1 Introduction

Non-trivial rational families \((f_t)\) normally contain specific maps of different character with most interesting and unexpected Julia sets:

  • totally disconnected Julia sets (Cantor sets) occur in any family \(z\mapsto z^d+t\);

  • Julia sets consisting of uncountably many (a Cantor set of) quasi-circles occur in the McMullen family \(z\mapsto z^m+t/z^n\), which was introduced in [8]. The number of papers on various features of this family is legion; [3] marks the preliminary end of a long list of papers.

  • Julia sets that are Sierpiński curves (Milnor and Tan Lei  [12] were the first to construct examples with this property) occur again in the McMullen family [16], the Morosawa-Pilgrim family \(z\mapsto t\left( 1+\frac{(4/27)z^3}{1-z}\right) \) [4, 17], and the family \(t\mapsto -\frac{t}{4}\frac{(z^2-2)^2}{z^2-1}\) [7].

  • In any reasonable family, copies of the Mandelbrot sets of the families \(z\mapsto z^d+t\) are dense in the bifurcation locus—the Mandelbrot set is universal [10].

Each of these families has just one free critical point (or several free critical points which have the same dynamical behaviour, this happens, for example, in the McMullen family; the quasi-conjugated family \(F_t(z)=z^m(1+t/z)^d\) has just one free critical point). In contrast to this the rational maps

$$\begin{aligned} f_t(z)=tz^m\left( \frac{1-z}{1+z}\right) ^n\quad (m\ge 2,~n\in \mathbb N,~d=m+n,~t\ne 0) \end{aligned}$$
(1)

in the family under consideration have two free critical points. In this paper we will give a complete description of the parameter plane and the various dynamical planes. For basic notations and results the reader is referred to the texts [1, 2, 9, 11, 15].

2 Notation

The rational map (1) has

  • two super-attracting fixed points 0 and \(\infty \) with corresponding basins \(\mathscr {A}_t\) and \(\mathscr {B}_t\), respectively. Then \(\mathscr {A}_t\), say, either is completely invariant or else has a single pre-image \(\mathscr {A}_t^*\) that is mapped in a (n : 1)-manner onto \(\mathscr {A}_t\), which will be written as

    $$\begin{aligned} \mathscr {A}_t^*\buildrel n:1 \over \longrightarrow \mathscr {A}_t; \end{aligned}$$
  • two free critical points

    $$\begin{aligned} \qquad \alpha =-\frac{n}{m}+\sqrt{1+\left( \frac{n}{m}\right) ^2}\quad \mathrm{and}\quad \beta =-\frac{n}{m}-\sqrt{1+\left( \frac{n}{m}\right) ^2} \end{aligned}$$

    and critical values

    $$\begin{aligned} v^\alpha _t=f_t(\alpha )=tv^\alpha _1\quad \mathrm{and}\quad v^\beta _t=f_t(\beta )=t v^\beta _1; \end{aligned}$$
  • two escape loci \(\mathbf {\Omega }^\alpha \) and \(\mathbf {\Omega }^\beta \), with \(t\in \mathbf {\Omega }^\alpha \) and \(t\in \mathbf {\Omega }^\beta \) if and only if \(f^k_t(\alpha )\rightarrow 0\) and \(f^k_t(\beta )\rightarrow \infty \), respectively, as \(k\rightarrow \infty \);

  • two residual sets \(\mathbf {\Omega }_\mathrm{res}^\alpha \) and \(\mathbf {\Omega }_\mathrm{res}^\beta \), with \(t\in \mathbf {\Omega }_\mathrm{res}^\alpha \) and \(t\in \mathbf {\Omega }_\mathrm{res}^\beta \) if and only if \(v^\beta _t\in \mathscr {A}_t\) and \(v^\alpha _t\in \mathscr {B}_t\), respectively.

The notation of the residual sets indicates that \(\mathbf {\Omega }_\mathrm{res}^\alpha \) is related to \(\mathbf {\Omega }^\alpha \) rather than \(\mathbf {\Omega }^\beta \). The open sets \(\mathbf {\Omega }^\alpha \) and \(\mathbf {\Omega }^\beta \) are in a natural way sub-divided into

  • \(\mathbf {\Omega }_0^\alpha \) resp. \(\mathbf {\Omega }_0^\beta \): \(v^\alpha _t\in \mathscr {A}_t\) resp. \(v^\beta _t\in \mathscr {B}_t\), and

  • \(\mathbf {\Omega }_k^\alpha \) resp. \(\mathbf {\Omega }_k^\beta \): \(f^k_t\left( v^\alpha _t\right) \in \mathscr {A}_t\), but \(f^{k-1}_t\left( v^\alpha _t\right) \notin \mathscr {A}_t\) resp. \(f^k_t\left( v^\beta _t\right) \in \mathscr {B}_t\), but \(f^{k-1}_t\left( v^\beta _t\right) \notin \mathscr {B}_t\) (\(k\ge 1)\).

Hitherto, \(f_t\) is hyperbolic and the Fatou set of \(f_t\) consists of the basins \(\mathscr {A}_t\) and \(\mathscr {B}_t\), and their pre-images, if any. However, there may and will be also other hyperbolic components. By \(\mathbf {W}_k^\alpha \) and \(\mathbf {W}^\beta _k\) we denote the open sets such that \(\alpha \) and \(\beta \) belongs to some (super-)attracting cycle of Fatou domains \(U_1,\ldots ,U_k\), respectively, not containing 0 and \(\infty \).

The bifurcation locus \(\mathbf {B}\) of the family \((f_t)_{0<|t|<\infty }\) is the set of t such that the Julia set \(\mathscr {J}_t\) does not move continuously over any neighbourhood of t, see McMullen [9]. In order that \(t\in \mathbf {B}\) it is necessary and sufficient that at least one of the free critical points is active. Thus \(\mathbf {B}=\mathbf {B}^\alpha \cup \mathbf {B}^\beta \), where \(t\in \mathbf {B}^\alpha \) resp. \(t\in \mathbf {B}^\beta \) means that \(\alpha \) resp. \(\beta \) is active. It is a priori not excluded that \(\mathbf {B}^\alpha \) and \(\mathbf {B}^\beta \) overlap. Although there is just one parameter plane, each point of this plane carries at least two pieces of information, so one could also speak of the \(v^\alpha _t\)- and \(v^\beta _t\)-plane.

We also set

$$\begin{aligned} Q_0(t)=v^\alpha _t=tv^\alpha _1\quad \mathrm{and}\quad Q_k(t)=f_t^k(v^\alpha _t)=f_t(Q_{k-1}(t))\quad (k\ge 1) \end{aligned}$$

and note that \(Q_k\) is a rational function of degree \(1+d+\cdots +d^k=\frac{d^{k+1}-1}{d-1}\) with a zero of order \(\frac{m^{k+1}-1}{m-1}\) at the origin.

From

$$\begin{aligned} -1/f_t(-1/z)=f_{(-1)^{d+1}/t}(z)\quad (d=m+n) \end{aligned}$$

it follows that \(f_t\) is conjugated to \(f_{1/t}\) if d is odd, and to \(f_{-1/t}\) if d is even, hence \(t\in \mathbf {\Omega }^\alpha \) if and only if \((-1)^{d+1}/t\in \mathbf {\Omega }^\beta \), and this is also true for \(\mathbf {\Omega }^\alpha _{k}\) and \(\mathbf {\Omega }^\beta _{k}\), \(\mathbf {\Omega }^\alpha _\mathrm{res}\) and \(\mathbf {\Omega }^\beta _\mathrm{res}\), \(\mathbf {W}_k^\alpha \) and \(\mathbf {W}_k^\beta \), and \(\mathbf {B}^\alpha \) and \(\mathbf {B}^\beta \). This also indicates that the circle \(|t|=1\) plays a distinguished role with strong impact on what follows.

Lemma 1

For every \(m\ge 2\), \(n\ge 1\) there exists some \(r>0\), such that for \(0<|t|\le 1\) the disc \(\triangle _{r|t|}:|z|<r|t|\) contains \(f_t(\overline{\triangle }_{r|t|}\cup [0,1])\), but does not contain \(v^\beta _t\).

Proof

We will first consider \(f_1\) and show that there exists some disc \(\triangle _{r}:|z|<r\) such that \(f_1(\overline{\triangle }_r\cup [0,1])\subset \triangle _r\) holds. This is easy to show if \(n<m\) for \(r=\frac{1}{3}\):

$$\begin{aligned} |f_1(z)|\le 3^{-m}2^n<\textstyle \frac{1}{3} \end{aligned}$$

holds if \(|z|\le \frac{1}{3}\) and \(m>n\ge 1\), and from

$$\begin{aligned} 0\le f_1(x)\le x^2\frac{1-x}{1+x}\le \textstyle \frac{1}{2}\left( 5\sqrt{5}-11\right) <\frac{1}{10}\quad (0\le x\le 1) \end{aligned}$$

the assertion follows.

We now consider the case \(n\ge m\). Then \(f_1(\overline{\triangle }_r)\subset \triangle _r\) holds as long as

$$\begin{aligned} g(r)=r^{m-1}\left( \frac{1+r}{1-r}\right) ^n<1, \end{aligned}$$

and \(f_1\) maps [0, 1] into \(\triangle _r\) provided

$$\begin{aligned} v^\alpha _1=\max _{0\le x\le 1}x^m\left( \frac{1-x}{1+x}\right) ^n<r. \end{aligned}$$

Since g is increasing this may be achieved if \(g(v^\alpha _1)<1\) holds. To prove this we note that \(\sqrt{1+\tau }-1=\frac{\tau }{2\sqrt{1+\theta \tau }}\) \((0<\theta <1,\) \(\tau =\frac{m^2}{n^2}\le 1)\) implies \(\frac{m}{2\sqrt{2}n}<\alpha <\frac{m}{2n}\), while from \(\log \frac{1-x}{1+x}< -2x\) (\(0<x<1\)) it follows that

$$\begin{aligned} v^\alpha _1<\left( \frac{m}{2n}\right) ^me^{-2\frac{m}{2\sqrt{2}}}=\left( \frac{m}{2e^{\frac{1}{\sqrt{2}}} n}\right) ^m<\left( \frac{m}{4n}\right) ^m=\mu ^m. \end{aligned}$$

Moreover, from

$$\begin{aligned} \log \frac{1+x}{1-x}=2x\left( 1+\frac{1}{3}x^2+\frac{1}{5}x^4+\cdots \right) \le 2x\left( 1+\frac{x^2}{3}\frac{1}{1-x^2}\right) \le \textstyle 2x\left( 1+\frac{1}{45}\right) , \end{aligned}$$

which holds for \(x=\left( \frac{m}{4n}\right) ^{m-1}\le \frac{1}{4}\), we obtain

$$\begin{aligned} \left( \frac{1+\mu ^m}{1-\mu ^m}\right) ^n=\left( \frac{1+\frac{m}{4}\frac{\mu ^{m-1}}{n}}{1-\frac{m}{4}\frac{\mu ^{m-1}}{n}}\right) ^n\le e^{\frac{23}{45}m\mu ^{m-1}}<\left( e^{(\frac{m}{4n})^{m-1}}\right) ^m. \end{aligned}$$

Thus \(g(v^\alpha _1)<1\) follows from \(\left( \frac{m}{4n}\right) ^{m-1}e^{(\frac{m}{4n})^{m-1}}\le \frac{1}{4}e^{\frac{1}{4}}<1\).

With this choice of \(r\in (0,1)\) it is clear that \(v^\beta _t\) belongs to \(\triangle _r\) if |t| is small. For individual \(0<|t|\le 1\), \(f_t(z)=tf_1(z)\) maps \(\overline{\triangle }_{r|t|}\cup [0,1]\) into \(\triangle _{r|t|}\), while \(v^\beta _t\notin \triangle _{r|t|}\) follows from \(|v^\beta _t|=|t|/v^\alpha _1>|t|>r|t|\). \(\Box \)

3 The escape loci

The purpose of Lemma 1 is twofold. First of all it shows that the critical points \(\alpha \) and \(\beta \) cannot be simultaneously active, and the bifurcation sets \(\mathbf {B}^\alpha \) and \(\mathbf {B}^\beta \) are separated by the unit circle \(|t|=1\). Secondly, the condition \(v^\beta _t\notin \triangle _{r|t|}\) (\(0<|t|\le 1)\) ensures that in an exhaustion \((D_\kappa )\) of \(\mathscr {A}_t\) starting with \(D_0=\triangle _{r|t|}\), \(D_\kappa \) is simply connected as long as \(\beta \notin D_\kappa \), and \(f_t:D_{\kappa }\buildrel d:1 \over \longrightarrow D_{\kappa -1}\) has degree \(d=m+n\). In particular, for \(t\in \mathbf {\Omega }^\alpha _\mathrm{res}\) there exists some simply connected and forward invariant domain \(D_\kappa \subset \mathscr {A}_t\) that contains \(v^\beta _t\) (Figs. 1, 2).

We note some more simple consequences of Lemma 1; our focus is on the critical point \(\alpha \) and the \({}^\alpha \)-sets.

  • \(\{t:0<|t|\le 1\}\subset \mathbf {\Omega }^\alpha _0\);

  • \(\overline{\mathbf {\Omega }^\alpha _\mathrm{res}}\subset \mathbb D;\)

  • \(\alpha \) is inactive on \(0<|t|\le 1\);

  • \(\overline{\bigcup _{k\ge 1}(\mathbf {\Omega }^\alpha _k\cup \mathbf {W}^\alpha _k)}\subset \{t:1<|t|<T\}\) for some \(T=T_{mn}>1\);

  • \(\mathbf {B}^\alpha \subset \{t:1<|t|<T\}\) for some \(T=T_{mn}>1\).

The consequences for the dynamical planes are as follows.

Theorem 1

For \(t\in \mathbf {\Omega }^\alpha _0\), the basin \(\mathscr {A}_t\) is completely invariant, and any other Fatou component is simply connected. Moreover,

  • for \(t\in \mathbf {\Omega }^\alpha _0\cap \mathbf {\Omega }^\beta _0\) also \(\mathscr {B}_t\) is completely invariant, the Julia set \(\mathscr {J}_t=\partial \mathscr {A}_t=\partial \mathscr {B}_t\) is a quasi-circle, and \(f_t\) is quasi-conformally conjugated to \(z\mapsto z^d\);

  • for \(t\in \mathbf {\Omega }^\alpha _\mathrm{res}\), \(\mathscr {A}_t\) is infinitely connected and the Fatou set consists of \(\mathscr {A}_t\), \(\mathscr {B}_t\), and the predecessors of \(\mathscr {B}_t\) of any order.

Proof

To prove complete invariance of \(\mathscr {A}_t\) we first assume \(0<|t|\le 1\). Then \(\mathscr {A}_t\) contains the interval [0, 1] by Lemma 1, hence is completely invariant. If, however, \(|t|>1\), then \(\mathscr {B}_t\) is completely invariant, and any other Fatou component is simply connected. Assuming \(1\not \in \mathscr {A}_t\) (\(t\in \mathbf {\Omega }^\alpha _0\), \(|t|>1\)) we obtain either \(f_t:\mathscr {A}_t^*\buildrel n:1 \over \longrightarrow \mathscr {A}_t\) with \(n=(n-1)+1\) critical points if \(\alpha \in \mathscr {A}_t^*\) or else \(f_t:\mathscr {A}_t\buildrel m:1 \over \longrightarrow \mathscr {A}_t\) with \(m=(m-1)+1\) critical points if \(\alpha \in \mathscr {A}_t\), this contradicting simple connectivity of both domains \(\mathscr {A}_t\) and \(\mathscr {A}_t^*\) by the Riemann–Hurwitz formula.

The first assertion is obvious since \(\mathscr {B}_t\) shares the properties of \(\mathscr {A}_t\) and \(f_t\) is hyperbolic.

The second assertion follows from the Riemann-Hurwitz formula, since \(f_t:\mathscr {A}_t\buildrel d:1 \over \longrightarrow \mathscr {A}_t\) has degree d and \(r=(m-1)+(n-1)+1+1=d\) critical points 0, 1 (if \(n>1\)), \(\alpha \), and \(\beta \).\(\Box \)

Fig. 1
figure 1

Left the \(\alpha \)-parameter plane for \(\displaystyle f_t(z)=tz^2\frac{1-z}{1+z}\) displaying the unit circle, \(\mathbf {\Omega }^\alpha \) (gray), \(\mathbf {\Omega }^\alpha _\mathrm{res}\) and \(\mathbf {\Omega }^\beta _\mathrm{res}\) (white, in and outside the unit circle), and \(\mathbf {W}^\alpha \) (black). Right a neighbourhood of the origin displaying \(\mathbf {\Omega }^\alpha _\mathrm{res}\) (gray) surrounded by points of \(\mathbf {\Omega }^\alpha _0\) (white), \(\mathbf {\Omega }^\beta _k\) (\(k\ge 1\), white, small), and \(\mathbf {W}^\beta \) (black)

Fig. 2
figure 2

Left the parameter plane of \(P_c(z)=cz^2(z+1)\). The escape region for \(P_c\) (gray), the white region with slit, and the black regions correspond to \(\mathbf {\Omega }^\alpha _\mathrm{res}\), \(\mathbf {\Omega }^\beta \cap \mathbb D\), and \(\mathbf {W}^\beta \), in case of \(m=2\), \(n=1\), respectively. The punctured disc \(0<|t|<1\) corresponds to \(\mathbb C{\setminus }[-2,0]\) in the c-plane. Right the parameter plane of \(P_{-\frac{1}{2}(t+2+\frac{1}{t})}(z)\) in \(-0.2<\mathrm{Re}\;t<0.25,\) \(-0.25<\mathrm{Im}\;t<0.25\) (see also Fig. 1 right)

Theorem 2

\(\mathbf {\Omega }^\alpha _0\cup \{0\}\), \(\mathbf {\Omega }^\alpha _\mathrm{res}\cup \{0\}\), and the connected components of \(\mathbf {\Omega }^\alpha _k\) \((k\ge 1)\) are simply connected domains. Riemann maps onto \(\mathbb D\) are given by any branch of \(\root m \of {E_0(t)}\), \(\root m \of {E_\mathrm{res}(t)}\), and \(\root n \of {E_{k}(t)}\), respectively.

For the proof we need two auxiliary results on the maps

$$\begin{aligned} \begin{array}{rcll} E_0(t)&{}=&{}t\left( \Phi _t(v^\alpha _t)\right) ^{m-1}&{}\left( t\in \mathbf {\Omega }^\alpha _0\right) ,\\ E_\mathrm{res}(t)&{}=&{}t\left( \Phi _t(v^\beta _t)\right) ^{m-1}&{}\left( t\in \mathbf {\Omega }^\alpha _\mathrm{res}\right) , \mathrm{~and}\\ E_k(t)&{}=&{}t^{\frac{1}{m-1}}\Phi _t\left( f^k(v^\alpha _t)\right) &{}\left( t\in \mathbf {\Omega }^\alpha _k,~k\ge 1\right) , \end{array}\end{aligned}$$
(2)

where \(\Phi _t\) denotes the Böttcher function to the fixed point \(z=0\). In the first step (Lemma 2) of the proof of Theorem 2 we will show that the functions (2) provide proper maps on \(\mathbb D{\setminus }\{0\}\) and \(\mathbb D\), respectively, which are only ramified over the origin. In the second step (Lemma 3) this will be used to show that the corresponding domains (with 0 included, if necessary) are simply connected.

The solution to Böttcher’s functional equation

$$\begin{aligned} \Phi _t(f_t(z))=t\Phi _t(z)^m\quad (\Phi _t(z)\sim z \mathrm{~as~}z\rightarrow 0) \end{aligned}$$
(3)

is locally given by

$$\begin{aligned} \Phi _t(z)=\lim _{k\rightarrow \infty }\root m^k \of {f_t^k(z)/t^{1+m+\cdots +m^{k-1}}}=t^{-\frac{1}{m-1}}\lim _{k\rightarrow \infty }\root m^k \of {f_t^k(z)}; \end{aligned}$$

it conjugates \(f_t\) to \(\zeta \mapsto \zeta ^m\). This conjugation holds throughout \(\mathscr {A}_t\) in the third case, when \(\Phi _t\) maps \(\mathscr {A}_t\) conformally onto the disc \(|z|<|t|^{-\frac{1}{m-1}}\); the maps \(E_k\) are analytic and well-defined on the components of \(\mathbf {\Omega }^\alpha _k\), \(k\ge 1\).

In the first case the conjugation holds on some simply connected neighbourhood of \(z=0\) that contains \(z=0\) and \(z=v^\alpha _t\), but does not contain \(z=1\). The analytic continuation of \(\Phi _t\) causes singularities at \(z=1\) and its preimages under \(f_t^k\), nevertheless \(|\Phi _t(z)|\) is well-defined on \(\mathscr {A}_t\) and \(|\Phi _t(z)|\rightarrow |t|^{-\frac{1}{m-1}}\) as \(z\rightarrow \partial \mathscr {A}_t\) holds anyway. Thus \(E_0(t)=t\Phi _t(v^\alpha _t)^{m-1}\) is holomorphic on \(\mathbf {\Omega }^\alpha _0\) and zero-free, with \(E_0(t)\sim t(v^\alpha _t)^{m-1}=f_1(\alpha )^{m-1}t^m\) as \(t\rightarrow 0\).

In the second case we construct an exhaustion \((D_\kappa )\) of \(\mathscr {A}_t\) such that \(f_t:D_\kappa \buildrel d:1 \over \longrightarrow D_{\kappa -1}\) has degree d and \(D_\kappa \) is simply connected for \(\kappa \le \kappa _0\) with \(v^\beta _t\in D_{\kappa _0}\) and \(\beta \in D_{\kappa _0+1}{\setminus } D_{\kappa _0}\). This is possible by Lemma 1, and the procedure applied to \(t^{-\frac{1}{m-1}}\Phi _t(v^\alpha _t)\) on \(\mathbf {\Omega }^\alpha _{0}\) also applies to \(t^{-\frac{1}{m-1}}\Phi _t(v^\beta _t)\) on \(\mathbf {\Omega }^\alpha _\mathrm{res}\).

Lemma 2

The functions in (2) are well-defined and provide proper maps from \(\mathbf {\Omega }^\alpha _0\cup \{0\}\), \(\mathbf {\Omega }^\alpha _\mathrm{res}\cup \{0\}\), and the connected components of \(\mathbf {\Omega }^\alpha _k\) with \(k\ge 1\), respectively, onto the unit disc \(\mathbb D\).

Proof

To prove that \(|E_0(t)|\rightarrow 1\) as \(t\in \mathbf {\Omega }^\alpha _0\) tends to \(\partial \mathbf {\Omega }^\alpha _0{\setminus }\{0\}\) we choose any disc \(\triangle _r:|z|<r\) that is invariant under \(f_t\) for every \(t\in \mathbf {\Omega }^\alpha _0\). This is possible since \(\mathbf {\Omega }^\alpha _0\) is contained in some disc \(|t|<T\), hence we may choose \(r<1\) such that \(Tr^{m-1}\left( \frac{1+r}{1-r}\right) ^n=1\) holds. By \(k=k(t)\) we denote the largest integer such that \(f_t^k(v^\alpha _t)\not \in \triangle _r\). Then \(k(t)\rightarrow \infty \) as \(t\rightarrow \partial \mathbf {\Omega }^\alpha _0{\setminus }\{0\}\), and \(|f_t^{k(t)}(v^\alpha _t)|\ge r\) implies

$$\begin{aligned} \liminf _{t\rightarrow \mathbf {\Omega }^\alpha _0{\setminus }\{0\}} |\Phi _t(v^\alpha _t)|\ge \lim _{t\rightarrow \mathbf {\Omega }^\alpha _0{\setminus }\{0\}} |t|^{-\frac{1}{m-1}}\root m^{k(t)} \of {r}=|t|^{-\frac{1}{m-1}}, \end{aligned}$$

while \(|\Phi _t(z)|<|t|^{-\frac{1}{m-1}}\) is always true. Thus \(E_0\) maps each connected component of \(\mathbf {\Omega }^\alpha _0\) properly onto \(\mathbb D{\setminus }\{0\}\). It follows that the origin is removable for (a zero of) \(E_0\), and \(\mathbf {\Omega }^\alpha _0\cup \{0\}\) is a domain which is mapped by \(E_0\) properly with degree m onto the unit disc \(\mathbb D\).

If \(t\in \mathbf {\Omega }^\alpha _k\) for some \(k\ge 1\), then again \(|E_k(t)|\) tends to 1 as \(t\rightarrow \partial \Omega \), where \(\Omega \) is any component of \(\mathbf {\Omega }^\alpha _k\). Thus \(E_k\) is a proper map of \(\Omega \) onto \(\mathbb D\). We will prove that \(E_k\) is ramified only over zero even for \(k\ge 0\), that is \(E_k'(t)=0\) implies \(E_k(t)=0\). This is a well-known procedure, the idea of which is due to Roesch [13], and outlined in detail for the Morosawa-Pilgrim family \(z\mapsto t\left( 1+\frac{(4/27)z^3}{1-z}\right) \) in [17, Lemma 2].

We take any \(t_0\in \mathbf {\Omega }^\alpha _k\) and choose \(\varepsilon >0\) such that for t sufficiently close to \(t_0\), the closed disc \(\triangle _{3\epsilon }:|w-v^\alpha _{t_0}|\le 3\varepsilon \) belongs to the Fatou component \(D_{t_0}\) of \(f_{t_0}\) containing \(v^\alpha _{t_0}\) (\(D_{t_0}\) is a predecessor of \(\mathscr {A}_{t_0}\) of order \(\ell \ge 0\)). Furthermore let \(\eta _t:\widehat{\mathbb C}\longrightarrow \widehat{\mathbb C}\) be any diffeomorphism such that \(\eta _t(w)\) depends analytically on t, \(\eta _t(w)=w\) holds on \(|w-v^\alpha _{t_0}|\ge 3\varepsilon \) and \(\eta _t(w)=w+(v^\alpha _t-v^\alpha _{t_0})\) on \(|w-v^\alpha _{t_0}|<\varepsilon .\) Then \(g_t=\eta _t\circ f_{t_0}:\widehat{\mathbb C}\longrightarrow \widehat{\mathbb C}\) is a quasi-regular map which equals \(f_{t_0}\) on \(\widehat{\mathbb C}{\setminus } f^{-1}_{t_0}(\triangle _{3\epsilon })\), and is analytic on \(\widehat{\mathbb C}{\setminus } f_{t_0}^{-1}(A)\) with \(A=\{w:\varepsilon \le |w-v^\alpha _{t_0}|\le 3\varepsilon \}\). To apply Shishikura’s qc-lemma [14] we need to know that \(g_t\) is uniformly K-quasi-regular, that is, all iterates \(g_t^p\) are K-quasi-regular with one and the same K. This is obviously true if the sets \(f_{t_0}^{-p}(A)\) (\(p=1,2,\ldots \)) are visited at most once by any iterate of \(g_t\). This is trivial if \(k\ge 1\): the sets \(f_{t_0}^{-p}(A)\) belong to different Fatou components, namely predecessors of \(D_{t_0}\) of order p. If \(k=0\) the argument is different. Let \(\triangle _0:|z|<\delta \) be such that \(f_{t_0}(\overline{\triangle }_0)\subset \triangle _0\) and set \(\triangle _\nu =f_{t_0}^{-1}(\triangle _{\nu -1})\). Then choosing \(\epsilon >0\) sufficiently small we have \(A\subset \triangle _{\ell }{\setminus }\overline{\triangle _{\ell -1}}\) for some \(\ell \) and \(f_{t_0}^{-p}(A)\subset \triangle _{\ell +p}{\setminus }\overline{\triangle _{\ell +p-1}}\). By the above mentioned qc-lemma, \(g_t\) is quasi-conformally conjugated to some rational function

$$\begin{aligned} R_t=h_t\circ g_t\circ h_t^{-1}. \end{aligned}$$

The quasi-conformal mapping \(h_t:\widehat{\mathbb C}\longrightarrow \widehat{\mathbb C}\) is uniquely determined by the normalisation \(h_t(z)=z\) for \(z=0,\alpha ,1\), and depends analytically on the parameter t. Also \(h_t\) is analytic on \(\widehat{\mathbb C}{\setminus }\overline{\bigcup _{p\ge 0}f_{t_0}^{-p}(A)}\), which, in particular, contains the points 0, \(v^\alpha _t\), and \(v^\alpha _{t_0}\). We set \(z_0=h_t(-1)\) to obtain \(R_t(z)=a(t)z^m\left( \frac{1-z}{z-z_0}\right) ^n.\) Since \(h_t(\alpha )=\alpha \), \(R_t\) has a critical point at \(z=\alpha \), and solving \(R_t'(\alpha )=0\) for \(z_0\) yields \(z_0=-1\), thus

$$\begin{aligned} R_t(z)=a(t)z^m\left( \frac{1-z}{1+z}\right) ^n. \end{aligned}$$

From \(R_t=h_t\circ \eta _t\circ f_{t_0}\) and \(h_t(\alpha )=\alpha \), however, it follows that

$$\begin{aligned} a(t)v^\alpha _1=R_t(\alpha )=h_t\circ \eta _t\circ f_{t_0}(\alpha )=h_t\circ \eta _t\left( v^\alpha _{t_0}\right) =h_t\left( v^\alpha _t\right) , \end{aligned}$$

hence \(R_t(z)=f_\tau (z)\) with \(\tau =\tau (t)={h_t(v^\alpha _t)}/{v^\alpha _1}\) and \(v^\alpha _\tau =h_t(v^\alpha _t);\) in particular, \(\tau \) depends analytically on t. On some neighbourhood of \(z=0\) we have

$$\begin{array}{rcl} \displaystyle (t_0/\tau )^{\frac{1}{m-1}}\Phi _{t_0}\circ h_t^{-1}\circ f_\tau &{}=&{}\displaystyle (t_0/\tau )^{\frac{1}{m-1}}\Phi _{t_0}\circ g_t\circ h_t^{-1}\\ &{}=&{}\displaystyle (t_0/\tau )^{\frac{1}{m-1}}\Phi _{t_0}\circ \eta _t\circ f_{t_0}\circ h_t^{-1}\\ &{}=&{}\displaystyle (t_0/\tau )^{\frac{1}{m-1}}\Phi _{t_0}\circ f_{t_0}\circ h_t^{-1}\\ &{}=&{}\displaystyle (t_0/\tau )^{\frac{1}{m-1}}t_0\left( \Phi _{t_0}\circ h_t^{-1}\right) ^{m}\\ &{}=&{}\displaystyle \tau \left( (t_0/\tau )^{\frac{1}{m-1}}\Phi _{t_0}\circ h_t^{-1}\right) ^{m}, \end{array}$$

hence \(\phi _\tau =(t_0/\tau )^{\frac{1}{m-1}}\Phi _{t_0}\circ h_t^{-1}\) solves Böttcher’s functional equation

$$\begin{aligned} \phi _\tau \circ f_\tau (z)=\tau (\phi _\tau (z))^m. \end{aligned}$$

Since \(\tau \) and \(h_t\) depend analytically on t, this is also true for \(h_t^{-1}\), which is not self-evident. Also from \(h_t(g_t(z))=f_\tau (h_t(z))\sim \tau h_t(z)^m\) and \(g_t(z)=f_{t_0}(z)\sim t_0z^m\) as \(z\rightarrow 0\) it follows that \(h_t(t_0z^m)\sim \tau h_t(z)^m\), hence \(h_t(z)\sim \root m-1 \of {t_0/\tau }z\), \(h^{-1}_t(z)\sim \root m-1 \of {\tau /t_0}z\) and \(\phi _\tau (z)\sim \lambda z\) as \(z\rightarrow 0\), with \(\lambda ^{m-1}=1\). This implies \(\phi _\tau =\lambda \Phi _\tau \) by uniqueness of the Böttcher coordinate, and from \(\tau (t_0)=t_0\) and analytic dependence on t it follows that \(\lambda =1\) and \(\phi _\tau =\Phi _\tau \), hence

$$\begin{array}{rcl} E_k(\tau )&{}=&{}\tau ^{\frac{1}{m-1}}\Phi _\tau (Q_k(\tau ))=\tau ^{\frac{1}{m-1}}\Phi _\tau (f^k_\tau (v^\alpha _\tau ))\\ &{}=&{} t_0^{\frac{1}{m-1}}\Phi _{t_0}\circ h_t^{-1}(f^k_\tau (v^\alpha _\tau ))=t_0^{\frac{1}{m-1}}\Phi _{t_0}\circ f^k_{t_0}\circ h_t^{-1}(v^\alpha _\tau )\\ &{}=&{}t_0^{\frac{1}{m-1}}\Phi _{t_0}(f^k_{t_0}(v^\alpha _t))\quad \mathrm{if~} k\ge 1,~\quad \mathrm{and}\\ E_0(\tau )&{}=&{}t_0(\Phi _{t_0}(v^\alpha _t))^m. \end{array}$$

Since \(t\mapsto \tau \) is locally univalent, \(E_k\) is univalent at \(t_0\) if and only if the map \(t\mapsto t_0^{\frac{1}{m-1}}\Phi _{t_0}(f_{t_0}^k(v^\alpha _t))\) is univalent on some neighbourhood of \(t_0\). If \(k\ge 1\), \(\Phi _{t_0}\) is univalent on \(\mathscr {A}_{t_0}\), and \(f_{t_0}^k\) is univalent on \(|z-v^\alpha _{t_0}|<\delta \) provided \(Q_k(t_0)=f^k_{t_0}(v^\alpha _{t_0})\ne 0\), while \(f_{t_0}^k\) is n-valent at \(v^\alpha _{t_0}\) if \(Q_k(t_0)=0\). In case of \(k=0\) we note that \(\Phi _{t_0}\) is locally univalent on some forward invariant domain D that contains 0 and \(v^\alpha _{t_0}\), and \(v^\alpha _t=tv^\alpha _1\ne 0\) is trivially univalent.

\(\Box \)

The proof of Theorem 2 will be finished by

Lemma 3

Let h be a proper map of degree m of the domain D onto the unit disc \(\mathbb D\), and assume that h is ramified exactly over zero, that is, \(h'(z)=0\) implies \(h(z)=0\). Then D is simply connected and h has a single zero on D.

Proof

Assume that h has zeros with multiplicities \(m_\nu \) (\(1\le \nu \le n)\). Then h has degree \(d=m_1+\cdots +m_n\) and \(r=d-n\) critical points. The Riemann-Hurwitz formula then yields \(\# D-2=-d+r=-n\), hence \(\#D=2-n\), which only is possible if \(n=1\) and \(\# D=1\).\(\Box \)

Remark

Each connected component of \(\mathbf {\Omega }^\alpha _k\) contains a zero of \(Q_k(t)=tf_1(Q_{k-1}(t))\) which is not a zero of \(Q_{k-1}\), hence is a zero of \(Q_{k-1}(t)-1\). Thus \(\mathbf {\Omega }^\alpha _k\) consists of at most \(\frac{d^k-1}{d-1}\) connected components.

4 The hyperbolic loci

The bifurcation locus \(\mathbf {B}^\beta \) is contained in some annulus \(\delta<|t|<1\), and this also holds for \(\mathbf {W}^\beta \). Hence (super-)attracting cycles \(U_1,\ldots ,U_k\) that contain the critical point \(\beta \) may occur only if \(\delta<|t|<1\).

Theorem 3

For \(0<|t|<1\), \(f_t\) is quasi-conformally conjugated to some polynomial

$$\begin{aligned} P_c(z)=cz^m(z+1)^n\quad (c=c_t\ne 0) \end{aligned}$$

with free critical point \(-\frac{m}{m+n}\). The basin \(\mathscr {A}_t\) is completely invariant, and simply connected if and only if \(t\not \in \mathbf {\Omega }^\alpha _\mathrm{res}.\) For \(t\not \in \mathbf {\Omega }^\beta _0\), the Fatou set consists of \(\mathscr {A}_t\), the simply connected basin \(\mathscr {B}_t\) and its pre-images and, additionally, of some (super-)attracting cycle of Fatou components and their pre-images if \(t\in \mathbf {W}^\beta \); the cycle absorbs the critical point \(\beta \).

Proof

To prove the second assertion we note that by Lemma 1 the pre-image D of the disc \(\triangle =\triangle _{r|t|}\) is a simply connected Jordan domain that contains \(\overline{\triangle }\cup [0,1]\), but does not contain \(v^\beta _t\). Then \(D_2=\widehat{\mathbb C}{\setminus }\overline{\triangle }\) is a backward invariant domain, and

$$\begin{aligned} f_t:D_1\buildrel d:1 \over \longrightarrow D_2\quad \left( D_1=f_t^{-1}(D_2)\right) \end{aligned}$$

is a polynomial-like mapping in the sense of [6], of degree \(d=m+n\), hence is hybrid equivalent to some polynomial P of degree d. We may assume that the quasi-conformal conjugation \(\psi _t\) with

$$\begin{aligned} \psi _t\circ f_t=P\circ \psi _t \end{aligned}$$

maps \(\infty , 0,\) and \( -1\) onto \(0,\infty ,\) and \(-1\), respectively. Thus P is given by \(P(z)=P_c(z)=cz^m(z+1)^n\), and \(\psi _t\), hence also \(c=c_t\) depends analytically on t. \(\Box \)

Remark

We note that \(D_2=D_2(|t|)=\{z:|z|>r|t|\}\) increases if |t| decreases, while \(D_1=f_t^{-1}(\widehat{\mathbb C}{\setminus }\overline{\triangle }_{r|t|})=f_1^{-1}(\widehat{\mathbb C}{\setminus }\overline{\triangle }_r)\) is independent of t. Thus the conformal modulus \(\mu (|t|)\) of \(D_2(|t|){\setminus }\overline{D_1}\) satisfies \(\mu (1)\le \mu (|t|)-\log \frac{1}{|t|}\le \log \frac{\inf _{z\in D_1}|z|}{r}\). The bifurcation locus of \(P_c\) corresponds conformally to the bifurcation locus \(\mathbf {B}^\beta \), and the hyperbolic components are just quasi-conformal images of the hyperbolic components of the quadratic family \(z\mapsto z^2+\xi \).

For \(t\in \mathbf {W}_k\), the multiplier map \(t\mapsto \lambda _t\) is an algebraic function of t. This is easily seen by writing the equations \(f^k_t(z)=z\) and \(\lambda =(f^k_t)'(z)\) as polynomial equations \(q_1(z,t)=0\) and \(q_2(z,t,\lambda )=0\), and computing the resultant \(R_f(t,\lambda )\) of \(q_1\) and \(q_2\) with respect to z. For example, in case of \(k=1\), \(m=2\), and \(n=1\) we obtain

$$\begin{aligned} R_f(t,\lambda )=\left[ -2+14t-2t^2\right] +\left[ 1-10t+t^2\right] \lambda +2t\lambda ^2=0. \end{aligned}$$

For \(P_c(z)=cz^2(z+1)\) we obtain in the same manner (multiplier \(\mu \))

$$\begin{aligned} R_P(c,\mu )=9+2c-(c+6)\mu +\mu ^2=0. \end{aligned}$$

Since the quasi-conformal conjugation respects multipliers (\(\lambda _t=\mu _{c_t})\), \(c_t\) is an algebraic function of t by the identity theorem; in the present case we obtain \((1+2t+t^2+2tc)^2=0\) by computing the resultant of \(R_f(t,\lambda )\) and \(R_P(c,\lambda )\) with respect to \(\lambda \), hence

$$\begin{aligned} t\mapsto c=c_t=-{\textstyle \frac{1}{2}}\left( t+2+\frac{1}{t}\right) \quad \left( \textstyle c=-\frac{9}{2}\leftrightarrow t=\frac{1}{2}\left( \sqrt{49}-\sqrt{45}\right) \right) \end{aligned}$$

maps \(0<|t|<1\) conformally onto \(\mathbb C{\setminus }[-2,0]\), see Fig. 2.

The following result was not explicitly stated but proved in [17]. The proof is an adaption of the procedure due to Douady [5], applied to the hyperbolic components of the quadratic family \(R_t(z)=z^2+t\) with one free critical point. The occurrence of several critical points requirers a slightly more sophisticated argument. The present version applies to a wider class of functions like \(R_t(z)=z^d+t\), \(R_t(z)=z^m+t/z^n\), \(R_t(z)=t\left( 1+\frac{((d-1)^{d-1}/d^d)z^d}{1-z}\right) \) (\(d\ge 3\)), \(R_t(z)=-\frac{t}{4}\frac{(z^2-2)^2}{z^2-1}\), the present family, and many others, to show that the hyperbolic components are simply connected and are mapped properly onto the unit disc by the multiplier map \(t\mapsto \lambda _t\).

Theorem 4

Let \((R_t)_{t\in T}\) be any family of rational maps that is analytically parametrised over some domain T. Suppose that each \(R_t\) has a (super-)attracting cycle \(U_0\buildrel m_1:1 \over \longrightarrow U_1\buildrel m_2:1 \over \longrightarrow \ldots \buildrel m_{n-1}:1 \over \longrightarrow U_{n-1}\buildrel m_{n}:1 \over \longrightarrow U_{n}=U_0,\) such that \(R_t^n\) has a single critical point \(c_t\in U_0\) of multiplicity \(m-1\), where \(m=m_1\cdots m_n\) is the degree of \(R^n_t:U_0\buildrel m:1 \over \longrightarrow U_0\). Assume also that the multiplier \(\lambda _t\) satisfies \(|\lambda _t|\rightarrow 1\) as \(t\rightarrow \partial T\). Then the multiplier map \(t\mapsto \lambda _t\) provides a proper map \(T\buildrel (m-1):1 \over \longrightarrow \mathbb D\) which is ramified just over \(w=0\), and T is simply connected.