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

Families of infinite parabolic IFS with overlaps: the approximating method

Liangang Ma Dept. of Mathematical Sciences, Binzhou University, Huanghe 5th Road No. 391, Binzhou 256600, Shandong, P. R. China [email protected]
Abstract.

This work is devoted to the study of families of infinite parabolic iterated function systems (PIFS) on a closed interval parametrized by vectors in d\mathbb{R}^{d} with overlaps. We show that the Hausdorff dimension and absolute continuity of ergodic projections through a family of infinite PIFS are decided a.e. by the growth rate of the entropy and Lyapunov exponents of the families of truncated PIFS with respect to the concentrating measures, under transversality of the family essentially. We also give an estimation on the upper bound of the Hausdorff dimension of parameters where the corresponding ergodic projections admit certain dimension drop. The setwise topology on the space of finite measures enables us to approximate a family of infinite systems by families of their finite truncated sub-systems, which plays the key role throughout our work.

The work is supported by ZR2019QA003 from SPNSF and 12001056 from NSFC

1. Introduction

This work should be understood in the background of families of overlapping iterated function systems. A typical family of overlapping iterated function system (IFS) on a compact metric space consists of a flow of finitely many strictly contractive endomorphisms on the space parametrized by some (time) parameter. The dimension of attractors and measures supported on them (especially the invariant or ergodic ones) is the focus in the theory of families of IFS. Besides calculating the dimension of attractors and measures supported on the attractors, another important problem is to decide whether the projective measures on the attractors from measures on the symbolic spaces are singular or absolutely continuous with respect to the Lebesgue measure, refer to [BRS, Hoc1, Hoc3, PSS, Shm2, Shm3, SS1, SSS]. See also [Fur] by Furstenberg, [Hoc4] by Hochman, [PS1] by Peres-Solomyak and [Shm4] by Shmerkin for more interesting questions on families of overlapping IFS. These problems are already very difficult in the case of Bernoulli convolutions, see for example [BV1, BV2, Hoc1, LPS, Shm1, Sol3, SS2, Var1, Var2, Var3] for recent progress on the topic.

Note that most of the above research are done on families of finite IFS, that is, every individual IFS in the family is constituted by finitely many maps. In this work we try to deal with families of infinite IFS. We choose the families of parabolic iterated function systems investigated by K. Simon, B. Solomyak and M. Urbański in [SSU1, SSU2] to demonstrate our ideas, which are probably applicable to some other families of infinite IFS. Our technique here is to utilize the sequence of concentrating measures on the attractors of families of finite sub-systems to approximate the measures on the attractors of families of infinite systems.

Let XX\subset\mathbb{R} be a closed interval with its Borel σ\sigma-algebra \mathcal{B}. Let ^(X)\mathcal{\hat{M}}(X) be the collection of all the finite Borel measures on (X,)(X,\mathcal{B}). For a set AdA\subset\mathbb{R}^{d} with dd\in\mathbb{N}, let HD(A)HD(A) be its Hausdorff dimension [Fal2].

1.1 Definition.

For a measure ν^(X)\nu\in\mathcal{\hat{M}}(X), the lower and upper Hausdorff dimension of ν\nu are defined respectively to be:

dimν=inf{HD(A):ν(A)>0,A},dim_{*}\nu=\inf\{HD(A):\nu(A)>0,A\in\mathcal{B}\},

and

dimν=inf{HD(A):ν(A)=ν(X),A}.dim^{*}\nu=\inf\{HD(A):\nu(A)=\nu(X),A\in\mathcal{B}\}.

The two values indicate the distribution of a measure on XX from two complementary points of views, with the possibility of the lower one being strictly less than the upper one. However, they coincide with each other for ergodic measures with respect to transformations of proper regularity (for example the C1C^{1} diffeomorphisms) on XX. See for example [HS, SimS2, You] on calculation of the Hausdorff dimension of measures in various circumstances. Following P. Mattila, M. Morán and J. M. Rey [MMR], the author discovers some semi-continuity of the measure-dimension mappings dimdim_{*} and dimdim^{*} on ^(X)\mathcal{\hat{M}}(X) equipped with the setwise topology [FKZ, Las], which is crucial in our approximating process.

Now we gradually introduce the families of parabolic iterated function systems. For ϑ(0,1]\vartheta\in(0,1], a point vXv\in X is called an indifferent point of a C1+ϑC^{1+\vartheta} map s:XXs:X\rightarrow X if

|s(v)|=1|s^{\prime}(v)|=1.

1.2 Definition.

A C1+ϑC^{1+\vartheta} map s:XXs:X\rightarrow X is called parabolic it satisfies the following conditions:

  • ss has only one indifferent point vv and it is fixed, that is, s(v)=vs(v)=v.

  • ss is contractive on X{v}X\setminus\{v\}, that is, 0<|s(x)|<10<|s^{\prime}(x)|<1 for any xX{v}x\in X\setminus\{v\}.

  • ss is well-behaved around vv, that is, s(x)s^{\prime}(x) is monotone on each component of X{v}X\setminus\{v\}.

  • There exist L11L_{1}\geq 1 and β<ϑ1ϑ\beta<\cfrac{\vartheta}{1-\vartheta} such that

    1L1lim infxv|s(x)s(v)||xv|βlim supxv|s(x)s(v)||xv|βL1\cfrac{1}{L_{1}}\leq\liminf_{x\rightarrow v}\cfrac{|s^{\prime}(x)-s^{\prime}(v)|}{|x-v|^{\beta}}\leq\limsup_{x\rightarrow v}\cfrac{|s^{\prime}(x)-s^{\prime}(v)|}{|x-v|^{\beta}}\leq L_{1}.

It is called hyperbolic if 0<|s(x)|<10<|s^{\prime}(x)|<1 for any xXx\in X. Parabolic maps with finitely many indifferent points can be handled in our context, with some technical modifications on the proofs.

1.3 Definition.

For ϑ(0,1]\vartheta\in(0,1] and a countable index set II (with at least two elements), a parabolic iterated function system (PIFS) is a collection of C1+ϑC^{1+\vartheta} maps

S={si:XX}iIS=\{s_{i}:X\rightarrow X\}_{i\in I}

satisfying the following conditions:

  • There is one and only one i1Ii_{1}\in I such that the map si1s_{i_{1}} is a parabolic map with its indifferent point vv.

  • The other maps {si}iI{i1}\{s_{i}\}_{i\in I\setminus\{i_{1}\}} are all hyperbolic.

The collection of PIFS S={si:XX}iIS=\{s_{i}:X\rightarrow X\}_{i\in I} satisfying

iI{i1}si(X)Xo{v}\cup_{i\in I\setminus\{i_{1}\}}s_{i}(X)\subset X^{o}\setminus\{v\}

is termed ΓX(ϑ)\Gamma_{X}(\vartheta). For an integer dd\in\mathbb{N} and an open set UdU\subset\mathbb{R}^{d}, we focus on families of PIFS instead of a single system in ΓX(ϑ)\Gamma_{X}(\vartheta) parametrized by the vector-time parameter 𝕥U\mathbb{t}\in U. The parabolic map is always fixed at any time in the family. Since families of finite PIFS (that is, #I<\#I<\infty) has been considered by Simon-Solomyak-Urbański, we focus on the families of infinite PIFS (#I=\#I=\infty) in the following. That is, we focus on families of PIFS

S𝕥={si𝕥:XX}iS^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}},

such that S𝕥ΓX(ϑ)S^{\mathbb{t}}\in\Gamma_{X}(\vartheta) with the parabolic map s1𝕥=s1s_{1}^{\mathbb{t}}=s_{1} remaining the same at any time 𝕥U\mathbb{t}\in U. Let

π𝕥:IX\pi_{\mathbb{t}}:I^{\infty}\rightarrow X

be the projection map and J𝕥=π𝕥(I)J_{\mathbb{t}}=\pi_{\mathbb{t}}(I^{\infty}) be the attractor ([Hut]) at time 𝕥U\mathbb{t}\in U. To achieve some reasonable conclusions on the parametrized family, obviously some control on the dependence of the family with respect to 𝕥\mathbb{t} is necessary. A family of PIFS S𝕥={si𝕥:XX}iS^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}} is said to satisfy the continuity condition if for any fixed ii\in\mathbb{N}, the map si𝕥s_{i}^{\mathbb{t}} depends continuously on the parameter 𝕥U\mathbb{t}\in U in the Banach space C1+ϑC^{1+\vartheta} equipped with the supreme norm. Let 𝔏d\mathfrak{L}^{d} be the Lebesgue measure on d\mathbb{R}^{d} for dd\in\mathbb{N}. The family is said to satisfy the transversality condition if there exists a constant C1C_{1} such that

𝔏d{𝕥U:|π𝕥(ω)π𝕥(τ)|r}C1r\mathfrak{L}^{d}\{\mathbb{t}\in U:|\pi_{\mathbb{t}}(\omega)-\pi_{\mathbb{t}}(\tau)|\leq r\}\leq C_{1}r

for any ω,τ,ω1τ1\omega,\tau\in\mathbb{N}^{\infty},\omega_{1}\neq\tau_{1} and any r>0r>0.

While the continuity condition comes more naturally, the transversality condition has been recognized to be an effective condition to achieve some uniform conclusions in various contexts of flows of dynamical systems. For example, see [Fal1, PolS, PS2, PS3, SimS1, Sol1, Sol2].

The following are some classical notions on the symbolic dynamics on \mathbb{N}^{\infty}. For an infinite word ω=ω1ω2\omega=\omega_{1}\omega_{2}\cdots\in\mathbb{N}^{\infty}, let

σ(ω)=ω2ω3\sigma(\omega)=\omega_{2}\omega_{3}\cdots

be the shift map. For a finite kk-word τk\tau\in\mathbb{N}^{k}, let

[τ]={ω:ω|k=τ}[\tau]=\{\omega\in\mathbb{N}^{\infty}:\omega|_{k}=\tau\}

be a cylinder set, in which ω|k=ω1ω2ωk\omega|_{k}=\omega_{1}\omega_{2}\cdots\omega_{k} is the kk-th restriction of the infinite word ω\omega. Let

=k=1k\mathbb{N}^{*}=\cup_{k=1}^{\infty}\mathbb{N}^{k}

be the collection of all the finite words. Now consider the σ\sigma-algebra \mathcal{B}_{\mathbb{N}^{\infty}} generated by all the cylinder sets on \mathbb{N}^{\infty}. Let hμ(σ)h_{\mu}(\sigma) be the entropy of the shift map σ\sigma with respect to the partition of \mathbb{N}^{\infty} by cylinder sets {[i]}i=1\{[i]\}_{i=1}^{\infty} for a measure μ^()\mu\in\mathcal{\hat{M}}(\mathbb{N}^{\infty}) ([Hoc2, Wal]).

One can treat μ\mu as the law of an infinite discrete-time stochastic process

𝕐={Yi}i\mathbb{Y}=\{Y_{i}\}_{i\in\mathbb{N}},

in which YiY_{i} is a discrete random variable on \mathbb{N} for ii\in\mathbb{N} under the natural projection with law μ(×××Yi××)\mu(\mathbb{N}\times\mathbb{N}\times\cdots\times Y_{i}\times\mathbb{N}\times\cdots). The sequence of infinite random variables is called independent if the finite variables

{Y1,Y2,,Yn}\{Y_{1},Y_{2},\cdots,Y_{n}\}

are independent for any nn\in\mathbb{N}, see [Tsi, Definition 3a(1),(b)]. The stochastic process 𝕐\mathbb{Y} is a Markov process under the hypothesis of independence, refer to [BHP, FKZ, Hai, HL].

In the following we focus on the measures on J𝕥J_{\mathbb{t}} projected from probability measures on \mathbb{N}^{\infty} through π𝕥\pi_{\mathbb{t}} for 𝕥U\mathbb{t}\in U. For a measure μ\mu on the symbolic space \mathbb{N}^{\infty}, consider its projection under π𝕥\pi_{\mathbb{t}}:

ν𝕥=μπ𝕥1\nu_{\mathbb{t}}=\mu\circ\pi_{\mathbb{t}}^{-1}

for any 𝕥U\mathbb{t}\in U. We are particularly interested in the cases when μ\mu is invariant or ergodic with respect to the shift map σ\sigma on \mathbb{N}^{\infty}. If μ\mu is invariant then ν𝕥\nu_{\mathbb{t}} is of pure type, see for example [JW]. Let

λμ𝕥(σ)=log|sω1(π𝕥σ(ω))|dμ(ω)\lambda^{\mathbb{t}}_{\mu}(\sigma)=-\int_{\mathbb{N}^{\infty}}\log|s_{\omega_{1}}^{\prime}(\pi_{\mathbb{t}}\circ\sigma(\omega))|d\mu(\omega)

be the average Lyapunov exponent of the family of PIFS S𝕥={si𝕥:XX}iS^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}} at time 𝕥U\mathbb{t}\in U.

Our first main result deals with the two basic questions regarding the projective measures {ν𝕥}𝕥U\{\nu_{\mathbb{t}}\}_{\mathbb{t}\in U}, these are, deciding their dimensions dimν𝕥,dimν𝕥dim_{*}\nu_{\mathbb{t}},dim^{*}\nu_{\mathbb{t}} and whether they are absolutely continuous or singular with respect to the Lebesgue measure for 𝕥U\mathbb{t}\in U. Let n={1,2,,n}\mathbb{N}_{n}=\{1,2,\cdots,n\} be the nn-th truncation of \mathbb{N} for nn\in\mathbb{N}. We leave the technical concept-the nn-th concentrating measure μn\mu_{n} (supported on n\mathbb{N}_{n}^{\infty}) of a finite measure μ\mu (supported on \mathbb{N}^{\infty}) to Section 4. The following result generalizes [SSU1, Theorem 2.3] from families of finite PIFS to families of infinite PIFS.

1.4 Theorem.

Let

{S𝕥={si𝕥:XX}iΓX(ϑ)}𝕥U\big{\{}S^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}}\in\Gamma_{X}(\vartheta)\big{\}}_{\mathbb{t}\in U}

be a family of infinite parabolic iterated function systems satisfying the continuity and transversality condition with respect to the vector-time parameter 𝕥U\mathbb{t}\in U. For an ergodic probability measure μ\mu on the symbolic space \mathbb{N}^{\infty} with positive entropy hμ(σ)h_{\mu}(\sigma), let μn\mu_{n} be its nn-th concentrating measure for any nn\in\mathbb{N}. If the sequence of infinite random variables in 𝕐\mathbb{Y} under the law μ\mu is independent and μ(n=1n)=1\mu(\cup_{n=1}^{\infty}\mathbb{N}_{n}^{\infty})=1, then

  1. (i).

    For Lebesgue a.e. 𝕥U\mathbb{t}\in U, dimν𝕥=dimν𝕥=limnmin{hμn(σ)λμn𝕥(σ),1}dim_{*}\nu_{\mathbb{t}}=dim^{*}\nu_{\mathbb{t}}=\lim_{n\rightarrow\infty}\min\Big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)},1\Big{\}},

  2. (ii).

    ν𝕥\nu_{\mathbb{t}} is absolutely continuous for Lebesgue a.e. 𝕥{𝕥:lim supn{hμn(σ)λμn𝕥(σ)}>1}\mathbb{t}\in\Big{\{}\mathbb{t}:\limsup_{n\rightarrow\infty}\big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)}\big{\}}>1\Big{\}},

in which ν𝕥=μπ𝕥1\nu_{\mathbb{t}}=\mu\circ\pi_{\mathbb{t}}^{-1} is the projective measure at time 𝕥U\mathbb{t}\in U.

1.5 Remark.

We are not sure whether the the hypothesis of independence of the sequence of infinite random variables under the law μ\mu can be removed or not, although it seems so alluring to us.

1.6 Remark.

As to the approximation of μ\mu, better results are possible if one does the approximation by the sequence of restrictions {μ|n}n=1\{\mu|_{\mathbb{N}_{n}^{\infty}}\}_{n=1}^{\infty} instead of the sequence of concentrating measures {μn}n=1\{\mu_{n}\}_{n=1}^{\infty}. However, the restrictions are not probabilities, while the concentrating measures take the advantage that they are all probability measures.

1.7 Remark.

Similar to [SSU1, Theorem 2.3], we can only guarantee the result holds a.e. instead of everywhere in Theorem 1.4, as it is still impossible to rule out the exact overlapping between images of {sω𝕥}ω\{s^{\mathbb{t}}_{\omega}\}_{\omega\in\mathbb{N}^{*}} for us for fixed 𝕥U\mathbb{t}\in U. There is an important conjecture that a dimension drop implies exact overlapping, which has been confirmed in the case of finite Bernoulli convolutions, see for example [Hoc1, Var3].

Note that the proof of [SSU1, Theorem 2.3] does not apply to Theorem 1.4. There are two main obstacles on application of Simon-Solomyak-Urbański’s techniques to the families of infinite PIFS. The first one is that hμ(σ)h_{\mu}(\sigma) can explode in case of #I=\#I=\infty, see [Hoc2]. If hμ(σ)=h_{\mu}(\sigma)=\infty for some ergodic μ^()\mu\in\mathcal{\hat{M}}(\mathbb{N}^{\infty}) with respect to the shift map, the Shannon-McMillan-Breiman Theorem is not known to hold in this case [Hay]. The second one is the regularity of the families of infinite PIFS. The collection of PIFS S={si:XX}iIS=\{s_{i}:X\rightarrow X\}_{i\in I} satisfying the following conditions is called ΓX(ϑ,V,γ,u,M)\Gamma_{X}(\vartheta,V,\gamma,u,M) in [SSU1].

  1. (a)

    There exists an open connected neighbourhood VV of vv, such that

    iI{i1}si(X)V=\cup_{i\in I\setminus\{i_{1}\}}s_{i}(X)\cap V=\emptyset.

  2. (b)

    There exists some γ(0,1)\gamma\in(0,1), such that

    supiI{i1}{si}γ\sup_{i\in I\setminus\{i_{1}\}}\{\|s_{i}^{\prime}\|\}\leq\gamma.

  3. (c)

    There exists some u(0,1)u\in(0,1), such that

    infiI,xX{|si(x)|}u\inf_{i\in I,x\in X}\{|s_{i}^{\prime}(x)|\}\geq u.

  4. (d)

    There exists M>0M>0 such that

    Sϑ=supiIsupx,yX{|si(x)si(y)||xy|ϑ}M\|S^{\prime}\|_{\vartheta}=\sup_{i\in I}\sup_{x,y\in X}\Big{\{}\cfrac{|s_{i}^{\prime}(x)-s_{i}^{\prime}(y)|}{|x-y|^{\vartheta}}\Big{\}}\leq M.

For a finite PIFS, SΓX(ϑ)S\in\Gamma_{X}(\vartheta) is enough to guarantee SΓX(ϑ,V,γ,u,M)S\in\Gamma_{X}(\vartheta,V,\gamma,u,M) for some parameters V,γ,u,MV,\gamma,u,M. However, this is not true for an infinite PIFS. Our technique of approximating measures on the infinite symbolic space by the sequence of concentrating measures on its finite subspaces successfully overcomes these obstacles.

Our next main result deals with local dimension of the exceptional parameters of a family of infinite PIFS {S𝕥}𝕥U\{S^{\mathbb{t}}\}_{\mathbb{t}\in U}. Let μ\mu be a measure on the symbolic space \mathbb{N}^{\infty} with its concentrating measures {μn}n\{\mu_{n}\}_{n\in\mathbb{N}}. For a subset GUG\subset U and 0<α<10<\alpha<1, in case the sequence {hμn(σ)λμn𝕥(σ)}n=1\Big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)}\Big{\}}_{n=1}^{\infty} admits a limit for any 𝕥G\mathbb{t}\in G, let

Eα,G:={𝕥G:dimν𝕥<limnmin{hμn(σ)λμn𝕥(σ),α}}E_{\alpha,G}:=\Big{\{}\mathbb{t}\in G:dim^{*}\nu_{\mathbb{t}}<\lim_{n\rightarrow\infty}\min\big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)},\alpha\big{\}}\Big{\}}

be the level-α\alpha exceptional parameters in GG. Let

Kα,G=min{sup𝕥Glimnhμn(σ)λμn𝕥(σ),α}+d1K_{\alpha,G}=\min\Big{\{}\sup_{\mathbb{t}\in G}\lim_{n\rightarrow\infty}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)},\alpha\Big{\}}+d-1.

For a parameter set AUdA\subset U\subset\mathbb{R}^{d}, let Nr(A)N_{r}(A) be the least number of open balls of radius r>0r>0 needed to cover the set AA. A family of PIFS S𝕥={si𝕥:XX}iS^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}} is said to satisfy the strong transversality condition if there exists a constant C2>0C_{2}>0 such that

Nr{𝕥U:|π𝕥(ω)π𝕥(τ)|r}C2r1dN_{r}\{\mathbb{t}\in U:|\pi_{\mathbb{t}}(\omega)-\pi_{\mathbb{t}}(\tau)|\leq r\}\leq C_{2}r^{1-d}

for any ω,τ,ω1τ1\omega,\tau\in\mathbb{N}^{\infty},\omega_{1}\neq\tau_{1} and any r>0r>0. Obviously this condition is stronger than the transversality condition. We have the following estimation on the upper bound of the Hausdorff dimension of the local level-α\alpha exceptional set.

1.8 Theorem.

Let

{S𝕥={si𝕥:XX}iΓX(ϑ)}𝕥U\big{\{}S^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}}\in\Gamma_{X}(\vartheta)\big{\}}_{\mathbb{t}\in U}

be a family of infinite parabolic iterated function systems satisfying the continuity and strong transversality condition with respect to the vector-time parameter 𝕥U\mathbb{t}\in U. For an ergodic probability measure μ\mu on the symbolic space \mathbb{N}^{\infty} with μ(n=1n)=1\mu(\cup_{n=1}^{\infty}\mathbb{N}_{n}^{\infty})=1, if the sequence of infinite random variables in 𝕐\mathbb{Y} under the law μ\mu is independent, then

(1.1) HD(Eα,G)Kα,GHD(E_{\alpha,G})\leq K_{\alpha,G}

for any 0<α<10<\alpha<1 and GUG\subset U such that limnhμn(σ)λμn𝕥(σ)\lim_{n\rightarrow\infty}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)} exists for any 𝕥G\mathbb{t}\in G, in which {μn}n\{\mu_{n}\}_{n\in\mathbb{N}} are the concentrating measures.

This result is a generalization of [SSU1, Theorem 5.3] from families of finite PIFS to families of infinite PIFS. See also [Kau, Theorem].

2. The setwise topology and semi-continuity of the measure-dimension mappings under it

Since we are considering families of IFS in this work, it would be very helpful if the measure-dimension mappings

dim:^(X)[0,)dim_{*}:\mathcal{\hat{M}}(X)\rightarrow[0,\infty)

or

dim:^(X)[0,)dim^{*}:\mathcal{\hat{M}}(X)\rightarrow[0,\infty)

admits some continuity property. To discuss the continuity problem, some appropriate topology on ^(X)\mathcal{\hat{M}}(X) is needed. Since the measures {ν𝕥}𝕥U^(X)\{\nu_{\mathbb{t}}\}_{\mathbb{t}\in U}\subset\mathcal{\hat{M}}(X) in Theorem 1.4 inherit the parametrization naturally from parametrization of the PIFS, one may try to endow the Euclidean metric on {ν𝕥}𝕥U\{\nu_{\mathbb{t}}\}_{\mathbb{t}\in U} through the parameter 𝕥Ud\mathbb{t}\in U\subset\mathbb{R}^{d} to discuss the continuity of the measure-dimension mappings. The measure-dimension mappings are not always continuous under the Euclidean metric on {ν𝕥}𝕥U\{\nu_{\mathbb{t}}\}_{\mathbb{t}\in U} in general consideration, although lower semi-continuity of them can be expected. One can see from the family of Bernoulli convolutions {νλ}λ^([0,1])\{\nu_{\lambda}\}_{\lambda}\subset\mathcal{\hat{M}}([0,1]) with the contraction ratio λ(0,1)\lambda\in(0,1). The measure-dimension mappings

dim=dim:(0,1)[0,1]dim_{*}=dim^{*}:(0,1)\rightarrow[0,1]

are not always continuous since there are dimension drops from 11 at the inverses of the Pisot parameters. However, they are lower semi-continuous [HS, Theorem 1.8].

Since we are dealing with measures originated from families of the infinite systems and their finite sub-systems simultaneously in this work, we will abandon the Euclidean metric on {ν𝕥}𝕥U^(X)\{\nu_{\mathbb{t}}\}_{\mathbb{t}\in U}\subset\mathcal{\hat{M}}(X) to discuss the continuity problem. A well-known topology on ^(X)\mathcal{\hat{M}}(X) is the weak topology.

2.1 Definition.

A sequence of bounded measures {νn^(X)}n=1\{\nu_{n}\in\mathcal{\hat{M}}(X)\}_{n=1}^{\infty} is said to converge weakly to ν^(X)\nu\in\mathcal{\hat{M}}(X), if

limnXf(x)𝑑νn=Xf(x)𝑑ν\lim_{n\rightarrow\infty}\int_{X}f(x)d\nu_{n}=\int_{X}f(x)d\nu

for any bounded continuous f:Xf:X\rightarrow\mathbb{R}.

Denote the convergence in this sense by νnwν\nu_{n}\stackrel{{\scriptstyle w}}{{\rightarrow}}\nu as nn\rightarrow\infty. However, The measure-dimension mappings dimdim^{*} and dimdim_{*} do not admit any semi-continuity under the weak topology on ^(X)\mathcal{\hat{M}}(X) [Ma, Theorem 3.1]. It turns out that the setwise topology is the most ideal topology for us to discuss the continuity of the measure-dimension mappings.

2.2 Definition.

A sequence of measures {νn^(X)}n=1\{\nu_{n}\in\mathcal{\hat{M}}(X)\}_{n=1}^{\infty} is said to converge setwisely to ν^(X)\nu\in\mathcal{\hat{M}}(X), if

limnνn(B)=ν(B)\lim_{n\rightarrow\infty}\nu_{n}(B)=\nu(B)

for any BB\in\mathcal{B}.

One is recommended to refer to [Doo, FKZ, GR, HL, Las, LY] for more equivalent descriptions of the setwise topology on ^(X)\mathcal{\hat{M}}(X) and its various applications. Denote the convergence in this sense by νnsν\nu_{n}\stackrel{{\scriptstyle s}}{{\rightarrow}}\nu as nn\rightarrow\infty. The measure-dimension mappings dimdim^{*} and dimdim_{*} are lower semi-continuous and upper semi-continuous under the setwise topology respectively on ^(X)\mathcal{\hat{M}}(X).

2.3 Theorem.

The measure-dimension mapping dimdim^{*} is lower semi-continuous, while dimdim_{*} is upper semi-continuous under the setwise topology on ^(X)\mathcal{\hat{M}}(X), that is, if νnsν\nu_{n}\stackrel{{\scriptstyle s}}{{\rightarrow}}\nu in ^(X)\mathcal{\hat{M}}(X) as nn\rightarrow\infty, then

(2.1) lim infndimνndimν\liminf_{n\rightarrow\infty}dim^{*}\nu_{n}\geq dim^{*}\nu

while

(2.2) lim supndimνndimν.\limsup_{n\rightarrow\infty}dim_{*}\nu_{n}\leq dim_{*}\nu.
2.4 Remark.

Similar results appear in [Ma, Theorem 2.8, 2.9] when one considers the measure-dimension mappings dimdim^{*} and dimdim_{*} on the probability space

(X)={ν:ν(X)=1,ν^(X)}\mathcal{M}(X)=\{\nu:\nu(X)=1,\nu\in\mathcal{\hat{M}}(X)\}

with XX being an arbitrary metric space. The proofs apply to the measure-dimension mappings on the space of finite measures ^(X)\mathcal{\hat{M}}(X) here, or even on the space of infinite measures.

Considering generalizations of measures, Theorem 2.3 is still true on the space of finitely additive measures (a finitely additive measure ν\nu on (X,)(X,\mathcal{B}) is a function from \mathcal{B} to [0,)[0,\infty) satisfying finite additivity instead of σ\sigma-additivity), but not true for a signed measure on (X,)(X,\mathcal{B}) under the Definition 1.1. Obviously a new meterage for the dimension of signed measures is needed. Let ^s(X)\mathcal{\hat{M}}_{s}(X) be the collection of all the finite signed Borel measures on XX.

2.5 Definition.

For a finite signed measure ν^s(X)\nu\in\mathcal{\hat{M}}_{s}(X), the lower and upper Hausdorff dimension of ν\nu are defined respectively to be:

dimsν=inf{HD(A):ν(A)>0 or ν(A)<0,A},dim_{s}\nu=\inf\{HD(A):\nu(A)>0\mbox{ or }\nu(A)<0,A\in\mathcal{B}\},

and

dimsν=max{inf{HD(A):ν(A)=max{ν(B)}B},inf{HD(A):ν(A)=min{ν(B)}B}}.dim^{s}\nu=\max\Big{\{}\inf\big{\{}HD(A):\nu(A)=\max\{\nu(B)\}_{B\in\mathcal{B}}\big{\}},\inf\big{\{}HD(A):\nu(A)=\min\{\nu(B)\}_{B\in\mathcal{B}}\big{\}}\Big{\}}.

Under this modified definition the measure-dimension mappings dimsdim_{s} and dimsdim^{s} from ^s(X)\mathcal{\hat{M}}_{s}(X) to [0,)[0,\infty) still bear some continuity.

2.6 Proposition.

The measure-dimension mapping dimsdim^{s} is lower semi-continuous, while dimsdim_{s} is upper semi-continuous under the setwise topology on ^s(X)\mathcal{\hat{M}}_{s}(X), that is, if νnsν\nu_{n}\stackrel{{\scriptstyle s}}{{\rightarrow}}\nu in ^s(X)\mathcal{\hat{M}}_{s}(X) as nn\rightarrow\infty, then

(2.3) lim infndimsνndimsν\liminf_{n\rightarrow\infty}dim^{s}\nu_{n}\geq dim^{s}\nu

while

(2.4) lim supndimsνndimsν.\limsup_{n\rightarrow\infty}dim_{s}\nu_{n}\leq dim_{s}\nu.
Proof.

The conclusion dimsdim_{s} is upper semi-continuous follows a similar argument as the proof of [Ma, Theorem 2.9]. Now we show dimsdim^{s} is lower semi-continuous under the setwise topology on ^s(X)\mathcal{\hat{M}}_{s}(X). Without loss of generality we assume

inf{HD(A):ν(A)=max{ν(B)}B}inf{HD(A):ν(A)=min{ν(B)}B}\inf\big{\{}HD(A):\nu(A)=\max\{\nu(B)\}_{B\in\mathcal{B}}\big{\}}\geq\inf\big{\{}HD(A):\nu(A)=\min\{\nu(B)\}_{B\in\mathcal{B}}\big{\}}

and

inf{HD(A):ν(A)=max{ν(B)}B}>0\inf\big{\{}HD(A):\nu(A)=\max\{\nu(B)\}_{B\in\mathcal{B}}\big{\}}>0, max{ν(B)}B>0\max\{\nu(B)\}_{B\in\mathcal{B}}>0.

Suppose conversely that (2.3) does not hold. Then we can find a subsequence {ni}i=1\{n_{i}\}_{i=1}^{\infty} and a real 0<a<inf{HD(A):ν(A)=max{ν(B)}B}0<a<\inf\big{\{}HD(A):\nu(A)=\max\{\nu(B)\}_{B\in\mathcal{B}}\big{\}} such that

(2.5) limidimsνni<a<dimsν=inf{HD(A):ν(A)=max{ν(B)}B}.\lim_{i\rightarrow\infty}dim^{s}\nu_{n_{i}}<a<dim^{s}\nu=\inf\big{\{}HD(A):\nu(A)=\max\{\nu(B)\}_{B\in\mathcal{B}}\big{\}}.

Now apply the Hahn decomposition theorem to the measure ν\nu, let X+XX^{+}\subset X be a positive set of ν\nu with ν(X+)=max{ν(B)}B\nu(X^{+})=\max\{\nu(B)\}_{B\in\mathcal{B}}. Under the above assumption we have

HD(X+)dimsν>a>dimsνniHD(X^{+})\geq dim^{s}\nu>a>dim^{s}\nu_{n_{i}}

for any ii large enough. Since limiνni(X+)=ν(X+)>0\lim_{i\rightarrow\infty}\nu_{n_{i}}(X^{+})=\nu(X^{+})>0, considering (2.5), for any ii large enough, there must exist Xi+X+X^{+}_{i}\subset X^{+}, such that

νni(Xi+)=0\nu_{n_{i}}(X^{+}_{i})=0 and HD(Xi+)=HD(X+)>a>HD(X+Xi+)HD(X^{+}_{i})=HD(X^{+})>a>HD(X^{+}\setminus X^{+}_{i}).

Now let X+=i=N(X+Xi+)X^{+}_{\infty}=\cup_{i=N}^{\infty}(X^{+}\setminus X^{+}_{i}) for some integer NN large enough. One can see that

X+X+X^{+}\setminus X^{+}_{\infty}\neq\emptyset

as

(2.6) HD(X+)a<HD(X+).HD(X^{+}_{\infty})\leq a<HD(X^{+}).

Moreover, we have

ν(X+X+)=limiνni(X+X+)=0\nu(X^{+}\setminus X^{+}_{\infty})=\lim_{i\rightarrow\infty}\nu_{n_{i}}(X^{+}\setminus X^{+}_{\infty})=0,

which forces

(2.7) ν(X+)=ν(X+).\nu(X^{+}_{\infty})=\nu(X^{+}).

Now (2.6) together with (2.7) contradict the fact that dimsν>adim^{s}\nu>a, which finishes the proof.

3. Absolute continuity of convergent sequences of measures under the setwise topology

In this section we discuss the relationship between absolute continuity of a sequence of measures and absolute continuity of its limit measure (under the weak or setwise topology, in case the sequence converges). We start from the following basic result.

3.1 Proposition.

For a sequence of measures νnsν\nu_{n}\stackrel{{\scriptstyle s}}{{\rightarrow}}\nu in ^(X)\mathcal{\hat{M}}(X) as nn\rightarrow\infty, if there exists a subsequence {νni}i=1\{\nu_{n_{i}}\}_{i=1}^{\infty} such that νni\nu_{n_{i}} is absolutely continuous with respect to some ϱ^(X)\varrho\in\mathcal{\hat{M}}(X) for any 1i<1\leq i<\infty, then ν\nu is absolutely continuous with respect to ϱ\varrho.

Proof.

We do this by reduction to absurdity. Suppose that ν\nu is not absolutely continuous with respect to ϱ\varrho, then there exists AA\in\mathcal{B} such that ν(A)>0\nu(A)>0 while ϱ(A)=0\varrho(A)=0. Since νnsν\nu_{n}\stackrel{{\scriptstyle s}}{{\rightarrow}}\nu as nn\rightarrow\infty, there exists NN large enough such that νnN(A)>0\nu_{n_{N}}(A)>0. Considering ϱ(A)=0\varrho(A)=0, this contradicts the fact that νnN\nu_{n_{N}} is absolutely continuous with respect to ϱ\varrho.

Due to Proposition 3.1, we make the following definition formally.

3.2 Definition.

A sequence of measures {νn}n=1\{\nu_{n}\}_{n=1}^{\infty} is said to be absolutely continuous with respect to some ϱ\varrho on ^(X)\mathcal{\hat{M}}(X) if it contains a subsequence such that every measure in the subsequence is absolutely continuous with respect to ϱ\varrho.

In fact Proposition 3.1 holds for any topological ambient space XX. However, it is possible that a sequence of absolutely continuous measures converges weakly to a measure which is not absolutely continuous, as one can see from the Example 3.3. Let 𝔏d|A\mathfrak{L}^{d}|_{A} be the restriction of 𝔏d\mathfrak{L}^{d} on a set AdA\subset\mathbb{R}^{d}.

3.3 Example.

Consider the Cantor middle-third set JJ appears as the attractor of the homogeneous IFS

S={s1(x)=13x,s2(x)=13x+23}S=\big{\{}s_{1}(x)=\frac{1}{3}x,s_{2}(x)=\frac{1}{3}x+\frac{2}{3}\big{\}}

on X=[0,1]X=[0,1]. Let

νn=12n𝔏1|ω2nsω(X)\nu_{n}=\frac{1}{2^{n}}\mathfrak{L}^{1}|_{\cup_{\omega\in\mathbb{N}_{2}^{n}}s_{\omega}(X)}.

Let ν\nu be the unique probability such that

ν=12νs11+12νs21\nu=\frac{1}{2}\nu\circ s_{1}^{-1}+\frac{1}{2}\nu\circ s_{2}^{-1}.

Unfortunately, the inverse of Proposition 3.1 is not always true, as one can see from the following example. Let δ{x}\delta_{\{x\}} be the Dirac probability measure at the point xXx\in X.

3.4 Example.

For X=[0,1]X=[0,1], let

νn=𝔏1|X+1nδ{1}\nu_{n}=\mathfrak{L}^{1}|_{X}+\frac{1}{n}\delta_{\{1\}}

be a sequence of measures in ^(X)\mathcal{\hat{M}}(X).

It is easy to justify that νns𝔏1|X\nu_{n}\stackrel{{\scriptstyle s}}{{\rightarrow}}\mathfrak{L}^{1}|_{X} as nn\rightarrow\infty in Example 3.4. However, there are no measure absolutely continuous with respect to 𝔏1|X\mathfrak{L}^{1}|_{X} at all in the sequence {νn}n=1\{\nu_{n}\}_{n=1}^{\infty}. One can see that in Example 3.4 the measures in the sequence are all of mixed type. In fact this is the only reason which prevents the inverse of Proposition 3.1 to be true.

3.5 Proposition.

For a sequence of measures νnsν\nu_{n}\stackrel{{\scriptstyle s}}{{\rightarrow}}\nu in ^(X)\mathcal{\hat{M}}(X) as nn\rightarrow\infty with ν(X)>0\nu(X)>0, if ν\nu is absolutely continuous with respect to some ϱ^(X)\varrho\in\mathcal{\hat{M}}(X) and the measures in the sequence are all of pure type with respect to ϱ\varrho, then there exists NN\in\mathbb{N} large enough such such that νn\nu_{n} is absolutely continuous with respect to ϱ^(X)\varrho\in\mathcal{\hat{M}}(X) for any nNn\geq N.

Proof.

Suppose in the contrary that the conclusion is not true, then we can find a subsequence {ni}i=1\{n_{i}\}_{i=1}^{\infty}\subset\mathbb{N} such that any measure in the sequence {νni}i=1\{\nu_{n_{i}}\}_{i=1}^{\infty} is singular with respect to ϱ\varrho. Denote an essential support of νni\nu_{n_{i}} by AiXA_{i}\subset X with ϱ(Ai)=0\varrho(A_{i})=0 for any 1i<1\leq i<\infty. Let A=i=1AiA=\cup_{i=1}^{\infty}A_{i}, according to the σ\sigma-additivity of ϱ\varrho, we have

(3.1) ϱ(A)=0.\varrho(A)=0.

Note that since νnsν\nu_{n}\stackrel{{\scriptstyle s}}{{\rightarrow}}\nu as nn\rightarrow\infty, we have

(3.2) ν(A)=limiνni(A)=limiνni(X)=ν(X)>0.\nu(A)=\lim_{i\rightarrow\infty}\nu_{n_{i}}(A)=\lim_{i\rightarrow\infty}\nu_{n_{i}}(X)=\nu(X)>0.

Now (3.1) together with (3.2) contradict the fact that ν\nu is absolutely continuous with respect to ϱ\varrho. ∎

Proposition 3.5 also holds for any topological ambient space XX, but it is not true for convergent sequences of measures under the weak topology on ^(X)\mathcal{\hat{M}}(X), as one can also see from our Example 3.3. The above results show that the absolute continuity of convergent sequences of measures and their limit measures are equivalent to each other under the setwise topology on ^(X)\mathcal{\hat{M}}(X), while this relationship is not true under the weak topology. This is another advantage that the setwise topology takes over the weak topology on ^(X)\mathcal{\hat{M}}(X).

4. The nn-th concentrating measures of probability on the symbolic spaces and some inherited properties

In this section we do the approximation of a probability measure μ\mu on the infinite symbolic space \mathbb{N}^{\infty} by the sequence of its concentrating measures {μn}n=1\{\mu_{n}\}_{n=1}^{\infty} supported on {n}n=1\{\mathbb{N}_{n}^{\infty}\}_{n=1}^{\infty} respectively. We will prove that some properties are inherited by the concentrating measures from μ\mu, under the hypothesis that the sequence of infinite random variables in the infinite stochastic process

𝕐={Yi}i\mathbb{Y}=\{Y_{i}\}_{i\in\mathbb{N}}

is independent under the law μ\mu. As indicated in Remark 1.6, one may be able to remove the independent hypothesis in due course if one tries the approximation of μ\mu by its restrictions {μ|n}n=1^(n)\{\mu|_{\mathbb{N}_{n}^{\infty}}\}_{n=1}^{\infty}\subset\mathcal{\hat{M}}(\mathbb{N}_{n}^{\infty}) instead of the concentrating measures {μn}n=1(n)\{\mu_{n}\}_{n=1}^{\infty}\subset\mathcal{M}(\mathbb{N}_{n}^{\infty}).

4.1 Definition.

For a probability measure μ\mu on the symbolic space \mathbb{N}^{\infty}, define its nn-th concentrating measure μn\mu_{n} for any nn\in\mathbb{N} to be the unique probability measure supported on n\mathbb{N}_{n}^{\infty} satisfying the following conditions.

  1. (1)

    On the first level cylinders,

    μn([i])=μ([i])\mu_{n}([i])=\mu([i]) for 1in11\leq i\leq n-1, μn([n])=i=nμ([i])\mu_{n}([n])=\sum_{i=n}^{\infty}\mu([i]).

  2. (2)

    On the second level cylinders,

    μn([ij])=μ([ij]) for 1i,jn1,μn([in])=k=nμ([ik]) for 1in1,μn([nj])=k=nμ([kj]) for 1jn1,μn([nn])=l=nk=nμ([kl]).\begin{array}[]{l}\mu_{n}([ij])=\mu([ij])\mbox{ for }1\leq i,j\leq n-1,\\ \mu_{n}([in])=\sum_{k=n}^{\infty}\mu([ik])\mbox{ for }1\leq i\leq n-1,\\ \mu_{n}([nj])=\sum_{k=n}^{\infty}\mu([kj])\mbox{ for }1\leq j\leq n-1,\\ \mu_{n}([nn])=\sum_{l=n}^{\infty}\sum_{k=n}^{\infty}\mu([kl]).\end{array}

  3. (3)

    For any qq-th cylinder [i1i2iq]n[i_{1}i_{2}\cdots i_{q}]\subset\mathbb{N}_{n}^{\infty} with i1,i2,,iqni_{1},i_{2},\cdots,i_{q}\in\mathbb{N}_{n} and qq\in\mathbb{N}, let

    #{1kq:ik=n}=l\#\{1\leq k\leq q:i_{k}=n\}=l.

    Number the indexes in {1kq:ik=n}\{1\leq k\leq q:i_{k}=n\} in increasing order as

    j1<j2<<jlj_{1}<j_{2}<\cdots<j_{l},

    that is, {j1,j2,,jl}={1kq:ik=n}\{j_{1},j_{2},\cdots,j_{l}\}=\{1\leq k\leq q:i_{k}=n\}. Then

    μn([i1i2iq])=nr1,r2,,rl<μ([i1ij11r1ij1+1ij21r2ij2+1ijl1rlijl+1iq]).\begin{array}[]{ll}&\mu_{n}([i_{1}i_{2}\cdots i_{q}])\\ =&\sum_{n\leq r_{1},r_{2},\cdots,r_{l}<\infty}\mu([i_{1}\cdots i_{j_{1}-1}r_{1}i_{j_{1}+1}\cdots i_{j_{2}-1}r_{2}i_{j_{2}+1}\cdots i_{j_{l}-1}r_{l}i_{j_{l}+1}\cdots i_{q}]).\end{array}

Since for any qq-level cylinder [i1i2iq]n[i_{1}i_{2}\cdots i_{q}]\subset\mathbb{N}_{n}^{\infty} with i1,i2,iqni_{1},i_{2},\cdots i_{q}\in\mathbb{N}_{n} and qq\in\mathbb{N}, the conditions in Definition 4.1 (1)-(3) imply

μn([i1i2iq])=j=1nμn([i1i2iqj])\mu_{n}([i_{1}i_{2}\cdots i_{q}])=\sum_{j=1}^{n}\mu_{n}([i_{1}i_{2}\cdots i_{q}j])

between cylinders of successive levels, the existence and uniqueness of the measure μn\mu_{n} are guaranteed by the Kolmogorov extension theorem, see for example [Tao, Theorem 2.4.3]. Moreover, μn\mu_{n} is invariant for any nn\in\mathbb{N} if μ\mu is invariant with respect to the shift map σ\sigma according to [Ber, Theorem 2]. In fact, more properties are inherited from μ\mu as the structure of the original measure is suitably preserved considering the structure of the concentrating measures {μn}n\{\mu_{n}\}_{n\in\mathbb{N}} on the cylinder sets, under the independent hypothesis of 𝕐\mathbb{Y} with respect to μ\mu.

4.2 Lemma.

For a measure μ\mu on the symbolic space \mathbb{N}^{\infty}, if the sequence of infinite random variables in 𝕐\mathbb{Y} under the law μ\mu is independent, then it is also independent under the concentrating law μn\mu_{n} for any nn\in\mathbb{N}.

Proof.

For any qq-word ω=i1i2iqnq\omega=i_{1}i_{2}\cdots i_{q}\in\mathbb{N}_{n}^{q}, let #{1kq:ik=n}=l0\#\{1\leq k\leq q:i_{k}=n\}=l\geq 0 and

{j1,j2,,jl}={1kq:ik=n}\{j_{1},j_{2},\cdots,j_{l}\}=\{1\leq k\leq q:i_{k}=n\}.

Since 𝕐\mathbb{Y} is independent under the law μ\mu, we have

μn([ω])=nr1<,nr2<,,nrl<μ([i1ij11r1ij1+1ij21r2ij2+1ijl1rlijl+1iq])=μ([i1])μ([ij11])(nr1<μ([r1]))μ([ij1+1])μ([ij21])(nr2<μ([r2]))μ([ij2+1])μ([ijl1])(nrl<μ([rl]))μ([ijl+1])μ([iq])=μn([i1])μn([ij11])μn([ij1])μn([ij1+1])μn([ij21])μn([ij2])μn([ij2+1])μn([ijl1])μn([ijl])μn([ijl+1])μn([iq]),\begin{array}[]{ll}&\mu_{n}([\omega])\\ =&\sum_{n\leq r_{1}<\infty,n\leq r_{2}<\infty,\cdots,n\leq r_{l}<\infty}\mu([i_{1}\cdots i_{j_{1}-1}r_{1}i_{j_{1}+1}\cdots i_{j_{2}-1}r_{2}i_{j_{2}+1}\cdots i_{j_{l}-1}r_{l}i_{j_{l}+1}\cdots i_{q}])\\ =&\mu([i_{1}])\cdots\mu([i_{j_{1}-1}])\big{(}\sum_{n\leq r_{1}<\infty}\mu([r_{1}])\big{)}\mu([i_{j_{1}+1}])\cdots\mu([i_{j_{2}-1}])\big{(}\sum_{n\leq r_{2}<\infty}\mu([r_{2}])\big{)}\mu([i_{j_{2}+1}])\\ &\cdots\mu([i_{j_{l}-1}])\big{(}\sum_{n\leq r_{l}<\infty}\mu([r_{l}])\big{)}\mu([i_{j_{l}+1}])\cdots\mu([i_{q}])\\ =&\mu_{n}([i_{1}])\cdots\mu_{n}([i_{j_{1}-1}])\mu_{n}([i_{j_{1}}])\mu_{n}([i_{j_{1}+1}])\cdots\mu_{n}([i_{j_{2}-1}])\mu_{n}([i_{j_{2}}])\mu_{n}([i_{j_{2}+1}])\cdots\\ &\mu_{n}([i_{j_{l}-1}])\mu_{n}([i_{j_{l}}])\mu_{n}([i_{j_{l}+1}])\cdots\mu_{n}([i_{q}]),\end{array}

which justifies the independence of the random variables in 𝕐\mathbb{Y} under the law μn\mu_{n} for any nn\in\mathbb{N}. ∎

The ergodicity is also inherited by the concentrating measures {μn}n\{\mu_{n}\}_{n\in\mathbb{N}} under the independent hypothesis of 𝕐\mathbb{Y}.

4.3 Lemma.

For an ergodic measure μ\mu on the symbolic space \mathbb{N}^{\infty}, if the sequence of infinite random variables in 𝕐\mathbb{Y} under the law μ\mu is independent, then the concentrating measure μn\mu_{n} is also ergodic on n\mathbb{N}_{n}^{\infty} for any nn\in\mathbb{N} with respect to the shift map σ\sigma.

Proof.

This is a standard proof in classical probability techniques, with all the above preparations. First, due to Lemma 4.2 and the independence assumption, the sequence of random variables {Yi}i=1\{Y_{i}\}_{i=1}^{\infty} in 𝕐\mathbb{Y} under the law μn\mu_{n} is independent for any nn\in\mathbb{N}. Now let n\mathcal{B}_{\mathbb{N}_{n}^{\infty}} be the σ\sigma-algebra generated by the cylinder sets in n\mathbb{N}_{n}^{\infty} for some nn\in\mathbb{N}. Let

k={nk×B:Bn}\mathcal{B}_{k}=\{\mathbb{N}_{n}^{k}\times B:B\in\mathcal{B}_{\mathbb{N}_{n}^{\infty}}\}.

The tail σ\sigma-algebra is define to be

T=kk\mathcal{B}_{T}=\cap_{k\in\mathbb{N}}\mathcal{B}_{k}.

Obviously the events

{An:σ1(A)=A}T\{A\in\mathcal{B}_{\mathbb{N}_{n}^{\infty}}:\sigma^{-1}(A)=A\}\subset\mathcal{B}_{T}

are all tail events. So according to the Kolmogorov’s 010-1 law (see for example [Tsi, Proposition 3b9] or [Shi]),

μn(A)=0\mu_{n}(A)=0 or 11

for any event in {An:σ1(A)=A}\{A\in\mathcal{B}_{\mathbb{N}_{n}^{\infty}}:\sigma^{-1}(A)=A\}. ∎

4.4 Remark.

Similar to Remark 1.5, we are wondering whether the result holds without the hypothesis of independence on the sequence of infinite random variables with law μ\mu, which essentially confines our main results to be under this condition.

Considering the limit behaviour of the sequence of concentrating measures {μn}n\{\mu_{n}\}_{n\in\mathbb{N}}, we have the following simple but important result.

4.5 Proposition.

For a probability measure μ\mu on the symbolic space \mathbb{N}^{\infty} such that μ(n=1n)=1\mu(\cup_{n=1}^{\infty}\mathbb{N}_{n}^{\infty})=1, let μn\mu_{n} be its nn-th concentrating measure on n\mathbb{N}_{n}^{\infty} for any nn\in\mathbb{N}. Then

μnsμ\mu_{n}\stackrel{{\scriptstyle s}}{{\rightarrow}}\mu

as nn\rightarrow\infty.

Proof.

It is obvious that for any finite word ω\omega\in\mathbb{N}^{*}, we have

limnμn([ω])=μ([ω])\lim_{n\rightarrow\infty}\mu_{n}([\omega])=\mu([\omega]),

as μn([ω])=μ([ω])\mu_{n}([\omega])=\mu([\omega]) for any nn large enough. Since μ\mu is supported on n=1n\cup_{n=1}^{\infty}\mathbb{N}_{n}^{\infty}, the convergence extends to all measurable sets in \mathbb{N}^{\infty} due to regularity of the measures, or [FKZ, Theorem 2.3]. ∎

4.6 Remark.

The convergence of the sequence of concentrating measures {μn}n\{\mu_{n}\}_{n\in\mathbb{N}} of a measure μ\mu is usually not true under the total variation (TV) topology on ()\mathcal{M}(\mathbb{N}^{\infty}), except in some trivial cases.

The entropy is also inherited by the sequence of concentrating measures {μn}n\{\mu_{n}\}_{n\in\mathbb{N}} from the original measure μ\mu, considering Proposition 4.5 as well as the following result.

4.7 Lemma.

For a finite measure μ^()\mu\in\mathcal{\hat{M}}(\mathbb{N}^{\infty}) and its concentrating measures {μn}n\{\mu_{n}\}_{n\in\mathbb{N}}, we have

limnhμn(σ)=hμ(σ)\lim_{n\rightarrow\infty}h_{\mu_{n}}(\sigma)=h_{\mu}(\sigma)

with respect to (,)(\mathbb{N}^{\infty},\mathcal{B}_{\mathbb{N}^{\infty}}).

Proof.

We justify the result in case of hμ(σ)=h_{\mu}(\sigma)=\infty and hμ(σ)<h_{\mu}(\sigma)<\infty respectively.

If

hμ(σ)=lim infk1kωkμ([ω])logμ([ω])=h_{\mu}(\sigma)=\liminf_{k\rightarrow\infty}-\frac{1}{k}\sum_{\omega\in\mathbb{N}^{k}}\mu([\omega])\log\mu([\omega])=\infty,

we have

1kωkμ([ω])logμ([ω])=-\frac{1}{k}\sum_{\omega\in\mathbb{N}^{k}}\mu([\omega])\log\mu([\omega])=\infty

for any kk\in\mathbb{N}. Then for any a>0a>0 and kk\in\mathbb{N}, we can find Na,kN_{a,k} large enough, such that

1kωnkμn([ω])logμn([ω])>a-\frac{1}{k}\sum_{\omega\in\mathbb{N}_{n}^{k}}\mu_{n}([\omega])\log\mu_{n}([\omega])>a

for any n>Na,kn>N_{a,k}. This is enough to force limnhμn(σ)=hμ(σ)=\lim_{n\rightarrow\infty}h_{\mu_{n}}(\sigma)=h_{\mu}(\sigma)=\infty by reduction to absurdity.

Now if

hμ(σ)=lim infk1kωkμ([ω])logμ([ω])<h_{\mu}(\sigma)=\liminf_{k\rightarrow\infty}-\frac{1}{k}\sum_{\omega\in\mathbb{N}^{k}}\mu([\omega])\log\mu([\omega])<\infty,

then for any small ϵ>0\epsilon>0, we can find KϵK_{\epsilon} large enough, such that

hμ(σ)1kωkμ([ω])logμ([ω])<hμ(σ)+ϵh_{\mu}(\sigma)\leq-\frac{1}{k}\sum_{\omega\in\mathbb{N}^{k}}\mu([\omega])\log\mu([\omega])<h_{\mu}(\sigma)+\epsilon

for any k>Kϵk>K_{\epsilon}. Now for fixed k>Kϵk>K_{\epsilon}, since ωkμ([ω])logμ([ω])<-\sum_{\omega\in\mathbb{N}^{k}}\mu([\omega])\log\mu([\omega])<\infty, we can find Nk,ϵN_{k,\epsilon} large enough such that

|ωkμ([ω])logμ([ω])ωnkμ([ω])logμn([ω])|<ϵ\big{|}\sum_{\omega\in\mathbb{N}^{k}}\mu([\omega])\log\mu([\omega])-\sum_{\omega\in\mathbb{N}_{n}^{k}}\mu_{(}[\omega])\log\mu_{n}([\omega])\big{|}<\epsilon

for any n>Nk,ϵn>N_{k,\epsilon}. This is enough to force limnhμn(σ)=hμ(σ)\lim_{n\rightarrow\infty}h_{\mu_{n}}(\sigma)=h_{\mu}(\sigma).

4.8 Remark.

In fact Lemma 4.7 holds for any sequence of measures converging to the finite measure μ\mu under the setwise topology in ^(X)\mathcal{\hat{M}}(X), not only for the sequence of its concentrating measures, for any topological ambient space XX. The version now is enough for our purpose in this work.

5. Setwise approximation of projective measures on the attractors of families of infinite PIFS

In this section we aim at proving Theorem 1.4. After the proof we formulate some interesting applications of Theorem 1.4 in some circumstances. We first show a result which links the upper and lower dimensions of the projective measure ν\nu on the attractor JJ of a single PIFS.

5.1 Lemma.

Let

S={si:XX}iIS=\{s_{i}:X\rightarrow X\}_{i\in I}

be a parabolic iterated function system with a countable index set II. For an ergodic measure μ\mu on II^{\infty}, let ν=μπ1\nu=\mu\circ\pi^{-1} be its projective measure. Then

dimν=dimνdim_{*}\nu=dim^{*}\nu.

Proof.

To prove the result, it suffices for us to prove

(5.1) dimνdimν.dim_{*}\nu\geq dim^{*}\nu.

Let JJ be the attractor of the PIFS. For a set AJA\subset J with ν(A)=μπ1(A)>0\nu(A)=\mu\circ\pi^{-1}(A)>0, consider the set

πσ1π1(A)=π(iIiπ1(A))=iIπ(iπ1(A))\pi\circ\sigma^{-1}\circ\pi^{-1}(A)=\pi(\cup_{i\in I}i\pi^{-1}(A))=\cup_{i\in I}\pi(i\pi^{-1}(A)).

Note that for any iIi\in I, due to π(ω)=sω|nπσn(ω)\pi(\omega)=s_{\omega|_{n}}\circ\pi\circ\sigma^{n}(\omega) for any ωI\omega\in I^{\infty}, we have

π(iπ1(A))=si(A)\pi(i\pi^{-1}(A))=s_{i}(A),

in which iπ1(A):={ωI:ω1=i,σ(ω)π1(A)}i\pi^{-1}(A):=\{\omega\in I^{\infty}:\omega_{1}=i,\sigma(\omega)\in\pi^{-1}(A)\}. Since 0<si(x)10<s_{i}^{\prime}(x)\leq 1 for any iIi\in I, we have

HD(π(iπ1(A)))=HD(A)HD\big{(}\pi(i\pi^{-1}(A))\big{)}=HD(A)

for any iIi\in I. Since II is countable, this forces

HD(πσ1π1(A))=HD(A)HD(\pi\circ\sigma^{-1}\circ\pi^{-1}(A))=HD(A).

Successively we can show HD(πσnπ1(A))=HD(A)HD(\pi\circ\sigma^{-n}\circ\pi^{-1}(A))=HD(A) for any nn\in\mathbb{N}, which forces

HD(nπσnπ1(A))=HD(A)HD(\cup_{n\in\mathbb{N}}\pi\circ\sigma^{-n}\circ\pi^{-1}(A))=HD(A).

Since μ\mu is ergodic with respect to σ\sigma, then

ν(nπσnπ1(A))=μπ1(nπσnπ1(A))μ(nσnπ1(A))=1.\begin{array}[]{ll}&\nu\big{(}\cup_{n\in\mathbb{N}}\pi\circ\sigma^{-n}\circ\pi^{-1}(A)\big{)}\\ =&\mu\circ\pi^{-1}\big{(}\cup_{n\in\mathbb{N}}\pi\circ\sigma^{-n}\circ\pi^{-1}(A)\big{)}\\ \geq&\mu\big{(}\cup_{n\in\mathbb{N}}\sigma^{-n}\circ\pi^{-1}(A)\big{)}\\ =&1.\end{array}

This implies the inequality (5.1).

Lemma 5.1 may not be true for some invariant measure μ\mu on the symbolic space. Apply it to families of PIFS, we have the following result instantly.

5.2 Corollary.

Let

{S𝕥={si𝕥:XX}iI}𝕥U\big{\{}S^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in I}\big{\}}_{\mathbb{t}\in U}

be a family of parabolic iterated function systems with a countable index set II. Let μ\mu be an ergodic measure on II^{\infty} and ν𝕥\nu_{\mathbb{t}} be the projective measure under π𝕥\pi_{\mathbb{t}} at time 𝕥\mathbb{t}, then

dimν𝕥=dimν𝕥dim_{*}\nu_{\mathbb{t}}=dim^{*}\nu_{\mathbb{t}}

for any 𝕥U\mathbb{t}\in U.

Note that the equality holds everywhere instead of Lebesgue a.e. with respect to 𝕥U\mathbb{t}\in U. Now we consider setwisely approximating the projective measure of μ\mu by the sequence of projective measures of its concentrating measures on the ambient space XX. The following result is an instant corollary of Proposition 4.5.

5.3 Corollary.

Let

S={si:XX}iS=\{s_{i}:X\rightarrow X\}_{i\in\mathbb{N}}

be an infinite parabolic iterated function system with its attractor JJ. For a probability measure μ\mu on the symbolic space \mathbb{N}^{\infty} such that μ(n=1n)=1\mu(\cup_{n=1}^{\infty}\mathbb{N}_{n}^{\infty})=1, let μn\mu_{n} be its nn-th concentrating measure on n\mathbb{N}_{n}^{\infty} for any nn\in\mathbb{N}. Then their projections {νn=μnπ1}n\{\nu_{n}=\mu_{n}\circ\pi^{-1}\}_{n\in\mathbb{N}} and ν=μπ1\nu=\mu\circ\pi^{-1} satisfy

νnsν\nu_{n}\stackrel{{\scriptstyle s}}{{\rightarrow}}\nu

as nn\rightarrow\infty on JJ.

As to absolute continuity of the projective measures on the ambient space, we have the following result.

5.4 Lemma.

Let

S={si:XX}iS=\{s_{i}:X\rightarrow X\}_{i\in\mathbb{N}}

be an infinite parabolic iterated function system with attractor JJ. For an ergodic probability measure μ\mu on the symbolic space \mathbb{N}^{\infty}, with μ(n=1n)=1\mu(\cup_{n=1}^{\infty}\mathbb{N}_{n}^{\infty})=1 and its nn-th concentrating measure μn\mu_{n} also being ergodic on n\mathbb{N}_{n}^{\infty} for any nn\in\mathbb{N}, consider their projections {νn=μnπ1}n\{\nu_{n}=\mu_{n}\circ\pi^{-1}\}_{n\in\mathbb{N}} and ν=μπ1\nu=\mu\circ\pi^{-1} on JJ. The sequence of measures {νn}n\{\nu_{n}\}_{n\in\mathbb{N}} is absolutely continuous with respect to 𝔏1\mathfrak{L}^{1} if and only if ν\nu is absolutely continuous with respect to 𝔏1\mathfrak{L}^{1}.

Proof.

First note that all the projective measures {νn}n\{\nu_{n}\}_{n\in\mathbb{N}} and ν\nu are of pure type since they are all projected from ergodic measures on the symbolic spaces. By Corollary 5.3 we have

νnsν\nu_{n}\stackrel{{\scriptstyle s}}{{\rightarrow}}\nu

as nn\rightarrow\infty on JJ. Then the absolute continuity of the sequence of measures {νn}n\{\nu_{n}\}_{n\in\mathbb{N}} and the absolute continuity of ν\nu with respect to 𝔏1\mathfrak{L}^{1} is equivalent to each other in virtue of Proposition 3.1 and Proposition 3.5. ∎

We will only use the result that absolute continuity of the sequence of measures {νn}n\{\nu_{n}\}_{n\in\mathbb{N}} implies the absolute continuity of ν\nu with respect to 𝔏1\mathfrak{L}^{1} in the following. This is true even if {μn}n\{\mu_{n}\}_{n\in\mathbb{N}} are not ergodic according to Proposition 3.1. Equipped with all the above results, now we are well prepared to prove Theorem 1.4.

Proof of Theorem 1.4:

Proof.

First note that since {S𝕥={si𝕥:XX}iΓX(ϑ)}𝕥U\big{\{}S^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}}\in\Gamma_{X}(\vartheta)\big{\}}_{\mathbb{t}\in U}, there exist sequences of positive real numbers {0<γn,un<1}n\{0<\gamma_{n},u_{n}<1\}_{n\in\mathbb{N}}, {Mn>0}n\{M_{n}>0\}_{n\in\mathbb{N}} (it is possible that these sequences satisfy sup{γn}n=1\sup\{\gamma_{n}\}_{n\in\mathbb{N}}=1, inf{un}n=0\inf\{u_{n}\}_{n\in\mathbb{N}}=0 and sup{Mn}n=\sup\{M_{n}\}_{n\in\mathbb{N}}=\infty) and a sequence of open neighbourhoods VnV_{n} of vv (it is also possible that n=1Vn={v}\cap_{n=1}^{\infty}V_{n}=\{v\}), such that its nn-th family of truncates

{Sn𝕥={si𝕥:XX}inΓX(ϑ,Vn,γn,un,Mn)}𝕥U\big{\{}S_{n}^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}_{n}}\in\Gamma_{X}(\vartheta,V_{n},\gamma_{n},u_{n},M_{n})\big{\}}_{\mathbb{t}\in U}

for any fixed nn\in\mathbb{N}. Denote their attractors by {Jn,𝕥J𝕥}n,𝕥U\{J_{n,\mathbb{t}}\subset J_{\mathbb{t}}\}_{n\in\mathbb{N},\mathbb{t}\in U}. Consider the projections of the concentrating measures

νn,𝕥=μnπ𝕥1\nu_{n,\mathbb{t}}=\mu_{n}\circ\pi_{\mathbb{t}}^{-1}

on the limit sets for 𝕥U,n\mathbb{t}\in U,n\in\mathbb{N}. Since the sequence of infinite random variables in 𝕐\mathbb{Y} is independent under the ergodic law μ\mu, the concentrating measures {μn}n\{\mu_{n}\}_{n\in\mathbb{N}} are also ergodic with respect to the shift map σ\sigma in virtue of Lemma 4.3. Moreover, according to Lemma 4.7, we have

hμn(σ)>0h_{\mu_{n}}(\sigma)>0

for nn large enough. Now apply the [SSU1, Theorem 2.3 (i)] and Corollary 5.2 to the ergodic projections {νn,𝕥}𝕥U\{\nu_{n,\mathbb{t}}\}_{\mathbb{t}\in U} originated from the family of finite PIFS

{Sn𝕥={si𝕥:XX}inΓX(ϑ,Vn,γn,un,Mn)}𝕥U\big{\{}S_{n}^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}_{n}}\in\Gamma_{X}(\vartheta,V_{n},\gamma_{n},u_{n},M_{n})\big{\}}_{\mathbb{t}\in U}

satisfying the continuity and transversality condition, we have

dimνn,𝕥=dimνn,𝕥=min{hμn(σ)λμn𝕥(σ),1}dim_{*}\nu_{n,\mathbb{t}}=dim^{*}\nu_{n,\mathbb{t}}=\min\Big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)},1\Big{\}}

for Lebesgue a.e. 𝕥U\mathbb{t}\in U and any nn\in\mathbb{N} large enough. In virtue of Corollary 5.3, we have

νn,𝕥sν𝕥\nu_{n,\mathbb{t}}\stackrel{{\scriptstyle s}}{{\rightarrow}}\nu_{\mathbb{t}}

as nn\rightarrow\infty on JJ for any 𝕥U\mathbb{t}\in U. Now apply the semi-continuity result Theorem 2.3 to the setwisely convergent sequence of projective measures {νn,𝕥}𝕥U\{\nu_{n,\mathbb{t}}\}_{\mathbb{t}\in U}, we have

(5.2) lim supnmin{hμn(σ)λμn𝕥(σ),1}=lim supndimνn,𝕥dimν𝕥=dimν𝕥lim infndimνn,𝕥=lim infnmin{hμn(σ)λμn𝕥(σ),1}\begin{array}[]{ll}\limsup_{n\rightarrow\infty}\min\Big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)},1\Big{\}}=\limsup_{n\rightarrow\infty}dim_{*}\nu_{n,\mathbb{t}}\leq dim_{*}\nu_{\mathbb{t}}\\ =dim^{*}\nu_{\mathbb{t}}\leq\liminf_{n\rightarrow\infty}dim^{*}\nu_{n,\mathbb{t}}=\liminf_{n\rightarrow\infty}\min\Big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)},1\Big{\}}\end{array}

for Lebesgue a.e. 𝕥U\mathbb{t}\in U and any nn\in\mathbb{N} large enough. Then Theorem 1.4 (i) is proved by letting nn\rightarrow\infty in (5.2), since the Lebesgue measure of unions of countablly many Lebesgue null sets is still null.

To prove Theorem 1.4 (ii), for fixed nn\in\mathbb{N}, apply [SSU1, Theorem 2.3 (ii)] to the ergodic projections {νn,𝕥}𝕥U\{\nu_{n,\mathbb{t}}\}_{\mathbb{t}\in U} originated from the nn-th truncated family of PIFS {Sn𝕥}𝕥U\{S_{n}^{\mathbb{t}}\}_{\mathbb{t}\in U} satisfying the continuity and transversality condition, we see that νn,𝕥\nu_{n,\mathbb{t}} is absolutely continuous for a.e. 𝕥{𝕥U:hμn(σ)λμn𝕥(σ)>1}\mathbb{t}\in\Big{\{}\mathbb{t}\in U:\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)}>1\Big{\}}. So in virtue of Lemma 5.4, ν𝕥\nu_{\mathbb{t}} is absolutely continuous for a.e. 𝕥\mathbb{t} in

{nj}j=1 is an infinite subsequence of {n}n=1j=1{𝕥U:hμnj(σ)λμnj𝕥(σ)>1}:=Ua\cup_{\{n_{j}\}_{j=1}^{\infty}\mbox{ is an infinite subsequence of }\{n\}_{n=1}^{\infty}}\cap_{j=1}^{\infty}\Big{\{}\mathbb{t}\in U:\cfrac{h_{\mu_{n_{j}}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n_{j}}}(\sigma)}>1\Big{\}}:=U_{a}.

It is easy to see that

{𝕥U:lim supnhμn(σ)λμn𝕥(σ)>1}Ua\Big{\{}\mathbb{t}\in U:\limsup_{n\rightarrow\infty}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)}>1\Big{\}}\subset U_{a}.

This justifies Theorem 1.4 (ii). ∎

One can see from the above proof that the limit of the sequence

{min{hμn(σ)λμn𝕥(σ),1}}n=1\Big{\{}\min\big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)},1\big{\}}\Big{\}}_{n=1}^{\infty}

always exists for Lebesgue a.e. 𝕥U\mathbb{t}\in U. More interesting results are possible according to different a.e. limit behaviours of the sequence {hμn(σ)λμn𝕥(σ)}𝕥U,n\Big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)}\Big{\}}_{\mathbb{t}\in U,n\in\mathbb{N}}.

In the following of the section we apply Theorem 1.4 to various families of PIFS to deduce some interesting results. The first application is on the Hausdorff dimension of the attractors of a family of infinite PIFS.

5.5 Corollary.

Let

{S𝕥={si𝕥:XX}iΓX(ϑ)}𝕥U\big{\{}S^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}}\in\Gamma_{X}(\vartheta)\big{\}}_{\mathbb{t}\in U}

be a family of infinite parabolic iterated function systems satisfying the continuity and transversality condition with respect to the vector-time parameter 𝕥U\mathbb{t}\in U. For an ergodic probability measure μ\mu on the symbolic space \mathbb{N}^{\infty} with positive entropy hμ(σ)h_{\mu}(\sigma) and μ(n=1n)=1\mu(\cup_{n=1}^{\infty}\mathbb{N}_{n}^{\infty})=1, let μn\mu_{n} be its nn-th concentrating measure for any nn\in\mathbb{N}. If the sequence of infinite random variables in 𝕐\mathbb{Y} under the law μ\mu is independent, then

  1. (i).

    HD(J𝕥)limnmin{hμn(σ)λμn𝕥(σ),1}HD(J_{\mathbb{t}})\geq\lim_{n\rightarrow\infty}\min\Big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)},1\Big{\}} for Lebesgue a.e. 𝕥U\mathbb{t}\in U.

  2. (ii).

    𝔏1(J𝕥)>0\mathfrak{L}^{1}(J_{\mathbb{t}})>0 for Lebesgue a.e. 𝕥{𝕥:lim supn{hμn(σ)λμn𝕥(σ)}>1}\mathbb{t}\in\Big{\{}\mathbb{t}:\limsup_{n\rightarrow\infty}\big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)}\big{\}}>1\Big{\}}.

Proof.

These results follow directly from Theorem 1.4 and the fact that supp(ν𝕥)J𝕥supp(\nu_{\mathbb{t}})\subset J_{\mathbb{t}} for any 𝕥U\mathbb{t}\in U. ∎

More interesting corollaries on the Hausdorff dimension of the attractors can be formulated from Theorem 1.4, considering different limit behaviours of the sequence {hμn(σ)λμn𝕥(σ)}𝕥U,n\Big{\{}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda^{\mathbb{t}}_{\mu_{n}}(\sigma)}\Big{\}}_{\mathbb{t}\in U,n\in\mathbb{N}} and the fact that supp(ν𝕥)supp(\nu_{\mathbb{t}}) is always contained in J𝕥J_{\mathbb{t}}.

Now we apply Theorem 1.4 or Corollary 5.5 to some families of infinite PIFS with exploding measures μ\mu on \mathbb{N}^{\infty}, that is, measures μ\mu satisfying

hμ(σ)=h_{\mu}(\sigma)=\infty.

There will be some stronger results available in due courses.

5.6 Corollary.

Let

{S𝕥={si𝕥:XX}iΓX(ϑ,V,γ,u,M)}𝕥U\big{\{}S^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}}\in\Gamma_{X}(\vartheta,V,\gamma,u,M)\big{\}}_{\mathbb{t}\in U}

be a family of infinite parabolic iterated function systems satisfying the continuity and transversality condition with respect to the vector-time parameter 𝕥U\mathbb{t}\in U for some fixed open neighbourhood VV of vv and some fixed parameters 0<ϑ,γ,u<1,M>00<\vartheta,\gamma,u<1,M>0. For an ergodic exploding probability measure μ\mu on the symbolic space \mathbb{N}^{\infty} with μ(n=1n)=1\mu(\cup_{n=1}^{\infty}\mathbb{N}_{n}^{\infty})=1, if the sequence of infinite random variables in 𝕐\mathbb{Y} under the law μ\mu is independent, then

  1. (i).

    For Lebesgue a.e. 𝕥U\mathbb{t}\in U, dimν𝕥=dimν𝕥=1dim_{*}\nu_{\mathbb{t}}=dim^{*}\nu_{\mathbb{t}}=1,

  2. (ii).

    ν𝕥\nu_{\mathbb{t}} is absolutely continuous for Lebesgue a.e. 𝕥U\mathbb{t}\in U,

in which ν𝕥=μπ𝕥1\nu_{\mathbb{t}}=\mu\circ\pi_{\mathbb{t}}^{-1} is the projective measure at time 𝕥\mathbb{t}.

Proof.

Note that since

(5.3) infn,xX{|(sn𝕥)(x)|}u\inf_{n\in\mathbb{N},x\in X}\{|(s_{n}^{\mathbb{t}})^{\prime}(x)|\}\geq u

for any 𝕥U\mathbb{t}\in U, we have

(5.4) λμ𝕥(σ)logu.\lambda^{\mathbb{t}}_{\mu}(\sigma)\leq-\log u.

Apply Theorem 1.4 to the projective measures {ν𝕥}𝕥U\{\nu_{\mathbb{t}}\}_{\mathbb{t}\in U} originated from the family of infinite parabolic iterated function systems

{S𝕥={si𝕥:XX}iΓX(ϑ,V,γ,u,M)}𝕥U\big{\{}S^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}}\in\Gamma_{X}(\vartheta,V,\gamma,u,M)\big{\}}_{\mathbb{t}\in U}

satisfying the continuity and transversality condition (note that ΓX(ϑ,V,γ,u,M)ΓX(ϑ)\Gamma_{X}(\vartheta,V,\gamma,u,M)\subset\Gamma_{X}(\vartheta)), the two conclusions follow instantly in virtue of (5.4), hμ(σ)=h_{\mu}(\sigma)=\infty and Lemma 4.7.

6. Dimensional estimates of the exceptional parameters

This section is devoted to estimation on the upper bound of the local (global) Hausdorff dimension of the exceptional parameters. We mean to prove Theorem 1.8 here. We start by showing several preceding results on the limit behaviours of the (families of truncated) finite sub-systems of an (family of) infinite PIFS.

6.1 Lemma.

Let

S={si:XX}iΓX(ϑ)S=\{s_{i}:X\rightarrow X\}_{i\in\mathbb{N}}\in\Gamma_{X}(\vartheta)

be a PIFS. Let μ\mu be a finite measure on \mathbb{N}^{\infty} with its concentrating measures {μn}n\{\mu_{n}\}_{n\in\mathbb{N}} on {n}n\{\mathbb{N}_{n}^{\infty}\}_{n\in\mathbb{N}} respectively. Considering the sequence of Lyapunov exponents {λμn(σ)}n=1\{\lambda_{\mu_{n}}(\sigma)\}_{n=1}^{\infty} of the truncated systems {Sn={si:XX}in}n\big{\{}S_{n}=\{s_{i}:X\rightarrow X\}_{i\in\mathbb{N}_{n}}\big{\}}_{n\in\mathbb{N}}, we have

limnλμn(σ)=λμ(σ)\lim_{n\rightarrow\infty}\lambda_{\mu_{n}}(\sigma)=\lambda_{\mu}(\sigma).

Proof.

First we represent λμ(σ)\lambda_{\mu}(\sigma) as

λμ(σ)=i=1[i]log|si(πσ(ω))|dμ(ω)\lambda_{\mu}(\sigma)=\sum_{i=1}^{\infty}-\int_{[i]}\log|s_{i}^{\prime}(\pi\circ\sigma(\omega))|d\mu(\omega).

Note that

(6.1) [i]log|si(πσ(ω))|dμ(ω)<-\int_{[i]}\log|s_{i}^{\prime}(\pi\circ\sigma(\omega))|d\mu(\omega)<\infty

for any 1i<1\leq i<\infty. We distinguish the case λμ(σ)=\lambda_{\mu}(\sigma)=\infty from the case λμ(σ)<\lambda_{\mu}(\sigma)<\infty. If λμ(σ)=\lambda_{\mu}(\sigma)=\infty, we will claim that for any a>0a>0, there exists an integer NaN_{a} large enough, such that for any n>Nan>N_{a}, we have

λμn(σ)>a\lambda_{\mu_{n}}(\sigma)>a.

This means limnλμn(σ)=\lim_{n\rightarrow\infty}\lambda_{\mu_{n}}(\sigma)=\infty. To see this, since λμ(σ)=\lambda_{\mu}(\sigma)=\infty, for any a>0a>0, there exists NN_{*} large enough, such that

i=1N[i]log|si(πσ(ω))|dμ(ω)>2a\sum_{i=1}^{N_{*}}-\int_{[i]}\log|s_{i}^{\prime}(\pi\circ\sigma(\omega))|d\mu(\omega)>2a.

Now choose some small 0<ϵ<a0<\epsilon<a, due to 6.1, for any 1iN1\leq i\leq N_{*}, we can find an integer NiN_{i} large enough, such that

ai:=|[i]log|si(πσ(ω))|dμ(ω)[i]log|si(πσ(ω))|dμn(ω)|<ϵ2ia_{i}:=\big{|}\int_{[i]}\log|s_{i}^{\prime}(\pi\circ\sigma(\omega))|d\mu(\omega)-\int_{[i]}\log|s_{i}^{\prime}(\pi\circ\sigma(\omega))|d\mu_{n}(\omega)\big{|}<\cfrac{\epsilon}{2^{i}}

for any n>Nin>N_{i}. Now let Na=max{Ni}i=1NN_{a}=\max\{N_{i}\}_{i=1}^{N_{*}}, we have

i=1N[i]log|si(πσ(ω))|dμn(ω)i=1N[i]log|si(πσ(ω))|dμ(ω)i=1Nai2ai=1Nϵ2i>a\begin{array}[]{ll}&\sum_{i=1}^{N_{*}}-\int_{[i]}\log|s_{i}^{\prime}(\pi\circ\sigma(\omega))|d\mu_{n}(\omega)\\ \geq&\sum_{i=1}^{N_{*}}-\int_{[i]}\log|s_{i}^{\prime}(\pi\circ\sigma(\omega))|d\mu(\omega)-\sum_{i=1}^{N_{*}}a_{i}\\ \geq&2a-\sum_{i=1}^{N_{*}}\cfrac{\epsilon}{2^{i}}\\ >&a\end{array}

for any n>Nan>N_{a}, which justifies the claim.

The case λμ(σ)<\lambda_{\mu}(\sigma)<\infty can be dealt with in a similar way, which is left to the interested readers. ∎

Similar to Remark 4.8, lemma 6.1 holds for any sequence of measures converging to the measure μ\mu under the setwise topology on ^(X)\mathcal{\hat{M}}(X), not only for the sequence of its concentrating measures, for any topological ambient space XX.

6.2 Lemma.

Let GUG\subset U be a subset. Consider a sequences of real functions {fn:G}n=1\{f_{n}:G\rightarrow\mathbb{R}\}_{n=1}^{\infty} such that

fn(𝕥)f(𝕥)f_{n}(\mathbb{t})\rightarrow f(\mathbb{t})

as nn\rightarrow\infty for some f(𝕥):Gf(\mathbb{t}):G\rightarrow\mathbb{R} at any 𝕥G\mathbb{t}\in G. Now if

HD({𝕥G:fn(𝕥)<0})hnHD(\{\mathbb{t}\in G:f_{n}(\mathbb{t})<0\})\leq h_{n},

for a sequence of reals {hn}n=1\{h_{n}\}_{n=1}^{\infty}, then we have

HD({𝕥G:f(𝕥)<0})lim supnhnHD(\{\mathbb{t}\in G:f(\mathbb{t})<0\})\leq\limsup_{n\rightarrow\infty}h_{n}.

Proof.

Note that since limnfn(𝕥)f(𝕥)\lim_{n\rightarrow\infty}f_{n}(\mathbb{t})\rightarrow f(\mathbb{t}) at any time 𝕥G\mathbb{t}\in G, then

{𝕥G:f(𝕥)<0}k=1n=k{𝕥G:fn(𝕥)<0}\{\mathbb{t}\in G:f(\mathbb{t})<0\}\subset\cap_{k=1}^{\infty}\cup_{n=k}^{\infty}\{\mathbb{t}\in G:f_{n}(\mathbb{t})<0\}.

This means

HD({𝕥G:f(𝕥)<0})inf{HD(n=k{𝕥G:fn(𝕥)<0})}k=1inf{sup{hn}n=k}k=1=lim supnhn.\begin{array}[]{ll}&HD(\{\mathbb{t}\in G:f(\mathbb{t})<0\})\\ \leq&\inf\big{\{}HD(\cup_{n=k}^{\infty}\{\mathbb{t}\in G:f_{n}(\mathbb{t})<0\})\big{\}}_{k=1}^{\infty}\\ \leq&\inf\big{\{}\sup\{h_{n}\}_{n=k}^{\infty}\big{\}}_{k=1}^{\infty}\\ =&\limsup_{n\rightarrow\infty}h_{n}.\end{array}

Be careful that Lemma 6.2 will not be true if the two << are both substituted by \leq in it. Now we are well prepared to prove Theorem 1.8.

Proof of Theorem 1.8:

Proof.

Since μ\mu is ergodic and 𝕐\mathbb{Y} is independent under the law μ\mu, according to Lemma 4.3, consider the sequence of measures {νn,𝕥}𝕥U,n\{\nu_{n,\mathbb{t}}\}_{\mathbb{t}\in U,n\in\mathbb{N}} projected from the ergodic concentrating measures {μn}n\{\mu_{n}\}_{n\in\mathbb{N}} of μ\mu. For fixed nn\in\mathbb{N}, as the family of nn-th truncated finite PIFS

{Sn𝕥={si𝕥:XX}inΓX(ϑ,Vn,γn,un,Mn)}𝕥U\big{\{}S_{n}^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}_{n}}\in\Gamma_{X}(\vartheta,V_{n},\gamma_{n},u_{n},M_{n})\big{\}}_{\mathbb{t}\in U}

satisfying the continuity and strong transversality condition with respect to the time parameter 𝕥U\mathbb{t}\in U for some open neighbourhood VnV_{n} of vv, some 0<γn,un<10<\gamma_{n},u_{n}<1 and Mn>0M_{n}>0, apply [SSU1, Theorem 5.3] to the family of truncated finite PIFS Sn𝕥S_{n}^{\mathbb{t}}, we have

(6.2) HD({𝕥G:dimνn,𝕥<min{hμn(σ)λμn𝕥(σ),α}})min{sup𝕥Ghμn(σ)λμn𝕥(σ),α}+d1HD\Big{(}\big{\{}\mathbb{t}\in G:dim^{*}\nu_{n,\mathbb{t}}<\min\{\frac{h_{\mu_{n}}(\sigma)}{\lambda_{\mu_{n}}^{\mathbb{t}}(\sigma)},\alpha\}\big{\}}\Big{)}\leq\min\Big{\{}\sup_{\mathbb{t}\in G}\frac{h_{\mu_{n}}(\sigma)}{\lambda_{\mu_{n}}^{\mathbb{t}}(\sigma)},\alpha\Big{\}}+d-1

for any nn\in\mathbb{N} and ϵ>0\epsilon>0. Note that

dimν𝕥=limndimνn,𝕥dim^{*}\nu_{\mathbb{t}}=\lim_{n\rightarrow\infty}dim^{*}\nu_{n,\mathbb{t}}

since limnνn,𝕥sν𝕥\lim_{n\rightarrow\infty}\nu_{n,\mathbb{t}}\stackrel{{\scriptstyle s}}{{\rightarrow}}\nu_{\mathbb{t}} and

limnmin{hμn(σ)λμn𝕥(σ),α}\lim_{n\rightarrow\infty}\min\{\cfrac{h_{\mu_{n}}(\sigma)}{\lambda_{\mu_{n}}^{\mathbb{t}}(\sigma)},\alpha\}

exists for any 𝕥G\mathbb{t}\in G since limnhμn(σ)λμn𝕥(σ)\lim_{n\rightarrow\infty}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda_{\mu_{n}}^{\mathbb{t}}(\sigma)} exists everywhere on GG. Moreover,

limn(min{supGhμn(σ)λμn𝕥(σ),α}+d1)=Kα,G\lim_{n\rightarrow\infty}\Big{(}\min\big{\{}\sup_{G}\cfrac{h_{\mu_{n}}(\sigma)}{\lambda_{\mu_{n}}^{\mathbb{t}}(\sigma)},\alpha\big{\}}+d-1\Big{)}=K_{\alpha,G}.

Apply Lemma 6.2 here, we get (6.3).

In the non-exploding case, Theorem 1.8 degenerates into a more simple version as following.

6.3 Corollary.

Let

{S𝕥={si𝕥:XX}iΓX(ϑ)}𝕥U\big{\{}S^{\mathbb{t}}=\{s_{i}^{\mathbb{t}}:X\rightarrow X\}_{i\in\mathbb{N}}\in\Gamma_{X}(\vartheta)\big{\}}_{\mathbb{t}\in U}

be a family of infinite parabolic iterated function systems satisfying the continuity and strong transversality condition with respect to the vector-time parameter 𝕥U\mathbb{t}\in U. For an ergodic probability measure μ\mu on the symbolic space \mathbb{N}^{\infty} satisfying μ(n=1n)=1\mu(\cup_{n=1}^{\infty}\mathbb{N}_{n}^{\infty})=1 and 0<hμ(σ),λμ𝕥(σ)<0<h_{\mu}(\sigma),\lambda^{\mathbb{t}}_{\mu}(\sigma)<\infty at any time 𝕥U\mathbb{t}\in U, if the sequence of infinite random variables in 𝕐\mathbb{Y} under the law μ\mu is independent, then

(6.3) HD({𝕥G:dimν𝕥<min{hμ(σ)λμ𝕥(σ),α}})min{sup𝕥Ghμ(σ)λμ𝕥(σ),α}+d1HD\Big{(}\Big{\{}\mathbb{t}\in G:dim^{*}\nu_{\mathbb{t}}<\min\big{\{}\cfrac{h_{\mu}(\sigma)}{\lambda^{\mathbb{t}}_{\mu}(\sigma)},\alpha\big{\}}\Big{\}}\Big{)}\leq\min\Big{\{}\sup_{\mathbb{t}\in G}\cfrac{h_{\mu}(\sigma)}{\lambda^{\mathbb{t}}_{\mu}(\sigma)},\alpha\Big{\}}+d-1

for any 0<α<10<\alpha<1 and GUG\subset U.

Proof.

Under the assumption 0<hμ(σ),λμ𝕥(σ)<0<h_{\mu}(\sigma),\lambda^{\mathbb{t}}_{\mu}(\sigma)<\infty at any time 𝕥U\mathbb{t}\in U, considering Lemma 4.7 and Lemma 6.1, the result follows instantly from Theorem 1.8.

In the exploding case, that is, either hμ(σ)=h_{\mu}(\sigma)=\infty or λμ𝕥(σ)=\lambda^{\mathbb{t}}_{\mu}(\sigma)=\infty for some 𝕥U\mathbb{t}\in U, the result depends heavily on the asymptotic behaviour of the sequence of concentrating measures of μ\mu, however, some estimations on the Hausdorff dimension of some form of exceptional parameters may still be possible upon reforming Lemma 6.2.

Note that most of the notions and results in this work apply to families of hyperbolic iterated function systems, which are families of iterated function systems constituted by contractive hyperbolic maps only.

References

  • [Ber] V. Berthé, Symbolic dynamics and representations, Les cours du C.I.R.M. Vol. 5 no 1 (2017) 1-16, Cours non^{o} I.
  • [BHP] R. Basu, J. Hermon and Y. Peres, Characterization of cutoff for reversible Markov chains, Ann. Probab. Volume 45, Number 3 (2017), 1448-1487.
  • [BRS] B. Bárány, M. Rams and K. Simon. On the dimension of self-affine sets and measures with overlaps. Proceedings of the American Mathematical Society, 144(10):4427-4440, 2016.
  • [BV1] E. Breuillard and P. Varjú, On the dimension of Bernoulli convolutions, Ann. Probab. Volume 47, Number 4 (2019), 2582-2617.
  • [BV2] E. Breuillard and P. Varjú, Entropy of Bernoulli convolutions and uniform exponential growth for linear groups, Journal d’Analyse Mathématique volume 140, pages 443-481(2020).
  • [Doo] J. Doob, Measure Theory, Graduate Texts in Mathematics, 143, Springer-Verlag New York, 1994.
  • [Fal1] K. Falconer, The Hausdorff dimension of self-affine fractals, Math. Proc. Cambridge Philos. Soc. 103, 1988, 339-350.
  • [Fal2] K. Falconer, Fractal Geometry, Mathematical Foundations and Applications, 3rd Edition, John Wiley & Sons, 2014.
  • [FKZ] E. Feinberg, P. Kasyanov and M. Zgurovsky, Convergence of probability measures and Markov decision models with incomplete information, Proceedings of the Steklov Institute of Mathematics, December 2014, Volume 287, Issue 1, pp 96-117.
  • [Fur] H. Furstenberg, Intersections of Cantor sets and transversality of semigroups, In Problems in analysis (Sympos. Salomon Bochner, Princeton Univ., Princeton, N.J., 1969), pages 41-59. Princeton Univ. Press, Princeton, N.J., 1970. 3, 4, 10.
  • [GR] J. K. Ghosh and R. V. Ramamoorthi, Bayesian Nonparametrics, Springer, New York, 2003.
  • [Hai] M. Hairer, Ergodic Properties of Markov Processes, Lecture given at The University of Warwick in Spring 2006, http://www.hairer.org/notes/Markov.pdf, 2018.
  • [Hay] N. Haydn, The Central Limit Theorem for uniformly strong mixing measures, Stochastics and Dynamics, Vol. 12, No. 4 (2012).
  • [HL] O. Hernandez-Lerma and J. Lasserre, Markov Chains and Invariant Probabilities, Progress in Mathematics, Birkhäuser Basel, 2003.
  • [Hoc1] M. Hochman, On self-similar sets with overlaps and inverse theorems for entropy, Ann. of Math. (2) 180, 773-822, 2014.
  • [Hoc2] M. Hochman, Lectures on dynamical systems and entropy, http://math.huji.ac.il/~mhochman/courses/dynamics2014/notes.5.pdf, June 27, 2014.
  • [Hoc3] M. Hochman, Self similar sets, entropy and additive combinatorics, Geometry and Analysis of Fractals, Volume 88, 2014, pp 225-252.
  • [Hoc4] M. Hochman, Some problems on the boundary of fractal geometry and additive combinatorics, Recent Developments in Fractals and Related Fields, pp 129-174, Editors: J. Barral and S. Seuret, Conference on Fractals and Related Fields III, île de Porquerolles, France, 2015.
  • [HS] M. Hochman and P. Shmerkin, Local entropy averages and projections of fractal measures, Ann. of Math. (2) 175 1001-1059, 2012.
  • [Hut] J. Hutchinson, Fractals and self-similarity, Indiana Univ. Math. J. 30 (1981), no. 5, 713-747.
  • [JW] B. Jessen and A. Wintner, Distribution functions and the Riemann zeta function, Trans. Amer. Math. Soc., 38(1):48-88, 1935.
  • [Kau] R. Kaufman, On Hausdorff dimension of projections, Mathematika 15 (1968), 153-155.
  • [Las] J. Lasserre, On the setwise convergence of sequences of measures, Journal of Applied Mathematics and Stochastic Analysis, 10:2 (1997), 131-136.
  • [LPS] E. Lindenstrauss, Y. Peres and W. Schlag, Bernoulli convolutions and an intermediate value theorem for entropy of K-partitions, J. d’Analyse Math 87 (2002), 337-367.
  • [LY] T. Linder and S. Yüksel, Optimization and convergence of observation channels in stochastic control, SIAM J. Control Optim. Vol. 50, No. 2, pp. 864-887.
  • [Ma] L. Ma, Continuity of measure-dimension mappings, arXiv:2105.05491 [math.DS], 2021.
  • [MMR] P. Mattila, M. Morán and J. M. Rey, Dimension of a measure, Studia Math. 142 (2000), no. 3, 219-233.
  • [PS1] Y. Peres and B. Solomyak, Problems on self-simiar and self-affine sets, an update, in ‘Fractals and Stochastics II, Proceedings of the Greifswald conference’ (Aug 1998), (2000), 95-106.
  • [PS2] Y. Peres and B. Solomyak, Absolute continuity of Bernoulli convolutions, a simple proof, Math. Research Letters 3:2 (1996), 231-239.
  • [PS3] Y. Peres and B. Solomyak, Self-similar measures and intersections of Cantor sets, Trans. Amer. Math. Soc. 350, no. 10 (1998), 4065-4087.
  • [PolS] M. Pollicott and K. Simon, The Hausdorff dimension of λ\lambda-expansions with deleted digits, Trans. Amer. Math. Soc. 347 (1995), 967-983.
  • [PSS] Y. Peres, K. Simon and B. Solomyak, Absolute continuity for random iterated function systems with overlaps, Journal of the London Math. Soc. (2) 74 (2006), 739-756.
  • [Shi] P. Shields, The theory of Bernoulli Shifts, Web Edition 1.01, https://www.impan.pl/~gutman/The%20Theory%20of%20Bernoulli%20Shifts.pdf.
  • [Shm1] P. Shmerkin, On the exceptional set for absolute continuity of Bernoulli convolutions, Geom. Funct. Anal. 24 no. 3(2014),946-958.
  • [Shm2] P. Shmerkin, A modified multifractal formalism for a class of self-similar measures with overlap, Asian J. Math., 9(3):323-348, 2005.
  • [Shm3] P. Shmerkin, Overlapping self-affine sets, Indiana Univ. Math. J., 55(4):1291-1331, 2006.
  • [Shm4] P. Shmerkin, Projections of self-similar and related fractals: a survey of recent developments, Fractal Geometry and Stochastics V, pages 53-74. Editors: Christoph Bandt, Kenneth J. Falconer, and Martina Zähle, Springer, 2015.
  • [SimS1] K. Simon and B. Solomyak, Hausdorff dimension for horseshoes in 3\mathbb{R}^{3}, Ergodic Th. and Dynam. Sys. 19 (1999), 1343-1363.
  • [SimS2] K. Simon and B. Solomyak, On the dimension of self-similar sets, Fractals 10 (2002), 59-65.
  • [Sol1] B. Solomyak, On the random series ±λi\sum\pm\lambda^{i} (an Erdös problem), Annals of Math. 142 (1995), 611-625.
  • [Sol2] B. Solomyak, Measure and dimension for some fractal families, Math. Proc. Cambridge Phil. Soc. 124, no. 3 (1998), 531-546.
  • [Sol3] B. Solomyak, Notes on Bernoulli convolutions, In Fractal Geometry and Applications: A Jubilee of Benoît Mandelbrot. Part 1. Proc. Sympos. Pure Math. 72 (2004) 207-230. Amer. Math. Soc., Providence, RI.
  • [SS1] P. Shmerkin and B. Solomyak, Absolute continuity of self-similar measures, their projections and convolutions, Transactions of the American Math. Society 368 (2016), 5125-5151.
  • [SS2] P. Shmerkin and B. Solomyak, Absolute continuity of complex Bernoulli convolutions, Math. Proc. Cambridge Philos. Soc. 161 (2016), no. 3, 435-453.
  • [SSS] S. Saglietti, P. Shmerkin and B. Solomyak, Absolute continuity of non-homogeneous self-similar measures, Adv. Math., 335:60-110, 2018.
  • [SSU1] K. Simon, B. Solomyak and M. Urbański, Invariant Measures for Parabolic IFS with Overlaps and Random Continued Fractions, Transactions of the American Mathematical Society, Vol. 353, No. 12 (Dec., 2001), pp. 5145-5164.
  • [SSU2] K. Simon, B. Solomyak and M. Urbański, Hausdorff dimension of limit sets for parabolic IFS with overlaps, Pacific J. Math. 201 (2001), 441-478.
  • [Tao] T. Tao, An introduction to measure theory, Graduate Studies in Mathematics, Volume: 126, American Math. Society, 2011.
  • [Tsi] B. Tsirelson, Lecture notes for Probability for mathematicians, Part A: Independence, 3, Infinite independent sequences, http://www.math.tau.ac.il/~tsirel/Courses/ProbForMath/A3.pdf.
  • [Var1] P. P. Varjú, Recent progress on Bernoulli convolutions, In European Congress of Mathematics: Berlin, July 18-22, 2016. Eur. Math. Soc., Zurich.
  • [Var2] P. P. Varjú, Absolute continuity of Bernoulli convolutions for algebraic parameters, J. Amer. Math. Soc. 32 (2019) 351-397.
  • [Var3] P. P. Varjú, On the dimension of Bernoulli convolutions for all transcendental parameters, Annals of Mathematics 189 (2019), 1001-1011.
  • [Wal] P. Walters, An introduction to ergodic theory, Graduate Texts in Mathematics, 79. Springer, New York-Berlin, 1982.
  • [You] L. Young, Dimension, entropy and Lyapunov exponents, Ergodic Theory Dynam. Systems 2(1982), no. 1, 109-124.