1 Introduction and main results

We will be concerned with the iteration of transcendental meromorphic functions \(f:\mathbb {C}\rightarrow \widehat{{\mathbb {C}}}:=\mathbb {C}\cup \{\infty \}\). The main objects studied here are the Fatou set F(f), consisting of all \(z\in {\mathbb {C}}\) for which the iterates \(f^k\) of f are defined and form a normal family in some neighborhood of z, and the Julia set \(J(f):=\widehat{{\mathbb {C}}}{\setminus } F(f)\). For an introduction to the dynamics of transcendental meromorphic functions we refer to [10].

The Speiser class \(\mathcal {S}\) consists of all transcendental meromorphic functions f for which the set \({\text {sing}}(f^{-1})\) of singularities of the inverse, i.e., the set of critical and asymptotic values of f, is finite. More precisely, if q denotes the cardinality of \({\text {sing}}(f^{-1})\), then we write \(f\in \mathcal {S}_q\). Equivalently, \(f\in \mathcal {S}\) if there exist finitely many points \(a_1,\dots ,a_q\in \widehat{{\mathbb {C}}}\) such that

$$\begin{aligned} f:\mathbb {C}{\setminus } f^{-1}(\{a_1,\dots ,a_q\})\rightarrow \widehat{{\mathbb {C}}}{\setminus }\{a_1,\dots ,a_q\} \end{aligned}$$

is a covering map. And if q is the minimal number with this property, then \(f\in \mathcal {S}_q\). The monodromy theorem implies that we always have \(q\ge 2\).

The Eremenko–Lyubich class \(\mathcal {B}\) consists of all transcendental meromorphic functions f for which \({\text {sing}}(f^{-1}){\setminus } \{\infty \}\) is a bounded subset of \(\mathbb {C}\). Thus

$$\begin{aligned} \mathcal {S}=\bigcup _{q=2}^\infty \mathcal {S}_q \subset \mathcal {B}. \end{aligned}$$

Both the Speiser and the Eremenko–Lyubich class play an important role in transcendental dynamics. The similarities and differences between these classes are addressed in a number of recent papers [2, 12,13,14, 21]. A survey of some of these as well as many other results concerning the dynamics of functions in \(\mathcal {S}\) and \(\mathcal {B}\) is given in [41].

Considerable attention has been paid to the Hausdorff dimension of Julia sets; see [30] and [45] for surveys. We will denote the Hausdorff dimension of a subset A of \(\mathbb {C}\) by \({\text {dim}}A\).

Baker [4] showed that if f is a transcendental entire function, then J(f) contains continua so that \({\text {dim}}J(f)\ge 1\). Stallard [44, Theorem 1.1] showed that for all \(d\in (1,2)\) there exists a transcendental entire function f with \({\text {dim}}J(f)=d\) while Bishop [15] constructed an example with \({\text {dim}}J(f)=1\). Stallard’s examples are actually in the Eremenko–Lyubich class \(\mathcal {B}\). Previously she had shown that \({\text {dim}}J(f)>1\) for entire \(f\in \mathcal {B}\). In particular, this is the case for entire functions in \(\mathcal {S}\).

Albrecht and Bishop [1] showed that given \(\delta >0\) there exists an entire function \(f\in \mathcal {S}\) such that \({\text {dim}}J(f)<1+\delta \). In fact, these were the first examples of entire functions in \(\mathcal {S}\) for which the Julia set has dimension strictly less than 2. In their examples the inverse has three finite singularities. Since every non-constant entire function has the asymptotic value \(\infty \) by Iversen’s theorem, and since we include \(\infty \) in \({\text {sing}}(f^{-1})\), their examples are in \(\mathcal {S}_4\).

For transcendental meromorphic functions f we have \({\text {dim}}J(f)>0\) by a result of Stallard [42]. On the other hand, she showed [43, Theorem 5] that for all \(d\in (0,1)\) there exists a transcendental meromorphic function f such that \({\text {dim}}J(f)=d\). Again her examples are in \(\mathcal {B}\). Together with her result covering the interval (1, 2) mentioned above, and since \(J(\exp z)=\mathbb {C}\) by a result of Misiurewicz [38] and \(J(\tan z)=\mathbb {R}\), it follows that for all \(d\in (0,2]\) there exists \(f\in \mathcal {B}\) such that \({\text {dim}}J(f)=d\).

We shall show that such examples also exist in the Speiser class \(\mathcal {S}\) and in fact in \(\mathcal {S}_3\).

Theorem 1.1

Let \(d\in (0,2]\). Then there exists a function \(f\in \mathcal {S}_3\) such that \({\text {dim}}J(f)=d\).

We note that if \(f\in \mathcal {S}_2\), then \({\text {dim}}J(f)>1/2\); see [8, Theorem 3.11] and [34, Remark 3.2 and Section 4.3]. On the other hand, Barański [8, Section 4] showed that for \(f_\lambda (z)=\lambda \tan z\in \mathcal {S}_2\) the function \(\lambda \mapsto {\text {dim}}J(f_\lambda )\) maps (0, 1] monotonically and continuously onto (1/2, 1]. It seems likely that for \(d\in (1,2]\) there also exists \(f\in \mathcal {S}_2\) such that \({\text {dim}}J(f)=d\). The main interest of Theorem 1.1 thus lies in the case where \(0<d\le 1/2\).

2 Preliminary results

The escaping set I(f) of a meromorphic function f is defined as the set of all \(z\in \mathbb {C}\) for which \(f^n(z) \rightarrow \infty \) as \(n\rightarrow \infty \). We always have \(I(f)\ne \emptyset \) and \(J(f)=\partial I(f)\). This was shown by Eremenko [22] for transcendental entire f and by Domínguez [17] for transcendental meromorphic f.

For \(f\in \mathcal {B}\) we have \(I(f)\subset J(f)\). This is due to Eremenko and Lyubich [24, Theorem 1] for entire f and to Rippon and Stallard [40, Theorem A] for meromorphic f.

Theorem 1.1 complements the result of [2] where it was shown that for all \(d\in [0,2]\) there exists a meromorphic function \(f\in \mathcal {S}\) such that \({\text {dim}}I(f)=d\).

An important concept in the theory of meromorphic functions is the order; see, e.g., [28]. The following result was proved in [11, Theorem 1.1]. Here and in the following \(\mathbb {N}=\{1,2,3,\dots \}\).

Lemma 2.1

Let \(f\in \mathcal {B}\) be of finite order \(\rho \). Suppose that \(\infty \) is not an asymptotic value of f and that there exists \(M\in \mathbb {N}\) such that all but finitely many poles of f have multiplicity at most M. Then

$$\begin{aligned} {\text {dim}}I(f)\le \frac{2M\rho }{2+M\rho }. \end{aligned}$$
(2.1)

We will not actually use this lemma. Instead we will use the following closely related result, which can be proved by the same method.

Lemma 2.2

Let f be as in the Lemma 2.1. Suppose, furthermore, that \(f(0)\ne 0\). For \(\lambda \in \mathbb {C}{\setminus }\{0\}\) define \(f_\lambda (z)=f(\lambda z)\). Then

$$\begin{aligned} \limsup _{\lambda \rightarrow 0} {\text {dim}}J(f_\lambda ) \le \frac{2M\rho }{2+M\rho }. \end{aligned}$$

Proof

It is easy to see that \(f_\lambda \) also has order \(\rho \), for all \(\lambda \in \mathbb {C}{\setminus }\{0\}\). We proceed as in [11]. Let \((a_j)\) be the sequence of poles of f and let \(m_j\) be the multiplicity of \(a_j\). Let \(b_j\in \mathbb {C}{\setminus }\{0\}\) be such that

$$\begin{aligned} f(z)\sim \left( \frac{b_j}{z-a_j}\right) ^{m_j} \quad \text {as}\ z\rightarrow a_j . \end{aligned}$$

Let

$$\begin{aligned} t> \frac{2M \rho }{2+ M \rho }. \end{aligned}$$

By [11, Lemma 3.1] we have

$$\begin{aligned} \sum _{j=1}^\infty \left( \frac{|b_j|}{|a_j|^{1+1/M}}\right) ^t < \infty . \end{aligned}$$
(2.2)

Let \(a_j^\lambda \) and \(b_j^\lambda \) be the corresponding values for \(f_\lambda \). Then \(a_j^\lambda =a_j/\lambda \) and \(b_j^\lambda =b_j/\lambda \) so that

$$\begin{aligned} \sum _{j=1}^\infty \left( \frac{|b_j^\lambda |}{|a_j^\lambda |^{1+1/M}}\right) ^t =|\lambda |^{t/M}\sum _{j=1}^\infty \left( \frac{|b_j|}{|a_j|^{1+1/M}}\right) ^t . \end{aligned}$$
(2.3)

As in [11] we choose \(R_0>|f(0)|\) such that \({\text {sing}}(f_\lambda ^{-1})={\text {sing}}(f^{-1})\subset D(0,R_0)\) and choose \(R\ge (16R_0)^M\). Here and in the following D(ar) denotes the open disk of radius r around a point \(a\in \mathbb {C}\). Choosing \(\lambda \) small we can achieve that \(f_\lambda (D(0,3R))\subset D(0,R)\) so that \(D(0,3R)\subset F(f_\lambda )\).

We put \(B(R):=\{z:|z|>R\}\cup \{\infty \}\) and, as in [11, pp. 5376f.], consider the collection \(E_l\) of all components V of \(f_\lambda ^{-l}(B(R))\) for which \(f_\lambda ^k(V)\subset B(3R)\) for \(0\le k\le l-1\). Then \(E_l\) is a cover of \(J(f_\lambda )\) and we have

$$\begin{aligned} \sum _{V\in E_l}\!\left( {\text {diam}}_{\chi }(V)\right) ^t = \frac{1}{M}\!\left( \frac{32}{(2R)^{1/M}24}\right) ^{\! t} \left( M (2^{1/M} 24)^{t} \sum _{j=1}^\infty \!\left( \frac{|b_j^\lambda |}{|a_j^\lambda |^{1+1/M}}\right) ^{\! t} \right) ^{\! l}. \end{aligned}$$

Here \({\text {diam}}_{\chi }(V)\) denotes the spherical diameter of V. Together with (2.2) and (2.3) the last estimate yields that if \(\lambda \) is sufficiently small, then

$$\begin{aligned} \sum _{V\in E_l}\left( {\text {diam}}_{\chi }(V)\right) ^t\rightarrow 0 \quad \text {as}\ l\rightarrow \infty . \end{aligned}$$

Hence \({\text {dim}}J(f_\lambda )\le t\) for small \(\lambda \). \(\square \)

It follows from a recent result of Mayer and Urbański [35] that Lemma 2.2 and (2.1) can be sharpened to

$$\begin{aligned} \lim _{\lambda \rightarrow 0} {\text {dim}}J(f_\lambda )={\text {dim}}I(f). \end{aligned}$$

In fact, their result says that \({\text {dim}}I(f)\) is the infimum of the set of all \(t>0\) for which (2.2) holds.

For entire functions, the following result can be found in [24, Section 3] and [21, Proposition 2.3].

Lemma 2.3

Let \(f,g\in \mathcal {S}_3\) and suppose that there exist homeomorphisms \(\psi :\widehat{{\mathbb {C}}}\rightarrow \widehat{{\mathbb {C}}}\) and \(\phi :\mathbb {C}\rightarrow \mathbb {C}\) such that \(\psi \circ f=g\circ \phi \). Then there exist a fractional linear transformation \(\alpha :\widehat{{\mathbb {C}}}\rightarrow \widehat{{\mathbb {C}}}\) and an affine map \(\beta :\mathbb {C}\rightarrow \mathbb {C}\) such that \(\alpha \circ f=g\circ \beta \).

Proof

Since \(f\in \mathcal {S}_3\) it follows from [21, Observation 1.10] that there exists a fractional linear transformation \(\alpha :\widehat{{\mathbb {C}}}\rightarrow \widehat{{\mathbb {C}}}\) which is isotopic to \(\psi \) relative to \({\text {sing}}(f^{-1})\). We follow the argument in the proof of [21, Proposition 2.3(a)]. Let \((\psi _t)_{t\in [0,1]}\) be the isotopy between \(\psi _0=\psi \) and \(\psi _1=\alpha \). By the isotopy lifting property, there exists a unique isotopy \((\phi _t)_{t\in [0,1]}\) in

$$\begin{aligned} U:=f^{-1}\!\left( \widehat{{\mathbb {C}}}{\setminus }{\text {sing}}(f^{-1})\right) \end{aligned}$$

such that \(\phi _0=\phi \) and \(\psi _t\circ f=g\circ \phi _t\) in U for all \(t\in [0,1]\). It remains to show that \(\phi _t\) extends continuously to the preimages of the singular values of f and coincides with \(\phi \) there for all t.

Let \(z_0\in \mathbb {C}\) be a preimage of \(v_0\in {\text {sing}}(f^{-1})\). We have to show that \(\phi _t(z)\rightarrow \phi (z_0)\) as \(z\rightarrow z_0\). We take a small neighborhood D of \(z_0\) such that \(f:D\rightarrow f(D)\) is a proper map, with no critical points except possibly at \(z_0\). We may also assume that \(f(z)\ne v_0\) for \(z\in {\overline{D}}{\setminus } \{z_0\}\). It then suffices to show that \(\phi _t(z)\in \phi (D)\) if z is sufficiently close to \(z_0\). By the continuity of f and the properties of an isotopy, we have

$$\begin{aligned} \psi _t(f(z))\rightarrow \psi _t(f(z_0))=\psi (v_0)\quad \text {as}\ z\rightarrow z_0, \end{aligned}$$

uniformly in \(t\in [0,1]\). It follows that

$$\begin{aligned} g(\phi _t(z))=\psi _t(f(z))\in \psi (f(D))=g(\phi (D)) \end{aligned}$$

if z is sufficiently close to \(z_0\). Thus \(\phi _t(z)\) can be obtained by analytic continuation of \(g^{-1}\) along the curve \(t\mapsto \psi _t(f(z))\). It follows that \(\phi _t(z)\in \phi (D)\) if z is sufficiently close to \(z_0\).

With \(\beta :=\phi _1\) we have \(\alpha \circ f=g\circ \beta \). If z is not a critical point of f, then \(\beta \) is of the form \(g^{-1}\circ \alpha \circ f\) near z for some branch of the inverse of g. Hence \(\beta \) is holomorphic. Since \(\beta :\mathbb {C}\rightarrow \mathbb {C}\) is a homeomorphism, this implies that \(\beta \) is affine. \(\square \)

The following result is due to Kotus and Urbański [29].

Lemma 2.4

Let f be an elliptic function. Let q be the maximal multiplicity of the poles of f. Then

$$\begin{aligned} {\text {dim}}J(f)> \frac{2q}{q+1}. \end{aligned}$$

We shall need a variation of this result.

Lemma 2.5

Let f be as in Lemma 2.4. If \((f_n)\) is a sequence of meromorphic functions which converges locally uniformly to f with respect to the spherical metric, then

$$\begin{aligned} {\text {dim}}J(f_n)> \frac{2q}{q+1} \end{aligned}$$

for all large n.

Lemma 2.5 can be deduced from the work of Kotus and Urbański; see Remark 2.1 below. But for the convenience of the reader we will include a proof of Lemma 2.5, thereby reproving Lemma 2.4. Here we will use the following result [25, Proposition 9.7].

Lemma 2.6

Let \(S_1,\dots ,S_m\) be contractions on a closed subset K of \(\mathbb {R}^d\) such that there exists \(b_1,\dots ,b_m\in (0,1)\) with

$$\begin{aligned} b_k|u-v|\le |S_k(u)-S_k(v)| \quad \text {for } u,v\in K \end{aligned}$$
(2.4)

and \(1\le k\le m\). Suppose that \(K_0\) is a non-empty compact subset of K with

$$\begin{aligned} K_0=\bigcup _{k=1}^m S_k(K_0) \end{aligned}$$

and \(S_j(K_0)\cap S_k(K_0)=\emptyset \) for \(j\ne k\). Let \(t>0\) with

$$\begin{aligned} \sum _{k=1}^m b_k^t=1. \end{aligned}$$

Then \({\text {dim}}K_0\ge t\).

Proof of Lemmas 2.4 and 2.5

Let \((a_j)\) be the sequence of poles of multiplicity q. For sufficiently large R there is a neighborhood \(U_j\) of \(a_j\) such that

$$\begin{aligned} f:U_j{\setminus }\{a_j\}\rightarrow \{z\in \mathbb {C}:|z|>R\} \end{aligned}$$

is a covering map of degree q. There exists \(r_1>0\) such that \(D(a_j,r_1)\subset U_j\) for all j. Let \(0<r_0<r_1\). Then there exists \(M\in \mathbb {N}\) such that

$$\begin{aligned} \overline{D(a_k,r_0)}\subset \overline{U_k}\subset f(D(a_j,r_0)) \end{aligned}$$

for all \(j,k\ge M\). Thus for \(j,k\ge M\) there exists \(V_{j,k}\subset D(a_j,r_0)\) such that \(f:V_{j,k}\rightarrow D(a_k,r_0)\) is biholomorphic.

Let \(W_k:=f^{-1}(V_{k,M})\cap D(a_M,r_0)\). Then

$$\begin{aligned} f^2:W_{k}\rightarrow D(a_M,r_0) \end{aligned}$$

is biholomorphic. Moreover, \(f^2\) extends to a bijective map from \(\overline{W_k}\) to \(K:=\overline{D(a_M,r_0)}\). Let \(S_k:K\rightarrow \overline{W_{k}}\) be the inverse function of \(f^2:\overline{W_{k}}\rightarrow K\). Then \(S_k\) extends to an injective map \(S_k:D(a_M,r_1)\rightarrow \mathbb {C}\). Choosing \(r_0\le (2-\sqrt{3})r_1\) we conclude that \(W_k=S_k(D(a_M,r_0))\) is convex; see [19, Theorem 2.13].

In order to apply Lemma 2.6 we note that since f has a pole of multiplicity q at \(a_j\), there exists \(c_1>0\) such that

$$\begin{aligned} |f'(z)|\le c_1 |f(z)|^{(q+1)/q} \quad \text {for}\ z\in D(a_j,r_0). \end{aligned}$$

This implies that there exists \(c_2>0\) such that

$$\begin{aligned} |f'(z)|\le c_2 |a_k|^{(q+1)/q} \quad \text {if}\ z\in D(a_j,r_0) \ \text {and}\ f(z)\in D(a_k,r_0). \end{aligned}$$

With \(c_3:= c_2^2 |a_M|^{(q+1)/q}\) this yields that

$$\begin{aligned} |(f^2)'(z)|=|f'(f(z))f'(z)|\le c_3 |a_k|^{(q+1)/q} \quad \text {for}\ z\in W_k. \end{aligned}$$

Since \(W_k\) is convex this yields that

$$\begin{aligned} |u-v|=|f^2(S_k(u))-f^2(S_k(v))|\le c_3 |a_k|^{(q+1)/q}|S_k(u)-S_k(v)| \quad \text {for}\ u,v\in K. \end{aligned}$$

It follows that (2.4) holds with

$$\begin{aligned} b_k:=\frac{1}{c_3 |a_k|^{(q+1)/q}} . \end{aligned}$$

Let now \(N\ge M\) and define \(t>0\) by

$$\begin{aligned} \sum _{k=M}^N b_k^t=1. \end{aligned}$$

It follows from Lemma 2.6 that the limit set of the iterated function system \(\{S_k:M\le k\le N\}\) has Hausdorff dimension at least t. It is easily seen that this limit set is contained in the Julia set. Thus \({\text {dim}}J(f)\ge t\). Since

$$\begin{aligned} \sum _{k=M}^\infty \frac{1}{|a_k|^2}=\infty \end{aligned}$$

we have \(t>2q/(q+1)\) if N is large enough. This yields Lemma 2.4.

To prove Lemma 2.5 we note that for large n there are domains \(W_k^n\) and \(U_M^n\) close to \(W_k\) and \(U_M\) such that

$$\begin{aligned} f_n^2:W_k^n\rightarrow U_M^n \end{aligned}$$

is biholomorphic, and the inverse function \(S_k^n\) satisfies (2.4) with a constant \(b_k^n\) instead of \(b_k\), with \(b_k^n\rightarrow b_k\) as \(n\rightarrow \infty \). The conclusion then follows again from Lemma 2.6. \(\square \)

Remark 2.1

The argument of Kotus and Urbański [29] is similar to the one used above, but they use an infinite iterated function system and apply results of Mauldin and Urbański [33] concerning such systems. However, it would suffice to consider a sufficiently large finite subsystem. This would yield Lemma 2.5 as above.

The proof actually yields that the hyperbolic dimension of f and \(f_n\), and not only the Hausdorff dimension of their Julia sets, have the given lower bound; see [9, 39] for a discussion of the hyperbolic dimension of meromorphic functions.

Lemma 2.7

Let \(f\in \mathcal {S}\). Suppose that f has an attracting fixed point whose immediate attracting basin contains all finite singularities of \(f^{-1}\). Suppose also that \(\infty \) is not an asymptotic value of f and that there exists a uniform bound on the multiplicities of the poles of f. Then J(f) is totally disconnected and F(f) is connected.

There are several closely related results in the literature, see [5, Theorem G], [18, Theorem A and Corollary  3.2] and [47, Theorem 2.7]. However, none of these results seems to apply exactly to the situation we have.

Zheng [47, Theorem 2.7] showed that the conclusion of Lemma 2.7 holds if \(\infty \notin {\text {sing}}(f^{-1})\). Thus his result would apply if the poles are assumed to be simple, while we allow multiple poles. Hawkins and Koss ([27, Theorem 3.12]; see also [26, Theorem 3.2]) do not require that the poles are simple, but they restrict to elliptic functions.

Our proof of Lemma 2.7 will use some ideas from the papers mentioned. A difference to the methods employed there, however, is that we will use the following consequence of the Grötzsch inequality; see [16, Section 5.2] and [36, Corollary A.7]. Here and in the following \({\text {mod}}(A)\) denotes the modulus of an annulus A.

Lemma 2.8

Let \((G_k)\) be a sequence of simply connected domains in \(\mathbb {C}\) such that \(A_k:=G_k{\setminus } \overline{G_{k+1}}\) is an annulus for all \(k\in \mathbb {N}\). Suppose that

$$\begin{aligned} \sum _{k=1}^\infty {\text {mod}}(A_k) =\infty . \end{aligned}$$

Then \(\bigcap _{k=1}^\infty G_k\) consists of a single point.

Proof of Lemma 2.7

Let \(\xi \) be the attracting fixed point whose attracting basin W contains all finite singularities of \(f^{-1}\). Then the postsingular set

$$\begin{aligned} P(f):=\overline{\bigcup _{n=0}^\infty f^n\!\left( {\text {sing}}(f^{-1}){\setminus }\{\infty \}\right) } \end{aligned}$$

is a compact subset of W. It is not difficult to see that there exist Jordan domains U and V such that

$$\begin{aligned} \{\xi \}\cup P(f)\subset U\subset {\overline{U}}\subset V \subset {\overline{V}} \subset W. \end{aligned}$$

Then \(A:=V{\setminus }{\overline{U}}\) is an annulus. Clearly, \(A\subset W\).

Let \((a_j)\) be the sequence of poles of f and let \(m_j\) denote the multiplicity of \(a_j\). By hypothesis, there exists \(M\in \mathbb {N}\) such that \(m_j\le M\) for all j. Let \(Y_j\) be the component of the preimage of \(\widehat{{\mathbb {C}}}{\setminus } {\overline{U}}\) that contains \(a_j\). Then \(f:Y_j{\setminus }\{a_j\}\rightarrow \mathbb {C}{\setminus } {\overline{U}}\) is a covering of degree \(m_j\). Putting \(B_j:=f^{-1}(A)\cap Y_j\) we find that \(B_j\) is an annulus with

$$\begin{aligned} {\text {mod}}(B_j)=\frac{1}{m_j}{\text {mod}}(A)\ge \frac{1}{M}{\text {mod}}(A). \end{aligned}$$

To prove that J(f) is totally disconnected, let \(z\in J(f)\). We want to show that the component of J(f) containing z consists of the point z only. First we note that for all \(n\in \mathbb {N}\) there exists \(j(n)\in \mathbb {N}\) such that \(f^n(z)\in Y_{j(n)}\). Let \(X_n\) be the component of \(f^{-n}(Y_{j(n)})\) containing z. Note that \(Y_{j(n)}\) and \(X_n\) are Jordan domains. Since \(\partial U\) and hence \(\partial X_n\) are contained in F(f), the component of J(f) containing z is contained in the intersection of the \(X_n\). It thus suffices to prove that this intersection consists of only one point.

In order to do so we note that since \(P(f)\subset U\), the map \(f^{n}:X_n\rightarrow Y_{j(n)}\) is biholomorphic. This implies that

$$\begin{aligned} C_n:=f^{-n}(B_{j(n)})\cap X_n= f^{-n-1}(A)\cap X_n \end{aligned}$$
(2.5)

is an annulus satisfying

$$\begin{aligned} {\text {mod}}(C_n)={\text {mod}}(B_{j(n)})\ge \frac{1}{M}{\text {mod}}(A). \end{aligned}$$

Moreover, \(X_n\) is equal to the union of \(C_n\) and the component of \(\mathbb {C}{\setminus } C_n\) that contains z.

Since the closure of A is a compact subset of the attracting basin W of \(\xi \), there exists \(p\in \mathbb {N}\) such that \(f^p(A)\subset U\). In particular,

$$\begin{aligned} f^p(A)\cap A=\emptyset . \end{aligned}$$
(2.6)

This implies that \(C_{n+p}\cap C_n=\emptyset \) for all \(n\in \mathbb {N}\). In fact, if \(w\in C_{n+p}\cap C_n\), then \(f^{n+p+1}(w)\in f^p(A)\cap A\) by (2.5), contradicting (2.6). It follows that

$$\begin{aligned} X_{n+p}\subset X_n{\setminus } C_n, \end{aligned}$$

and hence that

$$\begin{aligned} {\text {mod}}\!\left( X_n{\setminus }\overline{X_{n+p}}\right) \ge {\text {mod}}(C_n)\ge \frac{1}{M}{\text {mod}}(A). \end{aligned}$$

As already mentioned above it now follows from Lemma 2.8, applied with \(G_k=X_{1+pk}\), that J(f) is totally disconnected. Of course, this yields that F(f) is connected.

3 Proof of Theorem 1.1

Let G be the conformal map from the triangle with vertices 0, \(\pi /2\) and \(i\pi /2\) onto the lower half-plane such that

$$\begin{aligned} G(\pi /2)=0, \quad G(0)=1 \quad \text {and}\quad G(i\pi /2)=\infty . \end{aligned}$$

Extending this to the whole plane by reflections we obtain an elliptic function G. The critical values of G are 0, 1 and \(\infty \) so that \(G\in \mathcal {S}_3\). The zeros and poles of G have multiplicity 4 while the 1-points have multiplicity 2. We also note that \(G(\mathbb {R})=[0,1]\).

We may express G in terms of the Weierstrass \(\wp \)-function with periods \(\pi \) and \(\pi i\). The critical values of \(\wp \) are \(e_1\), \(e_2\), \(e_3\) and \(\infty \), with

$$\begin{aligned} e_2=\wp ((\pi +i \pi )/2)=0 \quad \text {and}\quad e_1=\wp (\pi /2)=-e_3=-\wp (i\pi /2) . \end{aligned}$$

It follows from this that

$$\begin{aligned} G(z)=\left( \frac{\wp (z+i\pi /2)}{e_1}\right) ^2 . \end{aligned}$$

First we construct an example of a function \(f\in \mathcal {S}_3\) where the Julia set is the whole sphere and thus has Hausdorff dimension 2. To this end we consider the function

$$\begin{aligned} f(z):=i G(\pi z/2). \end{aligned}$$

Then \(f\in \mathcal {S}_3\). The critical values of f are 0, i and \(\infty \) and we have \(f(0)=i\) and \(f(i)=\infty \). To prove that \(J(f)=\widehat{{\mathbb {C}}}\) we note that f has no wandering domains [6] and no Baker domains [40, Corollary to Theorem A]. All other types of components of F(f) are related to the singularities of \(f^{-1}\); see [10, Theorem 7] for the exact statement. Since the points in \({\text {sing}}(f^{-1})\cap \mathbb {C}=\{0,i\}\) are mapped to \(\infty \) by f or \(f^2\) this yields that \(F(f)=\emptyset \) and hence \(J(f)=\widehat{{\mathbb {C}}}\) as claimed.

To construct functions in \(\mathcal {S}_3\) for which the Julia set has Hausdorff dimension \(d\in (0,2)\), we consider, for \(p\in \mathbb {N}\) and small \(\eta \in (0,\pi /2)\), the function H defined by \(H(z):=\eta G(z)^p\). The critical values of H are 0, \(\eta \) and \(\infty \). Thus \(H\in \mathcal {S}_3\). We have \(H(0)=\eta \) and \(H(\pi /2)=0\). Also, H decreases in the interval \((0,\pi /2)\). Choosing \(\eta \) sufficiently small we can achieve that H has an attracting fixed point \(\xi \in (0,\pi /2)\) whose attracting basin contains \([0,\eta ]\) and thus, since \(H(\mathbb {R})=[0,\eta ]\), also contains \(\mathbb {R}\). Lemma 2.7 implies that this attracting basin coincides with F(H) and that J(H) is totally disconnected.

Since \(H((0,\pi /2))=(0,\eta )\) we actually have \(\xi \in (0,\eta )\). Choosing \(\eta \) small we can also achieve that \(H''(z)\ne 0\) for \(0<|z|\le \eta \). In fact, \(H''(x)<0\) for \(0<x\le \eta \).

Let now m be a (large) odd integer. Then

$$\begin{aligned} h_m(z)=H(m\arcsin (z/m)) . \end{aligned}$$

defines a meromorphic function \(h_m\in \mathcal {S}_3\). Similar examples were already considered by Teichmüller [46, p. 734], and later by Bank and Kaufman [7, Section 5], Langley [31, Section 2] and Eremenko [23].

The elliptic function \(H(z)=h_m(m\sin (z/m))\) has order 2. A result of Edrei and Fuchs [20, Corollary  1.2] thus yields that \(h_m\) has order 0. In fact, as in the papers cited above we find that there exists a constant c such that the Nevanlinna characteristic satisfies \(T(r,h_m)\sim c( \log r)^2\) as \(r\rightarrow \infty \).

For large m the function \(h_m\) has an attracting fixed point \(\xi _m\), with \(\xi _m\rightarrow \xi \) as \(m\rightarrow \infty \), such that the attracting basin of \(\xi _m\) contains the interval \([0,\eta ]\) and hence \(\mathbb {R}\). Lemma 2.7 implies that this attracting basin is connected and coincides with \(F(h_m)\). Choosing m large we can also achieve that \(h_m\) decreases in the interval \([0,\eta ]\) and that \(h_m''(x)<0\) for \(0<x\le \eta \). This implies that \(h_m'\) decreases in the interval \([0,\eta ]\).

The poles of H and \(h_m\) have multiplicity 4p. Except for the zeros at \(\pm m\), which have multiplicity 2p, the zeros of \(h_m\) also have multiplicity 4p. The \(\eta \)-points on the real axis have multiplicity 2, but if \(p\ge 2\), then H and \(h_m\) also have simple \(\eta \)-points (corresponding to the points where G takes the p-th roots of unity).

As the poles of H have multiplicity 4p, Lemma 2.4 yields that

$$\begin{aligned} {\text {dim}}J(H) > \frac{8p}{4p+1} . \end{aligned}$$

Moreover, it follows from Lemma 2.5 that if m is sufficiently large, then

$$\begin{aligned} {\text {dim}}J(h_m)> \frac{8p}{4p+1}. \end{aligned}$$
(3.1)

We fix such a value of m and, for \(\lambda \in (0,1]\), we put \(f_\lambda (z):=h_m(\lambda z)\) so that \(f_1=h_m\).

Since \(h_m\) has order 0, Lemma 2.2 yields that

$$\begin{aligned} \lim _{\lambda \rightarrow 0}{\text {dim}}J(f_\lambda ) = 0. \end{aligned}$$
(3.2)

We will show that the function \(\lambda \mapsto {\text {dim}}J(f_\lambda )\) is continuous in the interval (0, 1]. Since \(f_1=h_m\) it then follows from (3.1) and (3.2) that for all \(d\in (0,8p/(4p+1)]\) there exists \(\lambda \in (0,1]\) such that \({\text {dim}}J(f_\lambda )=d\). Since p can be chosen arbitrarily large, this yields the conclusion.

It remains to prove that \(\lambda \mapsto {\text {dim}}J(f_\lambda )\) is continuous. Recalling that \(h_m\) and \(h_m'\) are decreasing in the interval \([0,\eta ]\) we can deduce that \(f_\lambda \) has an attracting fixed point \(\zeta _\lambda \in (0,\eta )\) and that the multiplier

$$\begin{aligned} m_\lambda :=f_\lambda '(\zeta _\lambda )=\lambda h_m'(\lambda \zeta _\lambda ) \end{aligned}$$

is a decreasing function of \(\lambda \) in the interval (0, 1]. Note here that \(\zeta _1=\xi _m\) and \(m_1=h_m'(\xi _m)<0\). As before it follows from Lemma 2.7 that the attracting basin of \(\zeta _\lambda \) is connected and coincides with the Fatou set of \(f_\lambda \).

Let \(\lambda \in (0,1]\). Kœnigs’ theorem [37, Theorem 8.2] yields that there exists a function g holomorphic and injective in some neighborhood U of \(\zeta _\lambda \) such that \(g(\zeta _\lambda )=0\), \(g'(\zeta _\lambda )=1\) and

$$\begin{aligned} g(f_\lambda (g^{-1}(z)))=m_\lambda z \end{aligned}$$

for all \(z\in g(U)\). For \(\kappa \in (0,1]\) we put

$$\begin{aligned} \gamma =\frac{\log |m_\kappa |}{\log |m_\lambda |} -1 \end{aligned}$$

and define \(h:\mathbb {C}\rightarrow \mathbb {C}\), \(h(z)= z|z|^\gamma \). Then

$$\begin{aligned} h(m_\lambda h^{-1}(z))=m_\kappa z. \end{aligned}$$

With \(\phi =h\circ g:U\rightarrow \mathbb {C}\) we then have

$$\begin{aligned} \phi (f_\lambda (\phi ^{-1}(z)))=m_\kappa z \end{aligned}$$
(3.3)

for \(z\in \phi (U)\). The maps h and \(\phi \) are K-quasiconformal with

$$\begin{aligned} K= \max \left\{ \frac{\log |m_\kappa |}{\log |m_\lambda |}, \frac{\log |m_\lambda |}{\log |m_\kappa |}\right\} . \end{aligned}$$
(3.4)

For a detailed account of quasiconformal mappings, we refer to [32]. The complex dilatation \(\mu (z):=\phi _{{\overline{z}}}(z)/\phi _z(z)\) satisfies

$$\begin{aligned} \mu (f_\lambda (z))= \mu (z) \frac{f_\lambda '(z)}{\;\overline{f_\lambda '(z)}\;} \end{aligned}$$
(3.5)

if \(z,f_\lambda (z)\in U\). We may use (3.5) to extend \(\mu \) to \(\mathbb {C}\). More precisely, we put \(\mu (z)=0\) if \(z\in J(f_\lambda )\) or if \(z\in F(f_\lambda )\) and \((f_\lambda ^n)'(z)=0\) for some \(n\in \mathbb {N}\). For the remaining points \(z\in F(f_\lambda )\) we define

$$\begin{aligned} \mu (z) = \mu (f_\lambda ^n(z)) \frac{\;\overline{(f_\lambda ^n)'(z)}\;}{(f_\lambda ^n)'(z)}, \end{aligned}$$
(3.6)

where n is chosen so large that \(f_\lambda ^n(z)\in U\). Using (3.5) it is easily seen that \(\mu \) is well-defined, i.e., the definition does not depend on the value of n chosen in (3.6). We find that (3.5) holds for all z.

Let \(\psi :\mathbb {C}\rightarrow \mathbb {C}\) be the solution of the Beltrami equation

$$\begin{aligned} \mu (z)=\frac{\psi _{{\overline{z}}}(z)}{\psi _z(z)}, \end{aligned}$$

normalized by \(\psi (0)=0\) and \(\psi (\eta )=\eta \). It follows from (3.5) that

$$\begin{aligned} k:=\psi \circ f_\lambda \circ \psi ^{-1} \end{aligned}$$
(3.7)

is meromorphic. Since \(f_\lambda \) is symmetric with respect to the real axis, the same applies to g, \(\phi \), \(\mu \), \(\psi \) and k. By definition, \(f_\lambda \) is even. In order to show that k is also even, we first note that since \(f_\lambda \) is even, it follows from (3.5) that \(\mu \) is even. This implies that \(\psi (z)\) and \(\psi (-z)\) have the same complex dilatation. Hence there exists an affine map L such that \(\psi (z)=L(\psi (-z))\). Since \(\psi (0)=0\) we have \(L(0)=0\) so that L has the form \(L(z)=az\) for some \(a\in \mathbb {C}{\setminus }\{0\}\). We also see that L is real on the real axis so that \(a\in \mathbb {R}{\setminus }\{0\}\). Since \(a\psi (i)=L(\psi (i))=\psi (-i)=\overline{\psi (i)}\) we find that \(|a|=1\). As \(\psi \) is injective this implies that \(a=-1\) so that \(\psi \) is odd. Hence k is even.

Since the complex dilatations of \(\phi \) and \(\psi \) agree in U, we have \(\phi =\tau \circ \psi \) for some function \(\tau \) holomorphic and injective on \(\psi (U)\). Together with (3.3) this implies that

$$\begin{aligned} k(z)= \psi (f_\lambda (\psi ^{-1}(z)))=\tau ^{-1}(m_\kappa \tau (z)) . \end{aligned}$$

Thus \(\tau ^{-1}(0)=\psi (\zeta _\lambda )\) is a fixed point of k of multiplier \(m_\kappa \).

Another function with a fixed point of multiplier \(m_\kappa \) is \(f_\kappa \). We will show that \(k=f_\kappa \). In order to do so we note that both k and \(f_\kappa \) are in \(\mathcal {S}_3\), with critical values 0, \(\eta \) and \(\infty \). It follows from Lemma 2.3 and (3.7) that there exist a fractional linear transformation \(\alpha \) and an affine map \(\beta \) such that \(\alpha \circ k=f_\lambda \circ \beta \). Since all poles of k and \(f_\kappa \) have multiplicity 4p, all but two zeros of both functions have multiplicity 4p, and all \(\eta \)-points of both functions have multiplicity 2 or 1, we find that \(\alpha (0)=0\), \(\alpha (\eta )=\eta \) and \(\alpha (\infty )=\infty \). Hence \(\alpha (z)\equiv z\) so that \(k=f_\lambda \circ \beta \).

As \(\beta \) is affine we have \(-\beta (-z)=\beta (z)-2\beta (0)\). Noting that k and \(f_\lambda \) are even we deduce that

$$\begin{aligned} f_\lambda (\beta (z)-2\beta (0))= f_\lambda (-\beta (-z)) = f_\lambda (\beta (-z))=k(-z)=k(z)=f_\lambda (\beta (z)). \end{aligned}$$

Since periodic functions have order at least 1 while \(f_\lambda \) has order 0, this implies that \(\beta (0)=0\) so that \(\beta \) has the form \(\beta (z)=cz\) for some constant c. Thus \(k(z)=f_\lambda (cz)=h_m(\lambda cz)\). As k has an attracting fixed point of multiplier \(m_\kappa \) this yields that \(c=\kappa /\lambda \) so that \(k(z)=h_m(\kappa z)=f_\kappa (z)\).

Inserting \(k=f_\kappa \) in (3.7), we obtain

$$\begin{aligned} f_\kappa =\psi \circ f_\lambda \circ \psi ^{-1}. \end{aligned}$$

This implies that

$$\begin{aligned} J(f_\kappa )= \psi (J(f_\lambda )). \end{aligned}$$

As \(\psi \) is K-quasiconformal, and thus Hölder continuous with exponent 1/K, we deduce that

$$\begin{aligned} \frac{1}{K} {\text {dim}}J(f_\lambda )\le {\text {dim}}J(f_\kappa )\le K{\text {dim}}J(f_\lambda ). \end{aligned}$$
(3.8)

It follows from (3.4) that \(K\rightarrow 1\) as \(\kappa \rightarrow \lambda \). Thus we deduce from (3.8) that \({\text {dim}}J(f_\kappa )\rightarrow {\text {dim}}J(f_\lambda )\) as \(\kappa \rightarrow \lambda \). Hence \(\lambda \mapsto {\text {dim}}J(f_\lambda )\) is continuous. \(\square \)

Remark 3.1

A celebrated result of Astala [3, Corollary 1.3] says that (3.8) can be improved to

$$\begin{aligned} \frac{1}{K} \left( \frac{1}{{\text {dim}}J(f_\lambda )}-\frac{1}{2}\right) \le \frac{1}{{\text {dim}}J(f_\kappa )}-\frac{1}{2} \le K \left( \frac{1}{{\text {dim}}J(f_\lambda )}-\frac{1}{2}\right) . \end{aligned}$$

For our purposes, however, the weaker and simpler estimate (3.8) suffices.