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

Levy and Thurston obstructions of finite subdivision rules

Insung Park Department of Mathematics, Indiana University, Bloomington, IN 47405, USA park433@iu.edu
Abstract.

For a post-critically finite branched covering of the sphere that is a subdivision map of a finite subdivision rule, we define non-expanding spines which determine the existence of a Levy cycle in a non-exhaustive semi-decidable algorithm. Especially when a finite subdivision rule has polynomial growth of edge subdivisions, the algorithm terminates very quickly, and the existence of a Levy cycle is equivalent to the existence of a Thurston obstruction. In order to show the equivalence between Levy and Thurston obstructions, we generalize the arcs intersecting obstruction theorem by Pilgrim and Tan to a graph intersecting obstruction theorem. As a corollary, we prove that for a pair of post-critically finite polynomials, if at least one polynomial has core entropy zero, then their mating has a Levy cycle if and only if the mating has a Thurston obstruction.

1. Introduction

Obstructions for topological objects to have geometric structures are important subjects of study in topology and geometry. For example, the Geometrization Theorem is about topological obstructions for a 3-manifold to have one out of eight geometries. W. Thurston, who conjectured and proved a large part of the Geometrization Theorem, also proved a geometrization theorem, named Thurston’s characterization, in complex dynamics. He found obstructions, called Thurston obstructions, for a post-critically finite branched covering of the 22-sphere to be isotopic to a rational map [DH93]. Levy cycles were introduced at first as simple cases of Thurston obstructions in the study of the mating problem [Lev85, Tan92]. Recently, it turned out that a Levy cycle itself is an obstruction for a post-critically finite branched covering to be isotopic to an expanding dynamical system [BD18]. Therefore it is important to determine the existence of a Levy cycle as well as a Thurston obstruction for post-critically finite branched coverings. In this paper, we investigate a new method to detect the existence of a Levy cycle for a broad family of branched coverings, called subdivision maps of finite subdivision rules.

Obstructions of post-critically finite topological branched self-coverings of the sphere. A continuous map f:S2S2f:S^{2}\rightarrow S^{2} is a topological branched covering if it locally looks like zzdz\mapsto z^{d} for some integer d>0d>0. A point xS2x\in S^{2} is a critical point if ff is not locally injective at xx. The collection of the critical points Ωf\Omega_{f} is the critical set of ff and its forward orbit Pf:=k=1fk(Ωf)P_{f}:=\bigcup_{k=1}^{\infty}f^{\circ k}(\Omega_{f}) is the post-critical set. If PfP_{f} is finite, ff is a post-critically finite branched covering, or simply a Thurston map. A marked post-critically finite branched covering, is a map f:(S2,A)f:(S^{2},A)\righttoleftarrow such that APfA\supset P_{f}, |A|<|A|<\infty, and f(A)Af(A)\subset A. Every element aAa\in A is called a marked point and AA is called the set of marked points of f:(S2,A)f:(S^{2},A)\righttoleftarrow. Since f:(S2,A)f:(S^{2},A)\righttoleftarrow contains the information of being post-critically finite and the set of marked points, we often abbreviate it just as a branched covering and write more words when they are necessary.

Two branched coverings f:(S2,A)f:(S^{2},A)\righttoleftarrow and g:(S2,B)g:(S^{2},B)\righttoleftarrow are combinatorially equivalent ( by ϕ0\phi_{0} and ϕ1\phi_{1}) if there exist homeomorphisms ϕ0,ϕ1:(S2,A)(S2,B)\phi_{0},\phi_{1}:(S^{2},A)\rightarrow(S^{2},B) such that (1) ϕ0(A)=ϕ1(A)=B\phi_{0}(A)=\phi_{1}(A)=B, (2) ϕ1\phi_{1} is homotopic relative to AA to ϕ0\phi_{0}, and (3) the following diagram commutes.

(S2,Pf){(S^{2},P_{f})}(S2,Pg){(S^{2},P_{g})}(S2,Pf){(S^{2},P_{f})}(S2,Pg){(S^{2},P_{g})}ϕ0\scriptstyle{\phi_{0}}f\scriptstyle{f}g\scriptstyle{g}ϕ1\scriptstyle{\phi_{1}}

A post-critically finite topological branched covering which is not doubly covered by a torus endomorphism is combinatorially equivalent to a post-critically finite rational map if and only if it does not have a Thurston obstruction [DH93], see Section 7.1.

Definition 1.1 (Levy cycle).

A Levy cycle, or a Levy obstruction, of a post-critically finite branched covering f:(S2,A)f:(S^{2},A)\righttoleftarrow is a collection of simple closed curves {γ1,γ2,,γn}\{\gamma_{1},\gamma_{2},\dots,\gamma_{n}\} that are essential relative to AA with the following property: For each 1in1\leq i\leq n there is a connected component γi\gamma_{i}^{\prime} of f1(γi)f^{-1}(\gamma_{i}) which is isotopic to γi+1\gamma_{i+1} relative to AA, and f|γi:γiγif|_{\gamma_{i}^{\prime}}:\gamma_{i}^{\prime}\to\gamma_{i} is a homeomorphism.

Since a Levy cycle is a homeomorphically periodic cycle, a branched covering cannot be expanding along a Levy cycle. Schwartz lemma implies that every post-critically finite rational map g:(^,Pg)g:(\hat{\mathbb{C}},P_{g})\righttoleftarrow is expanding with respect to the conformal metric on ^Pg\hat{\mathbb{C}}\setminus P_{g}, except for a few cases. Therefore, Levy cycle is an example of Thurston obstruction. Shishikura and Tan found an example of mating of cubic polynomials that has a Thurston obstruction but does not have a Levy cycle [ST00]. Although Shishikura and Tan’s example is not conjugate to a rational map, it has an expanding metric, and many objects in the study of rational maps, such as Julia sets, are still well defined. These branched coverings are called Böttcher expanding maps, see [BD18] for Böttcher expanding maps. Rational maps are Böttcher expanding maps by Schwartz lemma. Recently, it was shown that a post-critically finite topological branched covering which is not doubly covered by a torus endomorphism is combinatorially equivalent to a Böttcher expanding map if and only if it does not have a Levy cycle [BD18]. Therefore, Thurston and Levy obstructions can be viewed as obstructions for conformal structures and expanding dynamics on branched coverings of the sphere, respectively.

Analogy with surface diffeomorphisms. There are analogues between surface diffeomorphisms and branched coverings of the sphere. Pseudo-Anosov maps are geometric in a sense that they are affine maps expanding along one dimension and contracting along the other one dimension with respect to appropriate conned Euclidean structures; rational maps are conformal geometric and Böttcher expanding maps are metric geometrically defined. In a pseudo-Anosov mapping class, there is a unique pseudo-Anosov map up to conjugation; in an isotopy class of post-critically finite topological branched coverings, a rational map or a Böttcher expanding map is unique up to conjugation if exists [DH93, BD18]. For non-periodic mapping classes, reducing multicurves are obstructions to pseudo-Anosov mapping classes; Thurston obstructions and Levy cycles are also multicurves, which are obstructions to being isotopic to rational maps and Böttcher expanding maps respectively.

In spite of this analogy, however, algorithms to determine the existence of obstructions for branched coverings of the sphere are relatively less studied compared with surface diffeomorphisms. Let us review some results on algorithms about branched coverings of the sphere. Exhaustive searches for Levy cycles or Thurston obstructions are decidable [BBY12], [BD18]. For topological polynomials, a non-exhaustive algorithm, that finds either Levy cycles if exist or Hubbard trees otherwise, was developed in [BLMW19]. D. Thurston’s positive characterization also gives a non-exhaustive algorithm to detect both Levy cycles and Thurston obstructions for hyperbolic post-critically finite branched coverings [Thu20]. Although these algorithms work efficiently for many examples in practice, no theoretical upper bound of the complexity is known for any of these algorithms. An upper bound for the computational complexity was studied for nearly Euclidean Thurston maps in [FPP18a]. Poirier proved that an abstract Hubbard tree HH is a Hubbard tree of a polynomial if and only if HH is expanding [Poi10]. This gives an efficient algorithm to check whether a Thurston obstruction (equivalently a Levy cycle in this case) exists, and one can easily find an upper bound for the complexity of this algorithm, though it is not stated in [Poi10]. In this paper, Theorem 6.21 provides a new non-exhaustive algorithm to detect Levy cycles when post-critically finite branched coverings are given as subdivision maps of finite subdivision rules. When edges have polynomial growth of subdivisions, Theorem 8.6 implies that this algorithm terminates very quickly, and the complexity is polynomial about the number of cells. But we do not compute the complexity in this paper.

Refer to caption
SS_{\mathcal{R}}
Refer to caption
(S)\mathcal{R}(S_{\mathcal{R}})
Refer to caption
2(S)\mathcal{R}^{2}(S_{\mathcal{R}})
Refer to caption
SS_{\mathcal{R}} with Julia set
Refer to caption
(S)\mathcal{R}(S_{\mathcal{R}}) with Jula set
Refer to caption
2(S)\mathcal{R}^{2}(S_{\mathcal{R}}) with Julia set
Figure 1. A finite subdivision rule of zz21z2+1z\mapsto\frac{z^{2}-1}{z^{2}+1}. The sphere is decomposed into two triangles in SS_{\mathcal{R}}. Each triangle subdivides into two triangles under the subdivision \mathcal{R}. The subdivision map f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\to S_{\mathcal{R}} sends each shaded or unshaded triangle in (S)\mathcal{R}(S_{\mathcal{R}}) to the shaded or unshaded triangle in SS_{\mathcal{R}} respectively.

Finite subdivision rules. A finite subdivision rule \mathcal{R} consists of a partition SS_{\mathcal{R}} of S2S^{2} into polygons and its subdivision (S)\mathcal{R}(S_{\mathcal{R}}) such that a subdivision map f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\to S_{\mathcal{R}} is homeomorpic on each open cell, see Figure 1 for an example and Section 4 for a precise definition. One can also see a finite subdivision rule as a sort of Markov partition. Because PfVert(S)P_{f}\subset\operatorname{Vert}(S_{\mathcal{R}}), the subdivision map is a post-critically finite topological branched covering. By iterating subdivisions, we have a further subdivision n(S)\mathcal{R}^{n}(S_{\mathcal{R}}) and an iterated map fn:n(S)Sf^{n}:\mathcal{R}^{n}(S_{\mathcal{R}})\to S_{\mathcal{R}} for each nn\in\mathbb{N}. It is an open question to determine which topological post-critically finite branched coverings are isotopic to subdivision maps of finite subdivision rules. See Section 4.1 for a list of topological branched coverings that can be represented as subdivision maps.

To detect a Levy cycle, for each n0n\geq 0 we define a level-nn non-expanding spine NnN^{n} which is a graph with a train-track structure encoding non-expanding parts of n(S)\mathcal{R}^{n}(S_{\mathcal{R}}), see Section 6. A finite set AVert(S)A\subset\operatorname{Vert}(S_{\mathcal{R}}) is called a set of marked points of \mathcal{R} if Pff(A)AP_{f}\cup f(A)\subset A. A point aAa\in A is called a Fatou point if its forward orbit contains a periodic critical point. Otherwise, aAa\in A is called a Julia point. We say that the level-nn non-expanding spine NnN^{n} is essential relative to AA if it contains (more precisely carries as a train-track) a closed curve that is homotopic relative to AA neither to a point nor to some iterate of a peripheral loop of a Julia point in AA.

Theorem 6.21.

Let \mathcal{R} be a finite subdivision rule and f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\to S_{\mathcal{R}} be its subdivision map which is not doubly covered by a torus endomorphism. Let AVert(S)A\subset\operatorname{Vert}(S_{\mathcal{R}}) be a set of marked points, i.e., Pff(A)AP_{f}\cup f(A)\subset A. Then the post-critically finite branched covering f:(S2,A)f:(S^{2},A)\righttoleftarrow has a Levy cycle if and only if the level-nn non-expanding spine NnN^{n} is essential relative to AA for every n0n\geq 0.

We first prove the equivalence between the existence of a Levy cycle and the existence of a sequence curves with certain properties in Section 5 using the theory of self-similar groups. Then we show in Section 6 the equivalence between the existence of such a sequence of curves and the level-nn non-expanding spine being essential at every level n0n\geq 0.

Algorithmic implication. Theorem 6.21 improves [BD18, Algorithm 5.5] by replacing the exhaustive semi-decidable search for nuclei of orbisphere bisets by checking if the non-expanding spines are essential, which terminates in finite time if there is no Levy cycle. There is an example showing that an arbitrarily higher level of non-expanding spine is required to be checked, see Proposition 9.9 in Section 9.3.

Question 1.2.

Is there an upper bound function U:++U:\mathbb{Z}_{+}\to\mathbb{Z}_{+} such that f:(S2,A)f:(S^{2},A)\righttoleftarrow has a Levy cycle if and only if NnN^{n} is essential relative to AA for every n<U(k)n<U(k) where kk is the number of tiles in \mathcal{R}?

Finite subdivision rules with polynomial growth of subdivisions. We will see that the growth of the subdivision of an edge is either exponential or polynomial in Theorem 3.6 and Proposition 8.2. If every edge has polynomial growth of subdivisions, then the level-nn non-expanding spines NnN^{n} are independent of n0n\geq 0. Hence the existence of a Levy cycle is decidable very quickly.

Theorem 8.6.

Let \mathcal{R} be a finite subdivision rule with polynomial growth of edge subdivisions and ff be its subdivision map which is not doubly covered by a torus endomorphism. Let AVert(S)A\subset\operatorname{Vert}(S_{\mathcal{R}}) be a set of marked point, i.e., f(A)PfAf(A)\cup P_{f}\subset A. Then the followings are equivalent.

  1. (1)

    The branched covering f:(S2,A)f:(S^{2},A)\righttoleftarrow does not have a Levy cycle.

  2. (2)

    The level-0 non-expanding spine N0N^{0} is essential relative to AA.

  3. (3)

    The branched covering f:(S2,A)f:(S^{2},A)\righttoleftarrow is combinatorially equivalent to a unique rational map up to conjugation by Möbius transformations.

Equivalence between Levy cycles and Thurston obstructions. Another important implication of Theorem 8.6 is the equivalence between the existence of a Levy cycle and the existence of a Thurston obstruction. As explained earlier, there are topological branched coverings which do not have a Levy cycle but have a Thurston obstruction [ST00]. For some families of post-critically finite topological branched coverings, e.g., post-critically finite topological polynomials or branched coverings of degree 22, the existence of a Thurston obstruction implies the existence of a Levy cycle, by Levy, Rees, Tan, and Berstein [Tan92, Hub16]. We add two new families to this list: subdivision maps with polynomial growth of edge subdivisions (Theorem 8.6) and matings of polynomials one of which has core entropy zero (Corollary 7.10).

Corollary 7.10.

Let ff and gg be post-critically finite hyperbolic (resp. possibly non-hyperbolic) polynomials such that at least one of ff and gg has core entropy zero. Then ff and gg are mateable if and only if the formal mating (resp. degenerate mating) does not have a Levy cycle.

The equivalence between the existence of a Levy cycle and of a Thurston obstruction follows from the graph intersecting obstruction theorem, which is a generalization of the arcs intersecting obstruction theorem by Pilgrim and Tan [PT98]. Here, htoph_{top} indicates the topological entropy.

Theorem 7.6 (Graph intersecting obstruction).

Let f:(S2,A)f:(S^{2},A)\righttoleftarrow be a post-critically finite branched covering and GG be a forward invariant graph such that htop(f|G)=0h_{top}(f|_{G})=0. Then every irreducible Thurston obstruction intersecting GG is a Levy cycle.

Examples: Critically fixed anti-holomorphic maps. In Section 9, we define an orientation reversing finite subdivision rule with no edge subdivision from every 22-vertex-connected planar graph GG. Then f2:(S2,Vert(G))f^{2}:(S^{2},\operatorname{Vert}(G))\righttoleftarrow and fτ:=τf:(S2,Vert(G))f_{\tau}:=\tau\circ f:(S^{2},\operatorname{Vert}(G))\righttoleftarrow are post-critically finite topological branched coverings, where τ\tau is an orientation-reversing automorphism of GG. Then we show in Theorem 9.4 that these maps do not have Levy cycles (or equivalently, Thurston obstructions) if and only if GG is 33-edge-connected. While this article was being written, two papers [LLMM19] and [Gey20] were published where it is shown that every critically fixed anti-holomorphic map is constructed in this way and a theorem almost same as Theorem 9.4 is proved.

Notation for integer intervals. We introduce a non-standard but intuitive notation for integer intervals to distinguish them from real intervals. For a<ba<b\in\mathbb{Z},

  • \cdot

    [a,b]:={a,a+1,,b}[a,b]_{\mathbb{Z}}:=\{a,a+1,\dots,b\}

  • \cdot

    [a,]:={a,a+1,}{}[a,\infty]_{\mathbb{Z}}:=\{a,a+1,\dots\}\cup\{\infty\}

  • \cdot

    [,b]:={}{,b1,b}[-\infty,b]_{\mathbb{Z}}:=\{-\infty\}\cup\{\cdots,b-1,b\}

  • \cdot

    [,]:=[-\infty,\infty]_{\mathbb{Z}}:=\mathbb{Z}

The interval [a,b][a,b] without the subscript Z indicates the real interval {x|axb}\{x\in\mathbb{R}\leavevmode\nobreak\ |\leavevmode\nobreak\ a\leq x\leq b\}.

Acknowledgements.

The author thanks Dylan Thurston and Kevin Pilgrim for helpful comments and discussions. Without their support and suggestions, this work would not have existed. The author also thanks Dzmitry Dudko for critical comments on the previous version of this article. The author thanks the reviewer for helpful comments and suggestions, which inspired the author to write Section 9.4. The author also thanks the developers of Xaos and Mathematica, which were used to draw Julia sets in the present article.

2. Monotonicity of lengths under subdivisions

In this section, we see combinatorial properties of CW-complex without dynamics. We follow some terminology defined in [FPP18b]. Let 𝒯\mathcal{T} be a finite CW-complex structure on S2S^{2}. A nn-gon, or a polygon if nn is not specified, is a 22-dimensional CW-complex structure on the closed 22-disc D2D^{2} whose 11-skeleton consists of nn edges on D2\partial D^{2}. For every closed 22-cell tt of 𝒯\mathcal{T}, there is a polygon 𝐭\mathbf{t} and a characteristic map ϕt:𝐭𝒯\phi_{t}:\mathbf{t}\to\mathcal{T} such that ϕt\phi_{t} is cell-wise homeomorphic and ϕt(𝐭)=t\phi_{t}(\mathbf{t})=t.

Definition 2.1 (Bands and bones).

A band of 𝒯\mathcal{T} is a triple (t;e1,e2)(t;e_{1},e_{2}), where tt is a closed 22-cell and e1e_{1} and e2e_{2} are edges on the boundary of tt. We allow e1=e2(=e)e_{1}=e_{2}(=e) only when two boundary edges of a polygon 𝐭\mathbf{t} is are identified to ee by the characteristic map ϕt:𝐭𝒯\phi_{t}:\mathbf{t}\to\mathcal{T}. We say that e1e_{1} and e2e_{2} are the sides of the band. The bone of (t;e1,e2)(t;e_{1},e_{2}) is the homotopy class (or ambiguously a representative of the class) of curves which are properly embedded into (t,t)(t,\partial t) with endpoints on the interiors of e1e_{1} and e2e_{2}.

Refer to caption
e1e_{1}
e2e_{2}
ee
Figure 2. Bones of bands. The figure on the right shows the case when two sides of the band are the same.

Let 𝒯(n)\mathcal{T}^{(n)} denote the nn-skeleton of 𝒯\mathcal{T}. Any curve γS2𝒯(0)\gamma\subset S^{2}\setminus\mathcal{T}^{(0)} transverse to 𝒯(1)\mathcal{T}^{(1)} is subdivided by 𝒯(1)\mathcal{T}^{(1)} into consecutive subcurves γ1,γ2,,γk\gamma_{1},\gamma_{2},\dots,\gamma_{k} such that each γi\gamma_{i} is a maximal subcurve embedded in a closed 22-cell. The set {γ1,,γk}\{\gamma_{1},\dots,\gamma_{k}\} is the 𝒯\mathcal{T}-decomposition of γ\gamma and each curve γi\gamma_{i} is a 𝒯\mathcal{T}-segment of γ\gamma.

If γ\gamma is not closed, then γ2,,γk1\gamma_{2},\dots,\gamma_{k-1} are called inner 𝒯\mathcal{T}-segments. The terminal 𝒯\mathcal{T}-segment γ1\gamma_{1} or γk\gamma_{k} is an outer 𝒯\mathcal{T}-segment if one of its endpoint is in the interior of a closed 2-cell; if both endpoints are on the 1-skeleton, then we still call them inner 𝒯\mathcal{T}-segments. If γ\gamma is closed, all segments are called inner segments. A curve γ\gamma is 𝒯\mathcal{T}-taut if every inner 𝒯\mathcal{T}-segment is the bone of a band, i.e., it cannot be pushed away from the 22-cell it is contained by an isotopy relative to 𝒯(0)\mathcal{T}^{(0)}.

Definition 2.2.

Two curves in S2𝒯(0)S^{2}\setminus\mathcal{T}^{(0)} are combinatorially equivalent relative to 𝒯\mathcal{T}, or simply 𝒯\mathcal{T}-combinatorially equivalent, if they are isotopic by a cellular isotopy of 𝒯\mathcal{T}, i.e., a isotopy from the identity map whose restriction to each cell XX is also an isotopy on XX.

Define the 𝒯\mathcal{T}-length of γ\gamma, denoted by l𝒯(γ)l_{\mathcal{T}}(\gamma), to be the number of inner 𝒯\mathcal{T}-segments. The 𝒯\mathcal{T}-length of a curve is an invariant of a combinatorial equivalence class. The following criterion is straightforward from the bigon criterion [FM12].

Proposition 2.3.

Let 𝒯\mathcal{T} be a finite CW-complex structure on S2S^{2}. Let γ\gamma be a curve in S2𝒯(0)S^{2}\setminus\mathcal{T}^{(0)} transverse to 𝒯(1)\mathcal{T}^{(1)}. Then l𝒯(γ)l_{\mathcal{T}}(\gamma) is minimized in its homotopy class within S2𝒯(0)S^{2}\setminus\mathcal{T}^{(0)}, relative to endpoints if γ\gamma is not closed, if and only if γ\gamma is taut. Moreover, in the homotopy class, the taut curve is unique up to 𝒯\mathcal{T}-combinatorial equivalence.

The following lemma is immediate.

Lemma 2.4.

Let 𝒯\mathcal{T} be a finite CW-complex structure on S2S^{2}. For every l>0l>0, there are only finitely many, possibly closed or non-closed, curves δ\delta in S2𝒯(0)S^{2}\setminus\mathcal{T}^{(0)} with l𝒯(δ)<ll_{\mathcal{T}}(\delta)<l up to combinatorial equivalence relative to 𝒯\mathcal{T}.

Definition 2.5 (Subbands).

Assume 𝒯\mathcal{T} is a finite CW-complex structure on S2S^{2} and 𝒯\mathcal{T}^{\prime} is its subdivision. A band (t;e1,e2)(t^{\prime};e^{\prime}_{1},e^{\prime}_{2}) of 𝒯\mathcal{T}^{\prime} is a subband of (t;e1,e2)(t;e_{1},e_{2}) of 𝒯\mathcal{T} if ttt^{\prime}\subset t and eieie_{i}^{\prime}\subset e_{i} for i=1,2i=1,2.

Proposition 2.6 (Monotonicity of lengths under refinements of CW-complexes).

Let 𝒯\mathcal{T} and 𝒯\mathcal{T}^{\prime} be finite CW-complex structures of the 22-sphere S2S^{2} such that 𝒯\mathcal{T}^{\prime} is a subdivision of 𝒯\mathcal{T}. Let γ\gamma be a 𝒯\mathcal{T}-taut curve and γ\gamma^{\prime} be a 𝒯\mathcal{T}^{\prime}-taut curve such that γ\gamma and γ\gamma^{\prime} are 𝒯\mathcal{T}-combinatorially equivalent. Then

l𝒯(γ)l𝒯(γ).l_{\mathcal{T}}(\gamma)\leq l_{\mathcal{T}^{\prime}}(\gamma^{\prime}).

Let γ1,,γk\gamma_{1},\dots,\gamma_{k} be the inner 𝒯\mathcal{T}-segments of γ\gamma and γ1,,γk\gamma^{\prime}_{1},\dots,\gamma^{\prime}_{k^{\prime}} be the inner 𝒯\mathcal{T}^{\prime}-segments of γ\gamma^{\prime}. Then the equality holds if and only if, under a proper reordering of indices, γi\gamma_{i} is a bone of (ti;ei,1,ei,2)(t_{i};e_{i,1},e_{i,2}) and γi\gamma_{i}^{\prime} is a bone of (ti;ei,1,ei,2)(t_{i}^{\prime};e_{i,1}^{\prime},e_{i,2}^{\prime}) such that (ti;ei,1,ei,2)(t_{i}^{\prime};e_{i,1}^{\prime},e_{i,2}^{\prime}) is a subband of (ti;ei,1,ei,2)(t_{i};e_{i,1},e_{i,2}) for any 1ik=k1\leq i\leq k=k^{\prime}

Proof.

Take unions of consecutive γi\gamma_{i}^{\prime}’s to get δ1,,δl\delta_{1},\dots,\delta_{l} such that each δj\delta_{j} is a 𝒯\mathcal{T}-segment of γ\gamma^{\prime}. If endpoints of δi\delta_{i} are on the same edge of 𝒯\mathcal{T} as below, then remove it by an isotopy pushing δi\delta_{i} away from the 22-cell that it was contained so as to make δi1δiδi+1\delta_{i-1}\cup\delta_{i}\cup\delta_{i+1} be properly embedded into the 22-cell where the subcurves δi1\delta_{i-1} and δi+1\delta_{i+1} were properly embedded as shown below.

Refer to caption
δi1\delta_{i-1}
δi\delta_{i}
δi+1\delta_{i+1}
Figure 3. The bold edge is an edge of 𝒯\mathcal{T} and the dotted edge is an edge of 𝒯\mathcal{T}^{\prime} that is not an edge of 𝒯\mathcal{T}.

Repeating this reduction, we obtain a 𝒯\mathcal{T}-taut curve γ¯\bar{\gamma} homotopic to γ\gamma. Let γ¯1,,γ¯m\bar{\gamma}_{1},\cdots,\bar{\gamma}_{m} be its subcurves with respect to 𝒯\mathcal{T} and γ¯i\bar{\gamma}_{i} be properly embedded into a band (t¯i;e¯i,1,e¯i,2)(\bar{t}_{i};\bar{e}_{i,1},\bar{e}_{i,2}). Since taut curves are unique in the homotopy class up to combinatorial equivalence, after reordering indices, we have k=mk=m and (t¯i;e¯i,1,e¯i,2)=(ti;ei,1,ei,2)(\bar{t}_{i};\bar{e}_{i,1},\bar{e}_{i,2})=(t_{i};e_{i,1},e_{i,2}). Then k=mlkk=m\leq l\leq k^{\prime}. The equality condition immediately follows from the constructions of δi\delta_{i} and γ¯i\bar{\gamma}_{i}. ∎

3. Directed graphs and topological entropy of graph maps

A directed graph will be used throughout this article to understand the dynamics of branched coverings. In this section, we review basic notions of directed graphs and prove properties that we need in subsequent sections.

Let GG be a finite directed graph. A path is a sequence of edges (e1,e2,,en)(e_{1},e_{2},\dots,e_{n}) such that the terminal vertex of eie_{i} is equal to the initial vertex of ei+1e_{i+1} for every i[1,n1]i\in[1,n-1]_{\mathbb{Z}}. The length of path is the number of edges in the sequence. The initial vertex of e1e_{1} is the initial vertex of the path and the terminal vertex of ene_{n} is the terminal vertex of the path. If the initial and terminal vertices of a path pp are vv and ww, then we call pp a path from vv to ww. Let pp and pp^{\prime} be paths of length nn and nn^{\prime} with n>nn^{\prime}>n. If the first nn subsequence of edges of pp^{\prime} is equal to the sequence of edges of pp, we say that pp^{\prime} is an extension of pp and pp is the first nn-restriction of pp^{\prime}.

A cycle is a path whose initial and terminal vertices coincide. A vertex is periodic if it is contained in a cycle and preperiodic if it is not periodic but there is a path from the vertex to a periodic vertex. For a subset WVert(G)W\subset\operatorname{Vert}(G), the subgraph generated by WW is the subgraph of GG consisting of WW and edges connecting vertices in WW.

Definition 3.1 (Recurrent paths).

For a periodic vertex vVert(G)v\in\operatorname{Vert}(G), a path pp from vv is recurrent if there exists a path from the terminal vertex ww of pp to vv, i.e., vv and ww are contained in one cycle. We also consider a periodic vertex as a recurrent path of length 0.

Definition 3.2 (Ideals).

Let GG be a directed graph. A subset XVert(G)X\subset\operatorname{Vert}(G) is an ideal if the following condition hold: For every vXv\in X, if there is a path from vv to ww for some wVert(G)w\in\operatorname{Vert}(G), then wXw\in X.

For vVert(G)v\in\operatorname{Vert}(G), the ideal generated by vv, denoted by v\left<v\right>, is the collection of vertices wVert(G)w\in\operatorname{Vert}(G) where there is a path from vv to ww.

Example 3.3.

Assume we have a directed graph as below.

v1{v_{1}}v2{v_{2}}v3{v_{3}}v7{v_{7}}v4{v_{4}}v5{v_{5}}v6{v_{6}}e1\scriptstyle{e_{1}}e2\scriptstyle{e_{2}}e3\scriptstyle{e_{3}}e4\scriptstyle{e_{4}}e7\scriptstyle{e_{7}}e5\scriptstyle{e_{5}}e6\scriptstyle{e_{6}}

Vertices v1v_{1} and v4v_{4} are neither periodic or preperiodic; v2,v3,v5,v_{2},v_{3},v_{5}, and v6v_{6} are periodic; v7v_{7} is preperiodic. Starting from v3v_{3}, the paths (e4,e3,e5)(e_{4},e_{3},e_{5}) and (e4,e3,e5,e6,e4,e3)(e_{4},e_{3},e_{5},e_{6},e_{4},e_{3}) are recurrent, but the paths (e4,e2,e1)(e_{4},e_{2},e_{1}) and (e4,e3,e5,e6,e4,e2)(e_{4},e_{3},e_{5},e_{6},e_{4},e_{2}) are not recurrent. There are only 44 ideals: {v4},{v1,v4},{v1,v2,v3,v4,v5,v6},\{v_{4}\},\{v_{1},v_{4}\},\{v_{1},v_{2},v_{3},v_{4},v_{5},v_{6}\}, and Vert(G)\operatorname{Vert}(G).

3.1. Adjacency matrices

Let GG be a finite directed graph and Vert(G)={v1,v2,,vn}\operatorname{Vert}(G)=\{v_{1},v_{2},\dots,v_{n}\}. The adjacency matrix AA of GG is defined by

Aij=thenumberofedgesfromvitovj,A_{ij}=\mathrm{the\leavevmode\nobreak\ number\leavevmode\nobreak\ of\leavevmode\nobreak\ edges\leavevmode\nobreak\ from\leavevmode\nobreak\ }v_{i}\mathrm{\leavevmode\nobreak\ to\leavevmode\nobreak\ }v_{j},

where the AijA_{ij} is the entry of ithi^{th}-row and jthj^{th}-column of AA. The adjacency matrix depends on the choice of indices of viv_{i}’s, and matrices defined by different choices of indices differ by conjugations by permutation matrices. In particular, if the index satisfies i<ji<j only when there exists a path from viv_{i} to vjv_{j}, then the adjacency matrix is an upper triangular block matrix.

(UTB-form) A=(A10A200Ak1000Ak),A=\left(\begin{array}[]{ccccc}A_{1}&*&\dots&*&*\\ 0&A_{2}&\dots&*&*\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&\dots&A_{k-1}&*\\ 0&0&\dots&0&A_{k}\end{array}\right),

where AiA_{i} is irreducible or a 0 matrix. A non-negative m×mm\times m square matrix MM is irreducible if for every 1i,jm1\leq i,j\leq m, there exists k1k\geq 1 such that (Mk)ij>0(M^{k})_{ij}>0. An irreducible non-negative matrix MM has a simple eigenvalue, called the Perron-Frobenius eigenvalue, which is a positive real number and equal to the spectral radius of MM. The spectral radius of AA is equal to the maximum of Perron-Frobenius eigenvalues of the irreducible AiA_{i}. See [BP94, Chapter 2].

Asymptotic growth of entries of AnA^{n}.. Let BB be a non-negative irreducible matrix and denote by vv and wTw^{T} the right and the left eigenvectors of BB which are normalized by wTv=1w^{T}v=1. By the Perron-Frobenius theorem, for the Perron-Frobenius eigenvalue ρ\rho of BB, we have

limnBn/ρn=vwT.\lim_{n\to\infty}B^{n}/\rho^{n}=vw^{T}.

Hence the (i,j)(i,j)-entry of BnB^{n} is asymptotic to ρnviwj\rho^{n}\cdot v_{i}\cdot w_{j}. This implies the following proposition. The proof is left to the reader.

Proposition 3.4.

Let AA be an N×NN\times N non-negative upper triangular block matrix such that A1,AkA_{1},\dots A_{k} are blocks on the diagonal like (UTB-form). Let AiA_{i} be an Ni×NiN_{i}\times N_{i} matrix so that N=i=1kNiN=\sum_{i=1}^{k}N_{i}. For i{1,,N}i\in\{1,\dots,N\}, let li[1,k]l_{i}\in[1,k]_{\mathbb{Z}} such that AliA_{l_{i}} contains the (i,i)(i,i)-entry of AA. Let Rn,iR_{n,i} (resp. Cn,iC_{n,i}) be the ithi^{th}-row (resp. column) sum of AnA^{n}, i.e., the sum of the entries in the ithi^{th}-row (resp. column) of AnA^{n}. Then

{limn1nlog(Rn,i)=maxlijklogρjlimn1nlog(Cn,i)=max1jlilogρj\left\{\begin{array}[]{l}\lim_{n\to\infty}\frac{1}{n}\cdot\log(R_{n,i})=\max_{l_{i}\leq j\leq k}\log\rho_{j}\\[2.0pt] \lim_{n\to\infty}\frac{1}{n}\cdot\log(C_{n,i})=\max_{1\leq j\leq l_{i}}\log\rho_{j}\end{array}\right.

where ρj\rho_{j} is the spectral radius of AjA_{j}.

3.2. The growth rate of the number of paths

Polynomial and exponential growth rate. Suppose we have a sequence {an|an>0,n1}\{a_{n}\leavevmode\nobreak\ |\leavevmode\nobreak\ a_{n}\in\mathbb{R}_{>0},\leavevmode\nobreak\ n\geq 1\}. We consider sequences which are uniformly bounded or diverge to the infinity when nn tends to the infinity. We say that the sequence has exponential growth, or grows exponentially fast, if there exists C>D>1C>D>1 such that for any sufficiently large n>0n>0 we have

D<|logan|<C,D<|\log a_{n}|<C,

and the sequence has polynomial growth of degree dd, or grows polynomially fast with degree dd, for a non-negative integer dd if there exists C>D>0C>D>0 such that for any sufficiently large n>0n>0 we have

Dnd<an<Cnd.D\cdot n^{d}<a_{n}<C\cdot n^{d}.
Lemma 3.5.

Suppose {an|n1,an0}\{a_{n}\leavevmode\nobreak\ |\leavevmode\nobreak\ n\geq 1,\leavevmode\nobreak\ a_{n}\geq 0\} is a non-decreasing sequence. If there exist k,l,m>0k,l,m>0 such that akn+lmnda_{k\cdot n+l}\geq m\cdot n^{d} (resp. akn+lmnda_{k\cdot n+l}\leq m\cdot n^{d}) for every n>0n>0, then there exists C>0C>0 such that an>Cnda_{n}>C\cdot n^{d} (resp. an<Cnda_{n}<C\cdot n^{d}) for any sufficiently large nn.

Proof.

Since the sequence {an}\{a_{n}\} is non-decreasing, we have a(k+1)nmnda_{(k+1)\cdot n}\geq m\cdot n^{d} for any n>l/kn>l/k. Then, for any sufficiently large n>0n>0, we have

ana(k+1)nk+1mnk+1d>m(k+2)dnd,a_{n}\geq a_{(k+1)\cdot\lfloor\frac{n}{k+1}\rfloor}\geq m\cdot{\left\lfloor\frac{n}{k+1}\right\rfloor}^{d}>\frac{m}{(k+2)^{d}}\cdot n^{d},

where x\lfloor x\rfloor means the largest integer less than or equal to xx. The same argument also works for the inverse direction. ∎

Let GG be a directed graph. Two paths in GG are different if their sequences of edges are different. Let P(v,n)P(v,n) be the number of different paths of length exactly nn starting from vv. The following criterion on the growth rate of P(v,n)P(v,n) is well-known. See [Sid00, Ufn82]. We restate the criterion with a slight improvement for the case of polynomial growth.

Theorem 3.6.

Let GG be a directed graph and vVert(G)v\in\operatorname{Vert}(G).

  • (1)

    If there exist two different cycles passing through vv, then P(v,n)P(v,n) has exponential growth.

  • (2)

    If there exists wVert(G)w\in\operatorname{Vert}(G) such that there are two different cycles starting from ww and a path from vv to ww, then P(v,n)P(v,n) has exponential growth.

  • (3)

    If vv does not satisfy (1)(1) or (2)(2), then P(v,n)P(v,n) has polynomial growth. Moreover, if the maximum number of (disjoint) cycles that a path from vv can intersect is d+1d+1, then the degree of the polynomial growth of P(v,n)P(v,n) is dd. When the maximum number of cycles is zero, then P(v,n)=0P(v,n)=0 for every sufficiently large nn and we define the degree of polynomial growth to be 1-1.

Proof.

The number of paths of length n\leq n is counted in [Ufn82] and we slightly modify it. If two cycles of lengths pp and qq pass through vv, then there are at least 2n2^{n} paths of length npqnpq starting from vv. Let MM be the maximal number of outgoing edges from one vertex in GG. Then the number of paths with length nn is less than MnM^{n}. This proves (1)(1) and (2)(2) is immediate from (1)(1).

Assume that vv does not satisfy (1)(1) or (2)(2). Let W=vW=\left<v\right> be the ideal generated by vv. There exist totally ordered finite subsets W1,WkW_{1},\cdots W_{k} such that vWiv\in W_{i} for each ii and W=i=1kWiW=\bigcup_{i=1}^{k}W_{i}. More precisely, we think of a directed graph HH obtained by collapsing each cycle in GG to a vertex. Since vv does not satisfy (1) or (2), the graph HH does not have any cycle. Let vv^{\prime} be the vertex of HH corresponding to vv. Then vv^{\prime} is the only minimal element of Vert(H)\operatorname{Vert}(H). Let w1,w2,,wmw_{1},w_{2},\dots,w_{m} be the maximal elements of Vert(H)\operatorname{Vert}(H). Choose any wVert(W)w\in\operatorname{Vert}(W).

Then the subgraph GiG_{i} generated by WiW_{i} is isomorphic to the graph in Figure 4. Since every path from vv is supported in some GiG_{i}, it suffices to show (3)(3) when GG is the graph of the type shown in Figure 4.

Refer to caption
vv
Figure 4. A graph with polynomial growth rate of P(v,n)P(v,n). Any arrow indicates paths, any dotted arrow means it may not exist but if exists it indicates a path, and any circle indicates a cycle. In each cycle, the incoming vertex and the outgoing vertex could be the same.

Assume GG is a graph of the type in Figure 4. Let it have d+1d+1 cycles of lengths p1,p2,,pd+1p_{1},p_{2},\dots,p_{d+1}. Let pMp_{M} and pmp_{m} be the maximum and minimum of {p1,,pd}\{p_{1},\dots,p_{d}\} and KK and LL be the number of vertices and edges of GG respectively. Let X(n)X(n) be the set of dd-tuples (n1,,nd)(n_{1},\dots,n_{d}) of non-negative integers satisfying n1++ndnn_{1}+\dots+n_{d}\leq n. The set X(n)X(n) has (n+dd)\binom{n+d}{d} elements.

Claim: For any n1n\geq 1, P(v,pmn)/K(n+dd)P(v,pMn+L)P(v,p_{m}\cdot n)/K\leq\binom{n+d}{d}\leq P(v,p_{M}\cdot n+L).

Proof of claim.

For any nn, we define an injective map from X(n)X(n) to the set of paths from vv of length pMn+Lp_{M}\cdot n+L as follow. For any (n1,,nd)X(n)(n_{1},\dots,n_{d})\in X(n), we have a path from vv which goes around the ii-th cycle nin_{i}-times for idi\leq d. The path has length not greater than pMn+Lp_{M}\cdot n+L. There is a unique extension of the path which (1) does not further go around the first d cycles, i.e., the additional rotations occur only in the last cycle, and (2) has length exactly equal to pMn+Lp_{M}\cdot n+L. We assign the extended path to the element (n1,,nd)X(n)(n_{1},\dots,n_{d})\in X(n). Hence we have (n+dd)P(v,pMn+L)\binom{n+d}{d}\leq P(v,p_{M}\cdot n+L).

Similarly, we define another map from the set of paths of length pmnp_{m}\cdot n to X(n)X(n) by assigning the numbers of times that each path goes around the first dd cycles. Suppose that two such paths p1p_{1} and p2p_{2} have the same image in X(n)X(n). Since p1p_{1} and p2p_{2} start from the same vertex vv and have the same length, there terminal vertex are the same if and only if p1p_{1} and p2p_{2} are the same. It follows that the number of preimage of the map is less than KK. ∎

As (n+dd)\binom{n+d}{d} is a degree dd polynomial in nn, it follows from Lemma 3.5 that the P(v,n)P(v,n) has polynomial growth of degree dd.as a sequence in nn. ∎

The polynomial growth rate of P(v,n)P(v,n) implies the recurrent extension is unique.

Proposition 3.7 (Extension of recurrent paths).

Let GG be a directed graph and vVert(G)v\in\operatorname{Vert}(G). Let γ\gamma be a path from vv of length m>0m>0. If γ\gamma is recurrent, then for any n>mn>m, γ\gamma has at least one extension to a recurrent path of length nn. Moreover, if P(v,n)P(v,n) grows polynomially fast as nn tends to \infty, then the extension is unique.

Proof.

Since γ\gamma is recurrent, the initial and the terminal points of γ\gamma belong to one cycle CC of GG. Then we can extend γ\gamma by repeatedly travelling along CC. If P(v,n)P(v,n) grows polynomially fast, then CC is the only cycle that passes through vv. Hence travelling along CC is the only way to extend γ\gamma. ∎

3.3. Graph maps with zero topological entropy

Let GG be a finite graph and Vert(G)\operatorname{Vert}(G) be its vertex set. A continuous map f:GGf:G\to G Markov if f(Vert(G))Vert(G)f(\operatorname{Vert}(G))\subset\operatorname{Vert}(G) and ff is a homeomorphism or constant on each component of Gf1(Vert(G))G\setminus f^{-1}(\operatorname{Vert}(G)). Then the edges form a Markov partition of ff. Denote by e1,e2,e_{1},e_{2},\dots the edges of GG. Note that every edge is mapped to a union of edges. The adjacency matrix AfA_{f} of f:GGf:G\to G is defined in such a way that f(ei)f(e_{i}) covers eje_{j} as many as the (i,j)(i,\,j)-entry of AfA_{f}. Under a suitable choice of indexing eie_{i}’s, we may assume that AfA_{f} is an upper triangular block matrix. Let A1,AkA_{1},\dots A_{k} be the block matrices on the diagonal as in (UTB-form).

The spectral radius λ\lambda of AfA_{f} is either equal to zero if every AiA_{i} is zero or equal to the maximum of Perron-Frobenius eigenvalues of the irreducible AiA_{i}’s. The topological entropy htop(f)h_{top}(f) of ff is equal to zero if λ=0\lambda=0 or equal to log(λ)\log(\lambda) if λ>0\lambda>0. The relationship between the topological entropy and the Perron-Frobenius eigenvalue follows from [MS80].

There is a directed graph 𝒟f\mathcal{D}_{f} such that Vert(𝒟f)=Edge(G)\operatorname{Vert}(\mathcal{D}_{f})=\operatorname{Edge}(G) and the directed edges are defined as follow: For ei,ejEdge(G)e_{i},e_{j}\in\operatorname{Edge}(G), if f(ei)f(e_{i}) covers eje_{j} exactly kk-times, then we draw kk directed edges of 𝒟f\mathcal{D}_{f} from eie_{i} to eje_{j}. Then AfA_{f} is equal to the adjacency matrix of 𝒟f\mathcal{D}_{f}. We refer to 𝒟f\mathcal{D}_{f} as the directed graph of the Markov map f:GGf:G\to G.

The following lemma is elementary, but we prove it for the sake of self-containedness.

Lemma 3.8.

Let MM be an irreducible non-negative integer matrix and λ(M)\lambda(M) be its Perron-Frobenius eigenvalue. Then λ(M)1\lambda(M)\geq 1 and the equality holds if and only if MM is a permutation matrix.

Proof.

Because the characteristic polynomial of MM is a monic polynomial with integer coefficients, the absolute value of the product of eigenvalues is a positive integer. Hence the spectral radius λ(M)\lambda(M) is at least one. If MM is a permutation matrix, then trivially λ(M)=1\lambda(M)=1. Assume λ(M)=1\lambda(M)=1. Let HH be a directed graph whose adjacency matrix is MM. Since MM is irreducible, for any pair (v,w)(v,w) of vertices of HH there exists a path p1p_{1} from vv to ww and a path p2p_{2} from ww to vv. If there exists a vertex xv,wx\neq v,w through which both paths p1p_{1} and p2p_{2} pass, there are two different cycles passing through xx. By Theorem 3.6, the number of length-nn paths P(x,n)P(x,n) grows exponentially fast. Since P(x,n)P(x,n) is the sum of entries in the row of MnM^{n} corresponding to the vertex vv, it follows that λ(M)>1\lambda(M)>1. So p1p_{1} and p2p_{2} form a cycle which passes through vertices exactly once. If there is a vertex xx that is not contained in this cycle, then there is a path from xx to vv and a path from vv to xx. Then two different cycles pass through vv so P(v,n)P(v,n) grows exponentially fast and λ(M)>1\lambda(M)>1. Hence HH is a cycle passing through every vertex exactly once, and MM is a permutation matrix. ∎

Proposition 3.9.

Let f:GGf:G\to G be a Markov map. Then the followings are equivalent.

  • (1)

    The topological entropy htop(f)h_{top}(f) is zero.

  • (2)

    Every irreducible block AiA_{i} of the upper-triangular block form of the adjacency matrix AfA_{f} is a permutation matrix.

  • (3)

    The directed graph 𝒟f\mathcal{D}_{f} of the adjacency matrix AfA_{f} has disjoint cycles, i.e., every pair of different cycles have disjoint vertices.

  • (4)

    There exists a positive integer dd such that (Afn)ij=O(nd)({A_{f}}^{n})_{ij}=O(n^{d}) for all i,ji,\,j.

Proof.

(3)(2)(1)(3)\Leftrightarrow(2)\Rightarrow(1) is trivial and (4)(3)(4)\Leftrightarrow(3) is immediate from Theorem 3.6. Assume htop(f)=0h_{top}(f)=0. Then the Perron-Frobenius eigenvalue of every irreducible block AiA_{i} of the adjacency matrix AfA_{f} is one. (1)(2)(1)\Rightarrow(2) follows from Lemma 3.8.

4. Finite subdivision rules

A finite subdivision rule \mathcal{R} consists of the following:

  • (1)

    a subdivision complex SS_{\mathcal{R}} which is a 22-dimensional finite CW-complex such that the underlying space is the union of its closed 22-cells, i.e., every 0 or 11-cell is on the boundary of a 22-cell,

  • (2)

    a subdivision (S)\mathcal{R}(S_{\mathcal{R}}) of SS_{\mathcal{R}}, that is a CW-complex for which every open cell is contained in an open cell of SS_{\mathcal{R}}, and

  • (3)

    a subdivision map f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\rightarrow S_{\mathcal{R}} which is continuous and cell-wise homeomorphic, i.e., its restriction to each open cell is a homeomorphism onto an open cell.

We say that \mathcal{R} is orientation-preserving if every 22-cell can be oriented in such a way that ff preserves the orientation. Similarly, \mathcal{R} is orientation-reversing if ff reverses the orientation on every cell. For any closed 22-cell tt of SS_{\mathcal{R}}, there exist a nn-gon 𝐭\bf{t} and the characteristic map ϕt:𝐭t\phi_{t}:\mathbf{t}\to t is cell-wise homeomorphic. The CW-complex 𝐭\mathbf{t} is called the tile type of the closed 22-cell tt. Similarly, for a characteristic map of a 11-cell ϕe:𝐞e\phi_{e}:\mathbf{e}\to e, 𝐞\mathbf{e} is the edge type of a closed 11-cell ee.

A 22-dimensional CW-complex XX is an \mathcal{R}-complex if it is the union of its closed 22-cells and there is a continuous cell-wise homeomorphism g:XSg:X\rightarrow S_{\mathcal{R}}. By pulling back the subdivision n(S)\mathcal{R}^{n}(S_{\mathcal{R}}) of SS_{\mathcal{R}} through gg for each n>0n>0, we also have a subdivision n(X)\mathcal{R}^{n}(X) of XX and a cell-wise homeomorphism fng:n(X)Sf^{\circ n}\circ g:\mathcal{R}^{n}(X)\rightarrow S_{\mathcal{R}}. For example, SS_{\mathcal{R}} itself and any tile type 𝐭\mathbf{t} are \mathcal{R}-complexes, so for any nn\in\mathbb{N} their level-nn subdivisions n(S)\mathcal{R}^{n}(S_{\mathcal{R}}) and n(𝐭)\mathcal{R}^{n}(\mathbf{t}) are defined. For any edge type 𝐞\mathbf{e}, its level-nn subdivision n(𝐞)\mathcal{R}^{n}(\mathbf{e}) is also similarly defined.

We call a closed 22-cell (resp. a closed 11-cell, a 0-cell) a tile (resp. an edge, a vertex) of a 22-dimensional complex. Every level-0 tile or edge of an \mathcal{R}-complex is also a \mathcal{R}-complex. See [CFP01] for more details on finite subdivision rules.

Notation.. As we wrote in the previous paragraph, we use bold fonts for the domains of characteristic maps and normal fonts for the corresponding closed cells in the CW-complexes. For example, for a closed 22-cell tt in a CW-complex XX, we write ϕt:𝐭t\phi_{t}:\mathbf{t}\to t for the characteristic map. Thus 𝐭\mathbf{t} is always homeomorphic to the closed 22-disk, but tt may not.

Remark 4.1.

Unlike in other articles on finite subdivision rules, every tile type is not assumed to have at least three vertices in this article. This modification allows the graphs in Theorem 9.4 to have bigon faces, see Example 9.8.

Subdivision maps as post-critically finite branched coverings. Throughout this article, we assume SS_{\mathcal{R}} is homeomorphic to the 22-sphere S2S^{2}. Considering (S)\mathcal{R}(S_{\mathcal{R}}) and SS_{\mathcal{R}} as different complexes on the same underlying 22-sphere, we may think of the subdivision map f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\rightarrow S_{\mathcal{R}} as a topological branched self covering of S2S^{2}. Since the set of critical points Ωf\Omega_{f} is a subset of the set of vertices of (S)\mathcal{R}(S_{\mathcal{R}}), ff is post-critically finite.

A set of marked points AA of \mathcal{R} is a subsets of Vert(S)\operatorname{Vert}(S_{\mathcal{R}}) with Pff(A)AP_{f}\cup f(A)\subset A. With a choice of a set of marked points AA, the subdivision map can be considered as a marked post-critically finite branched covering f:(S2,A)f:(S^{2},A)\righttoleftarrow.

4.1. Branched coverings represented as subdivision maps

If ff is a subdivision map, then the 11-skeleton S(1)S_{\mathcal{R}}^{(1)} is a graph such that (1) it contains PfP_{f}, (2) it is connected, and (3) it is forward invariant under ff. Conversely, if there is a graph satisfying the three conditions, then it defines a finite subdivision rule. Below is a list of some forward invariant graphs that are known to exist.

  • Spiders of polynomials [HS94].

  • Hubbard trees can be augmented to be invariant trees [ST19].

  • Jordan curves [BM17] and trees [Hlu17] of expanding Thurston maps.

  • Tischler graphs of critically fixed rational maps [PT98, Hlu19].

  • Jordan curves [GHMZ18] and trees [Hlu17] of Sierpiński Carpet rational maps.

  • Extended Newton graphs for post-critically finite Newton maps [LMS15].

  • A sufficiently large iterate fnf^{n} of any post-critically finite branched covering ff without a Levy cycle is homotopic to a subdivision map [FPP20]. Its 1-skeleton is invariant up to homotopy.

There are post-critically finite branched coverings whose any iterates cannot be represented as subdivision maps [FPP20, Section 4].

4.2. Combinatorial properties of \mathcal{R} and Levy and Thurston obstructions

A finite subdivision rule \mathcal{R} is edge-separating if for every tile type 𝐭\mathbf{t} and pair of disjoint closed edges ee and ee^{\prime} of 𝐭\mathbf{t}, there exists a positive integer nn such that no subtile of 𝐭\mathbf{t} in n(𝐭)\mathcal{R}^{n}(\mathbf{t}) contains both a subedge of ee and a subedge of ee^{\prime}. Similarly, \mathcal{R} is vertex-separating if for every tile type 𝐭\mathbf{t} and pair of vertices vv and ww of 𝐭\mathbf{t}, there exists a positive integer nn such that no subtile of 𝐭\mathbf{t} in n(𝐭)\mathcal{R}^{n}(\mathbf{t}) contains both vv and ww. These two separating conditions are a part of sufficient condition for a subdivision map not having a Levy cycle or a Thurston obstruction.

  • If \mathcal{R} is vertex-separating and edge-separating, then ff does not have a Levy cycle [FPP18b, Proposition 5.1]. There is a finite subdivision rule which is neither edge-separating nor vertex-separating but does not have a Levy cycle, see Example 4.2.

  • If \mathcal{R} is vertex-separating, edge-separating and conformal (we do not define this conformality in the article), then ff does not have a Thurston obstruction [CFKP03]. There is an example [CFKP03, Example 4.6] of finite subdivision rule which is not conformal but does not have a Thurston obstruction, thus it is combinatorially equivalent to a rational map. See [CFKP03] for a definition of conformal finite subdivision rules.

Example 4.2.

The finite subdivision rule RR given in Figure 5 is Example 5.3 of [FPP18b]. Its CW-complex 𝒮\mathcal{S}_{\mathcal{R}} of the 22-sphere consists of two square tiles. The edges of white and shaded tiles are glued to form a pillowcase. Since the shaded tile does not subdivide, \mathcal{R} is neither edge-separating or vertex-separating. However, it easily follows from Theorem 6.21 that the subdivision map does not have a Levy cycle.

Refer to caption
Figure 5. An example of expanding finite subdivision rule which is neither edge-separating nor vertex-separating.

4.3. Edges, bands, bones, and curves of subdivision complexes

For n0n\geq 0, a level-nn tile, edge, or band of \mathcal{R} is a tile, edge or band of n(S)\mathcal{R}^{n}(S_{\mathcal{R}}). See Definition 2.1 for definitions of bands and their bones. There is a bijection between level-0 tiles (resp. edge) and tile types (resp. edge types); a level-0 tile tt is the image of the tile type 𝐭\mathbf{t} under the characteristic map ϕt:𝐭t\phi_{t}:\mathbf{t}\to t.

We will use superscripts to indicate the level of tiles, edges, etc. Since frequently considering level-0 objects, we sometimes omit the superscript 0 for simplicity.

For n>mn>m, a level-nn tile tnt^{n} is a subtile of a level-mm tile tmt^{m} if tntmt^{n}\subset t^{m}. Let 𝐭\mathbf{t} be a tile type and tt be the corresponding level-0 tile. A level-nn tile tnt^{n} is of type 𝐭\mathbf{t} if fn(tn)=tf^{n}(t^{n})=t. Subedges and their types are similarly defined. A band type is a level-0 band. For a band type (t;e1,e2)(t;e_{1},e_{2}), a level-nn band (tn;e1n,e2n)(t^{n};e_{1}^{n},e_{2}^{n}) is of type (t;e1,e2)(t;e_{1},e_{2}) if the fnf^{n}-image of its bone is the bone of (t;e1,e2)(t;e_{1},e_{2}) (or, equivalently, if tnt^{n} is of type 𝐭\mathbf{t} and eine_{i}^{n} is of type 𝐞i\mathbf{e}_{i} for i=1,2i=1,2. For n>mn>m, a level-nn band (tn;e1n,e2n)(t^{n};e_{1}^{n},e_{2}^{n}) is a subband of a level-mm band (tm;e1m,e2m)(t^{m};e_{1}^{m},e_{2}^{m}) if tntmt^{n}\subset t^{m} and eineime_{i}^{n}\subset e_{i}^{m} for i=1,2i=1,2. If deg(f)=d\deg(f)=d, there are dnd^{n} level-nn tiles, edges, and bands of the same type.

Definition 4.3 (Abbreviations for level-nn bands and bones).

There are many level-nn bands that are not subbands of level-0 bands. However the only level-nn bands that we consider are level-nn subbands of level-0 bands. Since these objects will be very frequently used, for the sake of simple notation, by a level-nn band we mean a level-nn subband of a level-0 band. Similarly, by a level-nn bone we mean the bone of a level-nn subband of a level-0 band.

Definition 4.4 (Non-expanded level-nn curves).

Let \mathcal{R} be a finite subdivision rule. Let II be a closed interval [k,l][k,l], (,k](-\infty,k], [k,)[k,\infty), or (,)(-\infty,\infty) for k<lk<l\in\mathbb{Z}. For n0n\geq 0, a curve γn:In(S)\gamma^{n}:I\to\mathcal{R}^{n}(S_{\mathcal{R}}) is a non-expanded level-nn curve, if γn([i,i+1])\gamma^{n}([i,i+1]) is a level-nn bone for every ii\in\mathbb{Z} with [i,i+1]I[i,i+1]\subset I. A non-expanded level-nn curve is recurrent if it consists of level-nn bones that are recurrent. The recurrent bands and bones are defined in Definition 4.8.

4.4. Two directed graphs defined from finite subdivision rules

4.4.1. Directed graphs of edge subdivisions

Let \mathcal{E} be a directed graph such that Vert()\operatorname{Vert}(\mathcal{E}) is the same as the set of level-0 edges. To avoid confusion, we denote by [e][e] the vertex of \mathcal{E} corresponding to an edge ee. A directed edge from [e][e] to [e][e^{\prime}] corresponds to a level-11 subedge of ee of type 𝐞\mathbf{e}^{\prime}. We call \mathcal{E} the directed graph of edge subdivision of \mathcal{R}. The next proposition is straightforward from the definitions.

Proposition 4.5.

There is an 1-1 correspondence between the paths in \mathcal{E} of length nn starting from [e][e] and the level-nn subedges of ee. Thus, the number of level-nn subedges is equal to P([e],n)P_{\mathcal{E}}([e],n), the number of paths of length nn starting from [e][e].

Definition 4.6 (Periodic and recurrent edges).

Let \mathcal{R} be a finite subdivision rule and \mathcal{E} be the directed graph of edge subdivisions of \mathcal{R}. We define level-0 periodic edges and recurrent level-nn edges as follows.

  • A level-0 edge ee is periodically (resp. preperiodically) subdividing, or simply periodic (resp. preperiodic), if [e]Vert()[e]\in\operatorname{Vert}(\mathcal{E}) is periodic (resp. preperiodic). Equivalently, ee is periodic if and only if there exists a level-nn subedge of ee of type 𝐞\mathbf{e} for some n>0n>0.

  • A level-nn edge ene^{n} is a recurrent subedge of ee if it corresponds, by Proposition 4.5, to a recurrent path in \mathcal{E} which starts from [e][e] and has length nn, or equivalently if a further subdivision of ene^{n} contains a subedge of type 𝐞\mathbf{e}. If 𝐞\mathbf{e}^{\prime} is the type of a recurrent subedge of ee, then there is a cycle passing through both [e][e^{\prime}] and [e][e]. Only periodic level-0 edges have recurrent subedges.

  • We also refer to periodic level-0 edges as recurrent level-0 edges, which is sometimes useful for concise statements.

4.4.2. Directed graphs of bands

Let \mathcal{B} be a directed graph such that Vert()\operatorname{Vert}(\mathcal{B}) is the set of level-0 bands (t;e1,e2)(t;e_{1},e_{2}). To avoid confusion, we use bracket [(t;e1,e2)][(t;e_{1},e_{2})] to denote vertices of \mathcal{B}. Every directed edge from [(t;e1,e2)][(t;e_{1},e_{2})] to [(t;e1,e2)][(t^{\prime};e^{\prime}_{1},e^{\prime}_{2})] corresponds to a level-11 subband of (t;e1,e2)(t;e_{1},e_{2}) of type (t;e1,e2)(t^{\prime};e^{\prime}_{1},e^{\prime}_{2}). We call \mathcal{B} the directed graph of bands of \mathcal{R}. The following proposition is an analogue to Proposition 4.5.

Proposition 4.7.

There is an 1-1 correspondence between the paths in \mathcal{B} of length nn starting from [(t;e1,e2)][(t;e_{1},e_{2})] and the level-nn subbands of (t;e1,e2)(t;e_{1},e_{2}).

Definition 4.8 (Periodic and recurrent bands and bones).

Let \mathcal{R} be a finite subdivision rule and \mathcal{B} be the directed graph of bands of \mathcal{R}. We define level-0 periodic bands and level-nn recurrent subbands as we did for edges.

  • A level-0 band (t;e1,e2)(t;e_{1},e_{2}) is periodic (resp. preperiodic), if [(t;e1,e2)]Vert()[(t;e_{1},e_{2})]\in\operatorname{Vert}(\mathcal{B}) is periodic (resp. preperiodic). Equivalently, (t;e1,e2)(t;e_{1},e_{2}) is periodic if and only if there exists a level-nn band (tn;e1n,e2n)(t^{n};e_{1}^{n},e_{2}^{n}) of type (t;e1,e2)(t;e_{1},e_{2}) which is a subband of (t;e1,e2)(t;e_{1},e_{2}) for some n>0n>0.

  • A level-nn subband (tn;e1n,e2n)(t^{n};e_{1}^{n},e_{2}^{n}) of (t;e1,e2)(t;e_{1},e_{2}) is a recurrent subband of (t;e1,e2)(t;e_{1},e_{2}) if it corresponds, by Proposition 4.7, to a recurrent path of length nn starting from [(t;e1,e2)][(t;e_{1},e_{2})], or, equivalently, if (tn;e1n,e2n)(t^{n};e_{1}^{n},e_{2}^{n}) is a subband of (t;e1,e2)(t;e_{1},e_{2}) and has a subband in its further subdivision which is also a subband of (t;e1,e2)(t;e_{1},e_{2}). If (t;e1,e2)(t^{\prime};e^{\prime}_{1},e^{\prime}_{2}) is the type of a recurrent subband of (t;e1,e2)(t;e_{1},e_{2}), then there is a cycle passing through both [(t;e1,e2)][(t^{\prime};e^{\prime}_{1},e^{\prime}_{2})] and [(t;e1,e2)][(t;e_{1},e_{2})]. Only periodic level-0 bands have recurrent subbands.

  • We also refer to periodic level-0 bands as recurrent level-0 bands, which is sometimes useful for concise statements.

We say that a level-nn bone is recurrent if its corresponding level-nn band is recurrent.

A continuous map between two directed graphs is a graph homomorphism if it sends vertices to vertices and edges to edges preserving directions. From a finite subdivision rule \mathcal{R}, we have defined two directed graphs \mathcal{E} and \mathcal{B}. There are natural graph homomorphisms ι,τ:\iota,\tau:\mathcal{B}\rightarrow\mathcal{E} defined by ι([(t;e1,e2)])=[e1]\iota([(t;e_{1},e_{2})])=[e_{1}] and τ([(t;e1,e2)])=[e2]\tau([(t;e_{1},e_{2})])=[e_{2}]. The next lemma follows from the fact that ι\iota and τ\tau are homomorphisms.

Lemma 4.9.

If (t;e1,e2)(t;e_{1},e_{2}) is a periodic level-0 band, then e1e_{1} and e2e_{2} are periodic edges. If (tn;e1n,e2n)(t^{n};e_{1}^{n},e_{2}^{n}) is a level-nn recurrent subband of (t;e1,e2)(t;e_{1},e_{2}), then the sides e1ne_{1}^{n} and e2ne_{2}^{n} of (tn;e1n,e2n)(t^{n};e_{1}^{n},e_{2}^{n}) are level-nn recurrent subedges of e1e_{1} and e2e_{2}.

4.5. Parents and children

We define parents and children for various objects regarding finite subdivision rules. The following are some important properties of the parent-child relationship. For any i>0i>0, 𝐗𝐢\bf{X}^{i} stands for a level-ii object, which can be an edge, a band, or a curve consisting of the bones of bands.

  • (Transitivity) For n>m>l0n>m>l\geq 0, if 𝐗𝐧\bf{X}^{n} is a child of 𝐗𝐦\bf{X}^{m} and 𝐗𝐦\bf{X}^{m} is a child of object 𝐗𝐥\bf{X}^{l}, then 𝐗𝐧\bf{X}^{n} is a child of 𝐗𝐥\bf{X}^{l}. A similar statement holds for parents.

  • (Unique existence of parents) For n>m0n>m\geq 0 every level-nn object 𝐗𝐧\bf{X}^{n} has a unique level-mm parent 𝐗𝐦\bf{X}^{m}. If 𝐗𝐧\bf{X}^{n} is recurrent, then so is 𝐗𝐦\bf{X}^{m}.

  • (Existence of recurrent children) For n>m0n>m\geq 0 every level-mm recurrent 𝐗𝐦\bf{X}^{m} has at least one level-nn child 𝐗𝐧\bf{X}^{n} that is also recurrent. We note that it does not work for non-expanded curves consisting of more than one bones in general.

Edges. Suppose that a level-nn edge ene^{n} is a subedge of a level-mm edge eme^{m} where n>mn>m. Then we say that ene^{n} is a level-nn child of eme^{m} and eme^{m} is a level-mm parent of ene^{n}.

The transitivity is straightforward. If both ene^{n} and eme^{m} are subedges of a level-0 edge ee, then they correspond to directed paths in \mathcal{E} of length nn and mm, say pp and pp^{\prime} respectively, such that both pp and pp^{\prime} start from [e][e] and pp^{\prime} is the first length-mm restriction of pp. Then the unique existence of parents follow. The existence of recurrent children follows from Proposition 3.7.

Bands and bones. Suppose that a level-nn band bnb^{n} is a subband of a level-mm band bmb^{m} for some n>mn>m. Then we say that bnb_{n} is a child of bmb_{m} and bmb_{m} is a parent bnb_{n}. The transitivity, the unique existence of parents, and the existence of recurrent children follow from a similar argument used in the case of edges.

We define parents and children for bones according to the parents-children relationship of their corresponding bands.

Non-expanded curves. Let II be a closed interval with integer ends, such as [k,l][k,l], (,k](-\infty,k], [k,)[k,\infty), or (,)(-\infty,\infty) for k<lk<l\in\mathbb{Z}. For n>m0n>m\geq 0, let γn:In(S)\gamma^{n}:I\to\mathcal{R}^{n}(S_{\mathcal{R}}) and γm:Im(S)\gamma^{m}:I\to\mathcal{R}^{m}(S_{\mathcal{R}}) be level-nn and level-mm non-expanded curves respectively. Recall that γn([i,i+1])\gamma^{n}([i,i+1]) (resp. γm([i,i+1])\gamma^{m}([i,i+1])) is a level-nn (resp. level-mm) bone for every ii\in\mathbb{Z} with [i,i+1]I[i,i+1]\subset I. If γn([i,i+1])\gamma^{n}([i,i+1]) is a level-nn child of γn([i,i+1])\gamma^{n}([i,i+1]) for every ii\in\mathbb{Z} with [i,i+1]I[i,i+1]\subset I, then we say that γn\gamma^{n} is a level-nn child of γm\gamma^{m} and γm\gamma^{m} is a level-mm parent of γn\gamma^{n}.

The transitivity and the unique existence of parents follow from a similar argument used before. However, the existence of recurrent children does not work for curves in general; The level-nn children of level-mm bones constituting γm\gamma^{m} may not be joined as they are at level-mm.

Definition 4.10 (Genealogical sequence of non-expanded curves).

Let \mathcal{R} be a finite subdivision rule. Let II denote a closed interval [k,l][k,l], (,k](-\infty,k], [k,)[k,\infty), or (,)(-\infty,\infty) for k<lk<l\in\mathbb{Z}. A sequence of level-nn non-expanded curves {γn:In(S)}n0\{\gamma^{n}:I\to\mathcal{R}^{n}(S_{\mathcal{R}})\}_{n\geq 0} is genealogical if γn+1\gamma^{n+1} is a child of γn\gamma^{n} for every n0n\geq 0.

5. Levy cycle and genealogical sequence of homotopically infinite curves

The purpose of this section is to prove the following theorem.

Theorem 5.1.

Let \mathcal{R} be a finite subdivision rule and AVert(S)A\subset\operatorname{Vert}(S_{\mathcal{R}}) be a set of marked points. Suppose that the subdivision map f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\to S_{\mathcal{R}} is not doubly covered by a torus endomorphism. Then f:(S2,A)f:(S^{2},A)\righttoleftarrow has a Levy cycle if and only if there is a genealogical sequence of non-expanded recurrent bi-infinite curves {γn:(,)n(S)}\{\gamma^{n}:(-\infty,\infty)\to\mathcal{R}^{n}(S_{\mathcal{R}})\} such that each γn\gamma^{n} is homotopically infinite with respect to a hyperbolic orbisphere structure ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}} (hence with respect to any hyperbolic orbisphere structure because the definition of Levy cycles is independent of the choice of orbisphere structures).

The “ only if ” direction is not hard. We can use a Levy cycle to construct the desired genealogical sequence of non-expanded curves. The other direction, however, is non-trivial. Even if we have a genealogical sequence of non-expanded curves, it is difficult to explicitly find a Levy cycle. We prove the existence of a Levy cycle in a non-constructive way using an algebraic machinery, called self-similar groups [Nek05]. We use the term “orbisphere bisets” rather than self-similar groups in order to be consistent with our main reference [BD18].

5.1. Contracting orbisphere bisets

Let AA be a finite subset of the sphere S2S^{2}. An orbisphere structure on (S2,A)(S^{2},A) is an order function ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}}. We say that ord\mathrm{ord} is an orbisphere structure of a post-critically finite branched covering f:(S2,A)f:(S^{2},A)\righttoleftarrow if it satisfies

  1. (1)

    ord(a)degf(a)|ord(f(a))\mathrm{ord}(a)\cdot\deg_{f}(a)\leavevmode\nobreak\ |\leavevmode\nobreak\ \mathrm{ord}(f(a)) for every aS2a\in S^{2} where ord(a)=1\mathrm{ord}(a)=1 for aAa\notin A, and

  2. (2)

    ord(a)=\mathrm{ord}(a)=\infty only if aAa\in A is a Fatou point.

In (1), \infty is considered as a multiple of any integer or \infty itself. It follows that ord(a)=\mathrm{ord}(a)=\infty for every aa in a periodic cycle containing a critical point. The triple (S2,A,ord)(S^{2},A,\mathrm{ord}) is called an orbisphere.

The orbisphere group π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord}) of an orbisphere (S2,A,ord)(S^{2},A,\mathrm{ord}) is defined by

π1(S2,A,ord)=π1(S2A)/{γaord(a)|aAandord(a)}\pi_{1}(S^{2},A,\mathrm{ord})=\pi_{1}(S^{2}\setminus A)\left/\,\left<\{\gamma_{a}^{\mathrm{ord}(a)}\leavevmode\nobreak\ |\leavevmode\nobreak\ a\in A\leavevmode\nobreak\ \mathrm{and}\leavevmode\nobreak\ \mathrm{ord}(a)\neq\infty\}\right>\right.

where γa\gamma_{a} is a peripheral loop of aAa\in A and γaord(a)=1\gamma_{a}^{\mathrm{ord}(a)}=1 if ord(a)=\mathrm{ord}(a)=\infty.

Remark 5.2.

When A=PfA=P_{f}, the order of xx, ord(x)\mathrm{ord}(x), is usually defined as the least common multiple of {degfn(y)|yfn(x)forn}\{\deg_{f^{n}}(y)\leavevmode\nobreak\ |\leavevmode\nobreak\ y\in f^{-n}(x)\leavevmode\nobreak\ \mathrm{for}\leavevmode\nobreak\ n\in\mathbb{N}\}. When AA contains periodic points which do not belong to PfP_{f}, however, the least common multiples of their degrees equal to one so that it is not an orbisphere structure we care about. The reason that we require ord(a)>1\mathrm{ord}(a)>1 for every aAa\in A is that if ord(a)=1\mathrm{ord}(a)=1 then γa\gamma_{a} vanishes in π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord}) so that algebraic properties of π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord}) cannot carry any information of the aAa\in A.

The Euler characteristic χ(S2,A,ord)\chi(S^{2},A,\mathrm{ord}) of an orbisphere (S2,A,ord)(S^{2},A,\mathrm{ord}) is defined by

(1) χ(S2,A,ord)=2+aA(1ord(a)1).\chi(S^{2},A,\mathrm{ord})=2+\sum_{a\in A}\left(\frac{1}{\mathrm{ord}(a)}-1\right).

The orbisphere (S2,A,ord)(S^{2},A,\mathrm{ord}) is hyperbolic if χ(S2,A,ord)<0\chi(S^{2},A,\mathrm{ord})<0.

Let pp be a base point of π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord}). Define a set B(f,A,ord)B(f,A,\mathrm{ord}) by

{γ:[0,1]S2A|γ(0)=f(γ(1))=p}/homotopyrelativeto(A,ord).\left\{\gamma:[0,1]\to S^{2}\setminus A\leavevmode\nobreak\ |\leavevmode\nobreak\ \gamma(0)=f(\gamma(1))=p\right\}/\leavevmode\nobreak\ \mathrm{homotopy\leavevmode\nobreak\ relative\leavevmode\nobreak\ to}\leavevmode\nobreak\ (A,\mathrm{ord}).

By the homotopy relative to (A,ord)(A,\mathrm{ord}) we mean a homotopy relative to AA together with one more homotopy condition: For any aAa\in A with ord(a)<\mathrm{ord}(a)<\infty, the ord(a)th\mathrm{ord}(a)^{th} power of the peripheral loop of aa is considered to be homotopically trivial.

There is a natural π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord})-action on B(f,A,ord)B(f,A,\mathrm{ord}) from both left and right. More precisely, for γ1,γ2π1(S2,A,ord)\gamma_{1},\gamma_{2}\in\pi_{1}(S^{2},A,\mathrm{ord}) and for δB(f,A)\delta\in B(f,A), the product γ1δγ2\gamma_{1}\cdot\delta\cdot\gamma_{2} is the concatenation of γ1\gamma_{1}, δ\delta, and the lift of γ2\gamma_{2} through ff starting at the endpoint of δ\delta, in order. The left action is free, and the right action is transitive. The set B(f,A)B(f,A) equipped with the left and right actions is called the orbisphere biset of (S2,A,ord)(S^{2},A,\mathrm{ord}). If an orbisphere structure ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}} is given, we implicitly assume that B(f,A,ord)B(f,A,\mathrm{ord}) has the left and right π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord})-actions. When an orbisphere (S2,A,ord)(S^{2},A,\mathrm{ord}) is understood in the context, we simply write B(f)B(f) for B(f,A,ord)B(f,A,\mathrm{ord}).

Caution 5.3.

There are two conventions depending on whether you concatenate curves from right to left or from left to right in the operation of orbisphere group. Many documents, including [Nek05], follow the “from right to left” convention, but we will follow the “from left to right” convention for the sake of convenience in citing [BD18]. Thus a biset has a free left action and a transitive right action, which is opposite to a bimodule in [Nek05].

A tensor square B(f)B(f)B(f)\otimes B(f) can be defined in two different ways. Topologically, the tensor product δ1δ2\delta_{1}\otimes\delta_{2} for δ1,δ2B(f)\delta_{1},\delta_{2}\in B(f) is defined as a concatenation of δ1\delta_{1} and the lift of δ2\delta_{2} starting at the endpoint of δ1\delta_{1}. Algebraically, it is a quotient of B(f)×B(f)B(f)\times B(f) by the relation (δ1γ)δ2=δ1(γδ2)(\delta_{1}\cdot\gamma)\otimes\delta_{2}=\delta_{1}\otimes(\gamma\cdot\delta_{2}). The left and right actions naturally extend to B(f)B(f)B(f)\otimes B(f). Similarly, B(f)nB(f)^{\otimes n} has a left free and a right transitive π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord})-actions for any n1n\geq 1.

A basis XX for B(f)B(f) is a collection of representatives of left orbits of the biset B(f)B(f). Its cardinality |X||X| is the same as the degree of ff. For any n1n\geq 1, the tensor power XnX^{\otimes n} of XX is a basis for B(f)nB(f)^{\otimes n}. Topologically, a basis is a choice of curves from the base point pp of π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord}) to the dd preimages f1(p)={p1,p2,,pd}f^{-1}(p)=\{p_{1},p_{2},\dots,p_{d}\} where d=deg(f)d=\deg(f). Let δi\delta_{i} be a curve from pp to pip_{i} for i[1,d]i\in[1,d]_{\mathbb{Z}}. Then {δ1,δ2,,δd}\{\delta_{1},\delta_{2},\dots,\delta_{d}\} be a basis for B(f)B(f), and every basis of B(f)B(f) is of this form. Fix n1n\geq 1. Let i1,i2,,in[1,d]i_{1},i_{2},\dots,i_{n}\in[1,d]_{\mathbb{Z}}. We simply write

δi1i2in:=δi1δi2δin,\delta_{i_{1}i_{2}\dots i_{n}}:=\delta_{i_{1}}\otimes\delta_{i_{2}}\otimes\dots\otimes\delta_{i_{n}},

which gives a bijection ([1,d])nXn([1,d]_{\mathbb{Z}})^{n}\leftrightarrow X^{\otimes n}.

Definition 5.4 (Contracting biset and nucleus).

Let f:(S2,A)f:(S^{2},A)\righttoleftarrow be a marked post-critically finite branched covering and ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}} be an orbisphere structure. Let XX be a basis for B(f)B(f). The orbisphere biset B(f)B(f) is contracting if there exists a finite subset 𝒩π1(S2,A,ord)\mathcal{N}\subset\pi_{1}(S^{2},A,\mathrm{ord}) satisfying the following: For every gπ1(S2,A,ord)g\in\pi_{1}(S^{2},A,\mathrm{ord}), the inclusion Xng𝒩XnX^{\otimes n}\cdot g\subset\mathcal{N}\cdot X^{\otimes n} holds for every sufficiently large n>0n>0. The minimal 𝒩\mathcal{N} satisfying this property is the nucleus of (B(f),X)(B(f),X).

The contracting property does not depend on the choice of basis [Nek05, Corollary 2.11.7], but the nucleus does. See Remark 5.8 for the independence of the choice of orbisphere structures.

Definition 5.5 (Böttcher expanding map and Local rigidity).

Let f:(S2,A)f:(S^{2},A)\righttoleftarrow be a post-critically finite branched covering and ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}} be an orbisphere structure. Denote by AA^{\infty} the subset of AA consisting of aAa\in A with ord(a)=\mathrm{ord}(a)=\infty. Then f:(S2,A)f:(S^{2},A)\righttoleftarrow is Böttcher (metrically) expanding if there is a length metric μ\mu on S2AS^{2}\setminus A^{\infty}, satisfying the following conditions.

  • For every rectifiable curve γ:[0,1]S2A\gamma:[0,1]\to S^{2}\setminus A^{\infty}, the length of any lift of γ\gamma through ff is strictly less the length of γ\gamma, and

  • (Local rigidity near critical cycles) For every periodic point aAa\in A^{\infty}, the first return map of ff near aa is locally topologically conjugate to zzdega(fn)z\mapsto z^{\deg_{a}(f^{n})}, where nn is the period of aa.

Every Böttcher expanding map f:(S2,A)f:(S^{2},A)\righttoleftarrow also has the Fatou set and the Julia set, which have similar properties of the Fatou and Julia sets of rational maps, see [BD18].

A post-critically finite rational map ff is Böttcher expanding since it has the Böttcher coordinates and enjoys the Schwarz lemma about the conformal metric. The next theorem, which follows from [BD18, Theorem A, Corollary 1.2], is an analogue of Thurston’s characterization and rigidity.

Theorem 5.6 ([BD18, Theorem A, Corollary 1.2]).

Let f:(S2,A)f:(S^{2},A)\righttoleftarrow be a post-critically finite branched covering which is not doubly covered by a torus endomorphism and ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}} be an orbisphere structure. Then the following are equivalent

  • (1)

    f:(S2,A)f:(S^{2},A)\righttoleftarrow is combinatorially equivalent to a Böttcher expanding map.

  • (2)

    The orbisphere biset B(f,A,ord)B(f,A,\mathrm{ord}) is contracting.

  • (3)

    f:(S2,A)f:(S^{2},A)\righttoleftarrow has degree>1 and does not have a Levy-cycle.

Moreover, if exists, the Böttcher expanding map is unique in the combinatorial equivalent class up to topological conjugacy.

Remark 5.7.

In [BD18], the orbisphere structure used in Theorem A is required that ord(a)=\mathrm{ord}(a)=\infty if and only if aa is a periodic Fatou point, which is a little stronger than the definition of orbisphere structures in this paper. But this slight generalization follows almost immediately.

Remark 5.8.

The definition of B(f,A,ord)B(f,A,\mathrm{ord}) depends on the orbisphere structure ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}}, but the definition of Levy cycles of f:(S2,A)f:(S^{2},A)\righttoleftarrow doesn’t. Hence Theorem 5.6 implies that whether or not B(f,A,ord)B(f,A,\mathrm{ord}) is contracting is also independent of the choice of orbisphere structure.

5.2. Semi-conjugacy to Böttcher expanding maps

The idea of semi-conjugacy was introduced by Rees [Ree92] and Shishikura [Shi00] to show that, for any mateable pair of post-critically finite polynomials, the topological mating is topologically conjugate to the corresponding rational map. Then the idea was further developed by Cui-Peng-Tan [CPT12] to a form that can be applied for not only matings but also general post-critically finite branched coverings and rational maps. We slightly further generalize the theorem of Cui-Peng-Tan by applying Bartholdi-Dudko’s recent work on Böttcher expanding maps [BD18].

The next theorem is a generalization of [CPT12, Theorem 1.1, Corollary 1.2] replacing rational maps by Böttcher expanding maps.

Theorem 5.9 (Semi-conjugacies to Böttcher expanding maps).

Let f:(S2,A)f:(S^{2},A)\righttoleftarrow be a post-critically finite branched covering which is locally rigid near critical cycles. Suppose ff is combinatorially equivalent to a Böttcher expanding map F:(S2,B)F:(S^{2},B)\righttoleftarrow. Let F\mathcal{F}_{F} and 𝒥F\mathcal{J}_{F} denote the Fatou and the Julia sets of FF. Then there exists a semi-conjugacy h:(S2,A)(S2,B)h:(S^{2},A)\to(S^{2},B) from ff to FF, i.e., hf=Fhh\circ f=F\circ h, such that the following properties are satisfied.

  • h1(w)h^{-1}(w) is a singleton for wFw\in\mathcal{F}_{F} and a full continuum for w𝒥Fw\in\mathcal{J}_{F}.

  • For x,yS2x,y\in S^{2} with F(x)=yF(x)=y, the set h1(x)h^{-1}(x) is a connected component of f1(h1(y))f^{-1}(h^{-1}(y)). Moreover, the degree of the map f:h1(x)h1(y)f:h^{-1}(x)\to h^{-1}(y) is equal to degx(F)\deg_{x}(F); more precisely, for every wh1(y)w\in h^{-1}(y) we have

    zh1(x)f1(w)degzf=degx(F).\sum\limits_{z\in h^{-1}(x)\cap f^{-1}(w)}\deg_{z}f=\deg_{x}(F).
  • If ES2E\subset S^{2} is a continuum, then h1(E)h^{-1}(E) is a continuum.

  • f(h1(E))=h1(F(E))f(h^{-1}(E))=h^{-1}(F(E)) for every ES2E\subset S^{2}.

  • f1(E^)=f1(E)^f^{-1}(\widehat{E})=\widehat{f^{-1}(E)} for every ES2E\subset S^{2}, where E^:=h1(h(E))\widehat{E}:=h^{-1}(h(E)).

Proof.

In [CPT12], the complex structure of the Riemann sphere was used for two purposes: (i) the conformal metric is expanding, and (ii) there are Böttcher coordinates near critical cycles. Since Böttcher expanding maps also have these two properties, the proof in [CPT12] still works for this setting. For example,

  • In [CPT12], they use post-critically finite branched coverings on the Riemann sphere ^\hat{\mathbb{C}} that are holomorphic near critical cycles. Given a post-critically finite branched covering (on the topological sphere) which is locally rigid near critical cycles, we may define a holomorphic structure on the sphere so that the branched covering is holomorphic near critical cycles.

  • The orbifold metric in [CPT12, Section 2] can be replaced by the Riemannian orbifold metric in [BD18]. Then we still have the expansion property of homotopic lengths of paths.

Definition 5.10 (Homotopic length).

Let (X,μ)(X,\mu) be a metric space and γ\gamma be a curve joining xx to yy. Then the homotopic length lμ([γ])l_{\mu}([\gamma]) is the infimum of the lengths of rectifiable curves that joins xx to yy and homotopic to γ\gamma relative to {x,y}\{x,y\}.

Lemma 5.11.

Let f:(S2,A)f:(S^{2},A)\righttoleftarrow be a post-critically finite branched covering of degree d2d\geq 2 which is not doubly covered by a torus endomorphism. Suppose ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}} is an orbisphere structure and XX be a basis for the biset B(f)B(f). Suppose that ff does not have a Levy cycle such that there exists a semi-conjugacy h:(S2,A)(S2,B)h:(S^{2},A)\to(S^{2},B) where F:(S2,B)F:(S^{2},B)\righttoleftarrow is a Böttcher expanding map, expanding about a metric μ\mu, that is combinatorially equivalent to ff. Then there exists C>0C>0 such that lμ([h(w)])<Cl_{\mu}([h(w)])<C for every n>0n>0 and wXnw\in X^{\otimes n}. Here ww is considered as a curve joining the base point pp of π1(X,A,ord)\pi_{1}(X,A,\mathrm{ord}) to a point in the preimage fn(p)f^{-n}(p), as described in Section 5.1.

Proof.

Since the metric μ\mu blows up near marked points of infinite order, we should take a compact subset away the points of infinite order. Let B=h(A)B^{\infty}=h(A^{\infty}) where AA^{\infty} is the subset of AA consisting of elements having infinite order. There exists a small neighborhood UU of BB^{\infty} such that for M:=S2UM:=S^{2}\setminus U and M=f1(M)M^{\prime}=f^{-1}(M) we have MMM^{\prime}\subset M and f:MMf:M^{\prime}\to M being a branched covering which has a uniform expanding constant λ>1\lambda>1 in the following sense: For every curve γM\gamma\subset M and any of its lifting γ\gamma^{\prime} through ff, we have

λlμ([γ])<lμ([γ]).\lambda\cdot l_{\mu}([\gamma^{\prime}])<l_{\mu}([\gamma]).

Let X={δ1,δ2,,δd}X=\{\delta_{1},\delta_{2},\dots,\delta_{d}\} where each δi\delta_{i} joins the base point pp of π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord}) to one of the dd preimages f1(p)f^{-1}(p). Define D>0D>0 by

D=max1jdlμ([h(δj)]).D=\max_{1\leq j\leq d}l_{\mu}([h(\delta_{j})]).

Let w=δi1i2inXnw=\delta_{i_{1}i_{2}\dots i_{n}}\in X^{\otimes n} where il{1,2,,d}i_{l}\in\{1,2,\dots,d\}. Then ww is the concatenation of δi1\delta_{i_{1}}, a lift of δi2\delta_{i_{2}} through ff, a lift of δi3\delta_{i_{3}} through f2f^{2}, and so one. Every curve δi\delta_{i} and its any lifting can be contained in MM up to homotopy. Hence we have

lμ([h(w)]<D(1+1λ+1λ2+)=Dλλ1.l_{\mu}([h(w)]<D\cdot\left(1+\frac{1}{\lambda}+\frac{1}{\lambda^{2}}+\cdots\right)=D\cdot\frac{\lambda}{\lambda-1}.

5.3. Homotopically infinite non-expanded curves and Levy cycles

Definition 5.12 (Homotopically infinite curves).

Let (S2,A,ord)(S^{2},A,\mathrm{ord}) be a hyperbolic orbisphere and p:𝔻S2Ap:\mathbb{D}\to S^{2}\setminus A^{\infty} is the orbifold universal covering map. A closed curve γ:[0,1]S2A\gamma:[0,1]\to S^{2}\setminus A is homotopically infinite with respect to ord\mathrm{ord} if for a connected component γ~\widetilde{\gamma} of p1(γ)p^{-1}(\gamma), both ends of γ~\widetilde{\gamma} have a limit point on the boundary 𝔻\partial\mathbb{D}. A half-infinite curve γ:[0,)S2A\gamma:[0,\infty)\to S^{2}\setminus A (resp. bi-infinite curve γ:(,)S2A\gamma:(-\infty,\infty)\to S^{2}\setminus A) is homotopically infinte with respect to ord\mathrm{ord} if the end (resp. both ends) of its lift γ~\widetilde{\gamma} has a limit point.

The next proposition is immediate from standard properties of the hyperbolic geometry.

Proposition 5.13.

Let (S2,A,ord)(S^{2},A,\mathrm{ord}) be a hyperbolic orbisphere. A closed curve γ\gamma is homotopically infinite if and only if γ\gamma is neither homotopically trivial in S2AS^{2}\setminus A nor homotopic relative to AA to some iterate of the peripheral loop of aAa\in A with ord(a)<\mathrm{ord}(a)<\infty.

Each of the following two propositions is each direction of the equivalence in Theorem 5.1. We split them because the ideas of the proofs are quite different.

Proposition 5.14.

Let \mathcal{R} be a finite subdivision rule and AVert(S)A\subset\operatorname{Vert}(S_{\mathcal{R}}) be a set of marked points. Suppose that the subdivision map f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\to S_{\mathcal{R}} is not doubly covered by a torus endomorphism. If f:(S2,A)f:(S^{2},A)\righttoleftarrow has a Levy cycle, then there is a genealogical sequence of non-expanded closed curves {γn:In(S)}n0\{\gamma^{n}:I\to\mathcal{R}^{n}(S_{\mathcal{R}})\}_{n\geq 0} that are recurrent and homotopically infinite with respect to any hyperbolic orbisphere structure ord:A[0,]\mathrm{ord}:A\to[0,\infty]_{\mathbb{Z}}. Moreover, by iterating travelling along the closed curves, we may assume that each γn\gamma^{n} is a bi-infinite curve.

Proof.

Assume there exists a Levy cycle, i.e., there are an integer p>0p>0 and an essential simple closed curve γ\gamma of (S2,A)(S^{2},A) such that a connected component γ\gamma^{\prime} of fp(γ)f^{-p}(\gamma) is isotopic to γ\gamma relative to AA and deg(fp|γ)=1\deg(f^{p}|_{\gamma^{\prime}})=1. We may assume γ\gamma is SS_{\mathcal{R}}-taut so that γ\gamma^{\prime} is p(S)\mathcal{R}^{p}(S_{\mathcal{R}})-taut.

Claim: We may assume that γ\gamma and γ\gamma^{\prime} are SS_{\mathcal{R}}-combinatorially equivalent.

Proof of Claim.

For every k1k\geq 1, fkp(γ)f^{-kp}(\gamma) has a connected component γkp\gamma_{kp} that is isotopic to γ\gamma relative to AA and deg(fkp|γkp)=1\deg(f^{kp}|_{\gamma_{kp}})=1. For every k1k\geq 1, it follows from Proposition 2.6 that l0(γkp)lkp(γkp)l_{0}(\gamma_{kp})\leq l_{kp}(\gamma_{kp}) and from deg(fkp|γkp)=1\deg(f^{kp}|_{\gamma_{kp}})=1 that lkp(γkp)=l0(γ)l_{kp}(\gamma_{kp})=l_{0}(\gamma), where ln()l_{n}(\cdot) means ln(S)()l_{\mathcal{R}^{n}(S_{\mathcal{R}})}(\cdot). By Lemma 2.4, there exist k1>k2>0k_{1}>k_{2}>0 such that γk1p\gamma_{k_{1}p} and γk2p\gamma_{k_{2}p} are combinatorially equivalent relative to SS_{\mathcal{R}}. Then we can replace γ\gamma by γk2p\gamma_{k_{2}p} and pp by (k1k2)p(k_{1}-k_{2})p. ∎

It follows from the claim that we can parametrize γ\gamma and γ\gamma^{\prime} such that γ:Ip(S)\gamma^{\prime}:I\to\mathcal{R}^{p}(S_{\mathcal{R}}) is a level-pp non-expanded closed curve and γ:IS\gamma:I\to S_{\mathcal{R}} is the level-0 parent of γ\gamma^{\prime} for I=[0,l]I=[0,l] for some l>0l\in\mathbb{Z}_{>0}. Being essential relative to AA, γ\gamma and γ\gamma^{\prime} are, in particular, homotopically infinite relative to any hyperbolic orbisphere structure.

Let γ0:=γ\gamma^{0}:=\gamma and γp:=γ\gamma^{p}:=\gamma^{\prime}. By lifting an isotopy between γ0\gamma^{0} and γp\gamma^{p} through fpf^{p}, we have an isotopy from γp\gamma^{p} to a level-2p2p non-expanded curve γ2p:I2p(S)\gamma^{2p}:I\to\mathcal{R}^{2p}(S_{\mathcal{R}}) such that γp\gamma^{p} is the level-pp parent of γ2p\gamma^{2p}. This way, we obtain a sequence of level-kpkp non-expanded curves {γkp:Ikp(S)}k0\{\gamma^{kp}:I\to\mathcal{R}^{kp}(S_{\mathcal{R}})\}_{k\geq 0} such that (1) γkp\gamma^{kp} is the level-kpkp parent of γ(k+1)p\gamma^{(k+1)p} for every k0k\geq 0 and (2) fp:γ(k+1)pγkpf^{p}:\gamma^{(k+1)p}\to\gamma^{kp} is a homeomorphism. If we identify γp\gamma^{p} with γ0\gamma^{0} via an isotopy preserving the 1-skeleton of SS_{\mathcal{R}}, the map fp:γpγ0f^{p}:\gamma^{p}\to\gamma^{0} can be considered as a rotation of a circle of length ll by an integer. Hence, there exists k0>0k_{0}>0 such that for every n>0n>0 the level-nk0pnk_{0}p bone γk0np([i,i+1])\gamma^{k_{0}np}([i,i+1]) is mapped to γ0([i,i+1])\gamma^{0}([i,i+1]) by fk0npf^{k_{0}np}, which implies that γkp\gamma^{kp} is recurrent for every k0k\geq 0. For every m>0m>0 that is not a multiple of pp, we define γm\gamma^{m} as the level-mm parent of γkp\gamma^{kp} for some k>0k>0 with kp>mkp>m, which is well-defined up to m(S)\mathcal{R}^{m}(S_{\mathcal{R}})-combinatorial equivalence.

Since each γm\gamma^{m} is homotopic to an essential simple closed curve of (S2,A)(S^{2},A), it is homotopically infinite with respect to any hyperbolic orbisphere structure. ∎

Proposition 5.15.

Let \mathcal{R} be a finite subdivision rule and AVert(S)A\subset\operatorname{Vert}(S_{\mathcal{R}}) be a set of marked points. Suppose that the subdivision map f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\to S_{\mathcal{R}} is not doubly covered by a torus endomorphism. Let ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}} be a hyperbolic orbisphere structure. If there is a genealogical sequence of non-expanded bi-infinite curves {γn:(,)n(S)}n0\{\gamma^{n}:(-\infty,\infty)\to\mathcal{R}^{n}(S_{\mathcal{R}})\}_{n\geq 0} that are recurrent and homotopically infinite with respect to ord\mathrm{ord}, then f:(S2,A)f:(S^{2},A)\righttoleftarrow has a Levy cycle.

Proof.

Suppose that f:(S2,A)f:(S^{2},A)\righttoleftarrow does not have a Levy cycle. Then for any basis XX for the biset B(f)B(f), there is a nucleus 𝒩\mathcal{N}, which is a finite set. For every k>0k>0, we use the sequence of finite restrictions {γn|[0,k]}n0\{\gamma^{n}|_{[0,k]}\}_{n\geq 0} to obtain an element hk𝒩h_{k}\in\mathcal{N} such that hk\|h_{k}\|\to\infty. Here \|\cdot\| is a distance on π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord}) with respect to a generating set, which does not need to be specified. Then the nucleus 𝒩\mathcal{N} has infinitely many elements, which contradicts to the assumption that the biset is contracting.

Recall that we use pp for the base point of π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord}), each element of a basis XX for B(f)B(f) is a curve from pp to one of its ff-preimages f1(p)f^{-1}(p), and an element wXnw\in X^{\otimes n} is a concatenation of curves connecting an fif^{i}-preimage to fi+1f^{i+1}-preimage that are liftings of elements in XX.

Step 1: Construction of hkh_{k}. Fix k>0k>0. There is an infinite sequence n1<n2<n_{1}<n_{2}<\cdots, which depends on kk, so that γni([0,k])\gamma^{n_{i}}([0,k]) consists of the bones of the bands of the same types, i.e., there exists level-0 bands b0,b1,bk1b_{0},b_{1},\dots b_{k-1} (possibly repeated) such that for every j[0,k1]j\in[0,k-1], γni([j,j+1])\gamma^{n_{i}}([j,j+1]) is the bone of a level-nin_{i} band of type bjb_{j}, which is independent of i>0i>0.

For every level-0 edge ee, we fix a point meint(e)m_{e}\in\mathrm{int}(e) and call it the midpoint of ee. We assume that a bone of a level-0 band is chosen in the homotopy class in such a way that their endpoints are the midpoints of level-0 edges. For every level-0 edge ee, we also choose a path δe\delta_{e} from pp to mem_{e}. Then, for each level-0 band b=(t;e1,e2)b=(t;e_{1},e_{2}), we can assign an element

gbi:=δe1bone(bi)δ¯e2π1(S2,A,ord),g_{b_{i}}:=\delta_{e_{1}}\cdot bone(b_{i})\cdot\overline{\delta}_{e_{2}}\in\pi_{1}(S^{2},A,\mathrm{ord}),

where the overline ¯\overline{\,\cdot\,} means the reverse of the orientation of a curve and bone(bi)bone(b_{i}) means the bone of a band bib_{i}.

For gπ1(S,A,ord)g\in\pi_{1}(S,A,\mathrm{ord}), we define N(g)π1(S,A,ord)N(g)\subset\pi_{1}(S,A,\mathrm{ord}) by the collection of elements hh with the following property: For infinitely many n>0n>0 there exist v,wXnv,w\in X^{\otimes n} such that hv=wgh\cdot v=w\cdot g. We remark that N(g)𝒩N(g)\subset\mathcal{N}.

Refer to caption
δe\delta_{e}
δe\delta_{e^{\prime}}
wniw_{n_{i}}
vniv_{n_{i}}
δ~e\tilde{\delta}_{e}
δ~e\tilde{\delta}_{e^{\prime}}
Refer to caption
g~k\tilde{g}_{k}
Refer to caption
pp
pp
Figure 6. The left figure is gkg_{k} drawn in SS_{\mathcal{R}}, and the right figure is hk,nih_{k,n_{i}} drawn in ni(S)\mathcal{R}^{n_{i}}(S_{\mathcal{R}}). The bold line segments are portions of 11-skeletons of SS_{\mathcal{R}} and ni(S)\mathcal{R}^{n_{i}}(S_{\mathcal{R}}).

Claim: There exists C>0C>0, independent of kk, such that N(gk:=gb1gbk)N(g_{k}:=g_{b_{1}}\cdots g_{b_{k}}) contains at least one element hkh_{k} with d(hk,gk)<Cd(h_{k},g_{k})<C.

Proof of Claim.

Recall that for every i>0i>0 and j[0,k1]j\in[0,k-1]_{\mathbb{Z}}, γni([j,j+1])\gamma^{n_{i}}([j,j+1]) is the bone of a level-nin_{i} band of type bjb_{j}, which is independent of ii. Let ee and ee^{\prime} be level-0 edges such that their midpoints mem_{e} and mem_{e^{\prime}} are the endpoints of γ0([0,k])\gamma^{0}([0,k]).

Let wniXniw_{n_{i}}\in X^{\otimes n_{i}}, which will be specified soon. Let vniXniv_{n_{i}}\in X^{\otimes n_{i}} and hk,niπ1(S2,A,ord)h_{k,n_{i}}\in\pi_{1}(S^{2},A,\mathrm{ord}) be defined by

hk,nivni=wnigk.h_{k,n_{i}}\cdot v_{n_{i}}=w_{n_{i}}\cdot g_{k}.

Let g~k\tilde{g}_{k} be the lift of gkg_{k} through fnif^{n_{i}} starting from the terminal point of wniw_{n_{i}}. Let δ~e\tilde{\delta}_{e} and δ~e\tilde{\delta}_{e^{\prime}} be the parts of g~k\tilde{g}_{k} where δe\delta_{e} and δe\delta_{e^{\prime}} are lifted. We specify wniw_{n_{i}} as an element of XniX^{\otimes n_{i}} satisfying the following: the curve gkg_{k} with δe\delta_{e} and δe\delta_{e^{\prime}} being truncated is SS_{\mathcal{R}}-combinatorially equivalent to g~k\tilde{g}_{k} with δ~e\tilde{\delta}_{e} and δ~e\tilde{\delta}_{e^{\prime}} being truncated. See Figure 6.

Then hk,nih_{k,n_{i}} and gkg_{k} differ by the pre- and post-composition with loops wniδ~eδ¯ew_{n_{i}}\cdot\tilde{\delta}_{e}\cdot\overline{\delta}_{e} and δ¯eδev¯ni\overline{\delta}_{e^{\prime}}\cdot\delta_{e^{\prime}}\cdot\overline{v}_{n_{i}}. We have (1) a uniform upper bound on the homotopic length of vniv_{n_{i}} and wniw_{n_{i}} (in the projection to the Böttcher expanding map) by Lemma 5.11, (2) a uniform upper bound on the intersection between δe\delta_{e} and S(1)S_{\mathcal{R}}^{(1)}, and (3) the upper bound in (2) is also an upper bound of the intersection between any level-nn lift δ~e\tilde{\delta}_{e} and n(S)(1)\mathcal{R}^{n}(S_{\mathcal{R}})^{(1)}. Therefore, we have wniδ~eδ¯e,δ¯eδev¯ni<C/2\|w_{n_{i}}\cdot\tilde{\delta}_{e}\cdot\overline{\delta}_{e}\|,\|\overline{\delta}_{e^{\prime}}\cdot\delta_{e^{\prime}}\cdot\overline{v}_{n_{i}}\|<C/2 for some CC so that d(gk,hk,ni)<Cd(g_{k},h_{k,n_{i}})<C. Since there are only finitely many elements of π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord}) within the distance CC from gkg_{k}, there exists hkh_{k} such that hk=hk,nih_{k}=h_{k,n_{i}} for infinitely many ii’s. ∎

Step 2: Proof of hk\|h_{k}\|\to\infty. Let gkg_{k} and hkh_{k} be as defined in Step 1. Since d(hk,gk)<Cd(h_{k},g_{k})<C, it suffices to show gk\|g_{k}\|\to\infty as kk tends to \infty.

Let e0e_{0} and eke_{k} be the level-0 edges whose midpoints are the endpoints of γ0([0,k])\gamma^{0}([0,k]) so that gk=δe0γ0([0,k])δ¯ekg_{k}=\delta_{e_{0}}\cdot\gamma^{0}([0,k])\cdot\overline{\delta}_{e_{k}}. Then gk\|g_{k}\|\to\infty follows from the condition that γ0\gamma^{0} is homotopically infinite.

6. Non-expanding spines

From Theorem 5.1, we know that the existence of a Levy cycle is equivalent to the existence of a genealogical sequence of homotopically infinite recurrent non-expanded curves. Then, how can we detect the existence such a sequence? The direct search for the genealogical sequence could be more complicated then the search for the Levy cycle. One motivation for non-expanding spines is to have a simpler object with which we can efficiently detect the genealogical sequence.

Since recurrent non-expanded level-nn curves are concatenations of level-nn recurrent bones, i.e., the bones of level-nn recurrent bands, it is natural to consider the union of level-nn recurrent bones. The level-nn non-expanding spine NnN^{n} is, roughly speaking, the union of level-nn recurrent bones equipped with a natural train-track structure.

Motivation for the use of train-tracks. Let us see Figure 7. Let tt denote the hexagonal tile and e1,e2,e3e_{1},e_{2},e_{3} denote three of its boundary edges. Suppose that b1=(t;e1,e2)b_{1}=(t;e_{1},e_{2}) and b2=(t;e1,e3)b_{2}=(t;e_{1},e_{3}) are level-0 recurrent bands so that each has two level-1 recurrent subbands of type b1b_{1} and b2b_{2}. Then each level-1 recurrent subband of type b1b_{1} or b2b_{2} has two level-22 recurrent subbands of type b1b_{1} and b2b_{2}. Assume that we draw the bones of these bands. At level-0, we have two curves each of which joins e1e_{1} to e2e_{2} or e3e_{3}. As a “collection” of these two curves, we might consider a tripod whose leaves are on e1,e2,e_{1},e_{2}, and e3e_{3}. However, the tripod contains an unintended curve which joins e2e_{2} to e3e_{3}. To exclude such a curve, we use the idea of train-tracks, which makes the curve joining e2e_{2} and e3e_{3} illegal.

Refer to caption
e1e_{1}
e2e_{2}
e3e_{3}
e1e_{1}
e1e_{1}
e2e_{2}
e3e_{3}
e2e_{2}
e3e_{3}
e1e_{1}
Refer to caption
e2e_{2}
e3e_{3}
e3e_{3}
e2e_{2}
Figure 7.

6.1. Train-track

Let GG be a graph and vGv\in G be a point, which could be a vertex or a point in the interior of an edge. A direction at vv is a germ of continuous curves starting from vv. The number of directions at vv is equal to the number of connected components of U{v}U\setminus\{v\} where UU is a sufficiently small neighborhood of vv. Denote the set of directions at vv by DvD_{v}.

Definition 6.1 (Train-tracks and gates).

Let GG be a finite graph. For any vertex vv of GG. A train-track structure τ\tau on GG is an assignment of an equivalence relation on DvD_{v} for each vVert(G)v\in\operatorname{Vert}(G). A train-track is a finite graph GG equipped with a train-track structure τ\tau and denoted by (G,τ)(G,\tau). For any vGVert(G)v\in G\setminus\operatorname{Vert}(G), DvD_{v} has two directions and we define each equivalence class of DvD_{v} to have each direction. Below is a list of definitions regarding train-tracks.

  • Each equivalence class of DvD_{v} is called a gate at vv.

  • A train path is an oriented curve γ\gamma in GG such that at every vertex vv, the gate through which γ\gamma comes to vv is different from the gate through which γ\gamma goes out.

Definition 6.2 (Train-track map).

Let T1=(G1,τ1)T_{1}=(G_{1},\tau_{1}) and T2=(G2,τ2)T_{2}=(G_{2},\tau_{2}) be train-tracks. A train-track map ϕ:T1T2\phi:T_{1}\to T_{2} is a continuous map ϕ:G1G2\phi:G_{1}\to G_{2} that is locally injective on each edge such that for every vG1v\in G_{1} and d1,d2Dvd_{1},d_{2}\in D_{v}, ϕ(d1)\phi(d_{1}) and ϕ(d2)\phi(d_{2}) are in the same gate at ϕ(v)\phi(v) if d1d_{1} and d2d_{2} are in the same gate.

Definition 6.3 (Homotopy relative to (S,X)(\partial S,X)).

Let f,g:(S,S,X)(S,S,X)f,g:(S,\partial S,X)\to(S,\partial S,X) be two continuous maps. We say that ff and gg are homotopic (resp. isotopic) relative to (S,X)(\partial S,X) if there is a homotopy (resp. isotopy) {Ht:SS}t[0,1]\{H_{t}:S\to S\}_{t\in[0,1]} such that

  • H0=fH_{0}=f and H1=gH_{1}=g,

  • Ht|X:XX=idX:XXH_{t}|_{X}:X\to X=id_{X}:X\to X, and

  • For every t[0,1]t\in[0,1], HtH_{t} sends every connected component of SX\partial S\setminus X to itself.

Definition 6.4 (Graphs properly embedded in (S,S,X)(S,\partial S,X)).

Let SS be a compact surface with a finite set of marked points XX. Some marked points may be on the boundary S\partial S. A graph GSG\subset S is properly embedded in (S,S,X)(S,\partial S,X) if (G,G)(G,\partial G) is properly embedded in (S,S)(S,\partial S) and GSXG\subset S\setminus X, where G\partial G is the set of leaves of GG. We say that two graphs GG and HH properly embedded in (S,S,X)(S,\partial S,X) are homotopic (resp. isotopic) if there is an ambient homotopy {Ht:SS}t[0,1]\{H_{t}:S\to S\}_{t\in[0,1]} (resp. isotopy) relative to (S,X)(\partial S,X) such that H0=idS:SSH_{0}=id_{S}:S\to S and H1(G)=HH_{1}(G)=H.

Definition 6.5 (Train-tracks in surfaces).

Let SS be a compact surface possibly with boundary S\partial S and a finite set of marked points XSX\subset S. By a train-track in (S,S,X)(S,\partial S,X), we mean a train-track (G,τ)(G,\tau) of a graph GG satisfying (1) GG is properly embedded in (S,S,X)(S,\partial S,X) and (2) the train-track structure τ\tau is compatible with the planar structure in the following sense: For every vVert(G)v\in\operatorname{Vert}(G), DvD_{v} has a cyclic order defined by a local orientation of SS near vv. Then every gate at vv consists of edges that are consecutive respect to the cycle order. We note that the consecutiveness is independent of the choice of local orientations, thus the definition also works for non-orientable surface SS.

Remark 6.6.

Train-tracks are commonly used to describe complicated curves or foliations. For these purposes, train-tracks are often assumed to have additional properties, such as that every vertex vv has degree 33, the number of gates at each vertex is always two, and the complement of a train-track is hyperbolic, all of which are not assumed in this article. See [PH92].

Definition 6.7 (Carrying between train-tracks).

Let SS be a compact surface with a finite set of marked points XSX\subset S. Let T1T_{1} and T2T_{2} are train-tracks in (S,S,X)(S,\partial S,X). We say that T2T_{2} carries T1T_{1} if there is a train-track map ϕ:T1T2\phi:T_{1}\to T_{2} such that ϕ\phi can be extended to a map ϕ:SS\phi:S\to S that is ambient homotopic relative to (S,X)(\partial S,X) to the identity map. In particular, considering a possibly non-closed curve γ:IS\gamma:I\to S properly embedded in (S,S,X)(S,\partial S,X) as a train-track with no vertex, we can say that a train-track TT carries γ\gamma if γ\gamma is contained in SS up to homotopy in (S,S,X)(S,\partial S,X).

Definition 6.8 (Homotopically infinite train-tracks).

Let (S2,A,ord)(S^{2},A,\mathrm{ord}) be a hyperbolic orbisphere. A train-track TT in (S2,A)(S^{2},A) is homotopically infinite if TT carries a homotopically infinite closed curve with respect to the orbisphere structure ord\mathrm{ord}.

6.2. Decomposition of graph with crossing condition on the unit disk

In this subsection, we investigate a graph theoretic property which will be used to define non-expanding spines.

Let us consider a unit disk in the Euclidean plane and its boundary circle. A chord is an Euclidean line segment joining to point on the circle. We say that two chords intersect if they intersect in the interior of the disk. Similarly, for two sets of chords S1S_{1} and S2S_{2}, we say that S1S_{1} and S2S_{2} intersect if there exist chords s1S1s_{1}\in S_{1} and s2S2s_{2}\in S_{2} so that s1s_{1} and s2s_{2} intersect.

Fix n2n\geq 2 points on the boundary circle CC of a unit disk. There are n(n1)/2n(n-1)/2 different chords joining the nn points. Let SS be a collection of these chords. The collection SS can also be considered as a graph. We abuse notation and use SS to indicate the graph also. A decomposition of a graph GG is a collection of subgraphs G1,G2,,GkG_{1},G_{2},\dots,G_{k} which gives a partition on the set of edges.

Lemma 6.9.

Let v1,v2,,vnv_{1},v_{2},\dots,v_{n} be n2n\geq 2 points on a circle. Let SS be a collection of chords joining pairs of viv_{i}’s. Suppose SS satisfies the following crossing condition:

  • (Crossing condition) If two chords s,sSs,s^{\prime}\in S are intersecting, then all the six chords joining any pairs of four endpoints of s,ss,s^{\prime} are also contained in SS.

Then, as a graph, SS is decomposed into mutually non-intersecting (1) complete graphs with at least four vertices and (2) chords, which can also be considered as complete graphs with two vertices.

Proof.

The condition implies that if SS contains two intersecting chords, then it contains the complete graph of the four vertices. Suppose that a subset SS^{\prime} of SS forms a complete graph. We first show that if there is a chord ss in SSS\setminus S^{\prime} that intersects SS^{\prime}, then SS also contains the complete graph generated by S{s}S^{\prime}\cup\{s\}. Denote by vv and ww the endpoints of ss. Since ss intersects SS^{\prime}, there exists ss^{\prime} that intersects ss. Let vv^{\prime} and ww^{\prime} denote the endpoints of ss^{\prime}. For any vertex uu^{\prime} of SS^{\prime}, as a graph, that is not vv or ww, either uv¯\overline{u^{\prime}v^{\prime}} or uw¯\overline{u^{\prime}w^{\prime}} intersects ss. In particular, by the crossing condition, the chords uv¯\overline{u^{\prime}v} and uw¯\overline{u^{\prime}w} are contained in SS. Since uu^{\prime} was taken arbitrarily, the complete graph with vertex set Vert(S){v,w}\operatorname{Vert}(S^{\prime})\cup\{v,w\} is also contained in SS.

Then every complete graph in SS can be extended until when it does not intersect other chords in SS, which proves the conclusion. ∎

Proposition 6.10.

Let tt be an nn-gon for n2n\geq 2. For a curve γ\gamma properly embedded in (t,t,Vert(t))(t,\partial t,\operatorname{Vert}(t)), we call the boundary edges of tt that contain the endpoints of γ\gamma the side edges of γ\gamma. Let SS be a collection of homotopy classes of properly embedded curves joining different boundary edges. Suppose that SS satisfies the crossing condition:

  • (Crossing condition) If [α],[β]=1\left<[\alpha],[\beta]\right>=1, then SS contains the six homotopy classes of curves connecting any pairs of the four side edges of α\alpha and β\beta.

Then there is a train-track TT properly embedded in (t,t,Vert(t))(t,\partial t,\operatorname{Vert}(t)) such that for any curve γ\gamma properly embedded in (t,t,Vert(t))(t,\partial t,\operatorname{Vert}(t)), γ\gamma is carried by TT if and only if [γ]S[\gamma]\in S.

Proof.

We may consider tt as a closed disk. We also choose a point on each boundary edge and take a representative of a homotopy class of curves properly embedded in tt as a chord joining the chosen points on the edges. Then SS can be considered as a graph, see Figure 8. By Lemma 6.9, SS is decomposed into complete graphs (with at least 4 vertices) and curves that are mutually non-intersecting.

Suppose that SSS^{\prime}\subset S forms a complete graph with k4k\geq 4 vertices in the decomposition. Then we transform SS^{\prime} into a star-like tree TT^{\prime} whose leaves are the kk vertices of SS^{\prime}. We transform all the complete graphs in the decomposition of SS to star-like trees as above, see the second figure in Figure 8. Then we have a graph that is the union of some star-like trees and curves which can intersect only in the boundary of tt.

We define a train-track as follows:

  • (1)

    Let vv be the center of a star-like tree. We define gates at vv in such a way that each gate has only one edge.

  • (2)

    At each intersection in t\partial t, we zip the intersecting curves up a little bit as the third figure in Figure 8. More precisely, assume that kk edges, say e1,e2,eke_{1},e_{2},\dots e_{k} intersect at a boundary point pp. The transformation generates one vertex vv of degree d+1d+1; the vertex vv is incident to the (deformed) kk edges e1,e2,,eke_{1},e_{2},\dots,e_{k}, and to one new edge, say ee which joins pp to vv. There are two gates at vv: {thedirectionalongei|i[1,k]}\{\mathrm{the\leavevmode\nobreak\ direction\leavevmode\nobreak\ along\leavevmode\nobreak\ }e_{i}\leavevmode\nobreak\ |\leavevmode\nobreak\ i\in[1,k]_{\mathbb{Z}}\} and {thedirectionalonge}\{\mathrm{the\leavevmode\nobreak\ direction\leavevmode\nobreak\ along\leavevmode\nobreak\ }e\}.

Refer to caption
Figure 8. Transformation from a graph to a train-track. The dots on the boundary are vertices of a polygon. The graph contains two complete graphs with more than 3 vertices. These graphs are transformed into star-like trees of degree 44 and 55 respectively. We “zip-up” at boundary points to define a train-track structure.

It is immediate from the construction that the train-track satisfies the desired property. ∎

6.3. Non-expanding spines of tiles

Let tt be a level-0 tile of a finite subdivision rule \mathcal{R} and n0n\geq 0. For simplicity, we assume that tt is homeomorphic to a closed 2-disc, i.e., boundary edges are not identified. We will define the level-nn non-expanding spine of tt as a train-track properly embedded in tt which is roughly speaking the union of level-nn recurrent bones in tt. If boundary edges are identified, we first define a train-track in the closed 2-disc 𝐭\mathbf{t}, which is the domain of the characteristic map ϕt:𝐭t\phi_{t}:\mathbf{t}\to t, and then define the non-expanding spine as the image of the train-track by ϕt\phi_{t}.

For a level-0 tile tt and a level-nn band bn=(tn;e1n,e2n)b^{n}=(t^{n};e_{1}^{n},e_{2}^{n}), we say that bnb^{n} is a subband of tt if tntt^{n}\subset t and e1n,e2nte^{n}_{1},e^{n}_{2}\subset\partial t. We say that two level-nn subbands b0nb^{n}_{0} and b1nb^{n}_{1} of a level-0 tile tt intersect if their bones have non-zero intersection number, which must be one, of the homotopy classes of curves properly embedded in (t,n(t),Vert(n(t)))(t,\mathcal{R}^{n}(\partial t),\operatorname{Vert}(\mathcal{R}^{n}(\partial t))). It is immediate that if two level-nn subbands b0nb^{n}_{0} and b1nb^{n}_{1} intersect then they are bands of the same level-nn subtile of tt.

The next lemma, in particular the property (3), implies that the set of level-nn recurrent bones of tt satisfies the crossing condition in the statement of Lemma 6.9 or Proposition 6.10.

Lemma 6.11.

Let tt be a level-0 tile of a finite subdivision rule. For a level-0 recurrent band b0=(t;e0,1,e0,2)b_{0}=(t;e_{0,1},e_{0,2}), let A0:={bi=(t;ei,1,ei,2)|i=1,,k}A^{0}:=\{b_{i}=(t;e_{i,1},e_{i,2})\leavevmode\nobreak\ |\leavevmode\nobreak\ i=1,\dots,k\} be the collection of level-0 recurrent bands that intersect b0b_{0}. Suppose that A0A^{0} is non-empty. Then we have the following properties.

  • (1)

    For any n0n\geq 0 and i[0,k]i\in[0,k]_{\mathbb{Z}}, there is a level-nn subtile tnt^{n} of tt such that each bib_{i} has a unique level-nn subband bin=(tn;ei,1n,ei,2n)b^{n}_{i}=(t^{n};e^{n}_{i,1},e^{n}_{i,2}), which is also recurrent. That is, in the directed graph \mathcal{B} of bands, [bi][b_{i}] belongs to only one cycle, say CC, and there is no directed path from CC to another cycle.

  • (2)

    There exists p>0p>0 such that binpb^{np}_{i} is of type bib_{i} for every i[0,k]i\in[0,k]_{\mathbb{Z}} and any n>0n>0. Moreover, pp can be chosen as the common period of the cycles in \mathcal{B} containing [bi][b_{i}]’s.

  • (3)

    For every pair (i,j),(i,j)[0,k]×{1,2}(i,j),(i^{\prime},j^{\prime})\in[0,k]_{\mathbb{Z}}\times\{1,2\} with ei,jei,je_{i,j}\neq e_{i^{\prime},j^{\prime}}, the band (tn;ei,jn,ei,jn)(t^{n};e^{n}_{i,j},e^{n}_{i^{\prime},j^{\prime}}) is a level-nn recurrent subband of tt for every n0n\geq 0.

Proof.

For any n>0n>0, every level-nn child of b0b_{0} intersects any level-nn child of any bib_{i}. Hence, all the level-nn children of b0,b1,,bkb_{0},b_{1},\dots,b_{k} are bands of the same level-nn subtile, say tnt^{n}, of tt.

For n>0n>0, we define a set AnA^{n} as the collection of level-nn recurrent subbands of the level-0 bands in A0A^{0}. Since every recurrent level-nn band has at least one recurrent child at each higher level, we have a sequence of surjections A0A1A^{0}\leftarrow A^{1}\leftarrow\cdots which map recurrent subbands to their parents.

To show the uniqueness in (1), it suffices to show that the surjections (An+1AnA^{n+1}\to A^{n})’s are actually bijections. Since b0b_{0} is recurrent, for infinitely many n0>0n_{0}>0 the level-0 band b0b_{0} has a level-n0n_{0} recurrent subband b0n0b^{n_{0}}_{0} of type b0b_{0}. In particular, tn0t^{n_{0}} is of type 𝐭\mathbf{t}. Then the types of level-n0n_{0} subbands of bib_{i} for i[1,k]i\in[1,k]_{\mathbb{Z}} injectively correspond to bjb_{j} for j[1,k]j\in[1,k]_{\mathbb{Z}}. This implies that A0An0A^{0}\leftarrow\cdots\leftarrow A^{n_{0}} is a sequence of bijections. Since there are infinitely many such n0n_{0}’s, A0A1A^{0}\leftarrow A^{1}\leftarrow\cdots is a sequence of bijections also.

(2) Let pp be the least positive integer satisfying that b0p=(tp;e0,1p,e0,2p)b^{p}_{0}=(t^{p};e^{p}_{0,1},e^{p}_{0,2}) is of type b0=(t;e0,1,e0,2)b_{0}=(t;e_{0,1},e_{0,2}). Such a pp exists because b0b_{0} is recurrent. We claim that binpb^{np}_{i} is of type bib_{i} for every n1n\geq 1. Here is a sketch of the proof and we left the details to the reader. Since tiles and bands are objects embedded in the sphere, we can define an order on AnA^{n} according to how close binb^{n}_{i} is to b0nb^{n}_{0}. The order is preserved by the bijections Am+1AmA^{m+1}\to A^{m}’s we discussed in (1), which implies binpb^{np}_{i} is of type bib_{i} for every n>0n>0 and i[1,k]i\in[1,k]_{\mathbb{Z}}. By exchanging the roles of b0b_{0} with bib_{i} for any i[1,k]i\in[1,k]_{\mathbb{Z}}, we obtain that pp is the common period of the cycles in \mathcal{B} containing [bi][b_{i}]’s.

(3) Let p>0p>0 be the number determined in (2). It follows from (2) that for every n>0n>0 and (i,j)[0,k]×{1,2}(i,j)\in[0,k]_{\mathbb{Z}}\times\{1,2\}, the level-npnp subtile tnpt^{np} is of type 𝐭\mathbf{t} and its boundary edge ei,jnpe^{np}_{i,j} is a subedge of ei,je_{i,j} and of type 𝐞i,j\mathbf{e}_{i,j}. Hence, for every pair (i,j),(i,j)[0,k]×{1,2}(i,j),(i^{\prime},j^{\prime})\in[0,k]\times\{1,2\} with ei,jei,je_{i,j}\neq e_{i^{\prime},j^{\prime}}, the level-0 band (t;ei,j,ei,j)(t;e_{i,j},e_{i^{\prime},j^{\prime}}) has a level-npnp subband (tnp,ei,jnp,ei,jnp)(t^{np},e^{np}_{i,j},e^{np}_{i^{\prime},j^{\prime}}) which is of type (t;ei,j,ei,j)(t;e_{i,j},e_{i^{\prime},j^{\prime}}). Then (tn,ei,jn,ei,jn)(t^{n},e^{n}_{i,j},e^{n}_{i^{\prime},j^{\prime}}) is recurrent for every n1n\geq 1. ∎

Proposition 6.12.

Let tt be a level-0 tile of a finite subdivision rule \mathcal{R}. Then there is a train-track Tn(t)T^{n}(t) whose underlying graph is properly embedded in (t,n(t),Vert(n(t)))(t,\mathcal{R}^{n}(\partial t),\operatorname{Vert}(\mathcal{R}^{n}(\partial t))) such that any curve γ\gamma properly embedded in (t,n(t),Vert(n(t)))(t,\mathcal{R}^{n}(\partial t),\operatorname{Vert}(\mathcal{R}^{n}(\partial t))) is carried by Tn(t)T^{n}(t) if and only if γ\gamma is the bone of a level-nn recurrent subband of tt.

Proof.

We first assume that the tile tt is a closed disk, i.e., its boundary edges are not identified. It follows from Lemma 6.11-(3) that the collection of spines Bn(t)B^{n}(t) satisfies the crossing condition in Proposition 6.10. Hence we have the desired train-tack by Proposition 6.10.

If tt is not a closed disk. We can use Proposition 6.10 for its domain 𝐭\mathbf{t}, which is a closed disk, of the characteristic map ϕt:𝐭t\phi_{t}:\mathbf{t}\to t to obtain a train-track Tn(𝐭)T^{n}(\mathbf{t}) in 𝐭\mathbf{t}. Then we define Tn(t)T^{n}(t) as the ϕt\phi_{t}-image of Tn(𝐭)T^{n}(\mathbf{t}). If two level-nn edges that are identified by ϕt\phi_{t} have boundary points of Tn(𝐭)T^{n}(\mathbf{t}), then we also identify the boundary points when we define Tn(t)T^{n}(t). ∎

Definition 6.13 (Non-expanding spine of a tile).

Let tt be a level-0 tile of a finite subdivision rule \mathcal{R}. The level-nn train-track of tt, denoted by Tn(t)T^{n}(t), is a train-track properly embedded in (t,n(t),Vert(n(t)))(t,\mathcal{R}^{n}(t),\operatorname{Vert}(\mathcal{R}^{n}(t))) defined in Proposition 6.12. Simply, Tn(t)T^{n}(t) defined is by the following procedure:

  • (1)

    Draw the bones of level-nn recurrent subbands of tt.

  • (2)

    Merge intersecting bones, which form complete graphs by Lemma 6.9, to a star-like trees.

  • (3)

    Zip-up bones meeting at a boundary point of tt as in Figure 8.

Definition 6.14.

(Non-expanding spines) Let \mathcal{R} be a finite subdivision rule. For every n0n\geq 0, the level-nn non-expanding spine NnN^{n} of \mathcal{R} is a train-track that is defined by the union of level-nn non-expanding spines Tn(t)T^{n}(t) of all the level-0 tiles tt. When two tiles tt and tt^{\prime} have the common level-nn edge ene^{n} such that both Tn(t)T^{n}(t) and Tn(t)T^{n}(t^{\prime}) have boundary points on ene^{n}, then we identify the boundary points when we take the union to define NnN^{n}.

Remark 6.15.

From the definition, we may consider bones of level-nn recurrent bands as curves contained in the level-nn non-expanding spine NnN^{n} or non-expanding spine Tn(t)T^{n}(t) of tiles tt. We will in particular consider curves supported in NnN^{n} as a concatenation of bones of bands.

Definition 6.16 (Essential non-expanding spines).

Let \mathcal{R} be a finite subdivision rule and f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\to S_{\mathcal{R}} be its subdivision map. Let AVert(S)A\subset\operatorname{Vert}(S_{\mathcal{R}}) be a set of marked points. We say that the level-nn non-expanding spine NnN^{n} of \mathcal{R} is essential relative to AA if it contains (more precisely carries as a train-track) a closed curve that is homotopic relative to AA neither to a point nor to some iterate of a peripheral loop of a Julia point in AA.

Example 6.17.
Refer to caption
AA
CC
AA
AA
CC
AA
BB
DD
DD
BB
BB
BB
CC
CC
CC
CC
DD
BB
CC
DD
BB
AA
BB
AA
AA
DD
Refer to caption
Figure 9. An example of level-0,11 non expanding spines, which are non-essential.

See Figure 9. The upper two squares are level-0 tiles (or tile types). One tile is shaded and the other is not shaded. There are four level-0 edges (or edge types) A,B,C,A,\,B,\,C, and DD. The lower two squares are subdivisions at level-11. The bi-recurrent components of level-0 and level-11 non-expanding spine are both homotopic to a peripheral loop of a Julia vertex. By Theorem 6.21, the subdivision map does not have a Levy cycle. Later, in Example 8.7, we will show that the subdivision map also does not have a Thurston obstruction.

Definition 6.18 (Nested sequence of non-expanding spines).

For any n>m0n>m\geq 0, there is a natural map ϕmn:NnNm\phi^{n}_{m}:N^{n}\to N^{m} sending every level-nn recurrent bone to its level-mm parent. Since n(S)\mathcal{R}^{n}(S_{\mathcal{R}}) is a subdivision of m(S)\mathcal{R}^{m}(S_{\mathcal{R}}), we may consider NnN^{n} and NmN^{m} as train-tracks in the same complex m(S)\mathcal{R}^{m}(S_{\mathcal{R}}). Then the map ϕmn\phi^{n}_{m} is ambient homotopic to the the identity map relative to the 11-skeleton m(S)(1)\mathcal{R}^{m}(S_{\mathcal{R}})^{(1)} of the level-mm subdivision complex, i.e., there is an extension ϕmn:m(S)m(S)\phi^{n}_{m}:\mathcal{R}^{m}(S_{\mathcal{R}})\to\mathcal{R}^{m}(S_{\mathcal{R}}) that sends, possibly non-homeomorphically, every edge to the same edge and fixes vertices point-wisely. It is straightforward from definitions that the map ϕmn:NnNm\phi^{n}_{m}:N^{n}\to N^{m} is a train-track map. Then we have a sequence of train-track maps

N0ϕ01N1ϕ12N2ϕ23.N^{0}\xleftarrow[]{\phi^{1}_{0}}N^{1}\xleftarrow[]{\phi^{2}_{1}}N^{2}\xleftarrow[]{\phi^{3}_{2}}\cdots.

We call this sequence the nested sequence of non-expanding spines.

For example, in Figure 7, each tripod is mapped to a curve by ϕ01\phi^{1}_{0} and ϕ12\phi^{2}_{1}.

Proposition 6.19.

Let \mathcal{R} be a finite subdivision rule and NnN^{n} be the level-nn non-expanding spine of \mathcal{R}. For any n>mn>m, if NnN^{n} is essential relative to AA then NmN^{m} is also essential relative to AA

Proof.

Since NnN^{n} is essential, there is a close curve γ\gamma supported in NnN^{n} such that γ\gamma is neither homotopically trivial nor homotopic to some iterates of the peripheral loop of a Julia point in AA. Then ϕmn(γ)\phi^{n}_{m}(\gamma) is a closed curve supported in NmN^{m} with the same property. Then NmN^{m} is essential. ∎

Restriction of the ranges of non-expanded recurrent curves to NnN^{n}. Let \mathcal{R} be a finite subdivision rule. Let AVert(S)A\subset\operatorname{Vert}(S_{\mathcal{R}}) be a set of marked points. It is straightforward from Proposition 6.12 that a closed curve γ:IS2A\gamma:I\to S^{2}\setminus A is homotopic relative to AA to a level-nn recurrent non-expanded curve if and only if it is carried by NnN^{n} in (S2,A)(S^{2},A). Hence, when considering a level-nn non-expanded recurrent curve γ:In(S)\gamma:I\to\mathcal{R}^{n}(S_{\mathcal{R}}), we can restrict the range to Nnn(S)N^{n}\subset\mathcal{R}^{n}(S_{\mathcal{R}}) and think of γ\gamma as a curve supported in NnN^{n}.

Proposition 6.20.

Let \mathcal{R} be a finite subdivision rule and f:(S2,A)f:(S^{2},A)\righttoleftarrow be the subdivision map where AVert(S)A\subset\operatorname{Vert}(S_{\mathcal{R}}) is a set of marked points. Let ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}} be a hyperbolic orbisphere structure of f:(S2,A)f:(S^{2},A)\righttoleftarrow. Then the level-nn non-expanding spine NnN^{n} is homotopically infinite for every n0n\geq 0 if and only if there is a genealogical sequence of level-nn homotopically infinite non-expanded recurrent curves {γn:(,)Nnn(S)}n0\{\gamma^{n}:(-\infty,\infty)\to N^{n}\subset\mathcal{R}^{n}(S_{\mathcal{R}})\}_{n\geq 0}.

Proof.

If NnN^{n} is not homotopically infinite, then every closed curve is homotopically finite, i.e., either trivial or some iterate of the peripheral loop of aAa\in A with ord(a)<\mathrm{ord}(a)<\infty. Since NnN^{n} consists of finitely many bones, any infinite curve in NnN^{n} is approximated by closed curves. Then any infinite curve supported in NnN^{n} also cannot be homotopically infinite, so the desired genealogical sequence of curves does not exist.

Suppose that NnN^{n} is homotopically infinite for every n0n\geq 0. We are going to define a set CnC^{n} of homotopically infinite non-expanded recurrent curves supported NnN^{n} with maps {ϕmn:CnCm|n>m0}\{\phi^{n}_{m}:C^{n}\to C^{m}\leavevmode\nobreak\ |\leavevmode\nobreak\ n>m\geq 0\} which send any level-nn non-expanded recurrent curves to their level-mm parents. We will also (1) define a metric dnd_{n} on each CnC^{n} for which the continuity of ϕmn\phi^{n}_{m} easily follows and (2) show that every CnC^{n} is compact. Then the inverse limit of {ϕmn:CnCm|n>m0}\{\phi^{n}_{m}:C^{n}\to C^{m}\leavevmode\nobreak\ |\leavevmode\nobreak\ n>m\geq 0\} is non-empty whose every element yields a desired genealogical sequence of homotopically infinite non-expanded recurrent curves.

Recall that a level-nn non-expanded curve α:(,)n(S)\alpha:(-\infty,\infty)\to\mathcal{R}^{n}(S_{\mathcal{R}}) is defined to satisfy that for any nn\in\mathbb{Z}, α([n,n+1])\alpha([n,n+1]) is a level-nn bone. Let us define the set CnC^{n} as the collection of (parametrized) level-nn homotopically infinite non-expanded recurrent curves γn:(,)Nn\gamma^{n}:(-\infty,\infty)\to N^{n} up to reparametrization such that γn\gamma^{n} does not contain a homotopically finite closed curve, which is a closed curve whose free homotopy class corresponds to the conjugate class of a torsion element of π1(S2,A,ord)\pi_{1}(S^{2},A,\mathrm{ord}). Each CnC^{n} is non-empty because NnN^{n} is homotopically infinite. We define a metric dnd_{n} on CnC^{n} by dn(γn,δn)=2md_{n}(\gamma^{n},\delta^{n})=2^{-m} where γn,δnCn\gamma^{n},\delta^{n}\in C^{n} and m>0m>0 is the minimal integer satisfying γn([m,m])δn([m,m])\gamma^{n}([-m,m])\neq\delta^{n}([-m,m]) as unions of level-nn bones. It is easy to show that for αn,βn,γnCn\alpha^{n},\beta^{n},\gamma^{n}\in C^{n},

dn(αn,γn)=max(dn(αn,βn),dn(βn,γn))d_{n}(\alpha^{n},\gamma^{n})=\max(d_{n}(\alpha^{n},\beta^{n}),d_{n}(\beta^{n},\gamma^{n}))

so that dnd_{n} is indeed a metric. It is immediate from the definition that ϕmn:(Cn,dn)(Cm,dm)\phi^{n}_{m}:(C^{n},d_{n})\to(C^{m},d_{m}) is distance non-increasing. Hence ϕmn\phi^{n}_{m} is uniformly continuous.

Lastly, let us show that (Cn,dn)(C^{n},d_{n}) is sequentially compact. Suppose that {γinCn}i1\{\gamma^{n}_{i}\in C_{n}\}_{i\geq 1} is any sequence in CnC^{n}. Recall that γin([0,1])\gamma^{n}_{i}([0,1]) is a bone of level-nn band so that, in particular, γin(0)\gamma^{n}_{i}(0) is a point in Nnn(S)(1)N^{n}\cap\mathcal{R}^{n}(S_{\mathcal{R}})^{(1)}, which is a finite set. By dropping to a subsequence, we may assume that there exists xNnx\in N^{n} such that γin(0)=x\gamma^{n}_{i}(0)=x for any i>0i>0. Let p:𝔻S2Ap:\mathbb{D}\to S^{2}\setminus A^{\infty} is the orbifold universal covering map of (S2,A,ord)(S^{2},A,\mathrm{ord}), where A={aA|ord(a)=}A^{\infty}=\{a\in A\leavevmode\nobreak\ |\leavevmode\nobreak\ \mathrm{ord}(a)=\infty\}. Choose x~p1(x)\widetilde{x}\in p^{-1}(x). For every i>0i>0, there exists a unique lifting γ~in:(,)p1(Nn)\widetilde{\gamma}^{n}_{i}:(-\infty,\infty)\to p^{-1}(N^{n}) with γ~in(0)=x~\widetilde{\gamma}^{n}_{i}(0)=\widetilde{x}. Define S0:={γ~in}i1S_{0}:=\{\widetilde{\gamma}^{n}_{i}\}_{i\geq 1}.

Fix r>0r>0. Let B(x~,r)𝔻B(\widetilde{x},r)\subset\mathbb{D} denote the hyperbolic ball of radius r>0r>0 with the center at x~\widetilde{x}. For any i>0i>0 and some ki,li0k_{i},l_{i}\in\mathbb{Z}_{\geq 0}, we say that γ~in([ki,li])\widetilde{\gamma}^{n}_{i}([-k_{i},l_{i}]) is the longest initial subcurve of γ~in\widetilde{\gamma}^{n}_{i} staying in B(x~,r)B(\widetilde{x},r) if γ~in([ki,li])B(x~,r)\widetilde{\gamma}^{n}_{i}([-k_{i},l_{i}])\subset B(\widetilde{x},r) and γ~in(ki1),γ~in(li+1)B(x~,r)\widetilde{\gamma}^{n}_{i}(-k_{i}-1),\widetilde{\gamma}^{n}_{i}(l_{i}+1)\notin B(\widetilde{x},r). The intersection B(x~,r)p1(Nn)B(\widetilde{x},r)\cap p^{-1}(N^{n}) has at most finitely many edges of p1(Nn)p^{-1}(N^{n}). Since every element of CnC^{n} does not contain a homotopically finite closed curve, any γ~in\widetilde{\gamma}^{n}_{i} does not contain a closed curve. It follows that there exists m1<m2<m_{1}<m_{2}<\cdots such that every element of the subsequence {γ~min}i>0\{\widetilde{\gamma}^{n}_{m_{i}}\}_{i>0} of S0S_{0} has the same longest initial subcurve staying in B(x~,r)B(\widetilde{x},r).

Take a sequence 0<r1<r2<0<r_{1}<r_{2}<\cdots with rir_{i}\to\infty. For any m>0m>0 define SmS_{m} as a subsequence of Sm1S_{m-1} whose elements have the same longest initial subcurve staying in B(x~,rm)B(\widetilde{x},r_{m}). By taking the diagonal of a sequence of subsequences S0,S1,S2,S_{0},S_{1},S_{2},\dots, we have a subsequence {γ~min}\{\widetilde{\gamma}^{n}_{m_{i}}\} of S0S_{0} with the following property: There exist two strictly increasing sequences of positive integers {ai}i0\{a_{i}\}_{i\geq 0} and {bi}i0\{b_{i}\}_{i\geq 0} such that

  • (1)

    γ~min([ai,bi])\widetilde{\gamma}^{n}_{m_{i}}([-a_{i},b_{i}]) is the longest initial subcurve staying in B(x~,ri)B(\widetilde{x},r_{i}) and

  • (2)

    γ~mjn([ai,bi])=γ~min([ai,bi])\widetilde{\gamma}^{n}_{m_{j}}([-a_{i},b_{i}])=\widetilde{\gamma}^{n}_{m_{i}}([-a_{i},b_{i}]) for any j>ij>i, i.e., the initial subcurves are accumulated.

We define a curve γ~n:(,)p1(Nn)\widetilde{\gamma}^{n}:(-\infty,\infty)\to p^{-1}(N^{n}) by γ~n|[ai,bi]=γ~in|[ai,bi]\widetilde{\gamma}^{n}|_{[-a_{i},b_{i}]}=\widetilde{\gamma}^{n}_{i}|_{[-a_{i},b_{i}]} for every i1i\geq 1, which is well-defined by (2). It follows from (1) that for any i1i\geq 1 we have

d𝔻(x~,γ~n(ai1)),d𝔻(x~,γ~n(bi+1))>ri,d_{\mathbb{D}}(\widetilde{x},\widetilde{\gamma}^{n}(-a_{i}-1)),\leavevmode\nobreak\ d_{\mathbb{D}}(\widetilde{x},\widetilde{\gamma}^{n}(b_{i}+1))>r_{i},

where d𝔻d_{\mathbb{D}} is the hyperbolic metric on 𝔻\mathbb{D}. Hence, γn:=pγ~n\gamma^{n}:=p\circ\widetilde{\gamma}^{n} is homotopically infinite. Then γminγnCn\gamma^{n}_{m_{i}}\to\gamma^{n}\in C^{n}, which implies (Cn,dn)(C^{n},d_{n}) is sequentially compact.

Theorem 6.21.

Let \mathcal{R} be a finite subdivision rule and f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\to S_{\mathcal{R}} be its subdivision map which is not doubly covered by a torus endomorphism. Let AVert(S)A\subset\operatorname{Vert}(S_{\mathcal{R}}) be a set of marked points, i.e., Pff(A)AP_{f}\cup f(A)\subset A. Then the post-critically finite branched covering f:(S2,A)f:(S^{2},A)\righttoleftarrow has a Levy cycle if and only if the level-nn non-expanding spine NnN^{n} is essential relative to AA for every n0n\geq 0.

Proof.

Let ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}} be an orbisphere structure. Any multiplication of ord\mathrm{ord} by a positive integer gives rise to another orbisphere structure with strictly decreased Euler characteristic. Similarly, changing the order of every Fatou point in AA into the infinity also yields an orbisphere structure with strictly decreased Euler characteristic, if some order was actually changed. Hence, we always have a hyperbolic orbisphere structure ord:A[2,]\mathrm{ord}:A\to[2,\infty]_{\mathbb{Z}} with the property that ord(a)=\mathrm{ord}(a)=\infty if and only if aAa\in A is a Fatou point. Then a closed curve is homotopically infinite with respect to ord\mathrm{ord} if and only if it is neither homotopic relative to AA to a point nor to some iterate of a peripheral loop of a Julia point in AA. Then the theorem follows from Proposition 5.14, 5.15, and 6.20. ∎

7. Graph intersecting obstruction

7.1. Graph intersecting obstructions

Suppose that f:(S2,A)f:(S^{2},A)\righttoleftarrow is a post-critically finite branched covering. A graph GS2G\subset S^{2} is forward invariant under ff up to isotopy relative to AA if there exist a subgraph HH of f1(G)f^{-1}(G) and a homeomorphism ϕ:S2S2\phi:S^{2}\to S^{2} such that ϕ(H)=G\phi(H)=G and ϕ\phi is isotopic to the identity map relative to AA. A graph GG is forward invariant under ff if f(G)Gf(G)\subset G. A multicurve Γ\Gamma on (S2,A)(S^{2},A) is forward invariant under ff up to isotopy relative to AA if it is so as a graph. A multicurve Γ\Gamma is backward invariant under ff up to isotopy relative to AA, or ff-stable, if every connected component of f1(γ)f^{-1}(\gamma) for γΓ\gamma\in\Gamma is either isotopic relative to AA to an element of Γ\Gamma or peripheral to AA. When ff and AA are understood, we omit “under ff” and “relative to AA”.

Proposition 7.1.

Let f:(S2,A)f:(S^{2},A)\righttoleftarrow be a post-critically finite branched covering and GG be a graph that is forward invariant up to isotopy. Then there exists ι:(S2,A)\iota:(S^{2},A)\righttoleftarrow which is isotopic to idS2:(S2,A)id_{S^{2}}:(S^{2},A)\righttoleftarrow relative to AA, such that GG is forward invariant under a post-critically finite branched covering g:(S2,A)g:(S^{2},A)\righttoleftarrow defined by g:=fιg:=f\circ\iota. Especially, ff and gg are combinatorially equivalent by idS2id_{S^{2}} and ι\iota.

Proof.

Let HH be a subgraph of f1(G)f^{-1}(G) isotopic to GG rel Vert(G)\operatorname{Vert}(G). By extending the isotopy to S2S^{2}, we have ι:(S2,A)(S2,A)\iota:(S^{2},A)\to(S^{2},A) such that ι(G)=H\iota(G)=H and ι\iota and idS2id_{S^{2}} are isotopic relative to AA. Let g:=fιg:=f\circ\iota. Then idg=fιid\circ g=f\circ\iota and g(G)=f(ι(G))=f(H)Gg(G)=f(\iota(G))=f(H)\subset G. ∎

Due to Proposition 7.1, we may consider forward invariant graphs instead of graphs that are forward invariant up to isotopy when discussing properties of combinatorial equivalence classes, such as Levy cycles and Thurston obstructions.

Let Γ\Gamma be a multicurve in S2AS^{2}\setminus A. The Thurston linear transformation of Γ\Gamma is a linear map fΓ:ΓΓf_{\Gamma}:\mathbb{R}^{\Gamma}\to\mathbb{R}^{\Gamma} defined by

fΓ(γ)=γf1(γ)1deg(f|γ:γγ)[γ]Γf_{\Gamma}(\gamma)=\sum\limits_{\gamma^{\prime}\subset f^{-1}(\gamma)}\frac{1}{\deg(f|_{\gamma^{\prime}}:\gamma^{\prime}\to\gamma)}[\gamma^{\prime}]_{\Gamma}

where γ\gamma^{\prime} is a connected component of f1(γ)f^{-1}(\gamma) and [γ]Γ[\gamma^{\prime}]_{\Gamma} is an element of Γ\Gamma isotopic to γ\gamma^{\prime} if exists. If no such connected component exists, then the sum is defined to be zero. Since fΓf_{\Gamma} is a non-negative matrix, it has a non-negative real eigenvalue λ(fΓ)\lambda(f_{\Gamma}) that is the spectral radius of fΓf_{\Gamma}. If λ(fΓ)1\lambda(f_{\Gamma})\geq 1, then Γ\Gamma is a Thurston obstruction. A n×nn\times n non-negative square matrix MM is irreducible if for each i,ji,\,j with 1i,jn1\leq i,\,j\leq n there exists k1k\geq 1 such that the (i,j)(i,\,j)-entry of MkM^{k} is positive. An irreducible multicurve Γ\Gamma is a multicurve whose Thurston linear transformation fΓf_{\Gamma} is irreducible. An irreducible Thurston obstruction is an irreducible multicurve that is a Thurston obstruction.

Remark 7.2.

A Thurston obstruction Γ\Gamma is usually assumed to be ff-stable. For any multicurve Γ\Gamma with λ(fΓ)0\lambda(f_{\Gamma})\neq 0, there exists a sub-multicurve ΓΓ\Gamma^{\prime}\subset\Gamma such that Γ\Gamma^{\prime} is irreducible and λ(fΓ)=λ(fΓ)\lambda(f_{\Gamma^{\prime}})=\lambda(f_{\Gamma}). Such Γ\Gamma^{\prime} is determined as the multicurve of an irreducible diagonal block AiA_{i} of the upper-triangular block form (UTB-form) of fΓf_{\Gamma} with λ(Ai)=λ(fΓ)\lambda(A_{i})=\lambda(f_{\Gamma}). By Lemma 7.4, Γ\Gamma^{\prime} extends to a ff-invariant multicurve Γ′′\Gamma^{\prime\prime} with λ(fΓ′′)λ(fΓ)\lambda(f_{\Gamma^{\prime\prime}})\geq\lambda(f_{\Gamma}) . Hence we may drop the ff-condition condition from Thurston’s characterization.

Lemma 7.3.

If a multicurve Γ\Gamma is irreducible, then Γ\Gamma is forward invariant up to isotopy.

Proof.

For a contradiction, assume there exists γ\gamma such that for every γΓ\gamma^{\prime}\in\Gamma no connected component of f1(γ)f^{-1}(\gamma^{\prime}) is isotopic to γ\gamma. Then fΓ:ΓΓ{γ}Γf_{\Gamma}:\mathbb{R}^{\Gamma}\to\mathbb{R}^{\Gamma\setminus\{\gamma\}}\subset\mathbb{R}^{\Gamma}, thus fΓf_{\Gamma} is not irreducible. ∎

Lemma 7.4 ([Tan92, Lemma 2.2]).

For any multicurve Γ\Gamma of (S2,A)(S^{2},A) that is forward invariant up to isotopy, there exists a multicurve Γ\Gamma^{\prime} which is backward invariant up to isotopy such that ΓΓ\Gamma^{\prime}\supset\Gamma and λ(fΓ)λ(fΓ)\lambda(f_{\Gamma^{\prime}})\geq\lambda(f_{\Gamma}).

Proof.

Let Γ0=Γ\Gamma_{0}=\Gamma and Γn\Gamma_{n} be the set of homotopy classes of essential curves in fn(Γ0)f^{-n}(\Gamma_{0}). By the forward invariance up to isotopy, Γ0Γ1\Gamma_{0}\subset\Gamma_{1}\subset\dots is an increasing sequence of multicurves. Note that |A|3|A|-3 is the maximal number of non-homotopic essential simple closed curves that can be disjointly embedded into S2AS^{2}\setminus A. Hence there exists nn such that Γn\Gamma_{n} is ff-invariant. The inequality λ(fΓ)λ(fΓ)\lambda(f_{\Gamma^{\prime}})\geq\lambda(f_{\Gamma}) follows from the following: for non-negative square matrices MM and NN if MijNijM_{ij}\geq N_{ij} for every (i,j)(i,\,j) then λ(M)λ(N)\lambda(M)\geq\lambda(N). ∎

Theorem 7.5 (Arcs intersecting obstructions [PT98, Theorem 3.2]).

Let f:(S2,A)f:(S^{2},A)\righttoleftarrow be a post-critically finite branched covering and GG be an invariant graph such that f|G:GGf|_{G}:G\to G is a graph automorphism. Then every irreducible Thurston obstruction intersecting GG is a Levy cycle.

We generalize it to a case when GG is an ff-invariant graph with htop(f|G)=0h_{top}(f|_{G})=0.

Theorem 7.6 (Graph intersecting obstruction).

Let f:(S2,A)f:(S^{2},A)\righttoleftarrow be a post-critically finite branched covering and GG be a forward invariant graph such that htop(f|G)=0h_{top}(f|_{G})=0. Then every irreducible Thurston obstruction intersecting GG is a Levy cycle.

Remark 7.7.

The graphs in Theorem 7.5 and Theorem 7.6 are possibly disconnected. Moreover, the same statement works for graphs which are forward invariant up to isotopy by Proposition 7.1 with a slight modification to define htop(f|G)h_{top}(f|_{G}).

An arc of (S2,A)(S^{2},A) is a curve embedded in S2S^{2} such that its interior is embedded in S2AS^{2}\setminus A and its endpoints are in AA. A geometric intersection number γγ\gamma\cdot\gamma^{\prime} between curves (arcs and simple closed curves) is defined as the minimal number of intersection points in their isotopy classes relative to AA.

For a multicurve Γ\Gamma, the unweighted Thurston transformation f#,Γ:ΓΓf_{\#,\Gamma}:\mathbb{R}^{\Gamma}\to\mathbb{R}^{\Gamma} is defined by

f#,Γ(γ)=γf1(γ)[γ]Γf_{\#,\Gamma}(\gamma)=\sum\limits_{\gamma^{\prime}\subset f^{-1}(\gamma)}[\gamma^{\prime}]_{\Gamma}

where γ\gamma^{\prime} is a connected component of f1(γ)f^{-1}(\gamma) and [γ]Γ[\gamma^{\prime}]_{\Gamma} is an element of Γ\Gamma isotopic to γ\gamma^{\prime} if exists. If there is no such element, then the sum is defined to be zero. For every (i,j)(i,\,j), (1) 0(fΓ)ij(f#,Γ)ij0\leq(f_{\Gamma})_{ij}\leq(f_{\#,\Gamma})_{ij} and (2) (fΓ)ij=0(f_{\Gamma})_{ij}=0 if and only if (f#,Γ)ij=0(f_{\#,\Gamma})_{ij}=0. So f#,Γf_{\#,\Gamma} is irreducible if and only if fΓf_{\Gamma} is irreducible.

Proof of Theorem 7.6.

Let Edge(G)={e1,e2,,en}\mathrm{Edge}(G)=\{e_{1},e_{2},\dots,e_{n}\}. For any simple closed curve γS2A\gamma\subset S^{2}\setminus A, define

(γ)G=(#{γe1}#{γe2}#{γen})[γ]G=(γe1γe2γen)(\gamma)_{G}=\left(\begin{array}[]{c}\#\{\gamma\cap e_{1}\}\\ \#\{\gamma\cap e_{2}\}\\ \vdots\\ \#\{\gamma\cap e_{n}\}\end{array}\right)\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ [\gamma]_{G}=\left(\begin{array}[]{c}\gamma\cdot e_{1}\\ \gamma\cdot e_{2}\\ \vdots\\ \gamma\cdot e_{n}\end{array}\right)

where γei\gamma\cdot e_{i} means the geometric intersection number of γ\gamma and eie_{i} relative to AA. The ()G(\leavevmode\nobreak\ )_{G} and []G[\leavevmode\nobreak\ ]_{G} are linearly extended to weighted multicurves. Let TGT_{G} be the incidence matrix of f|Gf|_{G}. Let Γ={γ1,γ2,,γm}\Gamma=\{\gamma_{1},\gamma_{2},\dots,\gamma_{m}\} be an irreducible Thurston obstruction. From #{fk(γi)ej}fk(γi)ej\#\{f^{-k}(\gamma_{i})\cap e_{j}\}\geq f^{-k}(\gamma_{i})\cdot e_{j}, for every k1k\geq 1, we have

(2) TGk(γi)G=(fk(γi))G[fk(γi)]Gj=1m(f#,Γk)ij[γj]G.{T_{G}}^{k}\cdot(\gamma_{i})_{G}=(f^{-k}(\gamma_{i}))_{G}\geq[f^{-k}(\gamma_{i})]_{G}\geq\sum_{j=1}^{m}({f_{\#,\Gamma}}^{k})_{ij}[\gamma_{j}]_{G}.

The third term counts the intersection of GG with all connected components of fk(γi)f^{-k}(\gamma_{i}), but the last term counts the intersection of GG with connected components of fk(γi)f^{-k}(\gamma_{i}) isotopic relative to AA to some connected components of Γ\Gamma.

It follows from Proposition 3.9 that entries of TGk{T_{G}}^{k} grows at most polynomially fast, so (f#,Γk)ij(f_{\#,\Gamma}^{k})_{ij} grows at most polynomially fast too. Since f#,Γf_{\#,\Gamma} is an irreducible non-negative integer matrix, f#,Γf_{\#,\Gamma} is a permutation by Lemma 3.8. Recall that (1) 0(fΓ)ij(f#,Γ)ij0\leq(f_{\Gamma})_{ij}\leq(f_{\#,\Gamma})_{ij} and (2) (fΓ)ij>0(f_{\Gamma})_{ij}>0 if and only if (f#,Γ)ij>0(f_{\#,\Gamma})_{ij}>0. Hence the only way to have λ(fΓ)1\lambda(f_{\Gamma})\geq 1 is f#,Γ=fΓf_{\#,\Gamma}=f_{\Gamma}. Then Γ\Gamma is a Levy cycle. ∎

7.2. Application in the mating of polynomials

Formal mating. Let ff and gg be post-critically finite polynomials of degree dd. Consider ff and gg as maps from the complex plane \mathbb{C} to itself. Let ¯\overline{\mathbb{C}} be the compactification of \mathbb{C} by the circle S1S^{1} each point of which corresponds to a linear direction to the infinity. Then, ff and gg extend to the boundary S1S^{1} as the angle dd-times map. We can parametrize S1S^{1} by θ[0,1]/{01}\theta\in[0,1]/\{0\sim 1\} where θ\theta indicates the angle of an external ray. Let us use subscriptions f-_{f} and g-_{g} to distinguish two compactified complex planes where ff and gg act on respectively, such as 𝔻¯f:=fSf1\overline{\mathbb{D}}_{f}:=\mathbb{C}_{f}\cup S^{1}_{f}, 𝔻¯g:=gSg1\overline{\mathbb{D}}_{g}:=\mathbb{C}_{g}\cup S^{1}_{g}, f:𝔻¯ff:\overline{\mathbb{D}}_{f}\righttoleftarrow and g:𝔻¯gg:\overline{\mathbb{D}}_{g}\righttoleftarrow. Define a sphere Sfg2S^{2}_{f\uplus g} by gluing two compactified planes ¯f\overline{\mathbb{C}}_{f} and ¯g\overline{\mathbb{C}}_{g} by the equivalence relation θfθg\theta_{f}\sim-\theta_{g} for any θfSf1\theta_{f}\in S^{1}_{f} and θgSg1\theta_{g}\in S^{1}_{g} with θf=θg\theta_{f}=\theta_{g} as numbers in [0,1)[0,1). The dynamics of ff and gg also glue together to induce a dynamics fg:Sfg2f\uplus g:S^{2}_{f\uplus g}\righttoleftarrow, which is also a post-critically finite branched self-covering of the sphere. We call fg:Sfg2f\uplus g:S^{2}_{f\uplus g}\righttoleftarrow the formal mating of ff and gg.

Ray-equivalence class. Let fg:Sfg2f\uplus g:S^{2}_{f\uplus g}\righttoleftarrow be the formal mating of post-critically finite polynomials ff and gg. External rays of ff and gg forms a foliation on Sfg2(KfKg)S^{2}_{f\uplus g}\setminus(K_{f}\cup K_{g}) where KffK_{f}\subset\mathbb{C}_{f} and KfgK_{f}\subset\mathbb{C}_{g} are filled Julia sets. Every leaf of the foliation is called a ray-equivalence class of the formal mating fgf\uplus g. Each ray-equivalence class consists of external rays of ff and gg of the same period and pre-period.

Degenerate mating. If ff or gg (or both) is non-hyperbolic, there could be an obvious Levy cycle of fgf\uplus g which could be removed by collapsing some ray equivalence classes.

Let F:=fgF:=f\uplus g. Suppose that ff is not hyperbolic. Then the post-critical set PfP_{f} is in the Julia set 𝒥f\mathcal{J}_{f} so that each post-critical point of ff is contained in a ray-equivalence class. Suppose that there is a periodic ray-equivalence class ξ\xi that contains two points of PFP_{F} such that ξ\xi is topologically a tree. Then the boundary of a small neighborhood of ξ\xi generates a Levy cycle. Hence, we will collapse ξ\xi to a point. To obtain a topological branched covering on the quotient sphere, we need a little more careful construction as follows, see [Shi00] for details.

Let YY^{\prime} be the set of ray-equivalence classes containing at least to points in ΩFPF\Omega_{F}\cup P_{F}. Define YY be the set of ray-equivalence class ξ\xi^{\prime} containing at least one point of ΩFPF\Omega_{F}\cup P_{F} such that Fm(ξ)=Fn(ξ)F^{m}(\xi^{\prime})=F^{n}(\xi) for some m,n0m,n\geq 0 and ξY\xi\in Y. If YY\neq\emptyset and every element of YY is topologically a tree, then we define S2S^{\prime 2} as the quotient of Sfg2S^{2}_{f\uplus g} by collapsing every ray-equivalence class in YY to a point. The map FF induces a degree dd self-map on S2S^{\prime 2} which is not a branched covering near F1(ξ)F^{-1}(\xi) for ξY\xi\in Y. But we can take a homotopy near F1(ξ)F^{-1}(\xi) for ξY\xi\in Y to obtain a branched covering F:(S2,PF)F^{\prime}:(S^{\prime 2},P_{F^{\prime}})\righttoleftarrow, which is called the degenerate mating of ff and gg. We also denote the degenerate mating by fg:Sfg2f\uplus^{\prime}g:S^{2}_{f\uplus^{\prime}g}\righttoleftarrow. When both ff and gg are hyperbolic, the degenerate mating is equal to the formal mating.

Example 7.8 (f1/2f1/4f_{1/2}\uplus^{\prime}f_{1/4}).

For θ[0,1)\theta\in\mathbb{Q}\cap[0,1), let fθf_{\theta} denote the post-critically finite polynomial at the landing point of the external ray of angle θ\theta in the parameter plane of the quadratic polynomials z2+cz^{2}+c. Let f=f1/2f=f_{1/2} and g=f1/4g=f_{1/4}. Let us denote by Rf(θ)R_{f}(\theta) and Rg(θ)R_{g}(\theta) the external rays of ff and gg of angle θ\theta.

The set YY^{\prime} defined above consists of 3 ray-equivalence class: ξ0:=Rf(0)Rg(0)\xi_{0}:=R_{f}(0)\cup R_{g}(0), ξ1:=Rf(1/2)Rg(1/2)\xi_{1}:=R_{f}(1/2)\cup R_{g}(1/2), and ξ2:=Rf(1/4)Rf(3/4)Rg(1/4)Rg(3/4)\xi_{2}:=R_{f}(1/4)\cup R_{f}(3/4)\cup R_{g}(1/4)\cup R_{g}(3/4). The set YY has one more ray-equivalence class ξ3:=Rf(3/8)Rf(7/8)Rg(1/8)Rg(5/8)\xi_{3}:=R_{f}(3/8)\cup R_{f}(7/8)\cup R_{g}(1/8)\cup R_{g}(5/8) than YY^{\prime}.

Let F=fgF=f\uplus g be the formal mating. The boundary of a small disk neighborhood of ξ0\xi_{0} is a Levy cycle of period one. Let us also use ξi\xi_{i} to indicate the collapsed points in Sfg2S^{2}_{f\uplus^{\prime}g}. The degenerate mating FF^{\prime} maps ξi\xi_{i} to ξi1\xi_{i-1} for i=1,2,3i=1,2,3, where ξ2\xi_{2} and ξ3\xi_{3} are critical points of degree two.

Definition 7.9.

For post-critically finite polynomials ff and gg of the same degree, we say that ff and gg are mateable if the degenerate mating F:=fg:(SF2,PF)F^{\prime}:=f\uplus^{\prime}g:(S^{2}_{F^{\prime}},P_{F^{\prime}})\righttoleftarrow is combinatorially equivalent to a post-critically finite rational map.

Corollary 7.10.

Let ff and gg be post-critically finite hyperbolic (resp. possibly non-hyperbolic) polynomials such that at least one of ff and gg has core entropy zero. Then ff and gg are mateable if and only if the formal mating (resp. degenerate mating) does not have a Levy cycle.

Proof.

Assume ff and gg are hyperbolic and the core entropy of ff is zero. Suppose the formal mating of ff and gg does not have a Levy cycle but has a Thurston obstruction Γ\Gamma. We may assume that Γ\Gamma is irreducible. We can think of Hubbard trees HfH_{f} and HgH_{g} of ff and gg as invariant trees in the glued sphere Sfg2S^{2}_{f\uplus g}. By Theorem 7.6, Γ\Gamma is disjoint from HfH_{f}. Then Γ\Gamma yields a Thurston obstruction of the polynomial gg, which is a contradiction.

Suppose that ff and gg may not be hyperbolic and ff has core entropy zero. Let π:Sfg2Sfg2\pi:S^{2}_{f\uplus g}\to S^{2}_{f\uplus^{\prime}g} denote the projection from the sphere of the formal mating to the sphere of the degenerate mating. Let HfH_{f} and HgH_{g} denote the Hubbard tress embedded in Sfg2S^{2}_{f\uplus g}, and let HfH^{\prime}_{f} and HgH^{\prime}_{g} denote their π\pi-image in Sfg2S^{2}_{f\uplus^{\prime}g}. Some points of HfH_{f} and HgH_{g} are identified by π\pi, but HfH^{\prime}_{f} still has entropy zero. By the argument in the previous paragraph, if there is an irreducible Thurston Γ\Gamma obstruction of the degenerate mating that is not a Levy cycle, the Γ\Gamma is disjoint from HfH^{\prime}_{f}. For xSfg2x\in S^{2}_{f\uplus^{\prime}g}, if π1(x)\pi^{-1}(x) is not a singleton, then xHfHgx\in H^{\prime}_{f}\cap H^{\prime}_{g}. Hence the multicurve Γ\Gamma can be lifted to a Thurston obstruction of the formal mating fgf\uplus g with still being disjoint from HfH_{f}. Then Γ\Gamma again yields a Thurston obstruction of the polynomial gg, which is a contradiction. ∎

8. Finite subdivision rules with polynomial growth of edge subdivisions

Definition 8.1 (Polynomial growth of edge subdivisions).

Let \mathcal{R} be a finite subdivision rule and ee be a level-0 edge. The edge ee has sub-exponential growth of subdivisions if

limn#{level-nsubedges ofe}1/n=1.\lim\limits_{n\rightarrow\infty}\#\left\{\textup{level-}n\leavevmode\nobreak\ \textup{subedges\leavevmode\nobreak\ of}\leavevmode\nobreak\ e\right\}^{1/n}=1.

We say that \mathcal{R} has sub-exponential growth of edge subdivisions if every level-0 edge has sub-exponential growth of subdivisions. By Proposition 8.2, we can substitute the term “sub-exponential” for “polynomial ”.

Recall that we defined the directed graph of edge subdivisions \mathcal{E} in Section 4.4. Also recall that a level-0 edge ee is called periodic (or also called recurrent) if the corresponding vertex [e][e] in \mathcal{E} is contained in a cycle.

Proposition 8.2.

A finite subdivision rule \mathcal{R} has sub-exponential growth of edge subdivisions if and only if the cycles in \mathcal{E} are disjoint. In this case, for each level-0 edge ee, #{level-nsubedges ofe}\#\left\{\textup{level-}n\leavevmode\nobreak\ \textup{subedges\leavevmode\nobreak\ of}\leavevmode\nobreak\ e\right\} grows polynomially fast as nn\to\infty.

Proof.

It is straightforward from Theorem 3.6 and Proposition 4.5. ∎

Let f(1):(1)(1)f^{(1)}:\mathcal{R}^{(1)}\rightarrow\mathcal{R}^{(1)} be the restriction of ff to the 11-skeleton (1)\mathcal{R}^{(1)}. Then f(1)f^{(1)} is a Markov map. The adjacency matrix of the directed graph of edge subdivision \mathcal{E} coincides with the incidence matrix of the Markov map f(1)f^{(1)}. The following proposition is immediate from Proposition 3.9.

Proposition 8.3.

A finite subdivision rule \mathcal{R} has polynomial growth of edge subdivisions if and only if htop(f(1))=0h_{top}(f^{(1)})=0.

Let ee be a level-0 periodic edge. For every n>0n>0, ee has at least one level-nn child (subedge) that is recurrent, see Section 4.5. If ee has polynomial growth of subdivisions, then the recurrent subedges are unique at each level. The same statement also works for periodic bands.

Proposition 8.4 (Unique recurrent children).

Let \mathcal{R} be a finite subdivision rule and ee be a level-0 periodic edge with polynomial growth of subdivisions. For any n1n\in 1, ee has a unique level-nn subedge that is recurrent. For a level-0 periodic band (t;e1,e2)(t;e_{1},e_{2}), if e1e_{1} and e2e_{2} have polynomial growth of subdivisions, then for any n>0n>0 there exists a unique level-nn subband of (t;e1,e2)(t;e_{1},e_{2}) that is recurrent.

Proof.

By Proposition 8.2, there exists a unique cycle in \mathcal{E} passing through [e][e]. Hence, for any n>0n>0, there is only one path of length nn from [e][e] and supported within the cycle, which determines a unique level-nn recurrent subedge by Proposition 4.5. The uniqueness of recurrent subedge can also follow from Proposition 8.4.

If (t;e1,e2)(t;e_{1},e_{2}) is periodic, then it has at least one level-nn unique subband (tn;e1n,e2n)(t^{n};e_{1}^{n},e_{2}^{n}). By Lemma 4.9, the level-nn edges e1ne^{n}_{1} and e2ne^{n}_{2} are recurrent subedges of e1e_{1} and e2e_{2}, which are unique by the previous paragraph. Hence the recurrent subbands are unique at each level. ∎

Proposition 8.5.

Suppose a finite subdivision rule \mathcal{R} has polynomial growth of edge subdivisions. Then every train-track map ϕnn+1:Nn+1Nn\phi^{n+1}_{n}:N^{n+1}\to N^{n} in the nested sequence of non-expanding spines, defined in Definition 6.18,

N0ϕ01N1ϕ12N2ϕ23,N^{0}\xleftarrow[]{\phi^{1}_{0}}N^{1}\xleftarrow[]{\phi^{2}_{1}}N^{2}\xleftarrow[]{\phi^{3}_{2}}\cdots,

is a homeomorphism.

Proof.

Let (t;e1,e2)(t;e_{1},e_{2}) be a level-0 periodic band. It follows from Proposition 8.4 and Lemma 4.9 that for any nn there exists a unique level-nn recurrent band of (t;e1,e2)(t;e_{1},e_{2}) such that its sides are unique level-nn recurrent subedges of e1e_{1} and e2e_{2}. If two level-0 periodic bands share a side ee then the level-nn recurrent bands also share a side which is the level-nn recurrent subedge of ee. Hence NnN^{n} and N0N^{0} are made up of the same number of bones of bands which are glued in the same way. ∎

Theorem 8.6.

Let \mathcal{R} be a finite subdivision rule with polynomial growth of edge subdivisions and ff be its subdivision map which is not doubly covered by a torus endomorphism. Let AVert(S)A\subset\operatorname{Vert}(S_{\mathcal{R}}) be a set of marked point, i.e., f(A)PfAf(A)\cup P_{f}\subset A. Then the followings are equivalent.

  1. (1)

    f:(S2,A)f:(S^{2},A)\righttoleftarrow does not have a Levy cycle.

  2. (2)

    The level-0 non-expanding spine N0N^{0} does not carry a closed curve that is neither homotopic relative to AA to a point nor to some iterate of a peripheral loop of a Julia point in AA.

  3. (3)

    f:(S2,A)f:(S^{2},A)\righttoleftarrow is combinatorially equivalent to a unique rational map up to conjugation by Möbius transformations, i.e., ff does not have a Thurston obstruction.

Proof.

(1)(2)(1)\Leftrightarrow(2) follows from Theorem 6.21 and Proposition 8.5. The equivalence with (3)(3) follows from Theorem 7.6 and Proposition 8.3. ∎

Example 8.7 (Example 6.17 continued).

Removing the edge type CC from Figure 9, we have a finite subdivision rule with bounded edge subdivisions. Since the subdivision maps is unchanged, there is no Levy cycle by the discussion in Example 6.17. It follows from Theorem 8.6 that there is also no Thurston obstruction.

9. Examples

9.1. Critically fixed rational maps

A rational map is critically fixed if every critical point is a fixed point. It was recently shown that there is a one-to-one correspondence between critically fixed rational functions and planar graphs. The idea started from [PT98] and was completed in [Hlu19].

Theorem 9.1 (Hlushchanka, Pilgrim et. al.).

There is a one-to-one correspondence between the holomorphic conjugacy classes of critically fixed rational functions and the planar isotopy classes of connected planar graphs without loops.

Let GG be a planar graph without loops and ff be the corresponding critically fixed rational map in Theorem 9.1. At the end of this subsection, we construct a finite rule G\mathcal{R}_{G} such that (1) its subdivision map is ff and (2) every edge never subdivides.

Let ff be a critically fixed rational map. The Tischler graph of ff is a graph embedded in ^\hat{\mathbb{C}} whose edge set is the collection of fixed internal rays in the immediate attracting basins of all critical points. It follows from [Hlu19] that the Tischler graph any critically fixed rational map is connected.

To construct a critically fixed rational function from a planar graph without loops, we use blowing-up an arc construction, that is firstly introduced in [PT98].

Blowing-up an arc.. Let f:(S2,A)f:(S^{2},A)\righttoleftarrow be a post-critically finite branched covering and γ\gamma be an arc fixed by ff. Let DS2D\subset S^{2} be an open 22-disc contained in a small neighborhood of γ\gamma with γD\gamma\subset\partial D. Let γ=Dint(γ)\gamma^{\prime}=\partial D-\mathrm{int}(\gamma). Define an orientation-preserving continuous map g:S2DS2g:S^{2}\setminus D\rightarrow S^{2} in such a way that gg maps γ\gamma and γ\gamma^{\prime} to γ\gamma, with endpoints fixed. Define another orientation-preserving continuous map h:D¯S2h:\overline{D}\rightarrow S^{2} in a similar way so that hh maps γ\gamma and γ\gamma^{\prime} to γ\gamma, with endpoints fixed, and maps the DD to S2γS^{2}\setminus\gamma homeomorphically. A new branched covering fγ:(S2,A)f_{\gamma}:(S^{2},A)\righttoleftarrow is defined by fγ|S2D=fgf_{\gamma}|_{S^{2}\setminus D}=f\circ g and fγ|D¯=fhf_{\gamma}|_{\overline{D}}=f\circ h. We call fγf_{\gamma} the ff blown-up along an arc γ\gamma. Note that deg(fγ)=deg(f)+1\deg(f_{\gamma})=\deg(f)+1.

Let GG be a planar graph without loops and A=Vert(G)A=\operatorname{Vert}(G). Define a post-critically finite branched covering fG:(S2,A)f_{G}:(S^{2},A)\righttoleftarrow by blowing up the identity map idS2:(S2,A)id_{S^{2}}:(S^{2},A)\righttoleftarrow along all edges of GG. The combinatorial equivalence class dis independent of the order of blowing-up. Each vertex vv of GG is a critical point of fGf_{G} such that degv(fG)=deg(v)+1\deg_{v}(f_{G})=\deg(v)+1. If follows from [PT98, Corollary 3] that fGf_{G} is combinatorially equivalent to a rational map. Because fγf_{\gamma} fixes γ\gamma, the branched covering fGf_{G} is the identity on GG. Define a finite subdivision rule G\mathcal{R}_{G} such that

  • (1)

    SGS_{\mathcal{R}_{G}} be the CW-complex whose 11-skeleton is GG

  • (2)

    G(SG)\mathcal{R}_{G}(S_{\mathcal{R}_{G}}) be the CW-complex whose 11-skeleton is f1(G)f^{-1}(G), and

  • (3)

    fG:G(SG)SGf_{G}:\mathcal{R}_{G}(S_{\mathcal{R}_{G}})\rightarrow S_{\mathcal{R}_{G}} is the subdivision map of G\mathcal{R}_{G}.

Remark 9.2.

When an edge is blown-up, there are two choices for DD, depending on which side of γ\gamma the disk DD is. But the combinatorial equivalence class of the resulting branched covering is independent.

Example 9.3.

See Figure 10. The graph GG is a triangle with one more edge attached. Figure 10A indicates the disk that we use in the blowing-up the edge γ\gamma. The other two figures indicate the CW-complex structures at level-0 and 11. The shaded triangles in Figure 10C are mapped to the shaded triangle in Figure 10B under the subdivision map fGf_{G}.

Refer to caption
A
Refer to captionγ\gamma
B SGS_{\mathcal{R}_{G}}
Refer to caption
C G(SG)\mathcal{R}_{G}(S_{\mathcal{R}_{G}})
Figure 10.

9.2. Face-inversion constructions and critically fixed anti-rational maps

The construction in this section was also investigated in [Gey20] and [LLM20] in the study of critically fixed anti-rational maps.

Let GG be a finite graph in the 22-sphere S2S^{2}. The graph GG defines the CW-complex structure 𝒯\mathcal{T} with 𝒯(1)=G\mathcal{T}^{(1)}=G. A graph is kk-vertex-connected or kk-edge-connected if it is not disconnected by the removal of fewer than kk vertices or (open) edges respectively. For the characteristic map ϕt:𝐭t\phi_{t}:\mathbf{t}\rightarrow t of a closed 22-cell tt, we say that the boundary vertices or edges of tt are identified if more than one vertex or edge are identified under ϕt\phi_{t}. The followings are characterizations of 22- or 33-connectivity of graphs embedded in S2S^{2}.

  • GG is 22-vertex-connected if and only if boundary vertices of every 22-cell are not identified, i.e., the boundary of every 22-cell is a Jordan curve.

  • GG is 22-edge-connected if and only if the boundary edges of every 22-cell are not identified. The 22-vertex connectedness implies the 22-edge connectedness.

  • GG is 33-edge-connected if and only if it is 22-edge-connected and any two 22-cells do not share more than one edge. It is also equivalent to the dual graph having no cycle of length 2\leq 2.

Assume GG is 22-vertex-connected and deg(v)3\deg(v)\geq 3 for every vVert(G)v\in\operatorname{Vert}(G). Let tt be a 22-cell of 𝒯\mathcal{T} and σt\sigma_{t} be the reflection of S2S^{2} in t\partial t. This is possible because the 22-vertex-connectedness implies that t\partial t is a simple closed curve. Then σt(G)\sigma_{t}(G) is a graph isomorphic to GG such that σt(G)G=t\sigma_{t}(G)\cap G=\partial t. Define a graph HH by

H=tis a 2-cell of𝒯σt(G).H=\bigcup\limits_{t\leavevmode\nobreak\ \textup{is\leavevmode\nobreak\ a\leavevmode\nobreak\ 2-cell\leavevmode\nobreak\ of}\leavevmode\nobreak\ \mathcal{T}}\sigma_{t}(G).

Let 𝒯\mathcal{T}^{\prime} be the CW-complex structure on S2S^{2} with 𝒯(1)=H\mathcal{T}^{\prime(1)}=H. We define a finite subdivision rule as follows: Let S=𝒯S_{\mathcal{R}}=\mathcal{T} and (S)=𝒯\mathcal{R}(S_{\mathcal{R}})=\mathcal{T}^{\prime}. Define an orientation reversing branched self-covering f:S2f:S^{2}\righttoleftarrow defined by f|t=σt|tf|_{t}=\sigma_{t}|_{t} for every 22-cell tt of SS_{\mathcal{R}}. Then ff becomes a subdivision map f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\rightarrow S_{\mathcal{R}}. Note that every edge does not subdivide. The degree of ff is equal to the number of 22-cells of 𝒯\mathcal{T} minus one. We call \mathcal{R} the finite subdivision rule of face-inversion of GG. Every vertex vVert(G)v\in\operatorname{Vert}(G) is a fixed critical point of degv(f)=deg(v)1\deg_{v}(f)=\deg(v)-1.

A natural way to obtain an orientation preserving finite subdivision rule is to take square of the subdivision 2\mathcal{R}^{2} and the subdivision map f2:2(𝒯)𝒯f^{2}:\mathcal{R}^{2}(\mathcal{T})\rightarrow\mathcal{T}. We denote by sq\mathcal{R}_{sq} this squared orientation preserving subdivision rule. Another way is to post-compose with an orientation reversing automorphism of GG. An automorphism τAut(G)\tau\in\operatorname{Aut}(G) is called orientation reversing if it extends to an orientation reversing homeomorphism of S2S^{2}. For any orientation reversing automorphism τAut(G)\tau\in\operatorname{Aut}(G), we have an orientation preserving subdivision map fτ:=τf:(S)Sf_{\tau}:=\tau\circ f:\mathcal{R}(S_{\mathcal{R}})\rightarrow S_{\mathcal{R}} defined on the same subdivision complexes as \mathcal{R}. Denote this finite subdivision rule by τ\mathcal{R}_{\tau}.

Theorem 9.4.

Let GG be a 22-vertex-connected graph in S2S^{2} such that deg(v)3\deg(v)\geq 3 for every vVert(G)v\in\operatorname{Vert}(G). Let \mathcal{R} be the finite subdivision rule of the face-inversion of GG and f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\rightarrow S_{\mathcal{R}} be its subdivision map. Let τ\tau be any orientation reversing automorphism of GG. Then the followings are equivalent:

  • (1)

    GG is 33-edge-connected.

  • (2)

    f2:(S2,Vert(G))f^{2}:(S^{2},\operatorname{Vert}(G))\righttoleftarrow does not have a Levy cycle.

  • (2’)

    f2:(S2,Vert(G))f^{2}:(S^{2},\operatorname{Vert}(G))\righttoleftarrow does not have a Thurston obstruction.

  • (3)

    fτ:(S2,Vert(G))f_{\tau}:(S^{2},\operatorname{Vert}(G))\righttoleftarrow does not have a Levy cycle.

  • (3’)

    fτ:(S2,Vert(G))f_{\tau}:(S^{2},\operatorname{Vert}(G))\righttoleftarrow does not have a Thurston obstruction.

Proof.

A level-0 band b=(t;e1,e2)b=(t;e_{1},e_{2}) is non-separating if and only if there is another level-0 band b=(t;e1,e2)b^{\prime}=(t^{\prime};e^{\prime}_{1},e^{\prime}_{2}) such that e1=e1e_{1}=e_{1}^{\prime}, e2=e2e_{2}=e_{2}^{\prime} and ttt\neq t^{\prime}. If such bands bb and bb^{\prime} exist, the removal of two edges of GG intersecting the bones of these bands disconnects GG, i.e., GG is not 33-edge-connected. Conversely, if GG is not 33-edge-connected, then such level-0 bands bb and bb^{\prime} exist. Hence GG is 33-edge-connected if and only if every level-0 band is non-separating. In the case, the level-0 non-expanding spine for sq\mathcal{R}_{sq} is an empty set. Then (1)(2)(2)(1)\Rightarrow(2)\Leftrightarrow(2^{\prime}) follow from Theorem 8.6.

Assume GG is not 33-edge-connected so that there are bands bb and bb^{\prime} described as in the previous paragraph. The union of bones of bb and bb^{\prime} is a homotopically infinite circle contained in the level-0 non-expanding spine N0N^{0} of sq\mathcal{R}_{sq}. Hence (2)(1)(2)\Rightarrow(1) follows from Theorem 8.6.

The equivalence (2)(3)(2)\Leftrightarrow(3) follows from the fact that the subdivisions n(S)\mathcal{R}^{n}(S_{\mathcal{R}}) and τn(Sτ)\mathcal{R}_{\tau}^{n}(S_{\mathcal{R}_{\tau}}) have the same CW-complex structure. The level-2n2n non-expanding spine of τ\mathcal{R}_{\tau} is equal to the level-nn non-expanding spine of sq\mathcal{R}_{sq} for n0n\geq 0. ∎

Remark 9.5.

The equivalence (1)(2)(2)(1)\Leftrightarrow(2)\Leftrightarrow(2^{\prime}) is also shown in [Gey20, Theorem 5.8] and [LLM20, Proposition 4.10].

Remark 9.6.

For an orientation revering branched covering ff, f2f^{2} is combinatorially equivalent to a rational map if and only if ff is combinatorially equivalent to a anti-rational map. See [Gey20, Theorem 3.9] and [LLMM19, Proposition 6.1].

Remark 9.7.

Even if there exists a vertex vv with deg(v)=2\deg(v)=2, the construction is still well-defined, but vv is not a critical point. Note that such vertex vv can be removed from the vertex set without any change in the face-inversion construction.

Example 9.8.
Refer to caption
AA
BB
CC
DD
EE
Refer to caption
AA
AA
AA
AA
BB
BB
BB
BB
CC
CC
CC
DD
EE
EE
EE
EE
CC
DD
DD
DD
Refer to caption
AA
AA
AA
AA
BB
BB
BB
BB
CC
CC
CC
DD
EE
EE
EE
EE
CC
DD
DD
DD
Refer to caption
ff
Refer to caption
fτf_{\tau}
Figure 11. Finite subdivision rules defined from the face-inversion of a planar graph.
Refer to caption
Figure 12. Julia set of z1.50351z2(z21.15757z0.596204)z+0.133305z\mapsto-\frac{1.50351z^{2}(z^{2}-1.15757z-0.596204)}{z+0.133305}

See Figure 11. Let GG be the graph in the bottom and τ\tau the reflection along the middle horizontal line. Then the left and right subdivisions represent τ\mathcal{R}_{\tau} and \mathcal{R} respectively.

In order to obtain an explicit formula of fτ(z)=p(z)q(z)f_{\tau}(z)=\frac{p(z)}{q(z)}, we normalize three vertices on the axis of τ\tau in Figure 11 from left to right to 0,10,1, and \infty. Note that degz(fτ)\deg_{z}(f_{\tau}) is 22 at z=0z=0 or 11, and degz(fτ)\deg_{z}(f_{\tau}) is 33 at z=z=\infty. Since 0 and \infty are fixed points, p(z)p(z) is a quartic polynomial divided by z2z^{2}, and q(z)q(z) is a linear polynomial. We may assume that q(z)q(z) is monic. The conditions that (1) z=1z=1 is a critical fixed point and (2) the other two critical points are exchanged by f(z)f(z) give rise to a system of equation about coefficients of p(z)p(z) and q(z)q(z). Solving this numerically, we have

fτ(z)=1.50351z2(z21.15757z0.596204)z+0.133305.f_{\tau}(z)=-\frac{1.50351z^{2}(z^{2}-1.15757z-0.596204)}{z+0.133305}.

See Figure 12 for the Julia set.

9.3. Finite subdivision rules with essential non-expanding spines at higher levels

In this subsection, we prove Proposition 9.9 by constructing an example.

Proposition 9.9.

For every N>0N>0, there exists a finite subdivision rule N\mathcal{R}_{N} with the subdivision map fN:N(SN)SNf_{\mathcal{R}_{N}}:\mathcal{R}_{N}(S_{\mathcal{R}_{N}})\to S_{\mathcal{R}_{N}} of degree 66 such that (1) Vert(SN)=PfN\operatorname{Vert}(S_{\mathcal{R}_{N}})=P_{f_{\mathcal{R}_{N}}}, (2) the level-kk non-expanding spine NkN^{k} is essential relative to Vert(SN)\operatorname{Vert}(S_{\mathcal{R}_{N}}) for k<Nk<N, and (3) NkN^{k} is not essential relative to Vert(SN)\operatorname{Vert}(S_{\mathcal{R}_{N}}) for kNk\geq N.

Refer to caption

B

A

Refer to caption
ff
Refer to caption
Figure 13. A degree 66 finite subdivision rule with six tile types.

Let us see the finite subdivision rule \mathcal{R} in Figure 13. The 11-skeleton at level-0 is drawn by bold curves. The non-expanding spines N0N^{0} and N1N^{1} are drawn by dotted curves. The N0N^{0} is essential but N1N^{1} is homotopically trivial. Let ff be the subdivision map described in Figure 13. We modify the finite subdivision rule \mathcal{R} into \mathcal{R}^{\prime} as follows:

  1. (1)

    Change the labels AA and BB into A1A_{1} and B1B_{1}.

  2. (2)

    For 2in2\leq i\leq n, we draw n1n-1 copies of annuli consisting of AiA_{i} and BiB_{i} in a row on the left of the annulus consisting of A1A_{1} and B1B_{1} at level-0. Denote by SS_{\mathcal{R}^{\prime}} the modified level-0 CW-complex. Define (S)=f1(S)\mathcal{R}^{\prime}(S_{\mathcal{R}^{\prime}})=f^{-1}(S_{\mathcal{R}^{\prime}}).

    [Uncaptioned image]

    B2B_{2}

    A2A_{2}

    [Uncaptioned image]

    BnB_{n}

    AnA_{n}

    [Uncaptioned image]

    B1B_{1}

    A1A_{1}

    [Uncaptioned image]
  3. (3)

    Let σ\sigma be an orientation preserving homeomorphism of the 22-sphere such that σ(S)=S\sigma(S_{\mathcal{R}^{\prime}})=S_{\mathcal{R}^{\prime}} and σ(Ai)=Ai+1\sigma(A_{i})=A_{i+1} and σ(Bi)=Bi+1\sigma(B_{i})=B_{i+1} for any 1in1\leq i\leq n, where indices are considered modulo nn. That is, σ\sigma is a 1/n1/n-rotation. Define the subdivision map f:(S)Sf^{\prime}:\mathcal{R}^{\prime}(S_{\mathcal{R}^{\prime}})\to S_{\mathcal{R}^{\prime}} by f=σff^{\prime}=\sigma\circ f.

Let NiN^{\prime i} be the level-ii non-expanding spine of \mathcal{R}^{\prime}. The N0N^{\prime 0} is nn-copies of circles. The N1N^{\prime 1} is the union of n1n-1-copies of circles with three non-closed curves. Similarly, for k<nk<n, the level-kk non-expanding spine NkN^{\prime k} has nkn-k circles and some non-closed curves. Therefore, NiN^{\prime i} is essential if i<ni<n and non-essential if ini\geq n.

9.4. Edge-edge expansion vs. edge subdivisions

Let us further investigate the equivalence between the existence of Levy cycles and of Thurston obstructions. Recall that the coefficients of Thurston linear transformation are defined by

fΓ(γ)=γf1(γ)1deg(f|γ:γγ)[γ]Γ.f_{\Gamma}(\gamma)=\sum\limits_{\gamma^{\prime}\subset f^{-1}(\gamma)}\frac{1}{\deg(f|_{\gamma^{\prime}}:\gamma^{\prime}\to\gamma)}[\gamma^{\prime}]_{\Gamma}.

In the setting of finite subdivision rules, the summands 1deg\frac{1}{\deg} are related to the expansion between edges and the number of summands is related to the growth rate of edge subdivisions. Hence, we can expect there is no Thurston obstruction if the edge-edge expansion dominates the edge subdivisions. See [CPT16, Theorem 8.4] for a similar comparison.

Let \mathcal{R} be a finite subdivision rule.

Edge-edge expansion:

We say that \mathcal{R} is edge-edge λ\lambda-expanding for λ1\lambda\geq 1 if there exists C>0C>0 such that for any n0n\geq 0 and any bone γ\gamma of level-0 band, the level-nn subdivision complex n(S)\mathcal{R}^{n}(S_{\mathcal{R}}) subdivides γ\gamma into at least CλnC\cdot\lambda^{n} segments .

Edge subdivision rate:

For any level-0 edge ee, the exponential growth rate of subdivisions of ee is the number νe1\nu_{e}\geq 1 with limn(#{level-nsubedgesofe})1/n=νe\lim_{n\to\infty}(\#\{\mathrm{level\mbox{-}n\leavevmode\nobreak\ subedges\leavevmode\nobreak\ of\leavevmode\nobreak\ }e\})^{1/n}=\nu_{e}. The maximum ν:=maxνe\nu:=\max\leavevmode\nobreak\ \nu_{e} over level-0 edges ee is called the maximal exponential growth rate of edge subdivisions.

Proposition 9.10.

Let \mathcal{R} be a finite subdivision rule. Let ν\nu be the maximal exponential growth rate of edge subdivisions. If \mathcal{R} is edge-edge λ\lambda-expanding for some λ>1\lambda>1, then the non-expanding spine of \mathcal{R} is empty so that the subdivision map f:(S2,A)f:(S^{2},A)\righttoleftarrow does not have a Levy cycle for any set of marked points AA. Moreover, if λ>ν\lambda>\nu, then the subdivision map f:(S2,A)f:(S^{2},A)\righttoleftarrow does not have a Thurston obstruction for any set of marked points AA.

Proof.

Let Γ={γ1,γ2,,γk}\Gamma=\{\gamma_{1},\gamma_{2},\dots,\gamma_{k}\} be a multicurve of (S2,A)(S^{2},A). For a closed curve γ\gamma transverse to n(S)\mathcal{R}^{n}(S_{\mathcal{R}}), we denote by ln(γ)l_{n}(\gamma) the cardinality of the intersection (n(S))(1)γ(\mathcal{R}^{n}(S_{\mathcal{R}}))^{(1)}\cap\gamma. Let C0:=minijl0(γi)l0(γj)C_{0}:=\min_{i\neq j}\frac{l_{0}(\gamma_{i})}{l_{0}(\gamma_{j})}.

Let γiΓ\gamma_{i}\in\Gamma and γi\gamma_{i}^{\prime} be a connected component fn(γi)f^{-n}(\gamma_{i}) that is isotopic to γj\gamma_{j}. Since \mathcal{R} is edge-edge λ\lambda-expanding, there exists C1>0C_{1}>0, which is independent of the choices of γi\gamma_{i} and γi\gamma^{\prime}_{i}, such that

The first part about the emptiness of non-expanding spines immediately follows from the definition of non-expanding spine. Let us assume λ>ν\lambda>\nu and show the second part. Let Γ={γ1,γ2,,γk}\Gamma=\{\gamma_{1},\gamma_{2},\dots,\gamma_{k}\} be a multicurve of (S2,A)(S^{2},A). For a closed curve γ\gamma transverse to n(S)\mathcal{R}^{n}(S_{\mathcal{R}}), we denote by ln(γ)l_{n}(\gamma) the cardinality of the intersection (n(S))(1)γ(\mathcal{R}^{n}(S_{\mathcal{R}}))^{(1)}\cap\gamma. Let C0:=minijl0(γi)l0(γj)C_{0}:=\min_{i\neq j}\frac{l_{0}(\gamma_{i})}{l_{0}(\gamma_{j})}. Since \mathcal{R} is edge-edge λ\lambda-expanding, there exists C1>0C_{1}>0 such that for any γiΓ\gamma_{i}\in\Gamma and for any connected component γi\gamma_{i}^{\prime} of fn(γi)f^{-n}(\gamma_{i}), we have

ln(γi)C1λ˙nl0(γj)C0C1λ˙nl0(γi)\begin{array}[]{ccl}l_{n}(\gamma^{\prime}_{i})&\geq&C_{1}\dot{\lambda}^{n}\cdot l_{0}(\gamma_{j})\\ &\geq&C_{0}C_{1}\dot{\lambda}^{n}\cdot l_{0}(\gamma_{i})\end{array}

for any n0n\geq 0 where γjΓ\gamma_{j}\in\Gamma is isotopic to γi\gamma^{\prime}_{i}. Then deg(f:γiγi)C0C1λn\deg(f:\gamma_{i}^{\prime}\to\gamma_{i})\geq C_{0}C_{1}\lambda^{n}.

Let C2C_{2} be the minimal number satisfying |eΓ|C2|e\cap\Gamma|\leq C_{2} for any level-0 edge ee. It follows that (a) for any level-nn edge ene^{n}, |enfn(Γ)|C2|e^{n}\cap f^{-n}(\Gamma)|\leq C_{2}. By the definition of ν\nu, there exists C3>0C_{3}>0 such that for any level-0 edge ee and any n0n\geq 0, the number of level-nn subedges of ee is at most C3νnC_{3}\cdot{\nu^{\prime}}^{n} for some ν{\nu^{\prime}} with λ>ν>ν\lambda>\nu^{\prime}>\nu. Consider a concatenation α\alpha of level-0 edges connecting two points a1,a2Aa_{1},a_{2}\in A with a1a_{1} and a2a_{2} being in the different Jordan domains of γi\gamma_{i}. Note that (b) every simple closed curve isotopic to γi\gamma_{i} has at least one intersection point with α\alpha. We have

#{componentsoffn(Γ)homotopictoγi}thecardinalityoffn(Γ)αC2#{level-nsubedgesofα}C2C3|Edge(S)|νn.\begin{array}[]{ccl}\#\{\mathrm{components\leavevmode\nobreak\ of}\leavevmode\nobreak\ f^{-n}(\Gamma)\leavevmode\nobreak\ \mathrm{homotopic\leavevmode\nobreak\ to\leavevmode\nobreak\ }\gamma_{i}\}&\leq&\mathrm{the\leavevmode\nobreak\ cardinality\leavevmode\nobreak\ of\leavevmode\nobreak\ }f^{-n}(\Gamma)\cap\alpha\\ &\leq&C_{2}\cdot\#\{\mathrm{level\mbox{-}n\leavevmode\nobreak\ subedges\leavevmode\nobreak\ of\leavevmode\nobreak\ }\alpha\}\\ &\leq&C_{2}C_{3}\cdot|\operatorname{Edge}(S_{\mathcal{R}})|\cdot{\nu^{\prime}}^{n}.\end{array}

The first inequality follows from (b), the second follows from (a) and the third follows from the fact that α\alpha is a concatenation of at most |Edge(S)||\operatorname{Edge}(S_{\mathcal{R}})| edges at level-0.

Hence, every entry in the Thurston linear transformation is bounded from above by

C2C3|Edge(S)|νnC0C1λn,\frac{C_{2}C_{3}\cdot|\operatorname{Edge}(S_{\mathcal{R}})|\cdot{\nu^{\prime}}^{n}}{C_{0}C_{1}\lambda^{n}},

which tends to 0 as nn\to\infty. Then the map ff does not allow any Thurston obstruction. ∎

There are many ways to improve Proposition 9.10. Here are two possible directions.

Suggestion 1:

Proposition 9.10 can be compared with [FPP18b, Proposition 5.1], which states that if \mathcal{R} is edge separating and vertex separating, then the subdivision map does not have a Levy cycle. One difference is that the subdivision map in Proposition 9.10 has to be of hyperbolic-type, i.e., every critical point is a Fatou point, but [FPP18b, Proposition 5.1] works for any subdivision maps. The definition of edge separation in [FPP18b] is the edge-edge expansion, defined in this article, only for pairs of edges that do not share end points. The vertex separation might be necessary only for Julia vertices. One might be able to obtain a stronger result by combining these two propositions.

Suggestion 2:

It would be possible to combine Theorem 7.6 and Proposition 9.10 to obtain a stronger sufficient condition for the equivalence between the existence of Levy and Thurston obstructions. We might be able to (1) have the equivalence on the part where edges subdivide polynomially fast and (2) exclude Thurston obstructions where edges subdivide exponentially fast by assuming the condition in Proposition 9.10.

Example 9.11.
Refer to caption
uu
vv
ww
Figure 14. A finite subdivision rule with λ>ν\lambda>\nu

See Figure 14. We think of the doubles of the the left triangle and get the level-0 subdivision complex SS_{\mathcal{R}} with two tiles. Similarly, take the double of the right large triangle, which is subdivided into 12 small triangles, and define it as the level-11 complex (S)\mathcal{R}(S_{\mathcal{R}}). Then Figure 14 defines a finite subdivision rule \mathcal{R} with the subdivision map f:(S)Sf:\mathcal{R}(S_{\mathcal{R}})\to S_{\mathcal{R}} which is defined by a map sending each small triangle on the right to a triangle on the left or its copy with the types of edges being preserved. Then deg(f)=12\deg(f)=12 and ff has three critical values u,vu,v, and ww, which are vertices of the level-0 triangles. Moreover, f(u)=f(v)=f(w)=wf(u)=f(v)=f(w)=w.

It is immediate that the non-expanding spine is an empty set. By Theorem 6.21, ff does not have a Levy cycle. Since \mathcal{R} has exponential growth rate of edge subdivisions, we cannot apply Theorem 8.6 to claim that ff does not have a Thurston obstruction. However, it is easy to show that λ=4\lambda=4 and ν=2\nu=2, and then ff does not have a Thurston obstruction by Proposition 9.10.

References

  • [BBY12] Sylvain Bonnot, Mark Braverman, and Michael Yampolsky, Thurston equivalence to a rational map is decidable, Mosc. Math. J. 12 (2012), no. 4, 747–763, 884. MR 3076853
  • [BD18] Laurent Bartholdi and Dzmitry Dudko, Algorithmic aspects of branched coverings IV/V. Expanding maps, Trans. Amer. Math. Soc. 370 (2018), no. 11, 7679–7714. MR 3852445
  • [BLMW19] James Belk, Justin Lanier, Dan Margalit, and Rebecca R. Winarski, Recognizing topological polynomials by lifting trees, 2019, arXiv:1906.07680.
  • [BM17] Mario Bonk and Daniel Meyer, Expanding Thurston maps, Mathematical Surveys and Monographs, vol. 225, American Mathematical Society, Providence, RI, 2017. MR 3727134
  • [BP94] Abraham Berman and Robert J. Plemmons, Nonnegative matrices in the mathematical sciences, Classics in Applied Mathematics, vol. 9, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 1994, Revised reprint of the 1979 original. MR 1298430
  • [CFKP03] J. W. Cannon, W. J. Floyd, R. Kenyon, and W. R. Parry, Constructing rational maps from subdivision rules, Conform. Geom. Dyn. 7 (2003), 76–102. MR 1992038
  • [CFP01] J. W. Cannon, W. J. Floyd, and W. R. Parry, Finite subdivision rules, Conform. Geom. Dyn. 5 (2001), 153–196. MR 1875951
  • [CPT12] Guizhen Cui, Wenjuan Peng, and Lei Tan, On a theorem of Rees-Shishikura, Ann. Fac. Sci. Toulouse Math. (6) 21 (2012), no. 5, 981–993. MR 3088264
  • [CPT16] by same author, Renormalizations and wandering Jordan curves of rational maps, Comm. Math. Phys. 344 (2016), no. 1, 67–115. MR 3493138
  • [DH93] Adrien Douady and John H. Hubbard, A proof of Thurston’s topological characterization of rational functions, Acta Math. 171 (1993), no. 2, 263–297. MR 1251582
  • [FM12] Benson Farb and Dan Margalit, A primer on mapping class groups, Princeton Mathematical Series, vol. 49, Princeton University Press, Princeton, NJ, 2012. MR 2850125
  • [FPP18a] William Floyd, Walter Parry, and Kevin M. Pilgrim, Rationality is decidable for nearly euclidean thurston maps, 2018, arXiv:1812.01066.
  • [FPP18b] William J. Floyd, Walter R. Parry, and Kevin M. Pilgrim, Expansion properties for finite subdivision rules I, Sci. China Math. 61 (2018), no. 12, 2237–2266. MR 3881956
  • [FPP20] William Floyd, Walter Parry, and Kevin M. Pilgrim, Expansion properties for finite subdivision rules II, Conform. Geom. Dyn. 24 (2020), 29–50. MR 4051834
  • [Gey20] Lukas Geyer, Classification of critically fixed anti-rational maps, 2020, arXiv:2006.10788.
  • [GHMZ18] Yan Gao, Peter Haïssinsky, Daniel Meyer, and Jinsong Zeng, Invariant Jordan curves of Sierpiński carpet rational maps, Ergodic Theory Dynam. Systems 38 (2018), no. 2, 583–600. MR 3774834
  • [Hlu17] Mikhail Hlushchanka, Invariant graphs, tilings, and iterated monodromy groups, Ph.D. thesis, Jacobs University, 2017.
  • [Hlu19] by same author, Tischler graphs of critically fixed rational maps and their applications, 2019, arXiv:1904.04759.
  • [HS94] John H. Hubbard and Dierk Schleicher, The spider algorithm, Complex dynamical systems (Cincinnati, OH, 1994), Proc. Sympos. Appl. Math., vol. 49, Amer. Math. Soc., Providence, RI, 1994, pp. 155–180. MR 1315537
  • [Hub16] John Hamal Hubbard, Teichmüller theory and applications to geometry, topology, and dynamics. Vol. 2, Matrix Editions, Ithaca, NY, 2016, Surface homeomorphisms and rational functions. MR 3675959
  • [Lev85] Silvio Vieira Ferreira Levy, Critically finite rational maps (Thurston), Ph.D. thesis, Princeton University, 1985, p. 65. MR 2634168
  • [LLM20] Russell Lodge, Yusheng Luo, and Sabyasachi Mukherjee, Circle packings, kissing reflection groups and critically fixed anti-rational maps, 2020, arXiv:2007.03558.
  • [LLMM19] Russell Lodge, Mikhail Lyubich, Sergei Merenkov, and Sabyasachi Mukherjee, On dynamical gaskets generated by rational maps, Kleinian groups, and Schwarz reflections, 2019, arXiv:1912.13438.
  • [LMS15] Russell Lodge, Yauhen Mikulich, and Dierk Schleicher, Combinatorial properties of newton maps, 2015, arXiv:1510.02761.
  • [MS80] M. Misiurewicz and W. Szlenk, Entropy of piecewise monotone mappings, Studia Math. 67 (1980), no. 1, 45–63. MR 579440
  • [Nek05] Volodymyr Nekrashevych, Self-similar groups, Mathematical Surveys and Monographs, vol. 117, American Mathematical Society, Providence, RI, 2005. MR 2162164
  • [PH92] R. C. Penner and J. L. Harer, Combinatorics of train tracks, Annals of Mathematics Studies, vol. 125, Princeton University Press, Princeton, NJ, 1992. MR 1144770
  • [Poi10] Alfredo Poirier, Hubbard trees, Fund. Math. 208 (2010), no. 3, 193–248. MR 2650982
  • [PT98] Kevin M. Pilgrim and Lei Tan, Combining rational maps and controlling obstructions, Ergodic Theory Dynam. Systems 18 (1998), no. 1, 221–245. MR 1609463
  • [Ree92] Mary Rees, A partial description of parameter space of rational maps of degree two. I, Acta Math. 168 (1992), no. 1-2, 11–87. MR 1149864
  • [Shi00] Mitsuhiro Shishikura, On a theorem of M. Rees for matings of polynomials, The Mandelbrot set, theme and variations, London Math. Soc. Lecture Note Ser., vol. 274, Cambridge Univ. Press, Cambridge, 2000, pp. 289–305. MR 1765095
  • [Sid00] Said Sidki, Automorphisms of one-rooted trees: growth, circuit structure, and acyclicity, J. Math. Sci. (New York) 100 (2000), no. 1, 1925–1943, Algebra, 12. MR 1774362
  • [ST00] Mitsuhiro Shishikura and Lei Tan, A family of cubic rational maps and matings of cubic polynomials, Experiment. Math. 9 (2000), 29–53. MR 1758798
  • [ST19] Anastasia Shepelevtseva and Vladlen Timorin, Invariant spanning trees for quadratic rational maps, Arnold Math. J. 5 (2019), no. 4, 435–481. MR 4068871
  • [Tan92] Lei Tan, Matings of quadratic polynomials, Ergodic Theory Dynam. Systems 12 (1992), no. 3, 589–620. MR 1182664
  • [Thu20] Dylan P. Thurston, A positive characterization of rational maps, Ann. of Math. (2) 192 (2020), no. 1, 1–46. MR 4125449
  • [Ufn82] V. A. Ufnarovskiĭ, Criterion for the growth of graphs and algebras given by words, Mat. Zametki 31 (1982), no. 3, 465–472, 476. MR 652851