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

Dynamics of semigroups of Hénon maps

Sayani Bera Sayani Bera: School of Mathematical and Computational Sciences, Indian Association for the Cultivation of Science, Kolkata-700032, India sayanibera2016@gmail.com, mcssb2@iacs.res.in
Abstract.

The goal of this article is two fold. Firstly, we explore the dynamics of a semigroup of polynomial automorphisms of 2\mathbb{C}^{2}, generated by a finite collection of Hénon maps. In particular, we construct the positive and negative dynamical Green’s functions G𝒢±G_{\mathscr{G}}^{\pm} and the corresponding dynamical Green’s currents μ𝒢±\mu_{\mathscr{G}}^{\pm} for a semigroup 𝒮\mathcal{S}, generated by a collection 𝒢.\mathscr{G}. Using them, we show that the positive (or the negative) Julia set of the semigroup 𝒮\mathcal{S}, i.e., 𝒥𝒮+\mathcal{J}_{\mathcal{S}}^{+} (or 𝒥𝒮\mathcal{J}_{\mathcal{S}}^{-}) is equal to the closure of the union of individual positive (or negative) Julia sets of the maps, in the semigroup 𝒮\mathcal{S}. Furthermore, we prove that μ𝒢+\mu_{\mathscr{G}}^{+} is supported on the whole of 𝒥𝒮+\mathcal{J}_{\mathcal{S}}^{+} and is also the unique positive closed (1,1)(1,1)-current supported on 𝒥𝒮+\mathcal{J}_{\mathcal{S}}^{+}, satisfying a semi-invariance relation that depends on the generating set 𝒢\mathscr{G}.

Secondly, we study the dynamics of a non-autonomous sequence of Hénon maps, say {hk}\{h_{k}\}, contained in the semigroup 𝒮\mathcal{S}. Similarly, as above, here too, we construct the non-autonomous dynamical positive and negative Green’s function and the corresponding dynamical Green’s currents. Further, we use the properties of Green’s function to conclude that the non-autonomous attracting basin of any such sequence {hk}\{h_{k}\}, sharing a common attracting fixed point, is biholomorphic to 2.\mathbb{C}^{2}.

Key words and phrases:
Hénon maps, Semigroups, Julia and Fatou sets, Fatou-Bieberbach domains
2020 Mathematics Subject Classification:
Primary: 37F80, 32H50; Secondary: 37F44

1. Introduction

In this article, we study the dynamics of a semigroup of polynomial automorphisms of 2\mathbb{C}^{2}, generated by finitely many Hénon maps. To explain the setup, let 𝒢\mathscr{G} be a given finite collection of automorphisms of 2\mathbb{C}^{2}. We will consider the semigroup 𝒮\mathcal{S} generated by the elements of 𝒢\mathscr{G} under the composition operation, which will be denoted as

𝒮=𝒢 where 𝒢={𝖧i:1in0},\displaystyle\mathcal{S}=\langle\mathscr{G}\rangle\text{ where }\mathscr{G}=\{{\mathsf{H}}_{i}:1\leq i\leq n_{0}\}, (1.1)

and n01n_{0}\geq 1, a positive integer. Furthermore, we assume the maps 𝖧i\mathsf{H}_{i} are Hénon maps, i.e., for every 1in01\leq i\leq n_{0}, there exists mi1m_{i}\geq 1 such that

𝖧i=H1iH2iHmii\displaystyle\mathsf{H}_{i}=H_{1}^{i}\circ H_{2}^{i}\circ\cdots\circ H_{m_{i}}^{i} (1.2)

and Hji(x,y)H_{j}^{i}(x,y) is a map of the form

H(x,y)=(y,p(y)ax),\displaystyle H(x,y)=(y,p(y)-ax), (1.3)

where a0a\neq 0 and pp is a polynomial of degree at least 2, for every 1in01\leq i\leq n_{0} and 1jmi1\leq j\leq m_{i}.

Recall from [26], a map of the above form (1.3) was termed as generalised Hénon map, and as Hénon if p(y)=y2+cp(y)=y^{2}+c, classically. However, the maps of form (1.2) are more general than (1.3) and the methods to study maps of form (1.3), mostly generlises for the entire class. Hence by the terminology Hénon map we will mean maps of the form (1.2). Our interest to study the dynamics of semigroups generated by Hénon maps, is for the following facts

  • Firstly, a classical result of Friedland–Milnor [17] states that these maps are essentially the only class of polynomial automorphisms of 2\mathbb{C}^{2}, exhibiting interesting (iterative) dynamics and have been studied intensively. For instance, see [21], [4], [12] etc. Further, their dynamics are known to be connected to the dynamics of polynomial maps in \mathbb{C} (see [6]) and they also extend as (bi)rational maps of 2.\mathbb{P}^{2}.

  • Secondly, the study of dynamics of arbitrary families, i.e., non-iterative families of holomorphic self-maps (endomorphisms) is important from the point of view of complex function theory. In particular, the (non-autonomous) basins of attraction — see Theorem 1.1 below for definition — of a sequence of automorphisms of k\mathbb{C}^{k}, k2,k\geq 2, with a common attracting fixed point has lead to the construction of pathological domains in k\mathbb{C}^{k} (see [28], [1], [15]).

  • Also, it is conjectured — follows as a consequence of a conjecture, originally due to Bedford ([3],[16]) — that a non-autonomous basin of attraction of sequences of automorphisms of 2\mathbb{C}^{2}, that vary within an infinite (or finite) collection sharing a common uniformly attracting fixed point should be biholomorphic to 2\mathbb{C}^{2}. The same is true for autonomous (or iterative) basins of endomorphisms with an attracting fixed point (see [28, Theorem 9.1]).

To mention here in Section 7, we answer the above for a finite collection of Hénon maps by the methods developed in this article to study the semigroup 𝒮\mathcal{S}. It is stated as

Theorem 1.1.

Let 𝒮\mathcal{S} be as in (1.1), such that the generators, 𝖧i\mathsf{H}_{i}, 1in01\leq i\leq n_{0}, are attracting on a neighbourhood of origin, i.e., there exist r>0r>0 and 0<α<10<\alpha<1 such that

𝖧i(z)αz for every zB(0;r).\displaystyle\|\mathsf{H}_{i}(z)\|\leq\alpha\|z\|\text{ for every }z\in B(0;r). (1.4)

Then the non-autonomous basin of attraction at the origin of every sequence {hk}𝒮\{h_{k}\}\subset\mathcal{S} defined as Ω{hk}:={z2:hkhk1h1(z)0 as k}\Omega_{\{h_{k}\}}:=\{z\in\mathbb{C}^{2}:h_{k}\circ h_{k-1}\circ\cdots\circ h_{1}(z)\to 0\text{ as }k\to\infty\} is biholomorphic to 2\mathbb{C}^{2}.

Further, we study a few particular cases of an infinite collection or parametrised families of Hénon maps sharing a common attracting fixed point with ‘uniform bounds’. In particular, the following example is obtained as a consequence of Corollary 7.4 (also see Example 7.5), in comparison to Theorem 1.4 and 1.10 in [15].

Example 1.2.

Let Hk(x,y)=(aky,akx+y2)H_{k}(x,y)=(a_{k}y,a_{k}x+y^{2}) where pp is a polynomial of degree at least 22 and c<|ak|<dc<|a_{k}|<d for every k1k\geq 1, with 0<c<d<10<c<d<1. Then the basin of attraction of the sequence {Hk}\{H_{k}\}, i.e., Ω{Hk}\Omega_{\{H_{k}\}} (as defined in Theorem 1.1) is biholomorphic to 2\mathbb{C}^{2}.

  • Lastly, the study of dynamics of rational semigroups on 1\mathbb{P}^{1} is an interesting and widely studied area. This setup was introduced by Hinkkanen–Martin, in [20], motivated by their connection to the dynamics of Kleinian groups on the Riemann sphere, observed in [18].

Our primary goal in this article is, to explore the dynamics of a semigroup of Hénon maps both in 2\mathbb{P}^{2} and 2\mathbb{C}^{2}, motivated by the study of dynamics of rational semigroups in 1.\mathbb{P}^{1}. In particular, we will attempt to connect between results from iterative dynamics of Hénon maps of the form (1.2) and semigroup dynamics of rational maps in 1\mathbb{P}^{1} to the current setup. Later, we will generalise a few results appropriately in the setup of non-autonomous families to obtain the aforementioned applications.

Let XX be a complex manifold and 𝒮\mathcal{S} be an infinite family of holomorphic self-maps of X.X. The Fatou set for the family 𝒮\mathcal{S} is the largest open set of XX where the family 𝒮\mathcal{S} is normal, i.e.,

𝒮={zX:there exists a neighbourhood of z where the family 𝒮 is normal}.\mathcal{F}_{\mathcal{S}}=\{z\in X:\text{there exists a neighbourhood of }z\text{ where the family }\mathcal{S}\text{ is normal}\}.

The Julia set 𝒥𝒮\mathcal{J}_{\mathcal{S}}, is the complement of the Fatou set in X.X.

As reported earlier, the setup considering X=1X=\mathbb{P}^{1} and 𝒮\mathcal{S}, a semigroup generated by more than one rational map of degree at least 2, was introduced in [20] and later on has been explored extensively. A major difficulty in this framework — as compared to the iterative dynamics — is neither the Julia set nor the Fatou set is completely invariant, in general.

It is a classical result of Brolin [8] that says - if 𝒮\mathcal{S} is the semigroup of iterates of a (single) polynomial map pp of degree at least 2, the limiting distribution of points in the preimages of a generic point z1z\in\mathbb{P}^{1}, corresponds to the equilibrium measure of the Julia set. Further, the potential associated with this measure, i.e., the Green’s function of the Julia set can be constructed via the dynamics of pp. The equidistribution of the iterated preimages of a generic point z1z\in\mathbb{P}^{1}, i.e., the limiting distribution is independent of the (generic) zz, was established for the iterations of rational map of 1\mathbb{P}^{1} by Lyubich in [25]. Boyd in [7], extended Lyubich’s method and constructed an equidistributed measure supported on the Julia set of a finitely generated semigroup of rational maps (of degree at least 2) in 1\mathbb{P}^{1}. For a finitely generated semigroup of polynomials of degree at least 2, Boyd’s measure is not, in general, the equilibrium measure of its Julia set. Recently, in [24] the latter measure is interpreted as an equilibrium measure in the presence of an external field, which is given by a generalisation of the Greens function — attributed as ‘dynamical Greens function’.

To note, equidistributed measures exist for dynamics of certain meromorphic correspondences on compact connected Kähler manifolds, of appropriate intermediate degree (see [10]). However, birational maps of 2\mathbb{P}^{2} obtained from extension of Hénon maps do not belong to the above category. Also, for iterative families of a Hénon map the Julia set is captured via the support of a unique positive closed (1,1)(1,1)-current of mass 1, obtained by the action of ddcdd^{c}-operator on the pluri-complex Green’s function of the Julia set. Furthermore, it is an equidistributed current in 2\mathbb{C}^{2}, in the sense, that it can be recovered as a limit of appropriately weighted preimages of an algebraic variety in 2\mathbb{C}^{2} — see [4, Theorem 4.7] or [12, Corollary 6.7].

To mention here, construction of currents for non-autonomous families of Hénon maps have been done on an appropriate bounded region containing the origin, in [11], via the fact they are horizontal. Here, we construct a (similar) global equidistributed dynamical Green’s current on 2\mathbb{P}^{2} in Corollaries 5.8 and 6.11 — using the dynamical Green’s functions, constructed by generalising ideas from [7], [11], [24] — both for the semigroups 𝒮\mathcal{S} and non-autonomous families. Thus, obtaining the uniqueness of the currents upto a semi-invariance property for the semigroups 𝒮\mathcal{S}, stated in Corollary 1.6. Also, see Remark 6.12, for details.

Let us first recall a few important properties of iterations of a Hénon map 𝖧\mathsf{H}. The pluri-complex Green’s function (see [22] for the definition) associated to the Julia set of iterates of 𝖧\mathsf{H} or 𝖧1\mathsf{H}^{-1}, say G𝖧+G_{\mathsf{H}}^{+} or G𝖧G_{\mathsf{H}}^{-} respectively, can be recovered via the dynamics. In particular, if d𝖧d_{\mathsf{H}} is the degree of the map 𝖧\mathsf{H} then

G𝖧±(z)=limklog+𝖧±k(z)d𝖧k and μ𝖧±(z)=12πddc(G𝖧±),G_{\mathsf{H}}^{\pm}(z)=\lim_{k\to\infty}\frac{\log^{+}\|\mathsf{H}^{\pm k}(z)\|}{d_{\mathsf{H}}^{k}}\text{ and }\mu_{\mathsf{H}}^{\pm}(z)=\frac{1}{2\pi}dd^{c}(G_{\mathsf{H}}^{\pm}),

where log+x=max{logx,0}\log^{+}x=\max\{\log x,0\} for every x>0x>0 and \|\cdot\| be the supremum norm in 2\mathbb{C}^{2}. Also, μ𝖧±\mu_{\mathsf{H}}^{\pm} are the closed positive (1,1)(1,1)-(equidistributed) currents as mentioned previously. Note that the above definition holds for any norm on 2\mathbb{C}^{2}, however for the sake of convenience we will use the notation \|\cdot\| to denote the supremum norm, throughout this article.

Now, let 𝒮\mathcal{S} be the semigroup as introduced in (1.1), i.e., 𝒮=𝒢\mathcal{S}=\langle\mathscr{G}\rangle where 𝒢={𝖧i:1in0}\mathscr{G}=\{\mathsf{H}_{i}:1\leq i\leq n_{0}\} and 𝖧i\mathsf{H}_{i} are Hénon maps of the form (1.2) and of degree di2d_{i}\geq 2 for every 1in.1\leq i\leq n. We first generalise a few definitions and observe some basic results regarding the semigroup 𝒮\mathcal{S} in Section 2. Particularly, we note that 𝒮\mathcal{S} might have more than one generating set, however it has a unique minimal generating set.

In Section 3, we generalise the construction of positive and negative Green’s functions, i.e., the functions G𝖧±G_{\mathsf{H}}^{\pm} noted above, in the setup of the semigroup 𝒮.\mathcal{S}. To do the same, we define the total degree of the semigroup 𝒮\mathcal{S} with respect to the generating set 𝒢\mathscr{G} as D𝒢=i=1n0diD_{\mathscr{G}}=\sum_{i=1}^{n_{0}}d_{i}, and consider the sequence of plurisubharmonic functions Gk±G_{k}^{\pm} on 2\mathbb{C}^{2} defined as

Gk+(z)=1D𝒢kh𝒢klog+h(z) and Gk(z)=1D𝒢kh𝒢klog+h1(z),\displaystyle G_{k}^{+}(z)=\frac{1}{D_{\mathscr{G}}^{k}}\sum_{h\in\mathscr{G}_{k}}\log^{+}\|h(z)\|\text{ and }G_{k}^{-}(z)=\frac{1}{D_{\mathscr{G}}^{k}}\sum_{h\in\mathscr{G}_{k}}\log^{+}\|h^{-1}(z)\|, (1.5)

where 𝒢k\mathscr{G}_{k} denote the elements of the semigroup 𝒮\mathcal{S} of length kk with respect to the generating set 𝒢\mathscr{G}, i.e., 𝒢k={𝖧i1𝖧ik:1ijn0 and 1jk}.\mathscr{G}_{k}=\{\mathsf{H}_{i_{1}}\circ\cdots\circ\mathsf{H}_{i_{k}}:1\leq i_{j}\leq n_{0}\text{ and }1\leq j\leq k\}. We prove that the pointwise limits of the sequences {Gk±}\{G_{k}^{\pm}\} constructed in (1.5) exist, which is stated as

Theorem 1.3.

The sequences {Gk±}\{G_{k}^{\pm}\} converge pointwise to plurisubharmonic, continuous functions G𝒢±G_{\mathscr{G}}^{\pm} on 2\mathbb{C}^{2}, respectively.

Henceforth, the functions G𝒢±G_{\mathscr{G}}^{\pm} will be referred as the dynamical positive (or negative) Green’s function associated to the semigroup 𝒮\mathcal{S} generated by the set 𝒢={𝖧i:1in0}\mathscr{G}=\{{\mathsf{H}}_{i}:1\leq i\leq n_{0}\}. The need to specify the generating set 𝒢\mathscr{G} is important as the semigroup 𝒮\mathcal{S} may admit multiple generating sets. Also, note that the functions G𝒢±G_{\mathscr{G}}^{\pm} satisfy the following semi-invariance relation right by the construction (1.5) and Theorem 1.3.

Corollary 1.4.

i=1n0G𝒢+𝖧i(z)=D𝒢.G𝒢+(z)\sum_{i=1}^{n_{0}}G_{\mathscr{G}}^{+}\circ\mathsf{H}_{i}(z)=D_{\mathscr{G}}.G_{\mathscr{G}}^{+}(z) and i=1n0G𝒢𝖧i1(z)=D𝒢.G𝒢(z).\sum_{i=1}^{n_{0}}G_{\mathscr{G}}^{-}\circ\mathsf{H}_{i}^{-1}(z)=D_{\mathscr{G}}.G_{\mathscr{G}}^{-}(z).

Thus, as consequences of the proof of Theorem 1.3, we note that the functions G𝒢±G_{\mathscr{G}}^{\pm} admit logarithmic growth on appropriate regions, and the strong filled positive and negative Julia sets of the semigroup 𝒮\mathcal{S} are pseudoconcave sets (see Corollary 3.4 and Remark 3.5).

Next, we analyse the Julia sets 𝒥𝒮±\mathcal{J}_{\mathcal{S}}^{\pm} and the properties the dynamical Green’s (1,1)(1,1)-currents associated to the functions G𝒢±G_{\mathscr{G}}^{\pm}, defined as μ𝒢±=12πddcG𝒢±\mu_{\mathscr{G}}^{\pm}=\frac{1}{2\pi}dd^{c}G_{\mathscr{G}}^{\pm}. Consequently, in Section 4, we prove the analogue to Corollary 2.1 from [20] — an important fact from the dynamics of semigroups of the rational maps on 1\mathbb{P}^{1} — via the supports of μ𝒢±.\mu_{\mathscr{G}}^{\pm}.

Theorem 1.5.

The positive and negative Julia sets corresponding to the dynamics of the semigroup 𝒮\mathcal{S} is equal to the closure of the union of the (positive and negative) Julia sets of the elements of 𝒮\mathcal{S} respectively, i.e., 𝒥𝒮+=h𝒮Jh+¯ and 𝒥𝒮=h𝒮Jh¯.\displaystyle\mathcal{J}_{\mathcal{S}}^{+}=\overline{\bigcup_{h\in\mathcal{S}}J_{h}^{+}}\text{ and }\mathcal{J}_{\mathcal{S}}^{-}=\overline{\bigcup_{h\in\mathcal{S}}J_{h}^{-}}.

Further, the positive and the negative dynamical Green’s currents μ𝒢±\mu_{\mathscr{G}}^{\pm} are (1,1)(1,1)-closed positive currents of mass 1 supported (respectively) on the Julia sets, i.e., Supp (μ𝒢±)=𝒥𝒮±.\text{Supp }(\mu_{\mathscr{G}}^{\pm})=\mathcal{J}_{\mathcal{S}}^{\pm}.

Thus from Theorem 1.3 and the above, G𝒢±G_{\mathscr{G}}^{\pm} is actually pluriharmonic on the Fatou sets 𝒮±\mathcal{F}_{\mathcal{S}}^{\pm}. In Section 5, we study the extension of the currents μ𝒢±\mu_{\mathscr{G}}^{\pm} to 2\mathbb{P}^{2} and prove that they are limits of (weighted) equidistributed projective varieties, in the spirit of [12, Theorem 6.2]. Also, consequently we observe the following uniqueness of μ𝒢+\mu_{\mathscr{G}}^{+} upto a semi-invariance property.

Corollary 1.6.

The current μ𝒢+\mu_{\mathscr{G}}^{+} is the unique current of mass 1 supported on 𝒥𝒮+\mathcal{J}_{\mathcal{S}}^{+} and the current μ𝒢\mu_{\mathscr{G}}^{-} is the unique current of mass 1 supported on 𝒥𝒮\mathcal{J}_{\mathcal{S}}^{-} satisfying the following semi-invariance relations (respectively)

1D𝒢i=1n0𝖧i(μ𝒢+)=μ𝒢+ and 1D𝒢i=1n0𝖧i(μ𝒢)=μ𝒢.\displaystyle\frac{1}{D_{\mathscr{G}}}\sum_{i=1}^{n_{0}}\mathsf{H}_{i}^{*}(\mu_{\mathscr{G}}^{+})=\mu_{\mathscr{G}}^{+}\text{ and }\frac{1}{D_{\mathscr{G}}}\sum_{i=1}^{n_{0}}{\mathsf{H}_{i}}_{*}(\mu_{\mathscr{G}}^{-})=\mu_{\mathscr{G}}^{-}. (1.6)

In Section 6, we consider the dynamics of a non-autonomous sequence of Hénon maps, say {hk}𝒮\{h_{k}\}\in\mathcal{S} and prove that there exist plurisubharmonic and continuous dynamical (positive and negative) Green’s functions, denoted by 𝖦{hk}±\mathsf{G}_{\{h_{k}\}}^{\pm}, with logarithmic growth. Thus μ{hk}±=12πddc(𝖦{hk}±)\mu_{\{h_{k}\}}^{\pm}=\frac{1}{2\pi}dd^{c}(\mathsf{G}_{\{h_{k}\}}^{\pm}), are positive (1,1)(1,1)-currents of mass 1, supported on the positive and negative Julia sets of the sequence {hk}.\{h_{k}\}. Also, we obtain the analogs to the equidistribution results, i.e., Corollaries 5.8 and 5.10, in this setup of non-autonomous dynamics. However, we will discuss them briefly as most of the ideas are similar to that realised in Section 5 and, depends upon the existence of the dynamical Green’s function with suitable growth at infinity.

Finally, we study the non-autonomous attracting basin of a sequence of Hénon maps of the form (1.2), admitting a uniformly attracting behaviour, on a neighbourhood of the origin. Further, we prove Theorem 1.1, as an application of the existence of Green’s functions 𝖦{hk}±\mathsf{G}^{\pm}_{\{h_{k}\}} and enlist a few more applications, which follows from the technique. All of these affirmatively answers a few particular cases of the equivalent formulation of the Bedford Conjecture, in 2\mathbb{C}^{2} for Hénon maps — as alluded to in the beginning.

Acknowledgement

The author would like to thank the anonymous referee for carefully reading the manuscript and suggesting helpful comments.

2. Some basic definitions and preliminaries

In this section, we first observe a proposition about the generating set 𝒢\mathscr{G} of the semigroup 𝒮\mathcal{S} as in (1.1), which might not be unique, always. Recall the setup from Section 1, let 𝒢={𝖧i:1in0}\displaystyle\mathscr{G}=\{\mathsf{H}_{i}:1\leq i\leq n_{0}\} where 𝖧i\mathsf{H}_{i}’s are Hénon maps of the form (1.2), with degree di2d_{i}\geq 2. The total degree of the semigroup 𝒮\mathcal{S} with respect to the generating set 𝒢\mathscr{G} is D𝒢=i=1n0diD_{\mathscr{G}}=\sum_{i=1}^{n_{0}}d_{i} and 𝒢k\mathscr{G}_{k} is the set of all elements of length kk, k1k\geq 1 in the semigroup 𝒮\mathcal{S} with respect to 𝒢\mathscr{G}.

Proposition 2.1.

Let 𝒮\mathcal{S} be a finitely generated semigroup as in (1.1), then there exists a unique minimal set 𝒢0\mathscr{G}_{0} of maps of the form (1.2) that generates 𝒮\mathcal{S}, i.e., any set of generators 𝒢\mathscr{G} of 𝒮\mathcal{S} is a superset of 𝒢0.\mathscr{G}_{0}.

Proof.

For n1n\geq 1, let 𝒮(n)={H𝒮:degree of H is n}.\mathcal{S}(n)=\{H\in\mathcal{S}:\text{degree of }H\text{ is }n\}. Note that 𝒮(1)\mathcal{S}(1) is an empty set, as the degree of every element in the generating set 𝒢\mathscr{G} is at least 2.2. However, 𝒮(n)\mathcal{S}(n) for every n1n\geq 1, need not necessarily be empty but is always a finite set. We will construct the minimal generating set 𝒢0\mathscr{G}_{0} inductively, such that it terminates after finitely many steps. Let

A2:=𝒮(2),A3:=𝒮(3)A2,A4:=𝒮(4)A2A3,,An:=𝒮(n)i=2n1Ai.A_{2}:=\mathcal{S}(2),A_{3}:=\mathcal{S}(3)\setminus\big{\langle}A_{2}\big{\rangle},A_{4}:=\mathcal{S}(4)\setminus\big{\langle}A_{2}\cup A_{3}\big{\rangle},\ldots,A_{n}:=\mathcal{S}(n)\setminus\big{\langle}\bigcup_{i=2}^{n-1}A_{i}\big{\rangle}.

Since 𝒮\mathcal{S} is finitely generated, there exists an n01n_{0}\geq 1, such that An=A_{n}=\emptyset for n>n0n>n_{0} and An0.A_{n_{0}}\neq\emptyset. Let

𝒢0=i=2n0Ai.\mathscr{G}_{0}=\bigcup_{i=2}^{n_{0}}A_{i}.

Note that by construction, any element in 𝒢0\mathscr{G}_{0} is not generated by lower degree maps of form (1.2). Further as An=A_{n}=\emptyset for every n>n0n>n_{0}, 𝒢0\mathscr{G}_{0} is the minimal set generating 𝒮.\mathcal{S}.

Remark 2.2.

Thus the total degree of a semigroup 𝒮\mathcal{S} is dependent on the generating set 𝒢\mathscr{G} and is not unique, in general. Consequently, the sequence of plurisubharmonic functions {Gk±}\{G_{k}^{\pm}\} defined in (1.5) and the positive and negative dynamical Green’s function is also dependent on the generating set 𝒢\mathscr{G} of the semigroup 𝒮.\mathcal{S}.

Next, we revisit and introduce a few important definitions (and notations) with respect to the dynamics of the semigroup 𝒮\mathcal{S}, that are independent of the generating set 𝒢\mathscr{G}.

  • Let 𝒮\mathcal{S}^{-} denote the semigroup of maps comprising of the inverse of the maps that belong to 𝒮\mathcal{S} and 𝒢\mathscr{G}^{-} the inverse of the elements that belong to 𝒢\mathscr{G}, i.e.,

    𝒢={𝖧i1:1in0} and 𝒮=𝒢.\mathscr{G}^{-}=\{\mathsf{H}_{i}^{-1}:1\leq i\leq n_{0}\}\text{ and }\mathcal{S}^{-}=\langle\mathscr{G}^{-}\rangle.
  • The Fatou sets of 𝒮\mathcal{S} and 𝒮\mathcal{S}^{-}) — as stated in Section 1 — is denoted by 𝒮+\mathcal{F}_{\mathcal{S}}^{+} and 𝒮\mathcal{F}_{\mathcal{S}}^{-} respectively. The positive and negative Julia sets are denoted by 𝒥𝒮±.\mathcal{J}_{\mathcal{S}}^{\pm}.

  • We consider the following two alternatives for the filled positive and negative Julia sets.

    1. (1)

      The strong positive (or negative) filled Julia set is defined as the collection of all the points z2z\in\mathbb{C}^{2} such that for every sequence {hk}𝒮\{h_{k}\}\subset\mathcal{S}, the sequence {hk(z)}\{h_{k}(z)\} (or the sequence {hk1(z)}\{h_{k}^{-1}(z)\}, respectively) is bounded, i.e.,

      𝒦𝒮+={z2:for every sequence {hk}𝒮 the sequence {hk(z)} is bounded},\displaystyle{\mathcal{K}}^{+}_{\mathcal{S}}=\big{\{}z\in\mathbb{C}^{2}:\text{for every sequence }\{h_{k}\}\subset\mathcal{S}\text{ the sequence }\{h_{k}(z)\}\text{ is bounded}\big{\}},

      𝒦𝒮={z2:for every sequence {hk}𝒮 the sequence {hk1(z)} is bounded}.\displaystyle{\mathcal{K}}^{-}_{\mathcal{S}}=\big{\{}z\in\mathbb{C}^{2}:\text{for every sequence }\{h_{k}\}\subset\mathcal{S}\text{ the sequence }\{h_{k}^{-1}(z)\}\text{ is bounded}\big{\}}.

    2. (2)

      The weak positive filled Julia set is defined as the collection of all the points z2z\in\mathbb{C}^{2} such that there exists a sequence {hk}𝒮\{h_{k}\}\subset\mathcal{S} with hk𝒢nkh_{k}\in\mathscr{G}_{n_{k}}, where nkn_{k}\to\infty as kk\to\infty, and {hk(z)}\{h_{k}(z)\} is bounded. Similarly we define the weak negative filled Julia set, i.e.,

      𝐊𝒮+={z2: there exist hk𝒢nk such that nk and {hk(z)} is bounded},\displaystyle\mathbf{K}^{+}_{\mathcal{S}}=\big{\{}z\in\mathbb{C}^{2}:\text{ there exist }h_{k}\in\mathscr{G}_{n_{k}}\text{ such that }n_{k}\to\infty\text{ and }\{h_{k}(z)\}\text{ is bounded}\big{\}},

      𝐊𝒮={z2: there exist h~k𝒢nk such that nk and {h~k1(z)} is bounded}.\displaystyle\mathbf{K}^{-}_{\mathcal{S}}=\big{\{}z\in\mathbb{C}^{2}:\text{ there exist }\tilde{h}_{k}\in\mathscr{G}_{n_{k}}\text{ such that }n_{k}\to\infty\text{ and }\{\tilde{h}_{k}^{-1}(z)\}\text{ is bounded}\big{\}}.

    Note that 𝒦𝒮±𝐊𝒮±\mathcal{K}_{\mathcal{S}}^{\pm}\subset\mathbf{K}_{\mathcal{S}}^{\pm} and these sets are uniquely associated to the semigroup 𝒮\mathcal{S}.

  • Similarly as above we introduce the weak and strong escaping sets 𝒰𝒮±\mathcal{U}^{\pm}_{\mathcal{S}} and 𝐔𝒮±\mathbf{U}_{\mathcal{S}}^{\pm}

    𝒰𝒮±=2𝒦𝒮± and 𝐔𝒮±=2𝐊𝒮±.\mathcal{U}_{\mathcal{S}}^{\pm}=\mathbb{C}^{2}\setminus\mathcal{K}_{\mathcal{S}}^{\pm}\text{ and }\mathbf{U}_{\mathcal{S}}^{\pm}=\mathbb{C}^{2}\setminus\mathbf{K}_{\mathcal{S}}^{\pm}.

    Note that 𝐔𝒮+\mathbf{U}_{\mathcal{S}}^{+} is the Fatou component at infinity with respect to the dynamics of the semigroup 𝒮.\mathcal{S}. Similarly 𝐔𝒮\mathbf{U}_{\mathcal{S}}^{-} is the Fatou component at infinity for 𝒮\mathcal{S}^{-}.

  • Finally, we define the cumulative positive and negative Julia sets for the semigroup 𝒮\mathcal{S}, i.e.,

    𝐉𝒮+=h𝒮𝒥h+¯,𝐉𝒮=h𝒮𝒥h¯.\mathbf{J}_{\mathcal{S}}^{+}=\overline{\bigcup_{h\in\mathcal{S}}\mathcal{J}_{h}^{+}},\,\mathbf{J}_{\mathcal{S}}^{-}=\overline{\bigcup_{h\in\mathcal{S}}\mathcal{J}_{h}^{-}}.

Observe that, either of the sets 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+} or 𝒦𝒮\mathcal{K}_{\mathcal{S}}^{-} or both may be empty for some semigroups 𝒮\mathcal{S}, of form (1.1). However, this situation does not affect the dynamics as such.

We will now explore some important properties of the sets introduced above via the filtration properties of the elements of 𝒮\mathcal{S}, on appropriate domains. To discuss this in detail, let us first recall the definition of the sets VRV_{R} and VR±V_{R}^{\pm} for some R>0R>0, introduced in [21] (or [4]) for filtering the dynamics of (finite) compositions of generalised Hénon maps. They are VR={(x,y)2:max{|x|,|y|}R}V_{R}=\big{\{}(x,y)\in\mathbb{C}^{2}:\max\{|x|,|y|\}\leq R\big{\}}, the polydisk of radius RR and

VR+={(x,y)2:|y|max{|x|,R}},VR={(x,y)2:|x|max{|y|,R}}.\displaystyle V_{R}^{+}=\big{\{}(x,y)\in\mathbb{C}^{2}:|y|\geq\max\{|x|,R\}\big{\}},\;\;V_{R}^{-}=\big{\{}(x,y)\in\mathbb{C}^{2}:|x|\geq\max\{|y|,R\}\big{\}}.

Also recall the subsets 𝒢k\mathscr{G}_{k} of 𝒮=𝒢\mathcal{S}=\langle\mathscr{G}\rangle, for every k1k\geq 1, defined as

𝒢k={𝖧i1𝖧ik:1ijn0 and 1jk}, where 𝒢1=𝒢={𝖧i:1in0}.\mathscr{G}_{k}=\{\mathsf{H}_{i_{1}}\circ\cdots\circ\mathsf{H}_{i_{k}}:1\leq i_{j}\leq n_{0}\text{ and }1\leq j\leq k\},\text{ where }\mathscr{G}_{1}=\mathscr{G}=\{\mathsf{H}_{i}:1\leq i\leq n_{0}\}.

We first record the dynamical behaviour of the semigroup 𝒮\mathcal{S} on VR±V_{R}^{\pm} for an appropriate R>0R>0.

Lemma 2.3.

There exists R𝒮>0R_{\mathcal{S}}>0 such that for every R>R𝒮R>R_{\mathcal{S}},

h(VR+)¯VR+ and h1(VR¯)VR, i.e., VRh(VR+)= and VRh1(VR)=\overline{h(V_{R}^{+})}\subset V_{R}^{+}\text{ and }\overline{h^{-1}(V_{R}^{-}})\subset V_{R}^{-},\text{ i.e., }V_{R}\cap h(V_{R}^{+})=\emptyset\text{ and }V_{R}\cap h^{-1}(V_{R}^{-})=\emptyset

whenever h𝒮.h\in\mathcal{S}. Further let {hk}𝒮\{h_{k}\}\subset\mathcal{S} such that hk𝒢kh_{k}\in\mathscr{G}_{k} for every k1k\geq 1, then there exists a sequence positive real numbers RkR_{k}\to\infty satisfying

VRkhk(VR+)= and VRkhk1(VR)=.V_{R_{k}}\cap h_{k}(V_{R}^{+})=\emptyset\text{ and }V_{R_{k}}\cap h_{k}^{-1}(V_{R}^{-})=\emptyset.
Proof.

Recall from [4], for R>0R>0 (sufficiently large) there exists 0<m<M0<m<M such that

m|y|di<|π2𝖧i(x,y)|<M|y|di on VR+,m|x|di<|π1𝖧i1(x,y)|<M|x|di on VR.\displaystyle m|y|^{d_{i}}<|\pi_{2}\circ\mathsf{H}_{i}(x,y)|<M|y|^{d_{i}}\text{ on }V_{R}^{+},\;m|x|^{d_{i}}<|\pi_{1}\circ\mathsf{H}_{i}^{-1}(x,y)|<M|x|^{d_{i}}\text{ on }V_{R}^{-}.

Recall from (1.2)

𝖧i=H1iH2iHmii,\mathsf{H}_{i}=H_{1}^{i}\circ H_{2}^{i}\circ\cdots\circ H_{m_{i}}^{i},

where Hji(x,y)=(y,pj(y)ajx)H_{j}^{i}(x,y)=(y,p_{j}(y)-a_{j}x). Thus degree of π1𝖧i< the degree of π2𝖧i\pi_{1}\circ\mathsf{H}_{i}<\text{ the degree of }\pi_{2}\circ\mathsf{H}_{i}, for every 1in01\leq i\leq n_{0}. Now as VR+V_{R}^{+} is a closed subset of 2\mathbb{C}^{2} by definition, the above identity further implies that |π1𝖧i(z)|<|π2𝖧i(z)||\pi_{1}\circ\mathsf{H}_{i}(z)|<|\pi_{2}\circ\mathsf{H}_{i}(z)| and |π2𝖧i(z)|>R|\pi_{2}\circ\mathsf{H}_{i}(z)|>R, for zVR+z\in V_{R}^{+}, R>0R>0, sufficiently large. In particular, there exists R>0R>0, large enough such that

𝖧i(VR+)¯=𝖧i(VR+)int(VR+),𝖧i1(VR)¯=𝖧i1(VR)int(VR) for every 1in0.\displaystyle\overline{\mathsf{H}_{i}(V_{R}^{+})}=\mathsf{H}_{i}(V_{R}^{+})\subset\textsf{int}(V_{R}^{+}),\;\overline{\mathsf{H}_{i}^{-1}(V_{R}^{-})}=\mathsf{H}_{i}^{-1}(V_{R}^{-})\subset\textsf{int}(V_{R}^{-})\text{ for every }1\leq i\leq n_{0}. (2.1)

Let d0=min{di:1in0}2d_{0}=\min\{d_{i}:1\leq i\leq n_{0}\}\geq 2 and R𝒮>1R_{\mathcal{S}}>1 be sufficiently large such that 1<R𝒮<mR𝒮d0.1<R_{\mathcal{S}}<mR_{\mathcal{S}}^{d_{0}}. Hence from (2.1), for Rk=mRk1d0R_{k}=mR^{d_{0}}_{k-1} whenever k2k\geq 2 and R1>R𝒮R_{1}>R_{\mathcal{S}}. Thus the proof. ∎

The constant R𝒮>0R_{\mathcal{S}}>0 obtained in Lemma 2.3 is actually independent of the generators and will be referred along, as the radius of filtration for the semigroup 𝒮.\mathcal{S}.

Remark 2.4.

Note that in the above proof we may assume 0<m<1<M0<m<1<M, such that for every R>R𝒮R>R_{\mathcal{S}} and 1in01\leq i\leq n_{0}

m|y|di<|π2𝖧i(x,y)|<M|y|di on VR+, and m|x|di<|π1𝖧i1(x,y)|<M|x|di on VR.m|y|^{d_{i}}<|\pi_{2}\circ\mathsf{H}_{i}(x,y)|<M|y|^{d_{i}}\text{ on }V_{R}^{+},\text{ and }m|x|^{d_{i}}<|\pi_{1}\circ\mathsf{H}_{i}^{-1}(x,y)|<M|x|^{d_{i}}\text{ on }V_{R}^{-}.
Proposition 2.5.

The sets 𝐊𝒮±\mathbf{K}_{\mathcal{S}}^{\pm} and 𝒦𝒮±\mathcal{K}_{\mathcal{S}}^{\pm} are closed subsets of 2\mathbb{C}^{2} and 𝒦𝒮±𝐊𝒮±VRVR\mathcal{K}_{\mathcal{S}}^{\pm}\subset\mathbf{K}_{\mathcal{S}}^{\pm}\subset V_{R}\cup V_{R}^{\mp} (respectively) for RR𝒮.R\geq R_{\mathcal{S}}.

Proof.

Let U0=int(VR+)U_{0}=\textsf{int}(V_{R}^{+}) and let {𝐔k}\{\mathbf{U}_{k}\}, {𝒰k}\{\mathcal{U}_{k}\} be the sequences of open subsets defined as

𝐔k=h𝒢kh1(U0) and 𝒰k=h𝒢kh1(U0).\mathbf{U}_{k}=\bigcap_{h\in\mathscr{G}_{k}}h^{-1}(U_{0})\text{ and }\mathcal{U}_{k}=\bigcup_{h\in\mathscr{G}_{k}}h^{-1}(U_{0}).

Then

𝐔k¯=h𝒢kh1(U0¯) and 𝒰k¯=h𝒢kh1(U0¯).\overline{\mathbf{U}_{k}}=\bigcap_{h\in\mathscr{G}_{k}}h^{-1}(\overline{U_{0}})\text{ and }\overline{\mathcal{U}_{k}}=\bigcup_{h\in\mathscr{G}_{k}}h^{-1}(\overline{U_{0}}).

Since by (2.1), 𝖧i(z0)int(VR+)\mathsf{H}_{i}(z_{0})\in\textsf{int}(V_{R}^{+}) for every 1in01\leq i\leq n_{0} whenever z0VR+=U0¯z_{0}\in V_{R}^{+}=\overline{U_{0}}, we have U0¯𝖧i1(U0)\overline{U_{0}}\subset\mathsf{H}_{i}^{-1}(U_{0}). Hence U0¯𝐔1𝒰1.\overline{U_{0}}\subset\mathbf{U}_{1}\subset\mathcal{U}_{1}. Further for every h𝒢kh\in\mathscr{G}_{k}, h1(U0¯)h1𝖧i1(U0)h^{-1}(\overline{U_{0}})\subset h^{-1}\circ\mathsf{H}_{i}^{-1}(U_{0}) for every 𝖧i𝒢\mathsf{H}_{i}\in\mathscr{G}, where k1k\geq 1. Thus

𝐔k¯𝐔k+1 and 𝒰k¯𝒰k+1.\displaystyle\overline{\mathbf{U}_{k}}\subset\mathbf{U}_{k+1}\text{ and }\overline{\mathcal{U}_{k}}\subset\mathcal{U}_{k+1}. (2.2)

Let

𝐔+=k0𝐔k and 𝒰+=k0𝒰k.\mathbf{U}^{+}=\bigcup_{k\geq 0}\mathbf{U}_{k}\text{ and }{\mathcal{U}}^{+}=\bigcup_{k\geq 0}\mathcal{U}_{k}.

Observe that for every k1k\geq 1, h(𝐔k)VR+h(\mathbf{U}_{k})\subset V_{R}^{+} whenever h𝒢kh\in\mathscr{G}_{k}. Hence 𝐔+𝐔𝒮+.\mathbf{U}^{+}\subset\mathbf{U}_{\mathcal{S}}^{+}. Now for z2𝐔+z\in\mathbb{C}^{2}\setminus\mathbf{U}^{+}, note that hk(z)VRVRh_{k}(z)\in V_{R}\cup V_{R}^{-} for every sequence {hk}𝒮.\{h_{k}\}\subset\mathcal{S}. Let z0𝐔𝒮+(2𝐔+)z_{0}\in\mathbf{U}_{\mathcal{S}}^{+}\cap(\mathbb{C}^{2}\setminus\mathbf{U}^{+}). Then there exists a sequence {h~k}𝒮\{\tilde{h}_{k}\}\subset\mathcal{S} such that h~k𝒢nk\tilde{h}_{k}\in\mathscr{G}_{n_{k}} with nkn_{k}\to\infty as kk\to\infty and h~k(z0)VR\tilde{h}_{k}(z_{0})\in V_{R}^{-} for every k1k\geq 1, i.e., z0h~k1(VR)z_{0}\in\tilde{h}_{k}^{-1}(V_{R}^{-}). Hence by Lemma 2.3, z0kVRnk=2z_{0}\notin\cup_{k}V_{R_{n_{k}}}=\mathbb{C}^{2}, which is a contradiction! Thus 𝐔+=𝐔𝒮+\mathbf{U}^{+}=\mathbf{U}_{\mathcal{S}}^{+}. A similar argument works for 𝐔=𝐔𝒮\mathbf{U}^{-}=\mathbf{U}_{\mathcal{S}}^{-}.

Similarly for z𝒰k+z\in{\mathcal{U}_{k}^{+}} there exists hk𝒢kh_{k}\in\mathscr{G}_{k} such that hk(z)VR+h_{k}(z)\in V_{R}^{+}, hence 𝒰+𝒰𝒮+\mathcal{U}^{+}\subset\mathcal{U}_{\mathcal{S}}^{+}. Now for z~0𝒰𝒮+(2𝒰+)\tilde{z}_{0}\in\mathcal{U}_{\mathcal{S}}^{+}\cap(\mathbb{C}^{2}\setminus\mathcal{U}^{+}), as in the above case, there exists a sequence {h~k}\{\tilde{h}_{k}\} such that h~k𝒢nk\tilde{h}_{k}\in\mathscr{G}_{n_{k}} with nkn_{k}\to\infty as kk\to\infty and h~k(z~0)VR\tilde{h}_{k}(\tilde{z}_{0})\in V_{R}^{-} for every k1k\geq 1, i.e., z~0h~k1(VR)\tilde{z}_{0}\in\tilde{h}_{k}^{-1}(V_{R}^{-}). Hence by Lemma 2.3, z~0kVRnk=2\tilde{z}_{0}\notin\cup_{k}V_{R_{n_{k}}}=\mathbb{C}^{2}, which is a contradiction! Thus 𝒰+=𝒰𝒮+\mathcal{U}^{+}=\mathcal{U}_{\mathcal{S}}^{+}. A similar argument works for 𝒰=𝒰𝒮\mathcal{U}^{-}=\mathcal{U}_{\mathcal{S}}^{-}.

Note that the above observations also proves that VR±𝐔𝒮±𝒰𝒮±V_{R}^{\pm}\subset\mathbf{U}_{\mathcal{S}}^{\pm}\subset\mathcal{U}_{\mathcal{S}}^{\pm}, hence 𝒦𝒮±𝐊𝒮±VRVR\mathcal{K}_{\mathcal{S}}^{\pm}\subset\mathbf{K}_{\mathcal{S}}^{\pm}\subset V_{R}\cup V_{R}^{\mp} (respectively). ∎

Further, from the proof of Proposition 2.5 we get

Corollary 2.6.

The escaping sets of 𝒮\mathcal{S} can be further realised as

  1. (1)

    𝐔𝒮+=k1h𝒢kh1(VR+) and 𝐔𝒮=k1h𝒢kh(VR)\displaystyle\mathbf{U}_{\mathcal{S}}^{+}=\bigcup_{k\geq 1}\bigcap_{h\in\mathscr{G}_{k}}h^{-1}(V_{R}^{+})\text{ and }\;\mathbf{U}_{\mathcal{S}}^{-}=\bigcup_{k\geq 1}\bigcap_{h\in\mathscr{G}_{k}}h(V_{R}^{-});

  2. (2)

    𝒰𝒮+=k1h𝒢kh1(VR+) and 𝒰𝒮=k1h𝒢kh(VR)\displaystyle\mathcal{U}_{\mathcal{S}}^{+}=\bigcup_{k\geq 1}\bigcup_{h\in\mathscr{G}_{k}}h^{-1}(V_{R}^{+})\text{ and }\;\mathcal{U}_{\mathcal{S}}^{-}=\bigcup_{k\geq 1}\bigcup_{h\in\mathscr{G}_{k}}h(V_{R}^{-}).

Remark 2.7.

Note that 𝐊𝒮±int(𝒦𝒮±)=𝒰𝒮±𝐔𝒮±¯\mathbf{K}_{\mathcal{S}}^{\pm}\setminus\textsf{int}(\mathcal{K}_{\mathcal{S}}^{\pm})=\overline{\mathcal{U}_{\mathcal{S}}^{\pm}\setminus\mathbf{U}_{\mathcal{S}}^{\pm}} and the Julia sets 𝒥𝒮±𝐊𝒮±int(𝒦𝒮±)\mathcal{J}_{\mathcal{S}}^{\pm}\subset\mathbf{K}_{\mathcal{S}}^{\pm}\setminus\textsf{int}(\mathcal{K}_{\mathcal{S}}^{\pm}), however, they might not be equal. Thus so far we have the straightforward inclusion relation

𝐉𝒮±𝒥𝒮±𝐊𝒮±int(𝒦𝒮±).\mathbf{J}_{\mathcal{S}}^{\pm}\subset\mathcal{J}_{\mathcal{S}}^{\pm}\subset\mathbf{K}_{\mathcal{S}}^{\pm}\setminus\textsf{int}(\mathcal{K}_{\mathcal{S}}^{\pm}).

Also both 𝒦𝒮±,𝐊𝒮±𝒥𝒮±\partial\mathcal{K}_{\mathcal{S}}^{\pm},\partial\mathbf{K}_{\mathcal{S}}^{\pm}\subset\mathcal{J}_{\mathcal{S}}^{\pm}, i.e., they may be proper subsets 𝒥𝒮±\mathcal{J}_{\mathcal{S}}^{\pm}, unlike the iterative dynamics of Hénon maps. So it leads to the question: Is 𝒥𝒮±=𝐊𝒮±int(𝒦𝒮±)\mathcal{J}_{\mathcal{S}}^{\pm}=\mathbf{K}_{\mathcal{S}}^{\pm}\setminus\textsf{int}(\mathcal{K}_{\mathcal{S}}^{\pm}) or 𝒥𝒮±=𝒦𝒮±𝐊𝒮±\mathcal{J}_{\mathcal{S}}^{\pm}=\partial\mathcal{K}_{\mathcal{S}}^{\pm}\cup\partial\mathbf{K}_{\mathcal{S}}^{\pm}?

Refer to caption
Refer to caption
Figure 1. 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+}, 𝐊𝒮+\mathbf{K}_{\mathcal{S}}^{+} 𝒥𝒮+\mathcal{J}_{\mathcal{S}}^{+}, 𝐊𝒮\mathbf{K}_{\mathcal{S}}^{-} and 𝒥𝒮\mathcal{J}_{\mathcal{S}}^{-}, with 𝒦𝒮=\mathcal{K}_{\mathcal{S}}^{-}=\emptyset

The shaded region in Figure 1 corresponds to the weak filled Julia sets 𝐊𝒮±\mathbf{K}_{\mathcal{S}}^{\pm}, with the lighter shade representing the Fatou components and the darker shades representing the Julia sets 𝒥𝒮±\mathcal{J}_{\mathcal{S}}^{\pm} contained in them, respectively.

Finally we conclude this section, by observing the following crucial fact which will be very important for further computations

Lemma 2.8.

Let R>R𝒮R>R_{\mathcal{S}}, where R𝒮>0R_{\mathcal{S}}>0 is the radius of filtration for the semigroup 𝒮\mathcal{S} and CC be a compact subset of 2\mathbb{C}^{2}.

  1. (i)

    Then there exists a positive integer kC1k_{C}\geq 1 such that h(C)VRint(VR+)h(C)\subset V_{R}\cup\textsf{int}(V_{R}^{+}) for every h𝒢kh\in\mathscr{G}_{k} and kkC.k\geq k_{C}.

  2. (ii)

    Then there exists a positive integer k~C1\tilde{k}_{C}\geq 1 such that h1(C)VRint(VR)h^{-1}(C)\subset V_{R}\cup\textsf{int}(V_{R}^{-}) for every h𝒢kh\in\mathscr{G}_{k} and kk~C.k\geq\tilde{k}_{C}.

Proof.

Suppose the statement (i) is not true, i.e., there exist a sequence {zn}C\{z_{n}\}\subset C and a sequence {hn}\{h_{n}\} with hn𝒢knh_{n}\in\mathscr{G}_{k_{n}}, knk_{n}\to\infty as nn\to\infty such that hn(zn)VRh_{n}(z_{n})\in V_{R}^{-}. Then by Lemma 2.3, znhn1(VR)VRknz_{n}\in h_{n}^{-1}(V_{R}^{-})\subset V_{R_{k_{n}}}^{-}. As RknR_{k_{n}}\to\infty for nn\to\infty, zn\|z_{n}\|\to\infty. This contradicts the fact that {zn}\{z_{n}\} is contained in the compact set CC.

A similar argument works for part (ii). ∎

3. Proof of Theorem 1.3

In this section, we will first complete the proof of Theorem 1.3 and consequently observe a few important corollaries. To begin, let us recall the definition of the sequence of plurisubharmonic functions {Gk}\{G_{k}\}, introduced in Section 1,

Gk+(z)=1Dkh𝒢klog+h(z) and Gk(z)=1Dkh𝒢klog+h1(z),\displaystyle G_{k}^{+}(z)=\frac{1}{D^{k}}\sum_{h\in\mathscr{G}_{k}}\log^{+}\|h(z)\|\text{ and }G_{k}^{-}(z)=\frac{1}{D^{k}}\sum_{h\in\mathscr{G}_{k}}\log^{+}\|h^{-1}(z)\|,

where D=D𝒢D=D_{\mathscr{G}} is the total degree corresponding to the generating set 𝒢\mathscr{G} of the semigroup 𝒮.\mathcal{S}. First, we note the following straightforward consequence from the results in Section 2.

Lemma 3.1.

Fix an RR𝒮R\geq R_{\mathcal{S}}, the radius of filtration for the semigroup 𝒮.\mathcal{S}. Then

  • for a compact set C+𝐔𝒮+C^{+}\subset\mathbf{U}_{\mathcal{S}}^{+} there exists a positive integer 𝒩C+1\mathcal{N}_{C^{+}}\geq 1 such that h(C+)VR+h(C^{+})\subset V_{R}^{+} whenever h𝒢kh\in\mathscr{G}_{k}, k𝒩C+k\geq\mathcal{N}_{C^{+}};

  • for a compact set C𝐔𝒮C^{-}\subset\mathbf{U}_{\mathcal{S}}^{-} there exists a positive integer 𝒩C1\mathcal{N}_{C^{-}}\geq 1 such that h1(C)VRh^{-1}(C^{-})\subset V_{R}^{-} whenever h𝒢kh\in\mathscr{G}_{k}, k𝒩C.k\geq\mathcal{N}_{C^{-}}.

Proof.

By Corollary 2.6 and the proof of Proposition 2.5, there exists 𝒩C+,𝒩C1\mathcal{N}_{C^{+}},\mathcal{N}_{C^{-}}\geq 1 such that

C+h𝒢kh1(VR+),Ch𝒢lh(VR) whenever k𝒩C+ and l𝒩C.C^{+}\subset\bigcap_{h\in\mathscr{G}_{k}}h^{-1}(V_{R}^{+}),\;C^{-}\subset\bigcap_{h\in\mathscr{G}_{l}}h(V_{R}^{-})\text{ whenever }k\geq\mathcal{N}_{C^{+}}\text{ and }l\geq\mathcal{N}_{C^{-}}.\qed

Now we are ready to present the proof of Theorem 1.3, which involves some steps.

Proof of Theorem 1.3.

Let RR𝒮R\geq R_{\mathcal{S}} be as in Lemma 3.1.

Step 1: The sequence {Gk±}\{G_{k}^{\pm}\} converges uniformly to a pluriharmonic function on VR±V_{R}^{\pm}.

Suppose (x,y)VR+(x,y)\in V_{R}^{+} and the constants 0<m<1<M0<m<1<M be as assumed in Remark 2.4. Then

logm+dilog|y|<log|π2𝖧i(x,y)|<logM+dilog|y|\displaystyle\log m+{d_{i}}\log|y|<\log|\pi_{2}\circ\mathsf{H}_{i}(x,y)|<\log M+{d_{i}}\log|y| (3.1)

for every 1in01\leq i\leq n_{0}. Let h𝒢kh\in\mathscr{G}_{k}. Then h=𝖧j1𝖧jkh=\mathsf{H}_{j_{1}}\circ\cdots\circ\mathsf{H}_{j_{k}} where 1jin01\leq j_{i}\leq n_{0} for every 1ik1\leq i\leq k and the degree of hh (denoted by dhd_{h}) is given by the product dj1djnd_{j_{1}}\ldots d_{j_{n}}. Also, recall D=d1++dn0D=d_{1}+\cdots+d_{n_{0}}, the total degree of 𝒮=𝒢\mathcal{S}=\langle\mathscr{G}\rangle. Now by (2.1), for (x,y)VR+(x,y)\in V_{R}^{+} and k1k\geq 1,

Gk+(x,y)=h𝒢klog|π2h(x,y)|Dk.\displaystyle G_{k}^{+}(x,y)=\sum_{h\in\mathscr{G}_{k}}\frac{\log|\pi_{2}\circ h(x,y)|}{D^{k}}.

Hence Gk+G_{k}^{+} is pluriharmonic on VR+V_{R}^{+} for every k1k\geq 1 and

Gk+1+(x,y)=i=1n0h𝒢klog|π2𝖧ih(x,y)|Dk+1.G_{k+1}^{+}(x,y)=\sum_{i=1}^{n_{0}}\sum_{h\in\mathscr{G}_{k}}\frac{\log|\pi_{2}\circ\mathsf{H}_{i}\circ h(x,y)|}{D^{k+1}}.

Thus from (3.1)

Gk+1+(x,y)(n0D)k+1logM+i=1n0h𝒢kdilog|π2h(x,y)|Dk+1(n0D)k+1logM+Gk+(x,y).\displaystyle G_{k+1}^{+}(x,y)\leq\bigg{(}\frac{n_{0}}{D}\bigg{)}^{k+1}\log M+\sum_{i=1}^{n_{0}}\sum_{h\in\mathscr{G}_{k}}\frac{d_{i}\log|\pi_{2}\circ h(x,y)|}{D^{k+1}}\leq\bigg{(}\frac{n_{0}}{D}\bigg{)}^{k+1}\log M+G_{k}^{+}(x,y).

Similarly, by using the left inequality of (3.1) we have

(n0D)k+1logm+Gk+(x,y)Gk+1+(x,y)(n0D)k+1logM+Gk+(x,y).\displaystyle\bigg{(}\frac{n_{0}}{D}\bigg{)}^{k+1}\log m+G_{k}^{+}(x,y)\leq G_{k+1}^{+}(x,y)\leq\bigg{(}\frac{n_{0}}{D}\bigg{)}^{k+1}\log M+G_{k}^{+}(x,y).

Since D2n0D\geq 2n_{0}, the above inequality reduces to

|Gk+1+(x,y)Gk+(x,y)|(12)k+1M0\displaystyle|G_{k+1}^{+}(x,y)-G_{k}^{+}(x,y)|\leq\bigg{(}\frac{1}{2}\bigg{)}^{k+1}M_{0} (3.2)

where M0=max{|logm|,|logM|}M_{0}=\max\{|\log m|,|\log M|\} and (x,y)VR+.(x,y)\in V_{R}^{+}. Thus the sequence {Gk+}\{G_{k}^{+}\} is uniformly Cauchy on VR+V_{R}^{+}, and hence it converges uniformly to a pluriharmonic function G𝒢+G_{\mathscr{G}}^{+} on VR+.V_{R}^{+}. A similar argument on VRV_{R}^{-}, proves the same for G𝒢.G_{\mathscr{G}}^{-}.

Step 2: The sequence {Gk±}\{G_{k}^{\pm}\} converges uniformly to the pluriharmonic function G𝒢±G_{\mathscr{G}}^{\pm} on compact subsets of 𝐔𝒮±\mathbf{U}^{\pm}_{\mathcal{S}}, respectively.

As noted earlier similar arguments work on 𝐔𝒮\mathbf{U}_{\mathcal{S}}^{-}, so we complete the proof only for 𝐔𝒮+.\mathbf{U}_{\mathcal{S}}^{+}. Let CC be a compact subset of 𝐔𝒮+.\mathbf{U}^{+}_{\mathcal{S}}. By Lemma 3.1, there exists 𝒩C1\mathcal{N}_{C}\geq 1 such that h(C)VR+h(C)\subset V_{R}^{+} for every h𝒢kh\in\mathscr{G}_{k}, k𝒩C.k\geq\mathcal{N}_{C}. Note that 𝒢k\mathscr{G}_{k} has n0kn_{0}^{k} elements. Let Ch=h(C)VR+C_{h}=h(C)\subset V_{R}^{+} (by Lemma 3.1) for every h𝒢𝒩C.h\in\mathscr{G}_{\mathcal{N}_{C}}. Thus for zCz\in C

Gk+(z)=1D𝒩Ch𝒢𝒩CGk𝒩C+(h(z)).{G_{k}^{+}}(z)=\frac{1}{D^{\mathcal{N}_{C}}}\sum_{h\in\mathscr{G}_{\mathcal{N}_{C}}}{G_{k-{\mathcal{N}_{C}}}^{+}}(h(z)).

whenever k>𝒩C.k>\mathcal{N}_{C}. Now by Step 1, {Gk𝒩C+}\{G_{k-{\mathcal{N}_{C}}}^{+}\} is convergent on every ChC_{h}. Hence {Gk+}\{G_{k}^{+}\} converges uniformly on C𝐔𝒮+C\subset\mathbf{U}_{\mathcal{S}}^{+} to a pluriharmonic function and this completes Step 2. Thus, G𝒢±G_{\mathscr{G}}^{\pm} are (respectively) pluriharmonic on the 𝐔𝒮±\mathbf{U}_{\mathcal{S}}^{\pm}.

Also G𝒢±G_{\mathscr{G}}^{\pm} are pluriharmonic on int(𝒦𝒮±)\textsf{int}(\mathcal{K}_{\mathcal{S}}^{\pm}), as G𝒢±G_{\mathscr{G}}^{\pm} are identically zero in here. Next, for z02z_{0}\in\mathbb{C}^{2} and for k1k\geq 1, we define the following subsets of 𝒮\mathcal{S}, dependent on z0z_{0} as

𝒮b(z0)={h𝒮:h(z0)VRVR},\displaystyle\mathcal{S}^{b}(z_{0})=\{h\in\mathcal{S}:h(z_{0})\in V_{R}\cup V_{R}^{-}\},\; 𝒢kb(z0)=𝒮b(z0)𝒢k and\displaystyle\mathscr{G}^{b}_{k}(z_{0})=\mathcal{S}^{b}(z_{0})\cap\mathscr{G}_{k}\text{ and }
𝒮u(z0)={h𝒮:h(z0)int(VR+)},\displaystyle\mathcal{S}^{u}(z_{0})=\{h\in\mathcal{S}:h(z_{0})\in\textsf{int}(V_{R}^{+})\},\; 𝒢ku(z0)=𝒮u(z0)𝒢k.\displaystyle\mathscr{G}^{u}_{k}(z_{0})=\mathcal{S}^{u}(z_{0})\cap\mathscr{G}_{k}.

Note that by (2.1), the following inequality about the cardinality of the sets 𝒢kb(z0)\mathscr{G}^{b}_{k}(z_{0}) and 𝒢ku(z0)\mathscr{G}^{u}_{k}(z_{0}) is immediate for every k1k\geq 1

𝒢k+1b(z0)n0(𝒢kb(z0)),n0(𝒢ku(z0))𝒢k+1u(z0).\displaystyle\sharp\mathscr{G}^{b}_{k+1}(z_{0})\leq n_{0}(\sharp\mathscr{G}^{b}_{k}(z_{0})),\;n_{0}(\sharp\mathscr{G}^{u}_{k}(z_{0}))\leq\sharp\mathscr{G}^{u}_{k+1}(z_{0}). (3.3)

Now for z0𝐔𝒮+z_{0}\in\mathbf{U}_{\mathcal{S}}^{+}, there exists kz01k_{z_{0}}\geq 1 such that 𝒢kb(z0)=\mathscr{G}^{b}_{k}(z_{0})=\emptyset, for kkz0k\geq k_{z_{0}}. Otherwise from Lemma 2.8, there exists kz01k_{z_{0}}\geq 1 such that h(z0)VR¯h(z_{0})\in\overline{V_{R}} whenever h𝒢kb(z0)h\in\mathscr{G}^{b}_{k}(z_{0}) and kkz0k\geq k_{z_{0}}. Thus consider the following sequences of functions {Gkb}\{G^{b}_{k}\} and {Gku}\{G^{u}_{k}\} defined on 2\mathbb{C}^{2} as

Gkb(z)=h𝒢kb(z)log+h(z)Dk and Gku(z)=h𝒢ku(z)log+h(z)Dk.\displaystyle G^{b}_{k}(z)=\sum_{h\in{\mathscr{G}^{b}_{k}(z)}}\frac{\log^{+}\|h(z)\|}{D^{k}}\text{ and }G^{u}_{k}(z)=\sum_{h\in\mathscr{G}^{u}_{k}(z)}\frac{\log^{+}\|h(z)\|}{D^{k}}. (3.4)
Remark 3.2.

Note that Gk(z)=Gkb(z)+Gku(z)G_{k}(z)=G^{b}_{k}(z)+G^{u}_{k}(z) for every z2z\in\mathbb{C}^{2} and k1k\geq 1. Also if CC is a compact subset of 2\mathbb{C}^{2} then h(C)VRh(C)\subset V_{R} whenever h𝒢kbh\in\mathscr{G}_{k}^{b}, kkCk\geq k_{C}, as obtained in Lemma 2.8.

Step 3: The sequence {Gkb}\{G^{b}_{k}\} converges uniformly to zero on every compact subset C2.C\subset\mathbb{C}^{2}.

Since CC is a compact subset of 2\mathbb{C}^{2} from Lemma 2.8, for kkCk\geq k_{C}, we have

sup{h(z):zC and h𝒢kb(z)}R.\sup\big{\{}\|h(z)\|:z\in C\text{ and }h\in\mathscr{G}^{b}_{k}(z)\big{\}}\leq R.

Now, by (3.3) for every z2z\in\mathbb{C}^{2}, 𝒢kb(z)𝒢k=n0k.\sharp\mathscr{G}^{b}_{k}(z)\leq\sharp\mathscr{G}_{k}=n_{0}^{k}. Hence by the above claim for zCz\in C

Gkb(z)(n0D)klogR(12)klogR0G^{b}_{k}(z)\leq\bigg{(}\frac{n_{0}}{D}\bigg{)}^{k}\log R\leq\bigg{(}\frac{1}{2}\bigg{)}^{k}\log R\to 0

as k.k\to\infty. This completes Step 3.

Step 4: For every z0𝐊𝒮+𝒦𝒮+z_{0}\in\mathbf{K}_{\mathcal{S}}^{+}\setminus\mathcal{K}_{\mathcal{S}}^{+} there exist a constant M~>1\widetilde{M}>1 and a positive integer z01\ell_{z_{0}}\geq 1 such that for kz0,k\geq\ell_{z_{0}},

Gku(z0)M~(n0D)k+1Gk+1u(z0)(n0D)k+1M~+Gku(z0).\displaystyle G^{u}_{k}(z_{0})-\widetilde{M}\bigg{(}\frac{n_{0}}{D}\bigg{)}^{k+1}\leq G^{u}_{k+1}(z_{0})\leq\bigg{(}\frac{n_{0}}{D}\bigg{)}^{k+1}\widetilde{M}+G^{u}_{k}(z_{0}). (3.5)

Since z0𝐊𝒮+𝒦𝒮+z_{0}\in\mathbf{K}_{\mathcal{S}}^{+}\setminus\mathcal{K}_{\mathcal{S}}^{+} there exists a sequence {hn}𝒮\{h_{n}\}\subset\mathcal{S} such that hn(z0)\|h_{n}(z_{0})\|\to\infty as n.n\to\infty. In particular 𝒮u(z0)\mathcal{S}^{u}(z_{0})\neq\emptyset, i.e., by (3.3) there exist positive integers kz0,Nz01k_{z_{0}},N_{z_{0}}\geq 1 such that

1𝒢kz0u(z0)=Nz0<n0kz0.\displaystyle 1\leq\sharp\mathscr{G}_{k_{z_{0}}}^{u}(z_{0})=N_{z_{0}}<n_{0}^{k_{z_{0}}}.

Now note that from (2.1) it follows that 𝖧ih(z0)VR+\mathsf{H}_{i}\circ h(z_{0})\in V_{R}^{+} whenever h𝒢kz0uh\in\mathscr{G}_{k_{z_{0}}}^{u} and 1in0.1\leq i\leq n_{0}. Hence by (3.3) for every kkz0k\geq k_{z_{0}} we have

n0kkz0Nz0𝒢ku(z0)<n0k.\displaystyle n_{0}^{k-k_{z_{0}}}N_{z_{0}}\leq\sharp\mathscr{G}_{k}^{u}(z_{0})<n_{0}^{k}. (3.6)

As 0<m<10<m<1, from (3.6), (3.4) and (3.1) it follows that for every kkz0k\geq k_{z_{0}},

Gk+1u(z0)\displaystyle G^{u}_{k+1}(z_{0}) 1Dk+1i=1n0h𝒢ku(z0)log𝖧ih(z0)=i=1n0h𝒢ku(z0)log𝖧ih(z0)Dk+1,\displaystyle\geq\frac{1}{D^{k+1}}\sum_{i=1}^{n_{0}}\sum_{h\in\mathscr{G}^{u}_{k}(z_{0})}\log\|\mathsf{H}_{i}\circ h(z_{0})\|=\sum_{i=1}^{n_{0}}\sum_{h\in\mathscr{G}^{u}_{k}(z_{0})}\frac{\log\|\mathsf{H}_{i}\circ h(z_{0})\|}{D^{k+1}},
i=1n0h𝒢ku(z0)dilogh(z0)+logmDk+1=h𝒢ku(z0)logh(z0)Dk+n0(𝒢ku(z0))Dk+1logm\displaystyle\geq\sum_{i=1}^{n_{0}}\sum_{h\in\mathscr{G}^{u}_{k}(z_{0})}\frac{d_{i}\log\|h(z_{0})\|+\log m}{D^{k+1}}=\sum_{h\in\mathscr{G}^{u}_{k}(z_{0})}\frac{\log\|h(z_{0})\|}{D^{k}}+\frac{n_{0}(\sharp\mathscr{G}^{u}_{k}(z_{0}))}{D^{k+1}}\log m
Gku(z0)+(n0D)k+1logm.\displaystyle\geq G_{k}^{u}(z_{0})+\bigg{(}\frac{n_{0}}{D}\bigg{)}^{k+1}\log m.

Now from Lemma 2.8, there exists kz01k^{\prime}_{z_{0}}\geq 1 such that for h𝒢kb(z0)h\in\mathscr{G}_{k}^{b}(z_{0}), h(z0)R\|h(z_{0})\|\leq R whenever kkz0.k\geq k^{\prime}_{z_{0}}. Let z0=max{kz0,kz0}\ell_{z_{0}}=\max\{k_{z_{0}},k^{\prime}_{z_{0}}\} and B=max{𝖧i(z):zVR+1¯,1in0}B=\max\{\|\mathsf{H}_{i}(z)\|:z\in\overline{V_{R+1}},1\leq i\leq n_{0}\}. Further, for every k,l1k,l\geq 1, we introduce the following subsets 𝒢k,lu(z0)\mathscr{G}^{u}_{k,l}(z_{0}) of 𝒮\mathcal{S} defined as

𝒢k,lu(z0)={h1h2:h1𝒢l and h2𝒢ku(z0)}.\mathscr{G}^{u}_{k,l}(z_{0})=\{h_{1}\circ h_{2}:h_{1}\in\mathscr{G}_{l}\text{ and }h_{2}\in\mathscr{G}^{u}_{k}(z_{0})\}.

Now by (2.1), 𝒢k,lu(z0)𝒢k+lu(z0)\mathscr{G}^{u}_{k,l}(z_{0})\subset\mathscr{G}^{u}_{k+l}(z_{0}) and 𝒢k,lu(z0)=n0l(𝒢ku(z0))\sharp\mathscr{G}^{u}_{k,l}(z_{0})=n_{0}^{l}(\sharp\mathscr{G}^{u}_{k}(z_{0})) for kz0.k\geq\ell_{z_{0}}. Let h𝒢k+1u𝒢k,1uh\in\mathscr{G}^{u}_{k+1}\setminus\mathscr{G}^{u}_{k,1} then h=𝖧ih~h=\mathsf{H}_{i}\circ\tilde{h} for some h~𝒢kb\tilde{h}\in\mathscr{G}^{b}_{k} and 1in0.1\leq i\leq n_{0}. Thus by the above assumption, |π2h(z0)|B|\pi_{2}\circ h(z_{0})|\leq B. Since

(𝒢k+1u(z0)𝒢k,1u(z0))n0k+1n0(𝒢ku(z0))\sharp(\mathscr{G}^{u}_{k+1}(z_{0})\setminus\mathscr{G}^{u}_{k,1}(z_{0}))\leq n_{0}^{k+1}-n_{0}(\sharp\mathscr{G}^{u}_{k}(z_{0}))

and

Gk+1u(z0)=1Dk+1h𝒢k,1u(z0)logh(z0)+1Dk+1h𝒢k+1u(z0)𝒢k,1u(z0)logh(z0),\displaystyle G^{u}_{k+1}(z_{0})=\frac{1}{D^{k+1}}\sum_{h\in\mathscr{G}^{u}_{k,1}(z_{0})}\log\|h(z_{0})\|+\frac{1}{D^{k+1}}\sum_{h\in\mathscr{G}^{u}_{k+1}(z_{0})\setminus\mathscr{G}^{u}_{k,1}(z_{0})}{\log\|h(z_{0})\|},

we have the following inequations (as before)

Gk+1u(z0)\displaystyle G^{u}_{k+1}(z_{0}) =1Dk+1i=1n0h𝒢ku(z0)log𝖧ih(z0)+1Dk+1h𝒢k+1u(z0)𝒢k,1u(z0)logh(z0)\displaystyle=\frac{1}{D^{k+1}}\sum_{i=1}^{n_{0}}\sum_{h\in\mathscr{G}^{u}_{k}(z_{0})}\log\|\mathsf{H}_{i}\circ h(z_{0})\|+\frac{1}{D^{k+1}}\sum_{h\in\mathscr{G}^{u}_{k+1}(z_{0})\setminus\mathscr{G}^{u}_{k,1}(z_{0})}{\log\|h(z_{0})\|}
i=1n0h𝒢ku(z0)dilogh(z0)+logMDk+1+1Dk+1h𝒢k+1u(z0)𝒢k,1u(z0)logh(z0)\displaystyle\leq\sum_{i=1}^{n_{0}}\sum_{h\in\mathscr{G}^{u}_{k}(z_{0})}\frac{d_{i}\log\|h(z_{0})\|+\log M}{D^{k+1}}+\frac{1}{D^{k+1}}\sum_{h\in\mathscr{G}^{u}_{k+1}(z_{0})\setminus\mathscr{G}^{u}_{k,1}(z_{0})}{\log\|h(z_{0})\|}
Gku(z0)+n0(𝒢ku(z0))Dk+1logM+(𝒢k+1u(z0)𝒢k,1u(z0))Dk+1logB\displaystyle\leq G^{u}_{k}(z_{0})+\frac{n_{0}(\sharp\mathscr{G}^{u}_{k}(z_{0}))}{D^{k+1}}\log M+\frac{\sharp(\mathscr{G}^{u}_{k+1}(z_{0})\setminus\mathscr{G}^{u}_{k,1}(z_{0}))}{D^{k+1}}\log B
Gku(z0)+(n0D)k+1M~\displaystyle\leq G^{u}_{k}(z_{0})+\bigg{(}\frac{n_{0}}{D}\bigg{)}^{k+1}\widetilde{M}

where M~=max{|logM|,|logB|,|logm|}\widetilde{M}=\max\{|\log M|,|\log B|,|\log m|\}. This completes the proof of Step 4.

Since Gku0G^{u}_{k}\equiv 0 on 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+} and Gkb0G^{b}_{k}\equiv 0 on 𝐔𝒮+\mathbf{U}_{\mathcal{S}}^{+}, by Steps 2, 3 and 4, Gk+G_{k}^{+} converges pointwise to a non-negative function G𝒢+G_{\mathscr{G}}^{+} on 2\mathbb{C}^{2} which is identically zero on 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+} and pluriharmonic on 𝐔𝒮+.\mathbf{U}_{\mathcal{S}}^{+}.

Step 5: G𝒢+G_{\mathscr{G}}^{+} is non-negative, continuous and plurisubharmonic on 𝒰𝒮+.\mathcal{U}_{\mathcal{S}}^{+}.

Let CC be a compact subset of 𝒰𝒮+\mathcal{U}_{\mathcal{S}}^{+} and zC.z\in C. By the same reasoning as in Step 5, there exist kz,Nz1k_{z},N_{z}\geq 1 such that

1𝒢kzu(z)=Nzn0kz.\displaystyle 1\leq\sharp\mathscr{G}_{k_{z}}^{u}(z)=N_{z}\leq n_{0}^{k_{z}}.

Since 𝗂𝗇𝗍(VR+)\mathsf{int}(V_{R}^{+}) is open, there exists δz>0\delta_{z}>0 such that the closed ball B(z;δz)¯\overline{B(z;\delta_{z})} is contained in 𝒰𝒮+\mathcal{U}_{\mathcal{S}}^{+} and h(B(z;δz))𝗂𝗇𝗍(VR+)h(B(z;\delta_{z}))\subset\mathsf{int}(V_{R}^{+}) for every h𝒢kzu(z)h\in\mathscr{G}^{u}_{k_{z}}(z), i.e., for every ξB(z;δz)\xi\in B(z;\delta_{z})

𝒢kzu(z)𝒢kzu(ξ) and Nz𝒢kzu(ξ)n0kz.\mathscr{G}^{u}_{k_{z}}(z)\subseteq\mathscr{G}^{u}_{k_{z}}(\xi)\text{ and }N_{z}\leq\sharp\mathscr{G}_{k_{z}}^{u}(\xi)\leq n_{0}^{k_{z}}.

Since B(z;δz)¯\overline{B(z;\delta_{z})} is compact, by Lemma 2.8, there exists kz1k^{\prime}_{z}\geq 1, such that h(ξ)VRh(\xi)\in V_{R} whenever h𝒢kb(ξ)h\in\mathscr{G}^{b}_{k}(\xi) for every kkzk\geq k^{\prime}_{z} and ξB(z;δz)¯.\xi\in\overline{B(z;\delta_{z})}. Hence by exactly similar argument as in the proof of Step 4, for every kmax{kz,kz}k\geq\max\{k_{z},k^{\prime}_{z}\}

Gku(ξ)M~(n0D)k+1Gk+1u(ξ)(n0D)k+1M~+Gku(ξ)\displaystyle G^{u}_{k}(\xi)-\widetilde{M}\bigg{(}\frac{n_{0}}{D}\bigg{)}^{k+1}\leq G^{u}_{k+1}(\xi)\leq\bigg{(}\frac{n_{0}}{D}\bigg{)}^{k+1}\widetilde{M}+G^{u}_{k}(\xi)

where M~\widetilde{M} is as chosen in Step 4. Now for a given ϵ>0\epsilon>0, there exist a positive integer zumax{kz,kz}\ell^{u}_{z}\geq\max\{k_{z},k^{\prime}_{z}\} sufficiently large, such that for every p,qzup,q\geq\ell^{u}_{z} and for every ξB(z;δz)¯\xi\in\overline{B(z;\delta_{z})}

Gpu(ξ)Gqu(ξ)(12)zuM~ϵ2.\displaystyle\|G^{u}_{p}(\xi)-G^{u}_{q}(\xi)\|\leq\bigg{(}\frac{1}{2}\bigg{)}^{\ell^{u}_{z}}\widetilde{M}\leq\frac{\epsilon}{2}.

As CC is compact, there exists a finite collection {ziC:1iNC}\{z_{i}\in C:1\leq i\leq N_{C}\} such that CiB(zi;δzi).C\subset\cup_{i}B(z_{i};\delta_{z_{i}}). Let C~=iB(zi;δzi)¯\widetilde{C}=\overline{\cup_{i}B(z_{i};\delta_{z_{i}})}, then by Lemma 2.8, there exists Cϵmax{ziu:1iNC}\ell_{C}^{\epsilon}\geq\max\{\ell^{u}_{z_{i}}:1\leq i\leq N_{C}\} sufficiently large, such that for every ξC~\xi\in\widetilde{C} and for every p,qCϵp,q\geq\ell^{\epsilon}_{C}

Gpb(ξ)Gqb(ξ)<(12)Cϵ1logR<ϵ2.\displaystyle\|G^{b}_{p}(\xi)-G^{b}_{q}(\xi)\|<\bigg{(}\frac{1}{2}\bigg{)}^{\ell^{\epsilon}_{C}-1}\log R<\frac{\epsilon}{2}.

Thus for every ξC\xi\in C and p,qCϵp,q\geq\ell_{C}^{\epsilon}

Gp+(ξ)Gq+(ξ)<ϵ,\|G^{+}_{p}(\xi)-G^{+}_{q}(\xi)\|<\epsilon,

i.e., {Gk+}\{G_{k}^{+}\} is uniformly Cauchy on the compact set C.C. Further as {Gk+}\{G_{k}^{+}\} is a non-negative, subharmonic and continuous sequence of functions on 𝒰𝒮+\mathcal{U}_{\mathcal{S}}^{+}, so is G𝒢+.G_{\mathscr{G}}^{+}.

Step 6: G𝒢+G_{\mathscr{G}}^{+} is non-negative, non-constant, continuous and plurisubharmonic on 2.\mathbb{C}^{2}.

Note that if 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+} is empty there is nothing to proof. So we assume 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+}\neq\emptyset. By (3.2), it follows that on VR+V_{R}^{+}

G𝒢+log|y|M0i=0(n0D)i.\|G_{\mathscr{G}}^{+}-\log|y|\|\leq M_{0}\sum_{i=0}^{\infty}\bigg{(}\frac{n_{0}}{D}\bigg{)}^{i}.

Thus for (x,y)VR+(x,y)\in V_{R}^{+} with |y||y| sufficiently large, G𝒢+(x,y)G_{\mathscr{G}}^{+}(x,y) is both positive and non-constant. Now from Step 5, G𝒢+G_{\mathscr{G}}^{+} is continuous on 𝒰𝒮+\mathcal{U}_{\mathcal{S}}^{+} and it is identically zero on 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+}. Hence to establish the continuity of G𝒢+G_{\mathscr{G}}^{+}, it is sufficient to prove G𝒢+G_{\mathscr{G}}^{+} is continuous on 𝒦𝒮+=𝒰𝒮+.\partial\mathcal{K}_{\mathcal{S}}^{+}=\partial\mathcal{U}_{\mathcal{S}}^{+}.

We will prove it by contradiction, so suppose there exists a sequence {zk}𝒰𝒮+\{z_{k}\}\subset\mathcal{U}_{\mathcal{S}}^{+} such that G𝒢+(zk)>c>0G_{\mathscr{G}}^{+}(z_{k})>c>0 for every k1k\geq 1 and zkz0𝒦𝒮+.z_{k}\to z_{0}\in\partial\mathcal{K}_{\mathcal{S}}^{+}. Choose k01k_{0}\geq 1, large enough such that

c>M~2k0i=1(n0D)i,c>\frac{\widetilde{M}}{2^{k_{0}}}\sum_{i=1}^{\infty}\bigg{(}\frac{n_{0}}{D}\bigg{)}^{i},

where M~\widetilde{M} is as obtained in Step 4. Also, we may assume for every k1k\geq 1,

sup{h(zk):h𝒢k0b(zk)}R.\sup\big{\{}\|h(z_{k})\|:h\in\mathscr{G}^{b}_{k_{0}}(z_{k})\big{\}}\leq R.

Claim: For every k1k\geq 1, 𝒢k0u(zk).\mathscr{G}_{k_{0}}^{u}(z_{k})\neq\emptyset.

If not, then 𝒢k0u(zl)=\mathscr{G}_{k_{0}}^{u}(z_{l})=\emptyset for some fixed l1l\geq 1, i.e., 𝒢ku(zl)=\mathscr{G}_{k}^{u}(z_{l})=\emptyset for every 1kk0.1\leq k\leq k_{0}. However, there exists kl>k0k_{l}>k_{0} such that 𝒢klu(zl)\mathscr{G}_{k_{l}}^{u}(z_{l})\neq\emptyset, as G𝒢+(zl)>0G_{\mathscr{G}}^{+}(z_{l})>0. Let klk^{\prime}_{l} be the least of all such numbers, i.e., 𝒢ku(zl)\mathscr{G}_{k}^{u}(z_{l})\neq\emptyset for every kkl>k0.k\geq k_{l}^{\prime}>k_{0}. Thus for every kklk\geq k_{l}

Gku(zl)i=kl(n0D)iM~<c.G^{u}_{k}(z_{l})\leq\sum_{i=k_{l}}^{\infty}\bigg{(}\frac{n_{0}}{D}\bigg{)}^{i}\widetilde{M}<c.

As Gkb(zl)0G^{b}_{k}(z_{l})\to 0, the above thus proves that G𝒢+(zl)<cG_{\mathscr{G}}^{+}(z_{l})<c, which contradicts the assumption on the sequence {zk}\{z_{k}\}. Hence the claim follows.

Since 𝒢k0u(zk)𝒢k0\mathscr{G}_{k_{0}}^{u}(z_{k})\subset\mathscr{G}_{k_{0}} for every k1k\geq 1 and 𝒢k0\sharp\mathscr{G}_{k_{0}} is finite, there are only finitely many subsets of 𝒢k0\mathscr{G}_{k_{0}}. Thus there exists a subsequence {zkn}\{z_{k_{n}}\} of {zk}\{z_{k}\} such that the sets 𝒢k0u(zkn)\mathscr{G}_{k_{0}}^{u}(z_{k_{n}}) are equal for every n1.n\geq 1. Define the sequence {hl}\{h_{l}\} as hl=𝖧1lhh_{l}=\mathsf{H}_{1}^{l}\circ h for some h𝒢k0u(zkn)h\in\mathscr{G}_{k_{0}}^{u}(z_{k_{n}}). Since h(zkn)VR+h(z_{k_{n}})\in V_{R}^{+}, hl(zkn)VRl+h_{l}(z_{k_{n}})\in V_{R_{l}}^{+}, where {Rl}\{R_{l}\} is as obtained in Lemma 2.3, for every l,n1l,n\geq 1. As zknz0z_{k_{n}}\to z_{0}, this implies hl(z0)VRl+¯h_{l}(z_{0})\in\overline{V_{R_{l}}^{+}}, in particular it contradicts that z0𝒦𝒮+=𝒰𝒮+.z_{0}\in\partial\mathcal{K}_{\mathcal{S}}^{+}=\partial\mathcal{U}_{\mathcal{S}}^{+}.

Hence for every sequence {zk}𝒰𝒮+\{z_{k}\}\in\mathcal{U}^{+}_{\mathcal{S}} and zkz0z_{k}\to z_{0}, G𝒢+(zk)0G^{+}_{\mathscr{G}}(z_{k})\to 0, which consequently proves G𝒢+G_{\mathscr{G}}^{+} is continuous on 2.\mathbb{C}^{2}. Finally, as G𝒢+G_{\mathscr{G}}^{+} coincides with its upper semicontinuous regularisation of G𝒢+G_{\mathscr{G}}^{+} and satisfies the sub-mean value property on 𝒦𝒮+\partial\mathcal{K}_{\mathcal{S}}^{+}, G𝒢+G_{\mathscr{G}}^{+} is plurisubharmonic on 2.\mathbb{C}^{2}.

Similarly replicating Step 3,4,5 and 6 for the semigroup 𝒮\mathcal{S}^{-} with 𝒢\mathscr{G}^{-} as the generating set, gives that G𝒢G_{\mathscr{G}}^{-} is a plurisubharmonic continuous function on 2.\mathbb{C}^{2}.

Corollary 3.3.

There exist constants c𝒢±c_{\mathscr{G}}^{\pm}\in\mathbb{R} such that for (x,y)VR±(x,y)\in V_{R}^{\pm} (respectively).

G𝒢+(x,y)=log|y|+O(1) and G𝒢(x,y)=log|x|+O(1).G_{\mathscr{G}}^{+}(x,y)=\log|y|+O(1)\text{ and }G_{\mathscr{G}}^{-}(x,y)=\log|x|+O(1).
Proof.

It follows directly from the equation (3.2), in the proof Theorem 1.3, i.e., on VR+V_{R}^{+}

M0i=1k(n0D)i+log|y|Gk+(x,y)M0i=1k(n0D)i+log|y|.-M_{0}\sum_{i=1}^{k}\bigg{(}\frac{n_{0}}{D}\bigg{)}^{i}+\log|y|\leq G_{k}^{+}(x,y)\leq M_{0}\sum_{i=1}^{k}\bigg{(}\frac{n_{0}}{D}\bigg{)}^{i}+\log|y|.

Similarly the analogue to (3.2) on VRV_{R}^{-} gives the result for G𝒢.G_{\mathscr{G}}^{-}.

Corollary 3.4.

The functions G𝒢±>0G_{\mathscr{G}}^{\pm}>0 restricted to 𝒰𝒮±\mathcal{U}^{\pm}_{\mathcal{S}} (respectively).

Proof.

From Corollary 3.3 and Lemma 2.3, for z𝒰𝒮+z\in\mathcal{U}^{+}_{\mathcal{S}} there exists nz1n_{z}\geq 1 and hz𝒢nzh^{z}\in\mathscr{G}_{n_{z}} such that hz(z)VR+h^{z}(z)\in V_{R}^{+} and G𝒢+(hz(z))>0.G_{\mathscr{G}}^{+}(h^{z}(z))>0. Then from Corollary 1.4 we have G𝒢+(z)1DnzG𝒢+(hz(z))>0.\displaystyle G_{\mathscr{G}}^{+}(z)\geq\frac{1}{D^{n_{z}}}G^{+}_{\mathscr{G}}(h^{z}(z))>0.

Remark 3.5.

The above corollary also proves 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+} is pseudoconcave, provided it is non-empty.

4. Proof of Theorem 1.5

Recall from Remark 2.7, the cumulative (positive and negative) Julia sets are contained in the (positive and negative) Julia sets. Our goal in this section is to prove that these two sets are actually equal, by analysing the supports of the positive (1,1)(1,1)-currents on 2\mathbb{C}^{2} defined as

μ𝒢+=12πddcG𝒢+ and μ𝒢=12πddcG𝒢.\mu_{\mathscr{G}}^{+}=\frac{1}{2\pi}dd^{c}G_{\mathscr{G}}^{+}\text{ and }\mu_{\mathscr{G}}^{-}=\frac{1}{2\pi}dd^{c}G_{\mathscr{G}}^{-}.

Note that the above fact is also true for the dynamics of semigroups of rational functions in 1\mathbb{P}^{1}, and is proved using Ahlfor’s covering lemma in [20], a tool not available in higher dimensions. Instead we will use the Harnack’s inequality for harmonic functions and the properties of the dynamical Green’s functions G𝒢±.G_{\mathscr{G}}^{\pm}.

Theorem 4.1.

Let {G~k}\{\widetilde{G}_{k}\} be the sequence of plurisubharmonic functions on 2\mathbb{C}^{2} defined as

G~k±(z)=1Dkh𝒢kGh±(h±(z))=1Dkh𝒢kdhGh±(z).\widetilde{G}_{k}^{\pm}(z)=\frac{1}{D^{k}}\sum_{h\in\mathscr{G}_{k}}G_{h}^{\pm}(h^{\pm}(z))=\frac{1}{D^{k}}\sum_{h\in\mathscr{G}_{k}}d_{h}G_{h}^{\pm}(z).

Then the sequences {G~k±}\{\widetilde{G}_{k}^{\pm}\} converge uniformly to G𝒢±G_{\mathscr{G}}^{\pm} (respectively) on compact sets of 2\mathbb{C}^{2}.

Thus, we first observe Lemma 4.2, which is an extension of the proof of Theorem 1.3.

Lemma 4.2.

The sequences {Gk±}\{G_{k}^{\pm}\} — as in Section 3 — converge uniformly on compact subsets of 2\mathbb{C}^{2} to G𝒢±G_{\mathscr{G}}^{\pm}, respectively.

Proof.

As before, we only prove the convergence of {Gk+}\{G_{k}^{+}\} to G𝒢+G_{\mathscr{G}}^{+} by generalising the proof of Lemma 8.3.4 from [26]. The convergence of {Gk}\{G_{k}^{-}\} will follow likewise. Let CC be any compact subset of 2\mathbb{C}^{2}, then

  • If CC is contained in 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+}, Gk+(z)=Gkb(z)G_{k}^{+}(z)=G_{k}^{b}(z) (as defined in the proof of Theorem 1.3). Hence by Remark 3.2, Gk+G𝒢+G_{k}^{+}\to G_{\mathscr{G}}^{+} uniformly on CC;

  • If CC is contained in 𝒰𝒮+\mathcal{U}_{\mathcal{S}}^{+}, from the proof of Step 5 of Theorem 1.3, it follows that Gk+G𝒢+G_{k}^{+}\to G_{\mathscr{G}}^{+} uniformly on CC.

So we assume C𝒰𝒮+C\cap\mathcal{U}_{\mathcal{S}}^{+}\neq\emptyset and C𝒦𝒮+C\cap\mathcal{K}_{\mathcal{S}}^{+}\neq\emptyset and let zC.z\in C. By the above facts, for a given ϵ>0\epsilon>0 there exists k11k_{1}\geq 1, such that |Gkb(z)|<ϵ|G_{k}^{b}(z)|<\epsilon and h(z)<R\|h(z)\|<R for every h𝒢kb(z)h\in\mathscr{G}^{b}_{k}(z) whenever kk1.k\geq k_{1}. In particular, for zC𝒦𝒮+z\in C\cap\mathcal{K}_{\mathcal{S}}^{+} and kk1k\geq k_{1} |Gk+(z)G𝒢+(z)|<ϵ.|G_{k}^{+}(z)-G^{+}_{\mathscr{G}}(z)|<{\epsilon}. Recall the sets {𝒰k}\{\mathcal{U}_{k}\} from Proposition 2.5 defined as

𝒰k=h𝒢kh1(U0)\mathcal{U}_{k}=\bigcup_{h\in\mathscr{G}_{k}}h^{-1}(U_{0})

for every k1k\geq 1. Also from (2.2), we have 𝒰k¯𝒰k+1𝒰k+1¯\overline{\mathcal{U}_{k}}\subset\mathcal{U}_{k+1}\subset\overline{\mathcal{U}_{k+1}} and 𝒰𝒮+=k=1𝒰k\mathcal{U}_{\mathcal{S}}^{+}=\cup_{k=1}^{\infty}\mathcal{U}_{k}. Let

Ck:=C(𝒰k𝒰k1)C_{k}:=C\cap({\mathcal{U}_{k}}\setminus{\mathcal{U}_{k-1}})

for every k1k\geq 1. Since C𝒰𝒮+C\cap\mathcal{U}_{\mathcal{S}}^{+}\neq\emptyset and 𝒰k¯𝒰k+1\overline{\mathcal{U}_{k}}\subset\mathcal{U}_{k+1}, CkC_{k}’s are non-empty sets for k1k\geq 1, sufficiently large. Also k=1Ck=C𝒰𝒮+\cup_{k=1}^{\infty}C_{k}=C\cap\mathcal{U}_{\mathcal{S}}^{+}. Now for k>k1k>k_{1} and zCkz\in C_{k}, h(z)VRVRh(z)\in V_{R}\cup V_{R}^{-} whenever h𝒢ph\in\mathscr{G}_{p}, 1pk11\leq p\leq k-1, i.e., 𝒢pb(z)=𝒢p\mathscr{G}_{p}^{b}(z)=\mathscr{G}_{p} for all zCk.z\in C_{k}. Thus from Lemma 2.8 and Remark 3.2, for kk large enough, h(z)VRh(z)\in V_{R} whenever zCkz\in C_{k} and h𝒢k1.h\in\mathscr{G}_{k-1}. Let

B=sup{𝖧i(z):zVR+1¯,1in0}.B=\sup\{\|\mathsf{H}_{i}(z)\|:z\in\overline{V_{R+1}},1\leq i\leq n_{0}\}.

So for h𝒢kh\in\mathscr{G}_{k}, h(z)<B\|h(z)\|<B whenever zCkz\in C_{k}, kk sufficiently large. Let l>kl>k and zCkz\in C_{k}, then

Gl+(z)=1Dkh𝒢kGlk+(h(z))n0kDk(logB+logM)M~2k1.G_{l}^{+}(z)=\frac{1}{D^{k}}\sum_{h\in\mathscr{G}_{k}}G_{l-k}^{+}(h(z))\leq\frac{n_{0}^{k}}{D^{k}}(\log B+\log M)\leq\frac{\widetilde{M}}{2^{k-1}}.

where M~=max{|logM|,|logB|,|logm|}\widetilde{M}=\max\{|\log M|,|\log B|,|\log m|\} and m,Mm,M is as obtained in Remark 2.4. Now by continuity of G𝒢+G_{\mathscr{G}}^{+} for ϵ>0\epsilon>0 there exists a neighbourhood WW of C𝒦𝒮+C\cap\partial\mathcal{K}_{\mathcal{S}}^{+} such that |G𝒢+(z)|<ϵ/2|G_{\mathscr{G}}^{+}(z)|<\epsilon/2 for zWz\in W. Further choose k2k1k_{2}\geq k_{1} large enough, such that M~2k<ϵ/4\frac{\widetilde{M}}{2^{k}}<\epsilon/4 for every kk2k\geq k_{2} and

C~k2:=k=k2CkW.\widetilde{C}_{k_{2}}:=\bigcup_{k=k_{2}}^{\infty}C_{k}\subset W.

Then for every zC~k2z\in\widetilde{C}_{k_{2}}, |Gk+(z)G𝒢+(z)|<ϵ|G_{k}^{+}(z)-G_{\mathscr{G}}^{+}(z)|<\epsilon whenever kk2.k\geq k_{2}. Note that

(C𝒰𝒮+)C~k2C𝒰k21¯(C\cap\mathcal{U}_{\mathcal{S}}^{+})\setminus\widetilde{C}_{k_{2}}\subset C\cap\overline{\mathcal{U}_{k_{2}-1}}

and C𝒰k21¯C\cap\overline{\mathcal{U}_{k_{2}-1}} is a compact set contained in 𝒰𝒮+.\mathcal{U}_{\mathcal{S}}^{+}. Hence there k31k_{3}\geq 1 such that for every z(C𝒰𝒮+)C~k2z\in(C\cap\mathcal{U}_{\mathcal{S}}^{+})\setminus\widetilde{C}_{k_{2}}, |Gk+(z)G𝒢+(z)|ϵ.|G_{k}^{+}(z)-G_{\mathscr{G}}^{+}(z)|\leq\epsilon. Thus |Gk+G𝒢+|<ϵ|G^{+}_{k}-G_{\mathscr{G}}^{+}|<\epsilon on CC for kmax{k1,k2,k3}k\geq\max\{k_{1},k_{2},k_{3}\}. ∎

Remark 4.3.

Note that in the proof of Theorem 1.3, we use the pointwise convergence of Gk±G_{k}^{\pm} to prove G𝒢+G_{\mathscr{G}}^{+} is a continuous function. However, proof of Lemma 4.2 uses the continuity of G𝒢+G_{\mathscr{G}}^{+} crucially, to establish the convergence is uniform on compact subsets of 2\mathbb{C}^{2}.

Now we complete

Proof of Theorem 4.1.

We will show that for a given compact set C2C\subset\mathbb{C}^{2} and an ϵ>0\epsilon>0 there exists kCk_{C} such that |G~kGk|C<ϵ/2|\widetilde{G}_{k}-G_{k}|_{C}<\epsilon/2 for every k𝐤Ck\geq\mathbf{k}_{C}, and use Lemma 4.2.

As before for the compact subset CC by Lemma 2.8 and Remark 3.2, there exists k11k_{1}\geq 1 such that h(z)<R\|h(z)\|<R whenever h𝒢kb(z)h\in\mathscr{G}^{b}_{k}(z) for every zCz\in C and kk1.k\geq k_{1}. Now choose k2k_{2} such that M~2k2<ϵ\frac{\widetilde{M}}{2^{k-2}}<\epsilon for kk2k\geq k_{2} where M~\widetilde{M} as obtained in Lemma 4.2. Thus, for every zCz\in C with h𝒢kb(z)h\in\mathscr{G}^{b}_{k}(z) where k𝐤C=max{k1,k2}k\geq\mathbf{k}_{C}=\max\{k_{1},k_{2}\}

log+h(z)dh<M~2k<ϵ2.\displaystyle\frac{\log^{+}\|h(z)\|}{d_{h}}<\frac{\widetilde{M}}{2^{k}}<\frac{\epsilon}{2}. (4.1)

Step 1: For k𝐤Ck\geq\mathbf{k}_{C}, Gh+(z)<ϵ2{G_{h}^{+}(z)}<\frac{\epsilon}{2} for every zCz\in C and h𝒢kb(z)h\in\mathscr{G}_{k}^{b}(z), i.e., |log+h(z)dhGh+(z)|<ϵ.\displaystyle\bigg{|}\frac{\log^{+}\|h(z)\|}{d_{h}}-G_{h}^{+}(z)\bigg{|}<\epsilon.

If zKh+Cz\in K_{h}^{+}\cap C then Gh+(z)=0G_{h}^{+}(z)=0. So suppose zKh+Cz\notin K_{h}^{+}\cap C, i.e., h(z)VRh(z)\in V_{R}. Let h=hkh1(z)h=h_{k}\circ\cdots\circ h_{1}(z), where hi𝒢h_{i}\in\mathscr{G} for every 1ik.1\leq i\leq k. Further let 1lk1\leq l\leq k and nz1n_{z}\geq 1, the minimum positive integers, such that

wz:=hlh1hnz(z)VR+.w_{z}:=h_{l}\circ\ldots\circ h_{1}\circ h^{n_{z}}(z)\in V_{R}^{+}.

In particular, hjh1hn(z)VRh_{j}\circ\ldots\circ h_{1}\circ h^{n}(z)\in V_{R}, whenever 1nnz1\leq n\leq n_{z} and 1jl11\leq j\leq l-1. Then wz=hlh1hnz(z)<B\|w_{z}\|=\|h_{l}\circ\ldots\circ h_{1}\circ h^{n_{z}}(z)\|<B, where BB is as chosen in Lemma 4.2. If 1l<k1\leq l<k, then from (2.1)

mwzdl+1hl+1(wz)Mwzdl+1m\|w_{z}\|^{d_{l+1}}\leq\|h_{l+1}(w_{z})\|\leq M\|w_{z}\|^{d_{l+1}}

where dl+1d_{l+1} is the degree of hl+1h_{l+1}. As M>1M>1, we will consider a more robust bound to the above inequality, i.e., Mwzdl+1<Mwzdl+1M\|w_{z}\|^{d_{l+1}}<\|Mw_{z}\|^{d_{l}+1} and obtain the following

log+hi+lhl+1(wz)<dl+idl+1(logB+logM).\log^{+}{\|h_{i+l}\circ\cdots\circ h_{l+1}(w_{z})\|}<d_{l+i}\ldots d_{l+1}(\log B+\log M).

for 1ikl.1\leq i\leq k-l. By continuing to repeat the same argument, we get that for every j1j\geq 1

log+hj+nz(z)<dhj.dkdl+1(logB+logM)<dhj(logB+logM),\log^{+}{\|h^{j+n_{z}}(z)\|}<d_{h}^{j}.d_{k}\ldots d_{l+1}(\log B+\log M)<d_{h}^{j}(\log B+\log M),

where dh=dkd1d_{h}=d_{k}\ldots d_{1} is the degree of hh. Note that if l=kl=k, then the final bound on the above inequality is anyway true. Since dh2kd_{h}\geq 2^{k}, for every j>1j>1,

log+hj+nz(z)dhj+nzlogB+logMdhnzlogB+logMdhM~2k1<ϵ2,\frac{\log^{+}{\|h^{j+n_{z}}(z)\|}}{d_{h}^{j+n_{z}}}\leq\frac{\log B+\log M}{d_{h}^{n_{z}}}\leq\frac{\log B+\log M}{d_{h}}\leq\frac{\widetilde{M}}{2^{k-1}}<\frac{\epsilon}{2},

and thus Step 1 follows by taking the limit of jj\to\infty and (4.1).

Step 2: For h𝒢ku(z)h\in\mathscr{G}^{u}_{k}(z) and l2l\geq 2

logh(z)dhM~i=1l11dhiloghl(z)dhllogh(z)dh+M~i=1l11dhi.\displaystyle\frac{\log\|h(z)\|}{d_{h}}-\widetilde{M}\sum_{i=1}^{l-1}\frac{1}{d_{h}^{i}}\leq\frac{\log\|h^{l}(z)\|}{d_{h}^{l}}\leq\frac{\log\|h(z)\|}{d_{h}}+\widetilde{M}\sum_{i=1}^{l-1}\frac{1}{d_{h}^{i}}. (4.2)

We will prove the above by induction. Let l=2l=2 then by (2.1)

m1+i=2kdh(i)h(z)h2(z)M1+i=2kdh(i)h(z),\displaystyle{{m}^{1+\sum_{i=2}^{k}d_{h}(i)}}\|h(z)\|\leq\|h^{2}(z)\|\leq{M}^{1+\sum_{i=2}^{k}d_{h}(i)}\|h(z)\|, (4.3)

where d(i)=didkd(i)=d_{i}\cdots d_{k}, 2ik.2\leq i\leq k. Now, applying logarithm and dividing by dh2d_{h}^{2} to the right inequality of the identity (4.3), it follows that

logh2(z)dh2M~dh(i=1k1d1di)+logh(z)dhM~dh(i=1k12i)+logh(z)dhM~dh+logh(z)dh,\frac{\log\|h^{2}(z)\|}{d_{h}^{2}}\leq\frac{\widetilde{M}}{d_{h}}\bigg{(}\sum_{i=1}^{k}\frac{1}{d_{1}\cdots d_{i}}\bigg{)}+\frac{\log\|h(z)\|}{d_{h}}\leq\frac{\widetilde{M}}{d_{h}}\bigg{(}\sum_{i=1}^{k}\frac{1}{2^{i}}\bigg{)}+\frac{\log\|h(z)\|}{d_{h}}\leq\frac{\widetilde{M}}{d_{h}}+\frac{\log\|h(z)\|}{d_{h}},

as M~>max{|logm|,|logM|}\widetilde{M}>\max\{|\log m|,|\log M|\}. A similar argument applied to the left inequality of (4.3) along with the above observation gives

M~dh+logh(z)dhlogh2(z)dh2M~dh+logh(z)dh,-\frac{\widetilde{M}}{d_{h}}+\frac{\log\|h(z)\|}{d_{h}}\leq\frac{\log\|h^{2}(z)\|}{d_{h}^{2}}\leq\frac{\widetilde{M}}{d_{h}}+\frac{\log\|h(z)\|}{d_{h}},

which proves (4.2) for l=2.l=2. Now, assume (4.2) is true for some l2l\geq 2, by above

M~dh+loghl(z)dhloghl+1(z)dh2M~dh+loghl(z)dh.-\frac{\widetilde{M}}{d_{h}}+\frac{\log\|h^{l}(z)\|}{d_{h}}\leq\frac{\log\|h^{l+1}(z)\|}{d_{h}^{2}}\leq\frac{\widetilde{M}}{d_{h}}+\frac{\log\|h^{l}(z)\|}{d_{h}}.

Hence dividing further by dhl1d_{h}^{l-1} and substituting the assumption gives

logh(z)dhM~i=1l1dhiloghl(z)dhlM~dhlloghl+1(z)dhl+1M~dhl+loghl(z)dhllogh(z)dh+M~i=1l1dhi,\frac{\log\|h(z)\|}{d_{h}}-\widetilde{M}\sum_{i=1}^{l}\frac{1}{d_{h}^{i}}\leq\frac{\log\|h^{l}(z)\|}{d_{h}^{l}}-\frac{\widetilde{M}}{d_{h}^{l}}\leq\frac{\log\|h^{l+1}(z)\|}{d_{h}^{l+1}}\leq\frac{\widetilde{M}}{d_{h}^{l}}+\frac{\log\|h^{l}(z)\|}{d_{h}^{l}}\leq\frac{\log\|h(z)\|}{d_{h}}+\widetilde{M}\sum_{i=1}^{l}\frac{1}{d_{h}^{i}},

which proves the induction hypothesis and hence the Step 2.

Thus, by taking limit ll\to\infty on the identity (4.2), we have

|Gh+(z)logh(z)dh|2M~dhM~2k1<ϵ.\bigg{|}G_{h}^{+}(z)-\frac{\log\|h(z)\|}{d_{h}}\bigg{|}\leq\frac{2\widetilde{M}}{d_{h}}\leq\frac{\widetilde{M}}{2^{k-1}}<\epsilon.

Hence for zCz\in C and k𝐤Ck\geq\mathbf{k}_{C}

|G~k+(z)Gk+(z)|h𝒢kdhDk|Gh+(z)logh(z)dh|<ϵ.\bigg{|}\widetilde{G}_{k}^{+}(z)-G_{k}^{+}(z)\bigg{|}\leq\sum_{h\in\mathscr{G}_{k}}\frac{d_{h}}{D^{k}}\bigg{|}G_{h}^{+}(z)-\frac{\log\|h(z)\|}{d_{h}}\bigg{|}<\epsilon.\qed
Corollary 4.4.

Support of μ𝒢±\mu_{\mathscr{G}}^{\pm} is contained in the cumulative Julia sets 𝐉𝒮±.\mathbf{J}_{\mathcal{S}}^{\pm}.

Proof.

Let μk±=12πddcG~k±\mu_{k}^{\pm}=\frac{1}{2\pi}dd^{c}\widetilde{G}_{k}^{\pm} then from Lemma 3.6 of [4], it follows that

 supp (μk±)=h𝒢kJh±.\text{ supp }(\mu_{k}^{\pm})=\bigcup_{h\in\mathscr{G}_{k}}J_{h}^{\pm}.

Let SS be any positive (1,1)(1,1)-form supported in the complement of 𝐉𝒮+\mathbf{J}_{\mathcal{S}}^{+} then μk+(S)=0\mu_{k}^{+}(S)=0 for every k1.k\geq 1. By Theorem 4.1 and Corollary 3.6 of [9], μk+μ𝒢+\mu_{k}^{+}\to\mu^{+}_{\mathscr{G}}, in the sense of currents, i.e., μ𝒢+(S)=0.\mu^{+}_{\mathscr{G}}(S)=0. Hence the proof. A similar argument works for μ𝒢.\mu_{\mathscr{G}}^{-}.

Finally, we are ready to complete

Proof of Theorem 1.5.

Note that by Corollary 4.4, ddc(G𝒢+)=0dd^{c}(G_{\mathscr{G}}^{+})=0 everywhere in the complement of 𝐉𝒮+.\mathbf{J}_{\mathcal{S}}^{+}. Choose any ball 𝔹{\mathbb{B}} contained in 2𝐉𝒮+.\mathbb{C}^{2}\setminus\mathbf{J}_{\mathcal{S}}^{+}. As G𝒢+G_{\mathscr{G}}^{+} is continuous on 𝔹¯\overline{\mathbb{B}}, by uniqueness of solution to the Dirichlet problem it follows that G𝒢+G_{\mathscr{G}}^{+} is pluriharmonic on 𝔹\mathbb{B} and 2𝐉𝒮+.\mathbb{C}^{2}\setminus\mathbf{J}_{\mathcal{S}}^{+}.

Now suppose 𝒥𝒮+supp (μ𝒢+)\mathcal{J}_{\mathcal{S}}^{+}\setminus\text{supp }(\mu_{\mathscr{G}}^{+})\neq\emptyset and z0𝒥𝒮+supp (μ𝒢+)z_{0}\in\mathcal{J}_{\mathcal{S}}^{+}\setminus\text{supp }(\mu_{\mathscr{G}}^{+}). Then there exists r>0r>0, such that the ball of radius rr at z0z_{0}, B(z0;r)(supp (μ𝒢+))c.B(z_{0};r)\subset\big{(}\text{supp }(\mu_{\mathscr{G}}^{+})\big{)}^{c}. Let 0<r<r0<r^{\prime}<r. Since z0𝒥𝒮+z_{0}\in\mathcal{J}_{\mathcal{S}}^{+}, there exists a sequence {hn}𝒮\{h_{n}\}\subset\mathcal{S} that is neither locally uniformly bounded nor uniformly divergent to infinity on B(z0;r).B(z_{0};r^{\prime}). In particular, there exist sequences of points {zn}\{z_{n}\} and {wn}\{w_{n}\} in B(z0;r)B(z_{0};r^{\prime}) such that

hn(zn) is bounded and hn(wn)\|h_{n}(z_{n})\|\text{ is bounded and }\|h_{n}(w_{n})\|\to\infty

as n.n\to\infty. Note that without loss of generality we may assume, the length of hnh_{n}\to\infty as nn\to\infty. Now again, by Lemma 2.8 and Remark 3.2 the above may be modified further as – for nn sufficiently large,

hn(zn)VR and hn(wn)Vrn+,h_{n}(z_{n})\in V_{R}\text{ and }h_{n}(w_{n})\in V_{r_{n}}^{+},

where rnr_{n} is a sequence of positive real numbers that diverges to infinity as n.n\to\infty. Hence

G𝒢+hn(zn)<C0 and G𝒢+hn(wn).\displaystyle G_{\mathscr{G}}^{+}\circ h_{n}(z_{n})<C_{0}\text{ and }G_{\mathscr{G}}^{+}\circ h_{n}(w_{n})\to\infty. (4.4)

Also as G𝒢+G_{\mathscr{G}}^{+} is pluriharmonic on B(z0;r)B(z_{0};r) and plurisubharmonic on 2\mathbb{C}^{2}, by Corollary 1.4 we have G𝒢+hG^{+}_{\mathscr{G}}\circ h is pluriharmonic on B(z0;r)B(z_{0};r) for every h𝒮.h\in\mathcal{S}. Now by Harnack’s inequality (See Theorem 2.5, [19, Page 16]), there exists A>0A>0, a positive constant dependent on z0z_{0}, rr and rr^{\prime}, such that for every harmonic function uu on B(z0;r)B(z_{0};r)

supB(z0;r)u(z)AinfB(z0;r)u(z).\sup_{B(z_{0};r^{\prime})}u(z)\leq A\inf_{B(z_{0};r^{\prime})}u(z).

Hence 0G𝒢+hn(wn)AC00\leq G_{\mathscr{G}}^{+}\circ h_{n}(w_{n})\leq AC_{0} which contradicts (4.4). Hence supp (μ𝒢+)=𝐉𝒮+=𝒥𝒮+\text{supp }(\mu_{\mathscr{G}}^{+})=\mathbf{J}_{\mathcal{S}}^{+}=\mathcal{J}_{\mathcal{S}}^{+}.

Further μ𝒢+\mu_{\mathscr{G}}^{+} is a current of mass 1, follows from Theorem 4.1 and Corollary 3.6 in [9]. Similarly by analysing G𝒢G_{\mathscr{G}}^{-} and μ𝒢\mu_{\mathscr{G}}^{-} as above, we have supp (μ𝒢)=𝒥𝒮\text{supp }(\mu_{\mathscr{G}}^{-})=\mathcal{J}_{\mathcal{S}}^{-}. ∎

Remark 4.5.

Thus by Proposition 3.2 of [9], the measure μ𝒢:=μ𝒢+μ𝒢\mu_{\mathscr{G}}:=\mu_{\mathscr{G}}^{+}\wedge\mu_{\mathscr{G}}^{-} is a probability measure compactly supported on the intersection of the positive and negative Julia sets.

Corollary 4.6.

The Fatou component at infinity of the semigroup 𝒮\mathcal{S} and 𝒮\mathcal{S}^{-}, i.e.,

𝐔𝒮±=int (h𝒮Uh±).\displaystyle\mathbf{U}_{\mathcal{S}}^{\pm}=\text{int }\Big{(}\bigcap_{h\in\mathcal{S}}U_{h}^{\pm}\Big{)}. (4.5)
Proof.

Let h+\mathcal{F}^{+}_{h} denote the Fatou set corresponding to a h𝒮.h\in\mathcal{S}. Note that h+=Uh+hb\mathcal{F}^{+}_{h}=U_{h}^{+}\cup\mathcal{F}^{b}_{h} where Uh+U_{h}^{+} is the component at infinity and hb\mathcal{F}^{b}_{h} are the Fatou components contained in Kh+.K_{h}^{+}. Similarly h=Uhh1b\mathcal{F}^{-}_{h}=U_{h}^{-}\cup\mathcal{F}^{b}_{h^{-1}} where UhU_{h}^{-} is the component at infinity of h1h^{-1} whenever h𝒮.h\in\mathcal{S}. By Theorem1.5, it follows that

𝒮±=2𝐉𝒮+=int (h𝒮h±)=int (h𝒮(Uh±h±b)).\mathcal{F}_{\mathcal{S}}^{\pm}=\mathbb{C}^{2}\setminus\mathbf{J}_{\mathcal{S}}^{+}=\text{int }\Big{(}\bigcap_{h\in\mathcal{S}}\mathcal{F}_{h}^{\pm}\Big{)}=\text{int }\Big{(}\bigcap_{h\in\mathcal{S}}(U_{h}^{\pm}\cup\mathcal{F}_{h^{\pm}}^{b})\Big{)}.

Hence the components at infinity, corresponding to the dynamics of the semigroup 𝒮\mathcal{S} and 𝒮\mathcal{S}^{-} is given by (4.5). ∎

Remark 4.7.

Also note, if 𝒦𝒮+=𝐊𝒮+\mathcal{K}_{\mathcal{S}}^{+}=\mathbf{K}_{\mathcal{S}}^{+}, then Kh1+=Kh2+K_{h_{1}}^{+}=K_{h_{2}}^{+} and by Theorem 5.4 from [23], there exists m,n1m,n\geq 1 such that h1m=h2n.h_{1}^{m}=h_{2}^{n}. Thus 𝐉𝒮±=Jh±\mathbf{J}_{\mathcal{S}}^{\pm}=J_{h}^{\pm} for every h𝒮.h\in\mathcal{S}. In particular from Theorem 5.8, it follows that G𝒢±=Gh±G_{\mathscr{G}}^{\pm}=G_{h}^{\pm} for every h𝒮h\in\mathcal{S}, i.e., the Green’s function is unique.

However, the next corollary proves that the positive and negative Green’s functions obtained corresponding to the semigroup 𝒮\mathcal{S} is generally non-unique (i.e., whenever 𝒦𝒮+𝐊𝒮+\mathcal{K}_{\mathcal{S}}^{+}\subsetneq\mathbf{K}_{\mathcal{S}}^{+}), as a consequence of Corollary 1.4 and Corollary 3.3.

Corollary 4.8.

If 𝒦𝒮+𝐊𝒮+\mathcal{K}_{\mathcal{S}}^{+}\subsetneq\mathbf{K}_{\mathcal{S}}^{+} then the Green’s functions G𝒢±G_{\mathscr{G}}^{\pm} are non-unique and depends on the generating set 𝒢.\mathscr{G}.

Proof.

Suppose not, i.e., let the positive Green’s function be unique corresponding to semigroup 𝒮\mathcal{S}. By Proposition 2.1, it follows that 𝒮\mathcal{S} admits a minimal generating set 𝒢0.\mathscr{G}_{0}. Let 𝒢h=𝒢0h\mathscr{G}_{h}=\mathscr{G}_{0}\cup h, h𝒮𝒢0.h\in\mathcal{S}\setminus\mathscr{G}_{0}. Then 𝒮=𝒢0=𝒢h.\mathcal{S}=\langle\mathscr{G}_{0}\rangle=\langle\mathscr{G}_{h}\rangle. Then by assumption, G𝒢0+=G𝒢h+G_{\mathscr{G}_{0}}^{+}=G_{\mathscr{G}_{h}}^{+} and thus from Corollary 1.4, we have that

(D𝒢0+dh)G𝒢0+(z)=𝖧i𝒢0G𝒢0+(𝖧i(z))+G𝒢0+(h(z))=D𝒢0G𝒢0+(z)+G𝒢0+(h(z))(D_{\mathscr{G}_{0}}+d_{h})G_{\mathscr{G}_{0}}^{+}(z)=\sum_{\mathsf{H}_{i}\in\mathscr{G}_{0}}G_{\mathscr{G}_{0}}^{+}(\mathsf{H}_{i}(z))+G_{\mathscr{G}_{0}}^{+}(h(z))=D_{\mathscr{G}_{0}}G_{\mathscr{G}_{0}}^{+}(z)+G_{\mathscr{G}_{0}}^{+}(h(z))

where D𝒢0D_{\mathscr{G}_{0}} is the total degree of the generating set 𝒢0\mathscr{G}_{0} and dhd_{h} is the degree of h𝒮.h\in\mathcal{S}. Hence G𝒢0+(z)=dhG𝒢0+(h(z))G_{\mathscr{G}_{0}}^{+}(z)=d_{h}G_{\mathscr{G}_{0}}^{+}(h(z)), i.e., G𝒢0+(z)=0G_{\mathscr{G}_{0}}^{+}(z)=0 if zKh+.z\in K_{h}^{+}. But from Corollary 3.4, the above implies Kh+𝒦𝒮+K_{h}^{+}\subset\mathcal{K}_{\mathcal{S}}^{+} for every h𝒮h\in\mathcal{S}. Suppose 𝐊𝒮+𝒦𝒮+\mathbf{K}_{\mathcal{S}}^{+}\setminus\mathcal{K}_{\mathcal{S}}^{+}\neq\emptyset and z0𝐊𝒮+𝒦𝒮+z_{0}\in\mathbf{K}_{\mathcal{S}}^{+}\setminus\mathcal{K}_{\mathcal{S}}^{+}, then there exist sequence {hn}\{h_{n}\} and {h~n}\{\tilde{h}_{n}\} in SS such that both the lengths of hnh_{n} and h~n\tilde{h}_{n} goes to infinity as nn\to\infty. Further by Lemma 2.8 there exists 𝐤01\mathbf{k}_{0}\geq 1 such that for every n𝐤0n\geq\mathbf{k}_{0}, hn(z0)VR+ and h~n(z0)VR.h_{n}(z_{0})\in V_{R}^{+}\text{ and }\tilde{h}_{n}(z_{0})\in V_{R}. Then by Corollary 3.3

G𝒢0+(𝖧1lh𝐤0(z0))=log|π2𝖧1lh𝐤0(z0)|+O(1),G_{\mathscr{G}_{0}}^{+}(\mathsf{H}_{1}^{l}\circ h_{\mathbf{k}_{0}}(z_{0}))=\log|\pi_{2}\circ\mathsf{H}_{1}^{l}\circ h_{\mathbf{k}_{0}}(z_{0})|+O(1),

for every l1l\geq 1. Hence G𝒢0+(𝖧1lh𝐤0(z0))G_{\mathscr{G}_{0}}^{+}(\mathsf{H}_{1}^{l}\circ h_{\mathbf{k}_{0}}(z_{0}))\to\infty as ll\to\infty. Fix l11l_{1}\geq 1, sufficiently large such that G𝒢0+(𝖧1l1h𝐤0(z0))>B~G_{\mathscr{G}_{0}}^{+}(\mathsf{H}_{1}^{l_{1}}\circ h_{\mathbf{k}_{0}}(z_{0}))>\widetilde{B}, where B~=max{G𝒢0+(z):zVR}.\widetilde{B}=\max\{G_{\mathscr{G}_{0}}^{+}(z):z\in V_{R}\}. Let 𝐤1𝐤0\mathbf{k}_{1}\geq\mathbf{k}_{0}

𝗁1=𝖧1l1h𝐤0 and 𝗁2=h~𝐤1\mathsf{h}_{1}=\mathsf{H}_{1}^{l_{1}}\circ h_{\mathbf{k}_{0}}\text{ and }\mathsf{h}_{2}=\tilde{h}_{\mathbf{k}_{1}}

such that the degree of 𝗁2=d𝗁2>d𝗁1=\mathsf{h}_{2}=d_{\mathsf{h}_{2}}>d_{\mathsf{h}_{1}}= degree of 𝗁1\mathsf{h}_{1}. Since we have assumed that G𝒢0+G_{\mathscr{G}_{0}}^{+} is unique, it follows that

(d𝗁2d𝗁1)G𝒢0+(z0)=G𝒢0+(𝗁2(z0))G𝒢0+(𝗁1(z0)),(d_{\mathsf{h}_{2}}-d_{\mathsf{h}_{1}})G_{\mathscr{G}_{0}}^{+}(z_{0})=G_{\mathscr{G}_{0}}^{+}(\mathsf{h}_{2}(z_{0}))-G_{\mathscr{G}_{0}}^{+}(\mathsf{h}_{1}(z_{0})),

i.e., G𝒢0+(z0)<0G_{\mathscr{G}_{0}}^{+}(z_{0})<0, which is a contradiction! Thus 𝐊𝒮+=𝒦𝒮+\mathbf{K}_{\mathcal{S}}^{+}=\mathcal{K}_{\mathcal{S}}^{+}.

Now, if the negative Green’s function is unique, similar argument as above will imply 𝒦𝒮=𝐊𝒮\mathcal{K}_{\mathcal{S}}^{-}=\mathbf{K}_{\mathcal{S}}^{-}. Hence Kh1=Kh2K_{h_{1}}^{-}=K_{h_{2}}^{-} for every h1,h2𝒮h_{1},h_{2}\in\mathcal{S}. Now by Remark 4.7, the positive Green’s function will also be unique, i.e., 𝒦𝒮+=𝐊𝒮+\mathcal{K}_{\mathcal{S}}^{+}=\mathbf{K}_{\mathcal{S}}^{+}, which is again a contradiction! ∎

5. Equidistributed projective currents and proof of Corollary 1.6

Recall that every polynomial map g:22g:\mathbb{C}^{2}\to\mathbb{C}^{2}, i.e., g(x,y)=(g1(x,y),g2(x,y))g(x,y)=(g_{1}(x,y),g_{2}(x,y)) where g1g_{1} and g2g_{2} are polynomials in xx and yy, extends as a rational map g¯\bar{g} on 2.\mathbb{P}^{2}. Further in the homogeneous coordinates of 2\mathbb{P}^{2}, it is defined as

g¯[x:y:z]=[zdg1(xz,yz):zdg2(xz,yz):zd]\bar{g}[x:y:z]=\bigg{[}z^{d}g_{1}\Big{(}\frac{x}{z},\frac{y}{z}\Big{)}:z^{d}g_{2}\Big{(}\frac{x}{z},\frac{y}{z}\Big{)}:z^{d}\bigg{]}

where d=max{degree of g1, degree of g2}.d=\max\{\,\text{degree of }g_{1},\text{ degree of }g_{2}\,\}. Now for any map h𝒮h\in\mathcal{S}, h1h^{-1} is also a polynomial map. Hence both hh and h1h^{-1} extends as rational maps on 2\mathbb{P}^{2}, in the homogeneous coordinates. Further the degree of π2h\pi_{2}\circ h is strictly greater than π1h\pi_{1}\circ h and π2h(x,y)=ydh+ l.o.t.\pi_{2}\circ h(x,y)=y^{d_{h}}+\text{ l.o.t.} Hence the indeterminancy point of the rational map h¯\bar{h} in 2\mathbb{P}^{2} (for every h𝒮h\in\mathcal{S}) is I+=[1:0:0].I^{+}=[1:0:0]. A similar argument gives that the indeterminancy point of h1¯\overline{h^{-1}} is I=[0:1:0].I^{-}=[0:1:0]. Let 𝒮¯\overline{\mathcal{S}} and 𝒮¯\overline{\mathcal{S}^{-}} be the family of the rational maps on 2\mathbb{P}^{2} defined as

𝒮¯={h¯:h𝒮} and 𝒮¯={h1¯:h𝒮}.\overline{\mathcal{S}}=\{\bar{h}:h\in\mathcal{S}\}\text{ and }\overline{\mathcal{S}^{-}}=\{\overline{h^{-1}}:h\in\mathcal{S}\}.

Next, we study the dynamics of the above families in 2\mathbb{P}^{2} and generalise a few facts from [12]. Note that the line at infinity, except the point I+I^{+}, i.e., L+={[x:y:z]2:z=0}I+L_{\infty}^{+}=\{[x:y:z]\in\mathbb{P}^{2}:z=0\}\setminus I^{+} is contracted to II^{-} by every h¯𝒮¯.\bar{h}\in\overline{\mathcal{S}}. Similarly the line at infinity, except the point II^{-}, i.e., L={[x:y:z]2:z=0}IL_{\infty}^{-}=\{[x:y:z]\in\mathbb{P}^{2}:z=0\}\setminus I^{-} is contracted to I+I^{+} by every h¯𝒮¯\bar{h}\in\overline{\mathcal{S}^{-}}. Also, VR+V_{R}^{+} (respectively VRV_{R}^{-}) lies in the basis of attraction of II^{-} (respectively I+I^{+}) for every h¯𝒮¯\bar{h}\in\overline{\mathcal{S}} (respectively for every g¯𝒮¯\bar{g}\in\overline{\mathcal{S}^{-}}). Hence II^{-} is attracting fixed point for every h¯𝒮¯\bar{h}\in\overline{\mathcal{S}} and I+I^{+} is attracting fixed point for every g¯𝒮¯.\bar{g}\in\overline{\mathcal{S}^{-}}. Thus, we define the following sets.

  • 𝐔~𝒮+=int(h𝒮U~h+)\displaystyle\widetilde{\mathbf{U}}_{\mathcal{S}}^{+}=\text{int}\Big{(}\bigcap_{h\in\mathcal{S}}\widetilde{U}_{h}^{+}\Big{)}, where U~h+\widetilde{U}_{h}^{+}, where U~h+\widetilde{U}_{h}^{+} is the basin of attraction of II^{-} for h¯.\bar{h}.

  • 𝐔~𝒮=int(h𝒮U~h)\displaystyle\widetilde{\mathbf{U}}_{\mathcal{S}}^{-}=\text{int}\Big{(}\bigcap_{h\in\mathcal{S}}\widetilde{U}_{h}^{-}\Big{)}, where U~h\widetilde{U}_{h}^{-} , where U~h\widetilde{U}_{h}^{-} is the basin of attraction of I+I^{+} for h1¯.\overline{h^{-1}}.

Proposition 5.1.

The sets 𝐔~𝒮±2=𝐔𝒮±.\widetilde{\mathbf{U}}_{\mathcal{S}}^{\pm}\cap\mathbb{C}^{2}=\mathbf{U}_{\mathcal{S}}^{\pm}. Also, the closure of the sets 𝐊𝒮±\mathbf{K}_{\mathcal{S}}^{\pm} in 2\mathbb{P}^{2} is given by 𝐊𝒮±¯=𝐊𝒮±I±.\overline{\mathbf{K}_{\mathcal{S}}^{\pm}}=\mathbf{K}_{\mathcal{S}}^{\pm}\cup I^{\pm}.

Proof.

Let 𝒰~𝒮±\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm} be the basin of attraction of II^{\mp} for the family 𝒮¯\overline{\mathcal{S}} and 𝒮¯\overline{\mathcal{S}^{-}} in 2I±\mathbb{P}^{2}\setminus I^{\pm}, i.e.,

𝒰~𝒮+={𝐳¯2I+: a neighbourhood W of 𝐳¯ such that hn¯|WI for every {hn}𝒮}\widetilde{\mathcal{U}}_{\mathcal{S}}^{+}=\{\overline{\mathbf{z}}\in\mathbb{P}^{2}\setminus I^{+}:\exists\text{ a neighbourhood }W\text{ of }\overline{\mathbf{z}}\text{ such that }\overline{h_{n}}_{|W}\to I^{-}\text{ for every }\{h_{n}\}\subset\mathcal{S}\}

and

𝒰~𝒮={𝐳¯2I: a neighbourhood W of 𝐳¯ such that hn1¯|WI+ for every {hn}𝒮}.\widetilde{\mathcal{U}}_{\mathcal{S}}^{-}=\{\overline{\mathbf{z}}\in\mathbb{P}^{2}\setminus I^{-}:\exists\text{ a neighbourhood }W\text{ of }\overline{\mathbf{z}}\text{ such that }\overline{h_{n}^{-1}}_{|W}\to I^{+}\text{ for every }\{h_{n}\}\subset\mathcal{S}\}.

Observe that by definition, if (x,y)𝐔𝒮±(x,y)\in\mathbf{U}_{\mathcal{S}}^{\pm}, then [x:y:1]𝒰~𝒮±[x:y:1]\in\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm}. In particular 𝐔𝒮±𝒰~𝒮±2.\mathbf{U}_{\mathcal{S}}^{\pm}\subset\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm}\cap\mathbb{C}^{2}. Now for any point 𝐳¯0=[x0:y0:z0]L+\overline{\mathbf{z}}_{0}=[x_{0}:y_{0}:z_{0}]\in L_{\infty}^{+}, z0=0z_{0}=0 and |y0|0|y_{0}|\neq 0. Hence for every h𝒮h\in\mathcal{S}, h¯(𝐳¯0)=h¯[x0:y0:0]=[0:1:0]\bar{h}(\overline{\mathbf{z}}_{0})=\bar{h}[x_{0}:y_{0}:0]=[0:1:0] is immediate.

Claim: There exist open sets W±W^{\pm} containing L±L_{\infty}^{\pm} which is contained 𝒰~𝒮±\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm}, respectively.

Case 1: Suppose |x0|<|y0||x_{0}|<|y_{0}|, then choose a neighbourhood W𝐳¯0W_{\overline{\mathbf{z}}_{0}} of 𝐳¯0\overline{\mathbf{z}}_{0} such that |x|<|y||x|<|y| for every 𝐳¯=[x:y:z]W𝐳¯0\overline{\mathbf{z}}=[x:y:z]\in W_{\overline{\mathbf{z}}_{0}} and |z|<R1|y||z|<R^{-1}|y| if z0z\neq 0, where R>R𝒮R>R_{\mathcal{S}}, the radius of filtration as in Lemma 2.3. Hence for 𝐳¯W𝐳¯0L+\overline{\mathbf{z}}\in W_{\overline{\mathbf{z}}_{0}}\setminus L_{\infty}^{+}, 𝐳¯=[x:y:1]\overline{\mathbf{z}}=[x:y:1] such that (x,y)VR+(x,y)\in V_{R}^{+}, i.e, h¯(𝐳¯)[0:1:0]\bar{h}(\overline{\mathbf{z}})\to[0:1:0] as length of hh tends to infinity.

Case 2: Otherwise, there exists some α>1\alpha>1 such that |x0|<α|y0||x_{0}|<\alpha|y_{0}|. Note, we need to choose an appropriate neighbourhood of 𝐳¯0\overline{\mathbf{z}}_{0} contained in 𝒰~𝒮±.\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm}. We will do so in the light of Remark 5.3, which is a consequence of the following modification of Lemma 2.2 from [4].

Lemma 5.2.

Let H(x,y)=(y,p(y)ax)H(x,y)=(y,p(y)-ax) where pp is a polynomial of degree dH2d_{H}\geq 2 and a0.a\neq 0. Also, let RH>0R_{H}>0 be the radius of filtration for HH as obtained in Lemma 2.2 of [4]. For R>0R>0 and α>1\alpha>1 we define the following sets as

Vα,R+={(x,y)2:|x|α|y|,|y|α1R},Vα,R={(x,y)2:|x|α|y|,|y|>α1R}V_{\alpha,R}^{+}=\{(x,y)\in\mathbb{C}^{2}:|x|\leq\alpha|y|,|y|\geq\alpha^{-1}R\},\;V_{\alpha,R}^{-}=\{(x,y)\in\mathbb{C}^{2}:|x|\geq\alpha|y|,|y|>\alpha^{-1}R\}

and

V~α,R={(x,y)2:|y|α|x|,|x|α1R},V~α,R+={(x,y)2:|y|α|x|,|x|>α1R}.\widetilde{V}_{\alpha,R}^{-}=\{(x,y)\in\mathbb{C}^{2}:|y|\leq\alpha|x|,|x|\geq\alpha^{-1}R\},\;\widetilde{V}_{\alpha,R}^{+}=\{(x,y)\in\mathbb{C}^{2}:|y|\geq\alpha|x|,|x|>\alpha^{-1}R\}.

Then there exists an Rα>αRHR^{\alpha}>\alpha R_{H} such that H(Vα,Rα)VRH+H(V_{\alpha,R^{\alpha}})\subset V_{R_{H}}^{+} and H1(V~α,Rα)VRH.H^{-1}(\widetilde{V}_{\alpha,R^{\alpha}}^{-})\subset V_{R_{H}}^{-}.

Proof.

Note that

VRH+Vα,RH+,Vα,RHVRH,VRHV~α,RH and V~α,RH+VRH+.V_{R_{H}}^{+}\subset V_{\alpha,R_{H}}^{+},\;V_{\alpha,R_{H}}^{-}\subset V_{R_{H}}^{-},\;V_{R_{H}}^{-}\subset\widetilde{V}_{\alpha,R_{H}}^{-}\text{ and }\widetilde{V}_{\alpha,R_{H}}^{+}\subset V_{R_{H}}^{+}.

Also there exists an Rα>αRHR_{\alpha}>\alpha R_{H}, sufficiently large, and constant C1>0C_{1}>0 such that for (x,y)Vα,Rα+(x,y)\in V_{\alpha,R_{\alpha}}^{+}, i.e., |y|=α1R>α1Rα|y|=\alpha^{-1}R>\alpha^{-1}R_{\alpha} and |x|R|x|\leq R

|π2H(x,y)|>αdHC1RdH|a|R>α1R=|π1H(x,y)|.|\pi_{2}\circ H(x,y)|>\alpha^{-d_{H}}C_{1}R^{d_{H}}-|a|R>\alpha^{-1}R=|\pi_{1}\circ H(x,y)|.

Similarly there exists an R~α>αRH\widetilde{R}_{\alpha}>\alpha R_{H}, sufficiently large, and constant C2>0C_{2}>0 such that for (x,y)V~α,R~α(x,y)\in\widetilde{V}_{\alpha,\widetilde{R}_{\alpha}}^{-}, i.e., |x|=α1R>α1R~α|x|=\alpha^{-1}{R}>\alpha^{-1}\widetilde{R}_{\alpha} and |y|R|y|\leq R

|π1H1(x,y)|>αdHC2RdH|a|1R>α1R=|π2H1(x,y)|.|\pi_{1}\circ H^{-1}(x,y)|>\alpha^{-d_{H}}C_{2}R^{d_{H}}-|a|^{-1}R>\alpha^{-1}R=|\pi_{2}\circ H^{-1}(x,y)|.

Let RαR^{\alpha} be the maximum of RαR_{\alpha} and R~α\widetilde{R}_{\alpha}. Then H(Vα,Rα+)VR+H(V^{+}_{\alpha,R^{\alpha}})\subset V_{R}^{+} and H1(V~α,Rα)VR.H^{-1}(\widetilde{V}^{-}_{\alpha,R^{\alpha}})\subset V_{R}^{-}.

Remark 5.3.

By a similar technique as in the proof of Lemma 2.3, the above further assures that Rα>αR𝒮R^{\alpha}>\alpha R_{\mathcal{S}}, the radius of filtration of the semigroup 𝒮\mathcal{S}, such that h(Vα,Rα+)VR+h(V_{\alpha,R^{\alpha}}^{+})\subset V_{R}^{+} and h1(V~α,Rα)VRh^{-1}(\widetilde{V}_{\alpha,R^{\alpha}}^{-})\subset V_{R}^{-} for every h𝒮.h\in\mathcal{S}.

We now choose a neighbourhood W𝐳¯0W_{\overline{\mathbf{z}}_{0}} of 𝐳¯0\overline{\mathbf{z}}_{0} such that |x|<α|y||x|<\alpha|y| for every 𝐳¯=[x:y:z]W𝐳¯0\overline{\mathbf{z}}=[x:y:z]\in W_{\overline{\mathbf{z}}_{0}} and |z|Rα<α|y||z|{R^{\alpha}}<\alpha|y| if z0z\neq 0, where RαR^{\alpha} is as obtained in Remark 5.3. Hence for 𝐳¯W𝐳¯0L+\overline{\mathbf{z}}\in W_{\overline{\mathbf{z}}_{0}}\setminus L_{\infty}^{+}, 𝐳¯=[x:y:1]\overline{\mathbf{z}}=[x:y:1] where (x,y)Vα,Rα+(x,y)\in V^{+}_{\alpha,R^{\alpha}}, i.e, h¯(𝐳¯)[0:1:0]\bar{h}(\overline{\mathbf{z}})\to[0:1:0] as length of hh tends to infinity.

By similar arguments for h1h^{-1}, h𝒮h\in\mathcal{S} there exists an open set WW^{-} containing LL_{\infty}^{-} such that W𝒰~𝒮.W^{-}\subset\widetilde{\mathcal{U}}_{\mathcal{S}}^{-}. Further note that for 𝐳¯𝒰~𝒮±L±\overline{\mathbf{z}}\in\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm}\setminus L_{\infty}^{\pm}, 𝐳¯=[x:y:1]\overline{\mathbf{z}}=[x:y:1] such that (x,y)𝐔𝒮±.(x,y)\in\mathbf{U}_{\mathcal{S}}^{\pm}. Since 2=2L±I±\mathbb{P}^{2}=\mathbb{C}^{2}\sqcup L_{\infty}^{\pm}\sqcup I^{\pm}, 𝒰~𝒮±2=𝒰~𝒮±L±=𝐔𝒮±.\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm}\cap\mathbb{C}^{2}=\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm}\setminus L_{\infty}^{\pm}=\mathbf{U}_{\mathcal{S}}^{\pm}.

Finally, as a consequence of Corollary 4.6 and Proposition 5.5 of [12, Page 28] — which implies U~h±2=Uh±\widetilde{U}_{h}^{\pm}\cap\mathbb{C}^{2}=U_{h}^{\pm} for every h𝒮h\in\mathcal{S} — we can write

𝒰~𝒮±L±=𝒰~𝒮±2=𝐔𝒮±=inth𝒮Uh±=inth𝒮(U~h±2)=inth𝒮(U~h±L±)\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm}\setminus L_{\infty}^{\pm}=\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm}\cap\mathbb{C}^{2}=\mathbf{U}_{\mathcal{S}}^{\pm}=\text{int}\bigcap_{h\in\mathcal{S}}{U}_{h}^{\pm}=\text{int}\bigcap_{h\in\mathcal{S}}(\widetilde{U}_{h}^{\pm}\cap\mathbb{C}^{2})=\text{int}\bigcap_{h\in\mathcal{S}}(\widetilde{U}_{h}^{\pm}\setminus L_{\infty}^{\pm})

But L±W±𝒰~𝒮±L_{\infty}^{\pm}\subset W^{\pm}\subset\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm} and 𝒰~𝒮±U~h±\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm}\subset\widetilde{U}_{h}^{\pm} for every h𝒮h\in\mathcal{S}, i.e., W±W^{\pm} is contained in the interior of (h𝒮U~h)\Big{(}\bigcap_{h\in\mathcal{S}}\widetilde{U}_{h}\Big{)}. Hence the above identity reduces to

𝐔𝒮±=𝒰~𝒮±L±=(inth𝒮U~h±)L±=(inth𝒮U~h±)2=𝐔~𝒮±2.\displaystyle\mathbf{U}_{\mathcal{S}}^{\pm}=\widetilde{\mathcal{U}}_{\mathcal{S}}^{\pm}\setminus L_{\infty}^{\pm}=\Big{(}\text{int}\bigcap_{h\in\mathcal{S}}\widetilde{U}_{h}^{\pm}\Big{)}\setminus L_{\infty}^{\pm}=\Big{(}\text{int}\bigcap_{h\in\mathcal{S}}\widetilde{U}_{h}^{\pm}\Big{)}\cap\mathbb{C}^{2}=\widetilde{\mathbf{U}}_{\mathcal{S}}^{\pm}\cap\mathbb{C}^{2}. (5.1)

Now, since L±𝐔~𝒮±L_{\infty}^{\pm}\subset\widetilde{\mathbf{U}}_{\mathcal{S}}^{\pm}, it follows that 𝐊𝒮±¯𝐊𝒮±I±.\overline{\mathbf{K}_{\mathcal{S}}^{\pm}}\subset\mathbf{K}_{\mathcal{S}}^{\pm}\cup I^{\pm}. Also, Kh±𝐊𝒮±K_{h}^{\pm}\subset\mathbf{K}_{\mathcal{S}}^{\pm} for every h𝒮h\in\mathcal{S} and by Proposition 5.8 in [12, Page 29], I±Kh±¯I^{\pm}\in\overline{K_{h}^{\pm}}. Hence I±𝐊𝒮±¯I^{\pm}\in\overline{\mathbf{K}_{\mathcal{S}}^{\pm}}, which completes the proof. ∎

Remark 5.4.

As a consequence of Proposition 5.1, it follows that the basins of attraction of I±I^{\pm} for the families 𝒮¯\overline{\mathcal{S}} and 𝒮¯\overline{\mathcal{S}^{-}} are 𝐔~𝒮±\widetilde{\mathbf{U}}_{\mathcal{S}}^{\pm}, respectively. Further, the closure of the positive and negative Julia sets 𝐉𝒮±\mathbf{J}_{\mathcal{S}}^{\pm} in 2\mathbb{P}^{2}, i.e., 𝐉𝒮±¯=𝐉𝒮±I±.\overline{\mathbf{J}_{\mathcal{S}}^{\pm}}=\mathbf{J}_{\mathcal{S}}^{\pm}\cup I^{\pm}. Hence from Skoda-El-Mir extension Theorem (see [9]), the (1,1)(1,1)-currents μ𝒢±\mu_{\mathscr{G}}^{\pm} extends by 0 to positive closed (1,1)(1,1)-currents (will also be denoted by μ𝒢±\mu_{\mathscr{G}}^{\pm}) on 2\mathbb{P}^{2}. Now as G𝒢±G_{\mathscr{G}}^{\pm} are the logarithmic potential of μ𝒢±\mu_{\mathscr{G}}^{\pm} restricted to 2\mathbb{C}^{2} — from the observation in Example 3.7 in [12] — the functions g𝒢±(z)=G𝒢±(z)12log(z2+1)g_{\mathscr{G}}^{\pm}(z)=G_{\mathscr{G}}^{\pm}(z)-\frac{1}{2}\log(\|z\|^{2}+1) are the quasi-potentials corresponding to the currents μ𝒢±\mu_{\mathscr{G}}^{\pm} on 2\mathbb{P}^{2} (respectively).

Remark 5.5.

Note that the functions g𝒢±(z)g_{\mathscr{G}}^{\pm}(z) is uniformly bounded and pluriharmonic on VR𝒮±V_{R_{\mathcal{S}}}^{\pm} (respectively) from Corollary 3.3. Hence for every k1k\geq 1, on 𝐔k:=h𝒢kh1(VR+)\mathbf{U}_{k}:=\bigcap_{h\in\mathscr{G}_{k}}h^{-1}(V_{R}^{+})

G𝒢+(z)12log(z2+1)=1Dkh𝒢kG𝒢+(h(z))12log(z2+1) is pluriharmonic.G_{\mathscr{G}}^{+}(z)-\frac{1}{2}\log(\|z\|^{2}+1)=\frac{1}{D^{k}}\sum_{h\in\mathscr{G}_{k}}G_{\mathscr{G}}^{+}(h(z))-\frac{1}{2}\log(\|z\|^{2}+1)\text{ is pluriharmonic.}

Since 𝐔~𝒮+=𝐔𝒮+L+\widetilde{\mathbf{U}}_{\mathcal{S}}^{+}=\mathbf{U}_{\mathcal{S}}^{+}\cup L_{\infty}^{+} is an open set containing L+L_{\infty}^{+}, g𝒢+(z)g_{\mathscr{G}}^{+}(z) extends as a pluriharmonic function on 𝐔~𝒮+\widetilde{\mathbf{U}}_{\mathcal{S}}^{+}. A similar arguments gives g𝒢(z)g_{\mathscr{G}}^{-}(z) extends as a pluriharmonic function on 𝐔~𝒮\widetilde{\mathbf{U}}_{\mathcal{S}}^{-}.

Next, we prove a generalisation of Theorem 6.6 from [12] in our setup.

Proposition 5.6.

Let 𝒱\mathcal{V} be a neighbourhood of II^{-} and {Sk}\{S_{k}\}, a sequence of positive (1,1)(1,1)-closed currents of mass 1 in 2\mathbb{P}^{2} such that each SkS_{k}, k1k\geq 1, admits a quasi-potential uku_{k}, satisfying 0<|uk|A0<|u_{k}|\leq A (a constant) on 𝒱\mathcal{V}. Then there exists c>0c>0 such that for every 𝒞2\mathcal{C}^{2} test (1,1)(1,1)-form ϕ\phi on 2\mathbb{P}^{2}

|μkμ𝒢+,ϕ|ck2kϕ𝒞2, i.e., μk:=1Dkh𝒢kh¯(Sk)μ𝒢+.|\langle\mu_{k}^{*}-\mu_{\mathscr{G}}^{+},\phi\rangle|\leq\frac{ck}{2^{k}}\|\phi\|_{\mathcal{C}^{2}},\text{ i.e., }\mu_{k}^{*}:=\frac{1}{D^{k}}\sum_{h\in\mathscr{G}_{k}}\bar{h}^{*}(S_{k})\to\mu_{\mathscr{G}}^{+}.
Proof.

Note that we may assume that ddcuk=SkωFSdd^{c}u_{k}=S_{k}-\omega_{\text{FS}}, where wFSw_{\text{FS}} is the Fubini-Study (1,1)(1,1)-form on 2\mathbb{P}^{2} and 𝒱\mathcal{V} is a sufficiently small neighbourhood of II^{-} contained in 𝐔~𝒮+\widetilde{\mathbf{U}}_{\mathcal{S}}^{+}. In particular, 𝒱2VR+.\mathcal{V}\cap\mathbb{C}^{2}\subset V_{R}^{+}. Then for z𝒱VR𝒮+z\in\mathcal{V}\cap V_{R_{\mathcal{S}}}^{+}, and by the identities in Section 4, gh+(z)M~g_{h}^{+}(z)\leq\widetilde{M}

gk+(z):=G~k+(z)12log(z2+1)M~ for every h𝒢k,k1.g_{k}^{+}(z):=\widetilde{G}_{k}^{+}(z)-\frac{1}{2}\log(\|z\|^{2}+1)\leq\widetilde{M}\text{ for every }h\in\mathscr{G}_{k},~{}k\geq 1.

Thus by Remarks 5.4 and 5.5, the quasipotentials ukgh+u_{k}-g_{h}^{+}, ukgk+u_{k}-g_{k}^{+} and ukg𝒢+u_{k}-g_{\mathscr{G}}^{+} are uniformly bounded on 𝒱\mathcal{V}, by M~+A\widetilde{M}+A, with gk+g𝒢+g_{k}^{+}\to g_{\mathscr{G}}^{+} uniformly on 𝒱\mathcal{V}. Also by the proof of Lemma 5.2, let α1\alpha\geq 1 be such that supp(Sk)2VRαV~Rα\text{supp}(S_{k})\cap\mathbb{C}^{2}\subset V_{R_{\alpha}}\cup\widetilde{V}_{R_{\alpha}}^{-}, for every k1k\geq 1. Hence supp(h(Sk))2VRαVRα\text{supp}(h^{*}(S_{k}))\cap\mathbb{C}^{2}\subset V_{R_{\alpha}}\cup V_{R_{\alpha}}^{-} for very h𝒮h\in\mathcal{S}. Thus we refine 𝒱\mathcal{V} again, so that 𝒱2VRα+\mathcal{V}\cap\mathbb{C}^{2}\subset V_{R_{\alpha}}^{+} and by continuity

supp(μkμ𝒢+)𝒱=,supp(μkμk+)𝒱=,supp(1dhh¯(Sk)μh+)𝒱=.\text{supp}(\mu_{k}^{*}-\mu_{\mathscr{G}}^{+})\cap\mathcal{V}=\emptyset,~{}\text{supp}(\mu_{k}^{*}-\mu_{k}^{+})\cap\mathcal{V}=\emptyset,~{}\text{supp}\Big{(}\frac{1}{d_{h}}\bar{h}^{*}(S_{k})-\mu_{h}^{+}\Big{)}\cap\mathcal{V}=\emptyset.

for every h𝒮h\in\mathcal{S} and k1.k\geq 1. Hence

|μkμ𝒢+,ϕ|2=|μkμ𝒢+,ϕ|2𝒱 and |μkμk+,ϕ|2=|μkμk+,ϕ|2𝒱\displaystyle|\langle\mu_{k}^{*}-\mu_{\mathscr{G}}^{+},\phi\rangle|_{\mathbb{P}^{2}}=|\langle\mu_{k}^{*}-\mu_{\mathscr{G}}^{+},\phi\rangle|_{\mathbb{P}^{2}\setminus\mathcal{V}}\text{ and }|\langle\mu_{k}^{*}-\mu_{k}^{+},\phi\rangle|_{\mathbb{P}^{2}}=|\langle\mu_{k}^{*}-\mu_{k}^{+},\phi\rangle|_{\mathbb{P}^{2}\setminus\mathcal{V}} (5.2)

Also 𝖧i1(2𝒱)2𝒱\mathsf{H}_{i}^{-1}(\mathbb{P}^{2}\setminus\mathcal{V})\subset\mathbb{P}^{2}\setminus\mathcal{V} as I+I^{+} is a super attracting fixed point for every 𝖧i\mathsf{H}_{i}, 1in0.1\leq i\leq n_{0}. Hence the 𝒞1\mathcal{C}^{1}-norm of every h1¯\overline{h^{-1}} is bounded by 𝖬k\mathsf{M}^{k} for some 𝖬>0\mathsf{M}>0, whenever h𝒢kh\in\mathscr{G}_{k}. Since ukgh+u_{k}-g_{h}^{+}, ukgk+u_{k}-g_{k}^{+} and ukg𝒢+u_{k}-g_{\mathscr{G}}^{+} are d.s.h. functions in 2\mathbb{P}^{2}, by [12, Lemma 3.11] the DSH-norm (see [12, Section 3] for definition) of ukgh+u_{k}-g_{h}^{+}, ukgk+u_{k}-g_{k}^{+} and ukg𝒢+u_{k}-g_{\mathscr{G}}^{+} are uniformly bounded for very k1k\geq 1 and h𝒮.h\in\mathcal{S}. Hence by [12, Lemma 3.13] there exists a constant C00C_{0}\geq 0 such that

|μkμ𝒢+,ϕ|2𝒱=Dkh𝒢k|ukg𝒢+,ddc(ϕh1¯)|C0(n0D1)k(1+log+𝖬4k)ϕ𝒞2\displaystyle|\langle\mu_{k}^{*}-\mu_{\mathscr{G}}^{+},\phi\rangle|_{\mathbb{P}^{2}\setminus\mathcal{V}}={D^{-k}}\sum_{h\in\mathscr{G}_{k}}|\langle u_{k}-g_{\mathscr{G}}^{+},dd^{c}(\phi\circ\overline{h^{-1}})\rangle|\leq C_{0}(n_{0}D^{-1})^{k}(1+\log^{+}\mathsf{M}^{4k})\|\phi\|_{\mathcal{C}^{2}} (5.3)

which completes the proof. ∎

Remark 5.7.

Furthe Lemma 3.11 and Lemma 3.13 of [12] also gives that for C0>0C_{0}>0, where C0C_{0} is as obtained in the above proof of Proposition 5.6,

|μkμk+,ϕ|2𝒱=Dkh𝒢k|ukgh+,ddc(ϕh1¯)|C0(n0D1)k(1+log+𝖬4k)ϕ𝒞2,\displaystyle\bullet~{}|\langle\mu_{k}^{*}-\mu_{k}^{+},\phi\rangle|_{\mathbb{P}^{2}\setminus\mathcal{V}}={D^{-k}}\sum_{h\in\mathscr{G}_{k}}|\langle u_{k}-g_{h}^{+},dd^{c}(\phi\circ\overline{h^{-1}})\rangle|\leq C_{0}(n_{0}D^{-1})^{k}(1+\log^{+}\mathsf{M}^{4k})\|\phi\|_{\mathcal{C}^{2}},
|1dhh¯(Sk)μk+,ϕ|2𝒱=dh1|ukgh+,ddc(ϕh1¯)|C0dh1(1+log+𝖬4k)ϕ𝒞2.\displaystyle\bullet~{}\bigg{|}\langle\frac{1}{d_{h}}\bar{h}^{*}(S_{k})-\mu_{k}^{+},\phi\rangle\bigg{|}_{\mathbb{P}^{2}\setminus\mathcal{V}}={d_{h}^{-1}}|\langle u_{k}-g_{h}^{+},dd^{c}(\phi\circ\overline{h^{-1}})\rangle|\leq C_{0}d_{h}^{-1}(1+\log^{+}\mathsf{M}^{4k})\|\phi\|_{\mathcal{C}^{2}}.

Now, as a direct consequence of the above proposition we observe the following.

Corollary 5.8.

Let SS be a closed positive (1,1)(1,1)-current in 2\mathbb{P}^{2} of mass 1, such that support of SS does not contain the point [0:1:0][0:1:0] and h¯\bar{h} be the extension of hh to 2\mathbb{P}^{2}, h𝒮.h\in\mathcal{S}. Then

limk1D𝒢nh𝒢kh¯(S)μ𝒢+.\lim_{k\to\infty}\frac{1}{D_{\mathscr{G}}^{n}}\sum_{h\in\mathscr{G}_{k}}\bar{h}^{*}(S)\to\mu_{\mathscr{G}}^{+}.
Proof.

Note that if SS is an (1,1)(1,1) positive closed current of mass 1 in 2\mathbb{P}^{2}, and let uu be a quasi-potential associated to SS, i.e., uu is a quasi p.s.h function and ddcu=SωFS.dd^{c}u=S-\omega_{FS}. Then uu is bounded on a neighbourhood of II^{-} and by Proposition 5.6, the proof follows. ∎

Remark 5.9.

Since the analogue of Proposition 5.6, is true for the current μ𝒢\mu_{\mathscr{G}}^{-} as well, if SS is a positive (1,1)(1,1) current of mass 1 on 2\mathbb{P}^{2} then limk1D𝒢kh𝒢kh¯(S)μ𝒢.\displaystyle\lim_{k\to\infty}\frac{1}{D_{\mathscr{G}}^{k}}\sum_{h\in\mathscr{G}_{k}}\bar{h}_{*}(S)\to\mu_{\mathscr{G}}^{-}.

Thus, we conclude the uniqueness of μ𝒢±\mu_{\mathscr{G}}^{\pm} from Corollary 5.8.

Proof of Corollary 1.6.

Let SS be a positive (1,1)(1,1) closed current supported on 𝒥𝒮+\mathcal{J}_{\mathcal{S}}^{+} satisfying property (1.6). Then SS extends across I+I^{+} by zero as a closed (1,1)(1,1) current of mass 11 on 2\mathbb{P}^{2} that does not intersect I.I^{-}. Thus by Theorem 5.8, it follows that on 2\mathbb{C}^{2}

limk1D𝒢kh𝒢kh(S)μ𝒢+.\lim_{k\to\infty}\frac{1}{D_{\mathscr{G}}^{k}}\sum_{h\in\mathscr{G}_{k}}h^{*}(S)\to\mu_{\mathscr{G}}^{+}.

But from (1.6), 1D𝒢kh𝒢kh(S)=S\frac{1}{D_{\mathscr{G}}^{k}}\sum_{h\in\mathscr{G}_{k}}h^{*}(S)=S. Hence S=μ𝒢+.S=\mu_{\mathscr{G}}^{+}. A similar argument for μ𝒢.\mu_{\mathscr{G}}^{-}.

Finally, we end this section with the interpretation of Corollary 5.8 for algebraic varieties.

Corollary 5.10.

Let SS be an affine algebraic variety of codimension 1 in 2\mathbb{C}^{2}, then there exist non-zero constants 𝐜±>0\mathbf{c}^{\pm}>0 such that

limk1D𝒢nh𝒢kh[S]𝐜+μ𝒢+ and limk1D𝒢nh𝒢kh[S]𝐜μ𝒢.\lim_{k\to\infty}\frac{1}{D_{\mathscr{G}}^{n}}\sum_{h\in\mathscr{G}_{k}}h^{*}[S]\to\mathbf{c}^{+}\mu_{\mathscr{G}}^{+}\text{ and }\lim_{k\to\infty}\frac{1}{D_{\mathscr{G}}^{n}}\sum_{h\in\mathscr{G}_{k}}h_{*}[S]\to\mathbf{c}^{-}\mu_{\mathscr{G}}^{-}.
Proof.

Let SS be an algebraic variety of codimension 11 in 2\mathbb{C}^{2}, i.e., S={(x,y)2:p(x,y)=0}S=\{(x,y)\in\mathbb{C}^{2}:p(x,y)=0\} where pp is a polynomial of degree at least 1. Let p(x,y)=a,bpabxaybp(x,y)=\sum_{a,b\in\mathbb{N}}p_{ab}x^{a}y^{b} such that pab=0p_{ab}=0 whenever aa and bb is greater than some fixed positive integer. The degree of pp is 𝐝p=max{a+b:pab0}.\mathbf{d}_{p}=\max\{a+b:p_{ab}\neq 0\}.

Case 1: Let p(x,y)=cpy𝐝p+l.o.t.p(x,y)=c_{p}y^{\mathbf{d}_{p}}+\text{l.o.t.} then SS is a quasi-projective variety of 2\mathbb{P}^{2} of codimension 11 and S¯\bar{S} extends to 2\mathbb{P}^{2} as an analytic variety, that does not contain I.I^{-}. Hence the current of integration of [S¯][\bar{S}] is a closed positive (1,1)(1,1) current of finite mass, say 𝐜+\mathbf{c}^{+} (see [9, Page 140]). Thus from Theorem 5.8 it follows that limk1D𝒢kh𝒢kh[S]𝐜+μ𝒢+.\displaystyle\lim_{k\to\infty}\frac{1}{D_{\mathscr{G}}^{k}}\sum_{h\in\mathscr{G}_{k}}h^{*}[S]\to\mathbf{c}^{+}\mu_{\mathscr{G}}^{+}.

Case 2: For any polynomial pp, a generalisation of Proposition 4.2 in [4] (or Proposition 8.6.7 in [26]) gives that there exists k1k\geq 1 such that ph=php_{h}=p\circ h for all h𝒢kh\in\mathscr{G}_{k}, is as in Case 1, i.e.,

Claim: There exists k01k_{0}\geq 1 such that for every h𝒢kh\in\mathscr{G}_{k} and kk0k\geq k_{0},

ph(x,y)=cphy𝐝p~h+l.o.t.\displaystyle p\circ h(x,y)=c_{p_{h}}y^{\mathbf{d}_{\tilde{p}_{h}}}+\text{l.o.t.} (5.4)

For a positive integer i1i\geq 1, let λi(p)=max{a+ib:pab0}\lambda_{i}(p)=\max\{a+ib:p_{ab}\neq 0\} and

ρi(p)={(a,b):a+ib=λi(p) and pab0},\rho_{i}(p)=\{(a,b):a+ib=\lambda_{i}(p)\text{ and }p_{ab}\neq 0\},

i.e., the terms in the leading part of the polynomial pp with weight i.i. Let HH be a generalised Hénon map of the form (1.3) of degree dHd_{H}. We first note the following result, which is a rephrasing of Lemma 8.6.5 from [26].

Result. For a polynomial p(x,y)=a,bpabxaybp(x,y)=\sum_{a,b\in\mathbb{N}}p_{ab}x^{a}y^{b} the number of elements in the leading term of pHp\circ H in ii weight, i2i\geq 2, i.e., ρi(pH)\sharp\rho_{i}(p\circ H) satisfies the following inequality

ρi(pH)1+ρdH(p)1i.\sharp\rho_{i}(p\circ H)\leq 1+\frac{\sharp\rho_{d_{H}}(p)-1}{i}.

So if 𝖧\mathsf{H} is a map of the form (1.2), of degree sufficiently large, it follows from the above result, that the number of leading terms in any weight ii, i2i\geq 2 of the polynomial p𝖧p\circ\mathsf{H} is 1. Now if HH is a generalised Hénon map of form (1.3), then the degree of p𝖧H(x,y)p\circ\mathsf{H}\circ H(x,y) is λdH(p𝖧)\lambda_{d_{H}}(p\circ\mathsf{H}) and ρdH(p𝖧)=cp𝖧xayb\rho_{d_{H}}(p\circ\mathsf{H})=c_{p_{\mathsf{H}}}x^{a}y^{b} where a+dHb=λdH(p𝖧).a+d_{H}b=\lambda_{d_{H}}(p\circ\mathsf{H}). Hence p𝖧H(x,y)=cp𝖧cHyλdH+l.o.t.p\circ\mathsf{H}\circ H(x,y)=c_{p_{\mathsf{H}}}c^{\prime}_{H}y^{\lambda_{d_{H}}}+\text{l.o.t.} Since for very h𝒢kh\in\mathscr{G}_{k} degree of hh is greater than 2k2^{k}, from Lemma 8.6.6 of [26] there exists k01k_{0}\geq 1 such that the polynomial php\circ h has the desired form (5.4) whenever h𝒢kh\in\mathscr{G}_{k}, kk0.k\geq k_{0}.

Thus, Case 1 applied to every polynomial php_{h}, h𝒢k0h\in\mathscr{G}_{k_{0}}, proves Corollary 5.10 for μ𝒢+.\mu_{\mathscr{G}}^{+}.

6. Green’s functions for non-autonomous sequences in 𝒮\mathcal{S}

Let {hk}𝒮\{h_{k}\}\subset\mathcal{S}, where 𝒮\mathcal{S} is the semigroup of Hénon maps as defined in (1.1). Recall that to study dynamics of the sequence {hk}\{h_{k}\}, one needs to study the behaviour of the sequences {h(k)}\{h(k)\} and {h1(k)}\{h^{-1}(k)\} defined as

h(k):=hkh1 and h1(k)=hk1h11.h(k):=h_{k}\circ\cdots\circ h_{1}\text{ and }h^{-1}(k)=h_{k}^{-1}\circ\cdots\circ h_{1}^{-1}.

Now as each hi,i1h_{i},i\geq 1, is generated by the finitely many elements of 𝒢\mathscr{G}, there exist a sequence {h~k}𝒢\{\tilde{h}_{k}\}\subset\mathscr{G} and a sequence {nk}\{n_{k}\} of positive integers such that for every k1k\geq 1

h(k)=hkh1=h~nkh~1=h~(nk).\displaystyle h(k)=h_{k}\circ\cdots\circ h_{1}=\tilde{h}_{n_{k}}\circ\cdots\circ\tilde{h}_{1}=\tilde{h}(n_{k}). (6.1)

Hence with abuse of notation, we will assume that {hk}𝒢\{h_{k}\}\subset\mathscr{G}, i.e., the elements of the sequence {hk}\{h_{k}\} varies within the finite collection 𝒢={𝖧i:1in0}.\mathscr{G}=\{\mathsf{H}_{i}:1\leq i\leq n_{0}\}. Also analogue of the (positive and negative) escaping sets and the bounded sets for the sequence {hk}\{h_{k}\} is defined as

𝐔{hk}+={z2:h(k)(z) as k},𝐔{hk}={z2:h1(k)(z) as k}\mathbf{U}^{+}_{\{h_{k}\}}=\{z\in\mathbb{C}^{2}:h(k)(z)\to\infty\text{ as }k\to\infty\},\mathbf{U}^{-}_{\{h_{k}\}}=\{z\in\mathbb{C}^{2}:h^{-1}(k)(z)\to\infty\text{ as }k\to\infty\}

and

𝐊{hk}+\displaystyle\mathbf{K}^{+}_{\{h_{k}\}} ={z2:{h(k)(z)} is bounded},𝐊{hk}\displaystyle=\{z\in\mathbb{C}^{2}:\{h(k)(z)\}\text{ is bounded}\},\;\mathbf{K}^{-}_{\{h_{k}\}} ={z2:{h1(k)(z)} is bounded}.\displaystyle=\{z\in\mathbb{C}^{2}:\{h^{-1}(k)(z)\}\text{ is bounded}\}.

Since h(k)𝒢kh(k)\in\mathscr{G}_{k}, by Remark 2.4 the following inequality holds for every (x,y)VR+(x,y)\in V_{R}^{+}, R>R𝒮R>R_{\mathcal{S}} (sufficiently large)

m|y|dk<hk(x,y)=|π2hk(x,y)|<M|y|dk,\displaystyle m|y|^{d_{k}}<\|h_{k}(x,y)\|=|\pi_{2}\circ h_{k}(x,y)|<M|y|^{d_{k}}, (6.2)

where dkd_{k} is the degree of hk.h_{k}. Also for (x,y)VR(x,y)\in V_{R}^{-}

m|x|dk<hk1(x,y)=|π1hk1(x,y)|<M|x|dk.\displaystyle m|x|^{d_{k}}<\|h_{k}^{-1}(x,y)\|=|\pi_{1}\circ h_{k}^{-1}(x,y)|<M|x|^{d_{k}}. (6.3)

Also h(k)(VR+)int(VRk+)h(k)({V_{R}^{+}})\subset\textsf{int}(V_{R_{k}}^{+}) and h(k)1(VR)int(VRk)h(k)^{-1}({V_{R}^{-}})\subset\textsf{int}(V_{R_{k}}^{-}) where RkR_{k}\to\infty as k.k\to\infty.

Remark 6.1.

Thus int(VR±)𝐔{hk}±\textsf{int}(V_{R}^{\pm})\subset\mathbf{U}_{\{h_{k}\}}^{\pm}. Also we enlist the following observations on the escaping and non-escaping sets, which follows from the same arguments as in Proposition 2.5.

  • 𝐔{hk}+=k=0h(k)1(int(VR+))\displaystyle\mathbf{U}^{+}_{\{h_{k}\}}=\bigcup_{k=0}^{\infty}h(k)^{-1}\Big{(}\textsf{int}(V_{R}^{+})\Big{)} and 𝐔{hk}=k=0(h1(k))1(int(VR)).\displaystyle\mathbf{U}^{-}_{\{h_{k}\}}=\bigcup_{k=0}^{\infty}(h^{-1}(k))^{-1}\Big{(}\textsf{int}(V_{R}^{-})\Big{)}.

  • 𝐊{hk}±\mathbf{K}_{\{h_{k}\}}^{\pm} are closed subsets of 2\mathbb{C}^{2} and 𝐊{hk}±=2𝐔{hk}±.\mathbf{K}_{\{h_{k}\}}^{\pm}=\mathbb{C}^{2}\setminus\mathbf{U}_{\{h_{k}\}}^{\pm}.

  • 𝐔𝒮±𝐔{hk}±𝒰𝒮±\mathbf{U}_{\mathcal{S}}^{\pm}\subset\mathbf{U}_{\{h_{k}\}}^{\pm}\subset\mathcal{U}^{\pm}_{\mathcal{S}} and 𝒦𝒮±𝐊{hk}±𝐊𝒮±VRVR.\mathcal{K}_{\mathcal{S}}^{\pm}\subset\mathbf{K}_{\{h_{k}\}}^{\pm}\subset\mathbf{K}^{\pm}_{\mathcal{S}}\subset V_{R}\cup V_{R}^{\mp}.

Now as in Section 2, consider the following sequences of plurisubharmonic functions on 2\mathbb{C}^{2}

𝖦k+(z)=1𝐝klog+h(k)(z) and 𝖦k(z)=1𝐝klog+h1(k)(z),\displaystyle\mathsf{G}_{k}^{+}(z)=\frac{1}{\mathbf{d}_{k}}\log^{+}\|h(k)(z)\|\text{ and }\mathsf{G}_{k}^{-}(z)=\frac{1}{\mathbf{d}_{k}}\log^{+}\|h^{-1}(k)(z)\|, (6.4)

where 𝐝k=d1dk\mathbf{d}_{k}=d_{1}\ldots d_{k} is the degree of h(k).h(k). Then, we have an analogue to Theorem 1.3 here.

Theorem 6.2.

The sequences of functions {𝖦k±}\{\mathsf{G}_{k}^{\pm}\} converges pointwise to a plurisubharmonic continuous functions 𝖦{hk}±\mathsf{G}_{\{h_{k}\}}^{\pm} on 2\mathbb{C}^{2}, respectively. Further, 𝖦{hk}±\mathsf{G}^{\pm}_{\{h_{k}\}} is pluriharmonic on 𝐔{hk}±\mathbf{U}_{\{h_{k}\}}^{\pm} and int(𝐊{hk}±)\textsf{int}(\mathbf{K}_{\{h_{k}\}}^{\pm}).

The proof of the above theorem and other important results — obtained in this section — are essentially revisiting the techniques discussed through sections 3, 4 and 5, in the current non-autonomous dynamical setup. Hence the presentations will be mostly brief and sketchy. Also, note that Remark 6.4 and the definition of functions {𝖦k±}\{\mathsf{G}^{\pm}_{k}\} above, is valid for any non-autonomous sequence {hk}\{h_{k}\} of Hénon maps, satisfying the identities (6.2) and (6.3).

Proof.

Step 1: The sequence of functions {𝖦k+}\{\mathsf{G}^{+}_{k}\} converges uniformly on compact subsets of VR+V_{R}^{+} and the sequence of functions {𝖦k}\{\mathsf{G}^{-}_{k}\} converges uniformly on compact subsets of VR.V_{R}^{-}.

From the filtration identity (6.2) it follows that for (x,y)VR+(x,y)\in V_{R}^{+}

𝖦k1+(x,y)+logm𝐝k𝖦k+(x,y)𝖦k1+(x,y)+logM𝐝k.\displaystyle\mathsf{G}_{k-1}^{+}(x,y)+\frac{\log m}{\mathbf{d}_{k}}\leq\mathsf{G}_{k}^{+}(x,y)\leq\mathsf{G}_{k-1}^{+}(x,y)+\frac{\log M}{\mathbf{d}_{k}}.

As 𝐝k2k\mathbf{d}_{k}\geq 2^{k} for every k1k\geq 1, we have

|𝖦k1+(x,y)𝖦k+(x,y)|M0~2k,\big{|}\mathsf{G}_{k-1}^{+}(x,y)-\mathsf{G}_{k}^{+}(x,y)\big{|}\leq\frac{\widetilde{M_{0}}}{2^{k}},

where M0~=max{|logm|,|logM|}.\widetilde{M_{0}}=\max\{|\log m|,|\log M|\}. Thus for a given ϵ>0\epsilon>0 there exists m,n1m,n\geq 1, sufficiently large, |𝖦m+𝖦n+|ϵ|\mathsf{G}_{m}^{+}-\mathsf{G}_{n}^{+}|\leq\epsilon on VR+V_{R}^{+}. A similar argument works on VRV_{R}^{-}.

Step 2: The sequence of functions {𝖦k+}\{\mathsf{G}^{+}_{k}\} converges uniformly on compact subsets of 𝐔{hk}+\mathbf{U}_{\{h_{k}\}}^{+} and the sequence of functions {𝖦k}\{\mathsf{G}^{-}_{k}\} converges uniformly on compact subsets of 𝐔{hk}.\mathbf{U}_{\{h_{k}\}}^{-}.

Note that by Remark 6.1, for a given compact set C𝐔{hk}+C\subset\mathbf{U}_{\{h_{k}\}}^{+}, there exists C1\ell_{C}\geq 1, large enough such that h(C)(C)VR+h(\ell_{C})(C)\subset V_{R}^{+}. Thus by similar argument as above, for (x,y)C(x,y)\in C

|𝖦k1+(h(C)(x,y))𝖦k+(h(C)(x,y))|M0~2k.\big{|}\mathsf{G}_{k-1}^{+}(h(\ell_{C})(x,y))-\mathsf{G}_{k}^{+}(h(\ell_{C})(x,y))\big{|}\leq\frac{\widetilde{M_{0}}}{2^{k}}.

Now for a fixed 01\ell_{0}\geq 1 and k>1k>1 consider the sequence of functions defined as

𝖦k0(z)=𝐝0𝐝k+0log+h(k+0)h(0)1(z)\mathsf{G}_{k}^{\ell_{0}}(z)=\frac{\mathbf{d}_{\ell_{0}}}{\mathbf{d}_{k+\ell_{0}}}\log^{+}\|h(k+\ell_{0})h(\ell_{0})^{-1}(z)\|

As h(k+0)h(0)1=hk+0h1+0h(k+\ell_{0})h(\ell_{0})^{-1}=h_{k+\ell_{0}}\circ\cdots\circ h_{1+\ell_{0}}, the functions {𝖦k0}\{\mathsf{G}_{k}^{\ell_{0}}\} are pluriharmonic on VR+V_{R}^{+}, by the same argument as for {𝖦k+}\{\mathsf{G}_{k}^{+}\}. Since Ch(C)1(VR+)C\subset h(\ell_{C})^{-1}(V_{R}^{+}), also by the filtration identity (6.2)

|𝖦k1C(h(C)(x,y))𝖦kC(h(C)(x,y))|M0~2k\big{|}\mathsf{G}_{k-1}^{\ell_{C}}(h(\ell_{C})(x,y))-\mathsf{G}_{k}^{\ell_{C}}(h(\ell_{C})(x,y))\big{|}\leq\frac{\widetilde{M_{0}}}{2^{k}}

for every (x,y)C(x,y)\in C. Note that

𝖦k+C+(x,y)=𝖦kC(h(C)(x,y))𝐝C.\mathsf{G}_{k+\ell_{C}}^{+}(x,y)=\frac{\mathsf{G}_{k}^{\ell_{C}}\big{(}h(\ell_{C})(x,y)\big{)}}{\mathbf{d}_{\ell_{C}}}.

Hence for k1k\geq 1, sufficiently large and (x,y)C(x,y)\in C

|𝖦k+C1+(x,y)𝖦k+C+(x,y)|1𝐝CM0~2k.\big{|}\mathsf{G}_{k+\ell_{C}-1}^{+}(x,y)-\mathsf{G}_{k+\ell_{C}}^{+}(x,y)\big{|}\leq\frac{1}{\mathbf{d}_{\ell_{C}}}\frac{\widetilde{M_{0}}}{2^{k}}.

Thus {𝖦k+}\{\mathsf{G}_{k}^{+}\} converges to a pluriharmonic function on h(C)1(VR+)h(\ell_{C})^{-1}(V_{R}^{+}). Hence the function 𝖦{hk}+\mathsf{G}^{+}_{\{h_{k}\}} is pluriharmonic on 𝐔{hk}+\mathbf{U}_{\{h_{k}\}}^{+}. A similar proof works for 𝖦{hk}\mathsf{G}^{-}_{\{h_{k}\}} and 𝐔{hk}\mathbf{U}_{\{h_{k}\}}^{-}.

Step 3: Let 𝖦{hk}±:=limk𝖦k±\displaystyle\mathsf{G}_{\{h_{k}\}}^{\pm}:=\lim_{k\to\infty}\mathsf{G}_{k}^{\pm}, the pointwise limits of {𝖦k±}\{\mathsf{G}^{\pm}_{k}\}. Then both the limit functions 𝖦{hk}±\mathsf{G}_{\{h_{k}\}}^{\pm} are continuous and plurisubharmonic on 2.\mathbb{C}^{2}.

To complete the above, we first prove that 𝖦{hk}+\mathsf{G}^{+}_{\{h_{k}\}} is continuous on 2\mathbb{C}^{2}, in particular it is continuous on 𝐊{hk}+.\partial\mathbf{K}_{\{h_{k}\}}^{+}. Suppose not, then there exist a point z0𝐊{hk}+z_{0}\in\partial\mathbf{K}^{+}_{\{h_{k}\}} and a sequence {zn}𝐔{hk}+\{z_{n}\}\in\mathbf{U}^{+}_{\{h_{k}\}} such that znz0z_{n}\to z_{0} such that 𝖦{hk}+(zn)>c>0\mathsf{G}^{+}_{\{h_{k}\}}(z_{n})>c>0 for every n1.n\geq 1. Let B=max{log+𝖧i(z):zVR¯ and 1in0}.B=\max\{\log^{+}\|\mathsf{H}_{i}(z)\|:z\in\overline{V_{R}}\text{ and }1\leq i\leq n_{0}\}. Also let 𝐤11\mathbf{k}_{1}\geq 1, sufficiently large, such that

c2>M~2𝐤1\frac{c}{2}>\frac{\widetilde{M}}{2^{\mathbf{k}_{1}}}

where M~:=max{M0~,B}\widetilde{M}:=\max\{\widetilde{M_{0}},B\}. Further note that there exists 𝐤21\mathbf{k}_{2}\geq 1 such that h(k)(zn)VRVR+h(k)(z_{n})\in V_{R}\cup V_{R}^{+} for k𝐤2.k\geq\mathbf{k}_{2}. If not, then there exists a subsequence {kn}\{k_{n}\} of positive integers diverging to infinity and a subsequence {zln}\{z_{l_{n}}\} of {zn}\{z_{n}\} such that h(kn)(zln)VRh(k_{n})(z_{l_{n}})\in V_{R}^{-}, i.e., zlnh(kn)1(VR).z_{l_{n}}\in h(k_{n})^{-1}(V_{R}^{-}). Hence zlnint(VRkn)z_{l_{n}}\notin\textsf{int}(V_{R_{k_{n}}}). As RknR_{k_{n}}\to\infty, this would mean zln\|z_{l_{n}}\|\to\infty, which is a contradiction!

Claim: The sequence {h(k)(zn)}VR+\{h(k)(z_{n})\}\in V_{R}^{+} whenever kmax{𝐤1,𝐤2}k\geq\max\{\mathbf{k}_{1},\mathbf{k}_{2}\}. Suppose not, then there exist klmax{𝐤1,𝐤2}k_{l}\geq\max\{\mathbf{k}_{1},\mathbf{k}_{2}\} and zl{zk}z_{l}\in\{z_{k}\}, such that h(kl)(zl)VR.h(k_{l})(z_{l})\in V_{R}. Also from (6.2), h(k)(zl)VR+h(k)(z_{l})\notin V_{R}^{+} for every kkl.k\leq k_{l}. Let 𝐤l>kl\mathbf{k}_{l}>k_{l} be the minimum positive integer such that h(𝐤l)(zl)VR+h(\mathbf{k}_{l})(z_{l})\in V_{R}^{+}, i.e., h(𝐤l)(zl)B\|h(\mathbf{k}_{l})(z_{l})\|\leq B and h(k)(zl)VR+h(k)(z_{l})\in V_{R}^{+} for every k𝐤l.k\geq\mathbf{k}_{l}. Hence for every k>𝐤lk>\mathbf{k}_{l}

𝖦k+(zl)log+h(𝐤l)(zl)𝐝𝐤l+M0~i=𝐤l+1k1𝐝iM~i=𝐤lk1𝐝i.\mathsf{G}_{k}^{+}(z_{l})\leq\frac{\log^{+}\|h(\mathbf{k}_{l})(z_{l})\|}{\mathbf{d}_{\mathbf{k}_{l}}}+\widetilde{M_{0}}\sum_{i=\mathbf{k}_{l}+1}^{k}\frac{1}{\mathbf{d}_{i}}\leq\widetilde{M}\sum_{i=\mathbf{k}_{l}}^{k}\frac{1}{\mathbf{d}_{i}}.

Since 𝐝i2i\mathbf{d}_{i}\geq 2^{i} and 𝐤l1𝐤1\mathbf{k}_{l}-1\geq\mathbf{k}_{1}, the above simplifies to

𝖦k+(zl)M~i=𝐤lk12iM~2𝐤l1M~2𝐤1<c2.\mathsf{G}_{k}^{+}(z_{l})\leq\widetilde{M}\sum_{i=\mathbf{k}_{l}}^{k}\frac{1}{2^{i}}\leq\frac{\widetilde{M}}{2^{\mathbf{k}_{l}-1}}\leq\frac{\widetilde{M}}{2^{\mathbf{k}_{1}}}<\frac{c}{2}.

So 𝖦{hk}+(zl)<c\mathsf{G}_{\{h_{k}\}}^{+}(z_{l})<c, which is a contradiction to the assumption. Thus the claim follows.

As z0𝐊{hk}+𝐊{hk}+z_{0}\in\partial\mathbf{K}^{+}_{\{h_{k}\}}\subset\mathbf{K}^{+}_{\{h_{k}\}} (since it is closed), by Lemma 2.8 and Remark 3.2 there exists 𝐤01\mathbf{k}_{0}\geq 1 such that h(k)(z0)R\|h(k)(z_{0})\|\leq R for every k𝐤0.k\geq\mathbf{k}_{0}. Also, from the above Claim and (6.2) we may fix a k0>max{𝐤0,𝐤1,𝐤2}k_{0}>\max\{\mathbf{k}_{0},\mathbf{k}_{1},\mathbf{k}_{2}\} such that h(k0)(zn)int(VR+1+).h(k_{0})(z_{n})\in\textsf{int}(V_{R+1}^{+}). Since znz0z_{n}\to z_{0}, by continuity of h(k0)h(k_{0}) we have h(k0)(z0)VR+1+h(k_{0})(z_{0})\in V_{R+1}^{+}, i.e., h(k0)(z0)R+1>R\|h(k_{0})(z_{0})\|\geq R+1>R, which is not possible (as z0𝐊{hk}+z_{0}\in\mathbf{K}^{+}_{\{h_{k}\}}). Thus 𝖦{hk}+\mathsf{G}^{+}_{\{h_{k}\}} is continuous on 2.\mathbb{C}^{2}.

Now 𝖦{hk}+\mathsf{G}^{+}_{\{h_{k}\}} is pluriharmonic on U{hk}+U_{\{h_{k}\}}^{+}, and is identically zero, i.e, is also pluriharmonic in the interior of K{hk}±K_{\{h_{k}\}}^{\pm} (provided it is non-empty). Also it is continuous on 2\mathbb{C}^{2}, hence the upper semi-continuous regularisation of 𝖦{hk}+\mathsf{G}^{+}_{\{h_{k}\}} on 2\mathbb{C}^{2} matches with itself and Step 2 holds.

A similar argument will work for 𝖦{hk}\mathsf{G}^{-}_{\{h_{k}\}}, which completes the proof. ∎

Corollary 6.3.

There exist constants c{hk}±c_{\{h_{k}\}}^{\pm}\in\mathbb{R} such that for (x,y)VR±(x,y)\in V_{R}^{\pm} (respectively),

𝖦{hk}+(x,y)=log|y|+O(1) and 𝖦{hk}(x,y)=log|x|+O(1).\displaystyle\mathsf{G}^{+}_{\{h_{k}\}}(x,y)=\log|y|+O(1)\text{ and }\mathsf{G}^{-}_{\{h_{k}\}}(x,y)=\log|x|+O(1).
Proof.

The proof is same as the proof of Corollary 3.3. ∎

Lemma 6.4.

The sequences {𝖦k±}\{\mathsf{G}_{k}^{\pm}\} converges uniformly on compact subsets of 2\mathbb{C}^{2} to 𝖦{hk}±\mathsf{G}_{\{h_{k}\}}^{\pm}, respectively.

Proof.

The proof is again completely similar to the proof of Lemma 4.2, however, we revisit the steps briefly. Note that if C𝐔{hk}+C\subset\mathbf{U}_{\{h_{k}\}}^{+} then the uniform convergence is immediate, as the sequence {𝖦k+}\{\mathsf{G}^{+}_{k}\} is uniformly Cauchy on 𝐔{hk}+\mathbf{U}_{\{h_{k}\}}^{+}, by the proof of Theorem 6.2.

Next, let C𝐊{hk}+C\subset\mathbf{K}_{\{h_{k}\}}^{+} then h(k)𝒢kb(z)h(k)\in\mathscr{G}^{b}_{k}(z) for every zCz\in C and k1k\geq 1, where 𝒢kb(z)\mathscr{G}^{b}_{k}(z) is as introduced in the Step 3 of the proof of Theorem 1.3, and hence by Remark 3.2, there exists a positive integer 𝐤0(C)1\mathbf{k}_{0}(C)\geq 1 such that h(k)(C)VRh(k)(C)\subset V_{R} for every k𝐤0k\geq\mathbf{k}_{0}. Thus 𝖦k+|CR2k{\mathsf{G}_{k}^{+}}_{|C}\leq\frac{R}{2^{k}}, which proves the uniform convergence in this case.

Finally, let CC intersects both 𝐊{hk}+\mathbf{K}_{\{h_{k}\}}^{+} and 𝐔{hk}+\mathbf{U}_{\{h_{k}\}}^{+}, then the uniform convergence is immediate from above on C𝐊{hk}+C\cap\mathbf{K}^{+}_{\{h_{k}\}}, i.e., for a given ϵ>0\epsilon>0, there exists 𝐤01\mathbf{k}_{0}\geq 1 such that |𝖦k+(z)𝖦{hk}+(z)|ϵ|\mathsf{G}_{k}^{+}(z)-\mathsf{G}_{\{h_{k}\}}^{+}(z)|\leq\epsilon for every k𝐤0.k\geq\mathbf{k}_{0}. Further by Lemma 2.8 there exists 𝐤0(C)1\mathbf{k}_{0}(C)\geq 1 such that h(k)(C)VRVR+h(k)(C)\subset V_{R}\cup V_{R}^{+} for every k𝐤0(C)k\geq\mathbf{k}_{0}(C). Note that by (2.1),

h(k)1(VR+)=h(k)1(VR+)¯int(h(k+1)1(VR+)).\displaystyle h(k)^{-1}(V_{R}^{+})=\overline{h(k)^{-1}(V_{R}^{+})}\subset\textsf{int}(h(k+1)^{-1}(V_{R}^{+})). (6.5)

Now as in the proof of Lemma 4.2 we define the following subsets of CC

Ck=C(h(k)1(int(VR+))h(k1)1(int(VR+))) and C0=Cint(VR+),C_{k}=C\cap\big{(}h(k)^{-1}(\textsf{int}(V_{R}^{+}))\setminus h(k-1)^{-1}(\textsf{int}(V_{R}^{+}))\big{)}\text{ and }C_{0}=C\cap\textsf{int}(V_{R}^{+}),

i.e., k=0Ck=C𝐔{hk}+.\cup_{k=0}^{\infty}C_{k}=C\cap\mathbf{U}_{\{h_{k}\}}^{+}. Since C𝐔{hk}+C\cap\mathbf{U}_{\{h_{k}\}}^{+}\neq\emptyset, it follows from (6.5) that CkC_{k}’s are non-empty sets for k1k\geq 1, sufficiently large. Also, let B=max{𝖧i(z):zVR+1,1in0}B=\max\{\|\mathsf{H}_{i}(z)\|:z\in V_{R+1},1\leq i\leq n_{0}\} and C~k=i=kCi.\tilde{C}_{k}=\cup_{i=k}^{\infty}C_{i}. Then for zCkz\in C_{k}, 𝖦l+(z)logB𝐝l\mathsf{G}_{l}^{+}(z)\leq\frac{\log B}{\mathbf{d}_{l}} whenever 𝐤0(C)lk.\mathbf{k}_{0}(C)\leq l\leq k. Now for n1n\geq 1

𝖦l+n+(z)logB𝐝k+logMi=1n1𝐝k+i2M~𝐝kM~2k1,\mathsf{G}_{l+n}^{+}(z)\leq\frac{\log B}{\mathbf{d}_{k}}+\log{M}\sum_{i=1}^{n}\frac{1}{\mathbf{d}_{k+i}}\leq\frac{2\widetilde{M}}{\mathbf{d}_{k}}\leq\frac{\widetilde{M}}{2^{k-1}},

where M~=max{|logM|,|logm|,|logB|}.\widetilde{M}=\max\{|\log M|,|\log m|,|\log B|\}. Again, by the same arguments as in proof of Lemma 4.2, i.e., by continuity of 𝖦{hk}+\mathsf{G}^{+}_{\{h_{k}\}} and the above, there exists 𝐤11\mathbf{k}_{1}\geq 1 such that

|𝖦k+(z)𝖦{hk}+(z)|<ϵ|\mathsf{G}_{k}^{+}(z)-\mathsf{G}_{\{h_{k}\}}^{+}(z)|<\epsilon

whenever zC~𝐤1z\in\widetilde{C}_{\mathbf{k}_{1}} and k𝐤1.k\geq\mathbf{k}_{1}. Now, as

(𝐔{hk}+C)C~𝐤1Ch(𝐤11)1(VR+),(\mathbf{U}^{+}_{\{h_{k}\}}\cap C)\setminus\widetilde{C}_{\mathbf{k}_{1}}\subset C\cap h(\mathbf{k}_{1}-1)^{-1}(V_{R}^{+}),

and Ch(𝐤11)1(VR+)C\cap h(\mathbf{k}_{1}-1)^{-1}(V_{R}^{+}) is a compact set contained in 𝐔{hk}+\mathbf{U}_{\{h_{k}\}}^{+}, there exists 𝐤20\mathbf{k}_{2}\geq 0 such that |𝖦k+(z)𝖦{hk}+(z)|ϵ|\mathsf{G}_{k}^{+}(z)-\mathsf{G}_{\{h_{k}\}}^{+}(z)|\leq\epsilon whenever z(𝐔{hk}+C)C~𝐤1.z\in(\mathbf{U}^{+}_{\{h_{k}\}}\cap C)\setminus\widetilde{C}_{\mathbf{k}_{1}}. Since C=(C𝐊{hk}+)C~𝐤1((𝐔{hk}+C)C~𝐤1)C=(C\cap\mathbf{K}^{+}_{\{h_{k}\}})\cup\widetilde{C}_{\mathbf{k}_{1}}\cup((\mathbf{U}^{+}_{\{h_{k}\}}\cap C)\setminus\widetilde{C}_{\mathbf{k}_{1}}), for kmax{𝐤0,𝐤1,𝐤2}k\geq\max\{\mathbf{k}_{0},\mathbf{k}_{1},\mathbf{k}_{2}\}, we have |𝖦k+𝖦{hk}+|Cϵ.\big{|}\mathsf{G}_{k}^{+}-\mathsf{G}_{\{h_{k}\}}^{+}\big{|}_{C}\leq\epsilon.

Theorem 6.5.

For every k1k\geq 1, let Gh(k)±G^{\pm}_{h(k)} denote the Green’s function corresponding to the maps h(k)h(k) and h1(k).h^{-1}(k). Then the sequence {Gh(k)±}\{G^{\pm}_{h(k)}\} converge uniformly to 𝖦{hk}±\mathsf{G}_{\{h_{k}\}}^{\pm}, respectively, on compact subsets of 2.\mathbb{C}^{2}.

Proof.

This proof is again similar to the proof of Theorem 4.1. Let CC be a compact subset of 2\mathbb{C}^{2}, then let C1=𝐊{hk}+CC_{1}=\mathbf{K}_{\{h_{k}\}}^{+}\cap C. Since C1C_{1} is a compact set contained in 𝐊{hk}+\mathbf{K}_{\{h_{k}\}}^{+}, by Lemma 2.8 and Remark 3.2 there exists 𝐤C11\mathbf{k}_{C_{1}}\geq 1 such that h(k)(z)VRh(k)(z)\in V_{R} for k𝐤C1k\geq\mathbf{k}_{C_{1}}. Thus h(k)𝒢kb(z)h(k)\in\mathscr{G}^{b}_{k}(z) for zC1z\in C_{1} and h(k)𝒢ku(z~)h(k)\in\mathscr{G}^{u}_{k}(\tilde{z}) for z~CC1\tilde{z}\in C\setminus C_{1}, whenever k𝐤C1k\geq\mathbf{k}_{C_{1}}.

Now by Step 1 in the proof of Theorem 4.1 gives that, for a given ϵ>0\epsilon>0 there exists 𝐤1𝐤C1\mathbf{k}_{1}\geq\mathbf{k}_{C_{1}} such that for k𝐤1k\geq\mathbf{k}_{1},

|Gh(k)+(z)log+h(k)(z)𝐝k|C1=|Gh(k)+(z)𝖦k+(z)|C1<ϵ/2.\bigg{|}G^{+}_{h(k)}(z)-\frac{\log^{+}\|h(k)(z)\|}{\mathbf{d}_{k}}\bigg{|}_{C_{1}}=|G^{+}_{h(k)}(z)-\mathsf{G}^{+}_{k}(z)|_{C_{1}}<\epsilon/2.

Also as h(k)𝒢ku(z~)h(k)\in\mathscr{G}^{u}_{k}(\tilde{z}) for z~CC1\tilde{z}\in C\setminus C_{1} and k𝐤1k\geq\mathbf{k}_{1}, by Step 2 in the proof of Theorem 4.1

logh(k)(z~)𝐝kM~i=1l11𝐝kilogh(k)l(z~)𝐝kllogh(k)(z~)𝐝k+M~i=1l11𝐝ki,\frac{\log\|h(k)(\tilde{z})\|}{\mathbf{d}_{k}}-\widetilde{M}\sum_{i=1}^{l-1}\frac{1}{\mathbf{d}_{k}^{i}}\leq\frac{\log\|h(k)^{l}(\tilde{z})\|}{\mathbf{d}_{k}^{l}}\leq\frac{\log\|h(k)(\tilde{z})\|}{\mathbf{d}_{k}}+\widetilde{M}\sum_{i=1}^{l-1}\frac{1}{\mathbf{d}_{k}^{i}},

whenever l2l\geq 2. Hence there exists 𝐤2𝐤1\mathbf{k}_{2}\geq\mathbf{k}_{1} such that for k𝐤2k\geq\mathbf{k}_{2}

|Gh(k)+(z)log+h(k)(z)𝐝k|CC1=|Gh(k)+(z)𝖦k+(z)|CC12M~𝐝k<ϵ/2.\bigg{|}G^{+}_{h(k)}(z)-\frac{\log^{+}\|h(k)(z)\|}{\mathbf{d}_{k}}\bigg{|}_{C\setminus C_{1}}=|G^{+}_{h(k)}(z)-\mathsf{G}^{+}_{k}(z)|_{C\setminus C_{1}}\leq\frac{2\widetilde{M}}{\mathbf{d}_{k}}<\epsilon/2.

Finally, by Lemma 6.4, there exists 𝐤C𝐤2\mathbf{k}_{C}\geq\mathbf{k}_{2} such that the theorem holds. A similar argument will work for {Gh(k)}\{G_{h(k)}^{-}\} and 𝖦{hk}.\mathsf{G}^{-}_{\{h_{k}\}}.

Now 𝖦{hk}±0\mathsf{G}^{\pm}_{\{h_{k}\}}\equiv 0 in the int(𝐊{hk}±)\textsf{int}(\mathbf{K}_{\{h_{k}\}}^{\pm}), provided it is non-empty, hence 𝐊{hk}±\mathbf{K}_{\{h_{k}\}}^{\pm} are pseudoconcave subsets of 2\mathbb{C}^{2}. Also, as an immediate corollary to Theorems 6.2 and 6.5, we have the initial statement of the following.

Corollary 6.6.

The currents μ{hk}±:=12πddc(𝖦{hk}±)\displaystyle\mu_{\{h_{k}\}}^{\pm}:=\frac{1}{2\pi}dd^{c}(\mathsf{G}^{\pm}_{\{h_{k}\}}) are positive (1,1)(1,1) currents of mass 1, supported on 𝐉{hk}±=𝐊{hk}±\mathbf{J}_{\{h_{k}\}}^{\pm}=\partial\mathbf{K}_{\{h_{k}\}}^{\pm}, respectively. Also support of μ{hk}±=𝐉{hk}±\mu_{\{h_{k}\}}^{\pm}=\mathbf{J}_{\{h_{k}\}}^{\pm} and μ{hk}:=μ{hk}+μ{hk}\mu_{\{h_{k}\}}:=\mu_{\{h_{k}\}}^{+}\wedge\mu_{\{h_{k}\}}^{-} is a compactly supported probability measure.

Proof.

We only prove Supp μ{hk}±=𝐉{hk}±\text{Supp }\mu_{\{h_{k}\}}^{\pm}=\mathbf{J}_{\{h_{k}\}}^{\pm}, here. Since 𝖦{hk}+\mathsf{G}_{\{h_{k}\}}^{+} is non-constant on 2\mathbb{C}^{2} and attains the minimum value, i.e., zero, in the interior of neighborhood of a point z0𝐉{hk}+z_{0}\in\mathbf{J}_{\{h_{k}\}}^{+}, the function is strictly pluriharmonic at z0.z_{0}. As z0z_{0} is an arbitrary point on 𝐉{hk}+,\mathbf{J}_{\{h_{k}\}}^{+}, the support of μ{hk}+\mu_{\{h_{k}\}}^{+} is equal to 𝐉{hk}+\mathbf{J}_{\{h_{k}\}}^{+}. A similar argument will work for μ{hk}.\mu_{\{h_{k}\}}^{-}. Also μ{hk}\mu_{\{h_{k}\}} is a positive measure is immediate, and it is compactly supported follows from Remark 6.1. ∎

Remark 6.7.

Note that any subsequence of {h(k)}\{h(k)\} neither diverges to infinity nor is it bounded on any neighbourhood of a point z0𝐊{hk}+z_{0}\in\partial\mathbf{K}_{\{h_{k}\}}^{+}. Thus 𝐉{hk}+=𝐊{hk}+\mathbf{J}_{\{h_{k}\}}^{+}=\partial\mathbf{K}_{\{h_{k}\}}^{+} is contained in the Julia set for the dynamics of the non-autonomous family {hk}\{h_{k}\}. But note that

2𝐉{hk}+=int(𝐊{hk}+)𝐔{hk}+.\mathbb{C}^{2}\setminus\mathbf{J}_{\{h_{k}\}}^{+}=\textsf{int}(\mathbf{K}_{\{h_{k}\}}^{+})\cup\mathbf{U}_{\{h_{k}\}}^{+}.

Hence by Lemma 2.8 and Remark 3.2, int(𝐊{hk}+)\textsf{int}(\mathbf{K}_{\{h_{k}\}}^{+}) is contained in the Fatou set and thus the Julia set corresponding to the dynamics of {hk}\{h_{k}\} is equal to 𝐉{hk}+\mathbf{J}_{\{h_{k}\}}^{+}.

Remark 6.8.

Note that the two ‘crucial’ conditions required on a non-autonomous sequence of Hénon maps {hk}\{h_{k}\} of the form (1.2), to complete the proof of Theorem 6.2 and 6.5 are

  1. (i)

    The sequence {hk}\{h_{k}\} admits a uniform radius filtration R{hk}>1R_{\{h_{k}\}}>1 (in the above case it is the radius of filtration of the semigroup 𝒮\mathcal{S}, generated by 𝒢\mathscr{G}), such that for every R>R{hk}R>R_{\{h_{k}\}}

    • hk(VR+)¯VR+ and hk1(VR¯)VR.\displaystyle\overline{h_{k}(V_{R}^{+})}\subset V_{R}^{+}\text{ and }\overline{h^{-1}_{k}(V_{R}^{-}})\subset V_{R}^{-}.

    • there exists a sequence positive real numbers {Rk}\{R_{k}\} diverging to infinity, with R0=RR_{0}=R, satisfying VRkh(k)(VR+)= and VRkh1(k)(VR)=.\displaystyle V_{R_{k}}\cap h(k)(V_{R}^{+})=\emptyset\text{ and }V_{R_{k}}\cap h^{-1}(k)(V_{R}^{-})=\emptyset.

    • There exist uniform constants 0<m<1<M0<m<1<M, such that the filtration identities (6.2) and (6.3) are satisfied on VR±V_{R}^{\pm}, respectively.

  2. (ii)

    For every RR{hk}R\geq R_{\{h_{k}\}}, there exists a uniform constant BR=max{hk(z):zVR}<.B_{R}=\max\{\|h_{k}(z)\|:z\in V_{R}\}<\infty. The same holds in the above setup of Theorem 6.2 and 6.5, as the choices for hkh_{k} are finite, for every k1.k\geq 1.

Hence we have the following analogue of of Theorem 6.2 and 6.5 in a more general setup.

Remark 6.9.

Let {hk}\{h_{k}\} be a non-autonomous sequence of Hénon maps satisfying conditions (i) and (ii) of Remark 6.8 then

  • The sequences of plurisubharmonic function {𝖦k±}\{\mathsf{G}^{\pm}_{k}\}, as defined in (6.4) converges to a plurisubharmonic continuous functions 𝖦{hk}±\mathsf{G}_{\{h_{k}\}}^{\pm} on 2\mathbb{C}^{2}, respectively. Further 𝖦{hk}±\mathsf{G}^{\pm}_{\{h_{k}\}} is pluriharmonic on 𝐔{hk}±\mathbf{U}_{\{h_{k}\}}^{\pm} and int(𝐊{hk}±)\textsf{int}(\mathbf{K}_{\{h_{k}\}}^{\pm}), where 𝐔{hk}±\mathbf{U}_{\{h_{k}\}}^{\pm} and int(𝐊{hk}±)\textsf{int}(\mathbf{K}_{\{h_{k}\}}^{\pm}).

  • The sequences {Gh(k)±}\{G^{\pm}_{h(k)}\} converge uniformly to 𝖦{hk}±\mathsf{G}_{\{h_{k}\}}^{\pm}, respectively, on compact subsets.

Example 6.10.

Let Hk(x,y)=(aky,akx+p(y))H_{k}(x,y)=(a_{k}y,a_{k}x+p(y)) where pp is a polynomial of degree at least 22, then the sequence {H~k}\{\tilde{H}_{k}\} defined as below is a sequence of Hénon maps.

H~k(x,y):=H2kH2k1(x,y)\displaystyle\tilde{H}_{k}(x,y):=H_{2k}\circ H_{2k-1}(x,y) =(y,a2ka2k1x+p(y/a2k))(y,a2ka2k1x+a2kp(y)).\displaystyle=(y,a_{2k}a_{2k-1}x+p(y/a_{2k}))\circ(y,a_{2k}a_{2k-1}x+a_{2{k}}p(y)).

Further, if 0<c<|ak|<d0<c<|a_{k}|<d for every k1k\geq 1, the conditions (i) and (ii) in Remark 6.8 are satisfied, and by Remark 6.9, it is possible to construct the dynamical Green’s functions. However, the condition (i) in Remark 6.8 fails for Hk~1\tilde{H_{k}}^{-1}, if |ak|0|a_{k}|\to 0 (see Theorem 1.4 in [15]).

Also note that the functions 𝖦{hk}±\mathsf{G}_{\{h_{k}\}}^{\pm} admit logarithmic growth at infinity, and the closure of the sets 𝐊{hk}±\mathbf{K}_{\{h_{k}\}}^{\pm} in 2\mathbb{P}^{2} is 𝐊{hk}±I±\mathbf{K}_{\{h_{k}\}}^{\pm}\cup I^{\pm}, as defined in Section 5. Hence it is possible to generalise the results stated in Section 5, to the setup of dynamics of a non-autonomous sequence of Hénon maps {hk}𝒮\{h_{k}\}\subset\mathcal{S}. In particular, the analogue to Corollary 5.8 is

Corollary 6.11.

Let S±S^{\pm} be two closed positive (1,1)(1,1)-currents in 2\mathbb{P}^{2} of mass 1, such that the support of S+S^{+} does not contain the point [0:1:0][0:1:0] and the support of SS^{-} does not contain the point [1:0:0][1:0:0]. Also, let h(k)¯\overline{h(k)} denote the extension of h(k)h(k) to 2,\mathbb{P}^{2}, for every k1k\geq 1 then

limk1𝐝kh(k)¯(S+)μ{hk}+ and limk1𝐝kh1(k)¯(S)μ{hk}.\lim_{k\to\infty}\frac{1}{\mathbf{d}_{k}}\overline{h(k)}^{*}(S^{+})\to\mu_{\{h_{k}\}}^{+}\text{ and }\lim_{k\to\infty}\frac{1}{\mathbf{d}_{k}}\overline{h^{-1}(k)}^{*}(S^{-})\to\mu_{\{h_{k}\}}^{-}.

The proof is immediate from Remark 5.7 and Theorem 6.5. Also the proof of Corollary 6.11 does not generalises to general non-autonomous families of Hénon maps (observed in Remark 6.8), unlike Theorems 6.2 and 6.5. It crucially requires that h(k)𝒢kh(k)\in\mathscr{G}_{k}, k1.k\geq 1.

Remark 6.12.

Note that as mentioned in the introduction, the above result is a more explicit version of Theorem 5.1 in [11], for Hénon maps. The latter established the existence of similar non-autonomous currents for families of horizontal maps on appropriate subdomains of k\mathbb{C}^{k}, k2k\geq 2 and Hénon maps of the above form are indeed known to be horizontal on a large enough polydisc at the origin in 2\mathbb{C}^{2}, by [29]. Also, the construction and the convergence properties of similar Green’s current for parametrised families of skew-product of (monic) Hénon maps — of fixed degree — over compact complex manifolds, have been studied in [27].

7. Attracting basins of non-autonomous sequences in 𝒮\mathcal{S}

Let 𝒮\mathcal{S} be a semigroup generated by finitely many Hénon maps, having an attracting behaviour, i.e., satisfy (1.4) at the origin. Then for every i1i\geq 1, there exist r>0r>0 and 0<α<10<\alpha<1 such that 𝖧i(B(0;r))B(0;αr)\mathsf{H}_{i}(B(0;r))\subset B(0;\alpha r). In particular, for every sequence {hk}𝒮\{h_{k}\}\subset\mathcal{S}, hk(z)0h_{k}(z)\to 0 as nn\to\infty for zB(0;r)z\in B(0;r). Hence we have the following observations.

  • The strong filled positive Julia set 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+} is non-empty and contains a neighbourhood of the origin. Also, the strong filled negative Julia set 𝒦𝒮\mathcal{K}_{\mathcal{S}}^{-} is non-empty and contains the origin.

  • The basin of attraction at the origin of every h𝒮h\in\mathcal{S}, say Ωh\Omega_{h}, is a Fatou–Bieberbach domain, i.e., biholomorphic to 2\mathbb{C}^{2} (see [28] for the proof).

  • The non-autonomous basin of attraction at the origin for a sequence {hk}\{h_{k}\} — denoted by Ω{hk}\Omega_{\{h_{k}\}}, defined in statement of Theorem 1.1 — is an elliptic domain containing the origin as every hkh_{k} satisfies the uniform bound condition. (see [16],[14] for the result).

Lemma 7.1.

Ω{hk}𝐊{hk}+\partial\Omega_{\{h_{k}\}}\subset\partial\mathbf{K}_{\{h_{k}\}}^{+} for the non-autonomous dynamical system {hk}.\{h_{k}\}.

Proof.

Observe that by the argument as in Remark 6.7, Ω{hk}𝐔{hk}+=\partial\Omega_{\{h_{k}\}}\cap\mathbf{U}_{\{h_{k}\}}^{+}=\emptyset. Now, if z0Ω{hk}int(𝐊{hk}+)z_{0}\in\partial\Omega_{\{h_{k}\}}\cap\textsf{int}(\mathbf{K}_{\{h_{k}\}}^{+}), then there exists a neighbourhood of B(z0;δ)int(𝐊{hk}+)B(z_{0};\delta)\subset\textsf{int}(\mathbf{K}_{\{h_{k}\}}^{+}), i.e., the sequence {h(k)}\{h(k)\} is locally uniformly bounded and hence normal on B(z0;δ)B(z_{0};\delta). Then there exists a subsequence {nk}\{n_{k}\} such that h(nk)(z0)h(n_{k})(z_{0}) does not tend to 0, however h(nk)(z1)0h(n_{k})(z_{1})\to 0 whenever z1B(z0;δ)Ω{hk}+z_{1}\in B(z_{0};\delta)\cap\Omega_{\{h_{k}\}}^{+}. Thus {h(k)}\{h(k)\} is not normal on any neighbourhood of z0z_{0}. Hence by Remark 6.7, Ω{hk}𝐉{hk}+=𝐊{hk}+\partial\Omega_{\{h_{k}\}}\subset\mathbf{J}_{\{h_{k}\}}^{+}=\partial\mathbf{K}_{\{h_{k}\}}^{+}. ∎

Remark 7.2.

Note that in the setup of iterative dynamics of Hénon maps, the boundary of any attracting basin is equal to the Julia set (see [5, Theorem 2]). In the above lemma we only show that Ω{hk}\partial\Omega_{\{h_{k}\}} is properly contained in the Julia set. This leads to the question: Is 𝐉{hk}+=Ω{hk}\mathbf{J}_{\{h_{k}\}}^{+}=\partial\Omega_{\{h_{k}\}} in the non-autonomous setup?

Let Ωh(k)\Omega_{h(k)} be the basin of attraction of the origin for every h(k)h(k), k1k\geq 1, as origin is an attracting fixed point, i.e., the definitions are compared as

Ωh(k)={z2:h(k)n(z)0 as n} and Ω{hk}={z2:h(k)(z)0 as k}.\displaystyle\Omega_{h(k)}=\{z\in\mathbb{C}^{2}:h(k)^{n}(z)\to 0\text{ as }n\to\infty\}\text{ and }\Omega_{\{h_{k}\}}=\{z\in\mathbb{C}^{2}:h(k)(z)\to 0\text{ as }k\to\infty\}.
Lemma 7.3.

Let KK be a compact set contained in Ω{hk}\Omega_{\{h_{k}\}} then there exists a positive integer dependent on KK, i.e., N0(K)1N_{0}(K)\geq 1 such that KΩh(k)K\subset\Omega_{h(k)} for every kN0(K).k\geq N_{0}(K).

Proof.

Note that {hk}\{h_{k}\} varies within a collection of finitely many Hénon maps, {𝖧i:1in0}\{\mathsf{H}_{i}:1\leq i\leq n_{0}\}, each admitting an attracting fixed point at the origin. Thus there exists a neighbourhood B(0;r)B(0;r) at the origin and 0<α<10<\alpha<1 such that 𝖧i(B(0;r))B(0;αr)\mathsf{H}_{i}(B(0;r))\subset B(0;\alpha r) for every 1in0.1\leq i\leq n_{0}. In particular, B(0;r)B(0;r) is contained in attracting basin of the origin for every hh, h𝒮h\in\mathcal{S}. Since KΩ{hk}K\subset\Omega_{\{h_{k}\}} is compact, h(k)(w)B(0;r)h(k)(w)\in B(0;r) for every wKw\in K and kN0(K).k\geq N_{0}(K). Hence

Gh(k)+(h(k)(w))=0, i.e., Gh(k)+(w)=0G^{+}_{h(k)}(h(k)(w))=0,\text{ i.e., }G_{h(k)}^{+}(w)=0

for every wKw\in K and nN0(K).n\geq N_{0}(K). So Kint(Kh(k)+).K\subset\textsf{int}(K^{+}_{h(k)}). But h(k)(0)=0h(k)(0)=0, hence h(k)(Ωh(k))=Ωh(k)h(k)(\Omega_{h(k)})=\Omega_{h(k)}, and the above implies h(k)(K)Ωh(k)h(k)(K)\subset\Omega_{h(k)}. Thus KΩh(k)K\subset\Omega_{h(k)} for every kN0(K).k\geq N_{0}(K).

Next, we complete the proof of Theorem 1.1, by appealing to an idea used in [30].

Proof of Theorem 1.1.

Let {Kk}\{K_{k}\} be an exhaustion by compacts of Ω{hk}.\Omega_{\{h_{k}\}}. Then from Lemma 7.3, there exists an increasing sequence of positive integers {nk}\{n_{k}\} such that KkΩh(nk).K_{k}\subset\Omega_{h(n_{k})}. Since every Ωh(nk)\Omega_{h(n_{k})} is a Fatou-Bieberbach domain, our goal is to construct a sequence of biholomorphisms {ϕk}\{\phi_{k}\}, i.e., holomorphic maps that are both one-one and onto from Ωh(nk)\Omega_{h(n_{k})} to 2\mathbb{C}^{2}, appropriately and inductively, such that for a given summable sequence of positive real numbers {ρk}\{\rho_{k}\} the following holds

ϕk(z)ϕk+1(z)<ρk for zKk and ϕk1(z)ϕk+11(z)<ρk for zB(0;k).\displaystyle\|\phi_{k}(z)-\phi_{k+1}(z)\|<\rho_{k}\text{ for }z\in K_{k}\text{ and }\|\phi_{k}^{-1}(z)-\phi_{k+1}^{-1}(z)\|<\rho_{k}\text{ for }z\in B(0;k). (7.1)

Basic step: Let ϕ1:Ωh(n1)2\phi_{1}:\Omega_{h(n_{1})}\to\mathbb{C}^{2} be a biholomorphism. By results in [2, Theorem 2.1] for δ<ρ1/2\delta<\rho_{1}/2 there exists F2Aut(2)F_{2}\in{\rm Aut}(\mathbb{C}^{2}) such that

ϕ1(z)F2(z)\displaystyle\|\phi_{1}(z)-F_{2}(z)\| <δ for zK1(r) and ϕ11(z)F21(z)\displaystyle<\delta\text{ for }z\in K_{1}(r)\text{ and }\|\phi_{1}^{-1}(z)-F_{2}^{-1}(z)\| <δ for zB(0;1+r)\displaystyle<\delta\text{ for }z\in B(0;1+r) (7.2)

where K1(r)=zK1(B(z;r))Ωh(n1)K_{1}(r)=\cup_{z\in K_{1}}(B(z;r))\subset\Omega_{h(n_{1})} for some r>0r>0, i.e., an rr-neighbourhood of K1K_{1}, contained in Ωh(n1)\Omega_{h(n_{1})}. Since F2F_{2} is uniformly continuous on K1(r)K_{1}(r), there exists ϵ0>0\epsilon_{0}>0 such that for z,wK1(r)z,w\in K_{1}(r)

F2(z)F2(w)<δ whenever zw<ϵ0.\displaystyle\|F_{2}(z)-F_{2}(w)\|<\delta\text{ whenever }\|z-w\|<\epsilon_{0}. (7.3)

Let ϵ<min{ϵ0,r,δ}.\epsilon<\min\{\epsilon_{0},r,\delta\}. Then from [30, Lemma 4], there exists a biholomorphism ψ2:Ωh(n2)2\psi_{2}:\Omega_{h(n_{2})}\to\mathbb{C}^{2} such that

ψ2(z)z<ϵ for every zK1 and ψ21(z)z<ϵ for every zF21(B(0;1)).\displaystyle\|\psi_{2}(z)-z\|<\epsilon\text{ for every }z\in K_{1}\text{ and }\|\psi_{2}^{-1}(z)-z\|<\epsilon\text{ for every }z\in F_{2}^{-1}(B(0;1)). (7.4)

Thus for zK1z\in K_{1}, ψ2(z)K1(r)\psi_{2}(z)\in K_{1}(r) and by (7.3), (7.4) it follows that F2ψ2(z)F2(z)<δ.\|F_{2}\circ\psi_{2}(z)-F_{2}(z)\|<\delta. Hence from (7.2),

F2ψ2(z)ϕ1(z)<2δ<ρ1 for zK1.\|F_{2}\circ\psi_{2}(z)-\phi_{1}(z)\|<2\delta<\rho_{1}\text{ for }z\in K_{1}.

Also by (7.4), for zB(0;1)z\in B(0;1), ψ21F21F21(z)<ϵ<δ\|\psi_{2}^{-1}\circ F_{2}^{-1}-F_{2}^{-1}(z)\|<\epsilon<\delta. Again by (7.2),

ψ21F21(z)ϕ11(z)<2δ<ρ1 for zB(0;1).\|\psi_{2}^{-1}\circ F_{2}^{-1}(z)-\phi_{1}^{-1}(z)\|<2\delta<\rho_{1}\text{ for }z\in B(0;1).

Thus ϕ1\phi_{1} and ϕ2:=F2ψ2\phi_{2}:=F_{2}\circ\psi_{2} satisfies (7.1) for k=1.k=1.

Induction step: Suppose for N2N\geq 2, and there exist biholomorphisms ϕk:Ωh(nk)2\phi_{k}:\Omega_{h(n_{k})}\to\mathbb{C}^{2} such that (7.1) is satisfied for every 1kN1.1\leq k\leq N-1. Our goal is to construct ϕN+1\phi_{N+1} such that (7.1) holds for k=N.k=N. As before, for δ<ρN/2\delta<\rho_{N}/2, there exists FN+1Aut(2)F_{N+1}\in{\rm Aut}(\mathbb{C}^{2}) such that

ϕN(z)FN+1(z)<δ for zKN(r) and ϕN1(z)FN+11(z)\displaystyle\|\phi_{N}(z)-F_{N+1}(z)\|<\delta\text{ for }z\in K_{N}(r)\text{ and }\|\phi_{N}^{-1}(z)-F_{N+1}^{-1}(z)\| <δ for zB(0;N+r),\displaystyle<\delta\text{ for }z\in B(0;N+r),

where KN(r)K_{N}(r) is an rr-neighbourhood of KNK_{N}, contained in Ωh(kN)\Omega_{h(k_{N})} for some r>0r>0. Since FN+1F_{N+1} is uniformly continuous on KN(r)K_{N}(r), there exists ϵ0>0\epsilon_{0}>0 such that for z,wKN(r)z,w\in K_{N}(r)

FN+1(z)FN+1(w)<δ whenever zw<ϵ0.\displaystyle\|F_{N+1}(z)-F_{N+1}(w)\|<\delta\text{ whenever }\|z-w\|<\epsilon_{0}. (7.5)

Let ϵ<min{ϵ0,r,δ}.\epsilon<\min\{\epsilon_{0},r,\delta\}. Then again by [30, Lemma 4], there exists a biholomorphism ψN+1:Ωh(nN+1)2\psi_{N+1}:\Omega_{h(n_{N+1})}\to\mathbb{C}^{2} such that

ψN+1(z)z<ϵ for zKN and ψN+11(z)z<ϵ for zFN+11(B(0;N)).\displaystyle\|\psi_{N+1}(z)-z\|<\epsilon\text{ for }z\in K_{N}\text{ and }\|\psi_{N+1}^{-1}(z)-z\|<\epsilon\text{ for }z\in F_{N+1}^{-1}(B(0;N)). (7.6)

Thus for zKNz\in K_{N}, ψN+1(z)KN(r)\psi_{N+1}(z)\in K_{N}(r), and by (7.5), (7.6) it follows that

FN+1ψN+1(z)FN+1(z)<δ.\|F_{N+1}\circ\psi_{N+1}(z)-F_{N+1}(z)\|<\delta.

Hence FN+1ψN+1(z)ϕN(z)<2δ<ρN for zKN.\displaystyle\|F_{N+1}\circ\psi_{N+1}(z)-\phi_{N}(z)\|<2\delta<\rho_{N}\text{ for }z\in K_{N}. Also, similarly as above, by (7.6), for zB(0;N)z\in B(0;N), ψN+11FN+11(z)FN+11(z)<ϵ<δ\displaystyle\|\psi_{N+1}^{-1}\circ F_{N+1}^{-1}(z)-F_{N+1}^{-1}(z)\|<\epsilon<\delta and by assumption on FN+1F_{N+1},

ψN+11FN+11(z)ϕN1(z)<2δ<ρN.\|\psi_{N+1}^{-1}\circ F_{N+1}^{-1}(z)-\phi_{N}^{-1}(z)\|<2\delta<\rho_{N}.

Thus ϕN\phi_{N} and ϕN+1:=FN+1ψN+1\phi_{N+1}:=F_{N+1}\circ\psi_{N+1} satisfies (7.1) for k=N.k=N.

As {ρk}\{\rho_{k}\} is summable, the sequences {ϕk}\{\phi_{k}\} and {ϕk1}\{\phi^{-1}_{k}\} constructed converge on every compact subset of Ω{hk}\Omega_{\{h_{k}\}} and 2\mathbb{C}^{2}, i.e., there exist analytic limit maps ϕ:Ω{hk}2\phi:\Omega_{\{h_{k}\}}\to\mathbb{C}^{2} and ϕ~:22.\tilde{\phi}:\mathbb{C}^{2}\to\mathbb{C}^{2}. Since ϕ\phi is a limit of biholomorphisms, either ϕ\phi is one-one or Det Dϕ0\text{Det }D\phi\equiv 0 on Ω{hk}.\Omega_{\{h_{k}\}}.

Choose A>0A>0 and k1k\geq 1, sufficiently large, such that i=kρi<A/2\sum_{i=k}^{\infty}\rho_{i}<A/2. Also let K=ϕk1(B(0;A)K=\phi_{k}^{-1}(B(0;A). Then vol(K)>0\textsf{vol}(K)>0, and by (7.1), B(0;A/2)ϕ(K)B(0;3A/2)B(0;A/2)\subset\phi(K)\subset B(0;3A/2), i.e., vol(ϕ(K))>vol(B(0;A/2)).\textsf{vol}(\phi(K))>\textsf{vol}(B(0;A/2)). But if Det Dϕ0\text{Det }D\phi\equiv 0, then vol(ϕ(K))=0\textsf{vol}(\phi(K))=0, which is not true. Hence ϕ\phi is one-one on Ω{hk}.\Omega_{\{h_{k}\}}.

Finally, we prove that ϕ(Ω{hk})=2.\phi(\Omega_{\{h_{k}\}})=\mathbb{C}^{2}. So first, observe that as a consequence of Theorem 5.2 in [13], ϕn1\phi_{n}^{-1} converges uniformly to ϕ~\tilde{\phi} on compact subsets of 2\mathbb{C}^{2} and ϕ~1=ϕ\tilde{\phi}^{-1}=\phi on Ω{hk}.\Omega_{\{h_{k}\}}. Next we claim that for every positive integer N01N_{0}\geq 1, ϕ~(B(0,N0))Ω{hk}.\tilde{\phi}(B(0,N_{0}))\subset\Omega_{\{h_{k}\}}. Suppose not, then there exists z0ϕ~(B(0,N0))z_{0}\in\tilde{\phi}(B(0,N_{0})) such that 𝖦{hk}+(z0)>0.\mathsf{G}_{\{h_{k}\}}^{+}(z_{0})>0. Let w0=ϕ~1(z0)B(0;N0)w_{0}=\tilde{\phi}^{-1}(z_{0})\in B(0;N_{0}) and zk=ϕk1(w0).z_{k}=\phi_{k}^{-1}(w_{0}). Then zkΩh(k)z_{k}\in\Omega_{h(k)}, Gh(k)+(zk)=0G_{h(k)}^{+}(z_{k})=0, for every k1k\geq 1 and zkz0z_{k}\to z_{0} by (7.1). But by Theorem 6.5, Gh(k)+G_{h(k)}^{+} converges uniformly to 𝖦{hk}+\mathsf{G}^{+}_{\{h_{k}\}} on compact subsets of 2\mathbb{C}^{2}. Hence 𝖦{hk}+(z0)=0\mathsf{G}^{+}_{\{h_{k}\}}(z_{0})=0, which is a contradiction. Thus ϕ~(B(0,N0))int(𝐊{hk}+)\tilde{\phi}(B(0,N_{0}))\subset\textsf{int}(\mathbf{K}_{\{h_{k}\}}^{+}) with ϕ~(0)Ω{hk}\tilde{\phi}(0)\in\Omega_{\{h_{k}\}}. Hence by Lemma 7.1, ϕ~(2)Ω{hk}\tilde{\phi}(\mathbb{C}^{2})\subset\Omega_{\{h_{k}\}} or 2ϕ(Ω{hk})=2.\mathbb{C}^{2}\subset\phi(\Omega_{\{h_{k}\}})=\mathbb{C}^{2}.

Also, the following is immediate from the above proof, and the Remarks 6.8 and 6.9.

Corollary 7.4.

Let {𝖧k}\{\mathsf{H}_{k}\} be a sequence of Hénon maps of form (1.2), such that it satisfy

  • admits uniform filtration and bound conditions (i) and (ii) stated in Remark 6.8, and

  • is (upper) uniformly attracting on a neighbourhood of origin, i.e., satisfying (1.4).

Then the basin of attraction of the sequence {𝖧k}\{\mathsf{H}_{k}\} at the origin is biholomorphic to 2\mathbb{C}^{2}.

Further, for parametrised families of Hénon maps over compact manifolds, we have

Example 7.5.

Let MM be a compact complex manifold and :M×2M×2\mathcal{H}:M\times\mathbb{C}^{2}\to M\times\mathbb{C}^{2} be a skew product of Hénon map parametrised over MM, i.e., (λ,x,y)=(σ(λ),𝖧λ(x,y))\mathcal{H}(\lambda,x,y)=(\sigma(\lambda),\mathsf{H}_{\lambda}(x,y)) such that σ\sigma is an (holomorphic) endomorphism of MM and 𝖧λ\mathsf{H}_{\lambda} is a Hénon map of a fixed degree d2d\geq 2 for every λM\lambda\in M, i.e., the family {𝖧λ}λM\{\mathsf{H}_{\lambda}\}_{\lambda\in M} satisfies conditions (i) and (ii) of Remark 6.8. Further, if the family {Hλ}λM\{H_{\lambda}\}_{\lambda\in M} is uniformly attracting on a neighbourhood of origin, i.e, it satisfies (1.4), then for every p=(p1,0,0)p=(p_{1},0,0), the stable manifold Σs(p)\Sigma^{s}_{\mathcal{H}}(p), defined as

Σs(p):={(λ,z)M×2:k(λ,z)p as k},\Sigma^{s}_{\mathcal{H}}(p):=\{(\lambda,z)\in M\times\mathbb{C}^{2}:\mathcal{H}^{k}(\lambda,z)\to p\text{ as }k\to\infty\},

is biholomorphic to 2\mathbb{C}^{2}, provided it is non-empty. This, in fact also answers a few particular cases of Problem 38 and 39, stated in [1].

Finally, we conclude with an analytic property of the strong filled Julia set 𝒦𝒮+.\mathcal{K}_{\mathcal{S}}^{+}.

Proposition 7.6.

Suppose there exists 1ijn01\leq i\neq j\leq n_{0} such that K𝖧i+K𝖧j+K_{\mathsf{H}_{i}}^{+}\neq K_{\mathsf{H}_{j}}^{+}, then Ω{hk}𝒦𝒮+\Omega_{\{h_{k}\}}\not\subset\mathcal{K}_{\mathcal{S}}^{+}, the strong filled Julia set, for every {hk}𝒮.\{h_{k}\}\subset\mathcal{S}.

Proof.

We first claim that for every h𝒮h\in\mathcal{S}, Ωh𝒦𝒮+.\Omega_{h}\not\subset\mathcal{K}_{\mathcal{S}}^{+}. Note that by definition 𝒦𝒮+𝐊h+\mathcal{K}_{\mathcal{S}}^{+}\subset\mathbf{K}_{h}^{+} for every h𝒮.h\in\mathcal{S}. If Ωh𝒦𝒮+\Omega_{h}\subset\mathcal{K}_{\mathcal{S}}^{+} then Ωh=Jh+𝒦𝒮+.\partial\Omega_{h}=J_{h}^{+}\subset\mathcal{K}_{\mathcal{S}}^{+}. By assumption on 𝒮\mathcal{S}, there exists (at least one) g𝒮g\in\mathcal{S} such that Kg+Kh+.K_{g}^{+}\neq K_{h}^{+}. Now from the above Jh+𝒦𝒮+Kg+J_{h}^{+}\subset\mathcal{K}_{\mathcal{S}}^{+}\subset K_{g}^{+}, i.e., μh+\mu_{h}^{+} is a positive closed (1,1)(1,1) current supported on Kg+.K_{g}^{+}. Hence from Theorem 6.5 in [12], μh+=μg+\mu_{h}^{+}=\mu_{g}^{+} or Kh+=Kg+K_{h}^{+}=K_{g}^{+}, which is a contradiction to the assumption.

Now suppose there exists a sequence {hk}𝒮\{h_{k}\}\subset\mathcal{S}, such that Ω{hk}𝒦𝒮+.\Omega_{\{h_{k}\}}\subset\mathcal{K}_{\mathcal{S}}^{+}. Define the sequence h~k=h(k)h11\tilde{h}_{k}=h(k)\circ h_{1}^{-1}, for k2k\geq 2 Thus zΩ{hk}𝒦𝒮+z\in\Omega_{\{h_{k}\}}\subset\mathcal{K}_{\mathcal{S}}^{+}, h~k(z)\tilde{h}_{k}(z) is bounded, i.e., h(k)h11(z)h(k)\circ h_{1}^{-1}(z) is bounded. Let Fk(z):=h(k)h11(z)F_{k}(z):=\|h(k)\circ h_{1}^{-1}(z)\|. Then FkF_{k} is a sequence of positive pluri-subharmonic functions on Ω{hk}\Omega_{\{h_{k}\}}. Further on any compact subset of Ω{hk}\Omega_{\{h_{k}\}}, all FkF_{k}’s, except finitely many is bounded uniformly by R𝒮R_{\mathcal{S}}, where R𝒮R_{\mathcal{S}} is the radius of filtration of the semigroup 𝒮\mathcal{S} (as in Remark 2.4.

Next, let F(z)=lim supFk(z) for zΩ{hk}.F(z)=\limsup F_{k}(z)\text{ for }z\in\Omega_{\{h_{k}\}}. Hence from Theorem 2.6.3 in [22], the upper semicontinuous regularisation of FF, denoted by FF^{*} of FF is a bounded pluri-subharmonic function on Ω{hk}.\Omega_{\{h_{k}\}}. Also as F(z)=0F(z)=0 on B(0;r𝒮)B(0;r_{\mathcal{S}}), and the Lebesgue measure of the set {zΩ{hk}:F(z)F(z)}\{z\in\Omega_{\{h_{k}\}}:F(z)\neq F^{*}(z)\} is zero, it follows that F(z)=0F^{*}(z)=0 almost everywhere on B(0;r𝒮).B(0;r_{\mathcal{S}}). Since Ω{hk}\Omega_{\{h_{k}\}} is a Fatou-Bieberbach domain by Theorem 1.1, it cannot admit any non-constant bounded pluri-subharmonic function. Thus F0F^{*}\equiv 0 on Ω{hk}\Omega_{\{h_{k}\}}, i.e., h(k)(w)0h(k)(w)\to 0 for every wh11(Ω{hk})w\in h_{1}^{-1}(\Omega_{\{h_{k}\}}). Hence h11(Ω{hk})Ω{hk}.h_{1}^{-1}(\Omega_{\{h_{k}\}})\subset\Omega_{\{h_{k}\}}.

Let 𝒟B(0;r𝒮)Ω{hk}𝒦𝒮+\mathcal{D}\subset B(0;r_{\mathcal{S}})\subset\Omega_{\{h_{k}\}}\subset\mathcal{K}_{\mathcal{S}}^{+} be a relatively compact subset of an one-dimensional algebraic variety such that 𝒟Jh1=\partial{\mathcal{D}}\cap J_{h_{1}}^{-}=\emptyset and [𝒟]μh1=c0.[\mathcal{D}]\wedge\mu_{h_{1}}^{-}=c\neq 0. Thus by Corollary 1.7 of [6],

Sn=1dh1nh1n([𝒟])cμh1+ as n.S_{n}=\frac{1}{d_{h_{1}}^{n}}{h_{1}^{n}}^{*}([\mathcal{D}])\to c\mu_{h_{1}}^{+}\text{ as }n\to\infty.

Note SnS_{n}’s are positive (1,1)(1,1)-currents supported on h1n(𝒟)h1n(Ω{hk})Ω{hk}𝒦𝒮+h_{1}^{-n}(\mathcal{D})\subset h_{1}^{-n}(\Omega_{\{h_{k}\}})\subset\Omega_{\{h_{k}\}}\subset\mathcal{K}_{\mathcal{S}}^{+} (from the previous observation). Hence μh1+\mu_{h_{1}}^{+} is supported on 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+}, in particular Jh1+𝒦𝒮+J_{h_{1}}^{+}\subset\mathcal{K}_{\mathcal{S}}^{+} which is not possible from the claim above. ∎

Remark 7.7.

The above also proves that 𝒦𝒮+\mathcal{K}_{\mathcal{S}}^{+} cannot contain any (1,1)(1,1) positive closed current of finite mass and the positive Green’s function G𝒢+G_{\mathscr{G}}^{+} is unbounded on all — both autonomous and non-autonomous — basins of attraction.

References

  • [1] A. Abbondandolo, L. Arosio, J. E. Fornæss, P. Majer, H. Peters, J. Raissy and L. Vivas, A survey on non-autonomous basins in several complex variables, arXiv preprint arXiv:1311.3835 (2013).
  • [2] E. Andersén and L. Lempert, On the group of holomorphic automorphisms of n\mathbb{C}^{n}, Invent. Math. 110 (1992), no. 2, 371–388.
  • [3] E. Bedford, Open problem session of the Biholomorphic Mappings meeting at the American Institute of Mathematics, Palo Alto, CA, July 2000.
  • [4] E. Bedford and J. Smillie, Polynomial diffeomorphisms of 2\mathbb{C}^{2}: currents, equilibrium measure and hyperbolicity, Invent. Math. 103 (1991), no. 1, 69–99.
  • [5] by same author, Polynomial diffeomorphisms of 2\mathbb{C}^{2}. II: Stable manifolds and recurrence, Journal of the American Mathematical Society (1991), 657-679.
  • [6] by same author, Polynomial diffeomorphisms of 2\mathbb{C}^{2}. III. Ergodicity, exponents and entropy of the equilibrium measure, Math. Ann. 294 (1992), no. 3, 395–420.
  • [7] D. Boyd, An invariant measure for finitely generated rational semigroups, Complex Variables Theory Appl. 39 (1999), no. 3, 229–254.
  • [8] H. Brolin, Invariant sets under iteration of rational functions, Ark. Mat. 6 (1965), 103–144.
  • [9] J. P. Demailly, Complex analytic and differential geometry, Université de Grenoble I, 1997.
  • [10] T. C. Dinh and N. Sibony, Distribution des valeurs de transformations méromorphes et applications, Comment. Math. Helv. 81 (2006), no. 1, 221–258.
  • [11] by same author,Geometry of currents, intersection theory and dynamics of horizontal-like maps, Ann. Inst. Fourier 56 (2006), no. 2, 423–457.
  • [12] by same author, Rigidity of Julia sets for Hénon type maps, J. Mod. Dyn. 8 (2014), no. 3-4, 499–548.
  • [13] P. G. Dixon and J. Esterle, Michael’s problem and the Poincaré-Fatou-Bieberbach phenomenon, Bull. Amer. Math. Soc. (N.S.) 15 (1986), no. 2, 127–187.
  • [14] J. E. Fornæss and E. F. Wold, Non-autonomous basins with uniform bounds are elliptic, Proc. Amer. Math. Soc. 144 (2016), no. 11, 4709–4714.
  • [15] J. E. Fornæss, Short k\mathbb{C}^{k}, Complex analysis in several variables — Memorial Conference of Kiyoshi Oka’s Centennial Birthday, Adv. Stud. Pure Math.,42 (2004), Math. Soc. Japan, Tokyo, 95–108.
  • [16] J. E. Fornæss and B. Stensønes, Stable manifolds of holomorphic hyperbolic maps, Internat. J. Math. 15 (2004), no. 8, 749–758.
  • [17] S. Friedland and J. Milnor, Dynamical properties of plane polynomial automorphisms, Ergodic Theory Dynam. Systems 9 (1989), no. 1, 67–99.
  • [18] F. W. Gehring and G. J. Martin, Iteration theory and inequalities for Kleinian groups, Bull. Amer. Math. Soc. (N.S.) 21 (1989), no. 1, 57–63.
  • [19] D. Gilbarg and N. S. Trudinger, Elliptic partial differential equations of second order, Classics in Mathematics, Springer-Verlag, Berlin, 2001, Reprint of the 1998 edition.
  • [20] A. Hinkkanen and G. J. Martin, The dynamics of semigroups of rational functions. I, Proc. London Math. Soc. (3) 73 (1996), no. 2, 358–384.
  • [21] J. H. Hubbard and R. W. Oberste-Vorth, Hénon mappings in the complex domain. II. Projective and inductive limits of polynomials, Real and complex dynamical systems (Hillerød, 1993), NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., vol. 464, Kluwer Acad. Publ., Dordrecht, 1995, 89–132.
  • [22] M. Klimek, Pluripotential theory, London Mathematical Society Monographs (New Series), vol. 6, The Clarendon Press, Oxford University Press, New York, 1991, Oxford Science Publications.
  • [23] S. Lamy, L’alternative de Tits pour Aut[2]{\rm Aut}[\mathbb{C}^{2}], J. Algebra, 239 (2001), no. 2, 413–437.
  • [24] M. Londhe, Dynamics of polynomial semigroups: measures, potentials, and external fields, to appear in J. Anal. Math., arXiv preprint: arXiv:2004.02792.
  • [25] M. Lyubich, Entropy properties of rational endomorphisms of the Riemann sphere, Ergodic Theory Dynam. Systems 3 (1983), no. 3, 351–385.
  • [26] S. Morosawa, Y. Nishimura, M. Taniguchi, and T. Ueda, Holomorphic dynamics, Cambridge Studies in Advanced Mathematics, vol. 66, Cambridge University Press, Cambridge, 2000 (Translated from the 1995 Japanese original and revised by the authors).
  • [27] R. Pal and K. Verma, Dynamical properties of families of holomorphic mappings, Conform. Geom. Dyn. 19 (2015), 323–350.
  • [28] J. P. Rosay and W. Rudin, Holomorphic maps from n\mathbb{C}^{n} to n\mathbb{C}^{n}, Trans. Amer. Math. Soc. 310 (1988), no. 1, 47–86.
  • [29] N. Sibony, Dynamique des applications rationnelles de k\mathbb{P}^{k}, Dynamique et géométrie complexes (Lyon, 1997), ix–x, xi–xii, 97–185, Panor. Synthèses, 8, Soc. Math. France, Paris, 1999.
  • [30] W. F. Wold, Fatou-Bieberbach domains, Internat. J. Math. 16 (2005), no. 10, 1119–1130.