1 Background

The purpose of this note is to study the Fatou–Julia dichotomy, not for the iterates of a single holomorphic endomorphism of \(\mathbb C^k, \; k \ge 2\), but for a family \(\mathcal {F}\) of such maps. The Fatou set of \(\mathcal {F}\) will be by definition the largest open set where the family is normal, i.e., given any sequence in \(\mathcal {F}\) there exists a subsequence which is uniformly convergent or divergent on all compact subsets of the Fatou set, while the Julia set of \(\mathcal {F}\) will be its complement.

We are particularly interested in studying the dynamics of families that are semigroups generated by various classes of holomorphic endomorphisms of \(\mathbb C^k,\; k \ge 2.\) For a collection \(\{\psi _{\alpha }\}\) of such maps let

$$\begin{aligned} G=\langle \psi _{\alpha }\rangle \end{aligned}$$

denote the semigroup generated by them. The index set to which \(\alpha\) belongs is allowed to be uncountably infinite in general. The Fatou set and Julia set of this semigroup G will be henceforth denoted by F(G) and J(G),  respectively. Also for a holomorphic endomorphism \(\phi\) of \(\mathbb C^k ,\) \(F(\phi )\) and \(J(\phi )\), will denote the Fatou set and Julia set for the family of iterations of \(\phi .\) The \(\psi _{\alpha }\) that will be considered in the sequel will belong to one of the following classes:

  • \(\mathcal {E}_k{:}\;\) The set of holomorphic endomorphisms of \(\mathbb C^k\) which have maximal generic rank k.

  • \(\mathcal {I}_k{:}\;\) The set of injective holomorphic endomorphisms of \(\mathbb C^k.\)

  • \(\mathcal {V}_k{:}\;\) The set of volume preserving biholomorphisms of \(\mathbb C^k.\)

  • \(\mathcal {P}_k{:}\;\) The set of proper holomorphic endomorphisms of \(\mathbb C^k.\)

The main motivation for studying the dynamics of semigroups in higher dimensions comes from the results of Hinkkanen–Martin [7] and Fornaess–Sibony [5]. While [7] considers the dynamics of semigroups generated by rational functions on the Riemann sphere, [5] puts forth several basic results about the dynamics of the iterates of a single holomorphic endomorphism of \(\mathbb C^k,\; k \ge 2.\) Under such circumstances, it seemed natural to us to study the dynamics of semigroups in higher dimensions.

Section 2 deals with basic properties of F(G) and J(G) when G is generated by elements that belong to \(\mathcal {E}_k\) and \(\mathcal {P}_k.\) The main theorem in Sect. 3 states that if J(G) contains an isolated point, then G must contain an element that is conjugate to an upper triangular automorphism of \(\mathbb C^k\) and in Sect. 4, we discuss a few interesting examples of Julia set of a semigroup. Finally, we define recurrent domains for semigroups in Sect. 5 and provide a classification of such domains under some conditions which are generalizations of the corresponding statements of Fornaess–Sibony [5] for the iterates of a single holomorphic endomorphism of \(\mathbb C^k,\; k\ge 2.\) The classification for recurrent Fatou components for the iterates of holomorphic endomorphisms of \(\mathbb P^2\) and \(\mathbb P^k\) is studied in [4] and [3], respectively. In [4], Fornaess–Sibony also gave a classification of recurrent Fatou components for iterations of Hénon maps inside \(K^+\), which was initially considered by Bedford–Smillie in [1]. A classification for non-recurrent, non-wandering Fatou components of \(\mathbb P^2\) is given in [11], whereas a classification of invariant Fatou components for nearly dissipative Hénon maps is studied in [9].

2 Properties of the Fatou set and Julia set for a semigroup G

In this section, we will prove some basic properties of the Fatou set and the Julia set for semigroups.

Proposition 2.1

Let G be a semigroup generated by elements of \(\mathcal {E}_k\) where \(k \ge 2\) and for any \(\phi \in G\) define

$$\begin{aligned} \Sigma _{\phi }=\{z \in \mathbb C^k{:}\; \det {\phi (z)}=0\}. \end{aligned}$$

Then, for every \(\phi \in G\)

  1. (i)

    \(\phi (F(G)\ \Sigma _{\phi })\subset F(G).\)

  2. (ii)

    \(J(G)\cap \phi (\mathbb C^k)\subset \phi (J(G)),\) if G is generated by elements of \(\mathcal {P}_k\) or \(\mathcal {I}_k.\)

Proof

Note that \(\phi \in G\) is an open map at any point \(z \in F(G)\ \Sigma _{\phi }.\) Since for any sequence \(\psi _n \in G\), the sequence \(\psi _n \circ \phi\) has a convergent subsequence around a neighbourhood of z (say \(V_z\)), \(\psi _n\) also has a convergent subsequence on the open set \(\phi (V_z)\) containing \(\phi (z).\)

Now if G is generated by elements of \(\mathcal {P}_k\) or \(\mathcal {I}_k,\) then \(\phi\) is an open map at every point in \(\mathbb C^k.\) Then, the Fatou set is forward invariant and hence the Julia set is backward invariant in the range of \(\phi .\) \(\square\)

A family of endomorphisms \(\mathcal {F}\) in \(\mathbb C^k\) is said to be locally uniformly bounded on an open set \(\Omega \subset \mathbb C^k\) if for every point there exists a small enough neighbourhood of the point (say \(V\subset \Omega\)) such that \(\mathcal {F}\) restricted to V is bounded, i.e.,

$$\begin{aligned} \Vert f\Vert _{V}=\sup _{V}|f(z)|<M \end{aligned}$$

for some \(M>0\) and for every \(f \in \mathcal {F}.\)

Proposition 2.2

Let \(G=\langle \phi _1,\phi _2,\ldots ,\phi _n\rangle ,\) where each \(\phi _j \in \mathcal {E}_k\) and let \(\Omega _G\) be a Fatou component of G such that G is locally uniformly bounded on \(\Omega _G.\) Then for every \(\phi \in G\) the image of \(\Omega _G\) under \(\phi ,\) i.e., \(\phi (\Omega _G)\) is contained in Fatou set of G.

Proof

Let \(K \subset \subset \Omega _G\), i.e., K is a relatively compact subset of \(\Omega _G,\) then

Claim \(\Omega _G\) is a Runge domain, i.e., \(\hat{K} \subset \Omega _G\) where

$$\begin{aligned} \hat{K}{:}\; =\{z \in \mathbb C^k{:}\; |P(z)| \le \sup _{K}|P| ~\text{ for } \text{ every } \text{ polynomial }~ P\}. \end{aligned}$$

Let \(K_{\delta }=\{z \in \mathbb C^k{:}\; ~\text{ dist }(z,K)\le \delta \}\). Choose \(\delta >0\) such that \(K_{\delta } \subset \subset \Omega _G.\) Now note that \(\hat{K_\delta }\subset \subset \mathbb C^k\) , \(\hat{K_{\delta }}\supset \hat{K}\) and G is uniformly bounded on \(K_{\delta }.\) Pick \(\phi \in G.\) Then, there exists a polynomial endomorphism \(P_{\phi }\) of \(\mathbb C^k\) such that

$$\begin{aligned} |\phi (z)-P_{\phi }(z)| \le \epsilon \quad ~\text{ for } \text{ every }~z \in \hat{K_{\delta }},\\ \text{ i.e., } |P_{\phi }(z)|-\epsilon \le |\phi (z)| \le |P_{\phi }(z)|+\epsilon . \end{aligned}$$

Hence

$$\begin{aligned} |\phi (z)|&\le |P_{\phi }(z)|+\epsilon \le \sup _{K_\delta }|P_{\phi }(z)|+\epsilon \\&\le \sup _{K_\delta }|{\phi }(z)|+2 \epsilon \le M+2\epsilon \end{aligned}$$

for every \(z \in \hat{K_{\delta }}\) and some constant \(M > 0.\) So G is uniformly bounded on \(\hat{K_{\delta }}\) and \(\hat{K} \subset \Omega _G.\)

Let

$$\begin{aligned} \Sigma _i=\{z \in \mathbb C^k{:}\; \det {\phi _i(z)}=0\} \end{aligned}$$

for every \(1 \le i \le n\) and

$$\begin{aligned} \Sigma =\bigcup _{i=1}^n \Sigma _i. \end{aligned}$$

Thus \(\phi _i\) for every i, where \(1 \le i \le n\) is an open map in \(\Omega _G\ \Sigma\). Hence \(\phi _i(\Omega _G\ \Sigma )\) is contained inside a Fatou component say \(\Omega _i\) and G is locally uniformly bounded on each of \(\Omega _i\) for every \(1 \le i \le n,\) i.e., each \(\Omega _i\) is a Runge domain.

Now pick \(p \in \Omega _G \cap \Sigma .\) Since \(\Sigma\) is a set with empty interior, there exists a sufficiently small disc centred at p say \(\Delta _p\) such that \(\overline{\Delta }_p\ \{p\} \subset \Omega _G \ \Sigma .\) Then, \(\phi _i(\overline{\Delta }_p\ \{p\}) \subset \Omega _i\) for every \(1\le i \le n\) and since each \(\Omega _i\) is Runge \(\phi _i(p) \in \Omega _i,\) i.e., \(\phi _i(\Omega _G)\) is contained in the Fatou set for every \(1 \le i \le n.\) Now for any \(\phi \in G\) there exists a \(m > 0\) such that

$$\begin{aligned} \phi =\phi _{n_1}\circ \phi _{n_2}\circ \cdots \circ \phi _{n_m} \end{aligned}$$

where \(1 \le n_j \le n\) for every \(1 \le j \le m.\) Thus, applying the above argument repeatedly for each \(\phi _{n_j}(\tilde{\Omega }_j)\) where G is locally uniformly bounded on \(\tilde{\Omega }_j\) it follows that \(\phi (\Omega _G)\) is contained in the Fatou set of G. \(\square\)

Proposition 2.3

If \(G=\langle \phi _1,\phi _2,\ldots ,\phi _n\rangle\) where each \(\phi _i \in \mathcal {E}_k\) for every \(1 \le i \le n\) and let \(\Omega _G\) be a Fatou component of G. Then for any \(\phi \in G\) there exists a Fatou component of G, say \(\Omega _{\phi }\) such that \(\phi (\Omega _G) \subset \bar{\Omega }_{\phi }\) and

$$\begin{aligned} \partial \Omega _G\subset \bigcup _{i=1}^n \phi _i^{-1}(\partial \Omega _{\phi _i}). \end{aligned}$$

Proof

Let \(\phi \in G\) and let \(\Sigma _{\phi }\) denote the set of points in \(\mathbb C^k\) where the Jacobian of \(\phi\) vanishes. Since \(\Omega _G \ \Sigma _{\phi }\) is connected it follows that \(\phi (\Omega _G \ \Sigma _{\phi }) \subset \Omega _{\phi }\) where \(\Omega _{\phi }\) is a Fatou component of G and by continuity \(\phi (\Omega _G) \subset \bar{\Omega }_{\phi }.\)

Pick \(p \in \partial \Omega _G\) such that \(p \notin \partial \Omega _{\phi _i}\) for every \(1 \le i \le n.\) Since \(\phi _i(\Omega _G) \subset \bar{\Omega }_{\phi _i}\), \(\phi _i(p) \in \Omega _{\phi _i}\) for every \(1 \le i \le n.\) So there exists \(V_{\phi _i}\) an open neighbourhood of \(\phi _i(p)\) in \(\Omega _{\phi _i}\) for every i. Let \(V_p\) be a neighbourhood of p such that

$$\begin{aligned} \bar{V}_p\subset \bigcap _{i=1}^n \phi _i^{-1}(V_{\phi _i}). \end{aligned}$$

Let \(\{\psi _n\}\) be a sequence in G and without loss of generality it can be assumed that there exists a subsequence such that \(\psi _n=f_n \circ \phi _1.\) Now \(\phi _1(\bar{V}_p)\) is a compact subset in \(\Omega _1\) and \(f_n\) has a subsequence which either converges uniformly on \(\phi _1(\bar{V}_p)\) or diverges to infinity. Thus, \(V_p\) is contained in the Fatou set of G which is a contradiction! \(\square\)

The next observation is an extension of the fact that if \(\phi \in \mathcal {P}_k\), then \(F(\phi )=F(\phi ^n)\) for every \(n > 0\) for the case of semigroups.

Definition 2.4

Let G be a semigroup generated by endomorphisms of \(\mathbb C^k.\) A sub-semigroup H of G is said to have finite index if there is a finite collection of elements say \(\psi _1,\psi _2,\ldots ,\psi _ {m-1} \in G\) such that

$$\begin{aligned} G =\Big (\bigcup _{i=1}^{m-1} \psi _i \circ H\Big ) \cup H. \end{aligned}$$

The index of H in G is the smallest possible number m.

Definition 2.5

A sub-semigroup H of a semigroup G of endomorphisms of \(\mathbb C^k\) is of co-finite index if there is a finite collection of elements say \(\psi _1,\psi _2,\ldots ,\psi _{m-1} \in G\) such that either

$$\begin{aligned} \psi \circ \psi _{j} \in H \; \text{ or }\; \psi \in H \end{aligned}$$

for every \(\psi \in G\) and for some \(1 \le j \le m-1.\) The index of H in G is the smallest possible number m.

Proposition 2.6

Let G be a semigroup generated by proper holomorphic endomorphisms of \(\mathbb C^k\) and H be a sub-semigroup of G which has a finite (or co-finite) index in G. Then, \(F(G)=F(H)\) and \(J(G)=J(H).\)

Proof

From the definition itself it follows that \(F(G) \subset F(H).\) To prove the other inclusion, pick any sequence \(\{\phi _n \}\in G\). Since H has a finite index in G, there exists \(\psi _i\), \(1 \le i \le m-1\) such that

$$\begin{aligned} G =\left (\bigcup _{i=1}^{m-1} \psi _i \circ H\right) \cup H. \end{aligned}$$

So without loss of generality one can assume that there exists a subsequence say \(\phi _{n_k}\) with the property

$$\begin{aligned} \phi _{n_k}=\psi _1 \circ h_{n_k} \end{aligned}$$

where \(\{h_{n_k}\}\) is a sequence in H. Now on F(H), the sequence \(\{h_{n_k}\}\) has a convergent subsequence. Hence, so do \(\{\phi _{n_k}\}\) and \(\{\phi _n\}\) as \(\psi _1\) is a proper map in \(\mathbb C^k.\) \(\square\)

Let G be a semigroup

$$\begin{aligned} G=\langle \phi _1,\phi _2,\ldots ,\phi _m\rangle \end{aligned}$$

where \(\phi _i \in \mathcal {P}_k\), for every \(1 \le i \le m\) and each of these \(\phi _i\) commute with each other, i.e., \(\phi _i \circ \phi _j=\phi _j \circ \phi _i\) for \(i \ne j.\) Let H be a sub-semigroup of G defined as

$$\begin{aligned} H=\langle \phi _1^{l_1},\phi _2^{l_2},\ldots ,\phi _m^{l_m}\rangle \end{aligned}$$

where \(l_i>0\) for every \(1 \le i \le m.\) Then, H has a finite index in G and hence by Proposition 2.6 \(F(G)=F(H).\)

Corollary 2.7

Let \(\phi _i\) be elements in \(\mathcal {P}_k\) for \(1 \le i \le m\), \(l=(l_1,l_2,\ldots ,l_m)\) an \(m\) -tuple of positive integers and \(G_l=\langle \phi _1^{l_1},\phi _2^{l_2},\ldots , \phi _m^{l_m}\rangle .\) Then, \(F(G_l)\) and \(J(G_l)\) are independent of the \(m\) -tuple l, if \(\phi _i \circ \phi _j=\phi _j\circ \phi _i\) for every \(1 \le i,j\le m,\) i.e., given two \(m\) -tuples p and q, \(F(G_p)=F(G_q).\)

Proof

Since \(G_l\) has a finite index in G for every \(m\)-tuple \(l=(l_1,l_2,\ldots ,l_m)\), it follows that \(F(G_l)=F(G)\) and \(J(G_l)=J(G).\) \(\square\)

Example 2.8

Let \(G=\langle f,g \rangle\) where \(f(z_1,z_2)=(z_1^2,z_2^2)\) and \(g(z_1,z_2)=(z_1^2/a,z_2^2)\) for \(a \in \mathbb C\) such that \(|a|>1.\) Then, it is easy to check that

$$\begin{aligned} J(f)=\big \{|z_1|=1\big \} \times \big \{|z_2| \le 1 \big \} \cup \big \{|z_1|\le 1\big \}\times \big \{|z_2| =1\big \} \end{aligned}$$

and

$$\begin{aligned} J(g)=\big \{|z_1|=|a|\big \} \times \big \{|z_2| \le 1\big \} \cup \big \{|z_1|\le |a|\big \}\times \big \{|z_2| =1\big \}. \end{aligned}$$

Now consider the bidisc \(\{|z_1 |< 1, |z_2 | < 1\}.\) Clearly, this domain is forward invariant under both f and g. This shows that \(\{|z_1 |< 1, |z_2 | < 1\} \subset F(G)\). Similarly observe that

$$\begin{aligned} \{|z_2 |> 1\} \cup \{|z_1 | > |a|\} \subset F(G). \end{aligned}$$

We claim that

$$\begin{aligned} \big \{ 1 \le |z_1| \le |a|\big \} \times \big \{ |z_2| \le 1 \big \} \subset J(G). \end{aligned}$$

Note that \(\{|z_1|=|a|,|z_2| \le 1\}\) is contained inside J(G) and since J(G) is backward invariant it follows that

$$\begin{aligned} \{|z_1|=|a|^{1/2},|z_2| \le 1\} \subset f^{-1}(\{|z_1|=|a|,|z_2| \le 1\}) \subset J(G). \end{aligned}$$

So inductively we get that

$$\begin{aligned} \{|z_1|=|a|^t,|z_2| \le 1\} \subset J(G) \end{aligned}$$

for any \(t=k2^{-n}\) where \(1 \le k \le 2^n\) and \(n \ge 1.\) As \(\{k2^{-n}{:}\; 1 \le k \le 2^n,\; n \ge 1\}\) is dense in [0, 1], it follows that \(\{ 1 \le |z_1| \le |a|\} \times \{ |z_2| \le 1 \} \subset J(G).\) Thus, the Julia set of the semigroup G is not forward invariant and clearly from the above observations one can prove that

$$\begin{aligned} J(G)=\big \{ |z_1| \le 1 \big \} \times \big \{ |z_2| = 1 \big \}\cup \big \{ 1 \le |z_1| \le |a|\big \} \times \big \{ |z_2| \le 1 \big \}. \end{aligned}$$

Example 2.9

Let \(T_0 (z) = 1,\; T_1 (z) = z\) and \(T_{n+1}(z) = 2z T_n (z)-T_{n-1}(z)\) for \(n \ge 1\) and \(G =\langle f_0 , f_1 , f_2 , \ldots \rangle\), with \(f_i (z_1 , z_2 ) = (T_i (z_1 ), z_2^2 )\) for \(i \ge 0.\) Consider

$$\begin{aligned} G_1 =\langle T_0 (z_1 ), T_1 (z_1 ), T_2 (z_1 ), ... \rangle ,\; G_2 =\left\langle z_2^2 \right\rangle . \end{aligned}$$

Since any sequence in \(G_1\) is uniformly unbounded on the complement of \([-1,1],\) it follows that

$$\begin{aligned} J(G) = [-1, 1] \times \{|z_2 | \le 1\}. \end{aligned}$$

Also, as \(J(G_1)\subset \mathbb C\) is completely invariant so is J(G).

3 Isolated points in the Julia set of a semigroup G

Proposition 3.1

Let \(G=\langle \phi _1, \phi _2, \ldots \rangle\) where each \(\phi _i \in \mathcal E_k\) . If the Julia set J (G) contains an isolated point (say a), then there exists a neighbourhood \(\Omega _a\) of a such that \(\Omega _a\ \{a\} \subset F(G)\) and \(\psi \in G\) which satisfies \(\Omega _a \subset \subset \psi (\Omega _a).\) In particular, if G is a semigroup generated by proper maps, then \(\psi ^{-1}(a)=a\).

Proof

Assume \(a=0\) is an isolated point in the Julia set J(G). Then there exists a sufficiently small ball \(B(0,\epsilon )\) around 0 such that \(B(0,\epsilon )\ \{0\}\) is contained F(G). Let

$$\begin{aligned} A{:}\; =\{z{:}\; \epsilon /2\le |z|\le \epsilon \}. \end{aligned}$$

Then \(A \subset F(G).\)

Claim There exists a sequence \(\phi _n \in G\) such that \(\phi _n\) diverges to infinity on A.

Suppose not. Then for every sequence \(\{\phi _n\} \in G\), there exists a subsequence \(\{\phi _{n_k}\}\) which converges to a finite limit in A. By the maximum modulus principle

$$\begin{aligned} \Vert {\phi _{n_k}}\Vert _{B(0,\epsilon )}<M. \end{aligned}$$

By the Arzelá–Ascoli theorem, it follows that \(\phi _{n_k}\) is equicontinuous on \(B(0,\epsilon )\), which contradicts that \(0 \in J(G).\)

By the same reasoning as above there exists a sequence \(\{\phi _n\} \in G\) such that it diverges uniformly to infinity on A but does not diverge uniformly to infinity on \(B(0,\epsilon )\), since it would again imply that \(B(0,\epsilon )\) is contained in the Fatou set of G. Thus, there exists a sequence of points \(x_n\) in \(B(0,\epsilon )\) such that \(\phi _n(x_n)\) is bounded, i.e.,

$$\begin{aligned} |\phi _n(x_n)|< M \end{aligned}$$

for some large \(M>0.\) So we can choose a subsequence of this \(\{\phi _n\}\) and relabel it as \(\{\phi _n\}\) again such that it satisfies the following condition:

$$\begin{aligned} \phi _n(x_n) \rightarrow q \quad ~\text{ and } \quad~ x_n \rightarrow p \end{aligned}$$

where \(p \in \overline{B(0,\epsilon )}.\)

Claim \(p=0\).

Suppose not. Then \(\phi _n(p)\) is bounded. Let \(\widetilde{A}=\{z{:}\; \min (|p|, \epsilon /2)\le |z| \le \epsilon \}.\) Then \(\widetilde{A} \supseteq A.\) Now \(\phi _{n_k}(p)\) converges on \(\widetilde{A}\), then \(\phi _{n_k}\) on \(\widetilde{A}\) converges to a finite limit, and hence on A by the maximum modulus principle. This is a contradiction!

Since \({\phi _{n}|}_{\partial B(0,\epsilon )} \rightarrow \infty\) for large n

$$\begin{aligned} \Vert {\phi _n} \Vert _{\partial B(0,\epsilon )} \gg |q|. \end{aligned}$$

Thus for a sufficiently large \(R>0\) and n

$$\begin{aligned} B(0,|q|+R)\cap \phi _n(B(0,\epsilon ))\ne \emptyset . \end{aligned}$$

Now, if \(B(0,\epsilon ) \nsubseteq \phi _n(B(0,\epsilon ))\), then \(B(0,|q|+R)\nsubseteq \phi _n(B(0,\epsilon ))\) since \(B(0,\epsilon ) \subset B(0,|q|+R)\) for large \(R>0.\) Then there exists \(y_n \in \partial B(0,\epsilon )\) such that \(|\phi _n(y_n)| <|q|+R\), which is not possible. Hence \(B(0,\epsilon ) \subset \subset \phi _n(B(0,\epsilon ))\) for sufficiently large n. Relabel this \(\phi _n\) as \(\psi\) and consider the neighbourhood \(\Omega _0\) as \(B(0,\epsilon ).\)

Since \(0 \in B(0,\epsilon )\subset \psi (B(0,\epsilon ))\), there exists \(\alpha \in B(0,\epsilon )\) such that \(\psi (\alpha )=0.\) From Proposition 2.1 it follows that \(\alpha =0.\) \(\square\)

Theorem 3.2

Let \(G=\langle \phi _1, \phi _2, \ldots \rangle\) where each \(\phi _i \in \mathcal I_k.\) If the Julia set J(G) contains an isolated point, say a then there exists an element \(\psi \in G\) such that \(\psi\) is conjugate to an upper triangular automorphism.

Proof

Without loss of generality we can assume that \(a=0.\) Now by Proposition 3.1 it follows that there exists a sufficiently small ball \(B(0,\epsilon )\) around 0 and an element \(\psi \in G\) such that \(B(0,\epsilon ) \subset \subset \psi (B(0,\epsilon )).\) Since \(\psi\) is injective map in \(\mathbb C^k\), \(\psi (B(0,\epsilon ))\) is biholomorphic to \(B(0,\epsilon )\) and hence we can consider the inverse, i.e.,

$$\begin{aligned} \psi ^{-1}{:}\; \psi (B(0,\epsilon )) \rightarrow B(0,\epsilon ). \end{aligned}$$

Note that \(\psi (B(0,\epsilon ))\) is bounded and \(B(0,\epsilon )\) is compactly contained in \(\psi (B(0,\epsilon )).\) Therefore, there exists an \(\alpha >1\) such that the map defined by

$$\begin{aligned} \psi _{\alpha }=\alpha \psi ^{-1}(z) \end{aligned}$$

is a self-map of the bounded domain \(\psi (B(0,\epsilon ))\) with a fixed point at 0. Then by the Carathéodory–Cartan–Kaup–Wu Theorem (see Theorem 11.3.1 in [8]), it follows that all the eigenvalues of \(\psi _\alpha\) are contained in the unit disc. Hence 0 is a repelling fixed point for \(\psi\) and also is an isolated point in the Julia set of \(\psi .\)

Since \(B(0,\epsilon )\ \{0\} \in J(G)\), \(B(0,\epsilon )\ \{0\}\) is also contained in the Fatou set of \(\psi\) and using the same argument as in the Proposition 3.1, there exists a subsequence (say \(n_k\)) such that

$$\begin{aligned} \Vert \psi ^{n_k}\Vert _{\partial B(0,\epsilon )} \rightarrow \infty \end{aligned}$$

uniformly. Thus for any given \(R>0,\) there exists \(k_0\) large enough such that \(B(0,R) \subset \psi ^{n_{k_0}}(B(0,\epsilon )).\) Hence \(\psi\) is an automorphism of \(\mathbb C^k\) and the basin of attraction of \(\psi ^{-1}\) at 0 is all of \(\mathbb C^k.\) Now by the result of Rosay–Rudin ([10]) \(\psi\) is conjugate to an upper triangular map. \(\square\)

Remark 3.3

The proof here shows that there exists a sequence \(\phi _n \in G\) such that each \(\phi _n\) is conjugate to an upper triangular map.

Recall that a domain \(\omega\) is holomorphically homotopic to a point in a domain \(\Omega\) if there exists a continuous map \(h{:}\; [0,1]\times \bar{\omega }\rightarrow \Omega\) with \(h(1,z)=z\) and \(h(0,z)=p\) where \(p \in \omega\) and \(h(t,\cdot)\) is holomorphic in \(\omega\) for every \(t \in [0,1].\)

Proposition 3.4

Let \(\phi\) be a non-constant endomorphism of \(\mathbb C^k\) such that on a bounded domain \(U \subset F(\phi )\) , the map \(\phi\) is proper onto its image, \(U \subset \subset \phi (U)\) and U is holomorphically homotopic to a point in \(\phi (U)\) then

  1. (i)

    \(\phi\) has a fixed point, say p in U.

  2. (ii)

    \(\phi\) is invertible at its fixed points.

  3. (iii)

    The backward orbit of \(\phi\) at the fixed point in U is finite, i.e., \(O^- (p) \cap U\) is finite where

    $$\begin{aligned} O^-_{\phi } (p)=\{ z \in \mathbb C^k{:}\; \phi ^n(z)=p, n \ge 1\}. \end{aligned}$$

Proof

That the map \(\phi\) has a fixed point p in U follows from Lemma 4.3 in [5].

Without loss of generality we can assume \(p=0\). Consider \(\psi (z)=\phi (p+z)-p\) and \(\Omega =\{z-p{:}\; z \in U \}.\) Then, \(\psi\) is the required map with the properties \(\Omega \subset \subset \psi (\Omega )\) and 0 is a fixed point for \(\psi .\)

Suppose \(\psi\) is not invertible at 0,  i.e., \(A=D\psi (0)\) has a zero eigenvalue. Let \(\lambda _i\), \(1 \le i \le k\) be the eigenvalues of A. Therefore, there exist an \(\alpha\) such that \(0< \alpha < 1\) and \(1 <m \le k\) such that \(0=|\lambda _i|< \alpha\) for \(1 \le i \le m\) and \(|\lambda _i|> \alpha\) for \(m< i \le k.\) Choose \(\delta >0\) such that

$$\begin{aligned} 0< \Vert D_{\mathbb C}\psi (z)-A\Vert < \epsilon _0=\min \Big \{\alpha ,\big ||\lambda _i|-\alpha \big |\Big \} \end{aligned}$$

for \(z \in B(0, \delta )\) and \(m< i \le k.\) Let \(\Psi\) be a Lipschitz map in \(\mathbb C^k\) such that

$$\begin{aligned} Lip(\Psi )=\Vert A\Vert +\epsilon _0 \end{aligned}$$

and

$$\begin{aligned} \Psi \equiv \psi \;\; \text{ on } \; \; B(0,\delta ). \end{aligned}$$

Now

$$\begin{aligned} W_s^{\Psi }{:}\; =\{z \in \mathbb C^k{:}\;|\alpha ^n \Psi ^n(z)|\;\text{ is } \text{ bounded }\} \end{aligned}$$

can be realized as a graph of a continuous function (see [12]) \(G_{\Psi }{:}\; \mathbb C^m \rightarrow \mathbb C^{k-m}\) such that \(G_{\Psi }(0)=0.\) Since

$$\begin{aligned} {W_s^{\Psi }}={W_s^{\psi }} \; \quad \text {on} \; B(0,\delta /2) \end{aligned}$$

\(W_s^{\psi } \cap \Omega\) is an infinite non-empty set containing 0. Also \({\psi ^{n_k}}_{|\bar{\Omega }} \rightarrow \psi _0\) for some sequence \(n_k\) and \(\psi _0\) is holomorphic on the component (say \(F_0\)) of \(F(\psi )\) containing \(\Omega\). Let

$$\begin{aligned} W_1^{\psi }=\{z \in F_0{:}\;\psi ^{n_k}(z)\rightarrow 0 \;\text{ as }\; k \rightarrow \infty \}. \end{aligned}$$

Then \(W_s^{\psi } \cap F_0 \subset W_1^{\psi }\) and

$$\begin{aligned} W_1^{\psi }=\bigcap _{i=1}^k {\psi _{0,i}}^{-1}(0) \end{aligned}$$

where \({\psi _{0,i}}\) is the ith coordinate function of \(\psi _0.\) If \(W_1^{\psi } \cap \partial \Omega =\emptyset\) then \(W_1^{\psi } \cap \Omega\) and hence \(W_s^{\psi } \cap \Omega\) will have to be finite which is not true. Thus, there exists a positive integer \(n_0\) such that \(\psi ^{n_0}(\partial \Omega ) \cap \Omega \ne \emptyset\) but by assumption it follows that \(\Omega \subset \subset \psi ^n(\Omega )\) for all \(n \ge 1,\) i.e., \(\psi ^{n}(\partial \Omega ) \cap \Omega =\emptyset\) for all \(n > 0.\) This proves that A has no zero eigenvalues.

Note that this observation also reveals that \(W_1^{\psi } \cap \Omega\) has to be a finite set, and since

$$\begin{aligned} O^-_{\psi }(0) \subset W_1^{\psi } \end{aligned}$$

the backward orbit of 0 under \(\psi\) is finite. \(\square\)

Now we can state and prove Theorem 3.2 for semigroups generated by the elements of \(\mathcal {E}_k.\)

Theorem 3.5

Let \(G=\langle \phi _1, \phi _2, \ldots \rangle\) where each \(\phi _i \in \mathcal {E}_k.\) If the Julia set J(G) contains an isolated point (say a) then there exists a \(\psi \in G\) such that \(\psi\) is conjugate to an upper triangular automorphism.

Proof

Assume \(a=0.\) Then, as before by Proposition 3.1 there exists a map \(\psi \in G\) and a domain \(\Omega\) such that \(\Omega \subset \subset \psi (\Omega ).\)

If 0 is in the Julia set of \(\psi ,\) then 0 is an isolated point in \(J(\psi )\) and by applying Theorem 4.2 in [5], it follows that \(\psi\) is conjugate to an upper triangular automorphism.

Suppose \(\Omega \subset F(\psi ).\) By Proposition 3.4, \(\psi\) has a fixed point in \(\Omega ,\) i.e., \(\{\psi ^n\}\) has a convergent subsequence in \(\bar{\Omega }\).

Case 1 Suppose that \(G=\langle \phi _1,\phi _2, \ldots \rangle\) where each \(\phi _i \in \mathcal P_k.\)

Applying Proposition 3.1, we have that \(\psi ^{-1}(0)=0\) and there exists \(\psi \in G\) such that

$$\begin{aligned} \Omega \subset \subset B(0,R) \subset \subset \psi (\Omega ) \end{aligned}$$
(3.1)

where \(\Omega\) is a sufficiently small ball at 0 and \(R>0\) is a sufficiently large number. Now, let \(\omega\) is the component of \(\psi ^{-1}(B(0,R))\) in \(\Omega\) containing the origin. Also from Proposition 3.4 it follows that 0 is a regular point of \(\psi\), which implies that \(\psi\) is a biholomorphism on \(\omega .\) Define \(\Psi _\beta\) on \(\psi (\omega )\) as

$$\begin{aligned} \Psi _\beta (z)=\beta \psi ^{-1}(z) \end{aligned}$$

and note that \(\Psi _\beta\) is a self-map of B(0, R) for some \(\beta >1\) with a fixed point at 0. Then, the eigenvalues of \(D_{\mathbb C}{\Psi _{\beta }}(0)\) are in the closed unit disc, i.e.,

$$\begin{aligned} \beta |\lambda _i^{-1}| \le 1 \end{aligned}$$

where \(\lambda _i\) are eigenvalues of A. Hence 0 is a repelling fixed point for the map \(\psi\) and \(0 \notin F(\psi ).\) Since 0 is an isolated point in the Julia set of \(\psi\), by Theorem 4.2 in [5] \(\psi\) is conjugate to an upper triangular automorphism of \(\mathbb C^k.\)

Case 2 Suppose that \(G=\langle \phi _1,\phi _2, \ldots \rangle\) where each \(\phi _i \in \mathcal E_k.\)

As before by Proposition 3.1 there exists \(\psi \in G\) such that

$$\begin{aligned}\Omega \subset B(0,R) \subset \psi (\Omega )\end{aligned}$$

and let \(\omega\) be a component of \(\psi ^{-1} (B(0,R)) \subset \Omega .\) Then, \(\omega\) satisfies all the condition of Proposition 3.4 and hence there exists a fixed point p of \(\psi\) in \(\omega\) and \(O^-_{\psi }(p)\cap \omega\) is finite.

Claim \(\psi ^{-1}(p) \cap \omega =p\)

Suppose not, i.e.,

$$\begin{aligned} \#\{\psi ^{-1}(p)\}=\text {the cardinality of}\; \psi ^{-1}(p) =m \end{aligned}$$

and \(m \ge 2.\) Let \(a_1 \in \psi ^{-1}(p)\ \{p\}\) in \(\omega\) and define

$$\begin{aligned} S_1=O^-_{\psi }(a_1) \cap \omega . \end{aligned}$$

Then \(S_1 \subset O^-_{\psi }(p)\cap \omega .\) Now choose inductively \(a_n \in \psi ^{-1}(a_{n-1}) \ \{a_{n-1}\}\) for \(n \ge 2\) and define

$$\begin{aligned} S_n=O^-_{\psi }(a_n) \cap \omega . \end{aligned}$$

Then

$$\begin{aligned} S_n \subset S_{n-1} \; \; \text{ and } \; \; \bigcup _{i=1}^n S_i\subset O^-_{\psi }(p)\cap \omega \end{aligned}$$

for every \(n \ge 2.\) Note that \(a_n \notin S_n,\) otherwise there is a positive integer \(k_n >0\) such that \(\psi ^{k_n}(a_n)=a_n,\) i.e., \(a_n\) is a periodic point of \(\psi\), and

$$\begin{aligned} \psi ^{k_n+m}(a_n)=p \end{aligned}$$

for any \(m>n\). Since \(O^-_{\psi }(p)\cap \omega\) is finite it follows that \(S_n\) has to be empty for large n. This implies that there exists a \(n_0 \ge 1\) such that \(\psi ^{-1}(a_{n_0})=a_{n_0}\) and \(a_{n_0} \in \omega .\) But by Proposition 3.4 \(\psi\) is invertible at its fixed points which means that \(a_{n_0}\) is a regular value of \(\psi\) and

$$\begin{aligned} \#\{\psi ^{-1}(a_{n_0})\}=m \ge 2 \end{aligned}$$

which is a contradiction! Hence the claim.

Now by similar arguments as in the case of proper maps it follows that \(\psi\) is a biholomorphism from \(\omega\) to B(0, R) and p is a repelling fixed point of \(\psi\) and hence lies in \(J(\psi ) \subset J(G).\) Since \(\omega \cap J(G)=\{0\}\), we have \(p=0\) which is an isolated point in the Julia set of \(\psi\) and hence \(\psi\) is conjugate to an upper triangular automorphism. \(\square\)

4 Examples of semigroups and their Julia sets

Example 4.1

Consider the following lower triangular maps in \(\mathbb {C}^2\):

$$\begin{aligned} F_1(z,w)=(\lambda z, tw+p(z)), \ F_2(z,w)=(\mu z, sw+q(z)) \end{aligned}$$

where p and q are polynomials of degree d fixing the origin and \(|\lambda |, |\mu |,|s|, |t |>\theta >1\). Let \(G=\langle F_1,F_2\rangle\).

Note that for any sequence \(\{f_n\}\subseteq G\) and \((z,w)\ne 0\), \(|f_n(z,w)|\rightarrow \infty\) as \(n\rightarrow \infty\). It also can be checked that

$$\begin{aligned} \{(z,w)\in \mathbb {C}^2{:}\; z\ne 0\}\subseteq F(G) \quad \text { and } \quad 0\in J(G). \end{aligned}$$

Claim \(J(G)=\{0\}\).

If not, then there exists a point in J(G) apart from the origin and it must be of the form \((0,w_0)\) with \(w_0\ne 0\). Therefore, there exists a sequence \(\{(z_n,w_n)\}\) converging to \((0,w_0)\), \(\{f_n\}\subseteq G\) and \(M\ge 1\) such that

$$\begin{aligned} |f_n(z_n,w_n)|\le M, \text { i.e., } (z_n,w_n)\in f_n^{-1}(B(0;M)) \end{aligned}$$
(4.1)

for all \(n\ge 1\).

Let \(\tilde{G}=\langle F_1^{-1}, F_2^{-1} \rangle\) be the semigroup generated by \(F_1^{-1}\) and \(F_2^{-1}.\) Then \(\tilde{G}\) can be realized as:

$$\begin{aligned} \tilde{G}= \bigcup _{k=1}^{\infty } G_k \end{aligned}$$

where \(G_k \subseteq \tilde{G}\) is of the following form:

$$\begin{aligned} G_k=\{h_k \circ h_{k-1} \circ \cdots \circ h_1 {:}\; h_i=F_1^{-1} \text { or }F_2^{-1} \text { for every } 1 \le i \le k \}. \end{aligned}$$

Without loss generality we assume that \(f_n^{-1}\in G_n\) for all \(n\ge 1\).

Now note that \(F_1^{-1}\) and \(F_2^{-1}\) are lower triangular polynomial maps of the form

$$\begin{aligned} F_1^{-1}(z,w)=(\lambda ^{-1} z, t^{-1} w+\tilde{p}(z)), \;\; F_2^{-1}(z,w)=(\mu ^{-1} z, s^{-1} w+ \tilde{q}(z)) \end{aligned}$$

where \(\tilde{p}\) and \(\tilde{q}\) are polynomials of degree d preserving the origin. Let

$$\begin{aligned} \tilde{p}(z)=\sum _{i=1}^d C_iz^i \quad \text { and } \quad \tilde{q}(z)=\sum _{i=1}^d D_i z^i. \end{aligned}$$

Then choose C such that

$$\begin{aligned} C> \max _{1 \le i \le d}\{|C_i|,|D_i|\}. \end{aligned}$$

Induction statement: For every \((z,w) \in B((0,0);M)\) and for each \(h \in G_k\), \(k \ge 1\)

$$\begin{aligned} |\pi _1 \circ h(z,w)| \le \theta ^{-k}M \quad \text { and } \quad |\pi _2 \circ h(z,w)| \le \theta ^{-k}M + kC \theta ^{-(k-1)} M^d d . \end{aligned}$$
(4.2)

Clearly when \(k=1\) and \((z,w) \in B((0,0);M)\),

$$\begin{aligned} |\tilde{p}(z)|\le \sum _{i=1}^d|C_i| M^i< C M^d d. \end{aligned}$$

Similarly \(|\tilde{q}(z)|<C M^d d.\) Thus for \(h \in G_1\) as \(|\lambda |^{-1},|\mu |^{-1},|s|^{-1},|t|^{-1}< \theta ^{-1}<1\)

$$\begin{aligned} |\pi _1 \circ h(z,w)| \le \theta ^{-1}M \quad \text { and } \quad |\pi _2 \circ h(z,w)| \le \theta ^{-1}M + C M^d d. \end{aligned}$$

Hence the induction statement is true for \(k=1.\) Now assuming it to be true for some k we will show that it is true for \(k+1.\)

Let \(h \in G_{k+1}\) then \(h=F_1^{-1} \circ \tilde{h}\) or \(h = F_2^{-1}\circ \tilde{h}\) where \(\tilde{h} \in G_k.\) So we have

$$\begin{aligned} |\pi _1 \circ \tilde{h}(z,w)| \le \theta ^{-k}M \quad \text { and } \quad |\pi _2 \circ \tilde{h}(z,w)| \le \theta ^{-k}M + kC \theta ^{-(k-1)} M^d d. \end{aligned}$$

Assume that \(h=F_1^{-1}\circ \tilde{h}\) then

$$\begin{aligned} \pi _1 \circ h(z,w)&= \lambda ^{-1}\big (\pi _1 \circ \tilde{h}(z,w) \big )\nonumber \\ \pi _2 \circ h(z,w)&= t^{-1}\big (\pi _2 \circ \tilde{h}(z,w)\big )+ \tilde{p}\circ \pi _1 \circ \tilde{h}(z,w). \end{aligned}$$
(4.3)

Then clearly from the above observation if \((z,w) \in B((0,0);M)\) then

$$\begin{aligned} |\pi _1 \circ h(z,w)| \le \theta ^{-k-1}M. \end{aligned}$$

Since \(\theta ^{-1} <1\) and \(M>1\)

$$\begin{aligned} |\tilde{p}\circ \pi _1 \circ \tilde{h}(z,w)| \le \sum _{i=1}^d |C_i| (\theta ^{-k}M)^i \le C \theta ^{-k} M^d d. \end{aligned}$$

Now substituting this estimate on equation (4.3) we have

$$\begin{aligned} |\pi _2 \circ h(z,w)|&\le |\theta ^{-1}\big (\pi _2 \circ \tilde{h}(z,w)\big )|+ |\tilde{p}\circ \pi _1 \circ \tilde{h}(z,w)|\\&\le \theta ^{-k-1}M+kC\theta ^{-k} M^d d+C \theta ^{-k} M^d d \\&\le \theta ^{-k-1}M+(k+1)C\theta ^{-k} M^d d . \end{aligned}$$

Similarly if \(h=F_2^{-1} \circ \tilde{h}.\) Hence, the induction statement is true.

Now since \(f_k^{-1} \in G_k\), it follows from the induction statement (4.2) that for every \((z,w) \in B(0;M)\)

$$\begin{aligned} |\pi _1 \circ f_k^{-1}(z,w)| \le \theta ^{-k}M \text \quad { and } \quad |\pi _2 \circ f_k^{-1}(z,w)| \le \theta ^{-k}M + kC \theta ^{-(k-1)} M^d d . \end{aligned}$$

This implies that \((z_k,w_k)\rightarrow 0\) as \(k\rightarrow \infty\). Contradiction! Hence the claim follows.

Remark 4.2

Let \(G=\langle F_1, F_2,\cdots F_n\rangle\) for some \(n\ge 1\) where each \(F_i\) is a lower triangular polynomial map in \(\mathbb C^k\), \(k \ge 2\) having a repelling fixed point at the origin. Then using a similar set of arguments as above, it can be proved that \(J(G)=\{0\}\).

Remark 4.3

A large class of elementary polynomial automorphisms in the Friedland–Milnor classification ([6]) comprises of lower triangular polynomial automorphisms fixing the origin. Thus for a semigroup G which is finitely generated by such elementary maps, we get \(J(G)=\{0\}\).

Example 4.4

Let \(f_c\) denote the automorphism of \(\mathbb C^2\) of the form

$$\begin{aligned} f_c(z,w)=\big (z e^{ch(zw)}, we^{-ch(zw)}\big ) \end{aligned}$$

where \(c \in \mathbb C\) and h be a non-constant entire function on \(\mathbb {C}\). The Jacobian of \(f_c\) for every \(c\in \mathbb C\) is constant, i.e., \(Jf_c \equiv 1\) on \(\mathbb C^2.\) Consider the semigroup G:

$$\begin{aligned} G=\langle f_c{:}\; 1<c<\infty \rangle . \end{aligned}$$

Observe that

$$\begin{aligned} f_{c_2} \circ f_{c_1}(z,w)=\big (z e^{(c_1+c_2) h(zw)}, w e^{-(c_1+c_2) h(zw)}\big )=f_{c_1+c_2}(z,w). \end{aligned}$$

Hence, corresponding to any element \(f \in G\), there exists \(c_f>1\) such that

$$\begin{aligned} f(z,w)=\big (z e^{c_f h(zw)},we^{-c_f h(zw)}\big ). \end{aligned}$$

Since \(Jf \equiv 1\) for every \(f \in G\), no point is a repelling fixed point for any element of G. The proof of Theorem 3.2 shows that if the Julia set J(G) has an isolated point it should be a repelling fixed point for some element of G which is clearly not the case here. Thus, the Julia set J(G) should be perfect.

Claim If \(\text {Re }h(0)<0\), then the Julia set J(G) is exactly the following perfect set:

$$\begin{aligned} \{(z,w) \in \mathbb C^2{:}\; \text {Re }h(zw)=0 \} \cup \{(z,w) \in \mathbb C^2{:}\; w=0\}. \end{aligned}$$

Consider \(\{f_n\} \subset G\). Then each \(f_n\) can be thought of as

$$\begin{aligned} f_n(z,w)=\big (z e^{c_n h(zw)},we^{-c_n h(zw)}\big ). \end{aligned}$$

If there exists a subsequence \(c_{n_k}\rightarrow c \in \mathbb R^+\) then \(f_{n_k} \rightarrow f_c\) on compact subsets, otherwise \(c_n \rightarrow \infty\) as \(n\rightarrow \infty\).

Case 1 If \(\text {Re }h(zw)=0\) then \(\{f_{n}\}\) does not diverge to infinity as,

$$\begin{aligned} \Vert f_{n}(z,w) \Vert =\Vert (z,w)\Vert . \end{aligned}$$

But \(\pi _1(f_{n}(z,w))\) or \(\pi _2(f_{n}(z,w))\) diverges to infinity uniformly on a small enough neighbourhood of such a point, depending on whether \(\text {Re }h(zw)>0\) or \(\text {Re }h(zw)<0,\) respectively.

Case 2 If \(w=0\), then

$$\begin{aligned} \Vert f_{n}(z,0)\Vert =|z|e^{c_{n}\alpha }, \end{aligned}$$

i.e., \(\Vert f_{n}(z,0)\Vert \rightarrow 0\) as \(n \rightarrow \infty\) since \(\alpha =\text {Re }h(0)<0\). Now for every sufficiently small neighbourhood \(B_z\) around any (z, 0) there exists \((z,w')\in B_z\) such that \(w'\ne 0\) and \(\text {Re }h(zw')<0\). Therefore, \(\pi _2(f_{n}(z,w'))\) diverges to infinity as \(n\rightarrow \infty\).

Hence the claim follows.

Recall Examples 2.8 and 2.9. In each case G is a semigroup generated by maps of maximal generic rank in \(\mathbb C^2.\) So by Theorem 3.5 they should be perfect since none of the elements in the semigroup is conjugated to an upper triangular automorphism of \(\mathbb C^2\), which is exactly the case.

Example 4.5

Let \(f_1\) and \(f_2\) be the following maps in \(\mathbb C^2\) of maximal generic rank:

$$\begin{aligned} F_1(z,w)=(z,w^2), \ F_2(z,w)=(z,wz). \end{aligned}$$

Let G be the semigroup generated by them, i.e.,

$$\begin{aligned} G=\langle F_1,F_2\rangle . \end{aligned}$$

Then by Theorem 3.5, the Julia set for the semigroup G, i.e., J(G) should be perfect.

Claim The Julia set J(G) for the semigroup G is: (this is illustrated in Fig. 1)

$$\begin{aligned} \tilde{J}=\{(z,w) \in \mathbb C^2{:}\; 0 \le |z|\le 1, |w|\ge 1\} \cup \{(z,w) \in \mathbb C^2{:}\; 0 \le |w|\le 1, |z| \ge 1\}. \end{aligned}$$
Fig. 1
figure 1

The Julia set J(G)

Suppose \(\{f_n\}\) is a sequence from G. Then, there exist sequences of positive integers \(\{a_n\}\) and \(\{b_n\}\) such that

$$\begin{aligned} f_n(z,w)=(z,z^{a_n}w^{b_n}) \end{aligned}$$

and at least one of them is unbounded. Let \(\Delta ^2(0;1)\) denote the unit polydisc. Then we prove that the Fatou set

$$\begin{aligned} F(G) \supseteq \Delta ^2(0;1) \cup \{ (z,w) \in \mathbb C^2{:}\; |z|>1 \text{ and } |w|>1 \}. \end{aligned}$$

Case 1 In \(\Delta ^2(0;1)\)

$$\begin{aligned} \Vert {f_n}_{|\Delta ^2(0;1)}\Vert _{\infty }< 1. \end{aligned}$$

Thus G is locally uniformly bounded on \(\Delta ^2(0;1)\), and hence there exists a subsequence which converges uniformly on its compact subsets. So \(\Delta ^2(0;1) \subset F(G).\)

Case 2 Suppose \(|z|>1\) and \(|w|>1.\)

Then without loss of generality one can assume that there exists a subsequence \(\{a_{n_k}\}\) of \(\{a_n\}\) which diverges to \(\infty\) as \(k \rightarrow \infty .\) Thus

$$\begin{aligned} \Vert f_{n_k}(z,w)\Vert _{\infty } > |z|^{a_{n_k}} \rightarrow \infty , \end{aligned}$$

i.e., \(\{f_{n_k}\}\) diverges to \(\infty\) uniformly in a small enough neighbourhood of such a (zw). hence \((z,w) \in F(G).\)

Consider the set A defined as:

$$\begin{aligned} A=\big \{(z,w) \in \mathbb C^2{:}\; |z^{\frac{p}{2^q}} w|=1 \text { for some integers } p\ge 1 \text { and } q \ge 0\big \}. \end{aligned}$$

Since the set

$$\begin{aligned} \Big \{ \frac{p}{2^q}{:}\; p, q \text { integers with } p \ge 1 \text { and } q \ge 0 \Big \} \end{aligned}$$

is dense in the positive real axis, the set A is dense in \(\tilde{J}\) and \(\bar{A}=\tilde{J}.\) Also the Julia set of a semigroup is closed, so to prove the claim it is enough to prove that A is contained in J(G).

Now pick \((z_0,w_0) \in A\). Then \(|z_0^{p}w_0^{2^{q}}|=1\) for some \(p \ge 1\) and \(q \ge 0.\) The sequence

$$\begin{aligned} f_n(z,w)&=F_2^{p(2^{qn-q})} \circ F_1^{qn}(z,w)\\&=\big (z,wz^{p(2^{qn-q})}\big ) \circ \big (z,w^{2^{qn}}\big ) =\Big (z, \{z^p w^{2^q}\}^{2^{q(n-1)}}\Big ) \end{aligned}$$

for \(n \ge 1\). On every neighbourhood of \((z_0,w_0)\), there exists (zw) such that \(|z^pw^{2^q}|>1\) as well as (zw) such that \(|z^pw^{2^q}|<1.\) Thus, \((z_0,w_0)\) is contained in the Julia set and this completes the proof.

5 Recurrent and Wandering Fatou components of a semigroup G

As discussed in Section 1, we will be studying the properties of recurrent and wandering Fatou components of semigroup generated by entire maps of maximal generic rank on \(\mathbb {C}^k\). The wandering and the recurrent Fatou components for a semigroup G are defined as:

Definition 5.1

Let \(G=\langle \phi _1,\phi _2,\ldots \rangle\) where each \(\phi _i\in \mathcal {E}_k\). Given a Fatou component \(\Omega\) of G and \(\phi \in G\), let \(\Omega _{\phi }\) be the Fatou component of G containing \(\phi (\Omega \ \Sigma _\phi )\) where \(\Sigma _{\phi }\) is the set where the Jacobian of \(\phi\) vanishes. A Fatou component is wandering if the set \(\big \{ \Omega _{\phi }{:}\;\phi \in G\big \}\) contains infinitely many distinct elements.

Definition 5.2

Let \(G=\langle \phi _1,\phi _2,\ldots \rangle\) where each \(\phi _i\in \mathcal {E}_k\). A Fatou component \(\Omega\) of G is recurrent if for any sequence \(\{g_j\}_{j\ge 1}\subset G\), there exists a subsequence \(\{g_{j_m}\}\) and a point \(p\in \Omega\) (the point p depends on the chosen sequence) such that \(g_{j_m}(p)\rightarrow p_0 \in \Omega\).

Note that we assume here a stronger definition of recurrence than the existing definition for the case of iterations of a single holomorphic endomorphism of \(\mathbb C^k.\) The natural extension of this definition to the semigroup set up would have been the following, a Fatou component \(\Omega\) is recurrent if there is a point \(p \in \Omega\) and a sequence \(\phi _n \in \Omega\) such that \(\phi _n(p) \rightarrow p_0,\) where \(p_0 \in \Omega .\) If this definition of recurrence is adopted then it is possible that a Recurrent domain is Wandering. In particular, Theorem 5.3 in [7] gives an example of a polynomial semigroup \(G=\langle \phi _1,\phi _2, \ldots \rangle\) in \(\mathbb C\), such that there exists a Fatou component, (say \(\mathcal {B}\), which is conformally equivalent to a disc), that is wandering, but returns to the same component infinitely often. This means that there exists sequences say \(\phi _n^+ \in G\) and \(\phi _n^- \in G\) such that \(\phi _n^-(\mathcal {B}) \subset \mathcal {B}\) or \(\phi _n^ +(\mathcal {B})\) are contained in distinct Fatou components of G. This example can be easily adapted in higher dimensions.

Example 5.3

Consider the semigroup \(\mathcal {G}=\langle \Phi _1, \Phi _2, \ldots , \rangle\) generated by the maps

$$\begin{aligned} \Phi _i(z,w)=\left( \phi _i(z), w^2\right) \end{aligned}$$

where \(\phi _i\) are the polynomial maps as in Theorem 5.3 of [7]. Let \(\{\Phi _n^-\}_{n \ge 1} \subset G\) be the sequence that maps \(\mathcal {B} \times \mathbb {D}\) into itself and \(\{\Phi _n^+\}_{n \ge 1} \subset G\) be the sequence such that

$$\begin{aligned} \Phi ^+_i(\mathcal {B}\times \mathbb D) \cap \Phi ^+_j( \mathcal {B}\times \mathbb D) = \emptyset \end{aligned}$$

for every \(i \ne j.\) Also \(\mathcal {B}\times \mathbb D\) is a Fatou component of \(\mathcal {G}\) as any point on the boundary of \(\mathcal {B}\times \mathbb D\), is either in the Julia set of G or in the Julia set of the map \(z \rightarrow z^2\). Hence \(\mathcal {B}\times \mathbb D\) is a Fatou component which is wandering, but may be recurring as well if we adapt the classical definition of recurrence.

Hence, we work with a stronger definition of recurrence than the classical one. Next, we provide an alternative description for recurrent Fatou components of G.

Lemma 5.4

A Fatou component \(\Omega\) is recurrent if and only if for any sequence \(\{\phi _j\}\subset G\) , there exists a compact set \(K\subset \Omega\) and a subsequence \(\{\phi _{j_m}\}\) such that \(\phi _{j_m}(p_{j_m})\rightarrow p_0 \in \Omega\) for a sequence \(\{p_{j_m}\}\subset K\).

Proof

Take any sequence \(\{\phi _j\}\subset G\). Then, there exists a subsequence \(\{\phi _{j_m}\}\) and points \(\{p_{j_m}\}\subset K\) with K compact in \(\Omega\) such that

$$\begin{aligned} \phi _{j_m}(p_{j_m})\rightarrow p_0 \in \Omega . \end{aligned}$$

Without loss of generality we assume \(p_{j_m}\rightarrow q_0 \in K\). It follows that \(\phi _{j_m}(q_0)\rightarrow p_0 \in \Omega\) using the fact that any sequence of G is normal on the Fatou set of G. \(\square\)

Proposition 5.5

Let \(G=\langle \phi _1,\phi _2,\ldots ,\phi _m\rangle\) where each \(\phi _i \in \mathcal {E}_k\) for every \(1 \le i \le m\) . If \(\Omega\) is a recurrent Fatou component of G, then G is locally bounded on \(\Omega\) . Moreover, \(\Omega\) is pseudoconvex and Runge.

Proof

Assume G is not locally bounded on \(\Omega\). Then, there exists a compact set \(K\subset \Omega\) and \(\{g_r\} \subseteq G\) such that \(|g_r(z_r) |> r\) with \(z_r \in K\) for every \(r \ge 1\). Clearly, this cannot be the case since \(\Omega\) is a recurrent Fatou component, so we can always get a subsequence \(\{g_{r_k}\}\) from the sequence \(\{g_r\} \in G\) such that it converges to a holomorphic function uniformly on compact set in \(\Omega\) and in particular on K. From the proof of Proposition 2.2, it follows that local boundedness of G on \(\Omega\) implies that \(\Omega\) is polynomially convex. Hence \(\Omega\) is pseudoconvex.

Theorem 5.6

Let \(G= \langle \phi _1,\phi _2, \ldots \rangle\) where each \(\phi _i \in \mathcal {E}_k.\) Assume that \(\Omega\) is a recurrent Fatou component of G. If there exists a \(\phi \in G\) such that \(\phi (\Omega )\) is contained in the Fatou set of G,  i.e., \(\phi (\Omega ) \subset F(G)\) then one of the following is true

  1. (i)

    There exists an attracting fixed point (say \(p_0\) ) in \(\Omega\) for the map \(\phi .\)

  2. (ii)

    There exists a closed connected submanifold \(M_\phi \subset \Omega\) of dimension \(r_\phi\) with \(1 \le r_\phi \le k-1\) and an integer \(l_\phi >0\) such that

    1. (a)

      \(\phi ^{l_\phi }\) is an automorphism of \(M_{\phi }\) and \(\overline{\{\phi ^{nl_\phi }\}_{n \ge 1}}\) is a compact subgroup of \(\mathrm{Aut}(M_\phi ).\)

    2. (b)

      If \(f \in \overline{\{\phi ^n\}}\) , then f has maximal generic rank \(r_\phi\) in \(\Omega .\)

  3. (iii)

    \(\phi\) is an automorphism of \(\Omega\) and \(\overline{\{\phi ^n\}}\) is a compact subgroup of \(\mathrm{Aut}(\Omega ).\)

Proof

Since \(\Omega \subset F(G)\), there exists a recurrent Fatou component of the map \(\phi\) (say \(\Omega _{\phi }\)) such that \(\Omega \subset \Omega _{\phi },\) i.e., there exists an integer \(l\ge 1\) such that

$$\begin{aligned} \phi ^l(\Omega _\phi ) \cap \Omega _{\phi }\ne \emptyset \;\; \quad \text {and} \quad \;\; \phi ^{m}(\Omega _\phi ) \cap \Omega _{\phi }= \emptyset \end{aligned}$$

for \(0 \le m < l.\) So, if \(l>1\) then there do not exist any \(p \in \Omega\) such that any subsequence of \(\{\phi ^{lk+1}(p)\}_{k\ge 1}\) converges to a point in \(\Omega\). Hence \(l=1\) and by assumption it follows that \(\phi (\Omega ) \subset \Omega .\)

Let h be a limit function of \(\{\phi ^n\}\) of maximal rank (say \(r_\phi\)), i.e.,

$$\begin{aligned} h(p)=\lim _{j \rightarrow \infty } \phi ^{n_j}(p) \;\; \text {for every}\;\; p \in \Omega , \end{aligned}$$

where \(\{n_j\}\) is an increasing subsequence of natural numbers.

Case 1 If \(r_{\phi }=0.\) Then \(h(\Omega )=p_0\) for some \(p_0 \in \Omega\) since by recurrence there exists a point \(p \in \Omega\), such that \(\phi ^{n_j}(p) \rightarrow p_0\) and \(p_0 \in \Omega .\) Also \(h(p_0)=p_0.\) Then

$$\begin{aligned} \phi (p_0)= \phi (h(p_0))=h(\phi (p_0))=p_0, \end{aligned}$$

i.e., \(p_0\) is a fixed point of \(\phi .\) As some sequence of iterates of \(\phi\) converge to a constant function, \(p_0\) is an attracting fixed point for \(\phi .\)

Case 2 If \(r_{\phi } \ge 1.\) Then there exists an increasing subsequence \(\{m_j\}\) such that

$$\begin{aligned} p_j=m_{j+1}-m_j \end{aligned}$$

are increasing positive integers and the sequences \(\{\phi ^{m_j}\}\) and \(\{\phi ^{p_j}\}\) converge uniformly to the limit functions h and \(\tilde{h}\) respectively on the Fatou component \(\Omega .\) Since by recurrence \(h(\Omega ) \cap \Omega \ne \emptyset\), if \(p \in \Omega\) be such that \(p =h(q)\) for some \(q \in \Omega\) then

$$\begin{aligned} \tilde{h}(p)= \lim _{j \rightarrow \infty } \phi ^{m_{j+1}-m_j}(p)=\lim _{j \rightarrow \infty }\phi ^{m_{j+1}-m_j}\left( \phi ^{m_j}(q)\right) =p \end{aligned}$$

Define

$$\begin{aligned} M=\{x \in \Omega {:}\; \tilde{h}(x)=x \}. \end{aligned}$$

Claim M is a closed complex submanifold of \(\Omega .\)

Since \(h(\Omega ) \cap \Omega \subset M\), M is a variety of dimension \(\ge r_\phi\). But by the choice of h, the generic rank of \(\tilde{h} \le r_{\phi }\) and \(M \subset \tilde{h}(\Omega ) \cap \Omega .\) So the dimension of M is \(r_{\phi }.\) Now for any point in M,  the rank of the derivative matrix of \(\mathrm{Id}-\tilde{h}\) is greater than or equal to \(k-r_\phi\). Suppose for some \(x \in M\) the rank of \(D(\mathrm{Id}-\tilde{h})(x)>k-r_\phi ,\) then there exists a small neighbourhood of x, say \(V_x\) such that \(V_x \subset \Omega\) and

$$\begin{aligned} \text {rank of }\mathrm{Id}-\tilde{h}>k-r_\phi \;\; \text {for every }\;\; x \in V_x. \end{aligned}$$

Then \(\{\mathrm{Id}-\tilde{h}\}^{-1}(0) \cap V_x\) is a variety of dimension at most \(r_\phi -1,\) i.e., the dimension of M is strictly less than \(r_\phi ,\) which is a contradiction. Thus, the rank of \(\mathrm{Id}-\tilde{h}\) is \(k-r_\phi\) for every point in M and hence M is a closed submanifold of \(\Omega .\)

Step 1: Suppose that \(r_{\phi }=k.\)

Then clearly \(M=\Omega\) and \(\tilde{h}\) on \(\Omega\) is the identity map. Let \(h_2= \lim \phi ^{p_j-1}.\) Then

$$\begin{aligned} \tilde{h}(x)=h_2 \circ \phi (x)=x, \; \; \text {for every}\;\; x \in \Omega , \end{aligned}$$

i.e., \(\phi\) is injective on \(\Omega\) and \(\phi (\Omega )\) is an open subset of \(\Omega .\) Suppose there exists an \(x \in \Omega \ \phi (\Omega )\) then for a sufficiently small ball of radius \(r>0\) with \(B_r(x) \subset \Omega\)

$$\begin{aligned} \phi ^l(\Omega ) \cap B_r(x)=\emptyset \;\;\text {for every}\;\; l \ge 1. \end{aligned}$$

This contradicts that \(\phi ^{p_j} (x)\rightarrow x.\) Hence \(\phi\) is surjective on \(\Omega\) and hence an automorphism of \(\Omega .\)

Step 2: Suppose that \(1 \le r_{\phi } \le k-1.\) Let \(M_\phi\) denote an irreducible component of M. For every \(q \in M_{\phi }\), it follows that \(\phi ^{p_j}(q) \rightarrow q\) as \(j \rightarrow \infty .\) Since \(\phi (\Omega ) \subset \Omega\), we get \(\phi ^n(q) \in \Omega\) for every \(n \ge 1\) and

$$\begin{aligned} \tilde{h} \circ \phi ^n(q)=\phi ^n \circ \tilde{h}(q)=\phi ^n(q) \; \text {for every}\; q \in M_{\phi }, \end{aligned}$$

i.e., \(\phi ^n(M_\phi ) \subset M\) for every \(n \ge 1.\)

Claim There exists a positive integer \(l_\phi\) such that \(\phi ^{l_{\phi }}(M_{\phi }) \subset M_{\phi } .\)

Let \(p_0 \in M_\phi\) and \(\Delta \subset \Omega\) be a polydisk at \(p_0\) such that \(\Delta\) does not intersect the other components of \(M_{\phi } .\) Now choose \(\Delta ' \subset \Delta\), a sufficiently small polydisk such that \(\tilde{h}(\Delta ') \subset \Delta .\) Then \(\omega =\tilde{h}(\Delta ') \subset M_\phi\) is a \(r_\phi\)-dimensional manifold. Let \(\Delta ''\) be a \(r_\phi\)-dimensional polydisk inside \(\omega\) and \(\{w_l\}_{l \ge 1}\) be a sequence in \(\Delta ''\) such that it converges to some \(w_0 \in \Delta ''.\) But \(\phi ^{p_j}(w_{p_j}) \rightarrow w_0\) as \(j \rightarrow \infty\) hence

$$\begin{aligned} \phi ^{p_j}(M_\phi ) \cap \Delta \ne \emptyset ,\; \text {i.e.,}\; \phi ^{p_j} (M_\phi )\subset (M_\phi ) \end{aligned}$$

for j sufficiently large. Let \(l_\phi\) be the minimum value such that \(M_\phi\) is invariant under \(\phi ^{l_{\phi }}.\)

Claim \(\phi ^{l_\phi }\) is an automorphism of \(M_\phi .\)

Without loss of generality there exists a sequence \(\{k_j\}\) such that \(p_j=i_0+k_jl_\phi\) for some \(0 \le i_0 \le l_\phi -1,\) i.e.,

$$\begin{aligned} \phi ^{i_0} \circ \phi ^{k_jl_\phi }(x) \rightarrow x \;\; \text {for every} \;\; x \in M_{\phi }. \end{aligned}$$

As \(M_{\phi }\) is invariant under \(\phi ^{l_\phi }\), the sequence \(x_j=\phi ^{k_jl_\phi }(x)\) lies in \(M_{\phi }.\) Again as before let \(\Delta _x\) be a sufficiently small neighbourhood such that \(\Delta _x \subset \Omega\) and \(\Delta _x\) does not intersect the other components of M. Since \(\phi ^{i_0}(x_j) \in \Delta _x \cap M_\phi\) for large j, \(\phi ^{i_0}(M_\phi ) \subset M_\phi .\) But \(0 \le i_0 \le l_\phi -1\), i.e., \(i_0=0\) and \(\{\phi ^{k_jl_\phi }\}\) converges uniformly to the identity on \(M_{\phi }.\) Let \(\psi =\lim \phi ^{(k_j-1)l_\phi }\) then

$$\begin{aligned} \phi ^{l_\phi } \circ \psi (x)=\psi \circ \phi ^{l_\phi }(x)=x \;\; \quad \text {for every}\;\; x \in M_\phi . \end{aligned}$$

Hence \(\phi ^{l_\phi }\) is injective on \(M_\phi\) and \(\phi ^{l_\phi }(M_\phi )\) is an open subset in the manifold \(M_\phi\). Now as in Step 1 observe that \(\phi ^{k_jl_\phi }\) converges to the identity on \(M_\phi\) for an unbounded sequence \(\{k_j\}\), so \(\phi ^{l_\phi }\) is also surjective on \(M_{\phi }\). Thus the claim.

Let \(Y=\{\phi ^{nl_\phi }\}_{n \ge 1} \subset \mathrm{Aut}(M_\phi ).\)

Claim \(\bar{Y}\) is a locally compact subgroup of \(\mathrm{Aut}(M_\phi ).\)

For some \(\Psi \in Y\) and for a compact set \(K \subset M_\phi\) consider the neighbourhood of \(\Psi\) given by

$$\begin{aligned} V_\Psi (K, \epsilon )= \{\psi \in \mathrm{Aut}(M_\phi ){:}\; \Vert \psi (z)-\Psi (z)\Vert _{K} < \epsilon \}. \end{aligned}$$

One can choose \(\epsilon\) and K sufficiently small such that for every sequence \(\psi _j \in V_\Psi (K, \epsilon )\) there exists an open set \(U \subset \Omega\) such that \(\psi _j(U \cap M_\phi ) \subset \bar{V} \cap M_\phi \subset \Omega\), where V is some open subset of \(\Omega .\)

Since \(\psi _j=\phi ^{n_j l_\phi }\) for a sequence \(\{n_k\}\) and \(\Omega\) is a Fatou component, \(\psi _j\) has a convergent subsequence in \(\Omega .\) We choose appropriate subsequences such that the limit maps

$$\begin{aligned} \Psi _1=\lim _{j \rightarrow \infty } \phi ^{n_jl_{\phi }} \;\; \quad \text {and} \quad \;\; \Psi _2=\lim _{j \rightarrow \infty } \phi ^{(k_j-n_j)l_\phi } \end{aligned}$$

are defined on \(\Omega .\) Also as \(M_\phi\) is closed in \(\Omega\), \(\Psi _i(M_\phi ) \subset \overline{M_\phi }\) for every \(i=1,2\) where \(\overline{M_\phi }\) denote the closure of \(M_\phi\) in \(\mathbb C^k.\) Then \(\Psi _1(U) \subset \Omega\) and

$$\begin{aligned} \Psi _2 \circ \Psi _1(x)=x \;\; \quad \text {for every}\;\; x \in U \cap M_\phi . \end{aligned}$$
(5.1)

Since \(\Psi _1\) on \(M_\phi\) is a limit of automorphisms of \(M_\phi\), the Jacobian of \(\Psi _1\) on the manifold \(M_\phi\) is either non-zero at every point of \(M_\phi\) or vanishes identically. But by (5.1), \(\Psi _1\) restricted to \(U \cap M_{\phi }\) is injective, which is open in the manifold \(M_\phi ,\) i.e., \(\Psi _1\) is an open map of \(M_\phi\) and \(\Psi _1(M_\phi ) \subset M_{\phi }.\) So (5.1) is true for every \(x \in M_\phi .\) Now by the same arguments it follows that \(\Psi _2\) is an injective map from \(M_\phi\) such that \(\Psi _2(M_\phi ) \subset M_{\phi }.\) Hence

$$\begin{aligned} \Psi _2 \circ \Psi _1(x)=\Psi _1 \circ \Psi _2(x)=x \;\; \quad \text {for every}\;\; x \in M_\phi , \end{aligned}$$

i.e., \(\Psi _1\) is an automorphism of \(M_\phi .\) This proves that \(\bar{Y}\) is a locally compact subgroup of \(\mathrm{Aut(M_\phi )}.\)

Now since \(M_\phi\) is a complex manifold and \(\bar{Y}\) is a locally abelian subgroup of automorphisms of \(M_\phi\), by Theorem A in [2], it follows that \(\bar{Y}\) is a Lie group. Hence the component of \(\bar{Y}\) containing the identity is isomorphic to \(\mathbb T^l \times \mathbb R^m.\) Suppose \(\Psi\) is the isomorphism, then for some \(n >0\), \(\Psi (a,b)=\phi ^{nl_\phi }\). Now if \(b \ne 0\), then there does not exist an increasing sequence of \(k_j\) such that \(\phi ^{k_j l_\phi }\) converges to identity. This proves that the component of \(\bar{Y}\) containing the identity is compact and hence any component of \(\bar{Y}\) is compact by the same arguments. Also as \(M_\phi\) is contained in the Fatou set, the number of components of \(\bar{Y}\) is finite, thus \(\bar{Y}\) is a compact subgroup of \(\mathrm{Aut}(M_\phi ).\)

If \(r_\phi =k\), then \(M_\phi\) is \(\Omega\), then one can apply the same technique as discussed above to conclude that \(\overline{\{\phi ^n\}}\) is a closed compact subgroup of \(\mathrm{Aut}(\Omega ).\)

Finally, let f be a limit of \(\{\phi ^n\}_{n \ge 1},\) i.e.,

$$\begin{aligned} f(p)=\lim _{j \rightarrow \infty } \phi ^{n_j}(p) \;\; \quad \text {for every} \;\; p \in \Omega . \end{aligned}$$

Claim The generic rank of f is \(r_{\phi }.\)

By the definition of recurrence it follows that \(\Omega \subset \Omega _{\phi }\), where \(\Omega _\phi\) is a periodic Fatou component for \(\phi\) with period 1. Hence by Theorem 3.3 in [5] it follows that the limit maps of the set \(\{\phi ^n\}\) in \(\Omega _\phi\) have the same generic rank (say r). But \(\Omega\) is an open subset of the Fatou component \(\Omega _\phi\), so the rank of limit maps restricted to \(\Omega\) should be same, i.e., \(r=r_\phi\) and each limit map of \(\{\phi ^n\}\) has rank \(r_\phi .\) \(\square\)

By Proposition 5.5 a semigroup G is always locally uniformly bounded on a recurrent Fatou component semigroup G. If G is finitely generated by holomorphic endomorphisms of maximal rank k in \(\mathbb C^k\), then by Proposition 2.2 it follows that a recurrent Fatou component is mapped in the Fatou set by any element of G. Hence we have the following corollary.

Corollary 5.7

Let \(G= \langle \phi _1,\phi _2, \ldots ,\phi _m \rangle\) where each \(\phi _i \in \mathcal {E}_k\) for every \(1 \le i \le m.\) Assume that \(\Omega\) is a recurrent Fatou component of G then for every \(\phi \in G\) one of the following is true

  1. (i)

    There exists an attracting fixed point (say \(p_0\) ) in \(\Omega\) for the map \(\phi .\)

  2. (ii)

    There exists a closed connected submanifold \(M_\phi \subset \Omega\) of dimension \(r_\phi\) with \(1 \le r_\phi \le k-1\) and an integer \(l_\phi >0\) such that

    1. (a)

      \(\phi ^{l_\phi }\) is an automorphism of \(M_{\phi }\) and \(\overline{\{\phi ^{nl_\phi }\}_{n \ge 1}}\) is a compact subgroup of \(\mathrm{Aut}(M_\phi ).\)

    2. (b)

      If \(f \in \overline{\{\phi ^n\}}\) , then f has maximal generic rank \(r_\phi\) in \(\Omega .\)

  3. (iii)

    \(\phi\) is an automorphism of \(\Omega\) and \(\overline{\{\phi ^n\}}\) is a compact subgroup of \(\mathrm{Aut}(\Omega ).\)

Example 5.8

Let \(G=\langle \phi _1,\phi _2 \rangle\) be a semigroup of entire maps in \(\mathbb C^2\) generated by

$$\begin{aligned} \phi _1(z,w)=(w, \alpha z-w^2) \; \; \text{ and } \;\; \phi _2(z,w)=(zw,w) \end{aligned}$$

where \(0<\alpha < 1.\) Then G is locally uniformly bounded on a sufficiently small neighbourhood around the origin, and \(\phi (0)=0\) for every \(\phi \in G.\) So the Fatou component of G containing 0 (say \(\Omega _0\)) is recurrent. Now note that for \(\phi _2\)

$$\begin{aligned} r_{{\phi _2}}=1\;\; \text{ and } \;\;M_{\phi _2}=\{(0,w){:}\; w \in \mathbb C\} \cap \Omega _0, \end{aligned}$$

whereas for \(\phi _1\) the origin is an attracting fixed point. This illustrates the different behaviour of the sequences \(\{\phi _1^n\}\) and \(\{\phi _2^n\}\) (both of which are in G) on \(\Omega _0.\)

Note that for any other \(\phi \in G\) which is not of the form \(\phi _1^k,~k \ge 2\), contains a factor of \(\phi _2\) at least once. Since for a small enough ball (say B) around origin, \(\phi _2\) is contracting, and \(\phi _1(B) \subset B\) so there exists a constant \(0<a_\phi <1\) such that

$$\begin{aligned} |\phi (z)| \le a_{\phi } |z| \;\; \quad \text {for every} \;\; z \in B, \end{aligned}$$

i.e., the origin is an attracting fixed point.

Proposition 5.9

Let \(G=\langle \phi _1,\phi _2,\ldots ,\phi _m\rangle\) where each \(\phi _i\in \mathcal {V}_k\) for every \(1 \le i \le m\) and let \(\Omega\) be an invariant Fatou component of G. Then either \(\Omega\) is recurrent or there exists a sequence \(\{\phi _n\}\subset G\) converging to infinity.

Proof

If \(\Omega\) is not recurrent, then there exists a sequence \(\{\phi _n\}\subset G\) such that \(\{\phi _n\} \rightarrow \partial \Omega \cup \{\infty \}\) uniformly on compact sets of \(\Omega\). Assume \(\{\phi _{n_k}\}\) converges to a holomorphic function f on \(\Omega\). This implies that \(f(\Omega )\subset \partial \Omega\) contradicting the assumption that each \(\phi _{n_k}\) is volume preserving. Hence, \(\{\phi _{n_k}\}\) diverges to infinity uniformly on compact subsets of \(\Omega\). \(\square\)

Proposition 5.10

Let \(G=\langle \phi _1,\phi _2,\ldots ,\phi _m \rangle\) where each \(\phi _i\in \mathcal {V}_k\) for every \(1 \le i \le m\) and let \(\Omega\) be a wandering Fatou component of G. Then, there exists a sequence \(\{\phi _n\}\subset G\) converging to infinity.

Proof

Since \(\Omega\) is wandering, one can choose a sequence \(\{\phi _n \}\subset G\) so that

$$\begin{aligned} \Omega _{\phi _n}\cap {\Omega }_{\phi _m}=\emptyset \end{aligned}$$
(5.2)

for \(n\ne m\). If this sequence \(\{\phi _n\}\) does not diverge to infinity uniformly on compact subsets, some subsequence \(\{\phi _{n_k}\}\) will converge to a holomorphic function h on \(\Omega\). By abuse of notation, we denote \(\{\phi _{n_k}\}\) still by \(\{\phi _n \}\). Fix \(z_0\in \Omega\). Then for any given \(\epsilon\), there exists \(\delta\) such that

$$\begin{aligned} \left| \phi _{n_0}(z)-\phi _n(z)\right| <\epsilon \end{aligned}$$
(5.3)

for all \(n\ge n_0\) and for all \(z\in B(z_0,\delta )\). From (5.3) it follows that vol\((\cup _{n\ge n_o}\phi _n(B(z_0,\delta )))\) is finite. On the other hand, since each \({\phi _n}\) is volume preserving and (5.2) holds, we get

$$\begin{aligned} \text{ Vol }\Big (\bigcup _{n\ge n_o}\phi _n\big (B(z_0,\delta )\big )\Big ) =+\infty . \end{aligned}$$

Hence, we have proved the existence of a sequence in G converging to infinity. \(\square\)

6 Concluding remarks

As mentioned in the introduction, the classification of recurrent Fatou components for iterations of holomorphic endomorphisms of complex projective spaces has been studied in [4] and [3]. It would be interesting to explore the same question for semigroups of holomorphic endomorphisms of complex projective spaces. The main theorem in [4] and [3] is proved under the assumption that the given recurrent Fatou component is also forward invariant. The analogue of such a condition in the case of semigroups is not clear to us since we are then dealing with a family of maps none of which is distinguishable from the other.