leq\frac{1}{|b|^{\beta}}.

Next, by (18)

(PsnN(f))(x)=ωΩnN([ω])(Ps,ω,nN(f))(x).\left(P_{s}^{nN}\left(f\right)\right)^{\prime}(x)=\sum_{\omega\in\Omega^{nN}}\mathbb{Q}\left([\omega]\right)\left(P_{s,\omega,nN}\left(f\right)\right)^{\prime}(x). (24)

So, since f|b|||f^{\prime}||_{\infty}\leq|b| (as f(b)1||f||_{(b)}\leq 1) we have, as in Lemma 2.11, for some c1>0c_{1}>0

|(PsnN(f))(x)|ωΩnN([ω])(|b|c1|Ps,ω,nN(f)(x)|+ρnNPa,ω,nN(|f|)(x))\left|\left(P_{s}^{nN}\left(f\right)\right)^{\prime}(x)\right|\leq\sum_{\omega\in\Omega^{nN}}\mathbb{Q}\left([\omega]\right)\left(|b|\cdot c_{1}\cdot\left|P_{s,\omega,nN}\left(f\right)(x)\right|+\rho^{nN}P_{a,\omega,{nN}}\left(\left|f^{\prime}\right|\right)(x)\right)
c1|b|ωΩnN([ω])|Ps,ω,nN(f)(x)|+ρnN|b|enND|a|.\leq c_{1}\cdot|b|\sum_{\omega\in\Omega^{nN}}\mathbb{Q}\left([\omega]\right)\left|P_{s,\omega,nN}\left(f\right)(x)\right|+\rho^{nN}\cdot|b|\cdot e^{nND^{\prime}|a|}.

Thus,

1|b|(PsnNf)c1ωΩnN([ω])Ps,ω,nN(f)+ρnNenND|a|\frac{1}{|b|}||\left(P_{s}^{nN}f\right)^{\prime}||_{\infty}\leq c_{1}\sum_{\omega\in\Omega^{nN}}\mathbb{Q}\left([\omega]\right)||P_{s,\omega,nN}\left(f\right)||_{\infty}+\rho^{nN}e^{nND^{\prime}|a|}
c1ωΩnN([ω])Ps,ω,nN(f)+1ρ2N|b|2C(D|a||logρ|).\leq c_{1}\sum_{\omega\in\Omega^{nN}}\mathbb{Q}\left([\omega]\right)||P_{s,\omega,nN}\left(f\right)||_{\infty}+\frac{1}{\rho^{2N}}\cdot|b|^{2C(D^{\prime}|a|-|\log\rho|)}.

So, using similar ideas, for |a||a| small and |b||b| large and possibly a large CC,

PsNn(b)1|b|β,||P_{s}^{Nn}||_{(b)}\leq\frac{1}{|b|^{\beta^{\prime}}},

for some β>0\beta^{\prime}>0 and n=[CNlog|b|]n=[\frac{C}{N}\log|b|].

Finally, given mm\in\mathbb{N} we write

m=dN[CNlog|b|]+r for the corresponding d,0rN[CNlog|b|].m=d\cdot N[\frac{C}{N}\log|b|]+r\text{ for the corresponding }d,0\leq r\leq N\cdot[\frac{C}{N}\log|b|].

Let β′′>0\beta^{\prime\prime}>0 be such that CD|a|β<β′′CD^{\prime}|a|-\beta^{\prime}<-\beta^{\prime\prime} for all small enough |a||a|. Applying Lemma 2.11,

Psm(b)Psr(b)PsdN[CNlog|b|](b)C6eC(log|b|)D|a|(1|b|β)d||P_{s}^{m}||_{(b)}\leq||P_{s}^{r}||_{(b)}\cdot||P_{s}^{dN\cdot[\frac{C}{N}\log|b|]}||_{(b)}\leq C_{6}\cdot e^{C(\log|b|)\cdot D^{\prime}|a|}\cdot\left(\frac{1}{|b|^{\beta^{\prime}}}\right)^{d}
C6(|b|CD|a||b|β)dC6(1|b|β′′)dC6|b|β′′ρβ′′m,\leq C_{6}\cdot\left(\frac{|b|^{CD^{\prime}|a|}}{|b|^{\beta^{\prime}}}\right)^{d}\leq C_{6}\cdot\left(\frac{1}{|b|^{\beta^{\prime\prime}}}\right)^{d}\leq C_{6}\cdot|b|^{\beta^{\prime\prime}}\rho_{\beta^{\prime\prime}}^{m},

for some ρβ′′(0,1)\rho_{\beta^{\prime\prime}}\in(0,1). This is true for all β′′>0\beta^{\prime\prime}>0 with β′′\beta^{\prime\prime} small, and since

||||C1|b|||||(b)||\cdot||_{C^{1}}\leq|b|\cdot||\cdot||_{(b)}

Theorem 2.3 follows. \Box

2.5 The key Lemma

Given A>0A>0 let

𝒞A={fC1([0,1]):f>0, and |f(x)|Af(x),x[0,1]}.\mathcal{C}_{A}=\{f\in C^{1}([0,1]):\,f>0,\text{ and }\left|f^{\prime}(x)\right|\leq A\cdot f(x),\,\forall x\in[0,1]\}.

Note: If f𝒞Af\in\mathcal{C}_{A} and u,v[0,1]u,v\in[0,1] then

eA|uv|f(u)f(v)eA|uv|.e^{-A|u-v|}\leq\frac{f(u)}{f(v)}\leq e^{A|u-v|}.

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 N>0,A>1N>0,A>1 and 0<α<10<\alpha<1 such that for all s=a+ibs=a+ib with |a||a| small and |b||b| large:

For every ωΩ\omega\in\Omega there exist a finite set of bounded operators {NsJ}Js,ω\{N_{s}^{J}\}_{J\in\mathcal{E}_{s,\omega}} on C1([0,1])C^{1}([0,1]) such that:

  1. 1.

    The cone 𝒞A|b|\mathcal{C}_{A|b|} is stable under NsJN_{s}^{J} for all Js,ωJ\in\mathcal{E}_{s,\omega}.

  2. 2.

    For all H𝒞A|b|H\in\mathcal{C}_{A|b|} and Js,ωJ\in\mathcal{E}_{s,\omega},

    |NsJH|2𝑑μσNωαP0,ω,N(|H2|)𝑑μσNω.\int|N_{s}^{J}H|^{2}d\mu_{\sigma^{N}\omega}\leq\alpha\cdot\int P_{0,\omega,N}\left(\left|H^{2}\right|\right)d\mu_{\sigma^{N}\omega}.
  3. 3.

    Let H𝒞A|b|H\in\mathcal{C}_{A|b|} and fC1([0,1])f\in C^{1}([0,1]) be such that |f|H|f|\leq H and |f|A|b|H|f^{\prime}|\leq A|b|H. Then for every ωΩ\omega\in\Omega there exists Js,ωJ\in\mathcal{E}_{s,\omega} such that

    |Ps,ω,Nf|NsJH and |(Ps,ω,Nf)|A|b|NsJH.|P_{s,\omega,N}f|\leq N_{s}^{J}H\quad\text{ and }|(P_{s,\omega,N}f)^{\prime}|\leq A|b|N_{s}^{J}H.

As is customary in the literature (e.g. [39, 49]) we will refer to the operators constructed in Lemma 2.12 as Dolgopyat operators.

Proof that Lemma 2.12 implies Proposition 2.9 First, observe that for every ωΩ,s\omega\in\Omega,s\in\mathbb{C} and N,nN,n\in\mathbb{N} we have, by (17)

Ps,ω,nN=Ps,σnNω,NPs,σNω,NPs,ω,N.P_{s,\omega,nN}=P_{s,\sigma^{nN}\omega,N}\circ...\circ P_{s,\sigma^{N}\omega,N}\circ P_{s,\omega,N}.

Let fC1([0,1])f\in C^{1}([0,1]) with f0f\neq 0 and f(b)1||f||_{(b)}\leq 1. Put H=f(b)H=||f||_{(b)}. Then |f|H|f|\leq H and |f||b|f(b)A|b|H|f^{\prime}|\leq|b|\cdot||f||_{(b)}\leq A|b|H. By induction, for all n1n\geq 1 and every ω\omega there are Jωs,ω,JσNωs,σNω,,JσnNωs,σNnωJ_{\omega}\in\mathcal{E}_{s,\omega},J_{\sigma^{N}\omega}\in\mathcal{E}_{s,\sigma^{N}\omega},...,J_{\sigma^{nN}\omega}\in\mathcal{E}_{s,\sigma^{Nn}\omega} such that

|Ps,ω,nNf|NsJσnN(ω)NsJσ(n1)N(ω).NsJωH and |(Ps,ω,nN)|A|b|NsJσnN(ω)NsJσ(n1)N(ω).NsJωH.\left|P_{s,\omega,nN}f\right|\leq N_{s}^{J_{\sigma^{nN}\left(\omega\right)}}N_{s}^{J_{\sigma^{(n-1)N}\left(\omega\right)}}....N_{s}^{J_{\omega}}H\text{ and }\left|\left(P_{s,\omega,nN}\right)^{\prime}\right|\leq A|b|N_{s}^{J_{\sigma^{nN}\left(\omega\right)}}N_{s}^{J_{\sigma^{(n-1)N}\left(\omega\right)}}....N_{s}^{J_{\omega}}H.

So, by Lemma 2.12 and our equivariance relations,

|Ps,ω,nNf|2𝑑μσnNω\displaystyle\int\left|P_{s,\omega,nN}f\right|^{2}d\mu_{\sigma^{nN}\omega} =\displaystyle= |NsJn,ωN2Jn1,ω.NsJ1,ωH|2dμσnNω\displaystyle\int|N_{s}^{J_{n,\omega}}N_{2}^{J_{n-1,\omega}}....N_{s}^{J_{1,\omega}}H|^{2}d\mu_{\sigma^{nN}\omega}
\displaystyle\leq αP0,σN(n1)ω,N(|N2Jn1,ω.NsJ1,ωH|2)dμσnNω\displaystyle\alpha\cdot\int P_{0,\sigma^{N(n-1)}\omega,N}\left(|N_{2}^{J_{n-1,\omega}}....N_{s}^{J_{1,\omega}}H|^{2}\right)d\mu_{\sigma^{nN}\omega}
=\displaystyle= α(|N2Jn1,ω.NsJ1,ωH|2)dμσ(n1)Nω\displaystyle\alpha\cdot\int\left(|N_{2}^{J_{n-1,\omega}}....N_{s}^{J_{1,\omega}}H|^{2}\right)d\mu_{\sigma^{(n-1)N}\omega}
\displaystyle\leq αnP0,ω,N(|H|2)𝑑μσNω\displaystyle\alpha^{n}\cdot\int P_{0,\omega,N}\left(|H|^{2}\right)d\mu_{\sigma^{N}\omega}
=\displaystyle= αn|H|2𝑑μω\displaystyle\alpha^{n}\cdot\int|H|^{2}d\mu_{\omega}
\displaystyle\leq αn.\displaystyle\alpha^{n}.

\hfill{\Box}

2.6 The triple intersections property

The following Proposition allows for the construction, for every ω\omega, of a special partition of [0,1][0,1] that has the triple intersections property on KωK_{\omega}: This means that whenever a cell intersects KωK_{\omega}, two other nearby cells must also intersect KωK_{\omega}. 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 ω\omega, certain metric features of them, e.g. the size of the cells and the distance of the endpoints from KωK_{\omega}, are uniform across our model Σ\Sigma.

Proposition 2.13.

There exist constants A1,A1,A2>0A_{1}^{\prime},A_{1},A_{2}>0 such that for all ω\omega and every ϵ>0\epsilon>0 small enough:

There exists a finite collection of closed intervals (Vi)1iq(V_{i})_{1\leq i\leq q} ordered along [0,1][0,1] such that:

  1. 1.

    [0,1]=i=1qVi[0,1]=\bigcup_{i=1}^{q}V_{i}, and ijInt ViInt Vj=i\neq j\Rightarrow\text{Int }V_{i}\cap\text{Int }V_{j}=\emptyset.

  2. 2.

    For all 1iq1\leq i\leq q we have ϵA1|Vi|ϵA1\epsilon A_{1}^{\prime}\leq|V_{i}|\leq\epsilon A_{1}.

  3. 3.

    For all 1jq1\leq j\leq q such that VjKωV_{j}\cap K_{\omega}\neq\emptyset, either Vj1KωV_{j-1}\cap K_{\omega}\neq\emptyset and Vj+1KωV_{j+1}\cap K_{\omega}\neq\emptyset, or Vj2KωV_{j-2}\cap K_{\omega}\neq\emptyset and Vj1KωV_{j-1}\cap K_{\omega}\neq\emptyset, or Vj+1KωV_{j+1}\cap K_{\omega}\neq\emptyset and Vj+2KωV_{j+2}\cap K_{\omega}\neq\emptyset.

  4. 4.

    For all 1iq1\leq i\leq q such that ViKωV_{i}\cap K_{\omega}\neq\emptyset, we have

    dist(Vi,Kω)A2|Vi|\text{dist}(\partial V_{i},\,K_{\omega})\geq A_{2}|V_{i}|

It is critical to our argument is that the constants A1,A1,A2A_{1}^{\prime},A_{1},A_{2} 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 ω\omega and recall Definition 2.5 (cylinders of KωK_{\omega}). We require the following Lemmas:

Lemma 2.14.

There exists a constant B1B_{1} such that for all xKωx\in K_{\omega} and all r>0r>0, there exists a cylinder CαC_{\alpha} such that

CαB(x,r) and B1r|Cα|.C_{\alpha}\subseteq B(x,r)\text{ and }B_{1}\cdot r\leq|C_{\alpha}|.
Proof.

Consider a cylinder CαC_{\alpha} with xCαB(x,r)x\in C_{\alpha}\subseteq B(x,r), and assume |α||\alpha| is minimal. Since xKωx\in K_{\omega} such a cylinder exists. Let CαCβC_{\alpha}\subseteq C_{\beta} with |β|=|α|1|\beta|=|\alpha|-1. By minimality of |α||\alpha| we cannot have CβC_{\beta} included in B(x,r)B(x,r), and thus |Cβ|r|C_{\beta}|\geq r. Via Lemma 2.6 we obtain

C1δ1rC1δ1|Cβ||Cα|C^{-1}\delta_{1}r\leq C^{-1}\delta_{1}|C_{\beta}|\leq|C_{\alpha}|

as claimed. ∎

Lemma 2.15.

Let CβC_{\beta} be a cylinder. Then there a finite set of at least 33 words AβA_{\beta} such that

CβKωγAβCγ,C_{\beta}\cap K_{\omega}\subseteq\bigcup_{\gamma\in A_{\beta}}C_{\gamma},

where CγCβC_{\gamma}\subseteq C_{\beta} and |γ|=|β|+2|\gamma|=|\beta|+2.

Proof.

By definition, there is some βXn(ω)\beta\in X_{n}^{(\omega)} such that

Cβ:=fβ1(ω1)fβn(ωn)([0,1]).C_{\beta}:=f^{(\omega_{1})}_{\beta_{1}}\circ\circ\circ f^{(\omega_{n})}_{\beta_{n}}([0,1]).

Furthermore, by Theorem 2.4 Part (4)

CβKω=fβ1(ω1)fβn(ωn)(Kσn(ω)).C_{\beta}\cap K_{\omega}=f^{(\omega_{1})}_{\beta_{1}}\circ\circ\circ f^{(\omega_{n})}_{\beta_{n}}(K_{\sigma^{n}(\omega)}).

Now, by the definition of the model,

Kσn(ω)=ζX2(σnω)fζ1(ωn+1)fζ2(ωn+2)(Kσn+2(ω)),K_{\sigma^{n}(\omega)}=\bigcup_{\zeta\in X^{(\sigma^{n}\omega)}_{2}}f^{(\omega_{n+1})}_{\zeta_{1}}\circ f^{(\omega_{n+2})}_{\zeta_{2}}(K_{\sigma^{n+2}(\omega)}),

and 4|X2(σnω)|94\leq\left|X^{(\sigma^{n}\omega)}_{2}\right|\leq 9 by Theorem 2.4 Part (3). Note that we are also using Theorem 2.4 Part (4) to see that X2(σnω)X^{(\sigma^{n}\omega)}_{2} does not have exact overlaps (maps with different coding are not equal). Thus,

CβKωγX2(σnω)fβ1(ω1)fβn(ωn)fζ1(ωn+1)fζ2(ωn+2)([0,1]),C_{\beta}\cap K_{\omega}\subseteq\bigcup_{\gamma\in X^{(\sigma^{n}\omega)}_{2}}f^{(\omega_{1})}_{\beta_{1}}\circ\circ\circ f^{(\omega_{n})}_{\beta_{n}}\circ f^{(\omega_{n+1})}_{\zeta_{1}}\circ f^{(\omega_{n+2})}_{\zeta_{2}}([0,1]),

which is a union of 4k94\leq k\leq 9 cylinders that are contained in CβC_{\beta}. As required. ∎

Remark 2.16.

By Lemma 2.6, there are uniform constants B2,B3>0B_{2},B_{3}>0 independent of β,Aβ\beta,A_{\beta}, such that for all γAβ\gamma\in A_{\beta} we have

B2|Cβ||Cγ|B3|Cβ|.B_{2}|C_{\beta}|\leq|C_{\gamma}|\leq B_{3}|C_{\beta}|.

Applying Lemma 2.7, or Theorem 2.4 Part (4) directly, there is a uniform constant B4>0B_{4}>0 independent of β,Aβ\beta,A_{\beta}, such that for all γ1γ2Aβ\gamma_{1}\neq\gamma_{2}\in A_{\beta} we have

dist(Cγ1,Cγ2)B4|Cβ|.\text{dist}\left(C_{\gamma_{1}},\,C_{\gamma_{2}}\right)\geq B_{4}\left|C_{\beta}\right|.

Proof of Proposition 2.13 Let 0<ϵ10<\epsilon\ll 1. Set p:=[1ϵ]p:=[\frac{1}{\epsilon}]. First, we divide [0,1][0,1] into pp closed intervals JiJ_{i} with disjoint interiors such that

[0,1]=i=1pJi and ϵ|Ji|2ϵ.[0,1]=\bigcup_{i=1}^{p}J_{i}\text{ and }\epsilon\leq\left|J_{i}\right|\leq 2\epsilon.

For every 1ip1\leq i\leq p write Ji=[xi,xi+1]J_{i}=[x_{i},x_{i+1}]. We may assume x1=0x_{1}=0 and xp+1=1x_{p+1}=1.

We first deal with part (4): For 2ip2\leq i\leq p, if B(xi,ϵ8)Kω=B(x_{i},\frac{\epsilon}{8})\cap K_{\omega}=\emptyset we set x~i:=xi\tilde{x}_{i}:=x_{i}. Otherwise, let xiB(xi,ϵ8)Kωx_{i}^{\prime}\in B(x_{i},\frac{\epsilon}{8})\cap K_{\omega}. Applying Lemma 2.14, we find a cylinder CαB(xi,ϵ8)C_{\alpha}\subseteq B(x_{i}^{\prime},\frac{\epsilon}{8}) with |Cα|B1ϵ8|C_{\alpha}|\geq B_{1}\frac{\epsilon}{8}. By Lemma 2.15, there are consecutive cylinders Cγ1,Cγ2CαC_{\gamma_{1}},C_{\gamma_{2}}\subset C_{\alpha} such that

dist(Cγ1,Cγ2)B4|Cα|.\text{dist}\left(C_{\gamma_{1}},\,C_{\gamma_{2}}\right)\geq B_{4}\left|C_{\alpha}\right|.

Set

x~i:=12(maxCγ1+minCγ2).\tilde{x}_{i}:=\frac{1}{2}\left(\max C_{\gamma_{1}}+\min C_{\gamma_{2}}\right).

In both cases

|x~ixi|ϵ4 and dist(x~i,Kω)min(ϵ8,B1B4ϵ16).\left|\tilde{x}_{i}-x_{i}\right|\leq\frac{\epsilon}{4}\text{ and }\text{dist}\left(\tilde{x}_{i},K_{\omega}\right)\geq\min\left(\frac{\epsilon}{8},\,B_{1}B_{4}\frac{\epsilon}{16}\right).

As for the boundary points x1=0,xp+1=1x_{1}=0,x_{p+1}=1, by (3)

dist(Kω,[0,1])dist(K,[0,1])>0.\text{dist}\left(K_{\omega},\,\partial[0,1]\right)\geq\text{dist}\left(K,\,\partial[0,1]\right)>0.

So, upon taking ϵ1\epsilon\ll 1, we may put x~1=0\tilde{x}_{1}=0 and x~p+1=1\tilde{x}_{p+1}=1.

For all 1ip1\leq i\leq p set J~i:=[x~i,x~i+1]\tilde{J}_{i}:=[\tilde{x}_{i},\tilde{x}_{i+1}]. Then, writing B5=25min(B1B416,18)B_{5}=\frac{2}{5}\min\left(\frac{B_{1}B_{4}}{16},\frac{1}{8}\right) we have

ϵ2|J~i|52ϵ,dist(J~i,Kω)B5|J~i|, and [0,1]i=1pJ~i.\frac{\epsilon}{2}\leq\left|\tilde{J}_{i}\right|\leq\frac{5}{2}\epsilon,\,\text{dist}\left(\partial\tilde{J}_{i},K_{\omega}\right)\geq B_{5}\left|\tilde{J}_{i}\right|,\text{ and }[0,1]\subseteq\bigcup_{i=1}^{p}\tilde{J}_{i}.

Furthermore, the intervals J~i\tilde{J}_{i} still have disjoint interiors. Thus, this collection of intervals satisfies parts (1),(2), and (4) of the Proposition. Let us call these intervals JiJ_{i}.

We now deal with property (3): Fix JiJ_{i} such that KωJiK_{\omega}\cap J_{i}\neq\emptyset. Let xKωJix\in K_{\omega}\cap J_{i}, so that B(x,B5ϵ2)JiB(x,B_{5}\frac{\epsilon}{2})\subset J_{i}. By Lemma 2.14 and Lemma 2.15 there are 33 consecutive cylinders Cγ1,Cγ2,Cγ3C_{\gamma_{1}},C_{\gamma_{2}},C_{\gamma_{3}} such that

Cγ1Cγ2Cγ3B(x,B5ϵ2)Ji.C_{\gamma_{1}}\cup C_{\gamma_{2}}\cup C_{\gamma_{3}}\subseteq B(x,B_{5}\frac{\epsilon}{2})\subset J_{i}.

In addition, we have for i=1,2i=1,2

dist(Cγ1,Cγ2)B4B1B5ϵ2.\text{dist}\left(C_{\gamma_{1}},C_{\gamma_{2}}\right)\geq B_{4}\cdot B_{1}\cdot B_{5}\frac{\epsilon}{2}.

We set

yi:=12(maxCγ1+minCγ2) and zi:=12(maxCγ2+minCγ3),y_{i}:=\frac{1}{2}\left(\max C_{\gamma_{1}}+\min C_{\gamma_{2}}\right)\text{ and }z_{i}:=\frac{1}{2}\left(\max C_{\gamma_{2}}+\min C_{\gamma_{3}}\right),

and write

Ji1=[xi,yi],Ji2=[yi,zi],Ji3=[zi,xi+1].J_{i}^{1}=[x_{i},\,y_{i}],\,J_{i}^{2}=[y_{i},\,z_{i}],\,J_{i}^{3}=[z_{i},\,x_{i+1}].

Then for all j=1,2,3j=1,2,3 we have JijKJ_{i}^{j}\cap K\neq\emptyset, and

B1B2B5ϵ2|Jij|52ϵ and dist(J~ij,Kω)min(B5B4B1ϵ4,B5ϵ2).B_{1}B_{2}B_{5}\frac{\epsilon}{2}\leq\left|J_{i}^{j}\right|\leq\frac{5}{2}\epsilon\text{ and }\text{dist}\left(\partial\tilde{J}_{i}^{j},K_{\omega}\right)\geq\min\left(B_{5}B_{4}B_{1}\frac{\epsilon}{4},\,B_{5}\frac{\epsilon}{2}\right).

Finally, the set of intervals

{Ji:JiKω=}j=13{Jij:JiKω}\{J_{i}:\,J_{i}\cap K_{\omega}=\emptyset\}\bigcup_{j=1}^{3}\{J_{i}^{j}:\,J_{i}\cap K_{\omega}\neq\emptyset\}

now satisfy all the properties in Proposition 2.13. \Box

2.7 Proof of Lemma 2.12

Fix ω\omega and let s=a+ibs=a+ib. We begin by constructing the Dolgopyat operators as in Lemma 2.12. Let NN\in\mathbb{N} be sufficiently large in the sense of Theorem 2.4 part (5), and in other ways that will be specified soon, and let α1N,α2NXN(ω)\alpha_{1}^{N},\alpha_{2}^{N}\in X_{N}^{(\omega)} be the length NN words satisfying the conclusion of Theorem 2.4 Part (5). Let us fix ϵ,1|b|>0\epsilon^{\prime},\frac{1}{|b|}>0 uniformly in ω\omega, that are small enough (to be determined later). Let (Vi)1iq(V_{i})_{1\leq i\leq q} be a triadic partition as in Proposition 2.13 of KσNωK_{\sigma^{N}\omega} of modulus ϵ=ϵ|b|\epsilon=\frac{\epsilon^{\prime}}{|b|}. For all (i,j){1,2}×{1,,q}(i,j)\in\{1,2\}\times\{1,...,q\} set

Zji=fαiN(Vj).Z_{j}^{i}=f_{\alpha_{i}^{N}}\left(V_{j}\right).

By Proposition 2.13 part (4)

dist(KσNωVj,Vj)A2A1ϵ|b| whenever KσNωVj.\text{dist}\left(K_{\sigma^{N}\omega}\cap V_{j},\,\partial V_{j}\right)\geq A_{2}A_{1}^{\prime}\frac{\epsilon^{\prime}}{|b|}\text{ whenever }K_{\sigma^{N}\omega}\cap V_{j}\neq\emptyset.

Then there exists a cut off function χjC1([0,1])\chi_{j}\in C^{1}([0,1]) such that 0χj10\leq\chi_{j}\leq 1 on [0,1][0,1], χj=1\chi_{j}=1 on conv(KσNωVj)\text{conv}\left(K_{\sigma^{N}\omega}\cap V_{j}\right), and χj=0\chi_{j}=0 outside of Int(Vj)\text{Int}(V_{j}). Then there exists A3>0A_{3}>0 that depends only on the previous (uniform in ω\omega) constants such that

χjA3|b|ϵ.||\chi_{j}^{\prime}||_{\infty}\leq A_{3}\frac{|b|}{\epsilon^{\prime}}.

Define

Js,ω={(i,j){1,2}×{1,,q}:VjKσNω}.J_{s,\omega}=\{(i,j)\in\{1,2\}\times\{1,...,q\}:\,V_{j}\cap K_{\sigma^{N}\omega}\neq\emptyset\}.

Fix 0<θ<10<\theta<1 to be determined later. Let JJs,ω\emptyset\neq J\subseteq J_{s,\omega}. Define a function χJC1([0,1])\chi_{J}\in C^{1}([0,1]) by

χJ(x)=1θχjfαiN1(x) if (i,j)J and xZij, and χJ(x)=1 otherwise. \chi_{J}(x)=1-\theta\cdot\chi_{j}\circ f_{\alpha_{i}^{N}}^{-1}(x)\text{ if }(i,j)\in J\text{ and }x\in Z_{i}^{j},\,\text{ and }\chi_{J}(x)=1\text{ otherwise. }

Note that χJ\chi_{J} is well defined by the separation property in Theorem 2.4 Part (4), and since the VjV_{j}’s intersect potentially only at their endpoints by Proposition 2.13, where all the χj\chi_{j} vanish.

We can now define the Dolgopyat operators NsJN_{s}^{J} on C1([0,1])C^{1}([0,1]) by:

NsJ(f)(x):=Pa,ω,N(χJf).N_{s}^{J}\left(f\right)(x):=P_{a,\omega,N}\left(\chi_{J}\cdot f\right).

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 A>1A>1, NN\in\mathbb{N} and 0<θ<10<\theta<1 such that for s=a+ibs=a+ib with |a||a| sufficiently small and |b||b| sufficiently large, for every ω\omega,

  1. 1.

    The cone 𝒞A|b|\mathcal{C}_{A|b|} is stable under every NsJN_{s}^{J}.

  2. 2.

    If fC1([0,1])f\in C^{1}([0,1]) and H𝒞A|b|H\in\mathcal{C}_{A|b|} satisfy

    |f|H and |f|A|b|H|f|\leq H\text{ and }\left|f^{\prime}\right|\leq A|b|H

    then

    |(Ps,ω,N(f))(x)|A|b|NsJ(H)(x).\left|\left(P_{s,\omega,N}\left(f\right)\right)^{\prime}(x)\right|\leq A|b|N_{s}^{J}(H)(x).
  3. 3.

    If H𝒞A|b|H\in\mathcal{C}_{A|b|} then Pa,ω,N(H2)𝒞34A|b|P_{a,\omega,N}\left(H^{2}\right)\in\mathcal{C}_{\frac{3}{4}A|b|}.

Our proof is roughly based on [39, proof of equation (4)]:

Proof.

Fix H𝒞A|b|H\in\mathcal{C}_{A|b|} where AA is yet to be determined. Then for all x[0,1]x\in[0,1],

|NsJ(H)(x)|=|Pa,ω,N(χJH)(x)|\left|N_{s}^{J}\left(H\right)^{\prime}(x)\right|=\left|P_{a,\omega,N}\left(\chi_{J}\cdot H\right)^{\prime}(x)\right|
IXN(ω)ea2πc(I,x)|a2π(logfI)(x)|η(ω)(I)(HχJ)fI(x)+ea2πc(I,x)η(ω)(I)|((HχJ)fI)(x)|.\leq\sum_{I\in X_{N}^{(\omega)}}e^{a\cdot 2\pi\cdot c(I,x)}\left|a2\pi\cdot\left(\log f_{I}^{\prime}\right)^{\prime}(x)\right|\eta^{(\omega)}(I)\left(H\cdot\chi_{J}\right)\circ f_{I}(x)+e^{a\cdot 2\pi\cdot c(I,x)}\eta^{(\omega)}(I)\left|\left(\left(H\cdot\chi_{J}\right)\circ f_{I}\right)^{\prime}(x)\right|.

Now, for every IXN(ω)I\in X_{N}^{(\omega)}, by separation (Theorem 2.4 Part (4)) we have

|(χJfI)|θA3|b|ϵ.\left|\left(\chi_{J}\circ f_{I}\right)^{\prime}\right|\leq\theta A_{3}\frac{|b|}{\epsilon^{\prime}}.

Also, we can find a constant C~\tilde{C} uniform in NN, and aa such that if |a||a| is small enough then

|a2π(logfI)(x)|C~.\left|a2\pi\cdot\left(\log f_{I}^{\prime}\right)^{\prime}(x)\right|\leq\tilde{C}.

Therefore,

|NsJ(H)(x)|IXN(ω)ea2πc(I,x)η(ω)(I))(HχJ)fI(x)C~+ea2πc(I,x)η(ω)(I)(HχJ)fI(x)A|b|ρN\left|N_{s}^{J}\left(H\right)^{\prime}(x)\right|\leq\sum_{I\in X_{N}^{(\omega)}}e^{a\cdot 2\pi\cdot c(I,x)}\eta^{(\omega)}(I))\left(H\cdot\chi_{J}\right)\circ f_{I}(x)\cdot\tilde{C}+e^{a\cdot 2\pi\cdot c(I,x)}\eta^{(\omega)}(I)\left(H\cdot\chi_{J}\right)\circ f_{I}(x)\cdot A\cdot|b|\rho^{N}
+ea2πc(α1N,x)η(ω)(I)HfI(x)θA3|b|ϵ.+e^{a\cdot 2\pi\cdot c(\alpha_{1}^{N},x)}\eta^{(\omega)}(I)H\circ f_{I}(x)\cdot\theta A_{3}\frac{|b|}{\epsilon^{\prime}}.

Using that H=(χJH)χJ11θχJHH=\frac{\left(\chi_{J}H\right)}{\chi_{J}}\leq\frac{1}{1-\theta}\chi_{J}H we obtain

|NsJ(H)(x)|(C~|b|+A3θ(1θ)ϵ+AρN)|b|NsJ(H)(x)A|b|NsJ(H)(x),\left|N_{s}^{J}\left(H\right)^{\prime}(x)\right|\leq\left(\frac{\tilde{C}}{|b|}+A_{3}\frac{\theta}{(1-\theta)\epsilon^{\prime}}+A\rho^{N}\right)|b|N_{s}^{J}\left(H\right)(x)\leq A|b|N_{s}^{J}\left(H\right)(x),

assuming |b||b| is large enough, |a||a| is small enough, and

θmin(12,ϵA14A3) and ρNA12A.\theta\leq\min\left(\frac{1}{2},\,\epsilon^{\prime}\frac{A-1}{4A_{3}}\right)\text{ and }\rho^{N}\leq\frac{A-1}{2A}.

Note that the above calculation works for any A>1A>1. Now, if fC1([0,1])f\in C^{1}([0,1]) and H𝒞A|b|H\in\mathcal{C}_{A|b|} satisfy

|f|H and |f|A|b|H|f|\leq H\text{ and }\left|f^{\prime}\right|\leq A|b|H

then

|Ps,ω,N(f)(x)|IXN(ω)ea2πc(I,x)|(a+ib)2π(logfI)(x)|η(ω)(I)HfI(x)\left|P_{s,\omega,N}\left(f\right)^{\prime}(x)\right|\leq\sum_{I\in X_{N}^{(\omega)}}e^{a\cdot 2\pi\cdot c(I,x)}\left|(a+ib)2\pi\cdot\left(\log f_{I}^{\prime}\right)^{\prime}(x)\right|\eta^{(\omega)}(I)H\circ f_{I}(x)
+ea2πc(I,x)η(ω)(I)|(ffI)(x)|+e^{a\cdot 2\pi\cdot c(I,x)}\eta^{(\omega)}(I)\left|\left(f\circ f_{I}\right)^{\prime}(x)\right|
I𝒜Nea2πc(I,x)η(I)HfI(x)C^|b|+ea2πc(I,x)η(I)HfI(x)A|b|ρN,\leq\sum_{I\in\mathcal{A}^{N}}e^{a\cdot 2\pi\cdot c(I,x)}\eta(I)H\circ f_{I}(x)\cdot\hat{C}\cdot|b|+e^{a\cdot 2\pi\cdot c(I,x)}\eta(I)H\circ f_{I}(x)\cdot A|b|\rho^{N},

where C^\hat{C} is independent of NN and |b||b| is large enough. Note that χJ12\chi_{J}^{-1}\leq 2 when θ12\theta\leq\frac{1}{2} and so in this case H2χJHH\leq 2\chi_{J}H. So, under this assumption

|Ps,ω,N(f)(x)|A|b|NsJ(H)(x)\left|P_{s,\omega,N}\left(f\right)^{\prime}(x)\right|\leq A|b|N_{s}^{J}\left(H\right)(x)

as long as A4C^A\geq 4\hat{C} and ρN14\rho^{N}\leq\frac{1}{4}.

We thus fix Amax(2,4C^)A\geq\max\left(2,4\hat{C}\right) and take NN large enough so that ρNmin(A12A,14)\rho^{N}\leq\min\left(\frac{A-1}{2A},\frac{1}{4}\right) and fix θmin(12,ϵA14A3)\theta\leq\min\left(\frac{1}{2},\epsilon^{\prime}\frac{A-1}{4A_{3}}\right). A similar argument shows that this choice of parameters also yields Part (3).

2.7.2 Part 2: L2L^{2} contraction of the cones

We proceed to prove that the operators NsJN_{s}^{J} contract these cones in the L2L^{2} norms. We require the following Definition:

Definition 2.18.

A subset JJs,ωJ\subseteq J_{s,\omega} is called dense if for every 1jq1\leq j\leq q such that VjKσNωV_{j}\cap K_{\sigma^{N}\omega}\neq\emptyset there exists 1jq1\leq j^{\prime}\leq q such that:

i such that (i,j)J and |jj|2.\exists i\text{ such that }(i,j^{\prime})\in J\text{ and }|j^{\prime}-j|\leq 2.

For a dense subset JJ we write

WJ={xKσNω:(i,j)J,xVj}.W_{J}=\{x\in K_{\sigma^{N}\omega}:\,\exists(i,j)\in J,\,x\in V_{j}\}.

We need the following key Lemma, a variant of [39, Lemma 5.7]:

Lemma 2.19.

Let JJ be dense and fix HCA|b|H\in C_{A|b|}. Then there exists ϵ~>0\tilde{\epsilon}>0 independent of H,|b|,ω,NH,|b|,\omega,N and JJ such that

WJH𝑑μσNωϵ~KσNωH𝑑μσNω.\int_{W_{J}}H\,d\mu_{\sigma^{N}\omega}\geq\tilde{\epsilon}\int_{K_{\sigma^{N}\omega}}Hd\mu_{\sigma^{N}\omega}.
Proof.

Let

G:={1iq:ViKσNω}G:=\{1\leq i\leq q:V_{i}\cap K_{\sigma^{N}\omega}\neq\emptyset\}

and note that KσNωiGViK_{\sigma^{N}\omega}\subseteq\bigcup_{i\in G}V_{i}. For every iGi\in G, by density of JJ, there exists some index j(i)j(i) such that (i,j(i))J\left(i^{\prime},\,j(i)\right)\in J for some i{1,2}i^{\prime}\in\{1,2\} such that |j(i)i|2\left|j(i)-i\right|\leq 2. We thus get a function j:GGj:G\rightarrow G such that for every iGi\in G the set j1({i})j^{-1}\left(\{i\}\right) contains at most 55 elements.

Let r=3A1ϵ|b|r=3A_{1}\frac{\epsilon^{\prime}}{|b|}. For every iGi\in G fix some uiViKσNωu_{i}\in V_{i}\cap K_{\sigma^{N}\omega}. By Proposition 2.13,

Vj(i),ViB(ui,r).V_{j(i)},\,V_{i}\subseteq B(u_{i},\,r).

Moreover, by Proposition 2.13 part (4), for r=12A2A1ϵ|b|r^{\prime}=\frac{1}{2}A_{2}A_{1}^{\prime}\frac{\epsilon^{\prime}}{|b|} and some carefully chosen viKσNωVj(i)v_{i}\in K_{\sigma^{N}\omega}\cap V_{j(i)} we have

dist(KσNωVj(i),Vj(i))=dist(vi,Vj(i)) and so Vj(i)B(vi,r).\text{dist}\left(K_{\sigma^{N}\omega}\cap V_{j(i)},\,\partial V_{j(i)}\right)=\text{dist}\left(v_{i},\,\partial V_{j(i)}\right)\text{ and so }V_{j(i)}\supset B(v_{i},\,r^{\prime}).

Note that

B(vi,r)B(ui,r)B(vi,2r).B(v_{i},r^{\prime})\subset B(u_{i},r)\subset B(v_{i},2r).

So, by the Federer property of μσNω\mu_{\sigma^{N}\omega} proved in Theorem 2.4,

μσNω(B(ui,r))μσNω(B(vi,2r))C2rrμσNω(B(vi,r))C2rrμσNω(Vj(i)),\mu_{\sigma^{N}\omega}\left(B(u_{i},r)\right)\leq\mu_{\sigma^{N}\omega}\left(B(v_{i},2r)\right)\leq C_{\frac{2r}{r^{\prime}}}\mu_{\sigma^{N}\omega}\left(B(v_{i},r^{\prime})\right)\leq C_{\frac{2r}{r^{\prime}}}\mu_{\sigma^{N}\omega}\left(V_{j(i)}\right),

where we note that C2rrC_{\frac{2r}{r^{\prime}}} does not depend on σNω{\sigma^{N}\omega}.

Now, let HCA|b|H\in C_{A|b|}. Then, as long as |b||b| is large enough,

KσNωH𝑑μσNω\displaystyle\int_{K_{\sigma^{N}\omega}}Hd\mu_{\sigma^{N}\omega} =\displaystyle= iGViH𝑑μσNω\displaystyle\sum_{i\in G}\int_{V_{i}}Hd\mu_{\sigma^{N}\omega}
\displaystyle\leq iGB(ui,r)H𝑑μσNω\displaystyle\sum_{i\in G}\int_{B(u_{i},r)}Hd\mu_{\sigma^{N}\omega}
\displaystyle\leq iG(maxB(ui,r)H)μσNω(B(ui,r))\displaystyle\sum_{i\in G}\left(\max_{B(u_{i},r)}H\right)\mu_{\sigma^{N}\omega}\left(B(u_{i},r)\right)
\displaystyle\leq CiGe2A|b|r(minVj(i)H)μσNω(Vj(i))\displaystyle C^{\prime}\sum_{i\in G}e^{2A|b|r}\left(\min_{V_{j(i)}}H\right)\mu_{\sigma^{N}\omega}\left(V_{j(i)}\right)
\displaystyle\leq CeC′′iGVj(i)H𝑑μσNω\displaystyle C^{\prime}e^{C^{\prime\prime}}\sum_{i\in G}\int_{V_{j(i)}}Hd\mu_{\sigma^{N}\omega}
\displaystyle\leq CeC′′j:i,(i,j)JVj(i)H𝑑μσNω\displaystyle C^{\prime}e^{C^{\prime\prime}}\sum_{j:\exists i,(i,j)\in J}\int_{V_{j(i)}}Hd\mu_{\sigma^{N}\omega}
\displaystyle\leq 5CeC′′WJH𝑑μσNω.\displaystyle 5C^{\prime}e^{C^{\prime\prime}}\int_{W_{J}}Hd\mu_{\sigma^{N}\omega}.

Since the constant C,C′′C^{\prime},C^{\prime\prime} are uniform (in particular, in |b||b| and ω\omega), the proof is complete. ∎

Definition 2.20.

We define

s,ω:={JJs,ω:J is dense }.\mathcal{E}{s,\omega}:=\{J\subseteq J_{s,\omega}:\,J\text{ is dense }\}.

We can now prove the required contraction property in Lemma 2.12 Part (2).

Proposition 2.21.

There exists 0<α<10<\alpha<1 uniform in ω\omega such that for all s=a+ibs=a+ib with |a||a| small and |b||b| large, for all HCA|b|H\in C_{A|b|} and all Js,ωJ\in\mathcal{E}{s,\omega} we have

KσNω|NsJ(H)|2𝑑μσNωαKσNωP0,ω,N(H2)𝑑μσNω.\int_{K_{\sigma^{N}\omega}}\left|N_{s}^{J}\left(H\right)\right|^{2}d\mu_{\sigma^{N}\omega}\leq\alpha\int_{K_{\sigma^{N}\omega}}P_{0,\omega,N}\left(H^{2}\right)\,d\mu_{\sigma^{N}\omega}.

This is a randomized analogue of [39, Proposition 5.9].

Proof.

Let HCA|b|H\in C_{A|b|}. For every x[0,1]x\in[0,1], by the Cauchy-Schwartz inequality and the definition of NsJN_{s}^{J},

(NsJ(H))2(x)\left(N_{s}^{J}\left(H\right)\right)^{2}(x)

is bounded above by the product of

IXN(ω)ea2πc(I,x)η(ω)(I)χJ2fI(x)\sum_{I\in X_{N}^{(\omega)}}e^{a\cdot 2\pi\cdot c(I,x)}\eta^{(\omega)}(I)\chi_{J}^{2}\circ f_{I}(x) (25)

and

Pa,ω,N(H2)(x).P_{a,\omega,N}\left(H^{2}\right)(x).

For all xWJx\in W_{J} there exists i{1,2}i\in\{1,2\} such that

χJfαiN(x)=1θ.\chi_{J}\circ f_{\alpha_{i}^{N}}(x)=1-\theta.

Recall that for every IXN(ω)I\in X_{N}^{(\omega)} we have by (13)

|ac(I,x)||a|DN.\left|a\cdot c(I,x)\right|\leq|a|\cdot D^{\prime}\cdot N.

Therefore, recalling the notations from Section 2.3, if xWJx\in W_{J} then the sum in (25) is bounded by

e|a|DNθeN(miniI,1jkilog𝐩j(i)+|a|D).e^{|a|\cdot D^{\prime}\cdot N}-\theta\cdot e^{N\left(\min_{i\in I,1\leq j\leq k_{i}}\log\mathbf{p}_{j}^{(i)}+|a|\cdot D^{\prime}\right)}.

Since

KσNω(NsJ(H))2𝑑μσNω=WJ(NsJ(H))2𝑑μσNω+KσNωWJ(NsJ(H))2𝑑μσNω\int_{K_{\sigma^{N}\omega}}\left(N_{s}^{J}\left(H\right)\right)^{2}d\mu_{\sigma^{N}\omega}=\int_{W_{J}}\left(N_{s}^{J}\left(H\right)\right)^{2}d\mu_{\sigma^{N}\omega}+\int_{K_{\sigma^{N}\omega}\setminus W_{J}}\left(N_{s}^{J}\left(H\right)\right)^{2}d\mu_{\sigma^{N}\omega}

the previous discussion and the fact that e|a|DNe^{|a|\cdot D^{\prime}\cdot N} always bounds (25) show that

KσNω(NsJ(H))2𝑑μσNω(e|a|DNθeN(miniIlog𝐩j(i)+|a|D))WJPa,ω,N(H2)𝑑μσNω\int_{K_{\sigma^{N}\omega}}\left(N_{s}^{J}\left(H\right)\right)^{2}d\mu_{\sigma^{N}\omega}\leq\left(e^{|a|\cdot D^{\prime}\cdot N}-\theta\cdot e^{N\left(\min_{i\in I}\log\mathbf{p}_{j}^{(i)}+|a|\cdot D^{\prime}\right)}\right)\int_{W_{J}}P_{a,\omega,N}\left(H^{2}\right)d\mu_{\sigma^{N}\omega}
+e|a|DNKσNωWJPa,ω,N(H2)𝑑μσNω+e^{|a|\cdot D^{\prime}\cdot N}\cdot\int_{K_{\sigma^{N}\omega}\setminus W_{J}}P_{a,\omega,N}\left(H^{2}\right)d\mu_{\sigma^{N}\omega}
=e|a|DNKσNωPa,ω,N(H2)𝑑μσNωθeN(miniIlog𝐩j(i)+|a|D)WJPa,ω,N(H2)𝑑μσNω.=e^{|a|\cdot D^{\prime}\cdot N}\cdot\int_{K_{\sigma^{N}\omega}}P_{a,\omega,N}\left(H^{2}\right)d\mu_{\sigma^{N}\omega}-\theta\cdot e^{N\left(\min_{i\in I}\log\mathbf{p}_{j}^{(i)}+|a|\cdot D^{\prime}\right)}\int_{W_{J}}P_{a,\omega,N}\left(H^{2}\right)d\mu_{\sigma^{N}\omega}.

Now, by Lemma 2.17 Pa,ω,N(H2)C34A|b|P_{a,\omega,N}\left(H^{2}\right)\in C_{\frac{3}{4}A|b|} since HCa|b|H\in C_{a|b|}. Therefore, applying Lemma 2.19 to this function we obtain

KσNω(NsJ(H))2𝑑μσNω(e|a|DNϵ~θeN(miniIlog𝐩j(i)+|a|D))KσNωPa,ω,N(H2)𝑑μσNω.\int_{K\sigma^{N}\omega}\left(N_{s}^{J}\left(H\right)\right)^{2}d\mu_{\sigma^{N}\omega}\leq\left(e^{|a|\cdot D^{\prime}\cdot N}-\tilde{\epsilon}\cdot\theta\cdot e^{N\left(\min_{i\in I}\log\mathbf{p}_{j}^{(i)}+|a|\cdot D^{\prime}\right)}\right)\int_{K_{\sigma^{N}\omega}}P_{a,\omega,N}\left(H^{2}\right)d\mu_{\sigma^{N}\omega}.

Recall that

Pa,ω,N(H2)eND|a|P0,ω,N(H2).P_{a,\omega,N}\left(H^{2}\right)\leq e^{N\cdot D^{\prime}\cdot|a|}P_{0,\omega,N}\left(H^{2}\right).

So, if |a||a| is small enough there is some 0α<10\leq\alpha<1 such that

(e|a|DNϵ~θeN(miniIlog𝐩j(i)+|a|D))eND|a|α<1.\left(e^{|a|\cdot D^{\prime}\cdot N}-\tilde{\epsilon}\cdot\theta\cdot e^{N\left(\min_{i\in I}\log\mathbf{p}_{j}^{(i)}+|a|\cdot D^{\prime}\right)}\right)\cdot e^{N\cdot D^{\prime}\cdot|a|}\leq\alpha<1.

Therefore

KσNω(NsJ(H))2𝑑μσNωαKσNωP0,ω,N(H2)𝑑μσNω,\int_{K_{\sigma^{N}\omega}}\left(N_{s}^{J}\left(H\right)\right)^{2}d\mu_{\sigma^{N}\omega}\leq\alpha\cdot\int_{K_{\sigma^{N}\omega}}P_{0,\omega,N}\left(H^{2}\right)d\mu_{\sigma^{N}\omega},

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 HCA|b|,fC1([0,1])H\in C_{A|b|},f\in C^{1}\left([0,1]\right) be such that

|f|H, and |f|A|b|H.\left|f\right|\leq H,\text{ and }\left|f^{\prime}\right|\leq A|b|H.

For every j=1,2j=1,2 define functions Θj:[0,1]+\Theta_{j}:[0,1]\rightarrow\mathbb{R}_{+} via

Θ1(x):=|e(a+ib)2πc(α1N,x)η(ω)(α1N)ffα1N(x)+e(a+ib)2πc(α2N,x)η(ω)(α2N)ffα2N(x)|(12θ)ea2πc(α1N,x)η(ω)(α1N)Hfα1N(x)+ea2πc(α2N,x)η(ω)(α2N)Hfα2N(x)\Theta_{1}(x):=\frac{\left|e^{(a+ib)\cdot 2\pi\cdot c(\alpha_{1}^{N},x)}\eta^{(\omega)}(\alpha_{1}^{N})f\circ f_{\alpha_{1}^{N}}(x)+e^{(a+ib)\cdot 2\pi\cdot c(\alpha_{2}^{N},x)}\eta^{(\omega)}(\alpha_{2}^{N})f\circ f_{\alpha_{2}^{N}}(x)\right|}{(1-2\theta)e^{a\cdot 2\pi\cdot c(\alpha_{1}^{N},x)}\eta^{(\omega)}(\alpha_{1}^{N})H\circ f_{\alpha_{1}^{N}}(x)+e^{a\cdot 2\pi\cdot c(\alpha_{2}^{N},x)}\eta^{(\omega)}(\alpha_{2}^{N})H\circ f_{\alpha_{2}^{N}}(x)}
Θ2(x):=|e(a+ib)2πc(α1N,x)η(ω)(α1N)ffα1N(x)+e(a+ib)2πc(α2N,x)η(ω)(α2N)ffα2N(x)|ea2πc(α1N,x)η(ω)(α1N)Hfα1N(x)+(12θ)ea2πc(α2N,x)η(ω)(α2N)Hfα2N(x).\Theta_{2}(x):=\frac{\left|e^{(a+ib)\cdot 2\pi\cdot c(\alpha_{1}^{N},x)}\eta^{(\omega)}(\alpha_{1}^{N})f\circ f_{\alpha_{1}^{N}}(x)+e^{(a+ib)\cdot 2\pi\cdot c(\alpha_{2}^{N},x)}\eta^{(\omega)}(\alpha_{2}^{N})f\circ f_{\alpha_{2}^{N}}(x)\right|}{e^{a\cdot 2\pi\cdot c(\alpha_{1}^{N},x)}\eta^{(\omega)}(\alpha_{1}^{N})H\circ f_{\alpha_{1}^{N}}(x)+(1-2\theta)e^{a\cdot 2\pi\cdot c(\alpha_{2}^{N},x)}\eta^{(\omega)}(\alpha_{2}^{N})H\circ f_{\alpha_{2}^{N}}(x)}.

Then for θ\theta and ϵ\epsilon^{\prime} small enough, and for all |a||a| small enough and |b||b| large enough, for all jj such that VjKσNωV_{j}\cap K_{\sigma^{N}\omega}\neq\emptyset there exist jj^{\prime} such that |jj|2|j^{\prime}-j|\leq 2 and VjKσNωV_{j^{\prime}}\cap K_{\sigma^{N}\omega}\neq\emptyset and some i{1,2}i\in\{1,2\} such that: For every xVjx\in V_{j^{\prime}},

Θi(x)1.\Theta_{i}(x)\leq 1.

This is a model version of [39, Lemma 5.10]. For the proof, we require the following basic Lemmas from [39]:

Lemma 2.23.

[39, Lemma 5.11] Let ZIZ\subseteq I be an interval with |Z|c|b||Z|\leq\frac{c}{|b|}. Let H,fH,f be as in Lemma 2.22. Then for all cc small enough, either |f(u)|34H(u)\left|f(u)\right|\leq\frac{3}{4}H(u) for all uZu\in Z, or |f(u)|14H(u)\left|f(u)\right|\geq\frac{1}{4}H(u) for all uZu\in Z.

In the following arguments, for 0z0\neq z\in\mathbb{C} let arg(z)(π,π]\arg(z)\in(-\pi,\,\pi] be the unique real number such that |z|eiarg(z)=z|z|e^{i\arg(z)}=z.

Lemma 2.24.

[39, Lemma 5.12] Let 0z1,z20\neq z_{1},z_{2}\in\mathbb{C} be such that

|z1||z2|M and 0<ϵ|arg(z1)arg(z2)|2πϵ.\frac{|z_{1}|}{|z_{2}|}\leq M\text{ and }0<\epsilon\leq\left|\arg(z_{1})-\arg(z_{2})\right|\leq 2\pi-\epsilon.

Then there exists some 0<δ(M,ϵ)<10<\delta(M,\epsilon)<1 such that

|z1+z2|(1δ)|z1|+|z2|.\left|z_{1}+z_{2}\right|\leq(1-\delta)|z_{1}|+|z_{2}|.

Proof of Lemma 2.22 First, we choose ϵ\epsilon^{\prime} small enough so that the conclusion of Lemma 2.23 is true for all Z=Zji=fαiN(Vj)Z=Z_{j}^{i}=f_{\alpha_{i}^{N}}\left(V_{j}\right), and one checks that this does not change AA and NN (see the end of the proof of Lemma 2.17, and recall that fαiNf_{\alpha_{i}^{N}} are contractions). It is clear that |Zij||Vj||Z_{i}^{j}|\leq|V_{j}|. We add the assumption that

0<θ18 so that 12θ34.0<\theta\leq\frac{1}{8}\text{ so that }1-2\theta\geq\frac{3}{4}.

Let Vj,Vj+1,Vj+2V_{j},V_{j+1},V_{j+2} be a triad of intervals such that each of them intersects KσNωK_{\sigma^{N}\omega}. Write

Vj^=VjVj+1Vj+2.\widehat{V_{j}}=V_{j}\cup V_{j+1}\cup V_{j+2}.

If for some (i,j){1,2}×{j,j+1,j+2}(i,j^{\prime})\in\{1,2\}\times\{j,j+1,j+2\} we have |f(u)|34H(u)\left|f(u)\right|\leq\frac{3}{4}H(u) for all uZjiu\in Z_{j^{\prime}}^{i} then Θi(u)1\Theta_{i}(u)\leq 1 for all uZjiu\in Z_{j^{\prime}}^{i}, and we are done.

Otherwise, by Lemma 2.23 we have for all (i,j){1,2}×{j,j+1,j+2}(i,j^{\prime})\in\{1,2\}\times\{j,j+1,j+2\} and every uZjiu\in Z_{j^{\prime}}^{i},

|f(u)|14H(u).\left|f(u)\right|\geq\frac{1}{4}H(u).

We aim to make use of Lemma 2.24. For every xVj^x\in\widehat{V_{j}} set

z1(x):=e(a+ib)2πc(α1N,x)η(ω)(α1N)ffα1N(x),z2(x):=e(a+ib)2πc(α2N,x)η(ω)(α2N)ffα2N(x).z_{1}(x):=e^{(a+ib)\cdot 2\pi\cdot c(\alpha_{1}^{N},x)}\eta^{(\omega)}(\alpha_{1}^{N})f\circ f_{\alpha_{1}^{N}}(x),\quad z_{2}(x):=e^{(a+ib)\cdot 2\pi\cdot c(\alpha_{2}^{N},x)}\eta^{(\omega)}(\alpha_{2}^{N})f\circ f_{\alpha_{2}^{N}}(x).

Let

M=4e2N(miniI,1jkilog𝐩j(i)+|a|D)e2AϵA1.M=4e^{2N\left(-\min_{i\in I,1\leq j\leq k_{i}}\log\mathbf{p}_{j}^{(i)}+|a|\cdot D^{\prime}\right)}e^{2A\epsilon^{\prime}A_{1}}.

We claim that for j{j,j+1,j+2}j^{\prime}\in\{j,j+1,j+2\},

|z1(x)z2(x)|M for all xVj, or |z2(x)z1(x)|M for all xVj.\left|\frac{z_{1}(x)}{z_{2}(x)}\right|\leq M\text{ for all }x\in V_{j^{\prime}},\,\text{ or }\left|\frac{z_{2}(x)}{z_{1}(x)}\right|\leq M\text{ for all }x\in V_{j^{\prime}}.

Indeed,

14e2N(miniI,1jkilog𝐩j(i)+|a|D)Hα1N(x)Hα2N(x)|z1(x)z2(x)|4e2N(miniI,1jkilog𝐩j(i)+|a|D)Hα1N(x)Hα2N(x).\frac{1}{4}e^{-2N\left(-\min_{i\in I,1\leq j\leq k_{i}}\log\mathbf{p}_{j}^{(i)}+|a|\cdot D^{\prime}\right)}\frac{H\circ\alpha_{1}^{N}(x)}{H\circ\alpha_{2}^{N}(x)}\leq\left|\frac{z_{1}(x)}{z_{2}(x)}\right|\leq 4e^{2N\left(-\min_{i\in I,1\leq j\leq k_{i}}\log\mathbf{p}_{j}^{(i)}+|a|\cdot D^{\prime}\right)}\frac{H\circ\alpha_{1}^{N}(x)}{H\circ\alpha_{2}^{N}(x)}.

If for some x0Vjx_{0}\in V_{j^{\prime}} we have Hα1N(x0)Hα2N(x0)1\frac{H\circ\alpha_{1}^{N}(x_{0})}{H\circ\alpha_{2}^{N}(x_{0})}\leq 1 then for all xVjx\in V_{j^{\prime}} we obtain

Hα1N(x)Hα2N(x)eAA1ϵHα1N(x0)eAA1ϵHα2N(x0)e2AA1ϵ,\frac{H\circ\alpha_{1}^{N}(x)}{H\circ\alpha_{2}^{N}(x)}\leq\frac{e^{AA_{1}\epsilon^{\prime}}H\circ\alpha_{1}^{N}(x_{0})}{e^{-AA_{1}\epsilon^{\prime}}H\circ\alpha_{2}^{N}(x_{0})}\leq e^{2AA_{1}\epsilon^{\prime}},

and so |z1(x)z2(x)|M\left|\frac{z_{1}(x)}{z_{2}(x)}\right|\leq M. If Hα1N(x)Hα2N(x)1\frac{H\circ\alpha_{1}^{N}(x)}{H\circ\alpha_{2}^{N}(x)}\geq 1 for all xVjx\in V_{j^{\prime}} then

|z2(x)z1(x)|4e2N(miniIlog𝐩j(i)+|a|D)M.\left|\frac{z_{2}(x)}{z_{1}(x)}\right|\leq 4e^{2N\left(\min_{i\in I}\log\mathbf{p}_{j}^{(i)}+|a|\cdot D^{\prime}\right)}\leq M.

We next control the relative variations in the arguments of z1,z2z_{1},z_{2}. Since

|zi(x)|eN(miniI,1jkilog𝐩j(i)+aD)14HfαiN(x)>0, for all xVj^ and i=1,2,\left|z_{i}(x)\right|\geq e^{-N\left(\min_{i\in I,1\leq j\leq k_{i}}\log\mathbf{p}_{j}^{(i)}+a\cdot D\right)}\frac{1}{4}H\circ f_{\alpha_{i}^{N}}(x)>0,\quad\text{ for all }x\in\widehat{V_{j}}\text{ and }i=1,2,

there exist two C1C^{1} functions Li:Vj^L_{i}:\widehat{V_{j}}\rightarrow\mathbb{C} such that

Li(x)=zi(x)zi(x) and eLi(x)=zi(x) for all xVj^.L_{i}^{\prime}(x)=\frac{z_{i}^{\prime}(x)}{z_{i}(x)}\text{ and }e^{L_{i}(x)}=z_{i}(x)\text{ for all }x\in\widehat{V_{j}}.

For one possible construction see [39, Proof of Lemma 5.10]. For all xVj^x\in\widehat{V_{j}} set

Φ(x):=(L1(x))(L2(x)).\Phi(x):=\Im\left(L_{1}(x))-\Im(L_{2}(x)\right).

Then for xVj^x\in\widehat{V_{j}}

Φ(x)=(z1(x)z1(x)z2(x)z2(x))\Phi^{\prime}(x)=\Im\left(\frac{z_{1}^{\prime}(x)}{z_{1}(x)}-\frac{z_{2}^{\prime}(x)}{z_{2}(x)}\right)
=bddx(logfα1Nlogfα2N)(x)+((ffα1N)(x)ffα1N(x)(ffα2N)(x)ffα2N(x)).=b\frac{d}{dx}\left(\log f_{\alpha_{1}^{N}}^{\prime}-\log f_{\alpha_{2}^{N}}^{\prime}\right)(x)+\Im\left(\frac{\left(f\circ f_{\alpha_{1}^{N}}\right)^{\prime}(x)}{f\circ f_{\alpha_{1}^{N}}(x)}-\frac{\left(f\circ f_{\alpha_{2}^{N}}\right)^{\prime}(x)}{f\circ f_{\alpha_{2}^{N}}(x)}\right).

Now, by our assumptions on f,Hf,H

|(ffα1N)(x)ffα1N(x)(ffα2N)(x)ffα2N(x)|8A|b|ρN,\left|\frac{\left(f\circ f_{\alpha_{1}^{N}}\right)^{\prime}(x)}{f\circ f_{\alpha_{1}^{N}}(x)}-\frac{\left(f\circ f_{\alpha_{2}^{N}}\right)^{\prime}(x)}{f\circ f_{\alpha_{2}^{N}}(x)}\right|\leq 8A|b|\rho^{N},

and so by Theorem 2.4 Part (5) for all xVj^x\in\widehat{V_{j}}

m8AρN|Φ(x)||b|m+8AρN.m-8A\rho^{N}\leq\frac{\left|\Phi^{\prime}(x)\right|}{|b|}\leq m^{\prime}+8A\rho^{N}.

Let xVj,xVj+2x\in V_{j},x^{\prime}\in V_{j+2}. By the mean value Theorem

(m8AρN)A1ϵ|Φ(x)Φ(x)|(m+8AρN)3A1ϵ.\left(m-8A\rho^{N}\right)A_{1}^{\prime}\epsilon^{\prime}\leq\left|\Phi(x)-\Phi(x^{\prime})\right|\leq\left(m^{\prime}+8A\rho^{N}\right)3A_{1}\epsilon^{\prime}.

So, if NN is large enough then there are constants B1,B2>0B_{1},B_{2}>0 such that, independently of x,xx,x^{\prime} and |b||b|,

B1ϵ|Φ(x)Φ(x)|B2ϵ.B_{1}\epsilon^{\prime}\leq\left|\Phi(x)-\Phi(x^{\prime})\right|\leq B_{2}\epsilon^{\prime}.

We now pick ϵ\epsilon^{\prime} so that

(B2+B12)ϵπ,(B_{2}+\frac{B_{1}}{2})\epsilon^{\prime}\leq\pi,

and put ϵ=B1ϵ4\epsilon=B_{1}\frac{\epsilon^{\prime}}{4}.

Suppose, towards a contradiction, that there are xVjx\in V_{j} and xVj+2x^{\prime}\in V_{j+2} such that both

Φ(x),Φ(x)k[2kπϵ,2kπ+ϵ].\Phi(x),\Phi(x^{\prime})\in\bigcup_{k\in\mathbb{Z}}[2k\pi-\epsilon,2k\pi+\epsilon].

Since |Φ(x)Φ(x)|B2ϵ\left|\Phi(x)-\Phi(x^{\prime})\right|\leq B_{2}\epsilon^{\prime} we cannot have

Φ(x)[2k1πϵ,2k1π+ϵ] and Φ(x)[2k2πϵ,2k2π+ϵ]\Phi(x)\in[2k_{1}\pi-\epsilon,2k_{1}\pi+\epsilon]\text{ and }\Phi(x^{\prime})\in[2k_{2}\pi-\epsilon,2k_{2}\pi+\epsilon]

with k1k2k_{1}\neq k_{2}. Indeed, it will imply

2πB1ϵ22π2ϵ|Φ(x)Φ(x)|B2ϵ,2\pi-B_{1}\frac{\epsilon^{\prime}}{2}\leq 2\pi-2\epsilon\leq\left|\Phi(x)-\Phi(x^{\prime})\right|\leq B_{2}\epsilon^{\prime},

a contradiction. But then

B1ϵ|Φ(x)Φ(x)|2ϵ=B1ϵ/2,B_{1}\epsilon^{\prime}\leq\left|\Phi(x)-\Phi(x^{\prime})\right|\leq 2\epsilon=B_{1}\epsilon^{\prime}/2,

which is also a contradiction.

We conclude that there exists j{j,j+2}j^{\prime}\in\{j,j+2\} such that for all xVjx\in V_{j^{\prime}},

dist(Φ(x), 2π)>ϵ.\text{dist}\left(\Phi(x),\,2\pi\mathbb{Z}\right)>\epsilon.

Since eiΦ(x)=ei(arg(z1)arg(z2))e^{i\Phi(x)}=e^{i\left(\arg(z_{1})-\arg(z_{2})\right)}, the conditions of Lemma 2.24 are met. Thus, either for every xVjx\in V_{j^{\prime}}

|z1(x)+z2(x)|(1δ(M,ϵ))|z1(x)|+|z2(x)|,\left|z_{1}(x)+z_{2}(x)\right|\leq\left(1-\delta(M,\epsilon)\right)\left|z_{1}(x)\right|+\left|z_{2}(x)\right|,

or for every xVjx\in V_{j^{\prime}}

|z1(x)+z2(x)|(1δ(M,ϵ))|z2(x)|+|z1(x)|,\left|z_{1}(x)+z_{2}(x)\right|\leq\left(1-\delta(M,\epsilon)\right)\left|z_{2}(x)\right|+\left|z_{1}(x)\right|,

depending on whether

|z1(x)z2(x)|M for all xVj, or |z2(x)z1(x)|M for all xVj.\left|\frac{z_{1}(x)}{z_{2}(x)}\right|\leq M\text{ for all }x\in V_{j^{\prime}},\,\text{ or }\left|\frac{z_{2}(x)}{z_{1}(x)}\right|\leq M\text{ for all }x\in V_{j^{\prime}}.

Choosing 0<θ<12δ(M,ϵ)0<\theta<\frac{1}{2}\delta(M,\epsilon), we have Θi(x)1\Theta_{i}(x)\leq 1 for all xVjx\in V_{j^{\prime}} and some i{1,2}i\in\{1,2\}. \hfill{\Box}

Proof of Lemma 2.12 Part (3) Fix constants N,A,ϵ,θN,A,\epsilon^{\prime},\theta so that parts (1) and (2) of Lemma 2.12, and so that Lemma 2.22, all hold true. Let fC1([0,1],HCA|b|f\in C^{1}([0,1],H\in C_{A|b|} be such that

|f|H, and |f|A|b|H.\left|f\right|\leq H,\text{ and }\left|f^{\prime}\right|\leq A|b|H.

We aim to show that there exists a dense subset Js,ωJ\in\mathcal{E}_{s,\omega} such that

|Ps,ω,Nf|NsJH and |(Ps,ω,Nf)|A|b|NsJH.|P_{s,\omega,N}f|\leq N_{s}^{J}H\quad\text{ and }|(P_{s,\omega,N}f)^{\prime}|\leq A|b|N_{s}^{J}H.

Since the latter statement holds true for all JJ\neq\emptyset by Lemma 2.17, we focus on the first one.

Let

J:={(i,j):Θi(x)1 for all xVj}.J:=\{(i,j):\,\Theta_{i}(x)\leq 1\text{ for all }x\in V_{j}\}.

By Lemma 2.22 JJ is dense. Recall that by Theorem 2.4 Part (4), for all IJXN(ω)I\neq J\in X_{N}^{(\omega)},

fI([0,1])fJ([0,1])=.f_{I}([0,1])\cap f_{J}([0,1])=\emptyset. (26)

Let x[0,1]x\in[0,1]. If x Int Vjx\notin\text{ Int }V_{j} for all (i,j)J(i,j)\in J, then by the definition of χJ\chi_{J} and by (26),

χJfI(x)=1 for all IXN(ω).\chi_{J}\circ f_{I}(x)=1\text{ for all }I\in X_{N}^{(\omega)}.

Therefore,

|Ps,ω,N(f)(x)|\displaystyle\left|P_{s,\omega,N}\left(f\right)(x)\right| \displaystyle\leq IXN(ω)ea2πc(I,x)η(ω)(I)HfI(x)\displaystyle\sum_{I\in X_{N}^{(\omega)}}e^{a\cdot 2\pi\cdot c(I,x)}\eta^{(\omega)}(I)H\circ f_{I}(x)
=\displaystyle= IXN(ω)ea2πc(I,x)η(ω)(I)(HχJ)fI(x)\displaystyle\sum_{I\in X_{N}^{(\omega)}}e^{a\cdot 2\pi\cdot c(I,x)}\eta^{(\omega)}(I)(H\cdot\chi_{J})\circ f_{I}(x)
=\displaystyle= NsJ(H)(x).\displaystyle N_{s}^{J}\left(H\right)(x).

Now, suppose x Int Vjx\in\text{ Int }V_{j} for some (i,j)J(i,j)\in J.

Case 1 If (1,j)J(1,j)\in J but (2,j)J(2,j)\notin J then by (26) χJfI(x)=1\chi_{J}\circ f_{I}(x)=1 for all Iα1NI\neq\alpha_{1}^{N}: Indeed, (26) implies that fI(x)fα1N(Vj)f_{I}(x)\notin f_{\alpha_{1}^{N}}(V_{j}) since fI(x)fα1N([0,1])f_{I}(x)\notin f_{\alpha_{1}^{N}}([0,1]). Since Θ1(x)1\Theta_{1}(x)\leq 1 we find that

|Ps,ω,N(f)(x)|\displaystyle\left|P_{s,\omega,N}\left(f\right)(x)\right| \displaystyle\leq IαiN,IXN(ω)ea2πc(I,x)η(ω)(I)HfI(x)+\displaystyle\sum_{I\neq\alpha_{i}^{N},I\in X_{N}^{(\omega)}}e^{a\cdot 2\pi\cdot c(I,x)}\eta^{(\omega)}(I)H\circ f_{I}(x)+
(12θ)ea2πc(α1N,x)η(ω)(α1N)Hfα1N(x)+ea2πc(α2N,x)η(ω)(α2N)Hfα2N(x)\displaystyle(1-2\theta)e^{a\cdot 2\pi\cdot c(\alpha_{1}^{N},x)}\eta^{(\omega)}(\alpha_{1}^{N})H\circ f_{\alpha_{1}^{N}}(x)+e^{a\cdot 2\pi\cdot c(\alpha_{2}^{N},x)}\eta^{(\omega)}(\alpha_{2}^{N})H\circ f_{\alpha_{2}^{N}}(x)
\displaystyle\leq IαiN,IXN(ω)ea2πc(I,x)η(ω)(I)(HχJ)fI(x)+\displaystyle\sum_{I\neq\alpha_{i}^{N},I\in X_{N}^{(\omega)}}e^{a\cdot 2\pi\cdot c(I,x)}\eta^{(\omega)}(I)\left(H\cdot\chi_{J}\right)\circ f_{I}(x)+
ea2πc(α1N,x)η(ω)(α1N)(HχJ)fα1N(x)+ea2πc(α2N,x)η(ω)(α2N)(HχJ)fα2N(x)\displaystyle e^{a\cdot 2\pi\cdot c(\alpha_{1}^{N},x)}\eta^{(\omega)}(\alpha_{1}^{N})(H\cdot\chi_{J})\circ f_{\alpha_{1}^{N}}(x)+e^{a\cdot 2\pi\cdot c(\alpha_{2}^{N},x)}\eta^{(\omega)}(\alpha_{2}^{N})(H\cdot\chi_{J})\circ f_{\alpha_{2}^{N}}(x)
=\displaystyle= NsJ(H)(x).\displaystyle N_{s}^{J}\left(H\right)(x).

The case (2,j)J(2,j)\in J but (1,j)J(1,j)\notin J is similar.

Case 2 Suppose now both (1,j)J(1,j)\in J and (2,j)J(2,j)\in J. Then both Θ1(x),Θ2(x)1\Theta_{1}(x),\Theta_{2}(x)\leq 1. Taking half of the sum of these inequalities, we deduce that

|e(a+ib)2πc(α1N,x)η(ω)(α1N)ffα1N(x)+e(a+ib)2πc(α2N,x)η(ω)(α2N)ffα2N(x)|\left|e^{(a+ib)\cdot 2\pi\cdot c(\alpha_{1}^{N},x)}\eta^{(\omega)}(\alpha_{1}^{N})f\circ f_{\alpha_{1}^{N}}(x)+e^{(a+ib)\cdot 2\pi\cdot c(\alpha_{2}^{N},x)}\eta^{(\omega)}(\alpha_{2}^{N})f\circ f_{\alpha_{2}^{N}}(x)\right|
(1θ)ea2πc(α1N,x)η(ω)(α1N)Hfα1N(x)+(1θ)ea2πc(α2N,x)η(ω)(α2N)Hfα2N(x)\leq\left(1-\theta\right)e^{a\cdot 2\pi\cdot c(\alpha_{1}^{N},x)}\eta^{(\omega)}(\alpha_{1}^{N})H\circ f_{\alpha_{1}^{N}}(x)+\left(1-\theta\right)e^{a\cdot 2\pi\cdot c(\alpha_{2}^{N},x)}\eta^{(\omega)}(\alpha_{2}^{N})H\circ f_{\alpha_{2}^{N}}(x)
ea2πc(α1N,x)η(ω)(α1N)(HχJ)fα1N(x)+ea2πc(α2N,x)η(ω)(α2N)(HχJ)fα2N(x).\leq e^{a\cdot 2\pi\cdot c(\alpha_{1}^{N},x)}\eta^{(\omega)}(\alpha_{1}^{N})(H\cdot\chi_{J})\circ f_{\alpha_{1}^{N}}(x)+e^{a\cdot 2\pi\cdot c(\alpha_{2}^{N},x)}\eta^{(\omega)}(\alpha_{2}^{N})(H\cdot\chi_{J})\circ f_{\alpha_{2}^{N}}(x).

Since for every IXN(ω)I\in X_{N}^{(\omega)} with IαiNI\neq\alpha_{i}^{N} we have χJfI(x)=1\chi_{J}\circ f_{I}(x)=1, it follows that, similarly to case 1,

|Ps,ω,N(f)(x)|NsJ(H)(x).\left|P_{s,\omega,N}\left(f\right)(x)\right|\leq N_{s}^{J}\left(H\right)(x).

The proof is complete. \hfill{\Box}

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 Φ\Phi is C2C^{2}, and that ν𝐩\nu_{\mathbf{p}} is a self-conformal measure such that 𝐩\mathbf{p} is a strictly positive probability vector. Recall that =𝐩\mathbb{P}=\mathbf{p}^{\mathbb{N}} is our Bernoulli measure on 𝒜\mathcal{A}^{\mathbb{N}}.

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 t>0t>0 let

t={z:|z|<t}.\mathbb{C}_{t}=\{z\in\mathbb{C}:\,|\mathcal{R}z|<t\}.
Theorem 3.1.

Let ϵ,γ>0\epsilon,\gamma>0 be as in Theorem 2.8.

  1. 1.

    [12, Lemma 11.17] The transfer operator Ps+iθP_{s+i\theta} is a bounded operator on C1([0,1])C^{1}([0,1]) for every |s|<ϵ|s|<\epsilon and θ\theta, that depends analytically on s+iθs+i\theta.

  2. 2.

    [33, Proposition 4.1] There exists an analytic operator U(s+iθ)U(s+i\theta) for s+iθϵs+i\theta\in\mathbb{C}_{\epsilon} on C1([0,1])C^{1}([0,1]) such that, for s+iθϵs+i\theta\in\mathbb{C}_{\epsilon},

    (IP(s+iθ))1=1χ(s+iθ)N0+U(s+iθ), where N0(f)=f𝑑.(I-P_{-(s+i\theta)})^{-1}=\frac{1}{\chi(s+i\theta)}N_{0}+U(s+i\theta),\,\text{ where }N_{0}(f)=\int f\,d\mathbb{P}.
  3. 3.

    [34, Proposition 4.26] There exists some C>0C>0 such that for all s+iθϵs+i\theta\in\mathbb{C}_{\epsilon},

    U(s+iθ)C(1+|θ|)1+γ, where the operator norm is defined via (20).||U(s+i\theta)||\leq C(1+\left|\theta\right|)^{1+\gamma},\,\text{ where the operator norm is defined via \eqref{Eq B-Q norm}.}

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 UU when |θ|<R|\theta|<R for the R>0R>0 as in Theorem 2.8, and otherwise from a direct application of Theorem 2.8 via the identity

(IPs+iθ)1=n=0Ps+iθn.(I-P_{s+i\theta})^{-1}=\sum_{n=0}^{\infty}P_{s+i\theta}^{n}.

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 ff on [0,1]×[0,1]\times\mathbb{R} and (z,t)[0,1]×(z,t)\in[0,1]\times\mathbb{R}, define

Rf(z,t):=n=0f(η.z,σ(η,z)t)d𝐩n(η).Rf(z,\,t):=\sum_{n=0}^{\infty}\int f(\eta.z,\,\sigma(\eta,\,z)-t)d\mathbf{p}^{n}(\eta).

Since ff is positive, this sum is well defined. For every (z,x)[0,1]×(z,x)\in[0,1]\times\mathbb{R} let fz:f_{z}:\mathbb{R}\rightarrow\mathbb{R} be the function

fz(x):=f(z,x).f_{z}(x):=f(z,x).

We also define, for (z,θ)[0,1]×(z,\theta)\in[0,1]\times\mathbb{C} the Fourier transform

f^(z,θ):=eiθuf(z,u)𝑑u=fz^(θ).\hat{f}(z,\theta):=\int e^{i\theta u}f(z,u)\,du=\hat{f_{z}}(\theta).

The following Proposition is the main result of this Section:

Proposition 3.2.

Let ϵ\epsilon be as in Theorem 2.8, and let CC\subseteq\mathbb{R} be a compact set. Fix a non-negative bounded and continuous function f(y,x)f(y,x) on [0,1]×[0,1]\times\mathbb{R} such that fyCC3()f_{y}\in C_{C}^{3}(\mathbb{R}) for every y[0,1]y\in[0,1], and f^z(θ)C1([0,1])\hat{f}_{z}(\theta)\in C^{1}([0,1]) for every fixed θ\theta\in\mathbb{R}. Assume

supy(|fy(3)|L1+|fy|L1)<.\sup_{y}\left(|f_{y}^{(3)}|_{L^{1}}+|f_{y}|_{L^{1}}\right)<\infty.

Then for every z[0,1]z\in[0,1] and tt\in\mathbb{R}

Rf(z,t)=1χ[0,1]tf(y,u)𝑑u𝑑ν(y)+eϵ|t|O(eϵ|suppfz|(|(fz)(3)|L1+|fz|L1))Rf(z,t)=\frac{1}{\chi}\int_{[0,1]}\int_{-t}^{\infty}f(y,u)\,dud\nu(y)+e^{-\epsilon\cdot|t|}O\left(e^{\epsilon|\text{supp}f_{z}|}\left(|(f_{z})^{(3)}|_{L^{1}}+|f_{z}|_{L^{1}}\right)\right)

where |supp(fz)||\text{supp}(f_{z})| is the supremum of the absolute value of the elements of supp(f)\text{supp}(f).

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.

We begin by arguing, similarly to [33, Lemma 4.6] and [15, Proposition 4.14], that for every (z,t)[0,1]×(z,t)\in[0,1]\times\mathbb{R},

Rf(z,t)=1χ[0,1]tf(y,u)𝑑u𝑑ν(y)+lims0+12πeitθU(siθ)f^(z,θ)𝑑θ,Rf(z,t)=\frac{1}{\chi}\int_{[0,1]}\int_{-t}^{\infty}f(y,u)\,dud\nu(y)+\lim_{s\rightarrow 0^{+}}\frac{1}{2\pi}\int e^{-it\theta}U(s-i\theta)\hat{f}(z,\theta)\,d\theta, (27)

Indeed, for s0s\geq 0 write

Bsf(z,t)=esσ(i,z)f(i.z,σ(i,z)+t)d𝐩(i),B_{s}f(z,t)=\int e^{-s\sigma(i,\,z)}f(i.z,\,\sigma(i,z)+t)d\mathbf{p}(i),

and put B:=B0B:=B_{0}. Note that

Rf(z,t)=n0Bnf(z,t).Rf(z,-t)=\sum_{n\geq 0}B^{n}f(z,t).

By the monotone convergence Theorem, since f0f\geq 0 and σ(η,z)0\sigma(\eta,z)\geq 0 we have

lims0+n0esσ(η,z)f(η.z,σ(η,z)+t)d𝐩n(η)=n0f(η.z,σ(η,z)+t)d𝐩n(η).\lim_{s\rightarrow 0^{+}}\sum_{n\geq 0}\int e^{-s\sigma(\eta,\,z)}f(\eta.z,\,\sigma(\eta,\,z)+t)d\mathbf{p}^{n}(\eta)=\sum_{n\geq 0}\int f(\eta.z,\,\sigma(\eta,\,z)+t)d\mathbf{p}^{n}(\eta).

Thus,

n0Bn(f)(z,t)=lims0+n0Bsn(f)(z,t).\sum_{n\geq 0}B^{n}(f)(z,t)=\lim_{s\rightarrow 0^{+}}\sum_{n\geq 0}B_{s}^{n}(f)(z,t). (28)

Recall that fyCC3()L1()f_{y}\in C_{C}^{3}(\mathbb{R})\subset L^{1}(\mathbb{R}) for every y[0,1]y\in[0,1]. So, using the inverse Fourier transform, we have

n0Bsn(f)(z,t)=n0esσ(η,z)f(η.z,σ(η,z)+t)d𝐩n(η)\sum_{n\geq 0}B_{s}^{n}(f)(z,t)=\sum_{n\geq 0}\int e^{-s\sigma(\eta,\,z)}f(\eta.z,\,\sigma(\eta,\,z)+t)d\mathbf{p}^{n}(\eta)
=n0esσ(η,z)12πeiθ(σ(η,z)+t)f^(η.z,θ)dθd𝐩n(η).=\sum_{n\geq 0}\int e^{-s\sigma(\eta,\,z)}\frac{1}{2\pi}\int_{\mathbb{R}}e^{i\theta(\sigma(\eta,z)+t)}\hat{f}(\eta.z,\,\theta)d\theta d\mathbf{p}^{n}(\eta). (29)

We now argue that the sum in (29) is absolutely convergent: Since fyf_{y} is compactly supported for every y[0,1]y\in[0,1], fy^Cω()\hat{f_{y}}\in C^{\omega}(\mathbb{C}). Note that for every y[0,1]y\in[0,1], k{0}k\in\mathbb{N}\cup\{0\}, δ0\delta\geq 0, and θ\theta\in\mathbb{R},

|fy^(±iδ+θ)|eδ|suppfy|1|θ|k|(fy)(k)|L1\left|\hat{f_{y}}(\pm i\delta+\theta)\right|\leq e^{\delta|\text{supp}f_{y}|}\frac{1}{|\theta|^{k}}|(f_{y})^{(k)}|_{L^{1}}

So, plugging in k=0k=0 and k=3k=3,

|fy^(±iδ+θ)|eδ|suppfy|2|θ|3+1(|(fy)(3)|L1+|fy|L1)\left|\hat{f_{y}}(\pm i\delta+\theta)\right|\leq e^{\delta|\text{supp}f_{y}|}\frac{2}{|\theta|^{3}+1}\left(|(f_{y})^{(3)}|_{L^{1}}+|f_{y}|_{L^{1}}\right) (30)

Thus, as long as s>0s>0, putting C=2|θ|3+1𝑑θC=\int_{\mathbb{R}}\frac{2}{|\theta|^{3}+1}d\theta,

n0esσ(η,z)|f^(η.z,θ)|dθd𝐩n(η)Ceδ|suppfy|supy(|fy(3)|L1+|fy|L1)n0esσ(η,z)d𝐩n(η)\sum_{n\geq 0}\int e^{-s\sigma(\eta,\,z)}\int_{\mathbb{R}}|\hat{f}(\eta.z,\,\theta)|d\theta d\mathbf{p}^{n}(\eta)\leq C\cdot e^{\delta|\text{supp}f_{y}|}\cdot\sup_{y}\left(|f_{y}^{(3)}|_{L^{1}}+|f_{y}|_{L^{1}}\right)\cdot\sum_{n\geq 0}\int e^{-s\sigma(\eta,\,z)}\cdot d\mathbf{p}^{n}(\eta)
Ceδ|suppfy|supy(|fy(3)|L1+|fy|L1)n0esDn<.\leq C\cdot e^{\delta|\text{supp}f_{y}|}\cdot\sup_{y}\left(|f_{y}^{(3)}|_{L^{1}}+|f_{y}|_{L^{1}}\right)\cdot\sum_{n\geq 0}e^{-sD\cdot n}<\infty.

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 θ\theta\in\mathbb{R} f^(,θ)C1([0,1])\hat{f}(\cdot,\theta)\in C^{1}([0,1]), we can apply Theorem 3.1 to obtain

n0Bsn(f)(z,t)\displaystyle\sum_{n\geq 0}B_{s}^{n}(f)(z,t) =\displaystyle= 12πn0e(s+iθ)σ(η,z)f^(η.z,θ)d𝐩n(η)eitθdθ\displaystyle\frac{1}{2\pi}\int_{\mathbb{R}}\sum_{n\geq 0}\int e^{(-s+i\theta)\sigma(\eta,z)}\hat{f}(\eta.z,\,\theta)d\mathbf{p}^{n}(\eta)e^{it\theta}d\theta
=\displaystyle= 12πn0Ps+iθnf^(z,θ)eitθdθ\displaystyle\frac{1}{2\pi}\int_{\mathbb{R}}\sum_{n\geq 0}P_{-s+i\theta}^{n}\hat{f}(z,\,\theta)e^{it\theta}d\theta
=\displaystyle= 12π(1Ps+iθ)1f^(z,θ)eitθ𝑑θ\displaystyle\frac{1}{2\pi}\int_{\mathbb{R}}(1-P_{-s+i\theta})^{-1}\hat{f}(z,\,\theta)e^{it\theta}d\theta
=\displaystyle= 12π(1P(siθ))1f^(z,θ)eitθ𝑑θ\displaystyle\frac{1}{2\pi}\int_{\mathbb{R}}(1-P_{-(s-i\theta)})^{-1}\hat{f}(z,\,\theta)e^{it\theta}d\theta
=\displaystyle= 12π(1χ(siθ)N0+U(siθ))f^(z,θ)eitθ𝑑θ.\displaystyle\frac{1}{2\pi}\int_{\mathbb{R}}\left(\frac{1}{\chi(s-i\theta)}N_{0}+U(s-i\theta)\right)\hat{f}(z,\,\theta)e^{it\theta}d\theta.

Since 1siθ=0e(siθ)u𝑑u\frac{1}{s-i\theta}=\int_{0}^{\infty}e^{-(s-i\theta)u}du for s>0s>0, and since fy^L1()\hat{f_{y}}\in L^{1}(\mathbb{R}) for all yy by (30), we have

12πN0χ(siθ)f^(z,θ)eitθ𝑑θ\displaystyle\frac{1}{2\pi}\int_{\mathbb{R}}\frac{N_{0}}{\chi(s-i\theta)}\hat{f}(z,\,\theta)e^{it\theta}d\theta =\displaystyle= 12π1χ[0,1]f^(y,θ)siθeitθ𝑑θ𝑑ν(y)\displaystyle\frac{1}{2\pi}\frac{1}{\chi}\int_{[0,1]}\int_{\mathbb{R}}\frac{\hat{f}(y,\,\theta)}{s-i\theta}e^{it\theta}d\theta d\nu(y)
=\displaystyle= 1χ[0,1]0f(y,u+t)esu𝑑u𝑑ν(y).\displaystyle\frac{1}{\chi}\int_{[0,1]}\int_{0}^{\infty}f(y,\,u+t)e^{-su}dud\nu(y).

When s0+s\rightarrow 0^{+}, since ff is integrable with respect to ν×du\nu\times du, this converges to

1χ0f(y,u+t)𝑑u𝑑ν(y)\frac{1}{\chi}\int\int_{0}^{\infty}f(y,\,u+t)dud\nu(y)

by monotone convergence. Taking tally of our computations, (27) is proved.

Furthermore, using the bound on the norm of UU from Theorem 3.1 and (30), by dominated convergence

lims0+12πeitθU(siθ)f^(z,θ)𝑑θ=12πeitθU(iθ)f^(z,θ)𝑑θ.\lim_{s\rightarrow 0^{+}}\frac{1}{2\pi}\int e^{-it\theta}U(s-i\theta)\hat{f}(z,\theta)\,d\theta=\frac{1}{2\pi}\int e^{-it\theta}U(-i\theta)\hat{f}(z,\theta)\,d\theta.

Define

Tz(θ)=U(iθ)f^(z,θ).T_{z}(\theta)=U(-i\theta)\hat{f}(z,\theta).

Then TzT_{z} is a tempered distribution with an analytic continuation to |θ|<ϵ|\Im\theta|<\epsilon by Theorem 2.8 and Theorem 3.1, such that for any |b|<ϵ|b|<\epsilon we have Tz(+ib)L1T_{z}(\cdot+ib)\in L^{1}. Furthermore, by Theorem 3.1 and (30), for any |b|<ϵ|b|<\epsilon, supη<|b|||T(+iη)||L1<\sup_{\eta<|b|}||T(\cdot+i\eta)||_{L^{1}}<\infty (see the computation below). Applying [34, Lemma 4.28] we have, for all 0<δ<ϵ0<\delta<\epsilon

|12πeitθU(iθ)f^(z,θ)𝑑θ|=|Tˇ(t)|eδ|t|max|T(±iδ+θ)|L1(θ).\left|\frac{1}{2\pi}\int e^{-it\theta}U(-i\theta)\hat{f}(z,\theta)\,d\theta\right|=\left|\check{T}(t)\right|\leq e^{-\delta|t|}\max\left|T(\pm i\delta+\theta)\right|_{L^{1}(\theta)}.

Finally, by (30) and Theorem 3.1,

max|U(iθ±δ)f^(z,θ±iδ)|L1(θ)U(±δiθ)eδ|suppfz||θ|3+1(|(fz)(3)|L1+|fz|L1)𝑑θ\max\left|U(-i\theta\pm\delta)\hat{f}(z,\theta\pm i\delta)\right|_{L^{1}(\theta)}\leq\int\left||U(\pm\delta-i\theta)\right||\cdot\frac{e^{\delta|\text{supp}f_{z}|}}{|\theta|^{3}+1}\left(|(f_{z})^{(3)}|_{L^{1}}+|f_{z}|_{L^{1}}\right)\,d\theta
Ceδ|suppfz|(|(fz)(3)|L1+|fz|L1)1|θ|3+1(1+|θ|)1+γ𝑑θ=O(eδ|suppfz|(|(fz)(3)|L1+|fz|L1)),\leq C\cdot e^{\delta|\text{supp}f_{z}|}\left(|(f_{z})^{(3)}|_{L^{1}}+|f_{z}|_{L^{1}}\right)\int\frac{1}{|\theta|^{3}+1}\cdot(1+|\theta|)^{1+\gamma}\,d\theta=O\left(e^{\delta|\text{supp}f_{z}|}\left(|(f_{z})^{(3)}|_{L^{1}}+|f_{z}|_{L^{1}}\right)\right),

where in the last equality we used that 3(1+γ)>13-(1+\gamma)>1. 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 Φ\Phi is C2C^{2}, and that ν𝐩\nu_{\mathbf{p}} is a self-conformal measure such that 𝐩\mathbf{p} is a strictly positive probability vector. Recall that =𝐩\mathbb{P}=\mathbf{p}^{\mathbb{N}} is our Bernoulli measure on 𝒜\mathcal{A}^{\mathbb{N}}.

Recall the definition of the semigroup GG 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 c~\tilde{c}: It is defined as c~:G×𝒜\tilde{c}:G\times\mathcal{A}^{\mathbb{N}}\rightarrow\mathbb{R}

c~(I,ω)=logfI(xω).\tilde{c}(I,\omega)=-\log f^{\prime}_{I}(x_{\omega}). (31)

We can now define a random walk as follows: Recalling that σ:𝒜𝒜\sigma:\mathcal{A}^{\mathbb{N}}\rightarrow\mathcal{A}^{\mathbb{N}} is the left shift, for every nn\in\mathbb{N} we define a function on 𝒜\mathcal{A}^{\mathbb{N}} via

Sn(ω)=logfω|n(xσn(ω)).S_{n}(\omega)=-\log f^{\prime}_{\omega|_{n}}(x_{\sigma^{n}(\omega)}). (32)

Let X1:𝒜X_{1}:\mathcal{A}^{\mathbb{N}}\rightarrow\mathbb{R} be the random variable

X1(ω):=c~(ω1,σ(ω))=logfω1(xσ(ω)).X_{1}(\omega):=\tilde{c}(\omega_{1},\sigma(\omega))=-\log f^{\prime}_{\omega_{1}}(x_{\sigma(\omega)}). (33)

For every integer n>1n>1 we define

Xn(ω)=logfωn(xσn(xω))=X1σn1.X_{n}(\omega)=-\log f_{\omega_{n}}^{\prime}\left(x_{\sigma^{n}(x_{\omega})}\right)=X_{1}\circ\sigma^{n-1}.

Let κ\kappa be the law of the random variable X1X_{1}. Then for every nn, XnκX_{n}\sim\kappa. By uniform contraction there exists D,DD,D^{\prime}\in\mathbb{R} as in (13), so κ𝒫([D,D])\kappa\in\mathcal{P}([D,D^{\prime}]). In particular, the support of κ\kappa is bounded away from 0. It is easy to see that for every nn\in\mathbb{N} and ωA\omega\in A^{\mathbb{N}} we have

Sn(ω)=i=1nXi(ω).S_{n}(\omega)=\sum_{i=1}^{n}X_{i}(\omega).

Thus, in this sense SnS_{n} is a random walk.

Next, we define a function on 𝒜\mathcal{A}^{\mathbb{N}} that resembles a stopping time: For k>0k>0 let

τk(ω):=min{n:Sn(ω)k}.\tau_{k}(\omega):=\min\{n:S_{n}(\omega)\geq k\}.

We emphasize that kk is allowed to take non-integer values. We also recall that χ\chi is the corresponding Lyapunov exponent. Recalling (13), it is clear that for every k>0k>0 and ωA\omega\in A^{\mathbb{N}} we have

log|fω|τk(ω)(xστk(ω)(ω))|=Sτk(ω)(ω)[k,k+D].-\log|f_{\omega|_{\tau_{k}(\omega)}}^{\prime}(x_{\sigma^{\tau_{k}(\omega)}(\omega)})|=S_{\tau_{k}(\omega)}(\omega)\in[k,k+D^{\prime}].

We also define a local C3C^{3} norm on [1D,D+1][-1-D,D^{\prime}+1] by, for gC3()g\in C^{3}(\mathbb{R}),

gC3=maxi=0,1,2,3maxx[1D,D+1]|g(i)(x)|.||g||_{C^{3}}=\max_{i=0,1,2,3}\max_{x\in[-1-D,D^{\prime}+1]}\left|g^{(i)}(x)\right|. (34)

The following Theorem is an effective equidistribution result for the random variables SτkS_{\tau_{k}}:

Theorem 4.1.

Let ϵ>0\epsilon>0 be as in Proposition 3.2. Then for every k>D+1k>D^{\prime}+1, gC3()g\in C^{3}(\mathbb{R}), and ω𝒜\omega\in\mathcal{A}^{\mathbb{N}},

𝔼(g(Sτk(η)(η)k)|στk(η)η=ω)=1χDDy0g(x+y)𝑑x𝑑κ(y)+eϵk2O(gC3).\mathbb{E}\left(g\left(S_{\tau_{k}(\eta)}(\eta)-k\right)\big{|}\,\sigma^{\tau_{k}(\eta)}\eta=\omega\right)=\frac{1}{\chi}\int_{D}^{D^{\prime}}\int_{-y}^{0}g(x+y)\,dxd\kappa(y)+e^{-\frac{\epsilon k}{2}}O\left(||g||_{C^{3}}\right).

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 SτkS_{\tau_{k}} necessarily has slower equidistribution rates (no faster than polynomial is possible). This also makes our treatment more complicated since the random walk SnS_{n} 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 [0,1]×[0,1]\times\mathbb{R}, rather than just on \mathbb{R}. 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 ψ\psi be the even smooth bump function given by

ψ(x):=C0exp(11x2) if x(1,1), and 0 otherwise.\psi(x):=C_{0}\cdot\exp\left(\frac{-1}{1-x^{2}}\right)\text{ if }x\in(-1,1),\text{ and }0\text{ otherwise.}

Here C0C_{0} is chosen so that ψ\psi is also a probability density function. Next, for δ0\delta\neq 0 write

ψδ(x):=1δ2ψ(xδ2).\psi_{\delta}(x):=\frac{1}{\delta^{2}}\psi(\frac{x}{\delta^{2}}).

The following Proposition is an analogue of [36, Proposition 4.5]:

Proposition 4.2.

There exists some C1=C1(ψ)>0C_{1}=C_{1}(\psi)>0 such that, for our ϵ>0\epsilon>0 from Theorem 2.8:

For all k>0k>0, z[0,1]z\in[0,1], δ1\delta\leq 1, and b,ab,a such that ba2δb-a\geq 2\delta,

R(1[a,b])(z,k)3(ba)(1χ+eϵ(k+ba+2)CψO(1δ6+1)),R\left(1_{[a,b]}\right)(z,k)\leq 3(b-a)\left(\frac{1}{\chi}+e^{\epsilon\left(-k+b-a+2\right)}C_{\psi}O\left(\frac{1}{\delta^{6}}+1\right)\right),

where the function 1[a,b](z,x)=1[a,b](x)1_{[a,b]}(z,x)=1_{[a,b]}(x) is constant in zz (the [0,1][0,1] 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 x[a,b]x\in[a,b] then [xb,xa][x-b,x-a] contains either [0,δ][0,\delta] or [δ,0][-\delta,0]. Therefore, as ψδ\psi_{\delta} is even

ψδ1[a,b](x)=abψδ(xv)𝑑v0δψδ(v)𝑑v=12.\psi_{\delta}*1_{[a,b]}(x)=\int_{a}^{b}\psi_{\delta}(x-v)dv\geq\int_{0}^{\delta}\psi_{\delta}(v)dv=\frac{1}{2}.

So, as functions on \mathbb{R},

1[a,b]3ψδ1[a,b].1_{[a,b]}\leq 3\psi_{\delta}*1_{[a,b]}.

We proceed to bound R(ψδ1[a,b])R(\psi_{\delta}*1_{[a,b]}), where we note that ψδ1[a,b]\psi_{\delta}*1_{[a,b]} is a compactly supported function in C3()C^{3}(\mathbb{R}). By Proposition 3.2, since our function is constant in the [0,1][0,1]-coordinate,

R(ψδ1[a,b])(z,k)=1χkψδ1[a,b](x)𝑑x+eϵ|k|O(eϵ|suppψδ1[a,b]|(|(ψδ1[a,b])(3)|L1+|ψδ1[a,b]|L1)).R(\psi_{\delta}*1_{[a,b]})(z,k)=\frac{1}{\chi}\int_{-k}^{\infty}\psi_{\delta}*1_{[a,b]}(x)dx+e^{-\epsilon\cdot|k|}O\left(e^{\epsilon|\text{supp}\psi_{\delta}*1_{[a,b]}|}\left(|(\psi_{\delta}*1_{[a,b]})^{(3)}|_{L^{1}}+|\psi_{\delta}*1_{[a,b]}|_{L^{1}}\right)\right).

Up to global multiplicative constants, the first and third terms are less than ψδ1[a,b](x)𝑑x=(ba)\int\psi_{\delta}*1_{[a,b]}(x)dx=(b-a). For the second term, recall that

(ψδ1[a,b])(3)=ψδ(3)1[a,b](\psi_{\delta}*1_{[a,b]})^{(3)}=\psi_{\delta}^{(3)}*1_{[a,b]}

and so

(ψδ1[a,b])3)L1ψδ(3)L11[a,b]L1Cψ1δ23(ba).||(\psi_{\delta}*1_{[a,b]})^{3)}||_{L^{1}}\leq||\psi_{\delta}^{(3)}||_{L^{1}}\cdot||1_{[a,b]}||_{L^{1}}\leq C_{\psi}\cdot\frac{1}{\delta^{2\cdot 3}}\cdot(b-a).

Finally, ψδ1[a,b]\psi_{\delta}*1_{[a,b]} is supported on [a,b]+[δ2,δ2][a1,b+1][a,b]+[-\delta^{2},\delta^{2}]\subseteq[a-1,b+1]. 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 C>0C>0 such that for all ss, z[0,1]z\in[0,1], and kk\in\mathbb{R} we have

R(1[0,s])(z,k)Cmax{1,s},R(1_{[0,s]})(z,k)\leq C\cdot\max\{1,s\},

where the function 1[0,s](y,x)=1[0,s](x)1_{[0,s]}(y,x)=1_{[0,s]}(x) is constant in the [0,1][0,1] coordinate.

4.2 Residue process

Let ff be a non-negative bounded Borel function on [D,D]×[D,D^{\prime}]\times\mathbb{R}. For kk\in\mathbb{R} and z[0,1]z\in[0,1] define the residue operator by

Ef(z,k):=n0f(σ(i,η.z),σ(η,z)k)d𝐩n(η)d𝐩(i).Ef(z,k):=\sum_{n\geq 0}\int\int f(\sigma(i,\eta.z),\,\sigma(\eta,z)-k)d\mathbf{p}^{n}(\eta)d\mathbf{p}(i).

Recall that for every y[0,1]y\in[0,1] we let fy:f_{y}:\mathbb{R}\rightarrow\mathbb{R} be the function

fy(x):=f(y,x).f_{y}(x):=f(y,x).

The following Proposition is an analogue of [36, Proposition 4.7]:

Proposition 4.4.

(Residue process) Let CC\subseteq\mathbb{R} be a compact set. Fix a non-negative bounded, compactly supported, and continuous function f(y,x)f(y,x) on [D,D]×[D,D^{\prime}]\times\mathbb{R} such that fyCC3()f_{y}\in C_{C}^{3}(\mathbb{R}) for every y[D,D]y\in[D,D^{\prime}], and f^z(θ)C1([D,D])\hat{f}_{z}(\theta)\in C^{1}([D,D^{\prime}]) for every θ\theta\in\mathbb{R}. Assume

supy(|fy(3)|L1+|fy|L1)<.\sup_{y}\left(|f_{y}^{(3)}|_{L^{1}}+|f_{y}|_{L^{1}}\right)<\infty.

Then for every k>0k>0 and z[0,1]z\in[0,1],

Ef(z,k)=1χk𝒜[0,1]f(σ(i,y),x)𝑑ν(y)𝑑𝐩(i)𝑑x+eϵkO(eϵ|suppf|(supy,y[D,D]|(fy)(3)|L1+|fy|L1))Ef(z,k)=\frac{1}{\chi}\int_{-k}^{\infty}\int_{\mathcal{A}}\int_{[0,1]}f(\sigma(i,y),x)d\nu(y)d\mathbf{p}(i)dx+e^{-\epsilon\cdot k}O\left(e^{\epsilon|\text{supp}f|}\left(\sup_{y,y^{\prime}\in[D^{\prime},D]}|(f_{y})^{(3)}|_{L^{1}}+|f_{y^{\prime}}|_{L^{1}}\right)\right)
Proof.

For z[0,1]z\in[0,1] and uu\in\mathbb{R} define a non-negative bounded Borel function

Qf(z,u)=f(σ(i,z),u)𝑑𝐩(i).Qf(z,u)=\int f(\sigma(i,z),u)d\mathbf{p}(i).

Then

Ef(z,k)=n0Qf(η.z,σ(η,z)t)d𝐩n(η)=R(Qf)(z,k).Ef(z,k)=\sum_{n\geq 0}Qf(\eta.z,\sigma(\eta,z)-t)d\mathbf{p}^{n}(\eta)=R(Qf)(z,k).

Since Φ\Phi is a C2C^{2} IFS, the map zσ(i,z)C1([0,1],[D,D])z\mapsto\sigma(i,z)\in C^{1}([0,1],\,[D^{\prime},D]) for all i𝒜i\in\mathcal{A}, and therefore zf^σ(i,z)(θ)C1([0,1])z\mapsto\hat{f}_{\sigma(i,z)}(\theta)\in C^{1}([0,1]) for all i𝒜i\in\mathcal{A} and θ\theta\in\mathbb{R}. So, as our assumptions on ff imply that Qf(z,k)Qf(z,k) 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 ECE_{C} on real non-negative bounded Borel functions on [D,D]×[D,D^{\prime}]\times\mathbb{R} via, for (z,k)[0,1]×(z,k)\in[0,1]\times\mathbb{R},

ECf(z,k)=n0σ(η,z)<kσ(i,η.z)+σ(η,z)f(σ(i,η.z),σ(η,z)k)d𝐩(i)d𝐩n(η).E_{C}f(z,k)=\sum_{n\geq 0}\int_{\sigma(\eta,z)<k\leq\sigma(i,\eta.z)+\sigma(\eta,z)}f(\sigma(i,\eta.z),\,\sigma(\eta,z)-k)d\mathbf{p}(i)d\mathbf{p}^{n}(\eta).

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 C2>0C_{2}>0 such that for all kk\in\mathbb{R} and z[0,1]z\in[0,1] we have

EC(𝟏)(z,k)C2E_{C}(\mathbf{1})(z,k)\leq C_{2}
Proof.

Recalling (13), we have

EC(𝟏)(z,k)\displaystyle E_{C}(\mathbf{1})(z,k) =\displaystyle= n0𝐩×𝐩n({(i,η):σ(η,z)k[σ(i,η.z),0)})\displaystyle\sum_{n\geq 0}\mathbf{p}\times\mathbf{p}^{n}\left(\{(i,\eta):\sigma(\eta,z)-k\in[-\sigma(i,\eta.z),0)\}\right)
\displaystyle\leq n0𝐩×𝐩n({(i,η):σ(η,z)k[D,0]})\displaystyle\sum_{n\geq 0}\mathbf{p}\times\mathbf{p}^{n}\left(\{(i,\eta):\sigma(\eta,z)-k\in[-D^{\prime},0]\}\right)
=\displaystyle= R(1[D,0])(z,k)\displaystyle R\left(1_{[-D^{\prime},0]}\right)(z,k)
\displaystyle\leq Cmax{1,D},\displaystyle C\cdot\max\{1,D^{\prime}\},

where in the last inequality we applied Lemma 4.3 (or Proposition 4.2). ∎

By Lemma 4.5 the operator ECE_{C} is well defined for bounded Borel functions. For a function f(y,x)f(y,x) on 2\mathbb{R}^{2} we denote, for all xx,

fx(y)=f(y,x).f_{x}(y)=f(y,x).

We have the following analogue of [36, Proposition 4.10]:

Proposition 4.6.

Let K2K\subseteq\mathbb{R}^{2} be a compact set, and fix a bounded and continuous function f(y,x)f(y,x) on [D,D]×[D^{\prime},D]\times\mathbb{R} such that KK contains supp(f)\text{supp}(f), fyCP2K3()f_{y}\in C_{P_{2}K}^{3}(\mathbb{R}) for every y[D,D]y\in[D^{\prime},D], and fxC1([0,1])f_{x}\in C^{1}([0,1]) for every xx\in\mathbb{R}. Assume

supy(|fy(3)|L1+|fy|L1)<.\sup_{y}\left(|f_{y}^{(3)}|_{L^{1}}+|f_{y}|_{L^{1}}\right)<\infty.

Then for |K|<k|K|<k and every z[0,1]z\in[0,1] we have

ECf(z,k)=[0,1]𝒜σ(i,y)0f(σ(i,y),x)dxd𝐩(i)dν(y)+eϵk2O|K|((supy,y[D,D]||fy(3)||L1+||fy||L1)E_{C}f(z,k)=\int_{[0,1]}\int_{\mathcal{A}}\int_{-\sigma(i,y)}^{0}f(\sigma(i,y),x)dxd\mathbf{p}(i)d\nu(y)+e^{-\frac{\epsilon k}{2}}O_{|K|}(\left(\sup_{y,y^{\prime}\in[D^{\prime},D]}||f_{y}^{(3)}||_{L^{1}}+||f_{y^{\prime}}||_{L^{1}}\right)
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 ff is non-negative (with the same parameters assumed in the original Proposition).

The following Lemma relates the operators ECE_{C} and EE:

Lemma 4.8.

[36, Lemma 4.12] Under the assumptions of Proposition 4.6, let

fo(y,x):=1yx<0f(y,x).f_{o}(y,x):=1_{-y\leq x<0}f(y,x).

Then

ECf(z,k)=Efo(z,k).E_{C}f(z,k)=Ef_{o}(z,k).

Fix 1>δ>01>\delta>0 small enough so that |K|+δk|K|+\delta\leq k. We use ψδ\psi_{\delta} to regularize these functions; Let

fδ(y,x):=fo(y,xx1)ψδ(x1)𝑑x1=ψδfo(y,x)f_{\delta}(y,x):=\int f_{o}(y,x-x_{1})\psi_{\delta}(x_{1})dx_{1}=\psi_{\delta}*f_{o}(y,x)

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,

E(fδ)(z,k)=[0,1]𝒜σ(i,y)0f(σ(i,y),x)𝑑x𝑑𝐩(i)𝑑ν(y)+eϵkO(eϵ|suppfδ|(supy,y[D,D]|(fδ,y)(3)|L1+|fδ,y|L1))E(f_{\delta})(z,k)=\int_{[0,1]}\int_{\mathcal{A}}\int_{-\sigma(i,y)}^{0}f(\sigma(i,y),x)dxd\mathbf{p}(i)d\nu(y)+e^{-\epsilon\cdot k}O\left(e^{\epsilon|\text{supp}f_{\delta}|}\left(\sup_{y,y^{\prime}\in[D^{\prime},D]}|(f_{\delta,y})^{(3)}|_{L^{1}}+|f_{\delta,y^{\prime}}|_{L^{1}}\right)\right)
Proof.

We first wish to apply Proposition 4.4 to the function fδf_{\delta}. To this end, notice that by our assumptions for every y[D,D]y\in[D^{\prime},D] the function fyCP2K3()f_{y}\in C^{3}_{P_{2}K}(\mathbb{R}), and satisfies the required integrability conditions. Also, since fxC1([0,1])f_{x}\in C^{1}([0,1]) for every xx\in\mathbb{R}, it is clear that for every θ\theta\in\mathbb{R}, the function

z(fo^)z(θ)=eiθufo(z,u)𝑑u=z0eiθuf(z,u)𝑑uz\mapsto\left(\hat{f_{o}}\right)_{z}(\theta)=\int e^{i\theta u}f_{o}(z,u)\,du=\int_{-z}^{0}e^{i\theta u}f(z,u)\,du

belongs to C1([D,D])C^{1}([D,D^{\prime}]).

So, applying Proposition 4.4,

E(fδ)(z,k)=1χk𝒜[0,1]fδ(σ(i,y),x)𝑑ν(y)𝑑𝐩(i)𝑑x+eϵkO(eϵ|suppfδ|(supy,y|(fδ,y)(3)|L1+|fδ,y|L1))E(f_{\delta})(z,k)=\frac{1}{\chi}\int_{-k}^{\infty}\int_{\mathcal{A}}\int_{[0,1]}f_{\delta}(\sigma(i,y),x)d\nu(y)d\mathbf{p}(i)dx+e^{-\epsilon\cdot k}O\left(e^{\epsilon|\text{supp}f_{\delta}|}\left(\sup_{y,y^{\prime}}|(f_{\delta,y})^{(3)}|_{L^{1}}+|f_{\delta,y^{\prime}}|_{L^{1}}\right)\right)

Also, for every yP1supp(f)y\in P_{1}\text{supp}(f)

kfδ(y,x)𝑑x=ky0f(y,x1)ψδ(xx1)𝑑x1𝑑x=y0f(y,x1)kψδ(xx1)𝑑x𝑑x1.\int_{-k}^{\infty}f_{\delta}(y,x)dx=\int_{-k}^{\infty}\int_{-y}^{0}f(y,x_{1})\psi_{\delta}(x-x_{1})dx_{1}dx=\int_{-y}^{0}f(y,x_{1})\int_{-k}^{\infty}\psi_{\delta}(x-x_{1})dxdx_{1}.

Now, since kδ|K|k-\delta\geq|K| and yP1supp(f)y\in P_{1}\text{supp}(f) then

kx1k+yδ.-k-x_{1}\leq-k+y\leq-\delta.

It follows that for every x1P2(suppf)x_{1}\in P_{2}(\text{supp}f) we have

1kψδ(xx1)𝑑xδψδ=1.1\geq\int_{-k}^{\infty}\psi_{\delta}(x-x_{1})dx\geq\int_{-\delta}^{\infty}\psi_{\delta}=1.

So,

k𝒜[0,1]fδ(σ(i,y),x)𝑑ν(y)𝑑𝐩(i)𝑑x=[0,1]𝒜σ(i,y)0f(σ(i,y),x)𝑑x𝑑𝐩(i)𝑑ν(y)\int_{-k}^{\infty}\int_{\mathcal{A}}\int_{[0,1]}f_{\delta}(\sigma(i,y),x)d\nu(y)d\mathbf{p}(i)dx=\int_{[0,1]}\int_{\mathcal{A}}\int_{-\sigma(i,y)}^{0}f(\sigma(i,y),x)dxd\mathbf{p}(i)d\nu(y)

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 [1,1][-1,1].

Lemma 4.10.

Let φ\varphi be a C1()C^{1}(\mathbb{R}) function with φ<||\varphi^{\prime}||_{\infty}<\infty and φ1||\varphi||_{\infty}\leq 1. Let

φo(u)=1[a,b](u)φ(u).\varphi_{o}(u)=1_{[a,b]}(u)\varphi(u).

Then |ψδφo(u)φo(u)||\psi_{\delta}*\varphi_{o}(u)-\varphi_{o}(u)| is bounded by:

  • φδ||\varphi^{\prime}||_{\infty}\cdot\delta if u[a+δ,bδ]u\in[a+\delta,\,b-\delta].

  • 22 if u[aδ,a+δ]u\in[a-\delta,\,a+\delta] or u[bδ,b+δ]u\in[b-\delta,\,b+\delta].

  • ψδ1[a,b](u)\psi_{\delta}*1_{[a,b]}(u) if u<aδu<a-\delta or u>b+δu>b+\delta.

Proof.

If u[a+δ,bδ]u\in[a+\delta,b-\delta] then, since ψδ\psi_{\delta} is supported on [δ2,δ2][-\delta^{2},\delta^{2}]

|ψδφo(u)φo(u)|=|ψδ(t)(φo(ut)φo(u))𝑑t|\left|\psi_{\delta}*\varphi_{o}(u)-\varphi_{o}(u)\right|=\left|\int\psi_{\delta}(t)(\varphi_{o}(u-t)-\varphi_{o}(u))dt\right|
δδψδ(t)|φo(ut)φo(u)|𝑑t.\leq\int_{-\delta}^{\delta}\psi_{\delta}(t)\left|\varphi_{o}(u-t)-\varphi_{o}(u)\right|dt.

If |t|δ|t|\leq\delta then ut[a,b]u-t\in[a,b]. Since |φo(u)|φ|\varphi_{o}^{\prime}(u)|\leq||\varphi^{\prime}||_{\infty} for u[a,b]u\in[a,b] we obtain

δδψδ(t)|φo(ut)φo(u)|𝑑tδδψδ(t)|t|φ𝑑tδφ.\int_{-\delta}^{\delta}\psi_{\delta}(t)\left|\varphi_{o}(u-t)-\varphi_{o}(u)\right|dt\leq\int_{-\delta}^{\delta}\psi_{\delta}(t)\cdot\left|t\right|\cdot||\varphi^{\prime}||_{\infty}dt\leq\delta||\varphi^{\prime}||_{\infty}.

In the second case, we use the trivial bound

|ψδφo(u)φo(u)|2.|\psi_{\delta}*\varphi_{o}(u)-\varphi_{o}(u)|\leq 2.

In the third case, if u(,aδ)(b+δ,)u\in(-\infty,a-\delta)\cup(b+\delta,\infty) then φo(u)=0\varphi_{o}(u)=0, and so

|ψδφo||ψδ1[a,b]|.|\psi_{\delta}*\varphi_{o}|\leq|\psi_{\delta}*1_{[a,b]}|.

This gives the Lemma. ∎

Proof of Proposition 4.6 By Lemma 4.9, we only need to estimate E(|fδfo|)(z,k)E(|f_{\delta}-f_{o}|)(z,k).

Since fo(y,x)=1yx<0(x)f(y,x)f_{o}(y,x)=1_{-y\leq x<0}(x)f(y,x), applying Lemma 4.10, |fδfo|(x)|f_{\delta}-f_{o}|(x) is bounded by the sum of the following three terms (we sometimes omit the yy variable in the following computation):

  • supyfyδ\sup_{y}||f^{\prime}_{y}||_{\infty}\cdot\delta if x[y+δ,δ]x\in[-y+\delta,-\delta].

  • 22 if yδxy+δ-y-\delta\leq x\leq-y+\delta or δxδ-\delta\leq x\leq\delta

  • ψδ1[y,0](x)\psi_{\delta}*1_{[-y,0]}(x) if x<yδx<-y-\delta or x>δx>\delta.

By the definition of |K||K|, the first term is smaller than

supy|fy|δ1[|K|+δ,δ].\sup_{y}|f^{\prime}_{y}|_{\infty}\cdot\delta\cdot 1_{[-|K|+\delta,-\delta]}.

The third term is equal to

1[,yδ][δ,]ψδ1[y,0](x)=1[,yδ][δ,](x)y0ψδ(xx1)𝑑x11_{[-\infty,-y-\delta]\cup[\delta,\infty]}\psi_{\delta}*1_{[-y,0]}(x)=1_{[-\infty,-y-\delta]\cup[\delta,\infty]}(x)\int_{-y}^{0}\psi_{\delta}(x-x_{1})dx_{1}
=1[,yδ][δ,](x)xx+yψδ(x1)𝑑x1.=1_{[-\infty,-y-\delta]\cup[\delta,\infty]}(x)\int_{x}^{x+y}\psi_{\delta}(x_{1})dx_{1}.

This gives us

E(|fδfo|)(z,k)=n0|fδfo|(σ(i,η.z),σ(η,z)k)d𝐩n(η)d𝐩(i)E\left(\left|f_{\delta}-f_{o}\right|\right)(z,k)=\sum_{n\geq 0}\int\int|f_{\delta}-f_{o}|(\sigma(i,\eta.z),\,\sigma(\eta,\,z)-k)d\mathbf{p}^{n}(\eta)d\mathbf{p}(i)
n0(supy||fy||δ1[|K|,δ](σ(η,z)k)\leq\sum_{n\geq 0}\int(\sup_{y}||f^{\prime}_{y}||_{\infty}\cdot\delta\cdot 1_{[-|K|,-\delta]}(\sigma(\eta,z)-k)
+21[σ(i,η.z)δ,σ(i,η.z)+δ][δ,δ](σ(η,z)k)+2\cdot 1_{[-\sigma(i,\eta.z)-\delta,-\sigma(i,\eta.z)+\delta]\cup[-\delta,\delta]}(\sigma(\eta,z)-k)
+1[,σ(i,η.z)δ][δ,](σ(η,z)k)σ(η,z)kσ(η,z)k+σ(i,η.z)ψδ(x1)dx1)d𝐩n(η)d𝐩(i).+1_{[-\infty,-\sigma(i,\eta.z)-\delta]\cup[\delta,\infty]}(\sigma(\eta,z)-k)\int_{\sigma(\eta,z)-k}^{\sigma(\eta,z)-k+\sigma(i,\eta.z)}\psi_{\delta}(x_{1})dx_{1})d\mathbf{p}^{n}(\eta)d\mathbf{p}(i).

By Lemma 4.3 the first term is dominated by

supy|fy|δ|K|.\sup_{y}|f^{\prime}_{y}|_{\infty}\cdot\delta|K|.

For the second term, since σ\sigma is a cocycle,

1[σ(i,η.z)δ,σ(i,η.z)+δ](σ(η,z)k)𝑑𝐩n(η)𝐩(i)=1[δ,δ](σ(η,z)k)𝑑𝐩n+1(η).\int 1_{[-\sigma(i,\eta.z)-\delta,-\sigma(i,\eta.z)+\delta]}(\sigma(\eta,z)-k)d\mathbf{p}^{n}(\eta)\mathbf{p}(i)=\int 1_{[-\delta,\delta]}(\sigma(\eta,z)-k)d\mathbf{p}^{n+1}(\eta).

So, the second term is smaller than 4R(1[δ,δ])(z,k)4R\left(1_{[-\delta,\delta]}\right)(z,k). By proposition 4.2, this is dominated by

3δ(1χ+eϵkCψO(1δ6+1))3\delta\left(\frac{1}{\chi}+e^{-\epsilon\cdot k}C_{\psi}O\left(\frac{1}{\delta^{6}}+1\right)\right)

The third term, by a change of the order of integration (as in [36, Proof of Proposition 4.10]), is smaller than

[,δ]δ,]ψδ(x1)EC(1)(z,x1+k)𝑑x1.\int_{[-\infty,-\delta]\cup\delta,\infty]}\psi_{\delta}(x_{1})E_{C}(1)(z,x_{1}+k)dx_{1}.

By Lemma 4.5, this is smaller than

C2[,δ]δ,]ψδ(x1)𝑑x1.C_{2}\int_{[-\infty,-\delta]\cup\delta,\infty]}\psi_{\delta}(x_{1})dx_{1}.

This latter term is 0 since ψδ\psi_{\delta} is supported on [δ2,δ2][-\delta^{2},\delta^{2}].

Conclusion of proof Assuming 0<δ10<\delta\leq 1 is small enough, we have shown that

ECf(z,k)=[0,1]𝒜σ(i,y)0f(σ(i,y),x)𝑑x𝑑𝐩(i)𝑑ν(y)E_{C}f(z,k)=\int_{[0,1]}\int_{\mathcal{A}}\int_{-\sigma(i,y)}^{0}f(\sigma(i,y),x)dxd\mathbf{p}(i)d\nu(y)

with the sum following error terms: From Lemma 4.9 we have

eϵkO(eϵ|suppfδ|(supy,y|(fδ,y)(3)|L1+|fδ,y|L1))e^{-\epsilon\cdot k}O\left(e^{\epsilon|\text{supp}f_{\delta}|}\left(\sup_{y,y^{\prime}}|(f_{\delta,y})^{(3)}|_{L^{1}}+|f_{\delta,y^{\prime}}|_{L^{1}}\right)\right)

and from the previous argument above,

supy|fy|δ|K|+6δ(1χ+eϵkCψO(1δ6+1)).\sup_{y}|f^{\prime}_{y}|_{\infty}\cdot\delta|K|+6\delta\left(\frac{1}{\chi}+e^{-\epsilon\cdot k}C_{\psi}O\left(\frac{1}{\delta^{6}}+1\right)\right).

Note that, for every y[D,D]y\in[D,D^{\prime}],

|(fδ,y)(3)|L1=|(ψδfo(y,x))(3)|L1=|ψδ(fo(y,x))(3)|L1|ψδ|L1|(fo(y,x))(3)|L1|(fy)(3)|L1.|(f_{\delta,y})^{(3)}|_{L^{1}}=|(\psi_{\delta}*f_{o}(y,x))^{(3)}|_{L^{1}}=|\psi_{\delta}*(f_{o}(y,x))^{(3)}|_{L^{1}}\leq|\psi_{\delta}|_{L^{1}}\cdot|(f_{o}(y,x))^{(3)}|_{L^{1}}\leq|(f_{y})^{(3)}|_{L^{1}}.

Indeed, the second equality holds since ddxfo(y,x)\frac{d}{dx}f_{o}(y,x) 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

|fδ,y|L1|fy|L1.|f_{\delta,y}|_{L^{1}}\leq|f_{y}|_{L^{1}}.

Next, put

δ=eϵk12\delta=e^{-\frac{\epsilon k}{12}}

so that

eϵkδ6=eϵk2e^{-\epsilon k}\cdot\delta^{-6}=e^{-\frac{\epsilon k}{2}}

and the error becomes

eϵk2O|K|(supy,y[D,D]fyL1+fy(3)L1).e^{-\frac{\epsilon k}{2}}O_{|K|}\left(\sup_{y,y^{\prime}\in[D^{\prime},D]}||f_{y}||_{L^{1}}+||f_{y^{\prime}}^{(3)}||_{L^{1}}\right).

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 x[0,1]x\in[0,1] and kk\in\mathbb{R} we defined a cutoff operator

ECf(x,k)=n0c(η,x)<kc(i,η.x)+c(η,x)f(c(i,η.x),c(η,x)k)d𝐩(i)d𝐩n(η).E_{C}f(x,k)=\sum_{n\geq 0}\int_{c(\eta,x)<k\leq c(i,\eta.x)+c(\eta,x)}f(c(i,\eta.x),\,c(\eta,x)-k)d\mathbf{p}(i)d\mathbf{p}^{n}(\eta).

We can define a symbolic analogue via, for ω𝒜\omega\in\mathcal{A}^{\mathbb{N}},

ECf(ω,k)=n0c~(η,ω)<kc~(i,η.ω)+c~(η,ω)f(c~(i,η.ω),c~(η,ω)k)d𝐩(i)d𝐩n(η).E_{C}f(\omega,k)=\sum_{n\geq 0}\int_{\tilde{c}(\eta,\omega)<k\leq\tilde{c}(i,\eta.\omega)+\tilde{c}(\eta,\omega)}f(\tilde{c}(i,\eta.\omega),\,\tilde{c}(\eta,\omega)-k)d\mathbf{p}(i)d\mathbf{p}^{n}(\eta).

Then, by the definition of our cocycles cc (12) and c~\tilde{c} (31)

ECf(ω,k)=ECf(xω,k).E_{C}f(\omega,k)=E_{C}f(x_{\omega},k).

Thus, the conclusion of Proposition 4.6 applies to EC(ω,k)E_{C}(\omega,k) with the same error terms.

Now, let gC3()g\in C^{3}(\mathbb{R}). Let ρ\rho be a smooth cutoff function such that ρ|[0,D]=1\rho|_{[0,D^{\prime}]}=1, and such that it becomes 0 outside of [1,D+1][-1,D^{\prime}+1]. Let

f(y,x):=g(y+x)ρ(y)ρ(x+y).f(y,x):=g(y+x)\rho(y)\rho(x+y).

Then f(y,x)=g(x+y)f(y,x)=g(x+y) when y,x+y[0,D]y,x+y\in[0,D^{\prime}]. By definition,

𝔼(g(Sτk(η)(η)k)|στk(η)η=ω)=ECf(ω,k)=ECf(xω,k).\mathbb{E}\left(g\left(S_{\tau_{k}(\eta)}(\eta)-k\right)\big{|}\,\sigma^{\tau_{k}(\eta)}\eta=\omega\right)=E_{C}f(\omega,k)=E_{C}f(x_{\omega},k).

The function ff satisfies the conditions of Proposition 4.6, and so we can also apply this proposition to ECf(xω,k)E_{C}f(x_{\omega},k). The conclusion of Theorem 4.1 follows, with an error term of

eϵk2O|K|((supy,y[D,D]||fy(3)||L1+||fy||L1).e^{-\frac{\epsilon k}{2}}O_{|K|}(\left(\sup_{y,y^{\prime}\in[D,D^{\prime}]}||f_{y}^{(3)}||_{L^{1}}+||f_{y^{\prime}}||_{L^{1}}\right).

Since supp(ρ)[1,D+1]\text{supp}(\rho)\subseteq[-1,D^{\prime}+1], for every y[D,D]y\in[D,D^{\prime}] the function fyf_{y} is supported on [1D,1+D][-1-D,1+D^{\prime}]. So, we can take K=[D,D]×[1D,1+D]K=[D,D^{\prime}]\times[-1-D,1+D^{\prime}], and for all y[D,D]y\in[D,D^{\prime}]

fyL1=Oρ(maxx[1D,D+1]|g(x)|).||f_{y}||_{L^{1}}=O_{\rho}\left(\max_{x\in[-1-D,D^{\prime}+1]}\left|g(x)\right|\right).

Similarly, for all y[D,D]y^{\prime}\in[D,D^{\prime}]

fy(3)L1=Oρ(maxi=0,1,2,3maxx[1D,D+1]|g(i)(x)|).||f_{y^{\prime}}^{(3)}||_{L^{1}}=O_{\rho}\left(\max_{i=0,1,2,3}\max_{x\in[-1-D,D^{\prime}+1]}\left|g^{(i)}(x)\right|\right).

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 Φ\Phi is C2C^{2}, and that ν𝐩\nu_{\mathbf{p}} is a self-conformal measure such that 𝐩\mathbf{p} is a strictly positive probability vector. Recall that =𝐩\mathbb{P}=\mathbf{p}^{\mathbb{N}} is our Bernoulli measure on 𝒜\mathcal{A}^{\mathbb{N}}.

In this section we show that:

 There exists α>0 such that |q(ν)|O(1|q|α), as |q|.\text{ There exists }\alpha>0\text{ such that }\left|\mathcal{F}_{q}(\nu)\right|\leq O\left(\frac{1}{\left|q\right|^{\alpha}}\right),\text{ as }|q|\rightarrow\infty.

We will prove this under the additional (minor) assumption that Φ\Phi is orientation preserving; The general case is similar, but notationally heavier.

Let |q||q| be large and let k=k(q)log|q|k=k(q)\approx\log|q| to be chosen later. Let ϵ>0\epsilon>0 be as in Theorem 4.1. We define a stopping time βk:𝒜\beta_{k}:\mathcal{A}^{\mathbb{N}}\rightarrow\mathbb{N} by

βk(ω)=min{m:|fω|m(x0)|<ekϵk8}, where x0I is a prefixed point.\beta_{k}(\omega)=\min\{m:\,\left|f_{\omega|_{m}}^{\prime}(x_{0})\right|<e^{-k-\frac{\epsilon k}{8}}\},\,\text{ where }x_{0}\in I\text{ is a prefixed point}.

We require the following Theorem from [4], that relates the random variables SτkS_{\tau_{k}}, τk\tau_{k} (defined in Section 4), and βk\beta_{k}. For any tt\in\mathbb{R} let Mt:M_{t}:\mathbb{R}\rightarrow\mathbb{R} denote the scaling map Mt(x)=txM_{t}(x)=t\cdot x.

Theorem 5.1.

(Linearization) For any β(0,1)\beta\in(0,1),

|q(ν)|2|q(MeSτk(ω)fω|τk(ω)+1βk(ω)ν)|2𝑑(ω)+O(|q|e(k+kϵ8)βkϵ8).\left|\mathcal{F}_{q}(\nu)\right|^{2}\leq\int\left|\mathcal{F}_{q}(M_{e^{-S_{\tau_{k}(\omega)}}}\circ f_{\omega|_{\tau_{k}{(\omega)}+1}^{\beta_{k}(\omega)}}\nu)\right|^{2}d\mathbb{P}(\omega)+O\left(|q|e^{-(k+\frac{k\epsilon}{8})-\beta\cdot\frac{k\epsilon}{8}}\right).

Also, there exists a global constant C>1C^{\prime}>1 such that for every ω𝒜\omega\in\mathcal{A}^{\mathbb{N}}

|fω|τk(ω)+1βk(ω)(x)|=ΘC(eϵk8).\left|f_{\omega|_{\tau_{k}{(\omega)}+1}^{\beta_{k}(\omega)}}^{\prime}(x)\right|=\Theta_{C^{\prime}}\left(e^{-\frac{\epsilon k}{8}}\right). (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 Φ\Phi is orientation preserving. See [4, Corollary 4.6 and Remark 4.7] on how to remove this assumption.

Next, for every k>0k>0 we define a measurable partition 𝒫k\mathcal{P}_{k} of 𝒜\mathcal{A}^{\mathbb{N}} via the relation

ω𝒫kηστk(ω)ω=στk(η)η.\omega\sim_{\mathcal{P}_{k}}\eta\iff\sigma^{\tau_{k}(\omega)}\omega=\sigma^{\tau_{k}(\eta)}\eta.

Writing 𝔼𝒫k(ξ)()\mathbb{E}_{\mathcal{P}_{k}(\xi)}(\cdot) for the expectation with respect to the conditional measure of \mathbb{P} on a cell corresponding to a \mathbb{P}-typical ξ\xi, it follows from Theorem 5.1 and the law of total probability that: For any β(0,1)\beta\in(0,1),

|q(ν)|2𝔼𝒫k(ξ)(|q(MeSτk(ω)fξ|τk(ξ)+1βk(ξ)ν)|2)𝑑(ξ)+O(|q|e(k+kϵ8)βkϵ8).\left|\mathcal{F}_{q}(\nu)\right|^{2}\leq\int\mathbb{E}_{\mathcal{P}_{k}(\xi)}\left(\left|\mathcal{F}_{q}(M_{e^{-S_{\tau_{k}(\omega)}}}\circ f_{\xi|_{\tau_{k}{(\xi)}+1}^{\beta_{k}(\xi)}}\nu)\right|^{2}\right)d\mathbb{P}(\xi)+O\left(|q|e^{-(k+\frac{k\epsilon}{8})-\beta\cdot\frac{k\epsilon}{8}}\right). (36)

And, for every ξ𝒜\xi\in\mathcal{A}^{\mathbb{N}},

|fξ|τk(ξ)+1βk(ξ)(x)|=ΘC(eϵk8).\left|f_{\xi|_{\tau_{k}{(\xi)}+1}^{\beta_{k}(\xi)}}^{\prime}(x)\right|=\Theta_{C^{\prime}}\left(e^{-\frac{\epsilon k}{8}}\right).

Now, for every fixed ξ𝒜\xi\in\mathcal{A}^{\mathbb{N}} and k>0k>0, define the CωC^{\omega} function

gk,ξ(t)=|q(Me(tk)fξ|τk(ξ)+1βk(ξ)ν)|2.g_{k,\xi}(t)=\left|\mathcal{F}_{q}\left(M_{e^{(-t-k)}}\circ f_{\xi|_{\tau_{k}{(\xi)}+1}^{\beta_{k}(\xi)}}\nu\right)\right|^{2}.

Then by the definition of the local C3C^{3} norm on [1D,D+1][-1-D,D^{\prime}+1] as in (34), assuming |q|ek|q|\cdot e^{-k}\rightarrow\infty as qq\rightarrow\infty,

gk,ξC3O((|q|ek)3).||g_{k,\xi}||_{C^{3}}\leq O\left(\left(|q|\cdot e^{-k}\right)^{3}\right).

We emphasize that the bound above holds uniformly across ξ\xi. Applying Theorem 4.1 we obtain for a \mathbb{P}-typical ξ\xi

𝔼𝒫k(ξ)(|q(MeSτk(ω)fξ|τk(ξ)+1βk(ξ)ν)|2)\displaystyle\mathbb{E}_{\mathcal{P}_{k}(\xi)}\left(\left|\mathcal{F}_{q}(M_{e^{-S_{\tau_{k}(\omega)}}}\circ f_{\xi|_{\tau_{k}{(\xi)}+1}^{\beta_{k}(\xi)}}\nu)\right|^{2}\right) =\displaystyle= 𝔼𝒫k(ξ)(gk,ξ(Sτk(ω)k))\displaystyle\mathbb{E}_{\mathcal{P}_{k}(\xi)}\left(g_{k,\xi}\left(S_{\tau_{k}(\omega)}-k\right)\right)
=\displaystyle= 1χDDy0gk,ξ(x+y)𝑑x𝑑κ(y)+eϵk2O((|q|ek)3)\displaystyle\frac{1}{\chi}\int_{D}^{D^{\prime}}\int_{-y}^{0}g_{k,\xi}(x+y)\,dxd\kappa(y)+e^{-\frac{\epsilon k}{2}}O\left(\left(|q|\cdot e^{-k}\right)^{3}\right)
=\displaystyle= 1χDDy0gk,ξ(x+y)𝑑x𝑑κ(y)+eϵk2(|q|ek)3O(1).\displaystyle\frac{1}{\chi}\int_{D}^{D^{\prime}}\int_{-y}^{0}g_{k,\xi}(x+y)\,dxd\kappa(y)+e^{-\frac{\epsilon k}{2}}\left(|q|\cdot e^{-k}\right)^{3}O\left(1\right).

Plugging the above equality into (36), using the definition of gk,ξg_{k,\xi} and that if f0f\geq 0 then

1χDDy0f(x+y)𝑑x𝑑κ(y)1χ0Df(t)𝑑t,\frac{1}{\chi}\int_{D}^{D^{\prime}}\int_{-y}^{0}f(x+y)\,dxd\kappa(y)\leq\frac{1}{\chi}\int_{0}^{D^{\prime}}f(t)\,dt,

we obtain:

|q(ν)|2\displaystyle\left|\mathcal{F}_{q}(\nu)\right|^{2} \displaystyle\leq 𝔼𝒫k(ξ)(|q(MeSτk(ω)fξ|τk(ξ)+1βk(ξ)ν)|2)𝑑(ξ)+O(|q|e(k+kϵ8)βkϵ8)\displaystyle\int\mathbb{E}_{\mathcal{P}_{k}(\xi)}\left(\left|\mathcal{F}_{q}(M_{e^{-S_{\tau_{k}(\omega)}}}\circ f_{\xi|_{\tau_{k}{(\xi)}+1}^{\beta_{k}(\xi)}}\nu)\right|^{2}\right)d\mathbb{P}(\xi)+O\left(|q|e^{-(k+\frac{k\epsilon}{8})-\beta\cdot\frac{k\epsilon}{8}}\right)
=\displaystyle= 1χDDy0gk,ξ(x+y)𝑑x𝑑κ(y)𝑑(ξ)+O(|q|e(k+kϵ8)βkϵ8)+eϵk2(|q|ek)3O(1)\displaystyle\frac{1}{\chi}\int\int_{D}^{D^{\prime}}\int_{-y}^{0}g_{k,\xi}(x+y)\,dxd\kappa(y)d\mathbb{P}(\xi)+O\left(|q|e^{-(k+\frac{k\epsilon}{8})-\beta\cdot\frac{k\epsilon}{8}}\right)+e^{-\frac{\epsilon k}{2}}\left(|q|\cdot e^{-k}\right)^{3}O\left(1\right)
\displaystyle\leq 1χ0D|q(Me(tk)fξ|τk(ξ)+1βk(ξ)ν)|2𝑑t𝑑(ξ)+O(|q|e(k+kϵ8)βkϵ8)\displaystyle\frac{1}{\chi}\int\int_{0}^{D^{\prime}}\left|\mathcal{F}_{q}\left(M_{e^{(-t-k)}}\circ f_{\xi|_{\tau_{k}{(\xi)}+1}^{\beta_{k}(\xi)}}\nu\right)\right|^{2}dtd\mathbb{P}(\xi)+O\left(|q|e^{-(k+\frac{k\epsilon}{8})-\beta\cdot\frac{k\epsilon}{8}}\right)
+eϵk2(|q|ek)3O(1).\displaystyle+e^{-\frac{\epsilon k}{2}}\left(|q|\cdot e^{-k}\right)^{3}O\left(1\right).

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 ξ𝒜\xi\in\mathcal{A}^{\mathbb{N}}, k>0k>0, and r>0r>0

0D|q(Me(tk)fξ|τk(ξ)+1βk(ξ)ν)|2𝑑x=O(1r|q|e(k+ϵk8)+supyν(Br(y)))\int_{0}^{D^{\prime}}\left|\mathcal{F}_{q}\left(M_{e^{(-t-k)}}\circ f_{\xi|_{\tau_{k}{(\xi)}+1}^{\beta_{k}(\xi)}}\nu\right)\right|^{2}dx=O\left(\frac{1}{r|q|e^{-(k+\frac{\epsilon k}{8})}}+\sup_{y}\nu(B_{r}(y))\right)

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 |q(ν)|2\left|\mathcal{F}_{q}(\nu)\right|^{2} by the sum of the following terms. Every term is bounded with implicit dependence on 𝐩\mathbf{p} and the underlying IFS. For simplicity, we ignore global multiplicative constants so we omit the big-OO notation:

Linearization: For any prefixed β(0,1)\beta\in(0,1),

|q|e(k+kϵ8)βkϵ8;|q|e^{-(k+\frac{k\epsilon}{8})-\beta\cdot\frac{k\epsilon}{8}};

Equidistribution:

eϵk2(|q|ek)3;e^{-\frac{\epsilon k}{2}}\left(|q|\cdot e^{-k}\right)^{3};

Oscillatory integral: For every r>0r>0

1r|q|e(k+ϵk8)+supyν(Br(y)).\frac{1}{r|q|e^{-(k+\frac{\epsilon k}{8})}}+\sup_{y}\nu(B_{r}(y)).

Choice of parameters For |q||q| we choose k=k(q)k=k(q) that satisfies

|q|=ek+kϵ7.|q|=e^{k+\frac{k\epsilon}{7}}.

We also choose r=ekϵ100r=e^{-\frac{k\epsilon}{100}} and β=12\beta=\frac{1}{2}. Then we get:

Linearization:

|q|e(k+kϵ8)βkϵ8=ekϵ7kϵ8kϵ16, this decay exponentially fast in k.|q|e^{-(k+\frac{k\epsilon}{8})-\beta\cdot\frac{k\epsilon}{8}}=e^{\frac{k\epsilon}{7}-\frac{k\epsilon}{8}-\frac{k\epsilon}{16}},\,\text{ this decay exponentially fast in }k.

Equidistribution:

eϵk2(|q|ek)3=eϵk2+3kϵ7, this decay exponentially fast in k.e^{-\frac{\epsilon k}{2}}\left(|q|\cdot e^{-k}\right)^{3}=e^{-\frac{\epsilon k}{2}+\frac{3k\epsilon}{7}},\,\text{ this decay exponentially fast in }k.

Oscillatory integral: There is some d=d(ν)>0d=d(\nu)>0 such that

1r|q|e(k+ϵk8)+supyν(Br(y))1ekϵ100+kϵ7kϵ8+edϵk100, this decay exponentially fast in k.\frac{1}{r|q|e^{-(k+\frac{\epsilon k}{8})}}+\sup_{y}\nu(B_{r}(y))\leq\frac{1}{e^{-\frac{k\epsilon}{100}+\frac{k\epsilon}{7}-\frac{k\epsilon}{8}}}+e^{-\frac{d\epsilon k}{100}},\,\text{ this decay exponentially fast in }k.

Here we made use of [25, Proposition 2.2], where it is shown that there is some C>0C>0 such that for every r>0r>0 small enough supyν(Br(y))Crd\sup_{y}\nu(B_{r}(y))\leq Cr^{d}.

Finally, summing these error terms, we see that for some α>0\alpha>0 we have |q(ν)|=O(ekα)\left|\mathcal{F}_{q}(\nu)\right|=O\left(e^{-k\alpha}\right). Since as |q||q|\rightarrow\infty we have kC0log|q|k\geq C_{0}\cdot\log|q| for some uniform C0>0C_{0}>0, our claim follows. \Box

6 On the proof of Corollary 1.2

In this Section we prove Corollary 1.2: All self-conformal measures with respect to a Cω()C^{\omega}(\mathbb{R}) IFS that contains a non-affine map have polynomial Fourier decay: First, we show that any given Cω()C^{\omega}(\mathbb{R}) IFS is either not C2C^{2} conjugate to linear, or it is CωC^{\omega} 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 Φ\Phi, a C2()C^{2}(\mathbb{R}) IFS, is called linear if the following holds:

f′′(x)=0 for all xKΦ and fΦ.f^{\prime\prime}(x)=0\,\text{ for all }x\in K_{\Phi}\text{ and }f\in\Phi.

In particular, if Φ\Phi is Cω()C^{\omega}(\mathbb{R}) and linear then it is self-similar. Recall that an IFS Ψ\Psi is called C2C^{2} conjugate to Φ\Phi if there is a C2C^{2} diffeomorphsim hh between neighbourhoods of the corresponding attractors such that

Ψ=hΦh1:={hgh1}gΦ.\Psi=h\circ\Phi\circ h^{-1}:=\{h\circ g\circ h^{-1}\}_{g\in\Phi}.

The geometric properties of linear non self-similar Cr()C^{r}(\mathbb{R}) smooth IFSs when rωr\neq\omega 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 Φ\Phi be a Cω([0,1])C^{\omega}([0,1]) IFS. Then there is a dichotomy:

  1. 1.

    Φ\Phi is not C2C^{2} conjugate to linear, or

  2. 2.

    Φ\Phi is conjugate to a linear Cω()C^{\omega}(\mathbb{R}) IFS via an analytic map gg, that is a diffeomorphism on [0,1][0,1]. In particular, Φ\Phi is CωC^{\omega} 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 C2C^{2} version of the Claim holds if KΦK_{\Phi} 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 gCω([0,1])g\in C^{\omega}([0,1]) be a contracting map. Then there exists some non-trivial interval J[0,1]J\subseteq[0,1] and a diffeomorphism hCω([0,1],J)h\in C^{\omega}([0,1],J) such that hgh1h\circ g\circ h^{-1} is affine.

We also require the following Lemma:

Lemma 6.3.

Let Φ\Phi be a C2()C^{2}(\mathbb{R}) IFS that is C2C^{2} conjugate to a linear IFS. If there exists gΦg\in\Phi such that g′′=0g^{\prime\prime}=0 on its attractor KΦK_{\Phi}, then Φ\Phi is already linear; That is, for every fΦf\in\Phi, f′′=0f^{\prime\prime}=0 on KΦK_{\Phi}.

Proof.

Let hC2()h\in C^{2}(\mathbb{R}) be such that the IFS hΦh1h\circ\Phi\circ h^{-1} is linear. Let gΦg\in\Phi be such that g′′=0g^{\prime\prime}=0 on K:=KΦK:=K_{\Phi}. We first show that this implies that h′′h^{\prime\prime} vanishes on KK: By our assumption, for all zh(K)z\in h(K) we have

(hgh1)′′(z)=0.\left(h\circ g\circ h^{-1}\right)^{\prime\prime}(z)=0.

Writing h1(z)=yKh^{-1}(z)=y\in K, we compute:

log(hgh1)(z)=loghg(y)+logg(y)logh(y).\log\left(h\circ g\circ h^{-1}\right)^{\prime}(z)=\log h^{\prime}\circ g(y)+\log g^{\prime}(y)-\log h^{\prime}(y).

Combining the two previous displayed equations, it follow that

0=(hgh1)′′(z)(hgh1)(z)=ddzlog(hgh1)(z)=h′′g(y)g(y)hg(y)+g′′(y)g(y)h′′(y)h(y).0=\frac{\left(h\circ g\circ h^{-1}\right)^{\prime\prime}(z)}{\left(h\circ g\circ h^{-1}\right)^{\prime}(z)}=\frac{d}{dz}\log\left(h\circ g\circ h^{-1}\right)^{\prime}(z)=\frac{h^{\prime\prime}\circ g(y)\cdot g^{\prime}(y)}{h^{\prime}\circ g(y)}+\frac{g^{\prime\prime}(y)}{g^{\prime}(y)}-\frac{h^{\prime\prime}(y)}{h^{\prime}(y)}.

By our assumption g′′(y)=0g^{\prime\prime}(y)=0 since yKy\in K. We conclude that

h′′g(y)g(y)hg(y)=h′′(y)h(y).\frac{h^{\prime\prime}\circ g(y)\cdot g^{\prime}(y)}{h^{\prime}\circ g(y)}=\frac{h^{\prime\prime}(y)}{h^{\prime}(y)}. (37)

Finally, let xKx\in K. Since g′′(y)=0g^{\prime\prime}(y)=0 for all yKy\in K and for every kk\in\mathbb{N} we have for the kk-fold composition gk(x)Kg^{\circ k}(x)\in K, then for all kk\in\mathbb{N}, (gk)′′(x)=0\left(g^{\circ k}\right)^{\prime\prime}(x)=0. A similar argument shows that for every kk the IFS hΦkh1h\circ\Phi^{k}\circ h^{-1} is linear. It follows that in (37) we can substitute gkg^{\circ k} for gg. As the IFS Φ\Phi 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 h′′(y)=0h^{\prime\prime}(y)=0. We conclude that h′′h^{\prime\prime} vanishes on KK.

Finally, let fΦf\in\Phi. Then for all zh(K)z\in h(K) we have, for y=h1(z)Ky=h^{-1}(z)\in K

0=ddxlog(hfh1)(z)=h′′f(y)f(y)hf(y)+f′′(y)f(y)h′′(y)h(y).0=\frac{d}{dx}\log\left(h\circ f\circ h^{-1}\right)^{\prime}(z)=\frac{h^{\prime\prime}\circ f(y)\cdot f^{\prime}(y)}{h^{\prime}\circ f(y)}+\frac{f^{\prime\prime}(y)}{f^{\prime}(y)}-\frac{h^{\prime\prime}(y)}{h^{\prime}(y)}.

Since yKy\in K then f(y)Kf(y)\in K and so by the previous paragraph h′′f(y)=0h^{\prime\prime}\circ f(y)=0 and h′′(y)=0h^{\prime\prime}(y)=0. As |f|>0|f^{\prime}|>0 on [0,1][0,1] by assumption, this is only possible if f′′(y)=0f^{\prime\prime}(y)=0. It follows that f′′f^{\prime\prime} vanishes on KΦK_{\Phi}, as claimed. ∎

Proof of Claim 6.1 Suppose ΦCω()\Phi\in C^{\omega}(\mathbb{R}) is C2C^{2} conjugate to linear. Let gΦg\in\Phi be any map. By Proposition 6.2, there is some non-trivial interval J[0,1]J\subseteq[0,1] and a map hCω([0,1],J)h\in C^{\omega}([0,1],J) such that hgh1h\circ g\circ h^{-1} is affine. Then the IFS hΦh1h\circ\Phi\circ h^{-1} is CωC^{\omega} and contains an affine map. Since Φ\Phi is C2C^{2} conjugate to linear, so is hΦh1h\circ\Phi\circ h^{-1}. 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. \Box

6.2 Proof of Corollary 1.2

We now prove Corollary 1.2. Let Ψ\Psi be a Cω()C^{\omega}(\mathbb{R}) IFS, and assume Ψ\Psi contains a non-affine map. Recall that we are always assuming KΨK_{\Psi} is infinite. By Claim 6.1 there are two cases to consider:

The first alternative is that Ψ\Psi is not C2C^{2} conjugate to linear. Then, by Theorem 1.1, every self-conformal measure ν\nu admits some α>0\alpha>0 such that

|q(ν)|=O(1|q|α).\left|\mathcal{F}_{q}\left(\nu\right)\right|=O\left(\frac{1}{|q|^{\alpha}}\right).

The second alternative is that Ψ\Psi is CωC^{\omega} conjugate to a self-similar IFS Φ\Phi. Let gg denote the conjugating map. We have the following easy Lemma:

Lemma 6.4.

The analytic map gg is not affine.

Proof.

Suppose towards a contradiction that gg is affine. Recall that gg is a conjugating map between Ψ\Psi, a CωC^{\omega} IFS, and a self-similar IFS. So, both IFS’s in question are in fact self-similar. However, our standing assumption is that contains Ψ\Psi contains a non affine map. This is a contradiction. ∎

So, in the second alternative, Ψ\Psi is conjugate to a self-similar IFS Φ\Phi via a CωC^{\omega} map gg that is not affine. In particular,

|{x[0,1]:g′′(x)=0}|<.\left|\{x\in[0,1]:g^{\prime\prime}(x)=0\}\right|<\infty.

Since every self-conformal measure with respect to Ψ\Psi can be written as gμg\mu where μ\mu is a self-similar measure with respect to Φ\Phi, 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 μ\mu be a non-atomic self-similar measure with respect to Φ\Phi, and let gCω([0,1])g\in C^{\omega}([0,1]) be such that g0g^{\prime}\neq 0, and such that g′′0g^{\prime\prime}\neq 0 except for possibly finitely many points in [0,1][0,1]. Then there exists some α>0\alpha>0 such that

|q(gμ)|=O(1|q|α).\left|\mathcal{F}_{q}\left(g\mu\right)\right|=O\left(\frac{1}{|q|^{\alpha}}\right).

The proof of Corollary 1.2 is complete. \Box

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 C1+αC^{1+\alpha} 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 SL(2,)\text{SL}(2,\mathbb{R})-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 d\mathbb{R}^{d}. 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. LqL^{q} 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 M0M_{0}-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 d\mathbb{R}^{d}. 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 SL(2,)\text{SL}(2,\mathbb{R}) 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]