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

On invariant measures of "satellite" infinitely renormalizable quadratic polynomials

Genadi Levin Institute of Mathematics, The Hebrew University of Jerusalem, Givat Ram, Jerusalem, 91904, Israel levin@math.huji.ac.il  and  Feliks Przytycki Institute of Mathematics, Polish Academy of Sciences, Śniadeckich St., 8, 00-956 Warsaw, Poland feliksp@impan.pl
Abstract.

Let f(z)=z2+cf(z)=z^{2}+c be an infinitely renormalizable quadratic polynomial and JJ_{\infty} be the intersection of forward orbits of "small" Julia sets of its simple renormalizations. We prove that if ff admits an infinite sequence of satellite renormalizations, then every invariant measure of f:JJf:J_{\infty}\to J_{\infty} is supported on the postcritical set and has zero Lyapunov exponent. Coupled with [14], this implies that the Lyapunov exponent of such ff at cc is equal to zero, which answers partly a question posed by Weixiao Shen.

The first author partially supported by ISF grant 1226/17, Israel. The second author partially supported by National Science Centre, Poland, Grant OPUS 21 "Holomorphic dynamics, fractals, thermodynamic formalism”, 2021/41/B/ST1/00461.

1. Introduction

We consider the dynamics f:f:\mathbb{C}\to\mathbb{C} of a quadratic polynomial. Up to a linear change of coordinates, ff has the form fc(z)=z2+cf_{c}(z)=z^{2}+c for some cc\in\mathbb{C}. In this paper, which is the sequel of [9], we assume that ff is infinitely-renormalizable. Moreover, in the main results we assume that ff has infinitely many "satellite renormalizations", see e.g. [19], or below for definitions. Dynamics, geometry and topology of such system can be very non-trivial, in particular, due to the fact that different renormalization levels are largely independent.

Historically, the first example of infinitely-renormalizsable one-dimensional map was, probably, the Feigenbaum period-doubling quadratic polynomial fcFf_{c_{F}}, where cF=1.4c_{F}=-1.4... [6]. The Julia set of fcFf_{c_{F}} is locally connected [7] as it follows from so-called "complex bounds", a compactness property of renormalizations. This is a key tool since [28], in particular, in proving the Feigenbaum-Coullet-Tresser universality conjecture [28, 20, 15]. Perhaps, more striking for us are Douady-Hubbard’s examples, or alike, of infinitely-renormalizable quadratic polynomials with non-locally connected Julia sets [17, 27, 10, 11, 12, 4, 3]. As for the Feigenbaum polynomial fcFf_{c_{F}}, all the renormalizations of such maps are satellite, although, contrary to fcFf_{c_{F}}, combinatorics is unbounded (which, in turn, implies that those maps cannot have complex bounds [1]).

Dynamics of every holomorphic endomorphism of the Riemann sphere g:^^g:\hat{\mathbb{C}}\to\hat{\mathbb{C}} classically splits ^\hat{\mathbb{C}} into two subsets: the Fatou set F(g)F(g) and its complement the Julia set J(g)J(g), where F(g)F(g) is the maximal (possibly, empty) open set where the sequence of iterates gng^{n}, n=0,1,n=0,1,... forms a normal (i.e., a precompact) family. See e.g. [2], [16] for the Fatou-Julia theory and [26] for a recent survey.

If gg is a polynomial, then the Julia set J(g)J(g) coincides with the boundary of the basin of infinity A()={z|limngn(z)=}A(\infty)=\{z\in\mathbb{C}|\lim_{n\to\infty}g^{n}(z)=\infty\} of gg. The complement A(g)\mathbb{C}\setminus A(g) is called the filled Julia set K(g)K(g) of the polynomial gg. The compact K(g)K(g)\subset\mathbb{C} is connected if and only if it contains all critical points of gg in the complex plane.

A quadratic polynomial fcf_{c} with connected filled Julia set K(f)K(f) is renormalizable if, for some topological disks UVU\Subset V around the critical point 0 of fcf_{c}, and some p2p\geq 2 (period of the renormalization), the restriction F:=fcp:UVF:=f_{c}^{p}:U\to V is a proper branched covering map (called polynomial-like map) of degree 22 and the non-escaping set K(F)={zU:Fn(z)U for all n1}K(F)=\{z\in U:F^{n}(z)\in U\mbox{ for all }n\geq 1\} (called the filled Julia set of the polynomial-like map FF) is connected. The map F:UVF:U\to V is then a renormalization of fcf_{c} and the set K(F)K(F) is a "small" (filled) Julia set of fcf_{c}. By the theory of polynomial-like mappings [5], there is a quasiconformal homeomorphism of \mathbb{C}, which is conformal on K(F)K(F), that conjugates FF on a neighborhood of K(F)K(F) to a uniquely defined another quadratic polynomial fcf_{c^{\prime}} with connected filled Julia set. If fcf_{c^{\prime}} is renormalizable by itself, then fcf_{c} is called twice renormalizable, etc. If fcf_{c} admits infinitely many renormalizations, it is called infinitely-renormalizable. The renormalization F=fcpF=f^{p}_{c} is simple if any two sets fi(K(f))f^{i}(K(f)), fj(K(F))f^{j}(K(F)), 0i<jp10\leq i<j\leq p-1, are either disjoint or intersect each other at a unique point which does not separate either of them. A simple renormalization fpnf^{p_{n}} is called primitive if all sets fi(Kn)f^{i}(K_{n}), i=0,,pn1i=0,\cdots,p_{n}-1, are disjoints and satellite otherwise.

To state our main results, Theorems 1.1, let f(z)=z2+cf(z)=z^{2}+c be infinitely renormalizable. Then its Julia set J=J(f)J=J(f) coincides with the filled Julia set K(f)K(f) and is a nowhere dense compact full connected subset of \mathbb{C}. Let 1=p0<p1<<pn<1=p_{0}<p_{1}<...<p_{n}<... be the sequence of consecutive periods of simple renormalizations of ff and Jn0J_{n}\ni 0 denote the "small" Julia set of the nn-renormalization (where J0=JJ_{0}=J). Then pn+1/pnp_{n+1}/p_{n} is an integer, fpn(Jn)=Jnf^{p_{n}}(J_{n})=J_{n}, for any nn, and ff-orbits of JnJ_{n},

orb(Jn)=j0fj(Jn)=j=0pn1fj(Jn),orb(J_{n})=\cup_{j\geq 0}f^{j}(J_{n})=\cup_{j=0}^{p_{n}-1}f^{j}(J_{n}),

n=0,1,n=0,1,..., form a strictly decreasing sequence of compact subsets of \mathbb{C}. Let

J=n0orb(Jn)J_{\infty}=\cap_{n\geq 0}orb(J_{n})

be the intersection of the orbits of the "small" Julia sets JnJ_{n}. For every nn, repelling periodic orbits of ff are dense in orb(Jn)orb(J_{n}), while each component of JJ_{\infty} is wandering. In particular, JJ_{\infty} contains no periodic points of ff.

Let

P={fn(0)|n=1,2,}¯P=\overline{\{f^{n}(0)|n=1,2,...\}}

be the postcritical set of ff. Clearly,

PJ.P\subset J_{\infty}.

Moreover, the critical point 0 is recurrent, hence,

P=ω(0),P=\omega(0),

where ω(z)\omega(z) is the omega-limit set of a point zJz\in J.

We prove in [9] that JJ_{\infty} cannot contain any hyperbolic set. On the other hand, a hyperbolic set of a rational map always carries an invariant measure with a positive Lyapunov exponent. So a generalization of [9] would be that JJ_{\infty} never carries such a measure. Here we prove this generalization for a class of "satellite" infinitely-renormalizable quadratic polynomials:

Theorem 1.1.

Suppose that f(z)=z2+cf(z)=z^{2}+c admits infinitely many satellite renormalizations. Then f:JJf:J_{\infty}\to J_{\infty} has no invariant probability measure with positive Lyapunov exponent.

Remark 1.1.

Conjecturally, the same conclusion should hold for any infinitely-renormalizable f(z)=z2+cf(z)=z^{2}+c. One can show this assuming that the Julia set of ff is locally-connected (e.g., this is the case if ff admits complex bounds). Indeed, if f:JJf:J_{\infty}\to J_{\infty} had an invariant probability measure with positive Lyapunov exponent, then, taking a typical point of this measure and repeating the proof of [14], Corollary 5.5, we would conclude that the Julia set of ff is not locally-connected (in fact, JJ_{\infty} contains a non-trivial continuum). Thus the only open case remains when ff has only finitely many satellite renormalizations and JJ_{\infty} contains a non-trivial continuum.

Remark 1.2.

For every rational map f:f:\mathbb{C}\to\mathbb{C} (in particular, quadratic polynomial) and every invariant probability measure supported on Julia set Lyapunov exponents are non-negative, see [22] (compare a remark preceding Corollary 1.3). On the other hand, if ff is hyperbolic or non-uniformly hyperbolic (topologically Collet-Eckmann) Lyapunov exponents for all invariant probability measures supported on Julia set are positive and bounded away from 0, see [24].

Let us comment on the behavior of the restriction map f:JJf:J_{\infty}\to J_{\infty} where ff as in Theorem 1.1. First, by [9], the postcritical set PP must intersect the omega-limit set ω(x)\omega(x) of each xJx\in J_{\infty}. At the same time, dynamics and topology of the further restriction f:PPf:P\to P can vary. Indeed, there are infinitely renormalizable quadratic polynomials ff with all renormalizations being of satellite type such that at least one of the following holds111A more complete description of f:PPf:P\to P should follow from the methods developed in [3].:

(1) f:PPf:P\to P is not minimal. This case happens in Douady-Hubbard’s type examples. Indeed, by the basic construction [17], JJ_{\infty} then contains a closed invariant set XX (which is the limit set for the collection of α\alpha-fixed points of renormalizations) such that 0X0\notin X. By [9], XPX\cap P is non-empty. Thus XPX\cap P is an invariant non-empty proper compact subset of PP.

(2) PP is a so-called "hairy" Cantor set, in particular, PP contains uncountably many non-trivial continua. This case takes place following [3].

(3) PP is a Cantor set and f:PPf:P\to P is minimal; this happens whenever ff either admits complex bounds (which then imply J=PJ_{\infty}=P) or is robust [19]222The ”robustness” can happen without ”complex bounds” as it follows from [3] combined with [1].. Under either of the two conditions, f:PPf:P\to P is a minimal homeomorphism, which is topologically conjugate to xx+1x\mapsto x+1 acting on the projective limit of the sequence of groups {/pn}n=1\{\mathbb{Z}/p_{n}\mathbb{Z}\}_{n=1}^{\infty}; in particular, f:PPf:P\to P (hence, also f:JJf:J_{\infty}\to J_{\infty}, as it follows from the next Corollary 1.3) is uniquely ergodic in this case.

Theorem 1.1 yields the following dichotomy about the measurable dynamics of f:JJf:J\to J on the Julia set JJ of ff. Recall that, by [22], any invariant probability measure on the Julia set of a rational function has non-negative exponents.

Corollary 1.3.

Let μ\mu be an invariant probability ergodic measure of f:JJf:J\to J. Then either

  1. (i)

    supp(μ)J=\operatorname{supp}(\mu)\cap J_{\infty}=\emptyset and its Lyapunov exponent χ(μ)>0\chi(\mu)>0,

    or

  2. (ii)

    supp(μ)P\operatorname{supp}(\mu)\subset P and χ(μ)=0\chi(\mu)=0.

In particular, the set JPJ_{\infty}\setminus P is "measure invisible", see also Proposition 6.1 which is a somewhat stronger version of Corollary 1.3.

Corollary 1.4.

If ff admits infinitely many satellite renormalizations, then

(1.1) lim supn1nlog|(fn)(x)|0 for any xJ,\limsup_{n\to\infty}\frac{1}{n}\log|(f^{n})^{\prime}(x)|\leq 0\mbox{ for any }x\in J_{\infty},

and

(1.2) limn1nlog|(fn)(c)|=0.\lim_{n\to\infty}\frac{1}{n}\log|(f^{n})^{\prime}(c)|=0.

For the proof of Corollaries 1.3-1.4, see Sect. 6. The proof of Theorem 1.1 occupies sections 2-5.

As in [9], we use heavily a general result of [23] on the accessibility although the main idea of the proof is different. Indeed, in [9] we utilize the fact that the map cannot be one-to-one on an infinite hyperbolic set. At the present paper, to prove Theorem 1.1 we assign, loosely speaking, an external ray to a typical point of a hypothetical measure with positive exponent such that the family of such rays is invariant and has a controlled geometry. Given a satellite renormalization fpnf^{p_{n}} we use the measure and the above family of rays to choose a point xx and build a special domain that covers a "small" Julia set Jn,xxJ_{n,x}\ni x such that there is a univalent pullback of the domain by fpnf^{p_{n}} along the renormalization that enters into itself, leading to a contradiction. The choice of xx is ’probabilistic’, i.e., made from sets of positive measure, and the construction of the domain differs substantially depending on whether all satellite renormalizations of ff are doubling or not.

Acknowledgment. The conclusion (1.2) of Corollary 1.4 that the Lyapunov exponent at the critical value equals zero answers partly a question by Weixiao Shen333 Shen asked the following question, in relation with Corollary 5.5 of [14]: Is it possible that the upper Lyapunov exponent at a critical value vv of a polynomial gg is positive assuming that gg is infinitely-renormalizable around vv? See also Remark 1.1 which inspired the present work as well as the prior one [9]. The authors thank the referee for careful reading the paper and many helpful comments.

2. Preliminaries

Here we collect, for further references and use throughout the paper, necessary notations and general facts. (A)-(D) are slightly adapted versions of (A)-(D) in Sect. 2, [9] which are either well-known [19], [18], or are proved here.

Let f(z)=z2+cf(z)=z^{2}+c be infinitely renormalizable. We keep the notations of the Introduction.

(A). Let GG be the Green function of the basin of infinity A()={z|fn(z),n}A(\infty)=\{z|f^{n}(z)\to\infty,n\to\infty\} of ff with the standard normalization at infinity G(z)=ln|z|+O(1/|z|)G(z)=\ln|z|+O(1/|z|). The external ray RtR_{t} of argument t𝐒𝟏=/t\in{\bf S^{1}}=\mathbb{R}/\mathbb{Z} is a gradient line to the level sets of GG that has the (asymptotic) argument tt at \infty. G(z)G(z) is called the (Green) level of zA()z\in A(\infty) and the unique tt such that zRtz\in R_{t} is called the (external) argument (or angle) of zz. A point zJ(f)z\in J(f) is accessible if there is an external ray RtR_{t} which lands at (i.e., converges to) zz. Then tt is called an (external) argument (angle) of zz.

Let σ:𝐒𝟏𝐒𝟏\sigma:{\bf S^{1}}\to{\bf S^{1}} be the doubling map σ(t)=2t(mod1)\sigma(t)=2t(\mod 1). Then f(Rt)=Rσ(t)f(R_{t})=R_{\sigma(t)}.

Every point aa of a repelling cycle OaO_{a} of period pp is the landing point of an equal number vv, 1v<1\leq v<\infty, of external rays where vv coincides with the number of connected components of J(f){a}J(f)\setminus\{a\}. Their arguments are permuted by σp\sigma^{p} according to a rational rotation number r/qr/q (written in the lowest term); v/qv/q is the number of cycles of rays landing at aa. If v2v\geq 2, there is an alternative [18]:

r/q=0/1r/q=0/1, then v=2v=2 so that each of two external ray landing at aa is fixed by fpf^{p},

r/q0/1r/q\neq 0/1, i.e., q2q\geq 2, then v=qv=q, i.e., the arguments of qq rays landing at aa form a single cycle of σp\sigma^{p}.

(B). All periodic points of ff are repelling. Given a small Julia set JnJ_{n} containing 0, sets fj(Jn)f^{j}(J_{n}), 0j<pn0\leq j<p_{n}, are called small Julia sets of level nn. Each fj(Jn)f^{j}(J_{n}) contains pn+1/pn2p_{n+1}/p_{n}\geq 2 small Julia sets of level n+1n+1. We have Jn=JnJ_{n}=-J_{n}. Since all renormalizations are simple, for j0j\neq 0, the symmetric companion fj(Jn)-f^{j}(J_{n}) of fj(Jn)f^{j}(J_{n}) can intersect the orbit orb(Jn)=j=0pn1fj(Jn)orb(J_{n})=\cup_{j=0}^{p_{n}-1}f^{j}(J_{n}) of JnJ_{n} only at a single point which is periodic. On the other hand, since only finitely many external rays converge to each periodic point of ff, the set JJ_{\infty} contains no periodic points. In particular, each component KK of JJ_{\infty} is wandering, i.e., fi(K)fj(K)=f^{i}(K)\cap f^{j}(K)=\emptyset for all 0i<j<0\leq i<j<\infty. All this implies that {x,x}J\{x,-x\}\subset J_{\infty} if and only if xK0:=n=1Jnx\in K_{0}:=\cap_{n=1}^{\infty}J_{n}.

Given xJx\in J_{\infty}, for every nn, let jn(x)j_{n}(x) be the unique j{0,1,,pn1}j\in\{0,1,\cdots,p_{n}-1\} such that xfjn(x)(Jn)x\in f^{j_{n}(x)}(J_{n}). Let Jn,x=fjn(x)(Jn)J_{n,x}=f^{j_{n}(x)}(J_{n}) be a small Julia set of level nn containing xx and Kx=n0Jn,xK_{x}=\cap_{n\geq 0}J_{n,x}, a component of JJ_{\infty} containing xx.

In particular, K0=n0JnK_{0}=\cap_{n\geq 0}J_{n} is the component of JJ_{\infty} containing 0 and Kc=n=1f(Jn)K_{c}=\cap_{n=1}^{\infty}f(J_{n}), the component containing cc.

Note that either pnjn(x)p_{n}-j_{n}(x)\to\infty as nn\to\infty or pnjn(x)=Np_{n}-j_{n}(x)=N for some N0N\geq 0 and all nn, that is, fN(x)K0f^{N}(x)\in K_{0}. This is so since the sequence of the sets JnJ_{n} is non-increasing, hence Jn,xJ_{n,x} non-increasing, hence pnjn(x)p_{n}-j_{n}(x) (the time to reach JnJ_{n}) non-decreasing.

The map f:KxKf(x)f:K_{x}\to K_{f(x)} is two-to-one if x=0x=0 and one-to-one otherwise. Moreover, for every yJy\in J_{\infty}, f1(y)Jf^{-1}(y)\cap J_{\infty} consists of two points if yKcy\in K_{c} and consists of a single point otherwise. Denote

J=Jj=fj(K0).J_{\infty}^{\prime}=J_{\infty}\setminus\cup_{j=-\infty}^{\infty}f^{j}(K_{0}).

We conclude that:

f:JJf:J_{\infty}^{\prime}\to J_{\infty}^{\prime} is a homeomorphism. Given xJx\in J_{\infty}^{\prime} and m>0m>0, denote xm=fm(x)x_{m}=f^{m}(x) and

xm=f|Jm(x),x_{-m}=f|_{J_{\infty}^{\prime}}^{-m}(x),

that is, the only point fm(x)Jf^{-m}(x)\cap J_{\infty}.

(C). Given n0n\geq 0, the map fpn:f(Jn)f(Jn)f^{p_{n}}:f(J_{n})\to f(J_{n}) has two fixed points: the separating fixed point αn\alpha_{n} (that is, f(Jn){αn}f(J_{n})\setminus\{\alpha_{n}\} has at least two components) and the non-separating βn\beta_{n} (so that f(Jn)βnf(J_{n})\setminus\beta_{n} has a single component).

For every n>0n>0, there are 0<tn<t~n<10<t_{n}<\tilde{t}_{n}<1 such that two rays RtnR_{t_{n}} and Rt~nR_{\tilde{t}_{n}} land at the non-separating fixed point βnf(Jn)\beta_{n}\in f(J_{n}) of fpnf^{p_{n}} and the component Ωn\Omega_{n} of 𝐂(RtnRt~nβn){\bf C}\setminus(R_{t_{n}}\cup R_{\tilde{t}_{n}}\cup\beta_{n}) which does not contain 0 has two characteristic propertiers [18]:

(i) Ωn\Omega_{n} contains cc and is disjoint with the forward orbit of βn\beta_{n},

(ii) for every 1j<pn1\leq j<p_{n}, consider arguments (angles) of external rays which land at fj1(βn)f^{j-1}(\beta_{n}). The angles split 𝐒𝟏{\bf S^{1}} into finitely many arcs. Then the length of any such arc is bigger than the length of the arc

Sn,1=[tn,t~n]={t:RtΩn}.S_{n,1}=[t_{n},\tilde{t}_{n}]=\{t:R_{t}\subset\Omega_{n}\}.

Denote

tn=tn+t~ntn2pn,t~n=t~nt~ntn2pn.t_{n}^{\prime}=t_{n}+\frac{\tilde{t}_{n}-t_{n}}{2^{p_{n}}},\ \ \tilde{t}_{n}^{\prime}=\tilde{t}_{n}-\frac{\tilde{t}_{n}-t_{n}}{2^{p_{n}}}.

The rays RtnR_{t_{n}^{\prime}}, Rt~nR_{\tilde{t}_{n}^{\prime}} land at a common point βnfpn(βn)Ωn\beta_{n}^{\prime}\in f^{-p_{n}}(\beta_{n})\cap\Omega_{n}. Introduce an (unbounded) domain UnU_{n} with the boundary to be two curves RtnRt~nβnR_{t_{n}}\cup R_{\tilde{t}_{n}}\cup\beta_{n} and RtnRt~nβnR_{t_{n}^{\prime}}\cup R_{\tilde{t}_{n}^{\prime}}\cup\beta_{n}^{\prime}. Then cUnc\in U_{n} and fpn:UnΩnf^{p_{n}}:U_{n}\to\Omega_{n} is a two-to-one branched covering. Also,

f(Jn)={z:fkpn(z)U¯n,G(fkpn(z))<10,k=0,1,}.f(J_{n})=\{z:f^{kp_{n}}(z)\in\overline{U}_{n},G(f^{kp_{n}}(z))<10,k=0,1,...\}.

Let

sn,1=[tn,tn][t~n,t~n]s_{n,1}=[t_{n},t_{n}^{\prime}]\cup[\tilde{t}_{n}^{\prime},\tilde{t}_{n}]

so that sn,1Sn,1s_{n,1}\subset S_{n,1} and argument of any ray to f(Jn)f(J_{n}) lies in sn,1s_{n,1}.

Let us iterate this construction. Given 1jpn1\leq j\leq p_{n}, let Sn,jS_{n,j} be one of the two arcs of 𝐒𝟏{\bf S^{1}} with end points

tn,j=σj1(tn),t~n,j=σj1(t~n)t_{n,j}=\sigma^{j-1}(t_{n}),\tilde{t}_{n,j}=\sigma^{j-1}(\tilde{t}_{n})

such that arguments of any ray to fj(Jn)f^{j}(J_{n}) lies in Sn,jS_{n,j}. Let

sn,j=σj1(sn,1)=[tn,j,tn,j][t~n,j,t~n,j]s_{n,j}=\sigma^{j-1}(s_{n,1})=[t_{n,j},t_{n,j}^{\prime}]\cup[\tilde{t}_{n,j}^{\prime},\tilde{t}_{n,j}]

where tn,j=σj1(tn),t~n,j=σj1(t~n)t_{n,j}^{\prime}=\sigma^{j-1}(t_{n}^{\prime}),\tilde{t}_{n,j}^{\prime}=\sigma^{j-1}(\tilde{t}_{n}^{\prime}). Then

sn,jSn,js_{n,j}\subset S_{n,j}

and argument of any ray to fj(Jn)f^{j}(J_{n}) lies in fact in sn,js_{n,j}. Note that

(2.1) tn,jtn,j=t~n,jt~n,j=t~ntn2pnj+1<t~ntn<1/2.t_{n,j}^{\prime}-t_{n,j}=\tilde{t}_{n,j}-\tilde{t}_{n,j}^{\prime}=\frac{\tilde{t}_{n}-t_{n}}{2^{p_{n}-j+1}}<\tilde{t}_{n}-t_{n}<1/2.

So σj1:sn,1sn,j\sigma^{j-1}:s_{n,1}\to s_{n,j} is a homeomorphism and sn,js_{n,j} has two components (’windows’) [tn,j,tn,j][t_{n,j},t_{n,j}^{\prime}] and [t~n,j,t~n,j][\tilde{t}_{n,j}^{\prime},\tilde{t}_{n,j}] of equal length.

Let Un,j=fj1(Un)U_{n,j}=f^{j-1}(U_{n}) and βn,j=fj1(βn)\beta_{n,j}=f^{j-1}(\beta_{n}). The domain Un,jU_{n,j} is bounded by two rays Rtn,jRt~n,jR_{t_{n,j}}\cup R_{\tilde{t}_{n,j}} converging to βn,j\beta_{n,j} and completed by βn,j\beta_{n,j} along with two rays Rtn,jRt~n,jR_{t_{n,j}^{\prime}}\cup R_{\tilde{t}_{n,j}^{\prime}} completed by their common limit point fj1(βn)f^{j-1}(\beta_{n}^{\prime}) where tn,j=σj1(tn),t~n,j=σj1(t~n)t_{n,j}^{\prime}=\sigma^{j-1}(t_{n}^{\prime}),\tilde{t}_{n,j}^{\prime}=\sigma^{j-1}(\tilde{t}_{n}^{\prime}).

By (i)-(ii), for a fixed nn, domains Un,jU_{n,j}, 1jpn1\leq j\leq p_{n}, are pairwise disjoint.

Let Un,jpnU_{n,j-p_{n}} be a component of f(pnj)(Un)f^{-(p_{n}-j)}(U_{n}) which is contained in Un,jU_{n,j}. Then

(2.2) fpn:Un,jpnUn,jf^{p_{n}}:U_{n,j-p_{n}}\to U_{n,j}

is a two-to-one branched covering and

fj1(Jn)={z:fkpn(z)U¯n,jpn,G(fkpn(z))<10,k=0,1,}.f^{j-1}(J_{n})=\{z:f^{kp_{n}}(z)\in\overline{U}_{n,j-p_{n}},G(f^{kp_{n}}(z))<10,k=0,1,...\}.

Let sn,j1s^{1}_{n,j} be the set of arguments of rays entering Un,jpnU_{n,j-p_{n}}. Then sn,j1s^{1}_{n,j} consists of 44 components so that σpn\sigma^{p_{n}} map homeomorphically each of these components onto one of the ’windows’ of sn,js_{n,j}.

Furthermore, let

Ωn,j=fj1(Ωn).\Omega_{n,j}=f^{j-1}(\Omega_{n}).

Unlike the map (2.2), the map

(2.3) fpn:Un,jΩn,jf^{p_{n}}:U_{n,j}\to\Omega_{n,j}

is a two-to-one branched covering only assuming fj1:ΩnΩn,jf^{j-1}:\Omega_{n}\to\Omega_{n,j} is a homeomorphism, which holds if and only if σj1:Sn,1σj1(Sn,1)\sigma^{j-1}:S_{n,1}\to\sigma^{j-1}(S_{n,1}) is a homeomorphism. In the latter case,

σj1(Sn,1)=Sn,j.\sigma^{j-1}(S_{n,1})=S_{n,j}.

Primitive vs satellite renormalizations. Let n2n\geq 2 and rn/qnr_{n}/q_{n} be the rotation number of βn\beta_{n}. The next claim is well-known, we include the proof for reader’s convenience.

Lemma 2.1.
  1. (1)

    the renormalization fpnf^{p_{n}} is primitive if and only if rn/qn=0/1r_{n}/q_{n}=0/1, the period of βn\beta_{n} is pnp_{n} and βn\beta_{n} is the landing point of exactly two rays and they are fixed by fpnf^{p_{n}},

  2. (2)

    points βn\beta_{n}, n=1,2,n=1,2,\cdots are all different,

  3. (3)

    fpnf^{p_{n}} is satellite if and only if the α\alpha-fixed point αn1\alpha_{n-1} of fpn1:f(Jn1)f(Jn1)f^{p_{n-1}}:f(J_{n-1})\to f(J_{n-1}) coincides with the β\beta-fixed point βn\beta_{n} of fpn:f(Jn)f(Jn)f^{p_{n}}:f(J_{n})\to f(J_{n}). In particular, j=0qn1fjpn1(f(Jn))f(Jn1)\cup_{j=0}^{q_{n}-1}f^{jp_{n-1}}(f(J_{n}))\subset f(J_{n-1}) and pn=qnpn1p_{n}=q_{n}p_{n-1}. Moreover, each of pn1p_{n-1} points of the orbit of βn\beta_{n} is the landing points of precisely qnq_{n} rays which are permuted by fpn1f^{p_{n-1}} according to the rotation number rn/qnr_{n}/q_{n}. Completed by the landing point they split \mathbb{C} into qnq_{n} "sectors" such that the closure of each of them contains a unique "small" Julia set of level nn sharing a common point with the boundary of the "sector".

Proof.

(1). fpnf^{p_{n}} is satellite if and only if f(Jn)f(J_{n}) meets at βn\beta_{n} some other iterate of JnJ_{n}, hence, rn/qn0r_{n}/q_{n}\neq 0, and vice versa. (2). assume β:=βn=βm\beta:=\beta_{n}=\beta_{m} for some 0n<m0\leq n<m. As pn<pmp_{n}<p_{m}, the period of βm\beta_{m} is smaller than pnp_{n}. It follows that f(Jn)f(J_{n}) contains two small Julia sets of level mm that meet at β\beta, hence, β\beta separates f(Jn)f(J_{n}), a contradiction as βn\beta_{n} does not. (3). By (1), fpnf^{p_{n}} is satellite if and only if rn/qn0r_{n}/q_{n}\neq 0. Let p~n1=pn/qn\tilde{p}_{n-1}=p_{n}/q_{n}. Then p~n1\tilde{p}_{n-1} is an integer and is equal to the period of βn\beta_{n}. It follows that pnp_{n} sets f(Jn),f2(Jn),,fpn(Jn)f(J_{n}),f^{2}(J_{n}),\cdots,f^{p_{n}}(J_{n}) are split into p~n1\tilde{p}_{n-1} connected closed subsets EiE_{i}, i=1,,p~~n1i=1,\cdots,\tilde{\tilde{p}}_{n-1} where E1=j=0qn1fjp~n1(f(Jn))E_{1}=\cup_{j=0}^{q_{n}-1}f^{j\tilde{p}_{n-1}}(f(J_{n})) and Ei=fi1(E1)E_{i}=f^{i-1}(E_{1}), i=1,2,,p~n1i=1,2,\cdots,\tilde{p}_{n-1}. Moreover, 0Epn10\in E_{p_{n-1}} and f(Ei)=Ei+1f(E_{i})=E_{i+1}, i=1,,p~n11i=1,\cdots,\tilde{p}_{n-1}-1, f(Ep~n1)=E1f(E_{\tilde{p}_{n-1}})=E_{1}. By [19, Theorem 8.5], fp~n1f^{\tilde{p}_{n-1}} is a simple renormalization and EiE_{i}, i=1,,p~n1i=1,\cdots,\tilde{p}_{n-1} are subsets of its p~n1\tilde{p}_{n-1} small Julia sets. Since 1=p0<p1<1=p_{0}<p_{1}<... are all consecutive periods of simple renormalizations, then p~n1=pk\tilde{p}_{n-1}=p_{k} for some k<nk<n. Therefore, βn\beta_{n}-fixed point of fpn:f(Jn)f(Jn)f^{p_{n}}:f(J_{n})\to f(J_{n}) is αk\alpha_{k}-fixed point of fpk:f(Jpk)f(Jpk)f^{p_{k}}:f(J_{p_{k}})\to f(J_{p_{k}}). As all renormalizations are simple, if k<n1k<n-1 that would imply that βn=βn1==βk+1\beta_{n}=\beta_{n-1}=...=\beta_{k+1}, a contradiction with (2). The claim about "sectors" follows since each map fjf^{j} is one-to-one in a neighborhood of βn\beta_{n} and the closure of Ωn\Omega_{n} contains a single "small" Julia set f(Jn)f(J_{n}) of level nn sharing a common point with Ωn\partial\Omega_{n}. ∎

We need a more refined estimate provided the renormalization is not doubling. Assume fpnf^{p_{n}} is satellite so that pn1=pn/qnp_{n-1}=p_{n}/q_{n} with qn2q_{n}\geq 2 and the rotation number of βn\beta_{n} is rn/qn0/1r_{n}/q_{n}\neq 0/1.

Lemma 2.2.

Assume fpnf^{p_{n}} is satellite and qn=pn/pn13q_{n}=p_{n}/p_{n-1}\geq 3, i.e., fpnf^{p_{n}} is not doubling. Then

(2.4) σj1:Sn,1σj1Sn,1 is a homeomorphism for j=1,,pn1(qn2).\sigma^{j-1}:S_{n,1}\to\sigma^{j-1}S_{n,1}\mbox{ is a homeomorphism for }j=1,\cdots,p_{n-1}(q_{n}-2).

In particular, given ζ(0,1/3)\zeta\in(0,1/3), the length of σj1Sn,1\sigma^{j-1}S_{n,1} tends to zero as nn\to\infty uniformly in j=1,,[ζpn]j=1,\cdots,[\zeta p_{n}] (where [x][x] is the integer part of xx\in\mathbb{R}).

Moreover, for every 1jpn1(qn2)1\leq j\leq p_{n-1}(q_{n}-2), Sn,j=σj1(Sn,1)S_{n,j}=\sigma^{j-1}(S_{n,1}) and the map fpn:Un,jΩn,jf^{p_{n}}:U_{n,j}\to\Omega_{n,j} is a two-to-one branched covering such that

fj(Jn)={z:fkpn(z)U¯n,j,G(fkpn(z))<10,k=0,1,}.f^{j}(J_{n})=\{z:f^{kp_{n}}(z)\in\overline{U}_{n,j},G(f^{kp_{n}}(z))<10,k=0,1,...\}.
Proof.

Let g=fpn1:Un1Ωn1g=f^{p_{n-1}}:U_{n-1}\to\Omega_{n-1}. Then gg is a two-to-one covering of degree 22 and the critical value cc.

(1) Recall that sn1,1=[tn1,tn1][t~n1,t~n1]s_{n-1,1}=[t_{n-1},t_{n-1}^{\prime}]\cup[\tilde{t}_{n-1}^{\prime},\tilde{t}_{n-1}] consists of two ’windows’ so that σpn1\sigma^{p_{n-1}} is orientation preserving homeomorphism of either ’window’ onto Sn1,1=[tn1,t~n1]S_{n-1,1}=[t_{n-1},\tilde{t}_{n-1}].

(2) Consider qnq_{n} rays L1,,LqnL_{1},...,L_{q_{n}} to αn1\alpha_{n-1}. The map gg is a local homeomorphism near αn1\alpha_{n-1} which permutes the rays to αn1\alpha_{n-1} according to the rotation number ν:=rn/qn0,1/2\nu:=r_{n}/q_{n}\neq 0,1/2. In particular, gg maps any pair of adjacent rays to αn1\alpha_{n-1} onto another pair of adjacent rays to αn1\alpha_{n-1}.

Refer to caption
Figure 1. qn=3q_{n}=3.

(3) Not all arguments of these rays lie in a single ’window’ II of sn1,1s_{n-1,1} because otherwise, by (1), the set of those arguments would lie in the non-escaping set of an orientation preserving homeomorphism σpn1:ISn,1\sigma^{p_{n-1}}:I\to S_{n,1}, which consists of a fixed point of this map, a contradiction with the fact that qn>1q_{n}>1.

(4) The rays LjL_{j} split Un1U_{n-1} into qnq_{n} disjoint domains UjU^{j}, j=0,1,,qn1j=0,1,...,q_{n}-1. By the "ideal boundary" ^Uj\hat{\partial}{U^{j}} of UjU^{j} we will mean the usual (topological) boundary Uj\partial U^{j} (in our case, the set of boundary rays completed by their landing points) along with the "boundary at infinity" which is the set of arguments of rays entering UjU^{j}. Then define g^\hat{g} on ^Uj\hat{\partial}{U^{j}} to be gg on Uj\partial U^{j} and σpn1\sigma^{p_{n-1}} on the "boundary at infinity" of UjU^{j}.

(5) By (3), one of UjU^{j}, called U0U^{0}, has βn1\beta_{n-1} in its boundary, and another one, called Uqn1U^{q_{n}-1}, has βn1\beta_{n-1}^{\prime} in the boundary. In particular, the boundary of any other UjU^{j}, j0,qn1j\neq 0,q_{n}-1, consists of a pair of adjacent rays to αn1\alpha_{n-1} whose arguments belong to a single ’window’ of sn1,1s_{n-1,1}. Therefore, by (1), the rest of indices j=1,,qn2j=1,...,q_{n}-2 can be ordered in such a way that g^:^Uj^Uj+1\hat{g}:\hat{\partial}{U^{j}}\to\hat{\partial}{U^{j+1}} is a one-to-one map for j=1,,qn3j=1,\cdots,q_{n}-3 (note that the "boundary at infinity" of each UjU^{j}, 1jqn21\leq j\leq q_{n}-2, consists of a single "arc at infinity"). Therefore, g:UjUj+1g:U^{j}\to U^{j+1} is a homeomorphism for j=1,,qn3j=1,...,q_{n}-3. The map g^\hat{g} on ^Uqn2\hat{\partial}{U^{q_{n}-2}} is also a one-to-one map on its image W=g(Uqn2)W=g(U^{q_{n}-2}) where WW is bounded by two adjacent rays to αn1\alpha_{n-1}. WW cannot contain U0U^{0} because otherwise WW would contain βn1\beta_{n-1}^{\prime}, a contradiction. Thus WW must contain βn1\beta_{n-1}^{\prime}. That is, g(Uqn2)g(U^{q_{n}-2}) covers Uqn1U^{q_{n}-1}.

Thus, for j=1,,qn3j=1,\cdots,q_{n}-3, g:UjUj+1g:U^{j}\to U^{j+1} is a homeomorphism, and g:Uqn2Wg:U^{q_{n}-2}\to W is also a homeomorphism where the image W=g(Uqn2)W=g(U^{q_{n}-2}) covers Uqn1U^{q_{n}-1} and has two common rays with the boundary of Uqn1U^{q_{n}-1}.

(6) The critical value cc of gg has a unique preimage by g (the critical point of gg). As cΩnΩn1c\in\Omega_{n}\subset\Omega_{n-1} and Ωn\Omega_{n} is bounded by two adjacent rays to αn1\alpha_{n-1}, cUic\in U^{i} for some i{1,,qn1}i\in\{1,\cdots,q_{n}-1\}. If i>1i>1, then i11i-1\geq 1 while gg would not be a homeomorphism of Ui1U^{i-1} on its image. This shows that cU1=Ωnc\in U^{1}=\Omega_{n}.

Concluding, Uj=gj1(Ωn)U^{j}=g^{j-1}(\Omega_{n}), j=1,,qn2j=1,...,q_{n}-2, in particular,

Ωn,g(Ωn),,gqn3(Ωn)Un1\Omega_{n},g(\Omega_{n}),\cdots,g^{q_{n}-3}(\Omega_{n})\subset U_{n-1}

and gqn2:Ωngqn2(Ωn)g^{q_{n}-2}:\Omega_{n}\to g^{q_{n}-2}(\Omega_{n}) is a homeomorphism, that is, (2.4) holds. It implies the rest.

(D). Given a compact set YJ(f)Y\subset J(f) denote by (Y~)f(\tilde{Y})_{f} (or simply Y~\tilde{Y}, if the map is fixed) the set of arguments of the external rays which have their limit sets contained in YY. It follows from (C) that K~c=n=1{[tn,tn][t~n,t~n]}\tilde{K}_{c}=\bigcap_{n=1}^{\infty}\{[t_{n},t_{n}^{\prime}]\cup[\tilde{t}_{n}^{\prime},\tilde{t}_{n}]\}, i.e., it is either a single-point set or a two-point set.

Since K~c\tilde{K}_{c} contains at most two angles, KcK_{c} contains at most two different accessible points. More generally, given xJx\in J^{\prime}_{\infty} let

sn,jn(x)=[tn,jn(x),tn,jn(x)][t~n,jn(x),t~n,jn(x)].s_{n,j_{n}(x)}=[t_{n,j_{n}(x)},t_{n,j_{n}(x)}^{\prime}]\cup[\tilde{t}_{n,j_{n}(x)}^{\prime},\tilde{t}_{n,j_{n}(x)}].

Then sn+1,jn+1(x)sn,jn(x)s_{n+1,j_{n+1}(x)}\subset s_{n,j_{n}(x)} so that

s,x:=n>0sn,jn(x)s_{\infty,x}:=\cap_{n>0}s_{n,j_{n}(x)}

is not empty and consists of either one or two components. Since pnjn(x)p_{n}-j_{n}(x)\to\infty for xJx\in J^{\prime}_{\infty} we conclude using (2.1):

s,xs_{\infty,x} consists of either a single point or two different points. In particular, for any component KK of JJ_{\infty} which is not one of fj(K0)f^{-j}(K_{0}), j0j\geq 0, there is either one or two rays tending to KK.

From now on, μ\mu is an ff-invariant probability ergodic measures supported in JJ_{\infty}: suppμJ\operatorname{supp}\mu\subset J_{\infty}, and having a positive Lyapunov exponent

χ(μ):=log|f|dμ>0.\chi(\mu):=\int\log|f^{\prime}|d\mu>0.

(E). We start with the following basic statement. Parts (i)-(ii) are easy consequences of the invariance of μ\mu and (B) while (iii) is a part of Pesin’s theory as in [25, Theorem 11.2.3] coupled with the structure of f:JJf:J_{\infty}\to J_{\infty}, see (B). Recall that J=Jj=fj(K0)J_{\infty}^{\prime}=J_{\infty}\setminus\cup_{j=-\infty}^{\infty}f^{j}(K_{0}).

Proposition 2.3.

(i) For every nn and 0j<pn0\leq j<p_{n}, μ(fj(Jn))=1/pn\mu(f^{j}(J_{n}))=1/p_{n}.

(ii) μ\mu has no atoms and μ(K)=0\mu(K)=0 for every component KK of JJ_{\infty}.

(iii) μ(J)=1\mu(J_{\infty}^{\prime})=1 and f:JJf:J_{\infty}^{\prime}\to J_{\infty}^{\prime} is a μ\mu-measure preserving homeomorphism. There exists a measurable positive function r~(x)>0\tilde{r}(x)>0 on JJ_{\infty}^{\prime} such that for μ\mu-almost every xJx\in J_{\infty}^{\prime}, and all n𝐍n\in{\bf N}, if xnx_{-n} is the unique point of JJ_{\infty}^{\prime} with fn(xn)=xf^{n}(x_{-n})=x, then a (univalent) branch gn:B(x,r~(x))𝐂g_{n}:B(x,\tilde{r}(x))\to{\bf C} of fnf^{-n} is well-defined such that gn(x)=xng_{n}(x)=x_{-n},

Remark 2.4.

The branch gng_{n} of fnf^{-n} depends on nn and xnx_{-n} but it should be clear from the context which points xx and xnx_{-n} are meant.

Using the Birkhoff Ergodic Theorem and Egorov’s theorem, Proposition 2.3 implies immediately (e1)-(e3) of the next corollary. The proof of (e4)-(e5) is given right after it.

Corollary 2.5.

For every ϵ>0\epsilon>0, there exists a closed set Eϵ/2JE^{\prime}_{\epsilon/2}\subset J_{\infty}^{\prime} and constants ρ=ρ(ϵ)>0\rho=\rho(\epsilon)>0, κ=κ(ϵ)(0,1)\kappa=\kappa(\epsilon)\in(0,1) such that:

(e1e_{1}) μ(Eϵ/2)>1ϵ2\mu(E^{\prime}_{\epsilon/2})>1-\frac{\epsilon}{2},

(e2e_{2}) there exists another closed set E^ϵ/2\hat{E}_{\epsilon/2} such that Eϵ/2E^ϵ/2JE^{\prime}_{\epsilon/2}\subset\hat{E}_{\epsilon/2}\subset J^{\prime}_{\infty} as follows. For every xE^ϵ/2x\in\hat{E}_{\epsilon/2} and every m>0m>0 there exists a (univalent) branch gm:B(x,3ρ)𝐂g_{m}:B(x,3\rho)\to{\bf C} of fmf^{-m} such that gm(x)=xmg_{m}(x)=x_{-m} and |gm(x1)/gm(x2)|<2|g^{\prime}_{m}(x_{1})/g^{\prime}_{m}(x_{2})|<2, for every x1,x2B(x,2ρ)x_{1},x_{2}\in B(x,2\rho). Moreover, m1ln|gm(x)|χ(μ)m^{-1}\ln|g^{\prime}_{m}(x)|\to-\chi(\mu) as mm\to\infty uniformly in xEϵ/2x\in E^{\prime}_{\epsilon/2},

(e3e_{3}) for every xEϵ/2x\in E^{\prime}_{\epsilon/2} there exists a sequence of positive integers nj=nj(x)n_{j}=n_{j}(x), j=1,2,j=1,2,..., such that j/njκj/n_{j}\geq\kappa and fnj(x)E^ϵ/2f^{n_{j}}(x)\in\hat{E}_{\epsilon/2} for all jj,

(e4e_{4}) given xJx\in J_{\infty} and n0n\geq 0, let jn(x)j_{n}(x) be the unique 1j<pn1\leq j<p_{n} such that xfj(Jn)x\in f^{j}(J_{n}). Then pnjn(x)p_{n}-j_{n}(x)\to\infty as nn\to\infty uniformly in xEϵ/2x\in E^{\prime}_{\epsilon/2},

(e5e_{5}) for sn,jn(x)=[tn,jn(x),tn,jn(x)][t~n,jn(x),t~n,jn(x)]s_{n,j_{n}(x)}=[t_{n,j_{n}(x)},t_{n,j_{n}(x)}^{\prime}]\cup[\tilde{t}_{n,j_{n}(x)}^{\prime},\tilde{t}_{n,j_{n}(x)}], we have: sn+1,jn+1(x)sn,jn(x)s_{n+1,j_{n+1}(x)}\subset s_{n,j_{n}(x)} and

|tn,jn(x)tn,jn(x)|=|t~n,jn(x)t~n,jn(x)|0|t_{n,j_{n}(x)}-t_{n,j_{n}(x)}^{\prime}|=|\tilde{t}_{n,j_{n}(x)}^{\prime}-\tilde{t}_{n,j_{n}(x)}|\to 0

as nn\to\infty uniformly in xEϵ/2x\in E^{\prime}_{\epsilon/2}.

Proof of (e4e_{4})-(e5e_{5}): assuming the contrary in (e4e_{4}), we find some NN and sequences (nk)(n_{k})\subset\mathbb{N} and (xk)(x_{k}), xkEϵ/2x_{k}\in E^{\prime}_{\epsilon/2}, such that pnkjnk(xk)=Np_{n_{k}}-j_{n_{k}}(x_{k})=N (see (B)) hence, xkfN(Jnk)x_{k}\in f^{-N}(J_{n_{k}}), for all kk. Since Eϵ/2E_{\epsilon/2} is closed, one can assume xkxEϵ/2Jx_{k}\to x\in E^{\prime}_{\epsilon/2}\subset J^{\prime}_{\infty}. Hence, xfN(K0)x\in f^{-N}(K_{0}), a contradiction. Now, for (e5e_{5}) using (e4e_{4}), tn,jn(x)tn,jn(x)=t~n,jn(x)t~n,jn(x)<12pnjn(x)0t_{n,j_{n}(x)}^{\prime}-t_{n,j_{n}(x)}=\tilde{t}_{n,j_{n}(x)}-\tilde{t}_{n,j_{n}(x)}^{\prime}<\frac{1}{2^{p_{n}-j_{n}(x)}}\to 0 uniformly in xx.

3. External rays to typical points

We define a telescope following essentially [23]. Given xJ(f)x\in J(f), r>0r>0, δ>0\delta>0, k𝐍k\in{\bf N} and κ(0,1)\kappa\in(0,1), an (r,κ,δ,k)(r,\kappa,\delta,k)-telescope at xJx\in J is collections of times 0=n0<n1<<nk0=n_{0}<n_{1}<...<n_{k} and disks Bl=B(fnl(x),r)B_{l}=B(f^{n_{l}}(x),r), l=0,1,,kl=0,1,...,k such that, for every l>0l>0: (i) l/nl>κl/n_{l}>\kappa, (ii) there is a univalent branch gnl:B(fnl(x),2r)𝐂g_{n_{l}}:B(f^{n_{l}}(x),2r)\to{\bf C} of fnlf^{-n_{l}} so that gnl(fnl(x))=xg_{n_{l}}(f^{n_{l}}(x))=x and, for l=1,,kl=1,...,k, d(fnl1gnl(Bl),Bl1)>δd(f^{n_{l-1}}\circ g_{n_{l}}(B_{l}),\partial B_{l-1})>\delta (clearly, here fnl1gnlf^{n_{l-1}}\circ g_{n_{l}} is a branch of f(nlnl1)f^{-(n_{l}-n_{l-1})} that maps fnl(x)f^{n_{l}}(x) to fnl1(x)f^{n_{l-1}}(x)). The trace of the telescope is a collection of sets Bl,0=gnl(Bl)B_{l,0}=g_{n_{l}}(B_{l}), l=0,1,,kl=0,1,...,k. We have: Bk,0Bk1,0B1,0B0,0=B0=B(x,r)B_{k,0}\subset B_{k-1,0}\subset...\subset B_{1,0}\subset B_{0,0}=B_{0}=B(x,r).

By the first point of intersection of a ray RtR_{t}, or an arc of RtR_{t}, with a set EE we mean a point of RtER_{t}\cap E with the minimal level (if it exists).

Theorem 3.1.

[23] Given r>0r>0, κ(0,1)\kappa\in(0,1), δ>0\delta>0 and C>0C>0 there exist M>0M>0, l~,k~\tilde{l},\tilde{k}\in\mathbb{N} and K>1K>1 such that for every (r,κ,δ,k)(r,\kappa,\delta,k)-telescope the following hold. Let k>k~k>\tilde{k}. Let u0=uu_{0}=u be any point at the boundary of BkB_{k} such that G(u)CG(u)\geq C. Then there are indexes 1l1<l2<<lj=k1\leq l_{1}<l_{2}<...<l_{j}=k such that l1<l~l_{1}<\tilde{l}, li+1<Klil_{i+1}<Kl_{i}, i=1,,j1i=1,...,j-1 as follows. Let uk=gnk(u)Bk,0u_{k}=g_{n_{k}}(u)\in\partial B_{k,0} and let γk\gamma_{k} be an infinite arc of an external ray through uku_{k} between the point uku_{k} and \infty. Let uk,k=uku_{k,k}=u_{k} and, for l=1,,k1l=1,...,k-1, let uk,lu_{k,l} be the first point of intersection of γk\gamma_{k} with Bl,0\partial B_{l,0}. Then, for i=1,,ji=1,...,j,

G(uk,li)>M2nli.G(u_{k,l_{i}})>M2^{-n_{l_{i}}}.

Next corollary of Theorem 3.1 is a key one.

Proposition 3.1.

Given ϵ>0\epsilon>0 there exists a closed set EϵE_{\epsilon} as follows. First, μ(Eϵ)>1ϵ\mu(E_{\epsilon})>1-\epsilon and EϵEϵ/2E_{\epsilon}\subset E^{\prime}_{\epsilon/2} where Eϵ/2E^{\prime}_{\epsilon/2} is the set defined in (E) and satisfies (e1e_{1})-(e5e_{5}). There exists r=r(ϵ)>0r=r(\epsilon)>0 and, for each ν>0\nu>0 there is C(ν)>0C(\nu)>0 as follows.

(1) Let xEϵx\in E_{\epsilon}. Then xx is the landing point of an external ray Rt(x)R_{t(x)} of argument t(x)t(x). Moreover, the first intersection of Rt(x)R_{t(x)} with B(x,ν)\partial B(x,\nu) has the level at least C(ν)C(\nu).

(2) for each nn, a branch gn:B(x,2r)𝐂g_{n}:B(x,2r)\to{\bf C} of fnf^{-n} is well-defined such that gn(x)=xng_{n}(x)=x_{-n}, |gn(x1)/gn(x2)|<2|g^{\prime}_{n}(x_{1})/g^{\prime}_{n}(x_{2})|<2, for every x1,x2B(x,r)x_{1},x_{2}\in B(x,r) and n1ln|gn(x)|χ(μ)n^{-1}\ln|g^{\prime}_{n}(x)|\to-\chi(\mu) as mm\to\infty uniformly in xEϵx\in E_{\epsilon},

(3) if x=gn(x)Eϵx^{\prime}=g_{n}(x)\in E_{\epsilon}, then fn(Rt(x))=Rt(x)f^{n}(R_{t(x^{\prime})})=R_{t(x)}.

Proof.

(1)-(2) will hold already for the set Eϵ/2E^{\prime}_{\epsilon/2} which follows from Theorem 3.1 as in [23] and uses only that μ\mu has a positive exponent; (3) will follow in our case as we shrink a bit the set Eϵ/2E^{\prime}_{\epsilon/2} since each point xJx\in J_{\infty}^{\prime} admits at most two external arguments.

Here are details. Let r=ρ(ϵ)r=\rho(\epsilon) and κ=κ(ϵ)\kappa=\kappa(\epsilon) as in the properties (e2e_{2})-(e3e_{3}) of the set Eϵ/2E^{\prime}_{\epsilon/2}. Then, by (e2e_{2})-(e3e_{3}), there is δ>0\delta>0 such that, for each kk, every xEϵ/2x\in E^{\prime}_{\epsilon/2} admits (r,κ,δ,k)(r,\kappa,\delta,k)-telescope with the times 0=n0<n1<n2<<nk0=n_{0}<n_{1}<n_{2}<...<n_{k} that appear in the property (e3e_{3}) of Eϵ/2E^{\prime}_{\epsilon/2}. On the other hand, there exists Lr>0L_{r}>0 such that for every zJ(f)z\in J(f) there is a point u(z)B(z,r)u(z)\in\partial B(z,r) with the level G(u(z))>LrG(u(z))>L_{r}. This is so due to L>0{G(z)L}=A\bigcup_{L>0}\{G(z)\geq L\}=A_{\infty}.

Given this C=LrC=L_{r}, let MM, l~\tilde{l} and k~\tilde{k} be as in Theorem 3.1.

Let xEϵ/2x\in E^{\prime}_{\epsilon/2} and n1<n2<<nk<n_{1}<n_{2}<...<n_{k}<... as in (e3e_{3}). Fix k>k~k>\tilde{k}. Let Bk,0(x)Bk1,0(x)B1,0(x)B0,0(x)B_{k,0}(x)\subset B_{k-1,0}(x)\subset\cdots\subset B_{1,0}(x)\subset B_{0,0}(x) be the corresponding trace. By Theorem 3.1, there are 1l1,k(x)<l2,k(x)<<ljkx,k(x)=k1\leq l_{1,k}(x)<l_{2,k}(x)<\cdots<l_{j^{x}_{k},k}(x)=k such that l1,k(x)<l~l_{1,k}(x)<\tilde{l}, li+1,k(x)<Kli,k(x)l_{i+1,k}(x)<Kl_{i,k}(x), i=1,,jkx1i=1,\cdots,j^{x}_{k}-1. Let γk(x)\gamma_{k}(x) be an arc of an external ray between the point uk(x):=gnk(u(fnk(x))u_{k}(x):=g_{n_{k}}(u(f^{n_{k}}(x)) and \infty. Let uk,l(x)u_{k,l}(x) be the first intersection of γk(x)\gamma_{k}(x) with Bl,0(x)\partial B_{l,0}(x). Then, for i=1,,jkx1i=1,\cdots,j^{x}_{k}-1,

(3.1) G(uk,li,k(x))>M2nli,k(x)>M2li,k(x)/κ.G(u_{k,l_{i,k}}(x))>M2^{-n_{l_{i,k}(x)}}>M2^{-l_{i,k}(x)/\kappa}.

For all i=1,,jkx1i=1,\cdots,j^{x}_{k}-1,

(3.2) ili,k(x)<Kil~.i\leq l_{i,k}(x)<K^{i}\tilde{l}.

Denote by tk(x)t_{k}(x) the argument of an external ray that contains the arc γk(x)\gamma_{k}(x).

Now, given a sequence

(3.3) k1<k2<<km<k_{1}<k_{2}<...<k_{m}<...

such that k1>k~k_{1}>\tilde{k}, we get a sequence of arguments tkm(x)t_{k_{m}}(x) and a sequence of arcs γkm(x)\gamma_{k_{m}}(x) of external rays of the corresponding arguments tkm(x)t_{k_{m}}(x). Passing to a subsequence in the sequence (km)(k_{m}), if necessary, one can assume that tkm(x)t~(x)t_{k_{m}}(x)\to\tilde{t}(x), for some argument t~(x)\tilde{t}(x).

Fix any ν(0,r)\nu\in(0,r) and choose k~0>k~\tilde{k}_{0}>\tilde{k} such that,

2exp(Kk~02l~χ(μ))<ν and let C(ν)=M(21/κ)l~Kk~0.2\exp(-K^{\tilde{k}_{0}-2}\tilde{l}\chi(\mu))<\nu\mbox{ and let }C(\nu)=M(2^{-1/\kappa})^{\tilde{l}K^{\tilde{k}_{0}}}.

Then, by Theorem 3.1, for each km>k0k_{m}>k_{0}, the first intersection of the ray Rtkm(x)R_{t_{k_{m}}}(x) with the boundary of B(x,ν)B(x,\nu) has the level at least C(ν)C(\nu). It follows, for any 0<C<C(ν)0<C<C(\nu), the sequence of arcs of the rays Rtkm(x)R_{t_{k_{m}}(x)} between the levels CC and C(ν)C(\nu) do not exit B(x,ν)B(x,\nu) for all km>k0k_{m}>k_{0}. As tkm(x)t~(x)t_{k_{m}}(x)\to\tilde{t}(x), it follows that the arc of the ray Rt~(x)R_{\tilde{t}(x)} between levels CC and C(ν)C(\nu) stays in B(x,ν)B(x,\nu) too. As ν>0\nu>0 and C(0,C(ν))C\in(0,C(\nu)) can be chosen arbitrary small, Rt~(x)R_{\tilde{t}(x)} must land at xx and satisfy (1) with t(x)t(x) replaced by t~(x)\tilde{t}(x).

Let us call the above procedure of getting t~(x)\tilde{t}(x) from the constants rr, LrL_{r}, the point xEϵ/2x\in E^{\prime}_{\epsilon/2} and the sequence (3.3) the (r,Lr,x,(km))(r,L_{r},x,(k_{m}))-procedure.

Note that (2) is property (e2e_{2}) of the set Eϵ/2E^{\prime}_{\epsilon/2}.

In order to satisfy property (3), we shrink the set Eϵ/2E^{\prime}_{\epsilon/2} and correct t~(x)\tilde{t}(x) changing it to some t(x)t(x) (if necessary) as follows. Using the Birkhoff Ergodic Theorem and Egorov’s theorem, choose a closed subset EϵE_{\epsilon} of Eϵ/2E^{\prime}_{\epsilon/2} such that μ(Eϵ)>1ϵ\mu(E_{\epsilon})>1-\epsilon and, for each xEϵx\in E_{\epsilon}, the set 𝒩(x):={N:fN(x)Eϵ/2}\mathcal{N}(x):=\{N\in\mathbb{N}:f^{N}(x)\in E^{\prime}_{\epsilon/2}\} is infinite. Note that 𝒩(x){nk}k=1\mathcal{N}(x)\subset\{n_{k}\}_{k=1}^{\infty}. We have proved that, for each N𝒩(x)N\in\mathcal{N}(x), (1) holds for the point fN(x)f^{N}(x) instead of xx, in particular, t~(fN(x))\tilde{t}(f^{N}(x)) is an argument of fN(x)f^{N}(x). On the other hand, by (D1), each yEϵy\in E_{\epsilon} admits at most two external arguments, hence, all possible external arguments of the forward orbit fn(x)f^{n}(x), n0n\geq 0, belong to at most two different orbits of σ:S1S1\sigma:S^{1}\to S^{1}. Hence, there is one of those orbits, O={σn(t(x))}n0O=\{\sigma^{n}(t(x))\}_{n\geq 0} for some t(x)t(x), such that the intersection O{t~(fN(x)):N𝒩(x)}O\cap\{\tilde{t}(f^{N}(x)):N\in\mathcal{N}(x)\} is an infinite set, so that t~(fnkm(x)(x))=σnkm(x)(t(x))\tilde{t}(f^{n_{k_{m}(x)}}(x))=\sigma^{n_{k_{m}(x)}}(t(x)) for an infinite sequence (km(x))m1(k_{m}(x))_{m\geq 1}.

Let’s start over with the (r/2,C(r/2),x,(km(x)))(r/2,C(r/2),x,(k_{m}(x)))-procedure for the point xx and the sequence {kj(x)}\{k_{j}(x)\}. Then, by the construction, tkm(x)=t(x)t_{k_{m}(x)}=t(x) for all mm, hence, (1) holds with t(x)t(x) instead of the previous t~(x)\tilde{t}(x). If yEϵy\in E_{\epsilon} is any other point of the grand orbit {fn(x):n}\{f^{n}(x):n\in\mathbb{Z}\} (remember that f:JJf:J^{\prime}_{\infty}\to J^{\prime}_{\infty} is invertible), the (r/2,C(r/2),y,(km))(r/2,C(r/2),y,(k_{m}))-procedure works for yy with the same (perhaps, truncated) sequence k1(x)<k2(x)<k_{1}(x)<k_{2}(x)<..., which ensures that (3) holds (for the corrected arguments) too. ∎

Remark 3.2.

Given t(x)t(x), we cannot just set t(fn(x))=σn(t(x))t(f^{n}(x))=\sigma^{n}(t(x)) to satisfy property (3) because this would change κ\kappa in the definition of telescope, so we might loose property (1). Notice that correcting (flipping) t~(x)\tilde{t}(x) to t(x)t(x) does not change C(ν)C(\nu) The same for flipping any t(y)t(y) in the grand orbit of xx. But the flipping can make f(Rt(y))=Rt(fN(x)f^{\ell}(R_{t(y)})=R_{t(f^{N}(x)} for f(y)=fN(x)f^{\ell}(y)=f^{N}(x) where N=nkmN=n_{k_{m}} with G(Rt(f(y))B(f(y),r/2)>Lr/2G(R_{t(f^{\ell}(y))}\cap\partial B(f^{\ell}(y),r/2)>L_{r/2}, thus yielding (3).

4. Lemmas

Recall that for any zJz\in J^{\prime}_{\infty} we define zm=fm(z)z_{m}=f^{m}(z) for any mm\in\mathbb{Z}. This makes sense since ff is invertible on JJ^{\prime}_{\infty}, see (E).

Lemma 4.1.

Let z(k)j=0pnk1fj(Jnk)z(k)\in\cup_{j=0}^{p_{n_{k}}-1}f^{j}(J_{n_{k}}) where nkn_{k}\nearrow\infty.

(a) If z(k)zz(k)\to z then zJz\in J_{\infty}.

(b) zJn,xJz\in J_{n,x}\cap J^{\prime}_{\infty} yields z±pnJn,xz_{\pm p_{n}}\in J_{n,x}. If, additionally to (a), z(k)Jz(k)\in J^{\prime}_{\infty} for all kk and w(k)ww(k)\to w where w(k)=z(k)epnkw(k)=z(k)_{ep_{n_{k}}}, where ee is always either 11 or 1-1 then zz and ww are in the same component of JJ_{\infty}.

(c) If z(k)Eϵz(k)\in E_{\epsilon} for all kk and t(z(k))tt(z(k))\to t (where EϵE_{\epsilon}, t(z(k))t(z(k)) are defined in Proposition 3.1), then the ray RtR_{t} lands at the limit point zz. In particular, given σ>0\sigma>0 there is Δ(σ)>0\Delta(\sigma)>0 such that |x1x2|<σ|x_{1}-x_{2}|<\sigma for some x1,x2Eϵx_{1},x_{2}\in E_{\epsilon} whenever |t(x1)t(x2)|<Δ(σ)|t(x_{1})-t(x_{2})|<\Delta(\sigma).

Proof.

(a) Assume the contrary. Then there is nn such that d:=d(z,j=0pn1fj(Jn))>0d:=d(z,\cup_{j=0}^{p_{n}-1}f^{j}(J_{n}))>0. As, for any nknn_{k}\geq n, z(k)j=0pnk1fj(Jnk)z(k)\in\cup_{j=0}^{p_{n_{k}}-1}f^{j}(J_{n_{k}}) where the latter union is a subset of j=0pn1fj(Jn)\cup_{j=0}^{p_{n}-1}f^{j}(J_{n}), the distance between zz and zkz_{k} is at least dd, a contradiction.

(b) z±pnJn,xz_{\pm p_{n}}\in J_{n,x} by combinatorics and definitions of points zmz_{m}. In particular, for every kk, z(k)z(k) and w(k)w(k) are in the same component fjk(Jnk)f^{j_{k}}(J_{n_{k}}) of j=0pnk1fj(Jnk)\cup_{j=0}^{p_{n_{k}}-1}f^{j}(J_{n_{k}}). By (a), any limit set AA of the sequence of compacts (fjk(Jnk))(f^{j_{k}}(J_{n_{k}})) in the Hausdorff metric is a subset of JJ_{\infty}. On the other hand, AA is connected as each set fjk(Jnk)f^{j_{k}}(J_{n_{k}}) is connected. This proves (b).

(c) We prove only the first claim as the second one directly follows from it. Fix any ν(0,r)\nu\in(0,r) and choose k0k_{0} such that for any k>k0k>k_{0}, B(z(k),ν)B(z,11/10ν)B(z(k),\nu)\subset B(z,11/10\nu). Then, by Proposition 3.1, part (1), for each k>k0k>k_{0}, the first intersection of the ray Rt(zk)R_{t(z_{k})} with the boundary of B(z,(11/10)ν)B(z,(11/10)\nu) has the level at least C(ν)C(\nu). It follows, for any 0<C<C(ν)0<C<C(\nu), the sequence of arcs of the rays RtzkR_{t_{z_{k}}} between the levels CC and C(ν)C(\nu) do not exit B(z,(11/10)ν)B(z,(11/10)\nu) for all k>k0k>k_{0}. As ν>0\nu>0 and C(0,C~(ν))C\in(0,\tilde{C}(\nu)) can be chosen arbitrary small, RtR_{t} must land at zz. ∎

By lemma 4.1(c), if arguments t(x),t(x)t(x),t(x^{\prime}) of x,xEϵx,x^{\prime}\in E_{\epsilon} are close then x,xx,x^{\prime} are close as well.

Definition 4.2.

Given ϵ\epsilon and ρ\rho we define δ\delta as follows. First, for r^(0,1)\hat{r}\in(0,1) and C^>0\hat{C}>0, we define δ^=δ^(r^,C^)>0\hat{\delta}=\hat{\delta}(\hat{r},\hat{C})>0. Namely, let C0>0C_{0}>0 be so that the distance between the equipotential of level C0C_{0} and J(f)J(f) is bigger than 11. Then δ^=δ^(r^/2,C^)>0\hat{\delta}=\hat{\delta}(\hat{r}/2,\hat{C})>0 is such that for any C[C^,C0]C\in[\hat{C},C_{0}], if w1,w2w_{1},w_{2} lie on the same equipotential Γ\Gamma of level CC and the difference between external arguments of w1,w2w_{1},w_{2} is less than δ^\hat{\delta} then the length of the shortest arc of the equipotential Γ\Gamma between w1w_{1} and w2w_{2} is less than r^/2\hat{r}/2. Apply Lemma 4.1(c) with σ=ρ/4\sigma=\rho/4 and find the corresponding Δ(ρ/4)\Delta(\rho/4). Let

δ=δ(ϵ,ρ):=min{δ^(ρ,C(ρ/2)),Δ(ρ4)}\delta=\delta(\epsilon,\rho):=\min\{\hat{\delta}(\rho,C(\rho/2)),\Delta(\frac{\rho}{4})\}

where C(ν)C(\nu) is defined in Proposition 3.1.

In the next two lemmas we construct curves with special properties. The idea is as follows. Let xEϵJn,xx\in E_{\epsilon}\cap J_{n,x}. Then xpnJn,xx_{-p_{n}}\in J_{n,x}. It is easy to get a curve γ\gamma in A()A(\infty) as follows: begin with an arc from a point bRt(x)b\in R_{t(x)} to gpn(b)g_{p_{n}}(b) and then iterate this arc by gpng_{p_{n}}. In this way we get a curve γ\gamma such that gpn(γ)γg^{p_{n}}(\gamma)\subset\gamma, hence, γ\gamma lands at a fixed point aa of fpnf^{p_{n}}. We show in the next lemma (in a more general setting) that if both points x,xpnx,x_{-p_{n}} are either in the range of the covering (2.2) (condition (I)) or in the range of the covering (2.3) (condition (II)) then aJn,xa\in J_{n,x}. This implies that aa has to be the β\beta-fixed point of fpn:Jn,xJn,xf^{p_{n}}:J_{n,x}\to J_{n,x}. In Lemma 4.5 assuming additionally that fpnf^{p_{n}} is satellite, we ’rotate’ the curve γ\gamma by gpn1g_{p_{n-1}} to put the set Jn,xJ_{n,x} in a ’sector’ bounded by γ\gamma and by its ’rotation’. In Lemma 4.7-4.8 we consider the case of doubling for which the condition (II) usually does not hold.

In what follows, we use the following notation: given p,qp,q\in\mathbb{N} , let

Eϵ,p,q=j=0q1fjp(Eϵ).E_{\epsilon,p,q}=\cap_{j=0}^{q-1}f^{jp}(E_{\epsilon}).

It is a closed subset of EϵE_{\epsilon} of points xx such that xjpEϵx_{-jp}\in E_{\epsilon} for j=0,1,,q1j=0,1,\cdots,q-1. As f:JJf:J^{\prime}_{\infty}\to J^{\prime}_{\infty} is a μ\mu-automorphism, μ(Eϵ,p,q)>1qϵ\mu(E_{\epsilon,p,q})>1-q\epsilon. Notice that this bound is independent of pp.

For every n>0n>0 consider the closed set Eϵ,pn,qE_{\epsilon,p_{n},q}. Let xEϵ,pn,qx\in E_{\epsilon,p_{n},q}. Denote for brevity

xk:=xkpn and Rk:=Rt(xk),k=0,1,,q1.x^{k}:=x_{-kp_{n}}\mbox{ and }R^{k}:=R_{t(x^{k})},\ \ k=0,1,...,q-1.

By Lemma 4.1(b), xkJn,xx^{k}\in J_{n,x}. Hence, t(xk)sn,jn(x)Sn,jn(x)t(x^{k})\in s_{n,j_{n}(x)}\subset S_{n,j_{n}(x)}, 0kq10\leq k\leq q-1.

Recall that for a semi-open curve l:[0,1)l:[0,1)\to\mathbb{C}, we say that ll lands at, or tends to, or converges to a point zz\in\mathbb{C} if there exists limt1l(t)=z\lim_{t\to 1}l(t)=z. Then l(0)l(0), zz are endpoints of the curve and l(0)l(0) is called also the starting point of ll.

Lemma 4.3.

Fix ϵ>0\epsilon>0 and consider the set EϵE_{\epsilon} with the corresponding constant r(ϵ)>0r(\epsilon)>0. Fix ρ(0,r(ϵ)/3)\rho\in(0,r(\epsilon)/3). Let δ:=δ(ϵ,ρ)\delta:=\delta(\epsilon,\rho) from Definition 4.2. For every q2q\geq 2 there exist n~\tilde{n}, C~\tilde{C} as follows.

For every n>n~n>\tilde{n} consider the closed set Eϵ,pn,qE_{\epsilon,p_{n},q}. Fix 0i<jq10\leq i<j\leq q-1. Assume for an arbitrary nn as above, that either (I) t(xj)t(x^{j}) and t(xi)t(x^{i}) belong to a single component of sn,jn(x)s_{n,j_{n}(x)}, or (II) the map σjn(x)1:Sn,1Sn,jn(x)\sigma^{j_{n}(x)-1}:S_{n,1}\to S_{n,j_{n}(x)} is a homeomorphism and the length of the arc Sn,jn(x)S_{n,j_{n}(x)} is less than δ\delta.

Then:

(a) the map f(ji)pn:g(ji)pn(B(xi,ρ))B(xi,ρ)f^{(j-i)p_{n}}:g_{(j-i)p_{n}}(B(x^{i},\rho))\to B(x^{i},\rho) has a unique fixed point a=ana=a_{n} and aJn,xa\in J_{n,x},

(b) there is a semi-open simple curve

γpn,q,i,j(x)B(xi,ρ)A()\gamma_{p_{n},q,i,j}(x)\subset B(x^{i},\rho)\cap A(\infty)

such that:

  1. (1)

    it lands at aa and g(ji)pn(γpn,q,i,j(x))γpn,q,i,j(x)g_{(j-i)p_{n}}(\gamma_{p_{n},q,i,j}(x))\subset\gamma_{p_{n},q,i,j}(x). Another end point bb of γpn,q,i,j(x)\gamma_{p_{n},q,i,j}(x) lies in RiR^{i} and G(b)>C~/2G(b)>\tilde{C}/2,

  2. (2)

    γpn,q,i,j(x)=l0g(ji)pnl(L0L1)\gamma_{p_{n},q,i,j}(x)=\cup_{l\geq 0}g_{(j-i)p_{n}}^{l}(L_{0}\cup L_{1}) where the ’fundamental arc’ L0L1L_{0}\cup L_{1} consists of an arc L0L_{0} of an equipotential of the level at least C~/2\tilde{C}/2 that joins a point bRib\in R^{i} with a point b1Rjb_{1}\in R^{j}, being extended by an arc L1L_{1} of the ray RjR^{j} between points b1b_{1} and g(ji)pn(b)Rjg_{(j-i)p_{n}}(b)\in R^{j}; in particular, the Green function G(y)G(y) at a point yy is not increasing as yy moves from bb to aa along γpn,q,i,j(x)\gamma_{p_{n},q,i,j}(x),

  3. (3)

    the point aa is the landing point of a ray R(a)R(a) which is fixed by f(ji)pnf^{(j-i)p_{n}} and which is homotopic to γpn,q,i,j(x)\gamma_{p_{n},q,i,j}(x) through a family of curves in A()A(\infty) with the fixed end point aa.

  4. (4)

    arguments of all points of the curve g(ji)pn(γpn,q,i,j(x))g_{(j-i)p_{n}}(\gamma_{p_{n},q,i,j}(x)) lie in a single component of sn,jn(x)1s^{1}_{n,j_{n}(x)} in the case (I) and in a single component of sn,jn(x)s_{n,j_{n}(x)} in the case (II) (recall that sn,jn(x)1s^{1}_{n,j_{n}(x)} has 44 components and sn,jn(x)s_{n,j_{n}(x)} has 22 components, see Sect 2, (C)).

Besides,

(4.1) |axj|0 and log|(g(ji)pn)(xj)||(g(ji)pn)(a)|0|a-x^{j}|\to 0\mbox{ and }\log\frac{|(g_{(j-i)p_{n}})^{\prime}(x^{j})|}{|(g_{(j-i)p_{n}})^{\prime}(a)|}\to 0

as nn\to\infty, uniformly in xjx^{j} and qq.

(c) if ji=1j-i=1 then a=βn,jn(x)a=\beta_{n,j_{n}(x)} where βn,jn(x)=fjn(x)1(βn)\beta_{n,j_{n}(x)}=f^{j_{n}(x)-1}(\beta_{n}), the non-separating fixed point of fpn:Jn,xJn,xf^{p_{n}}:J_{n,x}\to J_{n,x}. Moreover,

χ(βn,jn(x)):=1pnlog|(fpn)(βn,jn(x))|=1pnlog|(fpn)(βn)|χ(μ)\chi(\beta_{n,j_{n}(x)}):=\frac{1}{p_{n}}\log|(f^{p_{n}})^{\prime}(\beta_{n,j_{n}(x)})|=\frac{1}{p_{n}}\log|(f^{p_{n}})^{\prime}(\beta_{n})|\to\chi(\mu)

as nn\to\infty.

Remark 4.4.

Note that aJa\notin J_{\infty} while x,x1,,xq1Jx,x^{1},...,x^{q-1}\in J_{\infty}.

Proof.

Fix n0n_{0} such that, for every n>n0n>n_{0} and xEϵx\in E_{\epsilon}, the length of each ’window’ of sn,jn(x)s_{n,j_{n}(x)} is less than δ\delta. Therefore, for n>n0n>n_{0}, in either case (I), (II),

(4.2) |t(xi)t(xj)|<δ,|t(x^{i})-t(x^{j})|<\delta,

which implies, in particular, that |xixj|<ρ/4|x^{i}-x^{j}|<\rho/4.

Denote Gn:=g(ji)pnG_{n}:=g_{(j-i)p_{n}} which is a holomorphic univalent function in B(xi,ρ)B(x^{i},\rho). Since gmg_{m} are uniform contractions there is n1n_{1} such that Gn(B(xi,ρ)¯)B(xi,ρ/2)G_{n}(\overline{B(x^{i},\rho)})\subset B(x^{i},\rho/2) whenever n>n1n>n_{1}. Let n~=max{n0,n1}\tilde{n}=\max\{n_{0},n_{1}\}.

Let also C~=C(ρ/2)\tilde{C}=C(\rho/2), where C(ν)C(\nu) is defined in Proposition 3.1.

Let a=ana=a_{n} be the unique fixed point of the latter map GnG_{n}. We construct the curve γpn,q,i,j(x)\gamma_{p_{n},q,i,j}(x) to the point aa as follows. First, joint a point bRib\in R^{i}, G(b)=(3/4)C~G(b)=(3/4)\tilde{C}, to a point b1Rjb_{1}\in R^{j} by an arc L0L_{0} of the equipotential {G(z)=(3/4)C~}\{G(z)=(3/4)\tilde{C}\}. By the choice of δ>0\delta>0, L0B(xi,ρ)L_{0}\subset B(x^{i},\rho). Secondly, connect b1b_{1} to the point g(ji)pn(b)Rjg_{(j-i)p_{n}}(b)\in R^{j} by an arc L1RjL_{1}\subset R^{j}. Let now γpn,q,i,j(x)=l0g(ji)pnl(L0L1)\gamma_{p_{n},q,i,j}(x)=\cup_{l\geq 0}g_{(j-i)p_{n}}^{l}(L_{0}\cup L_{1}). Then properties (1), (2) in (b) are immediate and (3) follows from general properties of conformal maps. Now, by Proposition 3.1(2) and (4.2), for all nn big enough, xj=g(ji)pn(xi)g(ji)pn(B(xi,ρ))B(xi,ρ)x^{j}=g_{(j-i)p_{n}}(x^{i})\in g_{(j-i)p_{n}}(B(x^{i},\rho))\subset B(x^{i},\rho), moreover, the modulus of the annulus B(xi,ρ)g(ji)pn(B(xi,ρ))B(x^{i},\rho)\setminus g_{(j-i)p_{n}}(B(x^{i},\rho)) tends to \infty as nn\to\infty. Therefore, (4.1) follows from Koebe and Proposition 3.1(2).

It remains to show the property (3) and that aJn,xa\in J_{n,x}. Consider the case (II), which is equivalent to say that the map σpn:sSn,jn(x)\sigma^{p_{n}}:s\to S_{n,j_{n}(x)} is a homeomorphism on each of two components ss of sn,jn(x)s_{n,j_{n}(x)}. Let Λ\Lambda be the set of arguments of points of the curve Γ:=g(ji)pn(γpn,q,i,j(x))\Gamma:=g_{(j-i)p_{n}}(\gamma_{p_{n},q,i,j}(x)). Let ss be a component that contains t(xj)t(x^{j}). Assume, by a contradiction, that Λ\Lambda contains tt which is in the boundary of ss. Then tt is the argument of a point of Gnl(L0)G_{n}^{l}(L_{0}), for some l1l\geq 1, hence, σl(ji)pn(t)\sigma^{l(j-i)p_{n}}(t) is simultaneously the argument of a point of L0L_{0} and in the boundary of Sn,jn(x)S_{n,j_{n}(x)}, a contradiction. The case (I) is similar. Property (3) is verified. In fact, we proved more: for k=0,1,,ji1k=0,1,\cdots,j-i-1, the set σkpn(Λ)\sigma^{kp_{n}}(\Lambda) is a subset of a single (depending on kk) component of sn,jn(x)s_{n,j_{n}(x)} in the case (II) and a single component of sn,jn(x)1s^{1}_{n,j_{n}(x)} in the case (I). This implies that all point fkpn(a)f^{kp_{n}}(a), 0kji10\leq k\leq j-i-1, of the cycle of fpnf^{p_{n}} containing aa belong to the closure of Un,jn(x)U_{n,j_{n}(x)} in the case (II) and to the closure of Un,jn(x)pnU_{n,j_{n}(x)-p_{n}} in the case (I). Therefore, this cycle lies in Jn,xJ_{n,x}, in particular, aJn,xa\in J_{n,x}.

Proof of (c): if ji=1j-i=1 then aa is a fixed point of fpn:Jn,xJn,xf^{p_{n}}:J_{n,x}\to J_{n,x} and, moreover, the ray R(a)R(a) lands at aa and is fixed by fpnf^{p_{n}}. Hence, the rotation number of aa w.r.t. the map fpn:Jn,xJn,xf^{p_{n}}:J_{n,x}\to J_{n,x} is zero. On the other hand, βn,jn(x)\beta_{n,j_{n}(x)} is the only such a fixed point, i.e., a=βn,jn(x)a=\beta_{n,j_{n}(x)} as claimed. Then (4.1) implies that χ(βn,jn(x))χ(μ)\chi(\beta_{n,j_{n}(x)})\to\chi(\mu). ∎

For the rest of the paper, let us fix QQ, ϵ\epsilon, rr, ρ\rho, n~\tilde{n}, C~\tilde{C} and δ\delta as follows:

QQ\in\mathbb{N}, Q>3Q>3, is such that

Q>4log2/χ(μ).Q>4\log 2/\chi(\mu).

This choice is motivated by the following

Fact ([21], [13], [8]): if a repelling fixed point zz of fnf^{n} is the landing point of qq rays, then χ(z):=(1/n)log|(fn)(z)|(2/q)log2\chi(z):=(1/n)\log|(f^{n})^{\prime}(z)|\leq(2/q)\log 2. Hence, if χ(z)>χ(μ)/2\chi(z)>\chi(\mu)/2, then q<Qq<Q.

Furthermore, fix ϵ>0\epsilon>0 such that 2100Qϵ<12^{100}Q\epsilon<1, apply Proposition 3.1 and Lemma 4.3 and find, first, r=r(ϵ)r=r(\epsilon), then fix ρ(0,r/32)\rho\in(0,r/32) and find the corresponding n~\tilde{n}, C~\tilde{C} and δ\delta.

Let

Xn=Eϵ,pn,4Eϵ,pn1,Q=i=03fipn(Eϵ)k=0Q1fkpn1(Eϵ).X_{n}=E_{\epsilon,p_{n},4}\cap E_{\epsilon,p_{n-1},Q}=\cap_{i=0}^{3}f^{ip_{n}}(E_{\epsilon})\cap_{k=0}^{Q-1}f^{kp_{n-1}}(E_{\epsilon}).

Let us analyze several possibilities.

Lemma 4.5.

There is n>n~n_{*}>\tilde{n} as follows. Let n>nn>n_{*} and xXnx\in X_{n}. Consider Jn,x=fjn(x)(Jn)fjn1(x)(Jn1)J_{n,x}=f^{j_{n}(x)}(J_{n})\subset f^{j_{n-1}(x)}(J_{n-1}) so that xJn,xx\in J_{n,x}.

Let x0=xx^{0}=x and x1=xpnx^{1}=x_{-p_{n}}. Assume that either (I) t(x0)t(x^{0}), t(x1)t(x^{1}) belong to a single component of sn,jn(x)s_{n,j_{n}(x)}, or (II) the map σjn(x)1:Sn,1Sn,jn(x)\sigma^{j_{n}(x)-1}:S_{n,1}\to S_{n,j_{n}(x)} is a homeomorphism and the length of the arc Sn,jn(x)S_{n,j_{n}(x)} is less than δ\delta.

Then:

(i) χ(βn,jn(x))=χ(βn)χ(μ)\chi(\beta_{n,j_{n}(x)})=\chi(\beta_{n})\to\chi(\mu) as nn\to\infty and χ(βn)>χ(μ)/2\chi(\beta_{n})>\chi(\mu)/2 for n>nn>n_{*}.

(ii) assume that fpnf^{p_{n}} is satellite, i.e., (by Lemma 2.1) βn\beta_{n} has period pn1p_{n-1}, qn2q_{n}\geq 2 with rotation number rn/qnr_{n}/q_{n} of βn\beta_{n}, and βn,jn(x)\beta_{n,j_{n}(x)} is the α\alpha (i.e., separating) fixed point of fpn1:Jn1,xJn1,xf^{p_{n-1}}:J_{n-1,x}\to J_{n-1,x}. Then qn<Qq_{n}<Q and

(4.3) |βn,jn(x)xkpn1|0,n, uniformly in xXn, 1kqn.|\beta_{n,j_{n}(x)}-x_{-kp_{n-1}}|\to 0,\ n\to\infty,\mbox{ uniformly in }x\in X_{n},\ 1\leq k\leq q_{n}.

There exist two simple semi-open curves γ(x)\gamma(x) and γ~(x)\tilde{\gamma}(x) that satisfy the following properties:

  1. (1)

    γ(x)\gamma(x) and γ~(x)\tilde{\gamma}(x) tend to βn,jn(x)\beta_{n,j_{n}(x)} and γ(x),γ~(x)B(x0,ρ)A()\gamma(x),\tilde{\gamma}(x)\subset B(x^{0},\rho)\cap A(\infty),

  2. (2)

    γ(x),γ~(x)\gamma(x),\tilde{\gamma}(x) consist of arcs of equipotentials and external rays; the starting point b1=b1(x)b_{1}=b_{1}(x) of γ(x)\gamma(x) lies in an arc of Rt(x1)R_{t(x^{1})} and the starting point b~1=b~1(x)\tilde{b}_{1}=\tilde{b}_{1}(x) of γ~(x)\tilde{\gamma}(x) lies in an arc of Rt(x~)R_{t(\tilde{x})} where x~=xipn1\tilde{x}=x_{-ip_{n-1}} for some i=i(x){1,,qn1}i=i(x)\in\{1,\cdots,q_{n}-1\}, such that levels of b1b_{1} and b~1\tilde{b}_{1} are equal and at least C~/4\tilde{C}/4,

  3. (3)

    one of the two curves (say, γ(x)\gamma(x)) is homotopic, through curves in A()A(\infty) tending to βn,jn(x)\beta_{n,j_{n}(x)}, to the ray Rtn,jn(x)=fjn(x)1(Rtn)R_{t_{n,j_{n}(x)}}=f^{j_{n}(x)-1}(R_{t_{n}}), and another one - to the ray Rt~n,jn(x)=fjn(x)1(Rt~n)R_{\tilde{t}_{n,j_{n}(x)}}=f^{j_{n}(x)-1}(R_{\tilde{t}_{n}});

  4. (4)

    γ(x)\gamma(x), γ~(x)Un1,jn1(x)\tilde{\gamma}(x)\subset U_{n-1,j_{n-1}(x)},

  5. (5)

    γ(x)Un,jn(x)\gamma(x)\subset U_{n,j_{n}(x)}, γ~(x)Un,jn(x~)\tilde{\gamma}(x)\subset U_{n,j_{n}(\tilde{x})}, in particular, γ(x),γ~(x)\gamma(x),\tilde{\gamma}(x) are disjoint; being completed by their common limit point βn,jn(x)\beta_{n,j_{n}(x)} and two other arcs: an arc of the ray Rt(x1)R_{t(x^{1})} from b1γ(x)b_{1}\in\gamma(x) to \infty and an arc of the ray Rt(x~)R_{t(\tilde{x})} from b~1γ~(x)\tilde{b}_{1}\in\tilde{\gamma}(x) to \infty, they split the plane into two domains such that one of them contains I:=Jn,xβn,jn(x)I:=J_{n,x}\setminus\beta_{n,j_{n}(x)} and another one contains all qn1q_{n}-1 other different iterates fkpn1(I)f^{kp_{n-1}}(I), 1kqn11\leq k\leq q_{n}-1. The intersection of closures of all those qnq_{n} sets consists of the fixed point βn,jn(x)\beta_{n,j_{n}(x)} of fpn1f^{p_{n-1}}.

Remark 4.6.

Beware that the point xx that determines both curves γ(x)\gamma(x), γ~(x)\tilde{\gamma}(x) does not belong to either of these curves.

Proof.

(i) follows from Lemma 4.3 where we take i=0,j=1i=0,j=1. Fix n>n~n_{*}>\tilde{n} such that χ(βn)>χ(μ)/2\chi(\beta_{n})>\chi(\mu)/2 for all n>nn>n_{*}.

Let us prove (ii). Here we build a "flower" of arcs at the β\beta fixed of the satellite fpnf^{p_{n}} starting with an arc which is fixed by fpnf^{p_{n}} and then "rotate" this arc by a branch of fpn1f^{-p_{n-1}} (for which the same β\beta point is also a fixed point, see (C)). Let γ(x):=γpn,1,0,1(x)\gamma^{\prime}(x):=\gamma_{p_{n},1,0,1}(x) where the latter curve is defined in Lemma 4.3. Then properties (1)-(3) of the curve γ(x)\gamma(x) are satisfied also for γ(x)\gamma^{\prime}(x). In particular, γ(x)\gamma^{\prime}(x) is homotopic to Rtn,jn(x)R_{t_{n,j_{n}(x)}}.

As both t~n,jn(x),tn,jn(x)\tilde{t}_{n,j_{n}(x)},t_{n,j_{n}(x)} are external arguments of βn,jn(x)\beta_{n,j_{n}(x)} which is a pn1p_{n-1}-periodic point of ff, there is i{1,,qn1}i\in\{1,\cdots,q_{n}-1\} such that σipn1(t~n,jn(x))=tn,jn(x)\sigma^{ip_{n-1}}(\tilde{t}_{n,j_{n}(x)})=t_{n,j_{n}(x)}. Now we use that xEϵ,pn1,Qx\in E_{\epsilon,p_{n-1},Q} and that qn<Qq_{n}<Q to prove (4.3). Indeed, for each k={1,,qn}k=\{1,\cdots,q_{n}\}, since f:JJf:J_{\infty}^{\prime}\to J_{\infty}^{\prime} is a homeomorphism and xkpn1Eϵx_{-kp_{n-1}}\in E_{\epsilon}, we have: gpn=g(qnk)pn1gkpn1g_{p_{n}}=g_{(q_{n}-k)p_{n-1}}\circ g_{kp_{n-1}}. Hence, if β=gkpn1(βn,jn(x))\beta^{\prime}=g_{kp_{n-1}}(\beta_{n,j_{n}(x)}), then βn,jn(x)=g(qn1k)pn1(β)\beta_{n,j_{n}(x)}=g_{(q_{n-1}-k)p_{n-1}}(\beta^{\prime}) implying that β=f(qnk)pn1(βn,jn(x))=βn,jn(x)\beta^{\prime}=f^{(q_{n}-k)p_{n-1}}(\beta_{n,j_{n}(x)})=\beta_{n,j_{n}(x)}. Then βn,jn(x),xkpn1gkpn1(B(x,ρ))\beta_{n,j_{n}(x)},x_{-kp_{n-1}}\in g_{kp_{n-1}}(B(x,\rho)) which along with Proposition 3.1, part (2) imply (4.3).

In turn, (4.3) implies that, provided nn is big, gkpn1:B(y,ρ/2)B(y,ρ/2)g_{kp_{n-1}}:B(y,\rho/2)\to B(y,\rho/2) uniformly in k=0,1,,qnk=0,1,\cdots,q_{n} where yy is either βn,jn(x)\beta_{n,j_{n}(x)} or xkpn1x_{-kp_{n-1}}.

Now we consider a curve gip~n(γ(x))g_{i\tilde{p}_{n}}(\gamma^{\prime}(x)) that starts at xip~nx_{-i\tilde{p}_{n}} and tends to βn,jn(x)\beta_{n,j_{n}(x)}. By Proposition 3.1 coupled with (4.3), one can join xipn1x_{-ip_{n-1}} by an arc of the ray Rt(xipn1)R_{t(x_{-ip_{n-1}})} inside of B(x,ρ/2)B(x,\rho/2) up to a point of level C~/4\tilde{C}/4. This will be the required curve γ~(x)\tilde{\gamma}(x). To get the curve γ(x)\gamma(x) we modify γ(x)=γpn,1,0,1(x)=l0gpnl(L0L1)\gamma^{\prime}(x)=\gamma_{p_{n},1,0,1}(x)=\cup_{l\geq 0}g_{p_{n}}^{l}(L_{0}\cup L_{1}) by cutting off the arc L0L_{0} of an equipotential: γ(x)=γ(x)L0\gamma(x)=\gamma^{\prime}(x)\setminus L_{0} (see Lemma 4.3 for details about L0L_{0}). Properties (1)-(5) follow.

Given a point x=x0x=x^{0} and nn such that xfj(Jn)Eϵ,pn,1x\in f^{j}(J_{n})\cap E_{\epsilon,p_{n},1}, where j=jn(x)j=j_{n}(x), let x1=xpnx^{1}=x_{-p_{n}} and t(x0)t(x^{0}), t(x1)t(x^{1}) the arguments of x0x^{0}, x1x^{1} as in Proposition 3.1. We call xx nn-friendly if t(x0)t(x^{0}) and t(x1)t(x^{1}) lie in the same component of sn,js_{n,j} and nn-unfriendly otherwise (or simply friendly and unfriendly if nn is clear from the context). The name reflects the fact that for an nn-friendly point xx the condition (I) of Lemma 4.5 always holds for x1=xx^{1}=x and x2=xpnx^{2}=x_{-p_{n}}, so Lemma 4.5 always applies.

When the rotation number of αn\alpha_{n} is equal to 1/21/2 we have:

Lemma 4.7.

There is C~3>0\tilde{C}_{3}>0 (depending only on fixed ϵ\epsilon and ρ\rho) as follows. Suppose that, for some n>n~n>\tilde{n}, the rotation number of the separating fixed point αn\alpha_{n} is equal to 1/21/2. Let z=z0fj(Jn)Eϵ,pn,3z=z^{0}\in f^{j}(J_{n})\cap E_{\epsilon,p_{n},3} and zi=zipnz^{i}=z_{-ip_{n}}, i=1,2,3i=1,2,3. Assume that all three points z0,z1,z2z^{0},z^{1},z^{2} are nn-unfriendly.

Then there exist two (semi-open) curves γn1/2(z)\gamma^{1/2}_{n}(z) and γ~n1/2(z)\tilde{\gamma}^{1/2}_{n}(z) consisting of arcs of rays and equipotentials with the following properties:

(i) γn1/2(z)B(z,ρ)\gamma^{1/2}_{n}(z)\subset B(z,\rho), γ~n1/2(z)B(z1,ρ)\tilde{\gamma}^{1/2}_{n}(z)\subset B(z^{1},\rho), moreover, arguments of points of γn1/2(z)\gamma^{1/2}_{n}(z) lie in one ’window’ of sn,js_{n,j} while arguments of points of γ~n1/2(x)\tilde{\gamma}^{1/2}_{n}(x) lie in another ’window’ of sn,js_{n,j},

(ii) γn1/2(z)\gamma^{1/2}_{n}(z) and γ~n1/2(z)\tilde{\gamma}^{1/2}_{n}(z) converge to a common point αn,j\alpha^{*}_{n,j} which is a fixed point of fpn:fj(Jn)fj(Jn)f^{p_{n}}:f^{j}(J_{n})\to f^{j}(J_{n}) (i.e., αn,j\alpha^{*}_{n,j} is either the non-separating fixed point βn,j\beta_{n,j} or the separating fixed point αn,j\alpha_{n,j},

(iii) starting points of γn1/2(z),γ~n1/2(z)\gamma^{1/2}_{n}(z),\tilde{\gamma}^{1/2}_{n}(z) have equal Green level which is bigger than C~3\tilde{C}_{3},

(iv) zkαn,j0z^{k}-\alpha^{*}_{n,j}\to 0, 0k30\leq k\leq 3, as nn\to\infty.

Proof.

As zEϵz\in E_{\epsilon}, lengths of ’windows’ of sn,jn(z)s_{n,j_{n}(z)} tend uniformly to zero as nn\to\infty. It follows from the definition of friendly-unfriendly points that t(z0),t(z2)t(z^{0}),t(z^{2}) are in one ’window’ of sn,js_{n,j} and t(z1),t(z3)t(z^{1}),t(z^{3}) are in another ’window’ of sn,js_{n,j}. Therefore, condition (I) of Lemma 4.3 holds for each pair z0,z2z^{0},z^{2} and z1,z3z^{1},z^{3}. Now, apply Lemma 4.3 to zEϵ,pn,3z\in E_{\epsilon,p_{n},3}, first, with i=0i=0, j=2j=2, and then with i=1i=1, j=3j=3. Let γn1/2(z)=γpn,3,0,2(z)\gamma^{1/2}_{n}(z)=\gamma_{p_{n},3,0,2}(z) and γ~n1/2(z)=γpn,3,1,3(z)\tilde{\gamma}^{1/2}_{n}(z)=\gamma_{p_{n},3,1,3}(z). Then (i),(iii) hold. To check (ii), note that these curves converge to some points α\alpha,α~fj(Jn)\tilde{\alpha}\in f^{j}(J_{n}) which are fixed by f2pnf^{2p_{n}} On the other hand, since the rotation number of αn\alpha_{n} is 1/21/2, fpn:fj(Jn)fj(Jn)f^{p_{n}}:f^{j}(J_{n})\to f^{j}(J_{n}) has no 22-cycle. Therefore, one must have either α=α~=βn,j\alpha=\tilde{\alpha}=\beta_{n,j} or α=α~=αn,j\alpha=\tilde{\alpha}=\alpha_{n,j}, i.e., (ii) holds too. As t(z0)t(z2)0t(z^{0})-t(z^{2})\to 0 and t(z1)t(z3)0t(z^{1})-t(z^{3})\to 0 as nn\to\infty, z0z2,z1z30z^{0}-z^{2},z^{1}-z^{3}\to 0, too, by Lemma 4.1. Besides, by (4.1), z2α,z3α~0z^{2}-\alpha,z^{3}-\tilde{\alpha}\to 0 as nn\to\infty. As α=α~=αn,j\alpha=\tilde{\alpha}=\alpha^{*}_{n,j}, (iv) also follows. ∎

The following is a consequence of Lemmas 4.3 and 4.7:

Lemma 4.8.

Let n>n~n>\tilde{n}. Assume that fpnf^{p_{n}} is satellite and doubling, i.e., βn=αn1\beta_{n}=\alpha_{n-1} and the rotation number of αn1\alpha_{n-1} is equal to 1/21/2 (in particular, pn=2pn1p_{n}=2p_{n-1}). For some 1jpn11\leq j\leq p_{n-1}, denote J:=fj(Jn1)J:=f^{j}(J_{n-1}). Let J1:=fj(Jn)J^{1}:=f^{j}(J_{n}), J0:=fj+pn1(Jn)J^{0}:=f^{j+p_{n-1}}(J_{n}) be the two small Julia sets of the next level nn which are contained in JJ (note that J0J^{0} contains the critical point and J1J^{1} contains the critical value of the map F:=fpn1:JJF:=f^{p_{n-1}}:J\to J). Let xJ1Eϵx\in J^{1}\cap E_{\epsilon} be such that all its 55 forward iterates xkpn1=Fk(x)Eϵx_{kp_{n-1}}=F^{k}(x)\in E_{\epsilon}, k=1,2,3,4,5k=1,2,3,4,5. Then there exist two simple semi-open curves Γn1/2(x)\Gamma_{n}^{1/2}(x), Γn1/2(x)\Gamma_{n}^{1/2}(x) consisting of arcs of rays and equipotentials that satisfy essentially conclusions of the previous lemma where nn is replaced by n1n-1, i.e.:

(i) Γn1/2(x),Γ~n1/2(x)B(x,3/2ρ)\Gamma^{1/2}_{n}(x),\tilde{\Gamma}^{1/2}_{n}(x)\subset B(x,3/2\rho), moreover, arguments of points of Γn1/2(x)\Gamma^{1/2}_{n}(x) lie in one ’window’ of sn1,jn1(x)s_{n-1,j_{n-1}(x)} while arguments of points of Γ~n1/2(x)\tilde{\Gamma}^{1/2}_{n}(x) lie in another ’window’ of sn1,jn1(x)s_{n-1,j_{n-1}(x)},

(ii) Γn1/2(x)\Gamma^{1/2}_{n}(x) and Γ~n1/2(x)\tilde{\Gamma}^{1/2}_{n}(x) converge to a common point βn1,jn1(x)\beta^{*}_{n-1,j_{n-1}(x)} which is a fixed point of fpn1:fj(Jn1)fj(Jn1)f^{p_{n-1}}:f^{j}(J_{n-1})\to f^{j}(J_{n-1}) (i.e., βn1,jn1(x)\beta^{*}_{n-1,j_{n-1}(x)} is either the non-separating fixed point βn1,jn1(x)\beta_{n-1,j_{n-1}(x)} or the separating fixed point αn1,jn1(x)\alpha_{n-1,j_{n-1}(x)},

(iii) starting points of Γn1/2(x),Γ~n1/2(x)\Gamma^{1/2}_{n}(x),\tilde{\Gamma}^{1/2}_{n}(x) have equal Green level which is bigger than C~3\tilde{C}_{3},

(iv) xkpn1βn1,jn1(x)0x_{kp_{n-1}}-\beta^{*}_{n-1,j_{n-1}(x)}\to 0, 0k30\leq k\leq 3 as nn\to\infty uniformly in xx.

Remark 4.9.

Condition Fk(x)EϵF^{k}(x)\in E_{\epsilon}, 0k50\leq k\leq 5, is equivalent to the following: xf5pn1(Eϵ,pn1,6)x\in f^{-5p_{n-1}}(E_{\epsilon,p_{n-1},6}).

Proof.

To fix the idea let’s replace fpn1:fj(Jn1)fj(Jn1)f^{p_{n-1}}:f^{j}(J_{n-1})\to f^{j}(J_{n-1}), using a conjugacy to a quadratic polynomial, by a quadratic polynomial (denoted also by FF) so that now F:JJF:J\to J where J=J(F)J=J(F) and F2F^{2} is satellite with two small Julia sets J0J^{0}, J1J^{1} that meet at the α\alpha-fixed point of FF and rays of arguments 1/31/3, 2/32/3 land at α\alpha . Here 0J00\in J^{0}, F(0)J1F(0)\in J^{1}, F:J1J0F:J^{1}\to J^{0} is a homeomorphism while F:J0J1F:J^{0}\to J^{1} is a two-to-one map. If a ray RtR_{t} of FF has its accumulation set in J1J^{1} then t[1/3,5/12][7/12,2/3]t\in[1/3,5/12]\cup[7/12,2/3] and if RtR_{t} accumulates in J0J^{0} then t[1/6,1/3][2/3,5/6]t\in[1/6,1/3]\cup[2/3,5/6]. This implies that if RtR_{t} lands at xJ1x\in J^{1} and tt lies in one of the two ’windows’ [0,1/2)[0,1/2), (1/2,1](1/2,1] then Rσ(t)R_{\sigma(t)} lands at J0J^{0} where σ(t)\sigma(t) must be in a different ’window’ (in other words, points of J0J^{0} are ’unfriendly’). Coming back to fpn1f^{p_{n-1}} this means that, for xJ1x\in J^{1}, t(x),t(F(x))t(x),t(F(x)) are always in different components (where by ’component’ we mean a component of sn1,js_{n-1,j}). Besides, for yJJy\in J_{\infty}\cap J, yy and F(y)F(y) are always in different JiJ^{i}, i=0,1i=0,1. This leaves us with the only possibilities:

(i) t(F(x)),t(F2(x))t(F(x)),t(F^{2}(x)) are in different components; this implies that t(x),t(F(x))t(x),t(F(x)) are in different components and t(F(x)),t(F2(x))t(F(x)),t(F^{2}(x)) are in different components, that is, points F3(x),F2(x),F(x)F^{3}(x),F^{2}(x),F(x) are all unfriendly;

(ii) t(F(x)),t(F2(x))t(F(x)),t(F^{2}(x)) are in the same components; there are two subcases:

(ii’) t(F3(x)),t(F4(x))t(F^{3}(x)),t(F^{4}(x)) are in different components, i.e., (i) holds with xx replaced by F2(x)F^{2}(x) which implies that F5(x),F4(x),F3(x)F^{5}(x),F^{4}(x),F^{3}(x) are all unfriendly;

(ii”) t(F3(x)),t(F4(x))t(F^{3}(x)),t(F^{4}(x)) are in the same component which then means that F2(x)F^{2}(x) and F4(x)F^{4}(x) are both friendly.

In the case (i) and (ii’), apply Lemma 4.7 with n1n-1 instead of nn to z=F3(x)z=F^{3}(x) and to z=F5(x)z=F^{5}(x), respectively, letting Γn1/2(x)=γn11/2(F3(x))\Gamma^{1/2}_{n}(x)=\gamma^{1/2}_{n-1}(F^{3}(x)), Γ~n1/2(x)=γ~n11/2(F3(x))\tilde{\Gamma}^{1/2}_{n}(x)=\tilde{\gamma}^{1/2}_{n-1}(F^{3}(x)) and Γn1/2(x)=γn11/2(F5(x))\Gamma^{1/2}_{n}(x)=\gamma^{1/2}_{n-1}(F^{5}(x)), Γ~n1/2(x)=γ~n11/2(F5(x))\tilde{\Gamma}^{1/2}_{n}(x)=\tilde{\gamma}^{1/2}_{n-1}(F^{5}(x)), respectively. In the case (ii”), apply Lemma 4.3 with pn1p_{n-1}, q=1q=1, i=0,j=0i=0,j=0, first, to the point F2(x)F^{2}(x) and then to the point F4(x)F^{4}(x) letting Γn1/2(x)=γpn1,1,0,1(F2(x))\Gamma^{1/2}_{n}(x)=\gamma_{p_{n-1},1,0,1}(F^{2}(x)), Γ~n1/2(x)=γpn1,1,0,1(F4(x))\tilde{\Gamma}^{1/2}_{n}(x)=\gamma_{p_{n-1},1,0,1}(F^{4}(x)). ∎

5. Proof of Theorem 1.1

Every invariant probability measure with positive Lyapunov exponent has an ergodic component with positive exponent. So let μ\mu be such an ergodic ff-invariant measure component supported in JJ_{\infty}. First, we have the following general

Remark 5.1.

Given xJx\in J^{\prime}_{\infty} such that r~(x)>0\tilde{r}(x)>0 as in Proposition 2.3, and given nn, the set Jn,x=fjn(x)(Jn)J_{n,x}=f^{j_{n}(x)}(J_{n}) cannot be covered by B(x,r~(x))B(x,\tilde{r}(x)) because otherwise the branch gpn:B(x,r~(x))𝐂g_{p_{n}}:B(x,\tilde{r}(x))\to{\bf C} of fpnf^{-p_{n}}, which sends xx to xpnJn,xx_{-p_{n}}\in J_{n,x} meets the critical value along the way so cannot be well-defined. Thus diamJn,x>r~(x)\operatorname{diam}J_{n,x}>\tilde{r}(x), for each nn, and diamKx=limdiamJn,xr~(x)\operatorname{diam}K_{x}=\lim\operatorname{diam}J_{n,x}\geq\tilde{r}(x). In particular, diamJn,xr(ϵ)\operatorname{diam}J_{n,x}\geq r(\epsilon) for all xEϵx\in E_{\epsilon} and nn.

We need to prove that ff has finitely many satellite renormalizations. Assuming the contrary, let 𝒮\mathcal{S} be an infinite subsequence such that fpnf^{p_{n}} is a satellite renormalization of ff for each n𝒮n\in\mathcal{S}.

We arrive at a contradiction by considering, roughly speaking, two alternative situations. In the first one, we find a point xEϵx\in E_{\epsilon}, nn, and two curves in BA()B\cap A(\infty) where B:=B(x,r~(x))B:=B(x,\tilde{r}(x)) that tend to the β\beta-fixed points of Jn,xJ_{n,x} such that another ends of the curves can be joined by an arc of equipotential in BB thus ’surrounding’ Jn,xJ_{n,x} by a ’triangle’ in BB which would be a contradiction as in Remark 5.1. The second situation is when the first one does not happen. Then we use several curves to ’surround’ Jn,xJ_{n,x} by a ’quadrilateral’ in BB, ending by the same conclusion. The curves we use have been constructed in Lemmas 4.5, 4.8.

The first situation happens in cases A and B1, and the second one in B2.

Case A: 𝒮\mathcal{S} contains an infinite sequence of indices of non-doubling renormalizations. Passing to a subsequence one can assume that fpnf^{p_{n}} is satellite not doubling for every n𝒮n\in\mathcal{S}.

Fix ζ=1/4\zeta=1/4. By Lemma 2.2, for each n𝒮n\in\mathcal{S} and each j=1,,[ζpn]j=1,\cdots,[\zeta p_{n}], the map σj1:Sn,1Sn,j\sigma^{j-1}:S_{n,1}\to S_{n,j} is a homeomorphism and the length |Sn,j|0|S_{n,j}|\to 0 as nn\to\infty uniformly in jj. Fix NN such that |Sn,j|<δ|S_{n,j}|<\delta for each n>Nn>N, n𝒮n\in\mathcal{S}. For n𝒮n\in\mathcal{S}, let

𝒞n={fj(Jn)|1j[ζpn]}.\mathcal{C}_{n}=\{f^{j}(J_{n})|1\leq j\leq[\zeta p_{n}]\}.

Let n,m𝒮n,m\in\mathcal{S}, m<nm<n. Denote p=pm,p~=pnp=p_{m},\tilde{p}=p_{n}, q=pn/pmq=p_{n}/p_{m}. The intersection 𝒞n𝒞m\mathcal{C}_{n}\cap\mathcal{C}_{m} contains all fj+kp(Jn)f^{j+kp}(J_{n}) with 1j[ζp]1\leq j\leq[\zeta p], j+kp[ζp~]j+kp\leq[\zeta\tilde{p}]. Hence,

#(𝒞n𝒞m)j=1[ζp][ζqjp][ζq1][ζp]\#(\mathcal{C}_{n}\cap\mathcal{C}_{m})\geq\sum_{j=1}^{[\zeta p]}[\zeta q-\frac{j}{p}]\geq[\zeta q-1][\zeta p]\geq
p~(ζp1pζq1qζq)ζ2p~\tilde{p}(\frac{\zeta p-1}{p}\frac{\zeta q-1}{q}-\frac{\zeta}{q})\sim\zeta^{2}\tilde{p}

as p,qp,q\to\infty. Therefore, fixing κ=ζ2/2\kappa=\zeta^{2}/2=1/8, there are m0,k0m_{0},k_{0} such that for each n,m𝒮n,m\in\mathcal{S}, m>m0m>m_{0}, n>m+k0n>m+k_{0},

μ(𝒞n𝒞m)>κ.\mu(\mathcal{C}_{n}\cap\mathcal{C}_{m})>\kappa.

Fix such n,mn,m, assume also that m>max{N,n}m>\max\{N,n_{*}\} where nn_{*} is defined in Lemma 4.5 and recall the set

Xn=Eϵ,pn,4Eϵ,pn1,Q=i=03fipn(Eϵ)k=0Q1fkpn1(Eϵ).X_{n}=E_{\epsilon,p_{n},4}\cap E_{\epsilon,p_{n-1},Q}=\cap_{i=0}^{3}f^{ip_{n}}(E_{\epsilon})\cap_{k=0}^{Q-1}f^{kp_{n-1}}(E_{\epsilon}).

Since μ(Xn)>1(Q+4)ϵ>1κ\mu(X_{n})>1-(Q+4)\epsilon>1-\kappa, there is xXn𝒞n𝒞mx\in X_{n}\cap\mathcal{C}_{n}\cap\mathcal{C}_{m} and, by the choice of nn, the assumption (II) of Lemma 4.5 holds for xx. Therefore, there exist two simple semi-open curves γ(x)\gamma(x) and γ~(x)\tilde{\gamma}(x) that satisfy the following properties: γ(x)\gamma(x) and γ~(x)\tilde{\gamma}(x) tend to βn,jn(x)\beta_{n,j_{n}(x)}, γ(x),γ~(x)B(x,ρ)A()\gamma(x),\tilde{\gamma}(x)\subset B(x,\rho)\cap A(\infty) and γ(x),γ~(x)\gamma(x),\tilde{\gamma}(x) consist of arcs of equipotentials and external rays; the starting point b1b_{1} of γ(x)\gamma(x) and the starting point b~1\tilde{b}_{1} of γ~(x)\tilde{\gamma}(x) have equal levels which is at least C~/4\tilde{C}/4; γ(x)\gamma(x), γ~(x)Un1,jn1(x)\tilde{\gamma}(x)\subset U_{n-1,j_{n-1}(x)}; finally, being completed by their common limit point βn,jn(x)\beta_{n,j_{n}(x)} and arcs of rays from b1γ(x)b_{1}\in\gamma(x) to \infty and from b~1γ~(x)\tilde{b}_{1}\in\tilde{\gamma}(x) to \infty, they split the plane into two domains such that one of them contains I:=Jn,xβn,jn(x)I:=J_{n,x}\setminus\beta_{n,j_{n}(x)} and another one contains all other iterates fkpn1(I)f^{kp_{n-1}}(I), 1kqn11\leq k\leq q_{n}-1. Now, since Un1,jn1(x)Um,jm(x)U_{n-1,j_{n-1}(x)}\subset U_{m,j_{m}(x)} and by the choice of mm, the distance between arguments of the points b1b_{1} and b~1\tilde{b}_{1} inside of Sn1,jn1(x)S_{n-1,j_{n-1}(x)} is less than δ\delta. By the definition of δ\delta, b1b_{1} and b~1\tilde{b}_{1} can be joined by an arc AnA_{n} of equipotential inside of B(x,ρ)Un1,jn1(x)B(x,\rho)\cap U_{n-1,j_{n-1}(x)}. Consider a Jordan domain ZnZ_{n} with the boundary to be the arc AnA_{n} and semi-open curves γ(x)\gamma(x), γ~(x)\tilde{\gamma}(x) completed by their common limit point βn,jn(x)\beta_{n,j_{n}(x)}. Then ZnB(x,ρ)Z_{n}\subset B(x,\rho). By the properties of the curves , Znβn,jn(x)Z_{n}\cup\beta_{n,j_{n}(x)} contains either Jn,xJ_{n,x} or its iterate fkpn1(Jn,x)f^{kp_{n-1}}(J_{n,x}), for some 1kqn11\leq k\leq q_{n}-1, in a contradiction with Remark 5.1.

Complementary to A is

Case B: for all big nn, every satellite renormalization fpnf^{p_{n}} is doubling, i.e., βn=αn1\beta_{n}=\alpha_{n-1} and pn=2pn1p_{n}=2p_{n-1} for every n𝒮n\in\mathcal{S}.

Let Yn1=Eϵ,pn1,6Y_{n-1}=E_{\epsilon,p_{n-1},6} and Y~n1=f5pn1(Yn1)\tilde{Y}_{n-1}=f^{-5p_{n-1}}(Y_{n-1}). Note that μ(Yn1)=μ(Y~n1)>16ϵ\mu(Y_{n-1})=\mu(\tilde{Y}_{n-1})>1-6\epsilon.

For every n𝒮n\in\mathcal{S}, let

Ln={0<j<pn1|μ(fj(Jn1)Y~n1)>1212ϵpn1}.L_{n}=\{0<j<p_{n-1}|\mu(f^{j}(J_{n-1})\cap\tilde{Y}_{n-1})>\frac{1-2^{12}\epsilon}{p_{n-1}}\}.

As μ(Y~n1)>16ϵ\mu(\tilde{Y}_{n-1})>1-6\epsilon, it follows,

#Ln>(13/211)pn1.\#L_{n}>(1-3/2^{11})p_{n-1}.

Since we are in case B, each fj(Jn1)f^{j}(J_{n-1}) contains precisely two small Julia sets fj(Jn),fj+pn1(Jn)f^{j}(J_{n}),f^{j+p_{n-1}}(J_{n}) of the next level nn each of them of measure 1/(2pn1)1/(2p_{n-1}). Hence, the measure of intersection of each of these small Julia sets with Y~n1\tilde{Y}_{n-1} is bigger than (1/2210ϵ)/pn1>0(1/2-2^{10}\epsilon)/p_{n-1}>0. By Lemma 4.8, choosing for every jLnj\in L_{n} a point xjfj(Jn1)Y~n1x_{j}\in f^{j}(J_{n-1})\cap\tilde{Y}_{n-1} we get a pair of curves Γn1/2(xj),Γ~n1/2(xj)\Gamma^{1/2}_{n}(x_{j}),\tilde{\Gamma}^{1/2}_{n}(x_{j}) consisting of arcs of rays and equipotentials as follows: (i) Γn1/2(xj),Γ~n1/2(xj)B(xj,3/2ρ)\Gamma^{1/2}_{n}(x_{j}),\tilde{\Gamma}^{1/2}_{n}(x_{j})\subset B(x_{j},3/2\rho), moreover, arguments of points of Γn1/2(xj)\Gamma^{1/2}_{n}(x_{j}) lie in one ’window’ of sn1,js_{n-1,j} while arguments of points of Γ~n1/2(xj)\tilde{\Gamma}^{1/2}_{n}(x_{j}) lie in another ’window’ of sn1,js_{n-1,j}, (ii) Γn1/2(xj)\Gamma^{1/2}_{n}(x_{j}) and Γ~n1/2(xj)\tilde{\Gamma}^{1/2}_{n}(x_{j}) converge to a common point βn1,j\beta^{*}_{n-1,j} which is a fixed point of fpn1:fj(Jn1)fj(Jn1)f^{p_{n-1}}:f^{j}(J_{n-1})\to f^{j}(J_{n-1}) (i.e., βn1,j\beta^{*}_{n-1,j} is either the non-separating fixed point βn1,j\beta_{n-1,j} or the separating fixed point αn1,j\alpha_{n-1,j}, (iii) start points of Γn1/2(xj),Γ~n1/2(xj)\Gamma^{1/2}_{n}(x_{j}),\tilde{\Gamma}^{1/2}_{n}(x_{j}) have equal Green level which is bigger than C~3\tilde{C}_{3}, (iv) xjβn1,j0x_{j}-\beta^{*}_{n-1,j}\to 0 as nn\to\infty uniformly in jj and xjx_{j}. We add one more property as follows. Let

Γn,j=Γn1/2(xj)βn1,jΓ~n1/2(xj).\Gamma_{n,j}=\Gamma^{1/2}_{n}(x_{j})\cup\beta^{*}_{n-1,j}\cup\tilde{\Gamma}^{1/2}_{n}(x_{j}).

Then: (v) Γn,j\Gamma_{n,j} is a simple curve; the level of zΓn,j{βn1,j}z\in\Gamma_{n,j}\setminus\{\beta^{*}_{n-1,j}\} is positive and decreases (not strickly) from C~3\tilde{C}_{3} to zero along Γn1/2(xj)\Gamma^{1/2}_{n}(x_{j}) and then increases from zero to C~3\tilde{C}_{3} along Γ~n1/2(xj)\tilde{\Gamma}^{1/2}_{n}(x_{j}); moreover, if j1,j2Lnj_{1},j_{2}\in L_{n}, j1j2j_{1}\neq j_{2}, then Γn,1\Gamma_{n,1}, Γn,j2\Gamma_{n,j_{2}} are either disjoint or meet at the unique common point βn1,j1=βn1,j2\beta_{n-1,j_{1}}=\beta_{n-1,j_{2}} and then disjoint with all others γn1,j\gamma_{n-1,j}, jj1,j2j\neq j_{1},j_{2}. This is because, by property (i), Γn,jUn1,j¯\Gamma_{n,j}\subset\overline{U_{n-1,j}} where (by (C), Sect 2) any two Un1,j¯\overline{U_{n-1,j}}, Un1,j~¯\overline{U_{n-1,\tilde{j}}}, jj~j\neq\tilde{j}, are either disjoint or meet at β:=βn1,j=βn1,j~\beta:=\beta_{n-1,j}=\beta_{n-1,\tilde{j}} in which case fpn1f^{p_{n-1}} is satellite. In the considered case, any satellite is doubling so ββn1,i\beta\neq\beta_{n-1,i} for all ii different from j,j~j,\tilde{j}.

We assign, for the use below, a ’small’ Julia set In,jI_{n,j} to each Γn,j\Gamma_{n,j} as follows: by the construction, βn1,j\beta^{*}_{n-1,j} is either the β\beta-fixed point of fj(Jn1)f^{j}(J_{n-1}), or the α\alpha-fixed point of fj(Jn1)f^{j}(J_{n-1}). In the former case, let In,j=fj(Jn1)I_{n,j}=f^{j}(J_{n-1}), and in the latter case, In,j=fj(Jn)I_{n,j}=f^{j}(J_{n}) (one of the two small Julai sets of the next level nn that are contained in fj(Jn1)f^{j}(J_{n-1}). Observe that In,jΓn1,j={βn1,j}I_{n,j}\cap\Gamma_{n-1,j}=\{\beta^{*}_{n-1,j}\} and is disjoint with any other Γn,j\Gamma_{n,j^{\prime}} provided Γn,j\Gamma_{n,j}, Γn,j\Gamma_{n,j^{\prime}} are disjoint.

There are two subcases B1-B2 to distinguish depending on whether arguments of end points of Γm,j\Gamma_{m,j} become close or not. If yes, then one can join the end points of some Γn,j\Gamma_{n,j} by an arc of equipotential inside of B(xj,2ρ)Γm,jB(x_{j},2\rho)\supset\Gamma_{m,j} to surround a small Julia set as in case A, which would lead to a contradiction. If no, the construction is more subtle: we build a domain (’quadrilateral’) in B(xj,2ρ)B(x_{j},2\rho) bounded by two disjoint curves as above completed by two arcs of equipotential that join ends of different curve, so that the obtained quadrilateral again contains a small Julia set.

B1: lim infn𝒮,jLn|Sn1,j|<δ\liminf_{n\in\mathcal{S},j\in L_{n}}|S_{n-1,j}|<\delta.

By property (i) listed above and the definition of δ\delta, there are a sequence (nk)𝒮(n_{k})\subset\mathcal{S}, jkLnkj_{k}\in L_{n_{k}} and xjkx_{j_{k}} as above, such that two ends of each curve Γnk,jk\Gamma_{n_{k},j_{k}} can be joined inside of B(xjk,ρ)B(x_{j_{k}},\rho) by an arc AkA^{k} of equipotential of fixed level C~3\tilde{C}_{3} such that all arguments of points in AkA^{k} belong to Snk1,jkS_{n_{k}-1,j_{k}}. Then we arrive at a contradiction as in case A.

Refer to caption
Figure 2. Left: Case A and Case B1, right: Case B2.

B2: |Sn1,j|δ|S_{n-1,j}|\geq\delta for all big n𝒮n\in\mathcal{S} and all jLnj\in L_{n}.

Fix n,m𝒮n,m\in\mathcal{S}, mn3m-n\geq 3. Define a subset of LnL_{n} as follows:

Lnm={0<j<pn1|μ(fj(Jn1)(Y~n1Y~m1))>1212ϵpn1}.L_{n}^{m}=\{0<j<p_{n-1}|\mu(f^{j}(J_{n-1})\cap(\tilde{Y}_{n-1}\cap\tilde{Y}_{m-1}))>\frac{1-2^{12}\epsilon}{p_{n-1}}\}.

As μ(Y~n1Y~m1)>112ϵ\mu(\tilde{Y}_{n-1}\cap\tilde{Y}_{m-1})>1-12\epsilon,

#Lnm>(13/210)pn1.\#L_{n}^{m}>(1-3/2^{10})p_{n-1}.

For each jLnmj\in L_{n}^{m} we define further

Ln,jm={0<k<pn1|fk(Jm1)L_{n,j}^{m}=\{0<k<p_{n-1}|f^{k}(J_{m-1})\subset
fj(Jn1),μ(fk(Jm1)(Y~n1Y~m1))>1216ϵpm1}.f^{j}(J_{n-1}),\mu(f^{k}(J_{m-1})\cap(\tilde{Y}_{n-1}\cap\tilde{Y}_{m-1}))>\frac{1-2^{16}\epsilon}{p_{m-1}}\}.

Then

#Ln,jm5\#L_{n,j}^{m}\geq 5

as otherwise #Ln,jm4\#L_{n,j}^{m}\leq 4 and, therefore, (1212ϵ)/pn1<4/pm1+(pm1/pn14)(1216ϵ)/pm1=218ϵ/pm1+(1216ϵ)/pn1(1-2^{12}\epsilon)/p_{n-1}<4/p_{m-1}+(p_{m-1}/p_{n-1}-4)(1-2^{16}\epsilon)/p_{m-1}=2^{18}\epsilon/p_{m-1}+(1-2^{16}\epsilon)/p_{n-1}, i.e., pm1/pn1<218ϵ/(216ϵ212ϵ)=4/(124)<8p_{m-1}/p_{n-1}<2^{18}\epsilon/(2^{16}\epsilon-2^{12}\epsilon)=4/(1-2^{-4})<8, a contradiction because pm1/pn12mn23p_{m-1}/p_{n-1}\geq 2^{m-n}\geq 2^{3}.

Fix jLnmj\in L_{n}^{m}. Thus Ln,jmL_{n,j}^{m} contains 55 pairwise different indices kik_{i}, 1k51\leq k\leq 5. As Ln,jmLmL_{n,j}^{m}\subset L_{m}, we find 55 curves Γm1,ki\Gamma_{m-1,k_{i}}. By property (v), if two of them meet, they are disjoint with all others. Therefore, there are at least 33 of them denoted by Γm1,ri\Gamma_{m-1,r_{i}}, i=1,2,3i=1,2,3, which are pairwise disjoint. Let wi,w~m,iw_{i},\tilde{w}_{m,i} be two ends of Γm1,ri\Gamma_{m-1,r_{i}}.

For each i=1,2,3i=1,2,3, arguments of points of wm,i,w~m,iw_{m,i},\tilde{w}_{m,i} lie in different ’windows’ of sm1,ris_{m-1,r_{i}}. On the other hand, by the choice of jj, sm1,risn1,jSn1,js_{m-1,r_{i}}\subset s_{n-1,j}\subset S_{n-1,j}. As nn is big enough, lengths of ’windows’ of sn1,js_{n-1,j} are less than δ\delta. But since we are in case B2, the length of Sn1,jS_{n-1,j} is bigger than δ\delta. One can assume, therefore, that, for i=1,2,3i=1,2,3, arguments of wm,iw_{m,i} lie in one window of sn1,js_{n-1,j} while arguments of w~m,i\tilde{w}_{m,i} are in another window. Therefore, differences of arguments of all wm,iw_{m,i} tend to zero as mm\to\infty, and the same for w~m,i\tilde{w}_{m,i}. As all wm,i,w~m,iEϵw_{m,i},\tilde{w}_{m,i}\in E_{\epsilon}, this implies by Lemma 4.1 that max1i,l3|wm,iwm,l|0\max_{1\leq i,l\leq 3}|w_{m,i}-w_{m,l}|\to 0. This along with property (iv) implies that γm1,riB(wm,1,2ρ)\gamma_{m-1,r_{i}}\subset B(w_{m,1},2\rho), i=1,2,3i=1,2,3, for all big mm. Since, for big mm, differences of arguments of all wm,iw_{m,i} are less than δ\delta, and the same for w~m,i\tilde{w}_{m,i}, one can joint all wm,iw_{m,i} by an arc DmD^{m} of equipotential of level C~3\tilde{C}_{3} and all w~m,i\tilde{w}_{m,i} by an arc D~m\tilde{D}^{m} of equipotential of the same level C~3\tilde{C}_{3} such that Dm,D~mB(w1,2ρ)D^{m},\tilde{D}^{m}\subset B(w_{1},2\rho). Let the end points of DmD^{m} be, say, wm,1w_{m,1} and wm,3w_{m,3}, so that wm,2Dmw_{m,2}\in D^{m} is in between. Since all 33 curves Γm1,ri\Gamma_{m-1,r_{i}}, i=1,2,3i=1,2,3, are pairwise disjoint, the end points of D~m\tilde{D}^{m} have to be then w~m,1\tilde{w}_{m,1} and w~m,3\tilde{w}_{m,3}, so that w~m,2D~m\tilde{w}_{m,2}\in\tilde{D}^{m} is in between. Therefore, we get a ’big’ quadrilateral Πm0B(wm,1,2ρ)\Pi_{m}^{0}\subset B(w_{m,1},2\rho) bounded by Dm,D~m,Γm,1,Γm,3D^{m},\tilde{D}^{m},\Gamma_{m,1},\Gamma_{m,3} where Γm,i:=Γm1,ri\Gamma_{m,i}:=\Gamma_{m-1,r_{i}}, i=1,2,3i=1,2,3. The curve Γm,2\Gamma_{m,2} splits Πm0\Pi_{m}^{0} into two ’small’ quadrilaterals Πm1,Πm2\Pi_{m}^{1},\Pi_{m}^{2} with a common curve Γm,2\Gamma_{m,2} in their boundaries. Recall now that the curve Γm,2\Gamma_{m,2} comes with a small Julia set Im,2I_{m,2} of level either m1m-1 or mm, such that Im,2Γm,2I_{m,2}\cap\Gamma_{m,2} is a single point while Im,2I_{m,2} is disjoint with Γm,1\Gamma_{m,1}, Γm,3\Gamma_{m,3}. Therefore, Im,2Πm0B(wm,1,2ρ)I_{m,2}\subset\Pi_{m}^{0}\subset B(w_{m,1},2\rho), a contradiction with Remark 5.1.

6. Proof of Corollaries 1.3-1.4

Corollary 1.3 follows directly from the following

Proposition 6.1.

Let ff be an infinitely renormalizable quadratic polynomial. Then conditions (1)-(4) are equivalent:

  1. (1)

    f:JJf:J_{\infty}\to J_{\infty} has no invariant probability measure with positive exponent,

  2. (2)

    for every neighborhood WW of PP and every α(0,1)\alpha\in(0,1) there exist m0m_{0} and n0n_{0} such that, for each mm0m\geq m_{0} and xorb(Jn)x\in orb(J_{n}) with nn0n\geq n_{0},

    #{i|0i<m,fi(x)W}m>α;\frac{\#\{i|0\leq i<m,f^{i}(x)\in W\}}{m}>\alpha;

    additionally, f:PPf:P\to P has no invariant probability measure with positive exponent,

  3. (3)

    every invariant probability measure of f:JJf:J_{\infty}\to J_{\infty} is, in fact, supported on PP and has zero exponent,

  4. (4)

    for every invariant probability ergodic measure μ\mu of ff on the Julia set JJ of ff, either supp(μ)J=\operatorname{supp}(\mu)\cap J_{\infty}=\emptyset and its Lyapunov exponent χ(μ)>0\chi(\mu)>0, or supp(μ)P\operatorname{supp}(\mu)\subset P and χ(μ)=0\chi(\mu)=0.

Proof.

(1)\Rightarrow(2). Assume the contrary. Let E=WE=\mathbb{C}\setminus W. Since WW is a neighborhood of a compact set PP, the Euclidean distance d(E,P)>0d(E,P)>0. By a standard normality argument, as all periodic points of ff are repelling, there are λ>1\lambda>1 and k0>0k_{0}>0 such that |(fk)(y)|>λ|(f^{k})^{\prime}(y)|>\lambda whenever y,fk(y)Ey,f^{k}(y)\in E and kk0k\geq k_{0}. As (2) does not hold, find α(0,1)\alpha\in(0,1), a sequence nkn_{k}\to\infty, points xkorb(Jnk)x_{k}\in orb(J_{n_{k}}) and a sequence mkm_{k}\to\infty such that, for each kk,

#{i:0i<mk,fi(xk)E}mkβ:=1α.\frac{\#\{i:0\leq i<m_{k},f^{i}(x_{k})\in E\}}{m_{k}}\geq\beta:=1-\alpha.

Fix a big kk such that βmk>3k0\beta m_{k}>3k_{0} and consider the times 0i1k<i2k<ilkk<mk0\leq i_{1}^{k}<i_{2}^{k}<...i_{l_{k}}^{k}<m_{k} where lk/mkβl_{k}/m_{k}\geq\beta such that fi(xk)Ef^{i}(x_{k})\in E. Let zk=fi1k(xk)z_{k}=f^{i_{1}^{k}}(x_{k}) so that zkEorb(Jn)z_{k}\in E\cap orb(J_{n}). Therefore, by the choice of λ\lambda and k0k_{0}, |(fmki1k)(zk)|λ~mkλ~mki1|(f^{m_{k}-i_{1}^{k}})^{\prime}(z_{k})|\geq\tilde{\lambda}^{m_{k}}\geq\tilde{\lambda}^{m_{k}-i_{1}} where λ~=λβ2k0>1\tilde{\lambda}=\lambda^{\frac{\beta}{2k_{0}}}>1. In this way we get a sequence of measures μk=1mki1ki=0mki1k1δfi(zk)\mu_{k}=\frac{1}{m_{k}-i_{1}^{k}}\sum_{i=0}^{m_{k}-i_{1}^{k}-1}\delta_{f^{i}(z_{k})} such that the Lyapunov exponent of μk\mu_{k} is at least logλ~>0\log\tilde{\lambda}>0. Passing to a subsequence one can assume that {μk}\{\mu_{k}\} converges weak-* to a measure μ\mu. Then μ\mu is an ff-invariant probability measure on J=orb(Jn)J_{\infty}=\cap orb(J_{n}) with the exponent at least logλ~>0\log\tilde{\lambda}>0, a contradiction with (1).

(2)\Rightarrow(3), by the Birkhoff Ergodic Theorem along with [22].

(3)\Rightarrow(4): let μ\mu be as in (4) and U¯P=\overline{U}\cap P=\emptyset for some open set UU with μ(U)>0\mu(U)>0. Let F:UUF:U\to U be the first return map equipped with the induced invariant measure μU\mu_{U}. By the Birkhoff Ergodic Theorem and by an argument as in (1)\Rightarrow(2), the exponent χF(μU)\chi_{F}(\mu_{U}) of FF w.r.t. μU\mu_{U} is strictly positive. Hence, χ(μ)=μ(U)χF(μU)\chi(\mu)=\mu(U)\chi_{F}(\mu_{U}) is positive too. This proves the implication.

And (4) obviously implies (1). ∎

Proof of Corollary 1.4.

If χ¯(x)\overline{\chi}(x) were strictly positive, for some xJx\in J_{\infty}, that would imply, by a standard argument (see the proof of Corollary 1.3), the existence of an ff-invariant measure with positive exponent supported in ω(x)J\omega(x)\subset J_{\infty}, with a contradiction to Theorem 1.1. This proves (1.1). By [14], lim infn1nlog|(fn)(c)|0\liminf_{n\to\infty}\frac{1}{n}\log|(f^{n})^{\prime}(c)|\geq 0. On the other hand, by (1.1), χ¯(c)0\overline{\chi}(c)\leq 0, which proves (1.2). ∎

References

  • [1] A. Blokh, G. Levin, L. Oversteegen, V. Timorin: The modulus of renormalization of rational functions: high periods imply small moduli., 2022, https://arxiv.org/pdf/2205.03157.pdf .
  • [2] L. Carleson, T. Gamelin: Complex Dynamics. Springer-Verlag, 1993.
  • [3] D. Cheraghi, M. Pedramfar, slides of talk at workshop "On geometric complexity of Julia sets", 18-23 March 2018, Będlewo, Poland, and video of talk at virtual workshop “Many faces of renormalization”, Simons Center for Geometry and Physics, 8-12 March 2021.
  • [4] D. Cheraghi, M. Shishikura: Satellite renormalization of quadratic polynomials. https://arxiv.org/abs/1509.07843.
  • [5] A. Douady, J. H. Hubbard: On the dynamics of polynomial-like mappings. Ann. Sci. Ec. Norm. Super. (4) 18 (1985), no. 2, 287-343.
  • [6] M. Feigenbaum: Quantitative universality for a class of nonlinear transformations. Journal of Statistical Physics. 19 (1) (1978), 25–52.
  • [7] J. Hu, Y. Jiang: The Julia set of the Feigenbaum quadratic polynomial is locally connected. Preprint, 1993.
  • [8] J. H. Hubbard: Local connectivity of Julia sets and bifurcation coci: three theorems of J.-C. Yoccoz. In: "Topological Methods in Modern Mathematics." Houston, TX:Publish or Perish, 1993.
  • [9] G. Levin, F. Przytycki: No hyperbolic sets in JJ_{\infty} for infinitely renormalizable quadratic polynomials. Israel J. Math. 251 (2022), no. 2, 635-656.
  • [10] G. Levin: Multipliers of periodic orbits of quadratic polynomials and the parameter plane. Israel Journal of Mathematics 170 (2009), 285-315.
  • [11] G. Levin: Rigidity and non-local connectivity of Julia sets of some quadratic polynomials. Comm. Math. Phys. 304 (2011), 2, 295-328.
  • [12] G. Levin: Addendum to: Rigidity and non-local connectivity of Julia sets of some quadratic polynomials. Comm. Math. Phys. 325 (2014), 3, 1171-1178.
  • [13] G. M. Levin: Bounds for multipliers of periodic points of holomorphic mappings. (Russian) Sibirsk. Mat. Zh. 31 (1990), no. 2, 104–110; translation in Siberian Math. J. 31 (1990), no. 2, 273-278.
  • [14] G. Levin, F. Przytycki, W. Shen: The Lyapunov exponent of holomorphic maps. Invent. Math. 205 (2016), 363-382.
  • [15] M. Lyubich: Feigenbaum-Coullet-Tresser universality and Milnor’s Hairiness Conjecture. Ann. Math. 149 (2) (1999), 319-420.
  • [16] J. Milnor: Dynamics in one complex variable. Princeton University Press, 2006.
  • [17] J. Milnor: Local connectivity of Julia sets: expository lectures. The Mandelbrot set, theme and variations, 67-116, London Math. Soc. Lecture Note Ser., 274, Cambridge Univ. Press, Cambridge, 2000.
  • [18] J. Milnor: Periodic orbits, external rays and the Mandelbrot set: an expository account. "Geometrie Complex et Systemes Dynamiques". Colloque en l’honnoeur d’Adrien Douady (Orsay, 1995). Asterisque 261, 277-331 (2000).
  • [19] C. McMullen: Complex Dynamics and Renormalization. Annals of Mathematical Studies, 135. Princeton Univ. Press, Princeton, New Jersey, 1994.
  • [20] C. McMullen: Renormalization and 3-Manifolds Which Fiber over the Circle. Annals of Mathematical Studies, 142. Princeton Univ. Press, Princeton, New Jersey, 1996.
  • [21] Ch. Pommerenke: On conformal mapping and iteration of rational functions. Complex Variables 5 (2-4) (1986), 117-126.
  • [22] F. Przytycki: Lyapunov characteristic exponents are nonnegative. Proc. Amer. Math. Soc. 119 (1993), no. 1, 309-317.
  • [23] F. Przytycki: Accessibility of typical points for invariant measures of positive Lyapunov exponents for iterations of holomorphic maps. Fund. Math. 144 (1994), no.3, 259-278
  • [24] F. Przytycki: Thermodynamic formalism methods in one-dimensional real and complex dynamics. Proceedings of the International Congress of Mathematicians 2018, Rio de Janeiro, Vol.2, 2081-2106.
  • [25] F. Przytycki, M. Urbański: Conformal Fractals: Ergodic Theory Methods. Cambridge, 2010.
  • [26] M. Rees: One hundred years of complex dynamics. Proc. R. Soc. A 472 (2016).
  • [27] D. E. K. Sorensen: Infinitely renormalizable quadratic polynomials with non-locally connected Julia set. J. Geom. Anal. 10 (2000), 169-208.
  • [28] D. Sullivan: Bounds, quadratic differentials and renormalization conjectures. In: Mathematics into the Twenty-First Century, AMS Centennial Publications 2, Providence, RI: Amer. Math. Soc., 1991.