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

Finite and infinite degree Thurston maps with a small postsingular set

Nikolai Prochorov Aix-Marseille Université, CNRS, Institut de Mathématiques de Marseille, 13003 Marseille, France nikolai.prochorov@etu.univ-amu.fr, prochorov41@gmail.com
Abstract.

We develop the theory of Thurston maps that are defined everywhere on the topological sphere S2S^{2} with a possible exception of a single essential singularity. We establish an analog of the celebrated W. Thurston’s characterization theorem for a broad class of such Thurston maps having four postsingular values. To achieve this, we analyze the corresponding pullback maps defined on the one-complex dimensional Teichmüller space. This analysis also allows us to derive various properties of Hurwitz classes of the corresponding Thurston maps.

Key words and phrases:
Thurston maps, postsingularly finite holomorphic maps, Teichmüller spaces, moduli spaces, pullback maps, Levy cycles.
2020 Mathematics Subject Classification:
Primary 37F20; Secondary 37F10, 37F15, 37F34.

1. Introduction

1.1. Thurston theory for finite and infinite degree maps

In the one-dimensional rational dynamics, the crucial role is played by the family of postcritically finite (or pcf in short) rational maps, i.e., maps with all critical points being periodic or pre-periodic. In this context, one of the most influential ideas has been to abstract from the rigid underlying complex structure and consider the more general setup of postcritically finite branched self-coverings of the topological 22-sphere S2S^{2}. Nowadays, orientation-preserving pcf branched covering maps f:S2S2f\colon S^{2}\to S^{2} of topological degree deg(f)2\deg(f)\geq 2 are called Thurston maps (of finite degree), in honor of William Thurston, who introduced them to deepen the understanding of the dynamics of postcritically finite rational maps on ^\widehat{{\mathbb{C}}}.

These ideas can be extended to the transcendental setting to explore the dynamics of postsingularly finite (psf in short) meromorphic maps. A meromorphic map is called postsingularly finite if it has finitely many singular values, and each of them eventually becomes periodic or lands on the essential singularity under the iteration. We can generalize the notion of a Thurston map to include postsingularly finite topologically holomorphic non-injective maps f:XS2f\colon X\to S^{2}, where XX is a punctured topological sphere, ff does not extend continuously to the entire S2S^{2} and meets a technical condition of being a parabolic type map; see Sections 2.2 and 2.3. Note that in this case the map ff must be transcendental, meaning that it has infinite topological degree. For simplicity, we will use the notation f:S2S2f\colon S^{2}\dashrightarrow S^{2} to indicate that the Thurston map ff, whether finite or infinite degree, might not be defined at a single point of S2S^{2}.

For a Thurston map f:S2S2f\colon S^{2}\dashrightarrow S^{2}, the postsingular set PfP_{f} is defined as the union of all orbits of the singular values of the map ff. It is important to note that some of these orbits might terminate after several iterations, if a singular value reaches the point where the map ff is not defined. The elements of the postsingular set PfP_{f} are called the postsingular values of the Thurston map ff. If the map ff is defined on the entire sphere S2S^{2}, we simply refer to its postcritical set and postcritical values, as the set of singular values of ff coincides with the set of its critical values. Two Thurston maps are called combinatorially (or Thurston) equivalent if they are conjugate up to isotopy relative to their postsingular sets; see Definition 2.7.

A fundamental question in this context is whether a given Thurston map ff can be realized by a psf meromorphic map with the same combinatorics, that is, if ff is combinatorially equivalent to a psf meromorphic map. If the Thurston map ff is not realized, then we say that ff is obstructed. William Thurston answered this question for Thurston maps of finite degree in his celebrated characterization of rational maps: if a finite degree Thurston map f:S2S2f\colon S^{2}\to S^{2} has a hyperbolic orbifold (this is always true, except for some well-understood special maps), then ff is realized by a pcf rational map if and only if ff has no Thurston obstruction [DH93]. Such an obstruction is given by a finite collection of disjoint simple closed curves in S2PfS^{2}-P_{f} with certain invariance properties under the map ff. In many instances, it suffices to restrict to simpler types of Thurston obstructions provided by Levy cycles or even Levy fixed curves; see Definition 2.10, and [Hub16, Theorem 10.3.8], [HP22, Corollary 1.5], or [Par23, Theorems 7.6 and 8.6] for examples of such cases.

The same characterization question can be also asked in the transcendental setting. The first breakthrough in this area was obtained in [HSS09], where it was shown that an exponential Thurston map is realized if and only if it has no Levy cycle. In this context, the exponential Thurston map is defined as a Thurston map with two singular values, both of which are omitted, and one of which is the only essential singularity of the considered Thurston map. Furthermore, the results of [She22] and [MPR24] suggest that a Thurston-like criterion for realizability may hold in a greater generality. However, the characterization question in the transcendental setting remains largely open, as many of the techniques used in Thurston theory for finite degree maps do not extend to this context.

Thurston theory lays out the relationship between the topological properties of a map, its dynamics, and its geometry in terms of the existence of a holomorphic realization. Further- more, it is strongly connected with the combinatorial and algebraic aspects of the dynamics of pcf rational and psf meromorphic maps. The results mentioned above have substantial applications for both rational and transcendental dynamics. For instance, Thurston’s characterization result has allowed to classify various families of postcritically finite rational maps or finite degree Thurston maps in terms of combinatorial models [Poi93, Poi10, BLMW22], [DMRS19, LMS22], [Hlu19, HP22]. Building on the result of [HSS09], similar classifications were obtained in [LSV08] in terms of kneading sequences and in [PRS21] in terms of homotopic Hubbard treed for the family of postsingularly finite exponential maps. Moreover, the concept of a homotopic Hubbard tree was extended to general postsingularly finite entire maps in [Pfr19] (see also [PPS21]). It is plausible that a Thurston-like criterion is the final missing ingredient for the complete classification of the family of all psf entire maps.

1.2. Pullback maps

The key method in determining whether a given finite or infinite degree Thurston map f:S2S2f\colon S^{2}\dashrightarrow S^{2} with the postsingular set Pf=AP_{f}=A is realized by a psf meromorphic map is the analysis of the dynamics of a holomorphic operator σf\sigma_{f}, known as the pullback map, defined on a complex manifold called the Teichmüller space 𝒯A\mathcal{T}_{A}; see Sections 2.4 and 2.5. Crucially, the Thurston map ff is realized if and only if the pullback map σf\sigma_{f} has a fixed point in 𝒯A\mathcal{T}_{A}; see 2.17. Moreover, the dynamics of the pullback map encodes many other properties of the corresponding Thurston map; see, for instance, [BEKP09], [KPS16], and [BDP24].

Instead of working directly in the Teichmüller space 𝒯A\mathcal{T}_{A}, it is often more convenient to work in a simpler complex manifold A\mathcal{M}_{A}, known as a moduli space. This space, roughly speaking, encodes all possible complex structures biholomorhic to a punctured Riemann sphere that can be put on the punctured topological sphere S2AS^{2}-A (see Section 2.4 for the precise definition). There is a natural projection map π:𝒯AA\pi\colon\mathcal{T}_{A}\to\mathcal{M}_{A} that is a holomorphic universal covering. However, the map σf\sigma_{f} rarely descends to a map on the moduli space A\mathcal{M}_{A}. Nevertheless, for a finite degree Thurston map f:S2S2f\colon S^{2}\to S^{2}, there exists a complex manifold 𝒲f\mathcal{W}_{f}, known as the Hurwitz space of the Thurston map ff, along with the holomorphic XX-map Xf:𝒲fAX_{f}\colon\mathcal{W}_{f}\to\mathcal{M}_{A}, a holomorphic covering map Yf:𝒲fAY_{f}\colon\mathcal{W}_{f}\to\mathcal{M}_{A}, called the YY-map, and a holomorphic covering ωf:𝒯A𝒲f\omega_{f}\colon\mathcal{T}_{A}\to\mathcal{W}_{f}, such that the following diagram commutes [Koc13, Section 2]:

𝒯A{\mathcal{T}_{A}}𝒯A{\mathcal{T}_{A}}𝒲f{\mathcal{W}_{f}}A{\mathcal{M}_{A}}A{\mathcal{M}_{A}}σf\scriptstyle{\sigma_{f}}π\scriptstyle{\pi}ωf\scriptstyle{\omega_{f}}π\scriptstyle{\pi}Yf\scriptstyle{Y_{f}}Xf\scriptstyle{X_{f}}
Figure 1. Fundamental diagram.

In other words, the pullback map σf\sigma_{f} is semi-conjugate to a self-correspondence XfYf1X_{f}\circ Y_{f}^{-1} of the moduli space A\mathcal{M}_{A}. If f:S2S2f\colon S^{2}\to S^{2} with Pf=AP_{f}=A is a finite degree Thurston map, then the YY-map Yf:𝒲fAY_{f}\colon\mathcal{W}_{f}\to\mathcal{M}_{A} has a finite topological degree; see [Koc13, Theorem 2.6]. This observation plays a crucial role in the proof of Thurston’s characterization of rational maps. In fact, it allows to conclude that for a Thurston map ff with a hyperbolic orbifold, the σf\sigma_{f}-orbit (σfn(τ))(\sigma_{f}^{\circ n}(\tau)) of τ𝒯A\tau\in\mathcal{T}_{A} converges (indicating that ff is realized), if the projection (π(σfn(τ)))(\pi(\sigma_{f}^{\circ n}(\tau))) of this orbit visits some compact set of the moduli space A\mathcal{M}_{A} infinitely many times; see [Hub16, Section 10.9 and Lemma 10.11.9] and [Sel12, Proof of Theorem 2.3, p. 20].

The objects introduced above, along with commutative diagram (1) and the fact that Yf:𝒲fAY_{f}\colon\mathcal{W}_{f}\to\mathcal{M}_{A} has a finite degree, have broad applications beyond the proof of Thurston’s characterization of rational maps; see, for example, [BN06, BEKP09, Sel12, Sel13, Koc13, Lod13, KPS16, FKK+17, Smi24a, Smi24b, BDP24]. These tools, for instance, allow to simultaneously study the entire Hurwitz (equivalence) class f{\mathcal{H}}_{f} of the finite degree Thurston map ff. Here, two Thurston maps f1:S2S2f_{1}\colon S^{2}\dashrightarrow S^{2} and f2:S2S2f_{2}\colon S^{2}\dashrightarrow S^{2} with Pf1=Pf2P_{f_{1}}=P_{f_{2}} are said to be Hurwitz equivalent if there exist orientation-preserving homeomorphisms ϕ1,ϕ2:S2S2\phi_{1},\phi_{2}\colon S^{2}\to S^{2} such that ϕ1|Pf1=ϕ2|Pf1\phi_{1}|P_{f_{1}}=\phi_{2}|P_{f_{1}} and ϕ1f1=f2ϕ2\phi_{1}\circ f_{1}=f_{2}\circ\phi_{2}; see [BN06, Lod13, KPS16] for examples of results on Hurwitz classes. For instance, one can pose a question whether a given Thurston map is totally unobstructed, i.e., f{\mathcal{H}}_{f} consists of only realized Thurston maps.

Commutative diagram (1) is particularly powerful in two specific cases. The first is when the Thurston map f:S2S2f\colon S^{2}\to S^{2} has the postcritical set Pf=AP_{f}=A consisting of exactly four points. In this situation, the spaces 𝒯A\mathcal{T}_{A}, A\mathcal{M}_{A}, and 𝒲f\mathcal{W}_{f} are simply Riemann surfaces. In fact, the Teichmüller space is biholomorphic to the unit disk 𝔻{\mathbb{D}} and the moduli space A\mathcal{M}_{A} is biholomorphic to the three punctured Riemann sphere Σ=^{0,1,}\Sigma=\widehat{{\mathbb{C}}}-\{0,1,\infty\}. This allows the use of powerful machinery of one-dimensional holomorphic dynamics to study pullback maps. For example, this approach was utilized in [Smi24a] to derive an alternative proof of Thurston’s characterization of rational maps in the case of four postcritical values, as well as in [Smi24a, Smi24b] to investigate the global curve attractor conjecture, which was ultimately resolved in [BDP24] for all pcf rational maps with four postcritical values. Secondly, when the XX-map XfX_{f} is injective, the “inverse” of σf\sigma_{f} descends to the so-called gg-map gf:=YfXf1g_{f}:=Y_{f}\circ X_{f}^{-1}, which is defined on the subset Xf(𝒲f)X_{f}(\mathcal{W}_{f}) of the moduli space A\mathcal{M}_{A}; see [Koc13, Section 5] for examples of finite degree Thurston maps that satisfy this condition.

The theory of moduli maps, as outlined above, is developed for finite degree Thurston maps but has not yet been established in the context of transcendental Thurston maps. In this paper, we consider a certain family of Thurston maps, that includes maps of both finite and infinite degree, with four postsingular values and show that the corresponding pullback maps admit an analogue of commutative diagram (1), where the XX-map is always injective, but the YY-map has infinite degree if the initial Thurston map f:S2S2f\colon S^{2}\dashrightarrow S^{2} is transcendental. Using tools of one-dimensional holomorphic dynamics and hyperbolic geometry, we establish a Thurston-like realizability criterion for this family of maps. In particular, we demonstrate that the obstacle of the YY-map having infinite degree can be finessed. Afterward, we illustrate how the developed machinery can be used to investigate the properties of the corresponding Hurwitz classes.

1.3. Main results

In this paper, we study the family of Thurston maps f:S2S2f\colon S^{2}\dashrightarrow S^{2} satisfying the following conditions:

  1. (A)

    the map ff has at most three singular values;

  2. (B)

    the postsingular set PfP_{f} consists of exactly four points;

  3. (C)

    there exists a set BPfB\subset P_{f} such that |B|=3|B|=3, SfBS_{f}\subset B, and |f1(B)¯Pf|=3|\overline{f^{-1}(B)}\cap P_{f}|=3.

Here, f1(B)¯\overline{f^{-1}(B)} is the closure of the set f1(B)f^{-1}(B) in the topology of S2S^{2}. In particular, it coincides with f1(B)f^{-1}(B) if the Thurston map ff has a finite degree; otherwise, it also includes the point where ff is not defined.

Clearly, conditions (B) and (C) are independent from the function-theoretical properties of the map ff. More specifically, if the map ff has at most three singular values, then these conditions can be verified by analyzing the dynamics of the map ff on the finite set PfP_{f}. For instance, in the case of entire Thurston maps (those that can be restricted to self-maps of 2{\mathbb{R}}^{2}; see Section 2.2) with three singular and four postsingular values, three out of seven possible postsingular portraits satisfy condition (C); see Example 4.13. More examples of families of Thurston maps that meet these conditions can be found in Section 4.3.

Although conditions (A)(C) are quite restrictive, there are still uncountably many pairwise combinatorially inequivalent both realized and obstructed Thurston maps that meet them; see Remark 4.11. Notably, these conditions are preserved under Hurwitz equivalence of Thurston maps. Finally, it is worth mentioning that all of our further results work in a slightly more general setting of marked Thurston maps satisfying analogous properties to (A)(C); see Sections 2.3 and 4.

1.3.1. Characterization problem

We establish an analog of Thurston’s characterization result for the family of Thurston maps that satisfy conditions (A)(C). In fact, we show that it is sufficient to consider one of the simplest types of Thurston obstructions – Levy fixed curves – to determine whether such a Thurston map is realized. For a Thurston map f:S2S2f\colon S^{2}\dashrightarrow S^{2}, a Levy fixed curve is a simple closed curve γS2Pf\gamma\subset S^{2}-P_{f} such that γ\gamma is essential, i.e., it cannot be shrinked to a point by a homotopy in S2PfS^{2}-P_{f}, and there exists another simple closed curve γ~f1(γ)\widetilde{\gamma}\subset f^{-1}(\gamma) such that γ\gamma and γ~\widetilde{\gamma} are homotopic in S2PfS^{2}-P_{f} and f|γ~:γ~γf|\widetilde{\gamma}\colon\widetilde{\gamma}\to\gamma is a homeomorphism. If the map ff is injective on one of the connected components of S2γ~S^{2}-\widetilde{\gamma}, then we say that the Levy fixed curve γ\gamma is weakly degenerate.

Main Theorem A.

Let f:S2S2f\colon S^{2}\dashrightarrow S^{2} be a Thurston map of finite or infinite degree that satisfies conditions (A)(C). Then ff is realized if and only if it has no weakly degenerate Levy fixed curve. Moreover, if ff is obstructed, then it has a unique Levy fixed curve up to homotopy in S2PfS^{2}-P_{f}; otherwise, it is realized by a psf meromorphic map that is unique up to Möbuis conjugation.

To prove Main Theorem A, we start by showing that the corresponding pullback map σf\sigma_{f} defined on the Teichmüller space 𝒯A𝔻\mathcal{T}_{A}\sim{\mathbb{D}}, where A=PfA=P_{f}, admits an analog of commutative diagram (1), where the XX-map is injective, the YY-map is a covering, potentially of infinite degree, and the analog of the Hurwitz space 𝒲f\mathcal{W}_{f} is a finitely or countably punctured Riemann sphere (see Proposition 4.1 and Remark 4.2). Further analysis reveals a crucial observation similar to that in the proof of Thurston’s characterization result: if the sequence (π(σfn(τ)))(\pi(\sigma_{f}^{\circ n}(\tau))) with τ𝒯A\tau\in\mathcal{T}_{A} visits a certain compact set of the moduli space A\mathcal{M}_{A} infinitely many times, then the σf\sigma_{f}-orbit (σfn(τ))(\sigma_{f}^{\circ n}(\tau)) of τ\tau converges to the unique fixed point of σf\sigma_{f} (see Claim 1 of the proof of Theorem 3.10).

Moreover, we establish a more refined result: if the pullback map σf\sigma_{f} does not have a fixed point (indicating that the Thurston map ff is obstructed), then the sequence (π(σfn(τ)))(\pi(\sigma_{f}^{\circ n}(\tau))) converges to the same “cusp” xAΣ={0,1,}x\in\partial\mathcal{M}_{A}\sim\partial\Sigma=\{0,1,\infty\} of the moduli space A\mathcal{M}_{A}, regardless of the choice of τ𝒯A\tau\in\mathcal{T}_{A}. Furthermore, the map gf=YfXf1g_{f}=Y_{f}\circ X_{f}^{-1} can be holomorphically extended to a neighborhood of this cusp and xx becomes a repelling fixed point of gfg_{f}. It is worth noting that these results hold not only for pullback maps, but also in a broader class of holomorphic self-maps of the unit disk; see Theorem 3.10. Finally, this analysis provides a sufficient control over the dynamics of σf\sigma_{f} to derive the existence of a Levy fixed curve for the obstructed Thurston map f:S2S2f\colon S^{2}\dashrightarrow S^{2}.

Main Theorem A is the first result addressing the characterization problem in the transcendental setting that is established with minimal reliance on the function-theoretical properties of the considered Thurston maps. This differs from [HSS09], which focuses on exponential Thurston maps, and [She22], which is primarily devoted to structurally finite Thurston maps. Note that the geometric and analytic properties of entire or meromorphic maps, even those with few singular values, can be highly varied and subtle; see [Bis15a, Bis15b, Bis17].

Main Theorem A also offers an alternative proof for the characterization of postsingularly finite exponential maps [HSS09, Theorem 2.4] for the case of four postsingular values; see Example 4.12. While the proof in [HSS09] relies on the intricate machinery of integrable quadratic differentials and their thick-thin decompositions, our approach uses more explicit techniques that shed the light on the geometry of the pullback dynamics on Teichmüller and moduli spaces. At the same time, Main Theorem A provides a novel proof of Thurston’s characterization of rational maps within a broad class of examples, and this proof does not rely on the fact that YY-map has a finite degree.

1.3.2. Hurwitz classes

Techniques explained in Section 1.3.1 also allow us to derive several properties of Hurwitz classes:

Main Theorem B.

Let f:S2S2f\colon S^{2}\dashrightarrow S^{2} be a Thurston map of finite or infinite degree that satisfies conditions (A)(C). Then

  1. (1)

    ff is totally unobstructed if and only if there are no two points a,bPfa,b\in P_{f} such that deg(f,a)=deg(f,b)=1\deg(f,a)=\deg(f,b)=1 and f({a,b})f(\{a,b\}) equals {a,b}\{a,b\} or Pf{a,b}P_{f}-\{a,b\};

  2. (2)

    if ff is not totally unobstructed, then its Hurwitz class f{\mathcal{H}}_{f} contains infinitely many pairwise combinatorially inequivalent obstructed Thurston maps;

  3. (3)

    if ff has infinite degree, then its Hurwitz class f{\mathcal{H}}_{f} contains infinitely many pairwise combinatorially inequivalent realized Thurston maps.

The main tool for proving Main Theorem B is the relationship between fixed points of the map gf=YfXf1g_{f}=Y_{f}\circ X_{f}^{-1} and the elements of the Hurwitz class f{\mathcal{H}}_{f}. Let f:S2S2f\colon S^{2}\dashrightarrow S^{2} be a Thurston map with Pf=AP_{f}=A satisfying assumptions (A)(C). We show that if the corresponding map gfg_{f} can be holomorphically extended to a neighborhood of xAΣ={0,1,}x\in\partial\mathcal{M}_{A}\sim\partial\Sigma=\{0,1,\infty\} and xx becomes a repelling fixed point of gfg_{f}, then the Hurwitz class f{\mathcal{H}}_{f} must contain an obstructed Thurston map. Moreover, ff is totally unobstructed if and only if none of the “cusps” of the moduli space A\mathcal{M}_{A} exhibit this behavior; see Proposition 4.6. Using additional properties of the map gfg_{f}, we establish a simple criterion, as in item (1) of Main Theorem B, for determining whether a given Thurston map with properties (A)(C) is totally unobstructed. Furthermore, with the understanding of possible obstructions provided by Main Theorem A, we can construct infinitely many pairwise combinatorially inequivalent obstructed Thurston maps within the Hurwitz class f{\mathcal{H}}_{f} starting from just one of them.

To prove item (3) of Main Theorem B, we show that a fixed point xAx\in\mathcal{M}_{A} of the map gfg_{f} corresponds to a realized Thurston map within the Hurwitz class f{\mathcal{H}}_{f}; see Proposition 4.5. Furthermore, only finitely many fixed points of gfg_{f} can correspond to the same Thurston map up to combinatorial equivalence (see the proof of Theorem 4.8). The desired result then follows because the map gfg_{f} has infinitely many fixed points when ff is transcendental; see Proposition 4.1 and Lemma A.2.

Similar connections between the fixed points of the map gfg_{f} and Thurston maps within the Hurwitz class f{\mathcal{H}}_{f} are already established in the context of finite degree Thurston maps (cf. [Koc13, Propositions 4.3 and 4.4] and [KPS16, Theorem 9.1]). However, their extensions to the setting of transcendental Thurston maps are novel contributions. Additionally, as an application of Main Theorem B, we can obtain the following result regarding the structure of parameter spaces of finite-type meromorphic maps (see Definition 4.9):

Corollary 1.1.

Let g:^g\colon{\mathbb{C}}\to\widehat{{\mathbb{C}}} be a transcendental meromorphic map such that |Sg|3|S_{g}|~{}\leq~{}3. Then its parameter space Par(g)\mathrm{Par}(g) contains infinitely many pairwise (topologically or conformally) non-conjugate psf maps with four postsingular values.

1.4. Organization of the paper

Our paper is organized as follows. In Section 2, we review some general background. In Section 2.1, we fix the notation and state some basic definitions. We discuss topologically holomorphic maps in Section 2.2. The necessary background on Thurston maps is covered in Section 2.3. Section 2.4 introduces the Teichmüller and moduli spaces of a marked topological sphere. Finally, in Section 2.5, we define pullback maps, discuss their basic properties and their relations with the associated Thurston maps.

In Section 3, we present several results concerning the hyperbolic geometry and dynamics of holomorphic self-maps of the unit disk. Section 3.1 provides tools for identifying obstructions for Thurston maps with four postsingular values. We establish some estimates for hyperbolic contraction of inclusion maps between two hyperbolic Riemann surfaces in Section 3.2. In Section 3.3, we investigate dynamics of holomorphic self-maps of the unit disk satisfying certain additional assumptions.

Further, in Section 4, we develop the Thurston theory for a family of Thurston maps satisfying condition (A)(C). In particular, in Section 4.1, we address the characterization problem for this class of Thurston maps and prove Main Theorem A. We study properties of Hurwitz classes and prove Main Theorem B and Corollary 1.1 in Section 4.2. Finally, in Section 4.3, we provide and analyze various examples.

Acknowledgments. I would like to express my deep gratitude to my thesis advisor, Dierk Schleicher, for introducing me to the fascinating world of Transcendental Thurston Theory. I am also profoundly thankful to Kevin Pilgrim and Lasse Rempe for the their valuable suggestions and for many helpful and inspiring discussions. I would like to thank Centre National de la Recherche Scientifique (CNRS) for supporting my visits to the University of Saarland and University of Liverpool, where these conversations took place. Special thanks go to Anna Jové and Zachary Smith for the discussions on the dynamics of holomorphic self-maps of the unit disk and their diverse applications.

2. Background

2.1. Notation and basic concepts

The sets of positive integers, non-zero integers, integers, real and complex numbers are denoted by {\mathbb{N}}, {\mathbb{Z}}^{*}, {\mathbb{Z}}, {\mathbb{R}}, and {\mathbb{C}}, respectively. We use the notation 𝕀:=[0,1]{\mathbb{I}}:=[0,1] for the closed unit interval on the real line, 𝔻:={z:|z|<1}{\mathbb{D}}:=\{z\in{\mathbb{C}}\colon|z|<1\} for the open unit disk in the complex plane, 𝔻:=𝔻{0}{\mathbb{D}}^{*}:={\mathbb{D}}-\{0\} for the punctured unit disk, :={0}{\mathbb{C}}^{*}:={\mathbb{C}}-\{0\} for the punctured complex plane, :={z:Im(z)>0}{\mathbb{H}}:=\{z\in{\mathbb{C}}:\operatorname{Im}(z)>0\} for the upper half-plane, ^:={}\widehat{{\mathbb{C}}}:={\mathbb{C}}\cup\{\infty\} for the Riemann sphere, and Σ\Sigma for the three-punctured Riemann sphere ^{0,1,}\widehat{{\mathbb{C}}}-\{0,1,\infty\}. The open and closed disks of radius r>0r>0 centered at 0 are denoted by 𝔻r{\mathbb{D}}_{r} and 𝔻¯r\overline{{\mathbb{D}}}_{r}, respectively. Finally, arg(z)[0,2π)\arg(z)\in[0,2\pi) and |z||z| denote the argument and the absolute value, respectively, of the complex number zz.

We denote the oriented 2-dimensional sphere by S2S^{2}. In this paper, we treat it as a purely topological object. In particular, our convention is to write g:^g\colon{\mathbb{C}}\to\widehat{{\mathbb{C}}} or g:^^g\colon\widehat{{\mathbb{C}}}\to\widehat{{\mathbb{C}}} to indicate that gg is holomorphic, and f:S2S2f\colon S^{2}\to S^{2} if ff is only continuous. The same rule applies to the notation g:^^g\colon\widehat{{\mathbb{C}}}\dashrightarrow\widehat{{\mathbb{C}}} and f:S2S2f\colon S^{2}\dashrightarrow S^{2} (see Section 2.2 for the details).

The cardinality of a set AA is denoted by |A||A| and the identity map on AA by idA\operatorname{id}_{A}. If f:UVf\colon U\to V is a map and WUW\subset U, then f|Wf|W stands for the restriction of ff to WW. If UU is a topological space and WUW\subset U, then W¯\overline{W} denotes the closure and W\partial W the boundary of WW in UU.

A subset DD of ^\widehat{{\mathbb{C}}} is called an open Jordan region if there exists an injective continuous map φ:𝔻¯^\varphi\colon\overline{{\mathbb{D}}}\to\widehat{{\mathbb{C}}} such that D=φ(𝔻)D=\varphi({\mathbb{D}}). In this case, D=φ(𝔻)\partial D=\varphi(\partial{\mathbb{D}}) is a simple closed curve in ^\widehat{{\mathbb{C}}}. A domain U^U\subset\widehat{{\mathbb{C}}} is called an annulus if ^U\widehat{{\mathbb{C}}}-U has two connected components. The modulus of an annulus UU is denoted by mod(U)\mathrm{mod}(U) (see [Hub06, Proposition 3.2.1] for the definition).

Let UU and VV be topological spaces. A continuous map H:U×𝕀VH\colon U\times{\mathbb{I}}\to V is called a homotopy from UU to VV. When U=VU=V, we simply say that HH is a homotopy in UU. Given a homotopy H:U×𝕀VH\colon U\times{\mathbb{I}}\to V, for each t𝕀t\in{\mathbb{I}}, we associate the time-tt map Ht:=H(t,):UVH_{t}:=H(t,\cdot)\colon U\to V. Sometimes it is convenient to think of the homotopy HH as a continuous family of its time maps (Ht)t𝕀(H_{t})_{t\in{\mathbb{I}}}. The homotopy HH is called an (ambient) isotopy if the map Ht:UVH_{t}\colon U\to V is a homeomorphism for each t𝕀t\in{\mathbb{I}}. Suppose AA is a subset of UU. An isotopy H:U×𝕀VH\colon U\times{\mathbb{I}}\to V is said to be an isotopy relative to AA (abbreviated “HH is an isotopy rel. AA”) if Ht(p)=H0(p)H_{t}(p)=H_{0}(p) for all pAp\in A and t𝕀t\in{\mathbb{I}}.

Given M,NUM,N\subset U, we say that MM is homotopic (in UU) to NN if there exists a homotopy H:U×𝕀UH\colon U\times{\mathbb{I}}\to U with H0=idUH_{0}=\operatorname{id}_{U} and H1(M)=NH_{1}(M)=N. Two homeomorphisms φ0,φ1:UV\varphi_{0},\varphi_{1}\colon U\to V are called isotopic (rel. AUA\subset U) if there exists an isotopy H:U×𝕀VH\colon U\times{\mathbb{I}}\to V (rel. AA) with H0=φ0H_{0}=\varphi_{0} and H1=φ1H_{1}=\varphi_{1}.

We assume that every topological surface XX is oriented. We denote by Homeo+(X)\operatorname{Homeo}^{+}(X) and Homeo+(X,A)\operatorname{Homeo}^{+}(X,A) the group of all orientation-preserving self-homeomorphisms of XX and the group of all orientation-preserving self-homeomorphisms of XX fixing AA pointwise, respectively. We use the notation Homeo0+(X,A)\operatorname{Homeo}^{+}_{0}(X,A) for the subgroup of Homeo+(X,A)\operatorname{Homeo}^{+}(X,A) consisting of all homeomorphisms isotopic rel. AA to idX\operatorname{id}_{X}.

2.2. Topologically holomorphic maps

In this section, we briefly recall the definition of a topologically holomorphic map and some of its basic properties (for more detailed discussion see [MPR24, Section 2.3]; see also [LP20]).

Definition 2.1.

Let XX and YY be two connected topological surfaces. A map f:XYf\colon X\to Y is called topologically holomorphic if it satisfies one of the following four equivalent conditions:

  1. (1)

    for every pXp\in X there exist dd\in{\mathbb{N}}, a neighborhood UU of xx, and two orientation-preserving homeomorphisms ψ:U𝔻\psi\colon U\to{\mathbb{D}} and φ:f(U)𝔻\varphi\colon f(U)\to{\mathbb{D}} such that ψ(p)=φ(f(p))=0\psi(p)=\varphi(f(p))=0 and (φfψ1)(z)=zd(\varphi\circ f\circ\psi^{-1})(z)=z^{d} for all z𝔻z\in{\mathbb{D}};

  2. (2)

    ff is continuous, open, discrete (i.e., f1(q)f^{-1}(q) is discrete in XX for very qYq\in Y), and for every pXp\in X such that ff is locally injective at pp, there exists a neighborhood UU of pp for which f|U:Uf(U)f|U\colon U\to f(U) is an orientation-preserving homeomorphism;

  3. (3)

    there exist Riemann surfaces SXS_{X} and SYS_{Y} and orientation-preserving homeomorphisms φ:YSY\varphi\colon Y\to S_{Y} and ψ:XSX\psi\colon X\to S_{X} such that φfψ1:SXSY\varphi\circ f\circ\psi^{-1}\colon S_{X}\to S_{Y} is a holomorphic map;

  4. (4)

    for every orientation-preserving homeomorphism φ:YSY\varphi\colon Y\to S_{Y}, where SYS_{Y} is a Riemann surface, there exist a Riemann surface SXS_{X} and an orientation-preserving homeomorphism ψ:XSX\psi\colon X\to S_{X} such that φfψ1:SXSY\varphi\circ f\circ\psi^{-1}\colon S_{X}\to S_{Y} is a holomorphic map.

Note that in condition (4) of Definition 2.1, the homeomorphism ψ\psi is defined uniquely up to post-composition with a conformal automorphism of SXS_{X} for fixed φ\varphi and SXS_{X}.

It is straightforward to define the concepts of regular, singular, critical, and asymptotic values, as well as regular and critical points and their local degrees (denoted by deg(f,)\deg(f,\cdot)) for the topologically holomorphic map ff (see [MPR24, Definition 2.7]). We denote by SfYS_{f}\subset Y the singular set of ff, i.e., the union of all singular values of the topologically holomorphic map f:XYf\colon X\to Y. We say that the map ff is of finite type or belongs to the Speiser class 𝒮\mathcal{S} if the set SfS_{f} is finite.

In this paper, we study topologically holomorphic maps f:XS2f\colon X\to S^{2}, where XX is either the sphere S2S^{2} or the punctured sphere S2{e}S^{2}-\{e\}. In the latter case, we assume that ff cannot be extended as a topologically holomorphic map to the entire sphere S2S^{2}. For the sake of simplicity, we are going to use the notation f:S2S2f\colon S^{2}\dashrightarrow S^{2} in order to indicate that ff might not be defined at a single point eS2e\in S^{2}. Similar to the holomorphic case, the point ee is referred as the essential singularity of the map ff. Likewise, for a holomorphic map gg defined everywhere on ^\widehat{{\mathbb{C}}} with the possible exception of a single essential singularity, we use the notation g:^^g\colon\widehat{{\mathbb{C}}}\dashrightarrow\widehat{{\mathbb{C}}}, and we say that g:^^g\colon\widehat{{\mathbb{C}}}\dashrightarrow\widehat{{\mathbb{C}}} is holomorphic.

It is possible to derive the following isotopy lifting property for topologically holomorphic maps as above in the case when they are of finite type (cf. [ERG15, Propostion 2.3]).

Proposition 2.2.

Let f:XS2f\colon X\to S^{2} and f~:X~S2\widetilde{f}\colon\widetilde{X}\to S^{2} be topologically holomorphic maps of finite type, where XX and X~\widetilde{X} are either topological spheres or punctured topological spheres. Suppose that φ0f=f~ψ0\varphi_{0}\circ f=\widetilde{f}\circ\psi_{0} for some φ0,ψ0Homeo+(S2)\varphi_{0},\psi_{0}\in\operatorname{Homeo}^{+}(S^{2}). Let AS2A\subset S^{2} be a finite set containing SfS_{f} and φ1Homeo+(S2)\varphi_{1}\in\operatorname{Homeo}^{+}(S^{2}) is isotopic rel. AA to φ0\varphi_{0}. Then φ1f=f~ψ1\varphi_{1}\circ f=\widetilde{f}\circ\psi_{1} for some ψ1Homeo+(S2)\psi_{1}\in\operatorname{Homeo}^{+}(S^{2}) isotopic rel. f1(A)(S2X)f1(A)¯f^{-1}(A)\cup(S^{2}-X)\supset\overline{f^{-1}(A)} to ψ0\psi_{0}.

Proof.

Let (φt)t𝕀(\varphi_{t})_{t\in{\mathbb{I}}} be the corresponding isotopy. From the definition of a singular value, it follows that the restrictions φtf|Y:YZ\varphi_{t}\circ f|Y\colon Y\to Z are covering maps for each t𝕀t\in{\mathbb{I}}, where Y:=Xf1(A)Y:=X-f^{-1}(A) and Z:=S2AZ:=S^{2}-A. Therefore, [ABF21, Lemma 2.7] implies the existence of an isotopy (ϕt)t𝕀(\phi_{t})_{t\in{\mathbb{I}}} in YY such that φtf=φ0fϕt\varphi_{t}\circ f=\varphi_{0}\circ f\circ\phi_{t}. Each homeomorphism ϕt:YY\phi_{t}\colon Y\to Y extends to a self-homeomorphism of the entire sphere S2S^{2} since all but at most one point of the set S2YS^{2}-Y are isolated. Moreover, it is straightforward to check that ϕt|f1(A)(S2X)=ϕ0|f1(A)(S2X)\phi_{t}|f^{-1}(A)\cup(S^{2}-X)=\phi_{0}|f^{-1}(A)\cup(S^{2}-X) for each t𝕀t\in{\mathbb{I}}. In other words, the homotopy (ϕt)t𝕀(\phi_{t})_{t\in{\mathbb{I}}} can be viewed as an isotopy in S2S^{2} rel. f1(A)(S2X)f^{-1}(A)\cup(S^{2}-X).

At the same time, φ0f=φ0fϕ0\varphi_{0}\circ f=\varphi_{0}\circ f\circ\phi_{0} and, therefore, we have the following:

φ1f=φ0fϕ1=(φ0fϕ0)ϕ01ϕ1=φ0fϕ01ϕ1=f~(ψ0ϕ01ϕ1).\varphi_{1}\circ f=\varphi_{0}\circ f\circ\phi_{1}=(\varphi_{0}\circ f\circ\phi_{0})\circ\phi_{0}^{-1}\circ\phi_{1}=\varphi_{0}\circ f\circ\phi_{0}^{-1}\circ\phi_{1}=\widetilde{f}\circ(\psi_{0}\circ\phi_{0}^{-1}\circ\phi_{1}).

Thus, we can set ψ1:=ψ0ϕ01ϕ1\psi_{1}:=\psi_{0}\circ\phi_{0}^{-1}\circ\phi_{1}, and (ψ0ϕ01ϕt)t𝕀(\psi_{0}\circ\phi_{0}^{-1}\circ\phi_{t})_{t\in{\mathbb{I}}} provides the required isotopy rel. f1(A)(S2X)f^{-1}(A)\cup(S^{2}-X). Clearly, if pS2p\in S^{2} is an accumulation point of the set f1(A)f^{-1}(A), then pXp\not\in X, which implies f1(A)¯f1(A)(S2X)\overline{f^{-1}(A)}\subset f^{-1}(A)\cup(S^{2}-X). Finally, ψ1\psi_{1} is orientation-preserving since ff and f~\widetilde{f} are local orientation-preserving homeomorphisms outside the sets of their critical points. ∎

Corollary 2.3.

Let f:XS2f\colon X\to S^{2} be a topologically holomorphic map of finite type, where X=S2X=S^{2} or X=S2{e}X=S^{2}-\{e\}, eS2e\in S^{2}, and AS2A\subset S^{2} be a finite set containing SfS_{f}. Suppose that γ0\gamma_{0} is a simple closed curve in S2AS^{2}-A, and let γ~0f1(γ)\widetilde{\gamma}_{0}\subset f^{-1}(\gamma) be a simple closed curve with deg(f|γ~0:γ~0γ0)=d\deg(f|\widetilde{\gamma}_{0}\colon\widetilde{\gamma}_{0}\to\gamma_{0})=d. If γ1\gamma_{1} is a simple closed curve that is homotopic in S2AS^{2}-A to γ0\gamma_{0}, then there exists a simple closed curve γ~1f1(γ1)\widetilde{\gamma}_{1}\subset f^{-1}(\gamma_{1}) such that γ~0\widetilde{\gamma}_{0} and γ~1\widetilde{\gamma}_{1} are homotopic in Xf1(A)S2f1(A)¯X-f^{-1}(A)\subset S^{2}-\overline{f^{-1}(A)} and deg(f|γ~1:γ~1γ1)=d\deg(f|\widetilde{\gamma}_{1}\colon\widetilde{\gamma}_{1}\to\gamma_{1})=d.

Proof.

According to [Bus10, Theorem A.3] (see also [FM12, Sections 1.2.5 and 1.2.6]), there exists an isotopy (φt)t𝕀(\varphi_{t})_{t\in{\mathbb{I}}} rel. AA in S2S^{2} such that φ0=idX\varphi_{0}=\operatorname{id}_{X} and φ1(γ0)=γ1\varphi_{1}(\gamma_{0})=\gamma_{1}. Since φ0=idS2\varphi_{0}=~{}\operatorname{id}_{S^{2}} is orientation-preserving, then φt\varphi_{t} is also orientation-preserving for each t𝕀t\in{\mathbb{I}}. By Proposition 2.2, there exists a homeomorphism ψ1Homeo+(S2,f1(A)(S2X))\psi_{1}\in\operatorname{Homeo}^{+}(S^{2},f^{-1}(A)\cup(S^{2}-X)) such that φ1f=fψ1\varphi_{1}\circ f=f\circ\psi_{1}. Thus, we can take γ~1:=ψ1(γ~0)\widetilde{\gamma}_{1}:=\psi_{1}(\widetilde{\gamma}_{0}). Finally, Xf1(A)S2f1(A)¯X-f^{-1}(A)\subset S^{2}-\overline{f^{-1}(A)}, since any accumulation point pS2p\in S^{2} of the set f1(A)f^{-1}(A) cannot be in XX. ∎

Due to the Uniformization Theorem and item (4) of Definition 2.1, in the case when X=S2X=S^{2}, a topologically holomorphic map f:XS2f\colon X\to S^{2} can be written as f=φgψ1f=\varphi\circ g\circ\psi^{-1}, where g:^^g\colon\widehat{{\mathbb{C}}}\to\widehat{{\mathbb{C}}} is a non-constant rational map and φ,ψ:S2^\varphi,\psi\colon S^{2}\to\widehat{{\mathbb{C}}} are orientation-preserving homeomorphisms. In fact, in this case f:S2S2f\colon S^{2}\to S^{2} is simply a branched self-covering of S2S^{2}, which is always of finite type and has finite topological degree.

Similarly, in the case when X=S2{e}X=S^{2}-\{e\}, we can write ff as φgψ1\varphi\circ g\circ\psi^{-1} such that g:R^g\colon R\to\widehat{{\mathbb{C}}} is a non-constant meromorphic map, φ:S2^\varphi\colon S^{2}\to\widehat{{\mathbb{C}}} and ψ:S2{e}R\psi\colon S^{2}-\{e\}\to R are orientation-preserving homeomorphisms, where R=R={\mathbb{C}} or R=𝔻R={\mathbb{D}}. Suppose that the map ff is of finite type. Then the image of ψ\psi above does not depend on the choice of the homeomorphism φ\varphi (see [Ere04, pp. 3-4]; it essentially follows from Proposition 2.2 and some well-known facts from the theory of quasiconformal mappings). Thus, finite-type topologically holomorphic maps for which the image of ψ\psi is {\mathbb{C}} are referred to as parabolic type maps, while those for which the image of ψ\psi is 𝔻{\mathbb{D}} are called hyperbolic type maps.

Further, we assume that every topologically holomorphic map f:S2S2f\colon S^{2}\dashrightarrow S^{2} we consider either has no essential singularities or is a finite-type topologically holomorphic map of parabolic type. Definition 2.1 and Great Picard’s Theorem imply that in any neighborhood of an essential singularity ee of such a map ff, every value is attained infinitely often with at most two exceptions. In particular, ff can have at most two omitted values, i.e., points pp in S2S^{2} such that the preimage f1(p)f^{-1}(p) is empty. Furthermore, each omitted value is an asymptotic value of ff. Additionally, observe that if AS2A\subset S^{2} is a finite set with |A|3|A|\geq 3 and SfAS_{f}\subset A, then the restriction

f|S2f1(A)¯:S2f1(A)¯S2Af|S^{2}-\overline{f^{-1}(A)}\colon S^{2}-\overline{f^{-1}(A)}\to S^{2}-A

is a covering map. Note that the closure f1(A)¯\overline{f^{-1}(A)} equals f1(A)f^{-1}(A) if f:S2S2f\colon S^{2}\to S^{2} has no essential singularity. Otherwise, f1(A)¯\overline{f^{-1}(A)} consists of f1(A)f^{-1}(A) and the essential singularity eS2e\in S^{2} of ff due to Great Picard’s Theorem and the assumption |A|3|A|\geq 3.

We say that a topologically holomorphic map f:S2S2f\colon S^{2}\dashrightarrow S^{2} is transcendental if it has an essential singularity. Given our previous assumptions on the map ff, this is equivalent to saying that ff has infinite topological degree. The map f:S2S2f\colon S^{2}\dashrightarrow S^{2} is called entire if either ff has finite topological degree and there exists a point pS2p\in S^{2} such that f1(p)={p}f^{-1}(p)=\{p\} (in which case ff is called a topological polynomial), or ff has infinite topological degree and f1(e)=f^{-1}(e)=\emptyset, where ee is the essential singularity of ff. We can view entire topologically holomorphic maps as topologically holomorphic self-maps of 2{\mathbb{R}}^{2}.

2.3. Thurston maps

Let f:S2S2f\colon S^{2}\dashrightarrow S^{2} be a topologically holomorphic map. Then the postsingular set PfP_{f} of the map ff is defined as

Pf:={qS2:q=fn(p) for some n0 and pSf}.P_{f}:=\{q\in S^{2}:q=f^{\circ n}(p)\text{ for some }n\geq 0\text{ and }p\in S_{f}\}.

In other words, the postsingular set PfP_{f} is the union of all forward orbits of the singular values of ff. It is worth noting that some of these orbits might terminate after several iterations if a singular value reaches the essential singularity of the map ff.

We say that f:S2S2f\colon S^{2}\dashrightarrow S^{2} is postsingularly finite (psf in short) if the set PfP_{f} is finite, i.e., ff has finitely many singular values and each of them eventually becomes periodic or lands on the essential singularity of ff under the iteration. Postsingularly finite topologically holomorphic maps of finite degree are also called postcritically finite (pcf in short), and their postsingular values are called postcritical, as their singular values are always critical.

Now we are ready to state one of the key definitions of this section.

Definition 2.4.

A non-injective topologically holomorphic map f:S2S2f\colon S^{2}\dashrightarrow S^{2} is called a Thurston map if it is postsingularly finite and either ff has no essential singularity or ff is a parabolic type map.

Given a finite set AS2A\subset S^{2} such that PfAP_{f}\subset A and every aAa\in A is either the essential singularity of ff or f(a)Af(a)\in A, we call the pair (f,A)(f,A) a marked Thurston map and AA its marked set.

We often consider marked Thurston maps in the same way as usual Thurston maps and use the notation f:(S2,A)f\colon(S^{2},A)\righttoleftarrow while still assuming that ff might not be defined on the entire sphere S2S^{2}. If no specific marked set is mentioned, we assume it to be PfP_{f}. Note that when the marked set AA contains the essential singularity of the map ff, the set AA is not forward invariant with respect to ff in the usual sense. However, if |A|3|A|\geq 3, then Af1(A)¯A\subset\overline{f^{-1}(A)}.

The dynamics of a Thurston map on its marked set or some other finite subsets of S2S^{2} can also be represented graphically, in a way that turns out to be useful in study. Suppose that f:S2S2f\colon S^{2}\dashrightarrow S^{2} is a Thurston map and AS2A\subset S^{2} is a finite set such that every aAa\in A is either the essential singularity of ff or f(a)Af(a)\in A. Then the dynamical portrait of the map ff on the set AA is a directed abstract graph with the vertex set AA, where for each vertex vAv\in A that is not the essential singularity of ff, there is a unique directed edge from vv to f(v)f(v), and if vAv\in A is the essential singularity of ff, there are no outgoing edges from vv. If the set AA coincides with the postsingular set PfP_{f}, the dynamical portrait of ff on the set AA is called the postsingular portrait of the Thurston map ff.

Definition 2.5.

Two Thurston maps f1:(S2,A)f_{1}\colon(S^{2},A)\righttoleftarrow and f2:(S2,A)f_{2}\colon(S^{2},A)\righttoleftarrow are called isotopic (rel. AA) if there exists ϕHomeo0+(S2,A)\phi\in\operatorname{Homeo}_{0}^{+}(S^{2},A) such that f1=f2ϕf_{1}=f_{2}\circ\phi.

Remark 2.6.

Let f1:(S2,A)f_{1}\colon(S^{2},A)\righttoleftarrow and f2:(S2,A)f_{2}\colon(S^{2},A)\righttoleftarrow be two Thurston maps satisfying the relation ϕ1f1=f2ϕ2\phi_{1}\circ f_{1}=f_{2}\circ\phi_{2} for some ϕ1,ϕ2Homeo0+(S2,A)\phi_{1},\phi_{2}\in\operatorname{Homeo}_{0}^{+}(S^{2},A). Then it follows from Proposition 2.2 that f1f_{1} and f2f_{2} are isotopic rel. AA.

The notion of isotopy for Thurston maps depends on their common marked set. Consequently, we sometimes refer to isotopy relative AA (or rel. AA for short) to specify which marked set is being considered. This applies to other notions introduced below that also depend on the choice of the marked set.

We say that two (marked) Thurston maps are combinatorially equivalent if they are “topologically conjugate up to isotopy”:

Definition 2.7.

Two Thurston maps f1:(S2,A1)f_{1}\colon(S^{2},A_{1})\righttoleftarrow and f2:(S2,A2)f_{2}\colon(S^{2},A_{2})\righttoleftarrow are called combinatorially (or Thurston) equivalent if there exist two Thurston maps f~1:(S2,A1)\widetilde{f}_{1}\colon(S^{2},A_{1})\righttoleftarrow and f~2:(S2,A2)\widetilde{f}_{2}\colon(S^{2},A_{2})\righttoleftarrow such that:

  • fif_{i} and f~i\widetilde{f}_{i} are isotopic rel. AiA_{i} for each i=1,2i=1,2, and

  • f~1\widetilde{f}_{1} and f~2\widetilde{f}_{2} are conjugate via a homeomorphsim ϕHomeo+(S2)\phi\in\operatorname{Homeo}^{+}(S^{2}), i.e., ϕf~1=f~2ϕ\phi\circ\widetilde{f}_{1}=\widetilde{f}_{2}\circ\phi, such that ϕ(A1)=A2\phi(A_{1})=A_{2}.

Remark 2.8.

Definition 2.7 can be reformulated in a more classical way. Thurston maps f1:(S2,A1)f_{1}\colon(S^{2},A_{1})\righttoleftarrow and f2:(S2,A2)f_{2}\colon(S^{2},A_{2})\righttoleftarrow are combinatorially equivalent if and only if there exist two homeomorphisms ϕ1,ϕ2Homeo+(S2)\phi_{1},\phi_{2}\in\operatorname{Homeo}^{+}(S^{2}) such that ϕ1(A1)=ϕ2(A1)=A2\phi_{1}(A_{1})=\phi_{2}(A_{1})=A_{2}, ϕ1\phi_{1} and ϕ2\phi_{2} are isotopic rel. AA, and ϕ1f1=f2ϕ2\phi_{1}\circ f_{1}=f_{2}\circ\phi_{2}.

Remark 2.9.

If A1=Pf1A_{1}=P_{f_{1}} and A2=Pf2A_{2}=P_{f_{2}}, then the condition that ϕ(A1)=A2\phi(A_{1})=A_{2} in Definition 2.7 and the condition ϕ1(A1)=ϕ2(A1)=A2\phi_{1}(A_{1})=\phi_{2}(A_{1})=A_{2} in Remark 2.8 can be removed since they are automatically satisfied if all other conditions hold.

A Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is said to be realized if it is combinatorially equivalent to a postsingularly finite holomorphic map g:(^,P)g\colon(\widehat{{\mathbb{C}}},P)\righttoleftarrow. If f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is not realized, we say that it is obstructed.

Let AS2A\subset S^{2} be a finite set. We say that a simple closed curve γS2A\gamma\subset S^{2}-A is essential in S2AS^{2}-A if each connected component of S2γS^{2}-\gamma contains at least two points of the set AA. In other words, γ\gamma is essential in S2AS^{2}-A if it cannot be shrinked to a point via a homotopy in S2AS^{2}-A.

Definition 2.10.

Let f:(S2,A)f\colon(S^{2},A)\righttoleftarrow be a Thurston map. We say that a simple closed curve γ\gamma forms a Levy cycle for f:(S2,A)f\colon(S^{2},A)\righttoleftarrow if γ\gamma is essential in S2AS^{2}-A and there exists another simple closed curve γ~fn(γ)\widetilde{\gamma}\subset f^{-n}(\gamma) for some n1n\geq 1 such that γ\gamma and γ~\widetilde{\gamma} are homotopic in S2AS^{2}-A and deg(fn|γ~:γ~γ)=1\deg(f^{\circ n}|\widetilde{\gamma}\colon\widetilde{\gamma}\to\gamma)=1.

If n=1n=1 in Definition 2.10, then γ\gamma is called a Levy fixed (or Levy invariant) curve. Levy fixed curve γ\gamma is called weakly degenerate if ff is injective on one of the connected components of S2γ~S^{2}-\widetilde{\gamma}. If additionally the image of this connected component UU under ff contains the same points of the set AA as UU, i.e., UA=f(U)AU\cap A=f(U)\cap A, we say that γ\gamma is a degenerate Levy fixed curve for the Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow.

The following observation is widely known in the context of finite degree Thurston maps [Hub16, Exercise 10.3.6], and its proof extends to the case of transcendental Thurston maps as well (see Section 3.1 for the proof).

Proposition 2.11.

Let f:(S2,A)f\colon(S^{2},A)\righttoleftarrow be a Thurston map. If there exists a simple closed curve γS2\gamma\subset S^{2} forming a Levy cycle for f:(S2,A)f\colon(S^{2},A)\righttoleftarrow, then ff is obstructed rel. AA.

We require one more notion of equivalence between Thurston maps.

Definition 2.12.

We say that two Thurston maps f1:(S2,A)f_{1}\colon(S^{2},A)\righttoleftarrow and f2:(S2,A)f_{2}\colon(S^{2},A)\righttoleftarrow are pure Hurwitz equivalent (or simply Hurwitz equivalent) if there exist two homeomorphisms ϕ1,ϕ2Homeo+(S2,A)\phi_{1},\phi_{2}\in\operatorname{Homeo}^{+}(S^{2},A) such that ϕ1f1=f2ϕ2\phi_{1}\circ f_{1}=f_{2}\circ\phi_{2}.

If f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is a Thurston map, then the Hurwitz class f,A{\mathcal{H}}_{f,A} of ff is the union of all Thurston maps with the marked set AA that are Hurwitz equivalent to ff. If AA coincides with the postsingular set PfP_{f} of ff, we simply use the notation f{\mathcal{H}}_{f}. We say that a Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is totally unobstructed if every Thurston map in f,A{\mathcal{H}}_{f,A} is unobstructed.

Remark 2.13.

According to [FM12, Proposition 2.3], if two orientation-preserving homomeomorphisms φ:S2^\varphi\colon S^{2}\to\widehat{{\mathbb{C}}} and ψ:S2^\psi\colon S^{2}\to\widehat{{\mathbb{C}}} agree on the set AS2A\subset S^{2} with |A|3|A|\leq 3, they are isotopic rel. AA. This observation can be used to show that any Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is realized when the marked set AA contains three or fewer points. Similarly, in thise case, one can show that f,A{\mathcal{H}}_{f,A} consists of a single Thurston map up to a combinatorial equivalence rel. AA. However, when |A|=4|A|=4, the question of realizability already becomes significantly more challenging.

2.4. Teichmüller and moduli spaces

Let AS2A\subset S^{2} be a finite set containing at least three points. Then the Teichmüller space of the sphere S2S^{2} with the marked set AA is defined as

𝒯A:={φ:S2^ is an orientation-preserving homeomorphism}/\mathcal{T}_{A}:=\{\varphi\colon S^{2}\rightarrow\widehat{{\mathbb{C}}}\text{ is an orientation-preserving homeomorphism}\}/\sim

where φ1φ2\varphi_{1}\sim\varphi_{2} if there exists a Möbius transformation MM such that φ1\varphi_{1} is isotopic rel. AA to Mφ2M\circ\varphi_{2}.

Similarly, we define the moduli space of the sphere S2S^{2} with the marked set AA:

A:={η:A^ is injective}/,\mathcal{M}_{A}:=\{\eta\colon A\rightarrow\widehat{{\mathbb{C}}}\text{ is injective}\}/\sim,

where η1η2\eta_{1}\sim\eta_{2} if there exists a Möbius transformation MM such that η1=Mη2\eta_{1}=M\circ\eta_{2}.

Further, [][\cdot] denotes an equivalence class corresponding to a point of either the Teichmüller space 𝒯A\mathcal{T}_{A} or the moduli space A\mathcal{M}_{A}. Note that there is an obvious map π:𝒯AA\pi\colon\mathcal{T}_{A}\to\mathcal{M}_{A} defined as π([φ])=[φ|A]\pi([\varphi])=[\varphi|A]. According to [FM12, Proposition 2.3], when |A|=3|A|=3, both the Teichmüller space 𝒯A\mathcal{T}_{A} and the moduli space A\mathcal{M}_{A} are just single points. Therefore, for the rest of this section, we assume that |A|4|A|\geq 4.

It is known that the Teichmüller space 𝒯A\mathcal{T}_{A} admits a complete metric dTd_{T}, known as the Teichmüller metric [Hub06, Proposition 6.4.4]. Moreover, with respect to the topology induced by this metric, 𝒯A\mathcal{T}_{A} is a contractible space [Hub06, Corollary 6.7.2]. At the same time, both 𝒯A\mathcal{T}_{A} and A\mathcal{M}_{A} admit structures of (|A|3)(|A|-3)-complex manifolds (see [Hub06, Theorem 6.5.1]) so that the map π:𝒯AA\pi\colon\mathcal{T}_{A}\to\mathcal{M}_{A} becomes a holomorphic universal covering map [Hub16, Section 10.9].

Moreover, the complex structure of A\mathcal{M}_{A} is quite explicit in the general case. Let A={a1,a2,,ak,ak+1,ak+2,ak+3}A=\{a_{1},a_{2},\dots,a_{k},a_{k+1},a_{k+2},a_{k+3}\}, k1k\geq 1, where the indexing of the points of AA is chosen arbitrarily. Define the map h:Akkh\colon\mathcal{M}_{A}\to{\mathbb{C}}^{k}-\mathcal{L}_{k} by

h([φ])=(φ(a1),φ(a2),,φ(ak)),h([\varphi])=(\varphi(a_{1}),\varphi(a_{2}),\dots,\varphi(a_{k})),

where the representative φ:S2^\varphi\colon S^{2}\to\widehat{{\mathbb{C}}} is chosen so that φ(ak+1)=0,φ(ak+2)=1\varphi(a_{k+1})=0,\varphi(a_{k+2})=1, and φ(ak+3)=\varphi(a_{k+3})=\infty, and where k\mathcal{L}_{k} is the subset of k{\mathbb{C}}^{k} defined by

k:={(z1,z2,,zk)k:zi=zj for some ij, or zi=0, or zi=1}.\mathcal{L}_{k}:=\{(z_{1},z_{2},\dots,z_{k})\in{\mathbb{C}}^{k}:z_{i}=z_{j}\text{ for some $i\neq j$},\text{ or }z_{i}=0,\text{ or }z_{i}=1\}.

It is known that the map hh provides a biholomorphism between A\mathcal{M}_{A} and kk{\mathbb{C}}^{k}-\mathcal{L}_{k} (see [Hub16, Section 10.9]).

Our focus in this paper is on the case when |A|=4|A|=4. In this situation, the Teichmüller space 𝒯A\mathcal{T}_{A} is biholomorphic to 𝔻{\mathbb{D}}, with the metric dTd_{T} coinciding with the usual hyperbolic metric on 𝔻{\mathbb{D}}; see [Hub06, Corollary 6.10.3 and Theorem 6.10.6]. Furthermore, the moduli space A\mathcal{M}_{A} is biholomorphic to the three-punctured Riemann sphere Σ\Sigma.

2.5. Pullback maps

In this section, we illustrate how the notions introduced in Section 2.4 can be applied for studying the properties of Thurston maps. Most importantly, using Definition 2.1 and Proposition 2.2, we can introduce the following crucial concept (see [MPR24, Proposition 2.21] for the proof; note that this is where the parabolic type condition in the Definition 2.4 plays a crucial role).

Proposition 2.14.

Suppose that f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is a Thurston map, or fHomeo+(S2)f\in\operatorname{Homeo}^{+}(S^{2}) and f(A)=Af(A)=A, where 3|A|<3\leq|A|<\infty. Let φ:S2^\varphi\colon S^{2}\to\widehat{{\mathbb{C}}} be an orientation-preserving homeomorphism. Then there exists an orientation-preserving homeomorphism ψ:S2^\psi\colon S^{2}\to\widehat{{\mathbb{C}}} such that gφ:=φfψ1:^^g_{\varphi}:=\varphi\circ f\circ\psi^{-1}\colon\widehat{{\mathbb{C}}}\dashrightarrow\widehat{{\mathbb{C}}} is holomorphic. In other words, the following diagram commutes

(S2,A){(S^{2},A)}(^,ψ(A)){(\widehat{{\mathbb{C}}},\psi(A))}(S2,A){(S^{2},A)}(^,φ(A)){(\widehat{{\mathbb{C}}},\varphi(A))}ψ\scriptstyle{\psi}f\scriptstyle{f}gφ\scriptstyle{g_{\varphi}}φ\scriptstyle{\varphi}

The homeomorphism ψ\psi is unique up to post-composition with a Möbius transformation. Different choices of φ\varphi that represent the same point in 𝒯A\mathcal{T}_{A} yield maps ψ\psi that represent the same point in 𝒯A\mathcal{T}_{A}.

In other words, we have a well-defined map σf:𝒯A𝒯A\sigma_{f}\colon\mathcal{T}_{A}\to\mathcal{T}_{A} such that σf([φ])=[ψ]\sigma_{f}([\varphi])=[\psi], called the pullback map (or the σ\sigma-map) associated with the Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow. As φ\varphi ranges across all maps representing a single point in 𝒯A\mathcal{T}_{A}, the map gφg_{\varphi} is uniquely defined up to pre- and post-composition with Möbius transformations.

Remark 2.15.

Let ϕ:S2S2\phi\colon S^{2}\to S^{2} be an orientation-preserving homeomorphism with ϕ(A)=A\phi(A)=A and |A|3|A|\geq 3. It is straightforward to verify that if τ=[φ]𝒯A\tau=[\varphi]\in\mathcal{T}_{A}, then σϕ([φ])=[φϕ]\sigma_{\phi}([\varphi])=[\varphi\circ\phi]. Moreover, if f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is a Thurston map, it is easy to see that σϕf=σfσϕ\sigma_{\phi\circ f}=\sigma_{f}\circ\sigma_{\phi} and σfϕ=σϕσf\sigma_{f\circ\phi}=\sigma_{\phi}\circ\sigma_{f}.

Proposition 2.16.

Suppose that we are in the setting of Proposition 2.14. If there exists a subset BAB\subset A, such that SfBS_{f}\subset B and |B|=3|B|=3, then gφg_{\varphi}, up to pre-composition with a Möbius transformation, depends only on φ|B\varphi|B. Furthermore, if there exists a subset CAC\subset A such that |C|=3|C|=3 and Cf1(B)¯C\subset\overline{f^{-1}(B)}, then gφg_{\varphi} is uniquely determined by φ|B\varphi|B and ψ|C\psi|C, and if Af1(B)¯A\subset\overline{f^{-1}(B)}, then σf\sigma_{f} is a constant map.

Proof.

Suppose that φ1,φ2,ψ1,ψ2:S2^\varphi_{1},\varphi_{2},\psi_{1},\psi_{2}\colon S^{2}\to\widehat{{\mathbb{C}}} are orientation-preserving homeomorphisms such that φ1|B=φ2|B\varphi_{1}|B=\varphi_{2}|B, and the maps gφ1=φ1fψ11g_{\varphi_{1}}=\varphi_{1}\circ f\circ\psi_{1}^{-1} and gφ2=φ2fψ21g_{\varphi_{2}}=\varphi_{2}\circ f\circ\psi_{2}^{-1} are holomorphic possibly outside of single points in ^\widehat{{\mathbb{C}}}. One can easily see that we have the following:

gφ1=(φ1φ21)gφ2(ψ2ψ11),g_{\varphi_{1}}=(\varphi_{1}\circ\varphi_{2}^{-1})\circ g_{\varphi_{2}}\circ(\psi_{2}\circ\psi_{1}^{-1}),

where the homeomorphism φ:=φ1φ21\varphi:=\varphi_{1}\circ\varphi_{2}^{-1} fixes each point of the set φ1(B)=φ2(B)\varphi_{1}(B)=\varphi_{2}(B). Since |B|=3|B|=3, [FM12, Proposition 2.3] implies that φ\varphi is isotopic rel. φ1(B)\varphi_{1}(B) to id^\operatorname{id}_{\widehat{{\mathbb{C}}}}. According to Proposition 2.2, this isotopy can be lifted, leading to the relation gφ1=gφ2ψg_{\varphi_{1}}=g_{\varphi_{2}}\circ\psi, where ψHomeo+(S2)\psi\in\operatorname{Homeo}^{+}(S^{2}). It is easy to see that the homeomorphism ψ\psi is a Möbius transformation.

Now, suppose that there exists a subset CAC\subset A such that |C|=3|C|=3 and Cf1(B)¯C\subset\overline{f^{-1}(B)}, and ψ1|C=ψ2|C\psi_{1}|C=\psi_{2}|C. Note that the homeomorphism ψ\psi is isotopic rel. gφ11(φ1(B))¯=ψ1(f1(B)¯)\overline{g_{\varphi_{1}}^{-1}(\varphi_{1}(B))}=\psi_{1}(\overline{f^{-1}(B)}) to ψ2ψ11\psi_{2}\circ\psi_{1}^{-1} due to Proposition 2.2. Consequently, ψ|ψ1(C)=(ψ2ψ11)|ψ1(C)=idψ1(C)\psi|\psi_{1}(C)=(\psi_{2}\circ\psi_{1}^{-1})|\psi_{1}(C)=\operatorname{id}_{\psi_{1}(C)}. Since ψ\psi is a Möbius transformation fixing three distinct points in ^\widehat{{\mathbb{C}}}, it must be the identity id^\operatorname{id}_{\widehat{{\mathbb{C}}}}. Thus, the maps gφ1g_{\varphi_{1}} and gφ2g_{\varphi_{2}} coincide.

If Af1(B)¯A\subset\overline{f^{-1}(B)}, then ψ2ψ11\psi_{2}\circ\psi_{1}^{-1} is isotopic rel. ψ1(A)\psi_{1}(A) to the Möbius transformation ψ=id^\psi=~{}\operatorname{id}_{\widehat{{\mathbb{C}}}}. Thus, σf([φ1])=[ψ1]=[ψ2]=σf([φ2])\sigma_{f}([\varphi_{1}])=[\psi_{1}]=[\psi_{2}]=\sigma_{f}([\varphi_{2}]) in the Teichmüller space 𝒯A\mathcal{T}_{A}, and the rest follows. ∎

The following observation provides the most crucial property of pullback maps (see [MPR24, Proposition 2.24] for the proof).

Proposition 2.17.

A Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow with |A|3|A|\geq 3 is realized if and only if the pullback map σf\sigma_{f} has a fixed point in the Teichmüller space 𝒯A\mathcal{T}_{A}.

To illustrate the principle formulated in Proposition 2.17, we present the following remark.

Remark 2.18.

Let f:(S2,A)f\colon(S^{2},A)\righttoleftarrow be a Thurston map with |A|3|A|\geq 3, and suppose that there is a subset BAB\subset A such that SfBS_{f}\subset B, |B|=3|B|=3, and Af1(B)¯A\subset\overline{f^{-1}(B)}. Then ff is realized rel. AA because σf\sigma_{f} is a constant map according to Proposition 2.16. Additionally, by applying Remark 2.15 and Proposition 2.16, it is easy to show that the Hurwitz class f,A{\mathcal{H}}_{f,A} consists of a single Thurston map, up to combinatorial equivalence rel. AA.

However, Thurston maps that satisfy these conditions are somewhat artificial. For such a marked Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow, it must hold PfBP_{f}\subset B and for every aABa\in A-B, either aa is the essential singularity of ff or f(a)Bf(a)\in B. For instance, this scenario is impossible for unmarked Thurston maps with at least four postsingular values.

Proposition 2.19.

Let f:(S2,A)f\colon(S^{2},A)\righttoleftarrow be a Thurston map with |A|3|A|\geq 3. Then the pullback map σf\sigma_{f} is holomorphic.

Proof.

This result is rather well-known in the context of finite degree Thurston maps (see [BCT14, Section 1.3] and [Hub16, Sections 10.6 and 10.7]), and it can be extended analogously to the transcendental setting (see, for instance, [Ast, Lemma 3.3]). ∎

Remark 2.20.

Proposition 2.19 and [Hub06, Corollary 6.10.7] imply that the map σf\sigma_{f} is 11-Lipschitz, meaning dT(σf(τ1),σf(τ2))dT(τ1,τ2)d_{T}(\sigma_{f}(\tau_{1}),\sigma_{f}(\tau_{2}))\leq d_{T}(\tau_{1},\tau_{2}) for every τ1,τ2𝒯A\tau_{1},\tau_{2}\in\mathcal{T}_{A}. In fact, in many cases, such as when the Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is transcendental, it can be shown (see [HSS09, Section 3.2], [Pfr19, Chapter 5.1], or [Ast, Sections 2.3 and 3.1]) that σf\sigma_{f} is actually distance-decreasing, i.e., dT(σf(τ1),σf(τ2))<dT(τ1,τ2)d_{T}(\sigma_{f}(\tau_{1}),\sigma_{f}(\tau_{2}))<d_{T}(\tau_{1},\tau_{2}) for every distinct τ1,τ2𝒯A\tau_{1},\tau_{2}\in\mathcal{T}_{A}. This property of pullback maps can be used to obtain certain rigidity results for transcendental postsingularly finite meromorphic maps (cf. [Hub16, Corollary 10.7.8] and [MPR24, Proposition 2.26]).

However, we do not require these results and the observation of Proposition 2.19 for our further arguments since we mostly work with families of Thurston maps satisfying additional assumptions. For these families, we will directly observe all the properties mentioned above.

3. Hyperbolic tools

Let UU be a hyperbolic Riemann surface, and let dUd_{U} denote the distance function of the hyperbolic metric on UU. For any rectifiable curve α\alpha in UU, we denote the length of α\alpha with respect to the hyperbolic metric by U(α)\ell_{U}(\alpha). When we refer to γ\gamma as a geodesic in UU, we always mean that γ\gamma is a geodesic with respect to the hyperbolic metric. Also, let BU(z,r)B_{U}(z,r) be the hyperbolic ball in UU with center zUz\in U and radius rr. If UU is a subset of {\mathbb{C}}, then the hyperbolic metric on UU, as a conformal metric, can be written as ρU(z)|dz|\rho_{U}(z)|dz|, where ρU:U[0,+)\rho_{U}\colon U\to[0,+\infty) is the density of the hyperbolic metric on UU.

For a holomorphic map g:UVg\colon U\to V between two hyperbolic Riemann surfaces, we denote by Dg(z)UV\|\mathrm{D}g(z)\|_{U}^{V} the norm of the derivative of gg with respect to the hyperbolic metrics on the domain UU and the range VV. More precisely, this norm is given by Dg(z)UV=Dzg(v)V/vU\|\mathrm{D}g(z)\|_{U}^{V}=\|\mathrm{D}_{z}g(v)\|_{V}/\|v\|_{U}, where vTzUv\in T_{z}U is any non-zero vector, and U\|\cdot\|_{U} represents the length of a tangent vector to UU with respect to the hyperbolic metric. If U=VU=V, we simply use the notation Dg(z)U\|\mathrm{D}g(z)\|_{U}.

Schwarz-Pick’s lemma [Hub06, Proposition 3.3.4] implies that for a holomorphic map g:UVg\colon U\to V between two hyperbolic Riemann surfaces, we have Dg(z)UV1\|\mathrm{D}g(z)\|_{U}^{V}\leq 1 for every zUz\in U. Furthermore, if gg is a covering map, this inequality becomes an equality; otherwise, gg is locally uniformly contracting, i.e., for every compact set KUK\subset U, there exists a constant λK<1\lambda_{K}<1 such that Dg(z)UVλK\|\mathrm{D}g(z)\|_{U}^{V}\leq\lambda_{K} for all zKz\in K. Suppose that α\alpha is a C1C^{1}-curve in UU and Dg(z)UVλ\|\mathrm{D}g(z)\|_{U}^{V}\leq\lambda for all zαz\in\alpha. Then it is straightforward to check that V(g(α))λU(α)\ell_{V}(g(\alpha))\leq\lambda\ell_{U}(\alpha). In particular, Schwarz-Pick’s lemma implies that the map g:UVg\colon U\to V is always 1-Lipschitz and, if gg is not a covering map, then gg is locally uniformly distance-decreasing.

We will be particularly interested in the case when the map gg mentioned above is simply the inclusion map I:UVI\colon U\hookrightarrow V. We are going to denote DI(z)UV\|\mathrm{D}I(z)\|_{U}^{V} by cUV(z)c_{U}^{V}(z). Clearly, if UVU\subset V\subset{\mathbb{C}}, then cUV(z)=ρV(z)/ρU(z)c_{U}^{V}(z)=\rho_{V}(z)/\rho_{U}(z).

3.1. Levy cycles

Let δ\delta be an essential simple closed curve in the punctured Riemann sphere X:=^PX:=\widehat{{\mathbb{C}}}-P, where 3|P|<3\leq|P|<\infty. According to [Hub06, Proposition 3.3.8], there exists a unique closed geodesic in XX that is homotopic (in XX) to δ\delta. Note that this geodesic should be necessarily simple [Hub06, Proposition 3.3.9].

In order to illustrate the utility of the hyperbolic tools, we prove that Levy cycles are obstructions for Thurston maps of both finite and infinite degree.

Proof of Proposition 2.11.

Let γ\gamma be a simple closed curve forming a Levy cycle for the Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow. We assume that n=1n=1 in Definition 2.10, i.e., γ\gamma is a Levy fixed curve. The general case can be handled in a similar manner.

Suppose that f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is combinatorially equivalent to a postsingularly finite holomorphic map g:(^,P)g\colon(\widehat{{\mathbb{C}}},P)\righttoleftarrow. Based on Definition 2.7 and Remark 2.8, it is easy to see that g:(^,P)g\colon(\widehat{{\mathbb{C}}},P)\righttoleftarrow also has a Levy fixed curve δ^P\delta\subset\widehat{{\mathbb{C}}}-P. Note that |P|4|P|\geq 4 since, otherwise, δ\delta would not be essential in ^P\widehat{{\mathbb{C}}}-P. According to the previous discussion and Corollary 2.3, we can assume that δ\delta is a simple closed geodesic in ^P\widehat{{\mathbb{C}}}-P.

Let δ~\widetilde{\delta} be a connected component of g1(δ)g^{-1}(\delta) such that δ\delta and δ~\widetilde{\delta} are homotopic in ^P\widehat{{\mathbb{C}}}-P, and deg(g|δ~:δ~δ)=1\deg(g|\widetilde{\delta}\colon\widetilde{\delta}\to\delta)=1. By Schwarz-Pick’s lemma, it follows that

^P(δ)=^g1(P)¯(δ~)>^P(δ~),\ell_{\widehat{{\mathbb{C}}}-P}(\delta)=\ell_{\widehat{{\mathbb{C}}}-\overline{g^{-1}(P)}}(\widetilde{\delta})>\ell_{\widehat{{\mathbb{C}}}-P}(\widetilde{\delta}),

where the last inequality is strict since g1(P)¯P\overline{g^{-1}(P)}-P is non-empty because of Great Picard’s theorem and the Riemann-Hurwitz formula [Hub06, Appendix A.3]. However, δ\delta is the unique geodesic in its homotopy class in ^P\widehat{{\mathbb{C}}}-P. Thus, ^P(δ)^P(δ~)\ell_{\widehat{{\mathbb{C}}}-P}(\delta)\leq\ell_{\widehat{{\mathbb{C}}}-P}(\widetilde{\delta}), and it leads to a contradiction. ∎

Let γ\gamma be an essential simple closed curve in S2AS^{2}-A, and τ=[φ]\tau=[\varphi] be a point in the Teichmüller space 𝒯A\mathcal{T}_{A}. We define lγ(τ)l_{\gamma}(\tau) as the length of the unique hyperbolic geodesic in ^φ(A)\widehat{{\mathbb{C}}}-\varphi(A) that is homotopic in ^φ(A)\widehat{{\mathbb{C}}}-\varphi(A) to φ(α)\varphi(\alpha). Additionally, we introduce wγ(τ):=loglγ(τ)w_{\gamma}(\tau):=\log l_{\gamma}(\tau). It is known that wγ:𝒯Aw_{\gamma}\colon\mathcal{T}_{A}\to{\mathbb{R}} is a 11-Lipschitz function [Hub06, Theorem 7.6.4].

Let XX be a hyperbolic Riemann surface, and suppose αX\alpha\subset X is a simple closed geodesic with X(α)<\ell_{X}(\alpha)<\ell^{*}, where :=log(3+22)\ell^{*}:=\log(3+2\sqrt{2}). In this case, we say that α\alpha is short. As stated in [Hub06, Proposition 3.3.8 and Corollary 3.8.7], two short simple closed geodesics on a hyperbolic Riemann surface XX are either disjoint and non-homotopic in XX, or they coincide. Therefore, [Hub06, Proposition 3.3.8] implies that a punctured Riemann sphere ^P\widehat{{\mathbb{C}}}-P, where 3|P|<3\leq|P|<\infty, can have at most |P|3|P|-3 distinct short simple closed geodesics.

The following result allows us to identify a Levy fixed curve for a Thurston map based on the behavior of the corresponding pullback map.

Proposition 3.1.

Let f:(S2,A)f\colon(S^{2},A)\righttoleftarrow be a Thurston map with |A|=4|A|=4, and τ=[φ]\tau=[\varphi] and σf(τ)=[ψ]\sigma_{f}(\tau)=[\psi] be points in the Teichmüller space 𝒯A\mathcal{T}_{A}, where the representatives φ,ψ:S2^\varphi,\psi\colon S^{2}\to\widehat{{\mathbb{C}}} are chosen so that the map g:=φfψ1:^^g:=\varphi\circ f\circ\psi^{-1}\colon\widehat{{\mathbb{C}}}\dashrightarrow\widehat{{\mathbb{C}}} is holomorphic. Suppose that there exists an annulus U^U\subset\widehat{{\mathbb{C}}} such that:

  • each connected component of ^U\widehat{{\mathbb{C}}}-U contains two points of ψ(A)\psi(A);

  • mod(U)>5πed0/\mathrm{mod}(U)>5\pi e^{d_{0}}/\ell^{*}, where d0=dT(τ,σf(τ))d_{0}=d_{T}(\tau,\sigma_{f}(\tau));

  • gg is defined and injective on UU.

Then f:(S2,A)f\colon(S^{2},A)\righttoleftarrow has a Levy fixed curve. Moreover, if gg is defined and injective outside a single connected component of ^U\widehat{{\mathbb{C}}}-U, then f:(S2,A)f\colon(S^{2},A)\righttoleftarrow has a weakly degenerate Levy fixed curve.

Proof.

Since gg is defined and injective on UU, the annulus UU contains at most 4 points of the set g1(φ(A))¯\overline{g^{-1}(\varphi(A))}. Therefore, we can find a parallel subannulus VV of UU such that mod(V)mod(U)/5>πed0\mathrm{mod}(V)\geq\mathrm{mod}(U)/5>\pi e^{d_{0}}\ell^{*}, and VV does not contain any points of g1(φ(A))¯\overline{g^{-1}(\varphi(A))}. Denote X:=^ψ(A)X:=\widehat{{\mathbb{C}}}-\psi(A), Y:=^φ(A)Y:=\widehat{{\mathbb{C}}}-\varphi(A), and Z:=^g1(φ(A))¯Z:=\widehat{{\mathbb{C}}}-\overline{g^{-1}(\varphi(A))}. In particular, g|Z:ZXg|Z\colon Z\to X is a holomorphic covering map.

Let α\alpha be a unique hyperbolic geodesic of VV. It is known that α\alpha is a simple closed curve that forms a core curve of the annulus VV, and its length in VV is given by V(α)=π/mod(V)\ell_{V}(\alpha)=\pi/\mathrm{mod}(V); see [Hub06, Proposition 3.3.7]. Let β\beta denote the curve g(α)g(\alpha). Since α\alpha is a simple closed curve and g|Vg|V is injective, then β\beta is also a simple closed curve. At the same time, β\beta must be essential in YY since, otherwise, α\alpha would not be essential in XX (see, for instance, [For91, Theorems 5.10 and 5.11]). Define α~:=ψ1(α)\widetilde{\alpha}:=\psi^{-1}(\alpha) and β~:=φ1(β)\widetilde{\beta}:=\varphi^{-1}(\beta), both of which are essential simple closed curves in S2AS^{2}-A. It is straightforward to verify that f(α~)=β~f(\widetilde{\alpha})=\widetilde{\beta} and deg(f|α~:α~β~)=1\deg(f|\widetilde{\alpha}\colon\widetilde{\alpha}\to\widetilde{\beta})=1.

According to Schwarz-Pick’s lemma and the choice of αZ\alpha\subset Z, we have the following inequality:

X(α)Z(α)V(α)=π/mod(V)<ed0.\ell_{X}(\alpha)\leq\ell_{Z}(\alpha)\leq\ell_{V}(\alpha)=\pi/\mathrm{mod}(V)<e^{-d_{0}}\ell^{*}.

Therefore, since the function wα~:𝒯Aw_{\widetilde{\alpha}}\colon\mathcal{T}_{A}\to{\mathbb{R}} is 1-Lipschitz, it follows that lα~(τ)<l_{\widetilde{\alpha}}(\tau)<\ell^{*}. Hence, there exists a simple closed geodesic δ\delta in YY, homotopic in YY to φ(α~)\varphi(\widetilde{\alpha}), such that Y(δ)<\ell_{Y}(\delta)<~{}\ell^{*}. At the same time, Y(β)=Z(α)<ed0\ell_{Y}(\beta)=\ell_{Z}(\alpha)<e^{-d_{0}}\ell^{*}. Therefore, since both β\beta and δ\delta are essential in YY, they are homotopic in YY to short simple closed geodesics. However, as discussed previously, the four-punctured Riemann sphere YY can have only one short simple closed geodesic. This implies that β\beta and δ\delta are homotopic in YY, which in turn means that the curves α~\widetilde{\alpha} and β~\widetilde{\beta} are homotopic in S2AS^{2}-A. Hence, β~\widetilde{\beta} provides a Levy fixed curve for the Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow. It is also straightforward to verify that if gg is defined and injective outside a single connected component of ^U\widehat{{\mathbb{C}}}-U, then this Levy fixed curve is weakly degenerate. ∎

The following result guarantees the uniqueness of a Levy fixed curve for a Thurston map with four marked points, given a specific technical condition.

Proposition 3.2.

Let f:(S2,A)f\colon(S^{2},A)\righttoleftarrow be a Thurston map with |A|=4|A|=4 that has a Levy fixed curve. Suppose that there exists a point τ\tau in the Teichmüller space 𝒯A\mathcal{T}_{A} such that the sequence (π(σfn(τ)))(\pi(\sigma_{f}^{\circ n}(\tau))) eventually leaves every compact subset of the moduli space A\mathcal{M}_{A}. Then f:(S2,A)f\colon(S^{2},A)\righttoleftarrow has a unique Levy fixed curve up to homotopy in S2AS^{2}-A.

Proof.

Let γ\gamma be a Levy fixed curve for the Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow. Choose a point μ𝒯A\mu\in\mathcal{T}_{A} such that lγ(μ)<l_{\gamma}(\mu)<\ell^{*}. Define τn:=σfn(τ)=[φn]\tau_{n}:=\sigma_{f}^{\circ n}(\tau)=[\varphi_{n}] and μn:=σfn(μ)=[ψn]\mu_{n}:=\sigma_{f}^{\circ n}(\mu)=[\psi_{n}] for all n0n\geq 0. According to Proposition 2.14, we can assume that gn:=ψnfψn+11:^^g_{n}:=\psi_{n}\circ f\circ\psi_{n+1}^{-1}\colon\widehat{{\mathbb{C}}}\dashrightarrow\widehat{{\mathbb{C}}} is holomorphic for every n0n\geq 0. We also define Xn:=^φn(A)X_{n}:=\widehat{{\mathbb{C}}}-\varphi_{n}(A), Yn:=^ψn(A)Y_{n}:=\widehat{{\mathbb{C}}}-\psi_{n}(A), and Zn:=^gn1(ψn(A))¯=^ψn+1(f1(A))¯Z_{n}:=\widehat{{\mathbb{C}}}-\overline{g_{n}^{-1}(\psi_{n}(A))}=\widehat{{\mathbb{C}}}-\overline{\psi_{n+1}(f^{-1}(A))} for all n0n\geq 0. In particular, gn|Zn:ZnYng_{n}|Z_{n}\colon Z_{n}\to Y_{n} is a holomorphic covering map.

Claim.

We have lγ(μn)<l_{\gamma}(\mu_{n})<\ell^{*} for all n0n\geq 0.

Proof.

Clearly, lγ(μ0)<l_{\gamma}(\mu_{0})<\ell^{*} by the choice of μ\mu. We will prove that lγ(μ1)<l_{\gamma}(\mu_{1})<\ell^{*}, and the rest easily follows by induction on nn. There exists a short simple closed geodesic β\beta in Y0Y_{0} such that β\beta and ψ0(γ)\psi_{0}(\gamma) are homotopic in Y0Y_{0}, and Y0(β)=lγ(μ0)\ell_{Y_{0}}(\beta)=l_{\gamma}(\mu_{0}). Define β~:=ψ01(β)\widetilde{\beta}:=\psi^{-1}_{0}(\beta). Since β~\widetilde{\beta} and γ\gamma are homotopic in S2AS^{2}-A, then by Corollary 2.3, there exists a simple closed curve α~f1(α~)\widetilde{\alpha}\subset f^{-1}(\widetilde{\alpha}) that is homotopic to γ\gamma in S2AS^{2}-A, since γ\gamma is a Levy fixed curve for f:(S2,A)f\colon(S^{2},A)\righttoleftarrow, and deg(f|α~:α~β~)=1\deg(f|\widetilde{\alpha}\colon\widetilde{\alpha}\to\widetilde{\beta})=1. Now, define α:=ψ1(α~)\alpha:=\psi_{1}(\widetilde{\alpha}), which is homotopic in Y1Y_{1} to ψ1(γ)\psi_{1}(\gamma). Clearly, g(α)=βg(\alpha)=\beta, deg(g|α:αβ)=1\deg(g|\alpha\colon\alpha\to\beta)=1, and αZ0\alpha\subset Z_{0}. Therefore, by Schwarz-Pick’s lemma, we have

Y1(α)Z0(α)=Y0(β)<.\ell_{Y_{1}}(\alpha)\leq\ell_{Z_{0}}(\alpha)=\ell_{Y_{0}}(\beta)<\ell^{*}.

Finally, lγ(μ1)=lα~(μ1)<l_{\gamma}(\mu_{1})=l_{\widetilde{\alpha}}(\mu_{1})<\ell^{*}. ∎

Since (π(τn))(\pi(\tau_{n})) eventually leaves every compact subset of A\mathcal{M}_{A}, Mumford’s compactness theorem [Hub06, Theorem 7.3.3] states that the length of the shortest simple closed geodesic δn\delta_{n} in XnX_{n} tends to zero. Then wφn1(δn)(μn)<logw_{\varphi_{n}^{-1}(\delta_{n})}(\mu_{n})<\log\ell^{*} for any sufficiently large nn, given that wφn1(δ):𝒯Aw_{\varphi_{n}^{-1}(\delta)}\colon\mathcal{T}_{A}\to{\mathbb{R}} is 1-Lipschitz and dT(μn,τn)dT(τ,μ)d_{T}(\mu_{n},\tau_{n})\leq d_{T}(\tau,\mu) by Proposition 2.19 and Schwarz-Pick’s lemma. Thus, ψn(γ)\psi_{n}(\gamma) and ψn(φn1(δn))\psi_{n}(\varphi_{n}^{-1}(\delta_{n})) are homotopic in YnY_{n} to short simple closed geodesics. Since Yn=^ψn(A)Y_{n}=\widehat{{\mathbb{C}}}-\psi_{n}(A) is a four-punctured Riemann sphere, it follows that γ\gamma and φn1(δn)\varphi_{n}^{-1}(\delta_{n}) are homotopic in S2AS^{2}-A for any nn large enough.

This argument applies to any Levy fixed curve of f:(S2,A)f\colon(S^{2},A)\righttoleftarrow. Therefore, it has to be unique up to homotopy in S2AS^{2}-A. ∎

Remark 3.3.

If an obstructed Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow with |A|3|A|\geq 3 has a finite degree and hyperbolic orbifold, then it is known (see [Hub16, Section 10.9 and Lemma 10.11.9] or [Sel12, Proof of Theorem 2.3, p. 20]) that every point τ\tau in the Teichmüller space 𝒯A\mathcal{T}_{A} satisfies the condition of Proposition 3.2.

Remark 3.4.

Suppose that we are in the setting of Proposition 3.2. It is clear that it suffices to require that the sequence (π(σfn(τ)))(\pi(\sigma_{f}^{\circ n}(\tau))) admits a subsequence eventually leaving every compact subset of the moduli space A\mathcal{M}_{A}.

3.2. Estimating contraction

The following proposition provides an estimate for the contraction of an inclusion between hyperbolic Riemann surfaces. Although this result is well-known [McM94, Theorem 2.25], we include the proof here for the sake of completeness.

Proposition 3.5.

Suppose that UU and VV are two hyperbolic Riemann surfaces such that UVU\subsetneq V. Let zUz\in U and s:=dV(z,VU)s:=d_{V}(z,V-U). Then

cUV(z)2r|logr|1r2,c_{U}^{V}(z)\leq\frac{2r|\log r|}{1-r^{2}},

where r:=tanh(s/2)r:=\tanh(s/2).

Proof.

Suppose that z0VUz_{0}\in V-U is a point such that s=dV(z,z0)s=d_{V}(z,z_{0}). Let p:𝔻Vp\colon{\mathbb{D}}\to V be a holomorphic universal covering with p(0)=z0p(0)=z_{0}. Let us chose wp1(z)w\in p^{-1}(z) such that d𝔻(0,w)=sd_{{\mathbb{D}}}(0,w)=~{}s. Finally, we denote by U~𝔻\widetilde{U}\subset{\mathbb{D}} the connected component of ww in p1(U)p^{-1}(U). Then we have the following commutative diagram:

𝔻{{\mathbb{D}}^{*}}U~{\widetilde{U}}U{U}𝔻{{\mathbb{D}}}V{V}p|U~\scriptstyle{p|\widetilde{U}}p\scriptstyle{p}

Applying Schwarz-Pick’s lemma and recalling that p:𝔻Vp\colon{\mathbb{D}}\to V and p|U~:U~Up|\widetilde{U}\colon\widetilde{U}\to U are holomorphic covering maps, we obtain the following:

cUV(z)=cU~𝔻(z)=ρ𝔻(w)ρU~(w)ρ𝔻(w)ρ𝔻(w).c_{U}^{V}(z)=c_{\widetilde{U}}^{{\mathbb{D}}}(z)=\frac{\rho_{{\mathbb{D}}}(w)}{\rho_{\widetilde{U}}(w)}\leq\frac{\rho_{{\mathbb{D}}}(w)}{\rho_{{\mathbb{D}}^{*}}(w)}.

The rest easily follows since ρ𝔻(w)=2/(1|w|2)\rho_{{\mathbb{D}}}(w)=2/(1-|w|^{2}), ρ𝔻(w)=1/|wlog|w||\rho_{{\mathbb{D}}^{*}}(w)=1/|w\log|w|| ([Hub06, Example 3.3.2]), and s=d𝔻(0,w)=2tanh1(|w|)s=d_{{\mathbb{D}}}(0,w)=2\tanh^{-1}(|w|) due to [Hub06, Exercise 2.1.8]. ∎

Remark 3.6.

Suppose that we are in the setting of Proposition 3.5. It shows that there exists an upper bound λ(s)<1\lambda(s)<1 for cUV(z)c_{U}^{V}(z) depending only on s=dV(z,VU)s=d_{V}(z,V-U) and not on UU, VV, or zz. Moreover, λ(s)|slog(s)|\lambda(s)\sim|s\log(s)| as s0s\to 0. In particular, λ(s)0\lambda(s)\to 0 as s0s\to 0.

Proposition 3.5 and Remark 3.6 allow us to estimate the contraction of certain inclusions between countably and finitely punctured Riemann spheres (see [Rem09, Lemma 2.1] for the result of a similar nature).

Proposition 3.7.

Suppose that (zn)(z_{n}) is a sequence in {\mathbb{C}} such that limnzn=\lim_{n\to\infty}z_{n}=\infty and limnlog|zn+1|/log|zn|=1\lim_{n\to\infty}\log|z_{n+1}|/\log|z_{n}|=1. Let PP be a finite set consisting of at least two elements of (zn)(z_{n}). Then cUV(z)c_{U}^{V}(z) converges to 0 as |z||z| tends to \infty for zUz\in U, where U:={zn:n}U:={\mathbb{C}}-\{z_{n}:n\in{\mathbb{N}}\} and V:=PV:={\mathbb{C}}-P.

Proof.

Without loss of generality, we can assume that P𝔻P\subset{\mathbb{D}} and the sequence of absolute values (|zn|)(|z_{n}|) is non-decreasing. Let Q:={zn:n}Q:=\{z_{n}:n\in{\mathbb{N}}\}, and let p:𝔻¯p\colon{\mathbb{H}}\to{\mathbb{C}}-\overline{{\mathbb{D}}} be the holomorphic universal covering defined as p(z)=exp(iz)p(z)=\exp(-iz) for zz\in{\mathbb{H}}.

According to Proposition 3.5 and Remark 3.6, it suffices to demonstrate that for any r>0r>0, the set xQPBV(x,r)\bigcup_{x\in Q-P}B_{V}(x,r) covers some punctured neighborhood of \infty in ^\widehat{{\mathbb{C}}}. Alternatively, it is sufficient to show that for any r>0r>0, there exists t>0t>0 so that

(1) xp1(Q𝔻¯)B(x,r)t:={z:Im(z)>t}.\displaystyle\bigcup_{x\in p^{-1}(Q-\overline{{\mathbb{D}}})}B_{{\mathbb{H}}}(x,r)\supset{\mathbb{H}}_{t}:=\{z\in{\mathbb{C}}:\operatorname{Im}(z)>t\}.

Indeed, p(t)=𝔻¯etp({\mathbb{H}}_{t})={\mathbb{C}}-\overline{{\mathbb{D}}}_{e^{t}} and p(B(x,r))BV(p(x),r)p(B_{{\mathbb{H}}}(x,r))\subset B_{V}(p(x),r) due to Schwarz-Pick’s lemma.

Note that for sufficiently large nn, we have that |zn|>1|z_{n}|>1 and

(2) π1(zn)={2πkarg(zn)+ilog(|zn|):k}=:{zn,k:k}.\displaystyle\pi^{-1}(z_{n})=\{2\pi k-\arg(z_{n})+i\log(|z_{n}|):k\in{\mathbb{Z}}\}=:\{z_{n,k}:k\in{\mathbb{Z}}\}.

Observation (2) implies that

(3) dn,k:=diam({z:2πkRe(z)2π(k+1),Im(zn,k)Im(z)Im(zn+1,k)})\displaystyle d_{n,k}:=\mathrm{diam}_{{\mathbb{H}}}\left(\left\{z:2\pi k\leq\operatorname{Re}(z)\leq 2\pi(k+1),\operatorname{Im}(z_{n,k})\leq\operatorname{Im}(z)\leq\operatorname{Im}(z_{n+1,k})\right\}\right)\leq
log(log|zn+1|log|zn|)+2πlog|zn|.\displaystyle\leq\log\left(\frac{\log|z_{n+1}|}{\log|z_{n}|}\right)+\frac{2\pi}{\log|z_{n}|}.

Estimate (3) is derived from the following well-known facts about the hyperbolic distance between two points z,wz,w\in{\mathbb{H}}:

  • if Re(z)=Re(w)\operatorname{Re}(z)=\operatorname{Re}(w), then d(z,w)=|log(Im(z)/Im(w))|d_{{\mathbb{H}}}(z,w)=|\log(\operatorname{Im}(z)/\operatorname{Im}(w))|, and

  • if Im(z)=Im(w)\operatorname{Im}(z)=\operatorname{Im}(w), then d(z,w)|Re(z)Re(w)|/Im(z)d_{{\mathbb{H}}}(z,w)\leq|\operatorname{Re}(z)-\operatorname{Re}(w)|/\operatorname{Im}(z). In fact, the upper estimate is the hyperbolic length of the horizontal segment connecting zz and ww.

Finally, estimate (3) implies that dn,kd_{n,k} tends to zero independently from kk when nn\to\infty. Thus, inclusion (1) holds for certain t>0t>0, and the desired result follows. ∎

Remark 3.8.

Suppose that we are in the setting of Proposition 3.7. Assume that we know that log|zn+1|/log|zn|\log|z_{n+1}|/\log|z_{n}| is uniformly bounded from above but does not necessarily converge to 11 as nn\to\infty. Following the proof of Proposition 3.7, one can show that there exists λ<1\lambda<1 so that cUV(z)<λc_{U}^{V}(z)<\lambda for any zWUz\in W\cap U, where W^W\subset\widehat{{\mathbb{C}}} is a neighborhood of \infty.

As an application of Proposition 3.7, we can obtain the following result.

Proposition 3.9.

Suppose that g:UVg\colon U\to V is a holomorphic covering map, where UVU\subset V is a domain of ^\widehat{{\mathbb{C}}} and V=^PV=\widehat{{\mathbb{C}}}-P with 3|P|<3\leq|P|<\infty. Let xPx\in P be an accumulation point of the set UU. Then exactly one of the following two possibilities is satisfied:

  1. (1)

    xx is an isolated removable singularity of g:UVg\colon U\to V, or

  2. (2)

    cUV(z)c_{U}^{V}(z) converges to 0 as |zx||z-x| tends to 0 for zUz\in U.

Proof.

Without loss of generality, we can assume that x=x=\infty. Suppose that condition (1) is not satisfied. We will prove that there exists a constant a>1a>1 such that for every sufficiently large r>0r>0, the annulus A(r/a,ra):={z:r/a<|z|<ra}A(r/a,ra):=\{z\in{\mathbb{C}}:r/a<|z|<ra\} contains a point of U{\mathbb{C}}-U (see [BR20, Lemma 3.2] for a similar result in the setting of entire maps). Suppose the contrary, i.e., that there exist sequences (an)(a_{n}) and (rn)(r_{n}) such that an>1a_{n}>1, ana_{n}\to\infty, rnr_{n}\to\infty, and An:=A(rn/an,rnan)UA_{n}:=A(r_{n}/a_{n},r_{n}a_{n})\subset U for every nn\in{\mathbb{N}}.

In particular, gg is defined and meromorphic on AnA_{n}, allowing us to consider the sequence of meromorphic functions gn:A(1/an,an)^g_{n}\colon A(1/a_{n},a_{n})\to\widehat{{\mathbb{C}}}, nn\in{\mathbb{N}}, where gn(z)=g(rnz)g_{n}(z)=g(r_{n}z) for every zA(1/an,an)z\in A(1/a_{n},a_{n}). By Montel’s theorem, the family (gn)(g_{n}) is normal since each gng_{n} omits the values in the set PP. Thus, there exists a subsequence of (gn)(g_{n}) that converges locally uniformly on the punctured plane {\mathbb{C}}^{*}. Without loss of generality, we assume that this subsequence coincides with the original sequence (gn)(g_{n}). Let h:^h\colon{\mathbb{C}}^{*}\to\widehat{{\mathbb{C}}} be the limiting function. If hh is not constant, then by Hurwitz’s theorem [Gam01, p. 231], hh also omits the values in the set PP. This leads to a contradiction because either hh would have an isolated essential singularity at 0 or \infty, contradicting Great Picard’s theorem, or both 0 and \infty would be removable singularities or poles of hh, and thus, hh is a rational map, which cannot omit any values.

Therefore, hh is a constant map. Denote by w^w\in\widehat{{\mathbb{C}}} its unique value. Choose a simply connected domain WwW\ni w such that W¯P={w}P\overline{W}\cap P=\{w\}\cap P. From the previous discussion, g(A(2rn/an,anrn/2))Wg(A(2r_{n}/a_{n},a_{n}r_{n}/2))\subset W for all nn large enough. Note that every connected component of g1(W)Ug^{-1}(W)\subset U is either a simply connected domain or a simply connected domain with a single puncture (see, for instance, [For91, Theorems 5.10 and 5.11]). Hence, for every sufficiently large nn, either all but at most one of the points in 𝔻anrn/2{\mathbb{D}}_{a_{n}r_{n}/2} belong to g1(W)g^{-1}(W), or the same is true for the set ^𝔻¯2rn/an\widehat{{\mathbb{C}}}-\overline{{\mathbb{D}}}_{2r_{n}/a_{n}}. The first case cannot happen for infinitely many nn, so there must be some nn\in{\mathbb{N}} for which the second case occurs. It immediately leads to a contradiction, as we initially assumed that xx is not an isolated removable singularity of the map g:UVg\colon U\to V.

Now, we can take znz_{n} to be a point of ^U\widehat{{\mathbb{C}}}-U in A(a2n,a2n+2)A(a^{2n},a^{2n+2}) for every sufficiently large nn. It is easy to see that the sequence (zn)(z_{n}) satisfies the conditions of Proposition 3.7. Given that

cUV(z)=ρV(z)ρU(z)ρV(z)ρQ(z),c_{U}^{V}(z)=\frac{\rho_{V}(z)}{\rho_{U}(z)}\leq\frac{\rho_{V}(z)}{\rho_{{\mathbb{C}}-Q}(z)},

where Q:=P{zn:n}Q:=P\cup\{z_{n}:n\in{\mathbb{N}}\}, item (2) follows from Proposition 3.7. ∎

3.3. Iteration on the unit disk

If h:𝔻𝔻h\colon{\mathbb{D}}\to{\mathbb{D}} is a non-injective holomorphic map, the Denjoy-Wolff theorem [Aba23, Theorem 3.2.1] states that any point z𝔻z\in{\mathbb{D}} converges under iteration of hh to a point z0𝔻¯z_{0}\in\overline{{\mathbb{D}}} that is independent of the initial choice of zz. In this section, we explore holomorphic maps on the unit disk that satisfy stronger assumptions, allowing us to achieve more precise control on the behavior of their orbits. Many examples of such maps will appear in Section 4. Specifically, many pullback maps of (marked) Thurston maps with a marked set AA, where |A|=4|A|=4, satisfy these conditions.

We say that z^z\in\widehat{{\mathbb{C}}} is a regular point of a holomorphic map g:UVg\colon U\to V, where UU and VV are domains of ^\widehat{{\mathbb{C}}}, if either zUz\in U and deg(g,z)=1\deg(g,z)=1, or zz is an isolated removable singularity of gg and, after extending gg holomorphically to a neighborhood of zz, deg(g,z)=1\deg(g,z)=1. A point z^z\in\widehat{{\mathbb{C}}} is a fixed point of the map g:UVg\colon U\to V if either zUz\in U and g(z)=zg(z)=z, or zz is an isolated removable singularity of gg and, after extending gg holomorphically to a neighborhood of zz, we have g(z)=zg(z)=z. The concepts of repelling or attracting fixed points can be generalized in a similar way.

Theorem 3.10.

Let h:𝔻𝔻h\colon{\mathbb{D}}\to{\mathbb{D}} be a holomorphic map, and π:𝔻V\pi\colon{\mathbb{D}}\to V and g:UVg\colon U\to V are holomorphic covering maps, where UVU\subset V is a domain of ^\widehat{{\mathbb{C}}} and V=^PV=\widehat{{\mathbb{C}}}-P with 3|P|<3\leq|P|<\infty. Suppose that π(h(𝔻))U\pi(h({\mathbb{D}}))\subset U and π=gπh\pi=g\circ\pi\circ h, i.e., the following diagram commutes:

𝔻{{\mathbb{D}}}h(𝔻){h({\mathbb{D}})}V{V}U{U}h\scriptstyle{h}π\scriptstyle{\pi}π\scriptstyle{\pi}g\scriptstyle{g}

If the map g:UVg\colon U\to V is non-injective, then exactly one of the following two possibilities is satisfied:

  1. (1)

    for every z𝔻z\in{\mathbb{D}}, the hh-orbit of zz converges to the unique fixed point of hh, or

  2. (2)

    the sequence (π(hn(z)))(\pi(h^{\circ n}(z))) converges to the same repelling fixed point xPx\in P of the map gg, regardless of the choice of z𝔻z\in{\mathbb{D}}.

Proof.

Given that the maps π:𝔻V\pi\colon{\mathbb{D}}\to V and gπ|U~:U~Vg\circ\pi|\widetilde{U}\colon\widetilde{U}\to V are covering maps, where U~:=π1(U)\widetilde{U}:=\pi^{-1}(U), it follows [Hat02, Section 1.3, Exercise 16] that h:𝔻U~h\colon{\mathbb{D}}\to\widetilde{U} is also a covering map and, in particular, U~\widetilde{U} is connected. Therefore, according to Schwarz-Pick’s lemma, we have Dh(z)𝔻=cU~𝔻(h(z))=cUV(π(h(z)))\|\mathrm{D}h(z)\|_{{\mathbb{D}}}=c_{\widetilde{U}}^{{\mathbb{D}}}(h(z))=c_{U}^{V}(\pi(h(z))) for every z𝔻z\in{\mathbb{D}}. At the same time, Great Picard’s theorem and the Riemann-Hurwitz formula [Hub06, Appendix A.3] imply that VUV-U contains at least one point. Hence, the inclusion I:UVI\colon U\hookrightarrow V is locally uniformly contracting with respect to the hyperbolic metrics on UU and VV. As a result, hh is locally uniformly contracting with respect to the hyperbolic metric on 𝔻{\mathbb{D}}. In particular, hh has at most one fixed point, and if there exists a point z𝔻z\in{\mathbb{D}} such that its orbit (hn(z))(h^{\circ n}(z)) converges in 𝔻{\mathbb{D}}, then every orbit of hh converges to the unique fixed point of hh.

Let us pick an arbitrary point z0𝔻z_{0}\in{\mathbb{D}}. Define zn=hn(z0)z_{n}=h^{\circ n}(z_{0}) and xn=π(hn(z0))x_{n}=\pi(h^{\circ n}(z_{0})) for n0n\geq 0. Connect the points z0z_{0} and z1z_{1} by the hyperbolic geodesic δ0𝔻\delta_{0}\subset{\mathbb{D}}. We denote by δn𝔻\delta_{n}\subset{\mathbb{D}} the curve hn(δ)h^{\circ n}(\delta) that connects znz_{n} and zn+1z_{n+1}. By Schwarz-Pick’s lemma, the sequence (𝔻(δn))(\ell_{{\mathbb{D}}}(\delta_{n})) is non-increasing. In particular, if d0=d𝔻(z0,z1)d_{0}=d_{{\mathbb{D}}}(z_{0},z_{1}), then d𝔻(zn,zn+1)𝔻(δn)d0d_{{\mathbb{D}}}(z_{n},z_{n+1})\leq\ell_{{\mathbb{D}}}(\delta_{n})\leq d_{0} and dV(xn,xn+1)V(π(δn))d0d_{V}(x_{n},x_{n+1})\leq\ell_{V}(\pi(\delta_{n}))\leq d_{0} for every n0n\geq 0.

Further, we structure the proof as a series of claims.

Claim 1.

If there exists a compact set KVK\subset V such that xnKx_{n}\in K for infinitely many nn, then the sequence (zn)(z_{n}) converges in 𝔻{\mathbb{D}}.

Proof.

Since V(π(δn))d0\ell_{V}(\pi(\delta_{n}))\leq d_{0}, we can enlarge KK so that π(δn)K\pi(\delta_{n})\subset K for infinitely many nn. First, we demonstrate that there exists λ<1\lambda<1 so that cUV(z)λc_{U}^{V}(z)\leq\lambda for all zKUz\in K\cap U. Indeed, as we mentioned earlier, there exists a point wUVw\in U-V. Therefore, for any zKz\in K, the distance dV(z,VU)dV(z,w)d_{V}(z,V-U)\leq d_{V}(z,w) is uniformly bounded from above. Hence, Proposition 3.5 and Remark 3.6 imply that such λ\lambda exists. Finally, from the previous discussions, if π(δn+1)K\pi(\delta_{n+1})\subset K, then 𝔻(δn+1)λ𝔻(δn)\ell_{{\mathbb{D}}}(\delta_{n+1})\leq\lambda\ell_{{\mathbb{D}}}(\delta_{n}). In particular, this shows that (𝔻(δn))(\ell_{{\mathbb{D}}}(\delta_{n})) converges to 0 as nn\to\infty.

There exists a subsequence of (xn)(x_{n}) that converges to a point xKx\in K. By enlarging KK, we can assume that BV(x,2r)KB_{V}(x,2r)\subset K for a certain r>0r>0. Now, we choose mm such that 𝔻(δm)<r(1λ)/2\ell_{{\mathbb{D}}}(\delta_{m})<r(1-\lambda)/2 and dV(xm,x)<rd_{V}(x_{m},x)<r. Notice that mm is chosen so that 𝔻(δm)i=0λi<r/2\ell_{{\mathbb{D}}}(\delta_{m})\sum_{i=0}^{\infty}\lambda^{i}<r/2. Using this fact and applying induction (see [Sel12, Proof of Theorem 2.3, p. 20] for a similar argument), it can be shown that π(δn)BV(x,r)\pi(\delta_{n})\subset B_{V}(x,r) and V(π(δn+1))λV(π(δn))\ell_{V}(\pi(\delta_{n+1}))\leq\lambda\ell_{V}(\pi(\delta_{n})) for all nmn\geq m. In other words, the distance between znz_{n} and zn+1z_{n+1} decreases exponentially, and the convergence follows from the completeness of the hyperbolic metric on 𝔻{\mathbb{D}}. ∎

Claim 2.

If the sequence (xn)(x_{n}) converges to xPx\in P along some subsequence, then the entire sequence (xn)(x_{n}) converges to xx.

Proof.

For any xPx\in P, we choose a neighborhood Vx^V_{x}\subset\widehat{{\mathbb{C}}} so that if xx′′x^{\prime}\neq x^{\prime\prime}, then dV(y1,y2)>d0d_{V}(y_{1},y_{2})>~{}d_{0} for any y1VxPy_{1}\in V_{x^{\prime}}-P and y2Vx′′Py_{2}\in V_{x^{\prime\prime}}-P. Now, suppose that the sequence (xn)(x_{n}) has subsequences that converge to different limits y^y^{\prime}\in\widehat{{\mathbb{C}}} and y′′^y^{\prime\prime}\in\widehat{{\mathbb{C}}}, respectively. If either of yy^{\prime} or y′′y^{\prime\prime} does not belong to PP, then by Claim 1, the sequence (zn)(z_{n}) must converge to a limit in 𝔻{\mathbb{D}}, which leads to a contradiction. If we instead assume that y,y′′Py^{\prime},y^{\prime\prime}\in P, then it follows that xnVxPVxx_{n}\in V-\bigcup_{x\in P}V_{x} for infinitely many nn. Once again, Claim 1 implies that (zn)(z_{n}) must converge to a limit in 𝔻{\mathbb{D}}, and it also leads to a contradiction. Thus, the entire sequence (xn)(x_{n}) converges to xPx\in P as nn\to\infty. ∎

Claim 3.

If the sequence (xn)(x_{n}) converges to xPx\in P, then xx is a fixed point of the map g:UVg\colon U\to V.

Proof.

Suppose xPx\in P is not an isolated removable singularity of gg. Observe that for sufficiently large nn, π(δn)\pi(\delta_{n}) lies within any given neighborhood of xx. Therefore, based on Proposition 3.9 and the fact that Dh(z)𝔻=cUV(π(h(z)))\|\mathrm{D}h(z)\|_{{\mathbb{D}}}=c_{U}^{V}(\pi(h(z))) for all z𝔻z\in{\mathbb{D}}, we have 𝔻(δn+1)λ𝔻(δn)\ell_{{\mathbb{D}}}(\delta_{n+1})\leq\lambda\ell_{{\mathbb{D}}}(\delta_{n}) for nn large enough and some λ<1\lambda<1. This means that the distance between znz_{n} and zn+1z_{n+1} decreases exponentially, so the sequence (zn)(z_{n}) must have a limit in 𝔻{\mathbb{D}}, leading to a contradiction. Therefore, xPx\in P is an isolated removable singularity of gg. Finally, since xn=g(xn+1)x_{n}=g(x_{n+1}) for all n0n\geq 0, xx must be a fixed point of the map gg. ∎

Claim 4.

If the sequence (xn)(x_{n}) converges to a fixed point xPx\in P of the map gg, then xx is a repelling fixed point of gg.

Proof.

Without loss of generality, we assume x=0x=0. Let us choose a continuous parametrization δ0:𝕀𝔻\delta_{0}\colon{\mathbb{I}}\to{\mathbb{D}} for the arc δ0\delta_{0}. We then define a continuous curve δ:[0,+)𝔻\delta\colon[0,+\infty)\to{\mathbb{D}} by setting δ(t+n)=hn(δ0(t))\delta(t+n)=h^{\circ n}(\delta_{0}(t)) for any non-negative integer nn and t𝕀t\in{\mathbb{I}}. It is straightforward to see that π(δ(t))\pi(\delta(t)) converges to 0 as tt tends to \infty, since π(δ(n))=xnx=0\pi(\delta(n))=x_{n}\to x=0 as nn\to\infty and V(π(δ([n,n+1])))d0\ell_{V}(\pi(\delta([n,n+1])))\leq d_{0} for all n0n\geq 0. Furthermore, note that π(δ(t))=g(π(δ(t+1))\pi(\delta(t))=g(\pi(\delta(t+1)) for t0t\geq 0. Applying the Snail Lemma [Mil06, Lemma 13.2, Corollorary 13.3] to the curve πδ\pi\circ\delta and the map gg in a neighborhood of 0, we obtain that either |g(0)|>1|g^{\prime}(0)|>1 or g(0)=1g^{\prime}(0)=1.

Suppose g(0)=1g^{\prime}(0)=1. Further, we assume that gg is extended holomorphically to a neighborhood 0. If we choose an arbitrary simply connected domain D^D\subset\widehat{{\mathbb{C}}} such that DP={0}D\cap P=\{0\}, then g1(D)g^{-1}(D) has a connected component DD^{\prime} containing 0. Moreover, DU{0}D^{\prime}\subset U\cup\{0\} is simply connected and g|D:DDg|D^{\prime}\colon D^{\prime}\to D is a biholomorphism that fixes 0. Similarly, we can define the local inverse branches φn:DU{0}\varphi_{n}\colon D\to U\cup\{0\} of gng^{\circ n} in a neighborhood of 0. In other words, φn:=(gn|Dn)1\varphi_{n}:=(g^{\circ n}|D_{n})^{-1}, where DnD_{n} is the connected component of 0 in gn(D)g^{-n}(D). In particular, φn\varphi_{n} is a biholomorphism, φn(0)=0\varphi_{n}(0)=0, and (φn)(0)=1/(gn)(0)=1(\varphi_{n})^{\prime}(0)=1/(g^{\circ n})^{\prime}(0)=1 for every nn\in{\mathbb{N}}.

Due to the previous discussions, U{0}U\cup\{0\} does not contain at least three points of ^\widehat{{\mathbb{C}}}. Hence, Montel’s theorem implies that the family (φn)(\varphi_{n}) is normal. Therefore, up to a subsequence, it converges locally uniformly on DD to some holomorphic map φ\varphi such that φ(0)=0\varphi(0)=0 and φ(0)=1\varphi^{\prime}(0)=1. Since φ(0)=1\varphi^{\prime}(0)=1, φ\varphi is injective in a neighborhood of 0. Thus, the iterates (gn)(g^{\circ n}) converge uniformly in a neighborhood of 0 to φ1\varphi^{-1}, up to a subsequence. However, if

g(z)=z+akzk+,g(z)=z+a_{k}z^{k}+\dots,

where ak0a_{k}\neq 0, k2k\geq 2, and three dots represent higher order terms, then

gn(z)=z+nakzk+,g^{\circ n}(z)=z+na_{k}z^{k}+\dots,

so (gn)(k)(0)(g^{\circ n})^{(k)}(0) diverges as nn\to\infty. This leads to a contradiction ruling out the possibility g(0)=1g^{\prime}(0)=1. Thus, x=0x=0 must be a repelling fixed point of the map gg. ∎

Claims 1-4 imply that if the sequence (hn(z0))(h^{\circ n}(z_{0})) diverges in 𝔻{\mathbb{D}}, then the sequence (π(hn(z0)))(\pi(h^{\circ n}(z_{0}))) converges to a repelling fixed point xPx\in P of the map gg. Furthermore, in this case (π(hn(z)))(\pi(h^{\circ n}(z))) converges to xx for every z𝔻z\in{\mathbb{D}}. Indeed, this follows easily since dV(π(hn(z0)),π(hn(z)))d_{V}(\pi(h^{\circ n}(z_{0})),\pi(h^{\circ n}(z))) is bounded from above by d𝔻(z0,z)d_{{\mathbb{D}}}(z_{0},z) due to Schwarz-Pick’s lemma. ∎

4. Thurston theory

In this section, we focus on the study of (marked) Thurston maps f:(S2,A)f\colon(S^{2},A)\righttoleftarrow that satisfy the following two conditions:

  1. (I)

    the marked set AA contains exactly four points, and

  2. (II)

    there exists a set BAB\subset A such that |B|=3|B|=3, SfBS_{f}\subset B, and |f1(B)¯A|=3|\overline{f^{-1}(B)}\cap A|=3.

It is evident that when the marked set AA coincides with the postsingular set PfP_{f}, conditions (I) and (II) are equivalent to conditions (A)(C) from Section 1.3. Also, it is worth noting that the case when |f1(B)¯A|=4|\overline{f^{-1}(B)}\cap A|=4 or, equivalently, Af1(B)¯A\subset\overline{f^{-1}(B)}, is rather trivial due to Remark 2.18.

Using the tools developed Section 3 and the theory of iteration of meromorphic functions, we analyze the corresponding pullbacks map defined on the one-complex dimensional Teichmüller space. It allows us to derived several properties of the corresponding Thurston maps and their Hurwitz classes. In particular, in Section 4.1 we prove Main Theorem A (see Theorem 4.4), and in Section 4.2 we prove Main Theorem B (see Theorem 4.8) and Corollary 1.1.

Let A={a1,a2,a3,a4}A=\{a_{1},a_{2},a_{3},a_{4}\}, B={ai1,ai2,ai3}B=\{a_{i_{1}},a_{i_{2}},a_{i_{3}}\}, and C:=f1(B)¯A={aj1,aj2,aj3}C:=\overline{f^{-1}(B)}\cap A=\{a_{j_{1}},a_{j_{2}},a_{j_{3}}\}, where i1<i2<i3i_{1}<i_{2}<i_{3} and j1<j2<j3j_{1}<j_{2}<j_{3}. Additionally, assume that ii and jj are the indices so that aiABa_{i}\in A-B and ajACa_{j}\in A-C. Under this conventions, we have that f(aj)=aif(a_{j})=a_{i} and deg(f,aj)=1\deg(f,a_{j})=1.

It is important to note that there may be multiple choices for the set BB. However, when |Sf|=3|S_{f}|=3, the set BB is uniquely determined by the properties described in condition (II). Also, another choice that we made, which will be relevant in our further arguments, is the indexing for the set AA.

Now we are ready to introduce the following objects:

  • the map πB:𝒯AΣ\pi_{B}\colon\mathcal{T}_{A}\to\Sigma, where Σ=^{0,1,}\Sigma=\widehat{{\mathbb{C}}}-\{0,1,\infty\}, is defined by πB([φ])=φ(ai)\pi_{B}([\varphi])=\varphi(a_{i}), where the representative φ:S2^\varphi\colon S^{2}\to~{}\widehat{{\mathbb{C}}} is chosen so that φ(ai1)=0\varphi(a_{i_{1}})=0, φ(ai2)=1\varphi(a_{i_{2}})=1, and φ(ai3)=\varphi(a_{i_{3}})=\infty;

  • the map πC:𝒯AΣ\pi_{C}\colon\mathcal{T}_{A}\to\Sigma is defined by πC([φ])=φ(aj)\pi_{C}([\varphi])=\varphi(a_{j}), where the representative φ:S2^\varphi\colon S^{2}\to~{}\widehat{{\mathbb{C}}} is chosen so that φ(aj1)=0\varphi(a_{j_{1}})=0, φ(aj2)=1\varphi(a_{j_{2}})=1, and φ(aj3)=\varphi(a_{j_{3}})=\infty;

  • the map ωf:𝒯AΣ\omega_{f}\colon\mathcal{T}_{A}\to\Sigma is defined by ωf=πCσf\omega_{f}=\pi_{C}\circ\sigma_{f};

  • the map Ff:^^F_{f}\colon\widehat{{\mathbb{C}}}\dashrightarrow\widehat{{\mathbb{C}}} is a unique (which follows from Proposition 2.16) holomorphic map such that Ff=φfψ1F_{f}=\varphi\circ f\circ\psi^{-1}, where φ,ψ:S2^\varphi,\psi\colon S^{2}\to\widehat{{\mathbb{C}}} are orientation-preserving homeomorphisms satisfying φ(ai1)=ψ(aj1)=0\varphi(a_{i_{1}})=\psi(a_{j_{1}})=~{}0, φ(ai2)=ψ(aj2)=1\varphi(a_{i_{2}})=\psi(a_{j_{2}})=1, and φ(ai3)=ψ(aj3)=\varphi(a_{i_{3}})=\psi(a_{j_{3}})=\infty;

  • the map Mi,j:ΣΣM_{i,j}\colon\Sigma\to\Sigma is defined as hihj1h_{i}\circ h_{j}^{-1}, where hi:AΣh_{i}\colon\mathcal{M}_{A}\to\Sigma is defined by hi([η])=η(ai)h_{i}([\eta])=\eta(a_{i}), with η:A^\eta\colon A\to\widehat{{\mathbb{C}}} chosen so that η(ai1)=0\eta(a_{i_{1}})=0, η(ai2)=1\eta(a_{i_{2}})=1, η(ai3)=\eta(a_{i_{3}})=\infty, and the map hj:AΣh_{j}\colon\mathcal{M}_{A}\to\Sigma is defined analogously.

  • the set Wf:=^Ff1({0,1,})¯W_{f}:=\widehat{{\mathbb{C}}}-\overline{F_{f}^{-1}(\{0,1,\infty\})} is a domain of ^\widehat{{\mathbb{C}}};

  • the map Gf:=FfMi,j1:Mi,j(Wf)ΣG_{f}:=F_{f}\circ M_{i,j}^{-1}\colon M_{i,j}(W_{f})\to\Sigma.

Of course, the maps FfF_{f} and GfG_{f}, as well as the domain WfW_{f}, also depend on the choice of the subset BB, and many other objects defined above depend on the indexing of AA. However, for the simplicity, we are going to exclude these dependencies from the notation. Throughout this section, we will maintain the notation and conventions established above. Specifically, if f~:(S2,A)\widetilde{f}\colon(S^{2},A)\righttoleftarrow is any Thurston map that is Hurwitz equivalent to f:(S2,A)f\colon(S^{2},A)\righttoleftarrow, then we use the same set BB and the same indexing of the set AA when we work with the map f~\widetilde{f} as we do it with ff.

Proposition 4.1.

The objects introduced above satisfy the following properties:

  1. (1)

    diagram (2) commutes.

    𝒯A{\mathcal{T}_{A}}𝒯A{\mathcal{T}_{A}}Wf{W_{f}}Σ{\Sigma}Σ{\Sigma}σf\scriptstyle{\sigma_{f}}πB\scriptstyle{\pi_{B}}ωf\scriptstyle{\omega_{f}}πB\scriptstyle{\pi_{B}}Ff\scriptstyle{F_{f}}Mi,j\scriptstyle{M_{i,j}}
    Figure 2. Fundamental diagram for Thurston maps satisfying conditions (I) and (II).
  2. (2)

    SFf{0,1,}S_{F_{f}}\subset\{0,1,\infty\} and Ff:WfΣF_{f}\colon W_{f}\to\Sigma is a covering map;

  3. (3)

    πB:𝒯AΣ\pi_{B}\colon\mathcal{T}_{A}\to\Sigma, πC:𝒯AΣ\pi_{C}\colon\mathcal{T}_{A}\to\Sigma, and ωf:𝒯AWf\omega_{f}\colon\mathcal{T}_{A}\to W_{f} are holomorphic covering maps;

  4. (4)

    Mi,jM_{i,j} extends to a Möbius transformation such that {0,1,}\{0,1,\infty\} is an Mi,jM_{i,j}-invariant subset;

  5. (5)

    σf(𝒯A)𝒯A\sigma_{f}(\mathcal{T}_{A})\subsetneq\mathcal{T}_{A} is open and dense in 𝒯A\mathcal{T}_{A} and σf:𝒯Aσf(𝒯A)\sigma_{f}\colon\mathcal{T}_{A}\to\sigma_{f}(\mathcal{T}_{A}) is a holomorphic covering map;

  6. (6)

    the maps FfF_{f} and GfG_{f}, as well as the domain WfW_{f}, depend only on the Hurwitz equivalence class of the Thurston map ff.

Proof.

Let τ=[φ]𝒯A\tau=[\varphi]\in\mathcal{T}_{A} and σf(τ)=[ψ]𝒯A\sigma_{f}(\tau)=[\psi]\in\mathcal{T}_{A}, where the representatives φ,ψ:S2^\varphi,\psi\colon S^{2}\to\widehat{{\mathbb{C}}} are chosen so that φ(ai1)=ψ(aj1)=0\varphi(a_{i_{1}})=\psi(a_{j_{1}})=0, φ(ai2)=ψ(aj2)=1\varphi(a_{i_{2}})=\psi(a_{j_{2}})=1, φ(ai3)=ψ(aj3)=\varphi(a_{i_{3}})=\psi(a_{j_{3}})=\infty, and the map φfψ1:^^\varphi\circ f\circ\psi^{-1}\colon\widehat{{\mathbb{C}}}\dashrightarrow\widehat{{\mathbb{C}}} is holomorphic. In particular, φfψ1\varphi\circ f\circ\psi^{-1} coincides with the map FfF_{f}. Since, ωf([φ])=πC([ψ])=ψ(aj)\omega_{f}([\varphi])=\pi_{C}([\psi])=\psi(a_{j}), we have the following

(Ffωf)([φ])=Ff(ψ(aj))=φ(f(aj))=φ(ai)=πB(φ).(F_{f}\circ\omega_{f})([\varphi])=F_{f}(\psi(a_{j}))=\varphi(f(a_{j}))=\varphi(a_{i})=\pi_{B}({\varphi}).

Thus, πB=Ffωf\pi_{B}=F_{f}\circ\omega_{f}. At the same time, it is straightforward to verify that πB=Mi,jπC\pi_{B}=M_{i,j}\circ\pi_{C}. Finally, Mi,jωf=Mi,jπCσf=πBσfM_{i,j}\circ\omega_{f}=M_{i,j}\circ\pi_{C}\circ\sigma_{f}=\pi_{B}\circ\sigma_{f}, and item (1) follows.

Item (2) directly follows from the definition of a singular set. Maps πB\pi_{B} and πC\pi_{C} are holomorphic coverings, as discussed in Section 2.4. The map ωf:𝒯AWf\omega_{f}\colon\mathcal{T}_{A}\to W_{f} is also a holomorphic covering since both πB:𝒯AΣ\pi_{B}\colon\mathcal{T}_{A}\to\Sigma and Ff:WfΣF_{f}\colon W_{f}\to\Sigma are holomorphic covering maps, and πB=Ffωf\pi_{B}=F_{f}\circ\omega_{f} (see [Hat02, Section 1.3, Excercise 16]). Hence, item (3) follows.

The discussion of Section 2.4 shows that the maps hi,hj:AΣh_{i},h_{j}\colon\mathcal{M}_{A}\to\Sigma are holomorphic. Consequently, Mi,jM_{i,j} is a conformal automorphism of Σ\Sigma that extends to a Möbius transformation of ^\widehat{{\mathbb{C}}} permuting 0, 1, and \infty. Alternative way to prove item (4) is through direct computation. For example, when i=ji=j, then Mi,j=id^M_{i,j}=\operatorname{id}_{\widehat{{\mathbb{C}}}}; if i=1i=1 and j=2j=2, then M1,2(z)=z/(z1)M_{1,2}(z)=z/(z-1), and so on.

Since Mi,jωf=πBσfM_{i,j}\circ\omega_{f}=\pi_{B}\circ\sigma_{f}, where ωf\omega_{f} and πB\pi_{B} are holomorphic covering maps and Mi,jM_{i,j} is a Möbius transformation, it follows that σf:𝒯Aσf(𝒯A)\sigma_{f}\colon\mathcal{T}_{A}\to\sigma_{f}(\mathcal{T}_{A}) is also a holomorphic covering map, where σf(𝒯A)=πB1(Mi,j(Wf))\sigma_{f}(\mathcal{T}_{A})=\pi_{B}^{-1}(M_{i,j}(W_{f})). Note that by Great Picard’s Theorem and the Riemann-Hurwitz formula [Hub06, Appendix A.3], the set ^Wf\widehat{{\mathbb{C}}}-W_{f} contains at least one point different from the points {0,1,}\{0,1,\infty\}. Therefore, σf(𝒯A)\sigma_{f}(\mathcal{T}_{A}) is different from 𝒯A\mathcal{T}_{A}. Since WfW_{f} is open and dense in ^\widehat{{\mathbb{C}}}, it follows that σf(𝒯A)\sigma_{f}(\mathcal{T}_{A}) is open and dense in 𝒯A\mathcal{T}_{A}, establishing item (5).

Finally, let f~\widetilde{f} be a Thurston map Hurwitz equivalent rel. AA to ff. Suppose that ϕ1f~=fϕ2\phi_{1}\circ\widetilde{f}=f\circ\phi_{2}, where ϕ1,ϕ2Homeo0+(S2,A)\phi_{1},\phi_{2}\in\operatorname{Homeo}_{0}^{+}(S^{2},A). Then Ff=φ~f~ψ~1F_{f}=\widetilde{\varphi}\circ\widetilde{f}\circ\widetilde{\psi}^{-1}, where φ~=φϕ1\widetilde{\varphi}=\varphi\circ\phi_{1} and ψ~=ψϕ2\widetilde{\psi}=\psi\circ\phi_{2}. In particular, φ~(ai1)=ψ~(aj1)=0\widetilde{\varphi}(a_{i_{1}})=\widetilde{\psi}(a_{j_{1}})=0, φ~(ai2)=ψ~(aj2)=1\widetilde{\varphi}(a_{i_{2}})=\widetilde{\psi}(a_{j_{2}})=1, and φ~(ai3)=ψ~(aj3)=\widetilde{\varphi}(a_{i_{3}})=\widetilde{\psi}(a_{j_{3}})=\infty. Therefore, FfF_{f} and Ff~F_{\widetilde{f}} coincide, as well as WfW_{f} and Wf~W_{\widetilde{f}}, and item (6) follows. ∎

Remark 4.2.

Proposition 4.1 shows that for Thurston maps satisfying conditions (I) and (II), the corresponding pullback maps admit a commutative diagram analogous to diagram (1) from Section 1.2. In this context, the Möbius transformation Mi,jM_{i,j} serves the role of the XX-map, the map FfF_{f} takes the place of the YY-map, and the domain WfW_{f} is an analog of the Hurwitz space. In particular, σf\sigma_{f} has the “gg-map” Gf=FfMi,j1:Mi,j(Wf)ΣG_{f}=F_{f}\circ M_{i,j}^{-1}\colon M_{i,j}(W_{f})\to\Sigma.

In contrast to the finite degree case, when f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is transcendental, the map Ff:WfΣF_{f}\colon W_{f}\to\Sigma is a covering of infinite degree and WfW_{f} is a countably punctured Riemann sphere.

Remark 4.3.

Suppose that we are in the setting of Proposition 4.1. We have observed that σf(𝒯A)=πB1(Mi,j(Wf))\sigma_{f}(\mathcal{T}_{A})=\pi_{B}^{-1}(M_{i,j}(W_{f})) and the complement of the set Mi,j(Wf)M_{i,j}(W_{f}) in Σ\Sigma contains at least one point. Given that πB:𝒯AΣ\pi_{B}\colon\mathcal{T}_{A}\to\Sigma is a covering map of infinite degree, the pullback map σf\sigma_{f} has infinitely many omitted values, i.e., the points of 𝒯Aσf(𝒯A)\mathcal{T}_{A}-\sigma_{f}(\mathcal{T}_{A}), and they are not compactly contained in 𝒯A\mathcal{T}_{A}. Moreover, the set of omitted values of σf\sigma_{f} is not discrete in the Teichmüller space 𝒯A\mathcal{T}_{A} if the essential singularity of the map ff lies within the set S2AS^{2}-A.

4.1. Characterization problem

In this section, we present and prove a slightly stronger version of Main Theorem A utilizing the tools developed in Sections 3.1 and 3.3 along with the properties of pullback maps established in Proposition 4.1.

We recall that the Teichmüller space 𝒯A\mathcal{T}_{A}, where |A|=4|A|=4, is biholomorphic to the unit disk 𝔻{\mathbb{D}}, and the metric dTd_{T} defined in Section 2.4 coincides with the hyperbolic metric d𝔻d_{{\mathbb{D}}}. If f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is a Thurston map satisfying assumptions (I) and (II), then the the corresponding pullback map σf:𝒯A𝒯A\sigma_{f}\colon\mathcal{T}_{A}\to\mathcal{T}_{A} is holomorphic. It can be established in two ways: either through the general approach outlined in Proposition 2.19 or by more elementary methods as in item (5) of Proposition 4.1. It is worth mentioning that item (5) of Proposition 4.1 and Schwarz-Pick’s lemma imply that σf\sigma_{f} is distance-decreasing on the Teichmüller space 𝒯A\mathcal{T}_{A} (cf. Remark 2.20).

Theorem 4.4.

Let f:(S2,A)f\colon(S^{2},A)\righttoleftarrow be a Thurston map of finite or infinite degree that satisfies properties (I) and (II). Then ff is realized rel. AA if and only if it has no weakly degenerate Levy fixed curve. Moreover,

  1. (1)

    if ff is realized rel. AA by postsingularly finite holomorphic maps g1:(^,P1)g_{1}\colon(\widehat{{\mathbb{C}}},P_{1})\righttoleftarrow and g2:(^,P2)g_{2}\colon(\widehat{{\mathbb{C}}},P_{2})\righttoleftarrow, then g1g_{1} and g2g_{2} are conjugate by a Möbius transformation MM, i.e., Mg1=g2MM\circ g_{1}=g_{2}\circ M, such that M(P1)=P2M(P_{1})=P_{2};

  2. (2)

    if ff is obstructed rel. AA, then it has a unique Levy fixed curve up to homotopy in S2AS^{2}-~{}A.

Proof.

Suppose that the Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is realized. According to Proposition 2.11, f:(S2,A)f\colon(S^{2},A)\righttoleftarrow cannot have a Levy fixed curve. From Proposition 2.17, it follows that σf\sigma_{f} has a fixed point in the Teichmüller space 𝒯A\mathcal{T}_{A}. As it was discussed previously, σf\sigma_{f} is distance-decreasing on 𝒯A\mathcal{T}_{A}, which implies that it has a unique fixed point. Now, it is straightforward to verify using Proposition 2.14 that item (1) holds (cf. [Hub16, Corollary 10.7.8] and [MPR24, Proposition 2.26]).

Now, suppose that the Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is obstructed. Choose an arbitrary point τ0𝒯A\tau_{0}\in\mathcal{T}_{A} and define τn:=σfn(τ0)\tau_{n}:=\sigma_{f}^{\circ n}(\tau_{0}), xn:=πB(τn)x_{n}:=\pi_{B}(\tau_{n}), and yn:=πC(τn+1)y_{n}:=\pi_{C}(\tau_{n+1}) for every n0n\geq 0. Let τn=[φn]𝒯A\tau_{n}=[\varphi_{n}]\in\mathcal{T}_{A}, where the representative φn:S2^\varphi_{n}\colon S^{2}\to\widehat{{\mathbb{C}}} is chosen so that φn(ai1)=0\varphi_{n}(a_{i_{1}})=0, φn(ai2)=1\varphi_{n}(a_{i_{2}})=1, and φn(ai3)=\varphi_{n}(a_{i_{3}})=\infty. Denote by ψn:S2^\psi_{n}\colon S^{2}\to\widehat{{\mathbb{C}}} the unique (due to Proposition 2.14) orientation-preserving homeomorphism so that gn:=φnfψn1:^^g_{n}:=\varphi_{n}\circ f\circ\psi_{n}^{-1}\colon\widehat{{\mathbb{C}}}\dashrightarrow\widehat{{\mathbb{C}}} is holomorphic and ψn(aj1)=0\psi_{n}(a_{j_{1}})=0, ψn(aj2)=1\psi_{n}(a_{j_{2}})=1, and ψn(aj3)=\psi_{n}(a_{j_{3}})=\infty. According to the definition of the map FfF_{f}, it must coincide with the map gng_{n} for every n0n\geq 0. Moreover, it follows that τn+1=[ψn]\tau_{n+1}=[\psi_{n}]. From this, we observe that φn(ai)=xn\varphi_{n}(a_{i})=x_{n}, ψn(aj)=yn\psi_{n}(a_{j})=y_{n}, and Ff(yn)=xnF_{f}(y_{n})=x_{n} for every n0n\geq 0.

According to item (1) of Proposition 4.1, we have πB(σf(𝒯A))Mi,j(Wf)\pi_{B}(\sigma_{f}(\mathcal{T}_{A}))\subset M_{i,j}(W_{f}) and πB=GfπBσf\pi_{B}=G_{f}\circ\pi_{B}\circ\sigma_{f}. Since GfG_{f} is not injective, items (2)-(5) of Proposition 4.1 allow us to apply Theorem 3.10 to the pullback map σf\sigma_{f}. This shows that the sequence (xn)(x_{n}) converges to a repelling fixed point x{0,1,}x\in\{0,1,\infty\} of the map GfG_{f}. Given that xn+1=Mi,j(yn)x_{n+1}=M_{i,j}(y_{n}), the sequence (yn)(y_{n}) converges to a regular point y{0,1,}y\in\{0,1,\infty\} of the map FfF_{f} due to item (4) of Proposition 4.1.

We assume that the map FfF_{f} is extended holomorphically to a neighborhood of yy. Then there exists a disk D^D\subset~{}\widehat{{\mathbb{C}}} such that D{0,1,}={y}D\cap\{0,1,\infty\}=\{y\} and FfF_{f} is injective on DD. Consider another disk DD^{\prime} such that D¯D\overline{D^{\prime}}\subset D and the annulus DD¯D-\overline{D^{\prime}} has modulus greater than 5πed0/5\pi e^{d_{0}}/\ell^{*}, where d0=dT(τ0,τ1)d_{0}=d_{T}(\tau_{0},\tau_{1}) and =log(3+22)\ell^{*}=\log(3+2\sqrt{2}). Observe that ynDy_{n}\in D^{\prime} for all nn large enough and, in particular, each connected component of ^(DD¯)\widehat{{\mathbb{C}}}-(D-\overline{D^{\prime}}) contains two points of the set ψn(A)\psi_{n}(A).

Finally, by Schwarz-Pick’s lemma, we have dT(τn,τn+1)dT(τ0,τ1)d_{T}(\tau_{n},\tau_{n+1})\leq d_{T}(\tau_{0},\tau_{1}) for every n0n\geq 0. The existence of a weakly degenerate Levy fixed curve for the Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow then follows from Proposition 3.1, applied to τn=[φn]\tau_{n}=[\varphi_{n}] and τn+1=[ψn]\tau_{n+1}=[\psi_{n}], where nn is taken sufficiently large, and the annulus DD¯D-\overline{D^{\prime}} as above. The uniqueness part follows from Proposition 3.2 since the sequence (π(σfn(τ0)))(\pi(\sigma_{f}^{\circ n}(\tau_{0}))) clearly leaves every compact set of the moduli space A\mathcal{M}_{A}. ∎

4.2. Hurwitz classes

In Section 4.1, we demonstrated how Proposition 4.1, especially commutative diagram (2), can be helpful for studying Thurston maps that satisfy assumptions (I) and (II). In this section, we further develop this idea by showing the significance of the dynamical properties of the map GfG_{f} in understanding the Hurwitz class f,A{\mathcal{H}}_{f,A} of a Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow that satisfies properties (I) and (II). In particular, we prove Main Theorem B (see Theorem 4.8) and Corollary 1.1. However, before proceeding with their proofs, we present two propositions that relate the fixed points of the map GfG_{f} to the Thurston maps in the Hurwitz class f,A{\mathcal{H}}_{f,A}, which are either obstructed or realized depending on the properties of the corresponding fixed point.

Proposition 4.5.

Let f:(S2,A)f\colon(S^{2},A)\righttoleftarrow be a Thurston map that satisfies conditions (I) and (II). Suppose that xΣx\in\Sigma is a fixed point of the map GfG_{f}. Then there exists a homeomorphism ϕHomeo+(S2,A)\phi\in\operatorname{Homeo}^{+}(S^{2},A) such that the Thurston map f~:=ϕf:(S2,A)\widetilde{f}:=\phi\circ f\colon(S^{2},A)\righttoleftarrow is realized by a holomorphic map g:(^,P)g\colon(\widehat{{\mathbb{C}}},P)\righttoleftarrow, where P={0,1,,x}P=\{0,1,\infty,x\}.

Proof.

Proposition 4.1 suggests that there exist points τ0\tau_{0} and τ1\tau_{1} in the Teichmüller space 𝒯A\mathcal{T}_{A} so that σf(τ0)=τ1\sigma_{f}(\tau_{0})=\tau_{1} and πB(τ0)=πB(τ1)=x\pi_{B}(\tau_{0})=\pi_{B}(\tau_{1})=x. Indeed, choose an arbitrary point τ1\tau_{1} of πB1(x)\pi^{-1}_{B}(x). Since xMi,j(Wf)x\in M_{i,j}(W_{f}), σf1(τ1)\sigma_{f}^{-1}(\tau_{1}) is non-empty, and moreover, πB(σf1(τ1))={x}\pi_{B}(\sigma_{f}^{-1}(\tau_{1}))=\{x\}. Thus, we can take τ0\tau_{0} to be any point of σf1(τ1)\sigma_{f}^{-1}(\tau_{1}).

Let τ0=[φ]\tau_{0}=[\varphi] and τ1=[ψ]\tau_{1}=[\psi], where the representatives φ,ψ:S2^\varphi,\psi\colon S^{2}\to\widehat{{\mathbb{C}}} are chosen such that g:=φfψ1:^^g:=\varphi\circ f\circ\psi^{-1}\colon\widehat{{\mathbb{C}}}\dashrightarrow\widehat{{\mathbb{C}}} is holomorphic, φ(ai1)=ψ(ai1)=0\varphi(a_{i_{1}})=\psi(a_{i_{1}})=0, φ(ai2)=ψ(ai2)=1\varphi(a_{i_{2}})=\psi(a_{i_{2}})=1, φ(ai3)=ψ(ai3)=\varphi(a_{i_{3}})=\psi(a_{i_{3}})=\infty, and φ(ai)=ψ(ai)=x\varphi(a_{i})=\psi(a_{i})=x. It is straightforward to verify that g:(^,P)g\colon(\widehat{{\mathbb{C}}},P)\righttoleftarrow is a postsingularly finite holomorphic map, where P=φ(A)=ψ(A)={0,1,,x}P=\varphi(A)=\psi(A)=\{0,1,\infty,x\}.

Now, define ϕ:=ψ1φHomeo+(S2,A)\phi:=\psi^{-1}\circ\varphi\in\operatorname{Homeo}^{+}(S^{2},A) and f~:=ϕf:(S2,A)\widetilde{f}:=\phi\circ f\colon(S^{2},A)\righttoleftarrow. It is easy to see that ψf~ψ1=g\psi\circ\widetilde{f}\circ\psi^{-1}=g. Therefore, the Thurston map f~:(S2,A)\widetilde{f}\colon(S^{2},A)\righttoleftarrow is combinatorially equivalent to g:(^,P)g\colon(\widehat{{\mathbb{C}}},P)\righttoleftarrow. ∎

Proposition 4.6.

Let f:(S2,A)f\colon(S^{2},A)\righttoleftarrow be a Thurston map that satisfies conditions (I) and (II). If GfG_{f} has a repelling fixed point x{0,1,}x\in\{0,1,\infty\}, then there exists a homeomorphism ϕHomeo+(S2,A)\phi\in\operatorname{Homeo}^{+}(S^{2},A) such that the Thurston map f~:=ϕf:(S2,A)\widetilde{f}:=\phi\circ f\colon(S^{2},A)\righttoleftarrow is obstructed. Moreover, ff is totally unobstructed rel. AA if and only if none of the points 0, 11, or \infty is a repelling fixed point of the map GfG_{f}.

Proof.

Suppose that x{0,1,}x\in\{0,1,\infty\} is a repelling fixed point of the map GfG_{f}. Assume that GfG_{f} is extended holomorphically to a neighborhood xx. Let UMi,j(Wf){x}U\subset M_{i,j}(W_{f})\cup\{x\} be a neighborhood of xx where GfG_{f} is injective and U¯Gf(U)\overline{U}\subset G_{f}(U). Define the local inverse branch g:Gf(U)Ug\colon G_{f}(U)\to U of GfG_{f} at xx, i.e., g:=(Gf|U)1g:=(G_{f}|U)^{-1}. Note that every orbit of gg converges to xx, since gg is uniformly distance-decreasing with respect to the hyperbolic metric on UU according to Schwarz-Pick’s lemma, Proposition 3.5, and Remark 3.6.

Claim.

The distance dΣ(y,g(y))d_{\Sigma}(y,g(y)) converges to 0 as yGf(U)y\in G_{f}(U) tends to xx.

Proof.

Without loss of generality, assume that x=0x=0 and y,g(y)𝔻y,g(y)\in{\mathbb{D}}. Since x=0x=0 is a repelling fixed point of the map GfG_{f}, by choosing yy sufficiently small, we can ensure that λ|y||g(y)||y|\lambda|y|\leq|g(y)|\leq|y| for some λ\lambda, where 0<λ<10<\lambda<1. Let p:𝔻p\colon{\mathbb{H}}\to{\mathbb{D}} be the holomorphic universal covering defined as p(z)=exp(iz)p(z)=\exp(iz) for zz\in{\mathbb{H}}. Define y1:=arg(y)ilog(|y|)p1(y)y_{1}:=\arg(y)-i\log(|y|)\in p^{-1}(y) and y2:=arg(g(y))ilog(|g(y)|)p1(g(y))y_{2}:=\arg(g(y))-i\log(|g(y)|)\in p^{-1}(g(y)). Similarly to the proof of Proposition 3.7, we have

d(y1,y2)log(log|g(y)|log|y|)+2π|log|y||log(log|y|+logλlog|y|)+2π|log|y||.d_{{\mathbb{H}}}(y_{1},y_{2})\leq\log\left(\frac{\log|g(y)|}{\log|y|}\right)+\frac{2\pi}{|\log|y||}\leq\log\left(\frac{\log|y|+\log\lambda}{\log|y|}\right)+\frac{2\pi}{|\log|y||}.

This shows that the distance d(y1,y2)d_{{\mathbb{H}}}(y_{1},y_{2}) converges to 0 as yy approaches xx. According to Schwarz-Pick’s lemma, the same holds for d𝔻(y,g(y))d_{{\mathbb{D}}}(y,g(y)) and dΣ(y,g(y))d_{\Sigma}(y,g(y)). ∎

Therefore, by making UU even smaller, we can assume that dΣ(y,g(y))<dΣ(y,y)d_{\Sigma}(y,g(y))<d_{\Sigma}(y,y^{\prime}), where yGf(U)y\in G_{f}(U) is any point other than xx and yGf1(y)y^{\prime}\in G_{f}^{-1}(y) with yg(y)y\neq g(y).

Now, we choose x1U{x}x_{1}\in U-\{x\} and let x0=Gf(x1)x_{0}=G_{f}(x_{1}). Similarly to the proof of Proposition 4.5, there exists two points τ1\tau_{1} and τ2\tau_{2} of the Teichmüller space 𝒯A\mathcal{T}_{A} so that σf(τ0)=τ1\sigma_{f}(\tau_{0})=\tau_{1}, πB(τ0)=x0\pi_{B}(\tau_{0})=x_{0}, and πB(τ1)=x1\pi_{B}(\tau_{1})=x_{1}. Since πB:𝒯AΣ\pi_{B}\colon\mathcal{T}_{A}\to\Sigma is a holomorphic covering map, then there also exists a point τ𝒯A\tau\in\mathcal{T}_{A} so that dT(τ,τ1)=dΣ(x0,x1)d_{T}(\tau,\tau_{1})=d_{\Sigma}(x_{0},x_{1}) and πB(τ)=x0\pi_{B}(\tau)=x_{0}.

Let τ=[φ]\tau=[\varphi] and τ0=[ψ]\tau_{0}=[\psi], where the representatives φ,ψ:S2^\varphi,\psi\colon S^{2}\to\widehat{{\mathbb{C}}} are chosen so that φ|A=ψ|A\varphi|A=\psi|A. Define the homeomorphism ϕ:=φ1ψHomeo+(S2,A)\phi:=\varphi^{-1}\circ\psi\in\operatorname{Homeo}^{+}(S^{2},A). According to Remark 2.15, σϕ(τ)=τ0\sigma_{\phi}(\tau)=\tau_{0} and σf~(τ)=σf(σϕ(τ))=σf(τ0)=τ1\sigma_{\widetilde{f}}(\tau)=\sigma_{f}(\sigma_{\phi}(\tau))=\sigma_{f}(\tau_{0})=\tau_{1}, where f~:=ϕf:(S2,A)\widetilde{f}:=\phi\circ f\colon(S^{2},A)\righttoleftarrow is a Thurston map.

Schwarz-Pick’s lemma, along with items (3) and (5) of Proposition 4.1, implies that

dΣ(πB(σf~n(τ)),πB(σf~(n+1)(τ)))dT(τ,σf~(τ))=dT(τ,τ1)=dΣ(x0,x1).d_{\Sigma}(\pi_{B}(\sigma_{\widetilde{f}}^{\circ n}(\tau)),\pi_{B}(\sigma_{\widetilde{f}}^{\circ(n+1)}(\tau)))\leq d_{T}(\tau,\sigma_{\widetilde{f}}(\tau))=d_{T}(\tau,\tau_{1})=d_{\Sigma}(x_{0},x_{1}).

At the same time, it follows from items (1) and (6) of Proposition 4.1 that

πB(σf~(n+1)(τ))Gf1(πB(σf~n(τ))).\pi_{B}(\sigma_{\widetilde{f}}^{\circ(n+1)}(\tau))\in G_{f}^{-1}(\pi_{B}(\sigma_{\widetilde{f}}^{\circ n}(\tau))).

Since πB(τ)=x0Gf(U)\pi_{B}(\tau)=x_{0}\in G_{f}(U), and based on the previous assumptions, we have πB(σf~n(τ))=gn(x0)\pi_{B}(\sigma_{\widetilde{f}}^{\circ n}(\tau))=g^{\circ n}(x_{0}). Thus, πB(σf~n(τ))\pi_{B}(\sigma_{\widetilde{f}}^{\circ n}(\tau)) converges to xx as nn\to\infty. Given that σf~\sigma_{\widetilde{f}} is 1-Lipschitz, it cannot have a fixed point. Hence, by Proposition 2.17, the Thurston map f~:(S2,A)\widetilde{f}\colon(S^{2},A)\righttoleftarrow must be obstructed.

Suppose none of points 0, 11, or \infty is a repelling fixed point of the map GfG_{f}. Let f^\widehat{f} be any Thurston map Hurwitz equivalent rel. AA to ff. Then, according to item (6) of Proposition 4.1, we have Wf=Wf^W_{f}=W_{\widehat{f}} and Gf=Gf^G_{f}=G_{\widehat{f}}. Taking into account Proposition 4.1 and applying Theorem 3.10, we see that σf^\sigma_{\widehat{f}} must have a fixed point. Thus, it follows from Proposition 2.17 that f^\widehat{f} is realized rel.  AA. ∎

Remark 4.7.

It is clear that GfG_{f} extends to a postsingularly finite holomorphic map having at most one essential singularity and a postsingular set contained within {0,1,}\{0,1,\infty\}. Therefore, by Lemma A.1, every fixed point of GfG_{f} is either superattracting or repelling. Furthermore, the only possible superattracting fixed points are 0, 11, and \infty.

Now we are ready to state and prove a slightly stronger version of Main Theorem B.

Theorem 4.8.

Let f:(S2,A)f\colon(S^{2},A)\righttoleftarrow be a Thurston map of finite or infinite degree that satisfies conditions (I) and (II). Then

  1. (1)

    ff is totally unobstructed rel. AA if and only if there are no two points a,bAa,b\in A such that deg(f,a)=deg(f,b)=1\deg(f,a)=\deg(f,b)=1 and f({a,b})f(\{a,b\}) equals {a,b}\{a,b\} or A{a,b}A-\{a,b\};

  2. (2)

    if ff is not totally unobstructed rel. AA, then its Hurwitz class f,A{\mathcal{H}}_{f,A} contains infinitely many pairwise combinatorially inequivalent obstructed Thurston maps;

  3. (3)

    if ff has infinite degree, then its Hurwitz class f,A{\mathcal{H}}_{f,A} contains infinitely many pairwise combinatorially inequivalent realized Thurston maps.

Proof.

Without loss of generality, we assume that AA is indexed so that B={a1,a2,a3}B=\{a_{1},a_{2},a_{3}\}, and therefore, i=4i=4 (see the beginning of Section 4). We then analyze four different cases based on the value of j,1j4j,1\leq j\leq 4, to find out when one of the points 0, 11, or \infty is a repelling fixed point of the map GfG_{f} (this analysis will be needed to apply Proposition 4.6). We also recall that deg(f,aj)=1\deg(f,a_{j})=1 and f(aj)=aif(a_{j})=a_{i}.

  • For the case j=1j=1, we have M4,1(z)=1/zM_{4,1}(z)=1/z. Therefore,

    Gf(0)=Ff()=(φfψ1)()=φ(f(a4)).G_{f}(0)=F_{f}(\infty)=(\varphi\circ f\circ\psi^{-1})(\infty)=\varphi(f(a_{4})).

    This means that 0 is a fixed point of GfG_{f} is and only if f(a4)=φ1(0)=a1f(a_{4})=\varphi^{-1}(0)=a_{1}. Furthermore, according to Remark 4.7, 0 is a repelling fixed point of GfG_{f} if and only if f(a4)=a1f(a_{4})=a_{1} and a4a_{4} is a regular point of ff. Similarly, 11 is a repelling fixed point of GfG_{f} if and only if f(a3)=a2f(a_{3})=a_{2} and deg(f,a3)=1\deg(f,a_{3})=1. Lastly, \infty is a repelling fixed point of the map GfG_{f} if and only if f(a2)=a3f(a_{2})=a_{3} and deg(f,a2)=1\deg(f,a_{2})=1.

  • For the case j=2j=2, we have M4,2(z)=(z1)/zM_{4,2}(z)=(z-1)/z. Similarly to the previous case, one of the points 0, 11, and \infty is a repelling fixed point of GfG_{f} if and only if f(a3)=a1f(a_{3})=a_{1} and deg(f,a3)=1\deg(f,a_{3})=1, or f(a4)=a2f(a_{4})=a_{2} and deg(f,a4)=1\deg(f,a_{4})=1, or f(a1)=a3f(a_{1})=a_{3} and deg(f,a1)=1\deg(f,a_{1})=1.

  • For the case j=3j=3, we have M4,3(z)=1zM_{4,3}(z)=1-z. Here, one of the points 0, 11, and \infty is a repelling fixed point of GfG_{f} if and only if f(a2)=a1f(a_{2})=a_{1} and deg(f,a2)=1\deg(f,a_{2})=1, or f(a1)=a2f(a_{1})=a_{2} and deg(f,a1)=1\deg(f,a_{1})=1, or f(a4)=a3f(a_{4})=a_{3} and deg(f,a4)=1\deg(f,a_{4})=1.

  • For the case j=4j=4, we have M4,4=id^M_{4,4}=\operatorname{id}_{\widehat{{\mathbb{C}}}}. In this case, one of the points 0, 11, and \infty is a repelling fixed point of GfG_{f} if and only if f(ak)=akf(a_{k})=a_{k} and deg(f,ak)=1\deg(f,a_{k})=1 for some k=1,2,3k=1,2,3.

Summarizing the calculations above and applying Proposition 4.6, we obtain item (1).

To establish item (2), we largely follow the approach used in the proof of [KPS16, Theorem 9.2(V)]. Suppose that ff is not totally unobstructed rel. AA, i.e., there exists an obstructed Thurston map in f,A{\mathcal{H}}_{f,A}. Without loss of generality, we assume that this map is f:(S2,A)f\colon(S^{2},A)\righttoleftarrow. Theorem 4.4 shows that there exists a Levy fixed curve γ\gamma for f:(S2,A)f\colon(S^{2},A)\righttoleftarrow. Define fn:=Tγnf:(S2,A)f_{n}:=T_{\gamma}^{\circ n}\circ f\colon(S^{2},A)\righttoleftarrow, where nn\in{\mathbb{Z}} and TγHomeo+(S2,A)T_{\gamma}\in\operatorname{Homeo}^{+}(S^{2},A) is the Dehn twist about a curve γ\gamma. Clearly, each Thurston map fnf_{n} has a Levy fixed curve γ\gamma, and therefore, is obstructed rel. AA by Proposition 2.11. We will show that these Thurston maps are pairwise combinatorially inequivalent rel. AA.

Suppose the contrary. Then there exist two homeomorphisms ϕ1,ϕ2Homeo+(S2)\phi_{1},\phi_{2}\in\operatorname{Homeo}^{+}(S^{2}) such that ϕ1(A)=ϕ2(A)=A\phi_{1}(A)=\phi_{2}(A)=A, ϕ1\phi_{1} is isotopic rel. AA to ϕ2\phi_{2}, and fn=ϕ1fmϕ21f_{n}=\phi_{1}\circ f_{m}\circ\phi_{2}^{-1} for some mnm\neq n. Therefore, fn:(S2,A)f_{n}\colon(S^{2},A)\righttoleftarrow has ϕ1(γ)\phi_{1}(\gamma) as a Levy fixed curve. However, Theorem 4.4 states that this Levy fixed curve is unique up to homotopy in S2AS^{2}-A. This implies that γ\gamma and ϕ1(γ)\phi_{1}(\gamma) are homotopic in S2AS^{2}-A, and thus, ϕ1\phi_{1} and ϕ2\phi_{2} are isotopic rel. AA to TγkT_{\gamma}^{\circ k} for some kk\in{\mathbb{Z}}.

One can easily see that ff commutes with the Dehn twist TγT_{\gamma} up to isotopy rel. AA, meaning TγfT_{\gamma}\circ f is isotopic rel. AA to fTγf\circ T_{\gamma}. Indeed, up to isotopy rel. AA, we can assume that ff is the identity on a certain annulus in S2AS^{2}-A with a core curve γ\gamma. Considering the previous discussion, we conclude that ff is isotopic rel. AA to fTγ(mn)f\circ T_{\gamma}^{\circ(m-n)}. The following claim proves that it is not possible, and item (2) follows.

Claim.

Suppose that ff is isotopic rel. AA to fϕf\circ\phi, where ϕHomeo+(S2,A)\phi\in\operatorname{Homeo}^{+}(S^{2},A). Then ϕ\phi is isotopic rel. AA to idS2\operatorname{id}_{S^{2}}.

Proof.

According to Definition 2.5, we can assume without loss of generality that f=fϕf=f\circ~{}\phi. There exist orientation-preserving homeomorphisms φ,ψ:^S2\varphi,\psi\colon\widehat{{\mathbb{C}}}\to S^{2} such that the map g:=φfψ1:^^g:=\varphi\circ f\circ\psi^{-1}\colon\widehat{{\mathbb{C}}}\dashrightarrow\widehat{{\mathbb{C}}} is holomorphic. One can easily check that g=ghg=g\circ h, where h:=ψϕψ1:^^h:=\psi\circ\phi\circ\psi^{-1}\colon\widehat{{\mathbb{C}}}\to\widehat{{\mathbb{C}}}. Since hh must be a Möbius transformation and hh fixes the points of the set ψ(A)\psi(A), it follows that h=id^h=\operatorname{id}_{\widehat{{\mathbb{C}}}}. Thus, ϕ\phi is also the identity map, proving the claim. ∎

Lemma A.2 implies that the map GfG_{f} has infinitely many fixed points when the map ff is transcendental. According to Proposition 4.5, every such fixed point, apart from 0, 1, and \infty, corresponds to a realized Thurston map in f,A{\mathcal{H}}_{f,A}. However, some of these maps might be combinatorially equivalent rel. AA. Nevertheless, we will show that only finitely many of them can be pairwise combinatorially equivalent rel. AA.

Consider two Thurston maps f1f_{1} and f2f_{2} realized rel. AA by postsingularly finite holomorphic maps g1:(^,P1)g_{1}\colon(\widehat{{\mathbb{C}}},P_{1})\righttoleftarrow and g2:(^,P2)g_{2}\colon(\widehat{{\mathbb{C}}},P_{2})\righttoleftarrow, respectively, where P1={0,1,,x1}P_{1}=\{0,1,\infty,x_{1}\} and P2={0,1,,x2}P_{2}=\{0,1,\infty,x_{2}\}, with x1x_{1} and x2x_{2} being distinct fixed points of the map GfG_{f}. If f1f_{1} and f2f_{2} are combinatorially equivalent rel. AA, then f1f_{1} is realized rel. AA by both g1:(^,P1)g_{1}\colon(\widehat{{\mathbb{C}}},P_{1})\righttoleftarrow and g2:(^,P2)g_{2}\colon(\widehat{{\mathbb{C}}},P_{2})\righttoleftarrow. By item (1) of Theorem 4.4, there exists of a Möbius transformation MM such that Mg1=g2MM\circ g_{1}=g_{2}\circ M and M(P1)=P2M(P_{1})=P_{2}. In particular, MM is not the identity map and {0,1,}M({0,1,,x1})\{0,1,\infty\}\subset M(\{0,1,\infty,x_{1}\}). Since a Möbius transformation is uniquely determined by its values at three distinct points of ^\widehat{{\mathbb{C}}}, there can be at most 24 such Thurston maps that are pairwise combinatorially equivalent rel. AA, and item (3) follows ∎

Let us now proceed to prove Corollary 1.1 from Section 1.3.2. First of all, we recall the definition of a parameter space.

Definition 4.9.

Let g:^g\colon{\mathbb{C}}\to\widehat{{\mathbb{C}}} be a non-constant meromorphic map of finite type. Then the parameter space of gg is defined as follows:

Par(g):={φgψ:^ holomorphic for some φHomeo+(^) and ψHomeo+()}.\displaystyle\mathrm{Par}(g):=\{\varphi\circ g\circ\psi\colon{\mathbb{C}}\to\widehat{{\mathbb{C}}}\text{ holomorphic }\text{for some }\varphi\in\operatorname{Homeo}^{+}(\widehat{{\mathbb{C}}})\text{ and }\psi\in\operatorname{Homeo}^{+}({\mathbb{C}})\}.
Proof of Corollary 1.1.

Note that gg should have at least two singular values and, according to Great Picard’s Theorem, at most two exceptional values, i.e., points w^w\in\widehat{{\mathbb{C}}} such that the preimage g1(w)g^{-1}(w) is finite. Clearly, every exceptional value is an asymptotic value of gg. Therefore, by post-composing gg with a Möbius transformation, we can assume that {0,}Sg{0,1,}\{0,\infty\}\subset S_{g}\subset\{0,1,\infty\} and g1(1)g^{-1}(1) is infinite. By pre-composing gg with an affine transformation, we can assume that g(0)=1g(0)=1 and g(1)=1g(1)=1. Let x,y0,1,x,y\neq 0,1,\infty be two points in ^\widehat{{\mathbb{C}}} so that g(x)=yg(x)=y.

Next, choose four distinct point aa, bb, cc, and dd in S2S^{2}, and two orientation-preserving homeomorphisms φ,ψ:S2^\varphi,\psi\colon S^{2}\to\widehat{{\mathbb{C}}} such that:

  • φ(a)=0\varphi(a)=0, φ(b)=1\varphi(b)=1, φ(c)=y\varphi(c)=y, and φ(d)=\varphi(d)=\infty, and

  • ψ(a)=0\psi(a)=0, ψ(b)=x\psi(b)=x, ψ(c)=1\psi(c)=1, and ψ(d)=\psi(d)=\infty.

Then f:=φ1gψ:S2S2f:=\varphi^{-1}\circ g\circ\psi\colon S^{2}\dashrightarrow S^{2} is a topologically holomorphic map. Moreover, {a,d}Sf{a,b,d}\{a,d\}\subset S_{f}\subset\{a,b,d\}, and f(a)=bf(a)=b, f(b)=cf(b)=c, and f(c)=bf(c)=b, while dSfd\in S_{f} is the essential singularity of the map ff. In other words, ff is a Thurston map with the postsingular set Pf={a,b,c,d}P_{f}=\{a,b,c,d\}. Moreover, it is easy to see that ff satisfies conditions (I) and (II). Indeed, by setting B={a,b,d}B=\{a,b,d\}, we have f1(B)¯Pf={a,c,d}\overline{f^{-1}(B)}\cap P_{f}=\{a,c,d\}.

Item (3) of Theorem 4.8 implies that the Hurwitz class f{\mathcal{H}}_{f} contains infinitely many realized Thurston maps that are pairwise combinatorially inequivalent. Clearly, each of these maps is realized by a postsingularly finite map from Par(g)\mathrm{Par}(g). Obviously, these maps must be pairwise (topologically or conformally) non-conjugate, leading to the desired result. ∎

Remark 4.10.

If g:g\colon{\mathbb{C}}\to{\mathbb{C}} is a non-constant entire map of finite type, its entire parameter space is defined by

ParE(g):={φgψ: holomorphic for some φ,ψHomeo+()}.\displaystyle\mathrm{Par}_{E}(g):=\{\varphi\circ g\circ\psi\colon{\mathbb{C}}\to{\mathbb{C}}\text{ holomorphic }\text{for some }\varphi,\psi\in\operatorname{Homeo}^{+}({\mathbb{C}})\}.

Following the proof of Corollary 1.1, one can show that if gg is a transcendental and |Sg|3|S_{g}|\leq~{}3, then ParE(g)\mathrm{Par}_{E}(g) contains infinitely many postsingularly finite entire maps with four postsingular values that are pairwise non-conjugate.

Remark 4.11.

Using the framework of line complexes (see [GO08, Section XI] or [MPR24, Section 2.7]), it can be shown that there are uncountably many distinct parameter spaces. Therefore, Corollary 1.1 implies that conditions (I) and (II) are met by uncountably many pairwise combinatorially inequivalent realized Thurston maps with four postsingular values. Furthermore, by applying item (1) of Theorem 4.4, it is easy to verify that the Thurston maps constructed in the proof Corollary 1.1 are not totally unobstructed if z=1z=1 is a regular point of the map gg. This shows that the family of Thurston maps under consideration also includes uncountably many pairwise combinatorially inequivalent obstructed Thurston maps. According to Remark 4.10, the same observations hold even if we restrict to the class of entire Thurston maps.

4.3. Examples

In this section, we provide examples of several families of Thurston maps that satisfy conditions (I) and (II). We also demonstrate how the framework of Sections 4.1 and 4.2 applies to these rather concrete cases.

Example 4.12 (Exponential maps).

Let f:S2S2f\colon S^{2}\dashrightarrow S^{2} be an entire Thurston map with Pf={a,b,c,d}P_{f}=\{a,b,c,d\} and Sf={a,d}S_{f}=\{a,d\}, where dd is the essential singularity of ff. We recall that Thurston maps of this type are called exponential Thurston maps. It is easy to see that aa must be an omitted value of the map ff. In particular, ff has one of the following two dynamical portraits on the set {a,b,c}\{a,b,c\} as illustrated in Figure 3: either the singular value aa has pre-period 1 and period 2, or it has pre-period 2 and period 1.

a{a}b{b}c{c}
a{a}b{b}c{c}
Figure 3. Possible orbit of the singular value of an exponential Thurston map with four postsingular values.

In both cases, the map ff satisfies properties (I) and (II). In particular, Theorem 4.4 shows that ff is realized if and only if it has no Levy fixed curve, which must, in fact, be degenerate. This, in particular, provides a new proof for the more general result [HSS09, Theorem 2.4] in the case of four postsingular values.

According to item (1) of Theorem 4.8, if the singular value aa has pre-period 2 and period 1, then ff it totally unobstructed. However, if aa has pre-period 1 and period 2, the Thurston map ff is never totally unobstructed since f({b,c})={b,c}f(\{b,c\})=\{b,c\} and deg(f,b)=deg(f,c)=1\deg(f,b)=\deg(f,c)=1 because otherwise either bb or cc would be a singular value of ff. Moreover, in this case the Hurwitz class f{\mathcal{H}}_{f} contains infinitely many pairwise combinatorially inequivalent obstructed Thurston maps by item (2) of Theorem 4.8. In both cases, item (3) of Theorem 4.8 states that the Hurwitz class of ff contains infinitely many pairwise combinatorially inequivalent realized Thurston maps.

We further assume that a=a1a=a_{1}, b=a2b=a_{2}, c=a3c=a_{3}, and d=a4d=a_{4}, and adopt the notation introduced at the beginning of Section 4. Our goal is to derive explicit formula for the map GfG_{f}, analyze its dynamics, and observe the phenomena described in Propositions 4.5 and 4.6, as well as in the proof of Theorem 4.8.

First, we consider the case when the singular value aa has pre-period 1 and period 2. Let B={a,b,d}={a1,a2,a4}B=\{a,b,d\}=\{a_{1},a_{2},a_{4}\}, and then C=f1(B)¯Pf={a,c,d}={a1,a3,a4}C=\overline{f^{-1}(B)}\cap P_{f}=\{a,c,d\}=\{a_{1},a_{3},a_{4}\}. In particular, here i=3i=3 and j=2j=2.

Let us compute the map FfF_{f}. It is evident that FfF_{f} is a transcendental entire function. Moreover, SFf={0,}S_{F_{f}}=\{0,\infty\}. By the classical theory of covering maps, Ff(z)=αexp(λz)F_{f}(z)=\alpha\exp(\lambda z) for some α,λ\alpha,\lambda\in{\mathbb{C}}^{*}. Given that Ff(0)=1F_{f}(0)=1 and Ff(1)=1F_{f}(1)=1, it follows that α=1\alpha=1 and λ=2πik,k\lambda=2\pi ik,k\in{\mathbb{Z}}^{*}, where kk is determined by the Hurwitz equivalence class of ff, as stated in item (6) of Proposition 4.1. At the same time, Wf={l/k:l}W_{f}={\mathbb{C}}-\{l/k:l\in{\mathbb{Z}}\} and M3,2(z)=1/zM_{3,2}(z)=1/z. Therefore, Gf(z)=exp(2πik/z)G_{f}(z)=\exp(2\pi ik/z).

In particular, 0 is the essential singularity of GfG_{f}, 11 is a fixed repelling fixed point of GfG_{f} of multiplier 2πik2\pi ik, and Gf()=1G_{f}(\infty)=1. Moreover, by Lemma A.2, the map GfG_{f} has infinitely many repelling fixed points. Thus, Propositions 4.5 and 4.6 already imply that the Hurwitz class f{\mathcal{H}}_{f} contains both realized and obstructed Thurston maps.

Now, let the singular value aa has pre-period 2 and period 1. Let B={a,c,d}={a1,a3,a4}B=\{a,c,d\}=\{a_{1},a_{3},a_{4}\} and then C={b,c,d}={a2,a3,a4}C=\{b,c,d\}=\{a_{2},a_{3},a_{4}\}. Here, i=2i=2 and j=1j=1. Similarly to the previous case, we find that Ff(z)=exp(2πikz)F_{f}(z)=\exp(2\pi ikz) with kk\in{\mathbb{Z}}^{*}, Wf={l/k:l}W_{f}={\mathbb{C}}-\{l/k:l\in{\mathbb{Z}}\}, M2,1(z)=z/(z1)M_{2,1}(z)=z/(z-1), and Gf(z)=exp(2πik/(z1))G_{f}(z)=\exp(2\pi ik/(z-1)). In particular, 11 is the essential singularity of GfG_{f}, and Gf(0)=Gf()=1G_{f}(0)=G_{f}(\infty)=1. Thus, none of the points 0, 11, or \infty is a fixed point of the map GfG_{f}. Therefore, Proposition 4.6 implies that ff is indeed totally unobstructed.

Example 4.13 (Entire maps with three singular values).

Let f:S2S2f\colon S^{2}\dashrightarrow S^{2} be an entire Thurston map with the postsingular set Pf={a,b,c,d}P_{f}=\{a,b,c,d\}, where Sf={a,b,d}S_{f}=\{a,b,d\} and dd is the essential singularity of ff. If ff satisfies condition (II), then it should have one of the three (up to relabeling) possible dynamical portraits on the set {a,b,c}\{a,b,c\} as illustrated in Figure 4. Additionally, there are four more dynamical portraits when condition (II) is not satisfied.

a{a}b{b}c{c}
a{a}b{b}c{c}
a{a}b{b}c{c}
Figure 4. Possible dynamical portraits for an entire Thurston map with three singular and four postsingular values that satisfies condition (II).

Theorem 4.4 states that a Thurston map with one of the dynamical portraits as in Figure 4 is realized if and only if it has no weakly degenerate Levy fixed curve. Furthermore, according to item (1) of Theorem 4.8, ff is totally unobstructed for the first two dynamical portraits (from left to right) in Figure 4 if and only if deg(f,c)=1\deg(f,c)=1, and ff is always totally unobstructed for the third dynamical portrait.

Let a=a1a=a_{1}, b=a2b=a_{2}, c=a3c=a_{3}, and d=a4d=a_{4}. Then we take B=Sf={a,b,d}={a1,a2,a4}B=S_{f}=\{a,b,d\}=\{a_{1},a_{2},a_{4}\}, and then C={a,c,d}={a1,a3,a4}C=\{a,c,d\}=\{a_{1},a_{3},a_{4}\}. In particular, i=3i=3 and j=2j=2. It can can be verified that FfF_{f} is an entire function with SFf={0,1,}S_{F_{f}}=\{0,1,\infty\}. At the same time, M3,2(z)=1/zM_{3,2}(z)=1/z, and therefore Gf(z)=Ff(1/z)G_{f}(z)=F_{f}(1/z). In particular, 0 is the essential singularity of GfG_{f}. Thus, we have the following behavior of the map GfG_{f} on the “cusps” 0 and 1 of the moduli space ΣA\Sigma\sim\mathcal{M}_{A}:

  • for the first dynamical portrait, Gf()=0G_{f}(\infty)=0 and 11 is a fixed point of GfG_{f} that, according to Lemma A.1, is repelling if deg(f,c)=1\deg(f,c)=1, or superattracting otherwise;

  • for the second dynamical portrait, Gf()=1G_{f}(\infty)=1 and 11 is a fixed point of GfG_{f} that is repelling if and only if deg(f,c)=1\deg(f,c)=1, or superattracting otherwise;

  • for the third dynamical portrait, Gf(1)=0G_{f}(1)=0 and Gf()=1G_{f}(\infty)=1. In particular, neither of 0, 11, and \infty is a fixed point of the map GfG_{f}.

Example 4.14 (Non-entire examples).

Most of the observations in Example 4.13 do not depend on the condition f1(d)=f^{-1}(d)=\emptyset, i.e., that the Thurston map ff is entire. Furthermore, there are more non-entire examples of Thurston maps that satisfy conditions (I) and (II). For instance, if f:S2S2f\colon S^{2}\dashrightarrow S^{2} is a Thurston map with |Sf|3|S_{f}|\leq 3, where the postsingular set Pf={a,b,c,d}P_{f}=\{a,b,c,d\} does not contain an essential singularity (e.g., ff could be a finite degree map), and ff has one of the postsingular portraits shown in Figure 5. In particular, Theorem 4.4 provides a novel proof of celebrated Thurston’s characterization theorem [DH93, Theorem 1] for a specific family of finite degree Thurston maps with four postcritical values.

a{a}b{b}d{d}c{c}
a{a}b{b}c{c}d{d}
a{a}b{b}d{d}c{c}
Figure 5. Examples of postsingular portraits of non-entire Thurston maps that satisfy conditions (I) and (II).

Of course, there are more examples, e.g., a Thurston map ff with Pf={a,b,c,d}P_{f}=\{a,b,c,d\} satisfies condition (II) if Sf={a,b,c}S_{f}=\{a,b,c\}, dd is the essential singularity of the map ff, f(a)=df(a)=d, and bb and cc form a 2-cycle for the map ff.

Example 4.15 (Maps with three postsingular values).

Suppose that f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is a Thurston map such that |A|=4|A|=4 and |Pf|3|P_{f}|\leq 3. It is easy to see that if there exists a marked point aAPfa\in A-P_{f} that is not periodic (i.e., it is either pre-periodic or lands to the essential singularity of ff under the iteration), then the pullback map σf\sigma_{f} is constant by Proposition 2.16 and the Thurston map f:(S2,A)f\colon(S^{2},A)\righttoleftarrow is realized, as noted in Remark 2.18.

On the other hand, if every marked point aAPfa\in A-P_{f} is periodic, then conditions (I) and (II) are clearly satisfied. For instance, if |Pf|=3|P_{f}|=3, we can simply take B=PfB=P_{f}.

Appendix A Few facts about dynamics of meromorphic maps

We require the following two results regarding the dynamics of meromorphic maps. Although these results are mostly folklore, we provide short proofs for the completeness.

Lemma A.1.

Every postsingularly finite meromorphic function has only finitely many superattracting periodic orbits, and all other periodic orbits are repelling.

Proof.

Let g:^g\colon{\mathbb{C}}\to\widehat{{\mathbb{C}}} be a postsingularly finite meromorphic function. It is evident that gg can have only a finite number of superattracting periodic orbits, as the points of each such orbit belong to the postsingular set of gg.

Now, consider a periodic point zz\in{\mathbb{C}} of gg. If zz is attracting (but not superattracting), then according to [Ber93, Theorem 7], the corresponding cycle of immediate attracting basins contains a singular value wSgw\in S_{g} that has an infinite orbit, leading to a contradiction. Therefore, if zz is in the Fatou set of gg, then it is the center of a cycle of Siegel disks U1,U2,,UkU_{1},U_{2},\dots,U_{k}, and postsingular values of gg are dense in Ui\partial U_{i} for each i=1,2,,ki=1,2,\dots,k [Ber93, Theorem 7]. This implies that the postsingular set of gg would be infinite. If zz is in the Julia set of gg, then it is either a Cremer periodic point or it lies on the boundary of a cycle of parabolic basins [Mil06, Theorem 7.2]. In both cases, zz is an accumulation point of the postsingular values of gg (see [Ber93, Theorem 7] and [Eps93, Proposition 16]; see also [Eps93, Lemma 72]). Thus, zz is either a superattracting or repelling periodic point of the function gg. ∎

The next result that we require states that a transcendental meromorphic function of finite type has infinitely many repelling fixed points. This was established in the more general context of finite type maps in [Eps93, Proposition 14]. Furthermore, in the paper [Ben16], it was shown that the same is true for transcendental meromorphic functions of bounded type (i.e., having bounded singular set) under the assumption the \infty is a logarithmic singularity of the considered function. In the following lemma, we show that this assumption can be removed and, in fact, the result holds for an arbitrary transcendental meromorphic function of bounded type.

Lemma A.2.

Every transcendental meromorphic function of bounded type has infinitely many repelling fixed points.

Proof.

Let g:^g\colon{\mathbb{C}}\to\widehat{{\mathbb{C}}} be a transcendental meromorphic function of bounded type and D^D\subset\widehat{{\mathbb{C}}} be an open Jordan region containing \infty such that SgD¯={}D¯S_{g}\cap\overline{D}=\{\infty\}\cap\overline{D}. If f1(D)f^{-1}(D) has a connected component that is unbounded (in {\mathbb{C}}), then the result follows directly from [Ben16]. Now, suppose that every connected component of f1(D)f^{-1}(D) is bounded. In this case, f1(D)f^{-1}(D) has infinitely many connected components, and all but finitely many of them are compactly contained in DD. Let UU be one such component, i.e., UU is a connected component of g1(D)g^{-1}(D) such that U¯D\overline{U}\subset D. Let zz be a unique pole of gg in UU and let d:=deg(g,z)d:=\deg(g,z). Note that UU is an open Jordan region and g|U{z}:U{z}D{}g|U-\{z\}\colon U-\{z\}\to D-\{\infty\} is covering map of degree dd.

Consider a Jordan arc αD¯\alpha\subset\overline{D} connecting \infty with a point on D\partial D, with the conditions that |αD|=1|\alpha\cap\partial D|=1 and αU¯=\alpha\cap\overline{U}=\emptyset. Then g1(α)g^{-1}(\alpha) subdivides UU in dd simply connected domains U1,U2,,UdU_{1},U_{2},\dots,U_{d}. Furthermore, the restriction g|Ui:UiDαg|U_{i}\colon U_{i}\to D-\alpha is a biholomorphism. Proposition 3.5 and Remark 3.6 imply that the inverse (g|Ui)1(g|U_{i})^{-1} is uniformly distance-decreasing with respect to the hyperbolic metric on DαD-\alpha, because Ui=g(Dα)U_{i}=g(D-\alpha) is compactly contained in DαD-\alpha. Therefore, by the Banach fixed point theorem, gg has a fixed point in each UiU_{i} for i=1,2,,di=1,2,\dots,d. These fixed points are attracting for (g|Ui)1(g|U_{i})^{-1} and thus repelling for the map gg. By applying the same argument to every connected component of g1(D)g^{-1}(D) that is compactly contained in DD, we conclude that the map gg has infinitely many repelling fixed points. ∎

References

  • [Aba23] Marco Abate. Holomorphic dynamics on hyperbolic Riemann surfaces, volume 89 of De Gruyter Studies in Mathematics. De Gruyter, Berlin, [2023] ©2023.
  • [ABF21] Matthieu Astorg, Anna Miriam Benini, and Núria Fagella. Bifurcation loci of families of finite type meromorphic maps. Preprint arXiv:2107.02663, 2021.
  • [Ast] Matthieu Astorg. Bifurcations and wandering domains in holomorphic dynamics. Habilitation à Diriger des Recherches, Université d’Orléans, 2024. Available at https://www.idpoisson.fr/astorg/papiers/hdrastorg.pdf.
  • [BCT14] Xavier Buff, Guizhen Cui, and Lei Tan. Teichmüller spaces and holomorphic dynamics. In Handbook of Teichmüller theory. Vol. IV, volume 19 of IRMA Lect. Math. Theor. Phys., pages 717–756. Eur. Math. Soc., Zürich, 2014.
  • [BDP24] Laurent Bartholdi, Dzmitry Dudko, and Kevin M. Pilgrim. Correspondences on Riemann surfaces and non-uniform hyperbolicity. Preprint arXiv: 2407.15548, 2024.
  • [BEKP09] Xavier Buff, Adam Epstein, Sarah Koch, and Kevin Pilgrim. On Thurston’s pullback map. In Complex dynamics, pages 561–583. A K Peters, Wellesley, MA, 2009.
  • [Ben16] Anna Miriam Benini. A note on repelling periodic points for meromorphic functions with a bounded set of singular values. Rev. Mat. Iberoam., 32(1):267–274, 2016.
  • [Ber93] Walter Bergweiler. Iteration of meromorphic functions. Bull. Amer. Math. Soc. (N.S.), 29(2):151–188, 1993.
  • [Bis15a] Christopher J. Bishop. Constructing entire functions by quasiconformal folding. Acta Math., 214(1):1–60, 2015.
  • [Bis15b] Christopher J. Bishop. The order conjecture fails in 𝒮\mathcal{S}. J. Anal. Math., 127:283–302, 2015.
  • [Bis17] Christopher J. Bishop. Models for the Speiser class. Proc. Lond. Math. Soc. (3), 114(5):765–797, 2017.
  • [BLMW22] J. Belk, J. Lanier, D. Margalit, and R.R. Winarski. Recognizing topological polynomials by lifting trees. Duke Math. J., 171(17):3401–3480, 2022.
  • [BN06] L. Bartholdi and V. Nekrashevych. Thurston equivalence of topological polynomials. Acta Math., 197(1):1–51, 2006.
  • [BR20] Anna Miriam Benini and Lasse Rempe. A landing theorem for entire functions with bounded post-singular sets. Geom. Funct. Anal., 30(6):1465–1530, 2020.
  • [Bus10] P. Buser. Geometry and spectra of compact Riemann surfaces. Birkhäuser, Boston, MA, 2010.
  • [DH93] A. Douady and J.H. Hubbard. A proof of Thurston’s topological characterization of rational functions. Acta Mathematica, 171(2):263–297, 1993.
  • [DMRS19] K. Drach, Y. Mikulich, J. Rückert, and D. Schleicher. A combinatorial classification of postcritically fixed Newton maps. Ergod. Theory Dyn. Syst., 39(11):2983–3014, 2019.
  • [Eps93] A.L. Epstein. Towers of finite type complex analytic maps. PhD thesis, The City University of New York, 1993. Available at http://pcwww.liv.ac.uk/~lrempe/adam/thesis.pdf.
  • [Ere04] A. Eremenko. Geometric theory of meromorphic functions. In In the tradition of Ahlfors and Bers, III, volume 355 of Contemp. Math., pages 221–230. Amer. Math. Soc., Providence, RI, 2004.
  • [ERG15] Adam Epstein and Lasse Rempe-Gillen. On invariance of order and the area property for finite-type entire functions. Ann. Acad. Sci. Fenn. Math., 40(2):573–599, 2015.
  • [FKK+17] W. Floyd, G. Kelsey, S. Koch, R. Lodge, W. Parry, K.M. Pilgrim, and E. Saenz. Origami, affine maps, and complex dynamics. Arnold Math. J., 3(3):365–395, 2017.
  • [FM12] B. Farb and D. Margalit. A Primer on Mapping Class Groups. Princeton Math. Ser. 49. Princeton Univ. Press, Princeton, NJ, 2012.
  • [For91] Otto Forster. Lectures on Riemann surfaces, volume 81 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1991. Translated from the 1977 German original by Bruce Gilligan, Reprint of the 1981 English translation.
  • [Gam01] Theodore W. Gamelin. Complex analysis. Undergraduate Texts in Mathematics. Springer-Verlag, New York, 2001.
  • [GO08] Anatoly A. Goldberg and Iossif V. Ostrovskii. Value distribution of meromorphic functions, volume 236 of Translations of Mathematical Monographs. American Mathematical Society, Providence, RI, 2008. Translated from the 1970 Russian original by Mikhail Ostrovskii, With an appendix by Alexandre Eremenko and James K. Langley.
  • [Hat02] Allen Hatcher. Algebraic topology. Cambridge University Press, Cambridge, 2002.
  • [Hlu19] M. Hlushchanka. Tischler graphs of critically fixed rational maps and their applications. Preprint arXiv:1904.04759, 2019.
  • [HP22] M. Hlushchanka and N. Prochorov. Critically fixed Thurston maps: classification, recognition, and twisting. Preprint arXiv:2212.14759, 2022.
  • [HSS09] John Hubbard, Dierk Schleicher, and Mitsuhiro Shishikura. Exponential Thurston maps and limits of quadratic differentials. J. Amer. Math. Soc., 22(1):77–117, 2009.
  • [Hub06] John Hamal Hubbard. Teichmüller theory and applications to geometry, topology, and dynamics. Vol. 1. Matrix Editions, Ithaca, NY, 2006. Teichmüller theory, With contributions by Adrien Douady, William Dunbar, Roland Roeder, Sylvain Bonnot, David Brown, Allen Hatcher, Chris Hruska and Sudeb Mitra, With forewords by William Thurston and Clifford Earle.
  • [Hub16] J.H. Hubbard. Teichmüller theory and applications to geometry, topology, and dynamics, Volume 2: Surface homeomorphisms and rational functions. Matrix Editions, Ithaca, NY, 2016.
  • [Koc13] Sarah Koch. Teichmüller theory and critically finite endomorphisms. Advances in Mathematics, 248:573–617, 2013.
  • [KPS16] Sarah Koch, Kevin M. Pilgrim, and Nikita Selinger. Pullback invariants of Thurston maps. Trans. Amer. Math. Soc., 368(7):4621–4655, 2016.
  • [LMS22] R. Lodge, Y. Mikulich, and D. Schleicher. A classification of postcritically finite Newton maps. In In the tradition of Thurston II. Geometry and groups, pages 421–448. Springer, Cham, 2022.
  • [Lod13] R. Lodge. Boundary values of the Thurston pullback map. Conform. Geom. Dyn., 17:77–118, 2013.
  • [LP20] Rami Luisto and Pekka Pankka. Stoïlow’s theorem revisited. Expo. Math., 38(3):303–318, 2020.
  • [LSV08] Bastian Laubner, Dierk Schleicher, and Vlad Vicol. A combinatorial classification of postsingularly finite complex exponential maps. Discrete Contin. Dyn. Syst., 22(3):663–682, 2008.
  • [McM94] C.T. McMullen. Complex dynamics and renormalization. Annals Math. Studies 135. Princeton Univ. Press, Princeton, NJ, 1994.
  • [Mil06] John Milnor. Dynamics in one complex variable, volume 160 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, third edition, 2006.
  • [MPR24] Malavika Mukundan, Nikolai Prochorov, and Bernhard Reinke. Dynamical approximations of postsingularly finite entire maps. Preprint arXiv:2305.17793, 2024.
  • [Par23] Insung Park. Levy and Thurston obstructions of finite subdivision rules. Ergodic Theory and Dynamical Systems, pages 1–51, 12 2023.
  • [Pfr19] David Pfrang. Homotopy Hubbard Trees for Post-Singularly Finite Transcendental Entire Maps. PhD thesis, Jacobs University, Bremen, 2019. Available at https://d-nb.info/1205544992/34.
  • [Poi93] A. Poirier. On post-critically finite polynomials. PhD thesis, State University at New York at Stony Brook, 1993. Available at https://www.math.stonybrook.edu/alumni/1993-Poirier-Alfredo.pdf.
  • [Poi10] A. Poirier. Hubbard trees. Fundamenta Mathematicae, 208(3):193–248, 2010.
  • [PPS21] David Pfrang, Sören Petrat, and Dierk Schleicher. Dreadlock pairs and dynamic partitions for post-singularly finite entire functions. Preprint arXiv: 2109.06863, 2021.
  • [PRS21] D. Pfrang, M. Rothgang, and D. Schleicher. Homotopy Hubbard trees for post-singularly finite exponential maps. Preprint arXiv:1812.11831, 2021.
  • [Rem09] Lasse Rempe. Rigidity of escaping dynamics for transcendental entire functions. Acta Math., 203(2):235–267, 2009.
  • [Sel12] Nikita Selinger. Thurston’s pullback map on the augmented Teichmüller space and applications. Invent. Math., 189(1):111–142, 2012.
  • [Sel13] Nikita Selinger. Topological characterization of canonical Thurston obstructions. J. Mod. Dyn., 7(1):99–117, 2013.
  • [She22] Sergey Shemyakov. A topological characterization of certain postsingularly finite entire functions: transcendental dynamics and Thurston theory. PhD thesis, Université d’Aix-Marseille, 2022. Available at https://www.theses.fr/2022AIXM0017.
  • [Smi24a] Z. Smith. Thurston theory and polymorphic maps. PhD thesis, University of California, 2024.
  • [Smi24b] Zachary Smith. Curve attractors for marked rational maps. Preprint arXiv:2401.16636, 2024.