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

Closed continuations of Riemann surfaces

Makoto Masumoto and Masakazu Shiba

Abstract. Any open Riemann surface R0R_{0} of finite genus gg can be conformally embedded into a closed Riemann surface of the same genus, that is, R0R_{0} is realized as a subdomain of a closed Riemann surface of genus gg. We are concerned with the set 𝔐(R0)\mathfrak{M}(R_{0}) of such closed Riemann surfaces. We formulate the problem in the Teichmüller space setting to investigate geometric properties of 𝔐(R0)\mathfrak{M}(R_{0}). We show, among other things, that 𝔐(R0)\mathfrak{M}(R_{0}) is a closed Lipschitz domain homeomorphic to a closed ball provided that R0R_{0} is nonanalytically finite.
Keywords: Riemann surface, conformal embedding, Teichmüller space, quadratic differential, measured foliation, extremal length
MSC2020: 30Fxx, 32G15

1 Introduction

Let R0R_{0} be an open Riemann surface of finite genus gg. If g=0g=0, then the general uniformization theorem assures us that R0R_{0} is conformally equivalent to a domain on the Riemann sphere ^:={}\hat{\mathbb{C}}:=\mathbb{C}\cup\{\infty\}. In 1928 Bochner generalized it to the cases of higher genera in his study [10] on continuations of Riemann surfaces, showing that R0R_{0} can be realized as a subdomain of a closed Riemann surface of the same genus gg. While a closed Riemann surface of genus zero is essentially unique, this is not the case for closed Riemann surfaces of positive genus so that there may be two or more closed Riemann surfaces of genus gg with a domain conformally equivalent to R0R_{0} provided that g>0g>0. It is then natural to ask into which closed Riemann surfaces of genus gg the Riemann surface R0R_{0} can be conformally embedded.

Heins [22] tackled the problem for g=1g=1, and proved in 1953 that the set of closed Riemann surfaces of genus one which are continuations of R0R_{0} is relatively compact in the moduli space of genus one. Four years later Oikawa [45] formulated the problem in the context of Teichmüller spaces, and discovered that the set of marked closed Riemann surfaces of genus gg into which R0R_{0} can be mapped by a homotopically consistent conformal embedding is compact and connected in the Teichmüller space of genus gg. In 1987 the second author [56] improved Oikawa’s result in the case of genus one by deducing that Oikawa’s set is in fact a closed disk, which may degenerate to a singleton, with respect to the Teichmüller distance. The authors then have been developing the theory mainly in the framework of Torelli spaces (see, for example, Masumoto [37, 38], Schmieder-Shiba [51, 52, 53] and Shiba [57, 58, 60, 61, 62]). For other results related to conformal embeddings of Riemann surfaces see Bourque [12], Earle-Marden [14], Fehlmann-Gardiner [15], Hamano [19], Hamano-Shiba-Yamaguchi [20], Horiuchi-Shiba [23], Ioffe [25], Ito-Shiba [27, 28], Kahn-Pilgrim-Thurston [30], Masumoto [39, 40], Sasai [49, 50], Shiba [55, 59] and Shiba-Shibata [63, 64]. Applications of conformal embedding theory to hyperbolic geometry and holomorphic mappings are found in Bourque [11] and Masumoto [36, 38, 41, 42, 43].

In the present paper we address the problem for finite open Riemann surfaces of positive genus in the Teichmüller space setting. One of our main purposes is to generalize our previous results in [56] to the cases of higher genera.

To formulate the problem we introduce the category 𝔉g\mathfrak{F}_{g} of marked Riemann surfaces of finite genus gg and conformally compatible continuous mappings in §3. Its class of objects consists of equivalence classes 𝑺=[S,η]{\bm{S}}=[S,\eta], where SS is a Riemann surface of genus gg, not necessarily closed, and η\eta is a sense-preserving homeomorphism of Σ˙g\dot{\Sigma}_{g} into SS, where Σ˙g\dot{\Sigma}_{g} is a fixed surface obtained from a closed oriented topological surface Σg\Sigma_{g} of genus gg by deleting one point (see Definition 3.5). A continuous mapping of 𝑺\bm{S} into another marked Riemann surface 𝑺=[S,η]𝔉g{\bm{S}}^{\prime}=[S^{\prime},\eta^{\prime}]\in\mathfrak{F}_{g} is an equivalence class 𝒇=[f,η,η]{\bm{f}}=[f,\eta,\eta^{\prime}], where ff is a continuous mapping of SS into SS^{\prime} (see Definition 3.9). The composition 𝒇𝒇{\bm{f}}^{\prime}\circ{\bm{f}} is meaningful only for conformally compatible continuous mappings (see Definition 3.10). Also, a quadratic differential on 𝑺{\bm{S}} is an equivalence class 𝝋=[φ,η]{\bm{\varphi}}=[\varphi,\eta], where φ\varphi is a quadratic differential on SS (see §6).

The Teichmüller space 𝔗g\mathfrak{T}_{g} of genus gg is defined to be the full-subcategory of 𝔉g\mathfrak{F}_{g} whose class of objects is composed of all marked closed Riemann surfaces of genus gg. Equipped with the Teichmüller distance, 𝔗g\mathfrak{T}_{g} is a metric space homeomorphic to 2dg\mathbb{R}^{2d_{g}}, where dg=max{g,3g3}d_{g}=\max\{g,3g-3\}. In fact, as is well-known, 𝔗g\mathfrak{T}_{g} is a dgd_{g}-dimensional complex manifold biholomorphic to a bounded domain in dg\mathbb{C}^{d_{g}}. It is quite new to understand the Teichmüller space 𝔗g\mathfrak{T}_{g} as a full-subcategory of 𝔉g\mathfrak{F}_{g}. This setting fits our theory because we are required to deal with not only quasiconformal mappings between marked closed Riemann surfaces but also continuous mappings of marked compact bordered Riemann surfaces into marked compact Riemann surfaces.

Let 𝑹0=[R0,θ0]{\bm{R}}_{0}=[R_{0},\theta_{0}] be a marked finite open Riemann surface of positive genus gg; the fundamental group of R0R_{0} is finitely generated, or equivalently, R0R_{0} has finitely many handles and boundary components in the sense of Kerékjártó-Stoïlow (for the definition of a boundary component, see [3, I.36B]). It is said to be analytically finite if R0R_{0} is conformally equivalent to a Riemann surface obtained from a closed Riemann surface by removing finitely many points. For 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g} let CEmbhc(𝑹0,𝑹)\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) be the set of homotopically consistent conformal embeddings of 𝑹0{\bm{R}}_{0} into 𝑹\bm{R} (see Definition 3.12). By a closed continuation of 𝑹0{\bm{R}}_{0} we mean a pair (𝑹,𝜾)({\bm{R}},{\bm{\iota}}), where 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g} and 𝜾CEmbhc(𝑹0,𝑹){\bm{\iota}}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}). We are concerned with the set 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) of 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g} for which CEmbhc(𝑹0,𝑹)\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}})\neq\varnothing.

Theorem 1.1.

Let 𝐑0{\bm{R}}_{0} be a marked finite open Riemann surface of positive genus gg. Then 𝔐(𝐑0)\mathfrak{M}({\bm{R}}_{0}) is either a singleton or a closed Lipschitz domain homeomorphic to a closed ball in 2dg\mathbb{R}^{2d_{g}}. The former case occurs if and only if 𝐑0{\bm{R}}_{0} is analytically finite.

The theorem is proved in the final section, where we show a stronger assertion that there is a homeomorphism of 𝔗g\mathfrak{T}_{g} onto 2dg\mathbb{R}^{2d_{g}} that maps 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) onto a point or a closed ball. Being a closed Lipschitz domain, 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) satisfies inner and outer cone conditions unless 𝑹0{\bm{R}}_{0} is analytically finite. Actually, it satisfies an outer ball condition. If g=1g=1, then its boundary 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) is smooth and a sphere (or rather circle) with respect to the Teichmüller distance. On the other hand, if g>1g>1, then the boundary is nonsmooth for some 𝑹0{\bm{R}}_{0} and hence cannot be a sphere with respect to the Teichmüller distance.

For K1K\geqq 1 let 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) denote the set of 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g} into which 𝑹0{\bm{R}}_{0} can be mapped by a homotopically consistent KK-quasiconformal embedding. Thus 𝔐1(𝑹0)=𝔐(𝑹0)\mathfrak{M}_{1}({\bm{R}}_{0})=\mathfrak{M}({\bm{R}}_{0}). If K>1K>1, then 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) is the closed (logK)/2(\log K)/2-neighborhood of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}).

Theorem 1.2.

Let 𝐑0{\bm{R}}_{0} be a marked finite open Riemann surface of positive genus gg. If K>1K>1, then 𝔐K(𝐑0)\mathfrak{M}_{K}({\bm{R}}_{0}) is a closed domain homeomorphic to a closed ball in 2dg\mathbb{R}^{2d_{g}} and has a C1C^{1}-boundary.

Actually, there is a homeomorphism of 𝔗g\mathfrak{T}_{g} onto 2dg\mathbb{R}^{2d_{g}} such that 𝔐K(𝑹0)\mathfrak{M}_{K^{\prime}}({\bm{R}}_{0}), KKK^{\prime}\geqq K, are mapped onto concentric closed balls. Theorem 1.2 is not a consequence of Theorem 1.1 but a tool for the proof of Theorem 1.1.

For the proofs of Theorems 1.1 and 1.2 we introduce a procedure called self-weldings of marked compact bordered Riemann surfaces 𝑺=[S,η]{\bm{S}}=[S,\eta], which is also useful to construct various examples of closed continuations. Let A+(𝑺)A_{+}({\bm{S}}) denote the set of nonzero holomorphic quadratic differentials 𝝋=[φ,η]{\bm{\varphi}}=[\varphi,\eta] on 𝑺\bm{S} with φ0\varphi\geqq 0 along the border S\partial S. We use 𝝋A+(𝑺){\bm{\varphi}}\in A_{+}({\bm{S}}) to identify arcs on the border. This operation gives rise to a continuous mapping 𝜿\bm{\kappa} of 𝑺\bm{S} onto a marked closed Riemann surface 𝑹\bm{R} holomorphic and injective on the interior 𝑺{\bm{S}}^{\circ} of 𝑺\bm{S} together with a meromorphic quadratic differential 𝝍\bm{\psi} on 𝑹\bm{R}. The pair 𝑹,𝜿\langle{\bm{R}},{\bm{\kappa}}\rangle is referred to as a closed self-welding of 𝑺\bm{S}, and the quadratic differentials 𝝋\bm{\varphi} and 𝝍\bm{\psi} are called a welder of 𝑹,𝜿\langle{\bm{R}},{\bm{\kappa}}\rangle and the co-welder of 𝝋\bm{\varphi}, respectively. Important are the self-weldings with holomorphic co-welders. Such a self-welding is said to be regular, or 𝝋\bm{\varphi}-regular if we need to refer to the welder 𝝋\bm{\varphi}.

If 𝑹0{\bm{R}}_{0} is nonanalytically finite, that is, it is finite but not analytically finite, then it is considered as the interior of a marked compact bordered Riemann surface 𝑺\bm{S}, possibly, with finitely many points deleted. Each closed self-welding 𝑹,𝜿\langle{\bm{R}},{\bm{\kappa}}\rangle of 𝑺\bm{S} yields a closed continuation (𝑹,𝜿|𝑹0)({\bm{R}},{\bm{\kappa}}|_{{\bm{R}}_{0}}) of 𝑹0{\bm{R}}_{0}, called a closed self-welding continuation of 𝑹0{\bm{R}}_{0}. A welder of the self-welding is also referred to as a welder of the continuation. If 𝑹,𝜿\langle{\bm{R}},{\bm{\kappa}}\rangle is 𝝋\bm{\varphi}-regular, then (𝑹,𝜿|𝑹0)({\bm{R}},{\bm{\kappa}}|_{{\bm{R}}_{0}}) is said to be 𝝋\bm{\varphi}-regular. If this is the case, then 𝜿|𝑹0{\bm{\kappa}}|_{{\bm{R}}_{0}} is a Teichmüller conformal embedding of 𝑹0{\bm{R}}_{0} into 𝑹\bm{R}, and 𝝋\bm{\varphi} is an initial quadratic differential of 𝜿|𝑹0{\bm{\kappa}}|_{{\bm{R}}_{0}}.

Let A+(𝑹0)A_{+}({\bm{R}}_{0}) be the space of nonzero holomorphic quadratic differentials on 𝑹0{\bm{R}}_{0} that can be extended to elements in A+(𝑺)A_{+}({\bm{S}}). Denote by AL(𝑹)A_{L}({\bm{R}}) the set of 𝝋=[φ,θ0]A+(𝑹0){\bm{\varphi}}=[\varphi,\theta_{0}]\in A_{+}({\bm{R}}_{0}) such that for any horizontal trajectory aa of φ\varphi on S\partial S its φ\varphi-length is at most the half of the φ\varphi-length of the component of S\partial S including aa. Clearly, AL(𝑹0)A_{L}({\bm{R}}_{0}) is a proper subset of A+(𝑹0)A_{+}({\bm{R}}_{0}).

Theorem 1.3.

Let 𝐑0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface of positive genus gg. An element of A+(𝐑0)A_{+}({\bm{R}}_{0}) induces a closed regular self-welding continuation of 𝐑0{\bm{R}}_{0} if and only if it belongs to AL(𝐑0)A_{L}({\bm{R}}_{0}).

The theorem shows that the procedure of closed regular self-welding provides us with an explicit method of obtaining all Teichmüller conformal embeddings of 𝑹0{\bm{R}}_{0}. Their importance is clarified in the next theorem.

For 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}) and 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}) let CEmb𝝋(𝑹0,𝑹)\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0},{\bm{R}}) be the set of 𝜾CEmbhc(𝑹0,𝑹){\bm{\iota}}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) such that (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) is a closed 𝝋{\bm{\varphi}}-regular self-welding continuation of 𝑹0{\bm{R}}_{0}; it can be empty. Set CEmb𝝋(𝑹0)=𝑹𝔐(𝑹0)CEmb𝝋(𝑹0,𝑹)\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0})=\bigcup_{{\bm{R}}\in\mathfrak{M}({\bm{R}}_{0})}\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0},{\bm{R}}). If cardCEmb𝝋(𝑹0)>1\operatorname{card}\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0})>1, then 𝝋\bm{\varphi} is called exceptional. Let AE(𝑹0)A_{E}({\bm{R}}_{0}) be the set of exceptional quadratic differentials in AL(𝑹0)A_{L}({\bm{R}}_{0}). Define CEmbL(𝑹0)=𝝋AL(𝑹0)CEmb𝝋(𝑹0)\operatorname{CEmb}_{L}({\bm{R}}_{0})=\bigcup_{{\bm{\varphi}}\in A_{L}({\bm{R}}_{0})}\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0}) and CEmbE(𝑹0)=𝝋AE(𝑹0)CEmb𝝋(𝑹0)\operatorname{CEmb}_{E}({\bm{R}}_{0})=\bigcup_{{\bm{\varphi}}\in A_{E}({\bm{R}}_{0})}\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0}).

For 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}) let 𝔐𝝋(𝑹0)\mathfrak{M}_{\bm{\varphi}}({\bm{R}}_{0}) denote the set of 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}) for which CEmb𝝋(𝑹0,𝑹)\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0},{\bm{R}})\neq\varnothing. If 𝝋AL(𝑹0)AE(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0})\setminus A_{E}({\bm{R}}_{0}), then 𝔐φ(𝑹0)\mathfrak{M}_{\varphi}({\bm{R}}_{0}) is a singleton. Set 𝔐L(𝑹0)=𝝋AL(𝑹0)𝔐𝝋(𝑹0)\mathfrak{M}_{L}({\bm{R}}_{0})=\bigcup_{{\bm{\varphi}}\in A_{L}({\bm{R}}_{0})}\mathfrak{M}_{\bm{\varphi}}({\bm{R}}_{0}) and 𝔐E(𝑹0)=𝝋AE(𝑹0)𝔐𝝋(𝑹0)\mathfrak{M}_{E}({\bm{R}}_{0})=\bigcup_{{\bm{\varphi}}\in A_{E}({\bm{R}}_{0})}\mathfrak{M}_{\bm{\varphi}}({\bm{R}}_{0}).

Theorem 1.4.

Let 𝐑0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface of positive genus gg.

  • (i)

    The boundary 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) coincides with 𝔐L(𝑹0)\mathfrak{M}_{L}({\bm{R}}_{0}).

  • (ii)

    If 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}) and 𝑹𝔐𝝋(𝑹0){\bm{R}}\in\mathfrak{M}_{\bm{\varphi}}({\bm{R}}_{0}), then CEmb𝝋(𝑹0,𝑹)=CEmbhc(𝑹0,𝑹)\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0},{\bm{R}})=\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}).

  • (iii)

    If g3g\geqq 3, then 𝔐E(𝑹0)\mathfrak{M}_{E}({\bm{R}}_{0}) has a nonempty interior with respect to the relative topology on 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}).

We examine the behavior of the extremal length function to prove (i) and (ii) though they also follow from Bourque [11] and Kahn-Pilgrim-Thurston [30]. The set AE(𝑹0)A_{E}({\bm{R}}_{0}) is nowhere dense in AL(𝑹0)A_{L}({\bm{R}}_{0}) (Theorem 10.3). Nevertheless, quadratic differentials in AE(𝑹0)A_{E}({\bm{R}}_{0}) yield abundant boundary points of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) through closed regular self-weldings, as assertion (iii) claims. If g=1g=1, then AE(𝑹0)=A_{E}({\bm{R}}_{0})=\varnothing. If g=2g=2, then there are examples of 𝑹0{\bm{R}}_{0} for which 𝔐E(𝑹0)\mathfrak{M}_{E}({\bm{R}}_{0}) has a nonempty interior in 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}).

Let (Σg)\mathscr{MF}(\Sigma_{g}) be the space of measured foliations on Σg\Sigma_{g}. The horizontal foliation of a holomorphic quadratic differential 𝝍=[ψ,θ]{\bm{\psi}}=[\psi,\theta] on a marked closed Riemann surface 𝑹=[R,θ]{\bm{R}}=[R,\theta] determines an element 𝑹(𝝍)\mathcal{H}_{\bm{R}}({\bm{\psi}}) of (Σg)\mathscr{MF}(\Sigma_{g}) through θ\theta. This defines a homeomorphism 𝑹\mathcal{H}_{\bm{R}} of the space A(𝑹)A({\bm{R}}) of holomorphic quadratic differentials on 𝑹\bm{R} onto (Σg)\mathscr{MF}(\Sigma_{g}) by Hubbard-Masur [24].

Now, for 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}), taking a closed 𝝋\bm{\varphi}-regular self-welding continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) of 𝑹0{\bm{R}}_{0} and denoting by 𝝍\bm{\psi} the co-welder of 𝝋\bm{\varphi}, we set 𝑹0(𝝋)=𝑹(𝝍)\mathcal{H}_{{\bm{R}}_{0}}({\bm{\varphi}})=\mathcal{H}_{\bm{R}}({\bm{\psi}}). Then 𝑹0\mathcal{H}_{{\bm{R}}_{0}} is a well-defined mapping of AL(𝑹0)A_{L}({\bm{R}}_{0}) into (Σg){0}\mathcal{MF}(\Sigma_{g})\setminus\{0\}.

Theorem 1.5.

If 𝐑0{\bm{R}}_{0} is a marked nonanalytically finite open Riemann surface of positive genus gg, then 𝐑0\mathcal{H}_{{\bm{R}}_{0}} is a homeomorphism of AL(𝐑0)A_{L}({\bm{R}}_{0}) onto (Σg){0}\mathscr{MF}(\Sigma_{g})\setminus\{0\}.

The extremal length function Ext\operatorname{Ext} is a nonnegative continuous function on 𝔗g×(Σg)\mathfrak{T}_{g}\times\mathscr{MF}(\Sigma_{g}). Specifically, if (𝑹,)𝔗g×(Σg)({\bm{R}},\mathcal{F})\in\mathfrak{T}_{g}\times\mathscr{MF}(\Sigma_{g}), then Ext(𝑹,)=Ext(𝑹)\operatorname{Ext}({\bm{R}},\mathcal{F})=\operatorname{Ext}_{\mathcal{F}}({\bm{R}}) is exactly 𝑹1()𝑹\|\mathcal{H}_{\bm{R}}^{-1}(\mathcal{F})\|_{\bm{R}}, where 𝝋𝑺\|{\bm{\varphi}}\|_{\bm{S}} stands for the L1L^{1}-norm of a quadratic differential 𝝋\bm{\varphi} on a marked Riemann surface 𝑺\bm{S}.

For (Σg){0}\mathcal{F}\in\mathscr{MF}(\Sigma_{g})\setminus\{0\} let 𝔐(𝑹0)\mathfrak{M}_{\mathcal{F}}({\bm{R}}_{0}) denote the set of points of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) at which the restriction of Ext\operatorname{Ext}_{\mathcal{F}} to 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) attains its maximum. It is nonempty since 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is compact.

Theorem 1.6.

Let 𝐑0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface of positive genus gg.

  • (i)

    Let (Σg){0}\mathcal{F}\in\mathscr{MF}(\Sigma_{g})\setminus\{0\}, and set 𝝋=𝑹01(){\bm{\varphi}}=\mathcal{H}_{{\bm{R}}_{0}}^{-1}(\mathcal{F}). Then 𝔐(𝑹0)=𝔐𝝋(𝑹0)\mathfrak{M}_{\mathcal{F}}({\bm{R}}_{0})=\mathfrak{M}_{\bm{\varphi}}({\bm{R}}_{0}) and

    (1.1) maxExt(𝔐(𝑹0))=maxExt(𝔐(𝑹0))=𝝋𝑹0.\max\operatorname{Ext}_{\mathcal{F}}(\mathfrak{M}({\bm{R}}_{0}))=\max\operatorname{Ext}_{\mathcal{F}}(\partial\mathfrak{M}({\bm{R}}_{0}))=\|{\bm{\varphi}}\|_{{\bm{R}}_{0}}.
  • (ii)

    If g3g\geqq 3, then there exists (Σg){0}\mathcal{F}\in\mathscr{MF}(\Sigma_{g})\setminus\{0\} such that card𝔐(𝑹0)=20\operatorname{card}\mathfrak{M}_{\mathcal{F}}({\bm{R}}_{0})=2^{\aleph_{0}} while cardCEmbhc(𝑹0,𝑹)=1\operatorname{card}\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}})=1 for all 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}_{\mathcal{F}}({\bm{R}}_{0}).

  • (iii)

    Let 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g}. Then 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}) if and only if

    (1.2) Ext(𝑹)maxExt(𝔐(𝑹0))\operatorname{Ext}_{\mathcal{F}}({\bm{R}})\leqq\max\operatorname{Ext}_{\mathcal{F}}(\partial\mathfrak{M}({\bm{R}}_{0}))

    for all (Σg){0}\mathcal{F}\in\mathscr{MF}(\Sigma_{g})\setminus\{0\}.

Kahn-Pilgrim-Thurston [30, Theorem 1] gives a necessary and sufficient condition for 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g} to belong to 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) in terms of stretch factors. A stretch factor of a topological embedding of 𝑹0{\bm{R}}_{0} into 𝑹\bm{R} is defined with the extremal lengths of simple closed multi-curves on 𝑹0{\bm{R}}_{0}. Theorem 1.6 (iii) tells us which multi-curves we should take into account to calculate the stretch factor. Specifically, the authors of [30] employ the Jenkins-Strebel quadratic differentials in A+(𝑹0)A_{+}({\bm{R}}_{0}) to obtain the stretch factors. Our theorem asserts that those in AL(𝑹0)A_{L}({\bm{R}}_{0}) are sufficient.

Besides Theorems 1.31.6, the Teichmüller geodesic rays induced from the co-welders of elements of AL(𝑹0)A_{L}({\bm{R}}_{0}) play an important role in the proof of Theorem 1.2. We then apply analytic properties of the extremal length function to establish Theorem 1.1. Existence of linearly independent elements 𝝋jAL(𝑹0){\bm{\varphi}}_{j}\in A_{L}({\bm{R}}_{0}), j=1,2j=1,2, with 𝔐𝝋1(𝑹0)𝔐𝝋2(𝑹0)\mathfrak{M}_{{\bm{\varphi}}_{1}}({\bm{R}}_{0})\cap\mathfrak{M}_{{\bm{\varphi}}_{2}}({\bm{R}}_{0})\neq\varnothing is one of the biggest obstacles in the proof of Theorem 1.1.

Acknowledgments. This research is supported in part by JSPS KAKENHI Grant Numbers 18K03334 and 22K03556. The authors express sincere thanks to Hideki Miyachi, a brilliant expert on Teichmüller theory, for his invaluable comments and suggestions. Without his help they could not finish the article. They are also grateful to Shuhei Masumoto and Shingo Okuyama for formulating our theory within the scheme of category theory. The two colleagues help the authors write a clearer paper. Last but not least, the authors thank Ken-ichi Sakan for his continuing encouragement.

2 Preliminaries

In the present article we are concerned with conformal embeddings of a Riemann surface into another. We begin with confirming terminology and notation.

Let XX and YY be topological spaces. Denote by Cont(X,Y)\operatorname{Cont}(X,Y) the set of continuous mappings of XX into YY; we avoid employing the usual notation C(X,Y)C(X,Y) so as not to confuse it with the set of conformal mappings in the case where XX and YY are Riemann surfaces. The subset of homeomorphisms of XX onto YY is denoted by Homeo(X,Y)\operatorname{Homeo}(X,Y). A topological embedding of XX into YY is, by definition, a mapping ff of XX into YY for which the correspondence xf(x)x\mapsto f(x) defines an element of Homeo(X,f(X))\operatorname{Homeo}(X,f(X)), where f(X)f(X) is endowed with the relative topology induced from the topology of YY. We denote by TEmb(X,Y)\operatorname{TEmb}(X,Y) the set of topological embeddings of XX into YY. If XX and YY are oriented surfaces, then Homeo+(X,Y)\operatorname{Homeo}^{+}(X,Y) and TEmb+(X,Y)\operatorname{TEmb}^{+}(X,Y) denote the sets of sense-preserving elements in Homeo(X,Y)\operatorname{Homeo}(X,Y) and TEmb(X,Y)\operatorname{TEmb}(X,Y), respectively.

For fkCont(X,Y)f_{k}\in\operatorname{Cont}(X,Y), k=0,1k=0,1, we write f0f1f_{0}\simeq f_{1} if f0f_{0} is homotopic to f1f_{1}. Also, for two loops c0c_{0} and c1c_{1} on XX we use the notation c0c1c_{0}\simeq c_{1} to mean that c0c_{0} is freely homotopic to c1c_{1}. This usage of \simeq is consistent with the previous one as we can regard c0c_{0} and c1c_{1} as continuous mappings of the unit circle into XX.

A surface means a connected 22-manifold whose topology has a countable base. A surface with boundary is often called a bordered surface or a surface with border. As is well-known, every surface is triangulable, and can be endowed with conformal structure, or complex structure provided that it is orientable.

A Riemann surface is a connected complex manifold of dimension one (see Ahlfors-Sario [3, II.1E] and Strebel [65, Definition 1.1]), and a bordered Riemann surface is a connected 1-dimensional complex manifold with boundary (see [3, II.3A] and [65, Definition 1.2]). By abuse of language we refer to bordered Riemann surfaces also as Riemann surfaces. Thus, when we speak of a Riemann surface, it may be a bordered Riemann surface. A bordered Riemann surface is sometimes called a Riemann surface with border. A Riemann surface without border means a Riemann surface which is not a bordered Riemann surface. A Riemann surface without border is called closed (resp. open) if it is compact (resp. noncompact). Thus a compact Riemann surface is either a closed Riemann surface or a compact bordered Riemann surface. By a torus we mean a closed Riemann surface of genus one.

A holomorphic mapping of a Riemann surface into another is called a conformal embedding if it is a topological embedding at the same time, where holomorphic mappings of a bordered Riemann surface are supposed to be analytic on the border. A biholomorphic mapping of a Riemann surface onto another is also referred to as a conformal homeomorphism. Conformal homeomorphisms are conformal embeddings. Two Riemann surfaces are said to be conformally equivalent to each other if there is a conformal homeomorphism of one onto the other.

Example 2.1.

Set S={z1<|z|<2,0argz<π}S=\{z\in\mathbb{C}\mid 1<|z|<2,0\leqq\arg z<\pi\}; it is a bordered Riemann surface. Though the holomorphic mapping f:Sf:S\to\mathbb{C} defined by f(z)=z2f(z)=z^{2} is injective, it is not a conformal embedding since ff is not a homeomorphism of SS onto its image f(S)={w1<|w|<4}f(S)=\{w\in\mathbb{C}\mid 1<|w|<4\}.

For Riemann surfaces SS and RR the sets of conformal embeddings of SS into RR and conformal homeomorphisms of SS onto RR are denoted by CEmb(S,R)\operatorname{CEmb}(S,R) and CHomeo(S,R)\operatorname{CHomeo}(S,R), respectively. By a conformal automorphism of SS we mean a conformal homeomorphism of SS onto itself. Let Aut(S)\operatorname{Aut}(S) be the group of conformal automorphisms of SS. Thus Aut(S)=CHomeo(S,S)\operatorname{Aut}(S)=\operatorname{CHomeo}(S,S). Its subgroup consisting of κAut(S)\kappa\in\operatorname{Aut}(S) satisfying κidS\kappa\simeq\operatorname{id}_{S} is denoted by Aut0(S)\operatorname{Aut}_{0}(S).

Example 2.2.

Conformal automorphisms of \mathbb{C} are of the form κa,b:zaz+b\kappa_{a,b}:z\mapsto az+b, where (a,b)({0})×(a,b)\in(\mathbb{C}\setminus\{0\})\times\mathbb{C}. Since ({0})×(\mathbb{C}\setminus\{0\})\times\mathbb{C} is arcwise connected, we know that Aut0()=Aut()\operatorname{Aut}_{0}(\mathbb{C})=\operatorname{Aut}(\mathbb{C}). The translations κ1,b\kappa_{1,b}, bb\in\mathbb{C}, form a subgroup Auttr()\operatorname{Aut}_{\mathrm{tr}}(\mathbb{C}).

Example 2.3.

For each τ\tau\in\mathbb{H} let Γτ\Gamma_{\tau} denote the additive subgroup of \mathbb{C} generated by 11 and τ\tau. We endow the quotient group Tτ:=/ΓτT_{\tau}:=\mathbb{C}/\Gamma_{\tau} with conformal structure so that the natural projection Πτ:Tτ\Pi_{\tau}:\mathbb{C}\to T_{\tau} is a holomorphic universal covering map. Then TτT_{\tau} is a torus. As is well-known, each torus is conformally equivalent to some TτT_{\tau}. For later use we introduce a few more notations. We denote by AτA_{\tau} and BτB_{\tau} the simple loops on TτT_{\tau} defined by Aτ(t)=Πτ(t)A_{\tau}(t)=\Pi_{\tau}(t), Bτ(t)=Πτ(tτ)B_{\tau}(t)=\Pi_{\tau}(t\tau), t[0,1]t\in[0,1]. For any βTτ\beta\in T_{\tau} the translation ιβ:pp+β\iota_{\beta}:p\mapsto p+\beta is a conformal automorphism of TτT_{\tau} for which the image loops (ιβ)Aτ(\iota_{\beta})_{*}A_{\tau} and (ιβ)Bτ(\iota_{\beta})_{*}B_{\tau} are freely homotopic to AτA_{\tau} and BτB_{\tau}, respectively. If bΠτ1(β)b\in\Pi_{\tau}^{-1}(\beta), then we have Πτκ1,b=ιβΠτ\Pi_{\tau}\circ\kappa_{1,b}=\iota_{\beta}\circ\Pi_{\tau}. The set Auttr(Tτ)\operatorname{Aut}_{\mathrm{tr}}(T_{\tau}) of translations of TτT_{\tau} is a subgroup of Aut0(Tτ)\operatorname{Aut}_{0}(T_{\tau}). In fact, Auttr(Tτ)\operatorname{Aut}_{\mathrm{tr}}(T_{\tau}) is identical with Aut0(Tτ)\operatorname{Aut}_{\mathrm{0}}(T_{\tau}) (see Corollary 3.3 (ii)).

Definition 2.4 (continuation).

Let SS be a Riemann surface. A continuation of SS is, by definition, a pair (R,ι)(R,\iota) where RR is a Riemann surface and ι\iota is a conformal embedding of SS into RR.

Two continuations (Rj,ιj)(R_{j},\iota_{j}), j=1,2j=1,2, of SS are defined to be equivalent to each other if there is κCHomeo(R1,R2)\kappa\in\operatorname{CHomeo}(R_{1},R_{2}) such that ι2=κι1\iota_{2}=\kappa\circ\iota_{1}. A continuation (R,ι)(R,\iota) of SS is called closed (resp. open, compact) if RR is closed (resp. open, compact). If ι(S)\iota(S) is dense in RR, then we say that (R,ι)(R,\iota) is a dense continuation of SS. A continuation (R,ι)(R,\iota) of SS is said to be genus-preserving if each connected component of Rι(S)R\setminus\iota(S) is topologically embedded into \mathbb{C}. In the case where SS is of finite genus, a continuation (R,ι)(R,\iota) of SS is genus-preserving if and only if RR is of the same genus as SS. Observe that if SS is a bordered Riemann surface, then there is a genus-preserving open continuation (R,ι)(R,\iota) of SS such that ι(S)\iota(S) is a retract of RR. If (R,ι)(R,\iota) is a continuation of SS and (R,ι)(R^{\prime},\iota^{\prime}) is a continuation of RR, then (R,ιι)(R^{\prime},\iota^{\prime}\circ\iota) is a continuation of SS.

Remark.

In [48, III.3D] Rodin and Sario call a continuation (R,ι)(R,\iota) of SS compact if RR is closed. We employ the term “closed continuation” to remain consistent. In [62] genus-preserving closed continuations are referred to as closings.

Let ff be a quasiconformal embedding of a Riemann surface SS without border into a Riemann surface. If zz is a local parameter around a point pSp\in S and ww is a local parameter around f(p)f(p), then with respect to these parameters ff is regarded as a complex function w=f(z)w=f(z). The quantity ¯f/f=(fz¯dz¯)/(fzdz)\bar{\partial}f/\partial f=(f_{\bar{z}}\,d\bar{z})/(f_{z}\,dz) does not depend on a particular choice of ww and defines a measurable (1,1)(-1,1)-form μf\mu_{f} on SS called the Beltrami differential of ff. Then |μf||\mu_{f}| is a measurable function on SS, whose LL^{\infty}-norm μf\|\mu_{f}\|_{\infty} is less than one. The maximal dilatation K(f)K(f) of ff is defined by K(f)=(1+μf)/(1μf)K(f)=(1+\|\mu_{f}\|_{\infty})/(1-\|\mu_{f}\|_{\infty}). Denote by QCEmb(S,R)\operatorname{QCEmb}(S,R) and QCHomeo(S,R)\operatorname{QCHomeo}(S,R) the sets of quasiconformal embeddings of SS into RR and quasiconformal homeomorphisms of SS onto RR, respectively.

Let hjQCHomeo(S,Rj)h_{j}\in\operatorname{QCHomeo}(S,R_{j}), j=1,2j=1,2, where SS and RjR_{j} are Riemann surfaces without border. If μh1=μh2\mu_{h_{1}}=\mu_{h_{2}} almost everywhere on SS, then h2h11CHomeo(R1,R2)h_{2}\circ h_{1}^{-1}\in\operatorname{CHomeo}(R_{1},R_{2}). For any measurable (1,1)(-1,1)-form μ\mu on SS with μ<1\|\mu\|_{\infty}<1 there is a quasiconformal homeomorphism hh of SS onto a Riemann surface without border for which μh=μ\mu_{h}=\mu almost everywhere on SS.

It is convenient to define quasiconformal embeddings also on bordered Riemann surfaces. Let SS be a bordered Riemann surface. A topological embedding ff of SS into another Riemann surface RR is called quasiconformal if there are continuations (S,ι)(S^{\prime},\iota^{\prime}) and (R,κ)(R^{\prime},\kappa^{\prime}) of SS and RR, respectively, where SS^{\prime} and RR^{\prime} are Riemann surfaces without border, together with fQCEmb(S,R)f^{\prime}\in\operatorname{QCEmb}(S^{\prime},R^{\prime}) such that κf=fι\kappa^{\prime}\circ f=f^{\prime}\circ\iota^{\prime}. We can then speak of the maximal dilatation K(f)K(f). Note that K(f)K(f)K(f^{\prime})\geqq K(f). Each component of the border of SS is mapped by ff onto a simple arc or loop of vanishing area.

Let KK be a real constant with K1K\geqq 1. A quasiconformal embedding is called KK-quasiconformal if its maximal dilatation does not exceed KK. Conformal embeddings are 11-quasiconformal, and vice versa.

A subset SS^{\prime} of a Riemann surface SS is called a subsurface of SS if it carries its own conformal structure such that the inclusion mapping of SS^{\prime} into SS is a conformal embedding of SS^{\prime} into SS. Each component of the border of SS^{\prime}, if any, is an analytic curve on SS. Subdomains of SS, or nonempty connected open subsets of SS, are subsurfaces of SS.

Let SS be a bordered Riemann surface. We denote its border by S\partial S, and set S=SSS^{\circ}=S\setminus\partial S, called the interior of SS. Note that SS^{\circ} is a subsurface of SS and that S\partial S, which is also a topological boundary of SS^{\circ}, is a 1-dimensional real analytic submanifold of SS. If ι\iota denotes the inclusion mapping of SS^{\circ} into SS, then we call (S,ι)(S,\iota) the canonical continuation of SS^{\circ}. In the case where SS is a Riemann surface without border, we define S=\partial S=\varnothing and S=SS^{\circ}=S for convenience, and call SS^{\circ} the interior of SS.

Definition 2.5 (finite Riemann surface).

A Riemann surface is called finite if its fundamental group is finitely generated.

A Riemann surface SS is said to be analytically finite if there is a closed continuation (R,ι)(R,\iota) of SS such that Rι(S)R\setminus\iota(S) is a finite set, that is, card(Rι(S))<0\operatorname{card}(R\setminus\iota(S))<\aleph_{0}, where cardE\operatorname{card}E denotes the cardinal number of a set EE. A Riemann surface is said to be a nonanalytically finite if it is finite but not analytically finite.

Analytically finite Riemann surfaces are finite Riemann surfaces without border. Closed Riemann surfaces are analytically finite. If SS is an analytically finite Riemann surface and hh is a quasiconformal homeomorphism of SS onto another Riemann surface RR, then RR is also analytically finite.

Definition 2.6 (natural continuation).

Let SS be a finite open Riemann surface. A compact continuation (S˘,ι˘)(\breve{S},\breve{\iota}) of SS is called natural if (S˘)ι˘(S)(\breve{S})^{\circ}\setminus\breve{\iota}(S) contains at most finitely many points.

Clearly, (S˘,ι˘)(\breve{S},\breve{\iota}) is a dense continuation of SS. To obtain a natural compact continuation of SS we have only to consider the case where SS is nonanalytically finite. Then the universal covering Riemann surface of SS is conformally equivalent to the upper half plane :={zImz>0}\mathbb{H}:=\{z\in\mathbb{C}\mid\operatorname{Im}z>0\}. The covering transformation group Γ\Gamma is a torsion-free Fuchsian group of the second kind keeping \mathbb{H} invariant, and /Γ\mathbb{H}/\Gamma is conformally equivalent to SS. Let Ω(Γ)\Omega(\Gamma) denote the region of discontinuity of Γ\Gamma, and set S=(¯Ω(Γ))/ΓS^{\prime}=(\bar{\mathbb{H}}\cap\Omega(\Gamma))/\Gamma, where ¯\bar{\mathbb{H}} is the closure of \mathbb{H} in the extended complex plane ^:={}\hat{\mathbb{C}}:=\mathbb{C}\cup\{\infty\}. If SS^{\prime} is compact, then we set S˘=S\breve{S}=S^{\prime}. Otherwise, SS^{\prime} has finitely many punctures and we let S˘\breve{S} be the Riemann surface SS^{\prime} with the punctures filled in. In any case we have a conformal embedding ι˘\breve{\iota} of SS into S˘\breve{S} to obtain a desired continuation (S˘,ι˘)(\breve{S},\breve{\iota}). As is easily verified, two natural compact continuations of SS are equivalent to each other.

Example 2.7.

Let 𝔻\mathbb{D} and SS be the open unit disk and the open square (1,1)×(1,1)(-1,1)\times(-1,1) in the complex plane \mathbb{C}, respectively. They are subsurfaces of \mathbb{C}. Möbius transformations of the closure 𝔻¯\bar{\mathbb{D}} onto ¯\bar{\mathbb{H}} make 𝔻¯\bar{\mathbb{D}} a compact bordered Riemann surface so that the inclusion mapping ι\iota of 𝔻¯\bar{\mathbb{D}} into \mathbb{C} is a conformal embedding. In other words, (,ι)(\mathbb{C},\iota) is a continuation of 𝔻¯\bar{\mathbb{D}}. Thus 𝔻¯\bar{\mathbb{D}} is a subsurface of \mathbb{C}. Take ι˘CEmb(S,𝔻¯)\breve{\iota}\in\operatorname{CEmb}(S,\bar{\mathbb{D}}) such that ι˘(S)=𝔻\breve{\iota}(S)=\mathbb{D}. Then (𝔻¯,ι˘)(\bar{\mathbb{D}},\breve{\iota}) is a natural compact continuation of SS. The embedding ι˘\breve{\iota} is extended to a homeomorphism of the topological closure S¯=[1,1]×[1,1]\bar{S}=[-1,1]\times[-1,1] onto 𝔻¯\bar{\mathbb{D}}. Its inverse defines a topological embedding hh of 𝔻¯\bar{\mathbb{D}} into \mathbb{C}, but (,h)(\mathbb{C},h) is not a continuation of 𝔻¯\bar{\mathbb{D}}, for, hh is not analytic at four points on the border 𝔻\partial\mathbb{D}. Obviously, S¯{±1±i}\bar{S}\setminus\{\pm 1\pm i\} is a subsurface of \mathbb{C} though S¯\bar{S} is not. Also, if S˙=S{0}\dot{S}=S\setminus\{0\}, then (𝔻¯,ι˘|S˙)(\bar{\mathbb{D}},\breve{\iota}|_{\dot{S}}) is a natural compact continuation of S˙\dot{S}.

3 Categorical definition of Teichmüller spaces

Let R0R_{0} be a Riemann surface of finite genus. Then it is conformally embedded into a closed Riemann surface of the same genus by Bochner [10, Satz V], which gives rise to a genus-preserving closed continuation of R0R_{0}. We are interested in the set of closed Riemann surfaces of the same genus as R0R_{0} into which R0R_{0} can be conformally embedded, and will work in the context of Teichmüller theory.

There are several equivalent definitions of Teichmüller spaces for closed Riemann surfaces. In the present article we adopt the following definition. The point is how we should define continuous mappings between marked Riemann surfaces.

We start with introducing the category 𝔉g\mathfrak{F}_{g} of marked Riemann surfaces of genus gg and conformally compatible continuous mappings. Fix a closed oriented surface Σg\Sigma_{g} of positive genus gg, and remove one point from Σg\Sigma_{g} to obtain a once-punctured closed oriented surface Σ˙g\dot{\Sigma}_{g}. For the sake of definiteness, using the notations in Example 2.3, we set Σ1=Ti=T1\Sigma_{1}=T_{i}=T_{\sqrt{-1}} and Σ˙1=Σ1{Πi((1+i)/2)}\dot{\Sigma}_{1}=\Sigma_{1}\setminus\{\Pi_{i}((1+i)/2)\}.

Definition 3.1 (handle mark).

Let gg be a positive integer, and let SS be an oriented surface of genus gg or higher, possibly, with border. By a gg-handle mark of SS we mean an element of TEmb+(Σ˙g,S)\operatorname{TEmb}^{+}(\dot{\Sigma}_{g},S).

If η\eta is a gg-handle mark of SS, then η(Σ˙g)S\eta(\dot{\Sigma}_{g})\subset S^{\circ}. Thus η\eta is also considered as a gg-handle mark of SS^{\circ}.

Proposition 3.2.

Let χ\chi be a 11-handle mark of a Riemann surface SS of positive genus, and assume that κAut(S)\kappa\in\operatorname{Aut}(S) satisfies κχχ\kappa\circ\chi\simeq\chi.

  • (i)

    If SS is not a torus, then κ=idS\kappa=\operatorname{id}_{S}.

  • (ii)

    If SS is a torus, then κAuttr(S)\kappa\in\operatorname{Aut}_{\mathrm{tr}}(S).

Proof.

(i) The interior SS^{\circ} carries a hyperbolic metric, that is, a complete conformal metric of constant curvature 4-4. Let AA and BB be the hyperbolic geodesic loops on SS^{\circ} that are freely homotopic to χA1\chi_{*}A_{\sqrt{-1}} and χB1\chi_{*}B_{\sqrt{-1}}, respectively (for the definition of A1A_{\sqrt{-1}} and B1B_{\sqrt{-1}} see Example 2.3). Then κAA\kappa_{*}A\simeq A and κBB\kappa_{*}B\simeq B due to κχχ\kappa\circ\chi\simeq\chi. Since κ\kappa is isometric with respect to the hyperbolic metric, κA\kappa_{*}A and κB\kappa_{*}B are also hyperbolic geodesic loops and hence trace AA and BB in the same direction, respectively. Because AA and BB have exactly one point, say pp, in common, we know that κ(p)=p\kappa(p)=p, which implies κ\kappa coincides with idS\operatorname{id}_{S} on AA and BB. Therefore κ=idS\kappa=\operatorname{id}_{S} by the identity theorem.

(ii) We may assume that SS is identical with some TτT_{\tau} in Example 2.3. The conformal automorphism κ\kappa of TτT_{\tau} is lifted to a conformal automorphism κ~\tilde{\kappa} of \mathbb{C}. The 1-form ω~:=dz\tilde{\omega}:=dz on \mathbb{C} is projected to a holomorphic 1-form ω\omega on TτT_{\tau}. Since κAτ\kappa_{*}A_{\tau} is freely homotopic to AτA_{\tau}, the pull-back κω\kappa^{*}\omega of ω\omega via κ\kappa has the same period along AτA_{\tau} as ω\omega:

Aτκω=κAτω=Aτω.\int_{A_{\tau}}\kappa^{*}\omega=\int_{\kappa_{*}A_{\tau}}\omega=\int_{A_{\tau}}\omega.

As TτT_{\tau} is a torus, we obtain κω=ω\kappa^{*}\omega=\omega and hence κ~ω~=ω~\tilde{\kappa}^{*}\tilde{\omega}=\tilde{\omega}, or equivalently, κ~(z)=z+C\tilde{\kappa}(z)=z+C, zz\in\mathbb{C}, for some constant CC. Consequently, κ\kappa belongs to Auttr(Tτ)\operatorname{Aut}_{\mathrm{tr}}(T_{\tau}). ∎

Corollary 3.3.

Let SS be a Riemann surface of positive genus.

  • (i)

    If SS is not a torus, then Aut0(S)\operatorname{Aut}_{0}(S) is trivial.

  • (ii)

    If SS is a torus, then Aut0(S)=Auttr(S)\operatorname{Aut}_{0}(S)=\operatorname{Aut}_{\mathrm{tr}}(S).

Proof.

Choose a 11-handle mark χ\chi of SS. If κAut0(S)\kappa\in\operatorname{Aut}_{0}(S), then κχχ\kappa\circ\chi\simeq\chi. Thus the corollary follows at once from Proposition 3.2. ∎

Corollary 3.4.

Let χ1\chi_{1} be a 11-handle mark of a Riemann surface S1S_{1} of positive genus, and suppose that κ0,κ1CHomeo(S1,S2)\kappa_{0},\kappa_{1}\in\operatorname{CHomeo}(S_{1},S_{2}) satisfy κ0χ1κ1χ1\kappa_{0}\circ\chi_{1}\simeq\kappa_{1}\circ\chi_{1}.

  • (i)

    If S1S_{1} is not a torus, then κ0=κ1\kappa_{0}=\kappa_{1}.

  • (ii)

    If S1S_{1} is a torus, then κ0κ1\kappa_{0}\simeq\kappa_{1}.

Definition 3.5 (marked Riemann surface).

Consider all pairs (S,η)(S,\eta), where SS is a Riemann surface SS of genus gg and η\eta is a gg-handle mark of SS. We say that (S1,η1)(S_{1},\eta_{1}) is equivalent to (S2,η2)(S_{2},\eta_{2}) if there is κCHomeo(S1,S2)\kappa\in\operatorname{CHomeo}(S_{1},S_{2}) such that κη1η2\kappa\circ\eta_{1}\simeq\eta_{2}. We call each equivalence class 𝑺=[S,η]{\bm{S}}=[S,\eta] a marked Riemann surface of genus gg.

A marked Riemann surface 𝑺=[S,η]{\bm{S}}=[S,\eta] is called a marked bordered Riemann surface or a marked Riemann surface with border if SS is a bordered Riemann surface. Otherwise, 𝑺\bm{S} is referred to as a marked Riemann surface without border. If SS is closed (resp. open, compact), then 𝑺\bm{S} is called closed (resp. open, compact). We also say that 𝑺\bm{S} is a marked closed (resp. open, compact) Riemann surface. The meaning of a marked torus is obvious. If SS is finite (resp. analytically finite, nonanalytically finite), then 𝑺\bm{S} is said to be finite (resp. analytically finite, nonanalytically finite). Since η(Σ˙g)S\eta(\dot{\Sigma}_{g})\subset S^{\circ}, regarding η\eta as a topological embedding of Σ˙g\dot{\Sigma}_{g} into SS^{\circ}, we obtain a marked Riemann surface 𝑺=[S,η]{\bm{S}}^{\circ}=[S^{\circ},\eta] without border, which will be called the interior of 𝑺\bm{S}.

Remark.

Let (S,η),(S,η)𝑺(S,\eta),(S^{\prime},\eta^{\prime})\in{\bm{S}}. Suppose that κ0,κ1CHomeo(S,S)\kappa_{0},\kappa_{1}\in\operatorname{CHomeo}(S,S^{\prime}) satisfy κkηη\kappa_{k}\circ\eta\simeq\eta^{\prime} for k=0,1k=0,1. Choose a 11-handle mark χ\chi of Σ˙g\dot{\Sigma}_{g} to obtain a 11-handle mark ηχ\eta\circ\chi of SS. Since κ0(ηχ)κ1(ηχ)\kappa_{0}\circ(\eta\circ\chi)\simeq\kappa_{1}\circ(\eta\circ\chi), it follows from Corollary 3.4 that κ0κ1\kappa_{0}\simeq\kappa_{1} if 𝑺\bm{S} is a marked torus and that κ0=κ1\kappa_{0}=\kappa_{1} otherwise.

Example 3.6.

We have obtained a torus TτT_{\tau} for each τ\tau\in\mathbb{H} in Example 2.3. The homeomorphism z{(τ+i)z(τi)z¯}/2iz\mapsto\{(\tau+i)z-(\tau-i)\bar{z}\}/2i of \mathbb{C} onto itself induces ητHomeo+(Σ1,Tτ)\eta_{\tau}\in\operatorname{Homeo}^{+}(\Sigma_{1},T_{\tau}) and a marked torus 𝑻τ:=[Tτ,η˙τ]{\bm{T}}_{\tau}:=[T_{\tau},\dot{\eta}_{\tau}], where η˙τ=ητ|Σ˙1\dot{\eta}_{\tau}=\eta_{\tau}|_{\dot{\Sigma}_{1}}.

We show two propositions to verify that the marked closed Riemann surfaces defined above are essentially the same as the known ones. The second proposition also plays a fundamental role in our investigations on closed continuations.

Proposition 3.7.

Let DD be a subdomain of 𝔻\mathbb{D} such that 𝔻D\mathbb{D}\setminus D is compact, and set D~=D𝔻\tilde{D}=D\cup\partial\mathbb{D}. For each fTEmb(D~,𝔻¯)f\in\operatorname{TEmb}(\tilde{D},\bar{\mathbb{D}}) with f(𝔻)=𝔻f(\partial\mathbb{D})=\partial\mathbb{D} there are hHomeo(𝔻¯,𝔻¯)h\in\operatorname{Homeo}(\bar{\mathbb{D}},\bar{\mathbb{D}}) and HCont(D~×[0,1],𝔻¯)H\in\operatorname{Cont}(\tilde{D}\times[0,1],\bar{\mathbb{D}}) such that H(z,0)=h(z)H(z,0)=h(z), H(z,1)=f(z)H(z,1)=f(z) and H(ζ,t)=f(ζ)H(\zeta,t)=f(\zeta) for zD~z\in\tilde{D}, ζ𝔻\zeta\in\partial\mathbb{D} and t[0,1]t\in[0,1].

Proof.

For r[0,1]r\in[0,1] and ζ𝔻\zeta\in\partial\mathbb{D} set h(rζ)=rf(ζ)h(r\zeta)=rf(\zeta). Then hh defines a homeomorphism of 𝔻¯\bar{\mathbb{D}} onto itself with h=fh=f on 𝔻\partial\mathbb{D}. The correspondence

H:D~×[0,1](z,t)(1t)h(z)+tf(z)𝔻¯H:\tilde{D}\times[0,1]\ni(z,t)\mapsto(1-t)h(z)+tf(z)\in\bar{\mathbb{D}}

possesses the desired properties. ∎

Proposition 3.8.

Let S1S_{1} and S2S_{2} be oriented surfaces of genus gg possibly with border, and let f0f_{0} and f1f_{1} be topological embeddings of S1S_{1} into S2S_{2}. If S2S_{2} is closed and f0η1f1η1f_{0}\circ\eta_{1}\simeq f_{1}\circ\eta_{1} for a gg-handle mark η1\eta_{1} of S1S_{1}, then f0f1f_{0}\simeq f_{1}.

Proof.

Take hTEmb+(𝔻¯,Σg)h\in\operatorname{TEmb}^{+}(\bar{\mathbb{D}},\Sigma_{g}) for which ΣgΣ˙gh(𝔻)\Sigma_{g}\setminus\dot{\Sigma}_{g}\subset h(\mathbb{D}). The simple loop γ1:=η1h(𝔻)\gamma_{1}:=\eta_{1}\circ h(\partial\mathbb{D}) divides S1S_{1} into two surfaces, one of which, say S1S^{\prime}_{1}, is homeomorphic to Σ˙g\dot{\Sigma}_{g} while the other, say D1D_{1}, is planar, that is, of genus zero. Since f0η1f1η1f_{0}\circ\eta_{1}\simeq f_{1}\circ\eta_{1}, there is FCont(S1×I,S2)F^{\prime}\in\operatorname{Cont}(S^{\prime}_{1}\times I,S_{2}), where I=[0,1]I=[0,1], such that F(,t)=ft|S1F^{\prime}(\,\cdot\,,t)=f_{t}|_{S^{\prime}_{1}} for t=0,1t=0,1.

Introducing a conformal structure into each SjS_{j}, let Πj:𝕌SjSj\Pi_{j}:\mathbb{U}_{S_{j}}\to S_{j} be a holomorphic universal covering map with covering transformation group Γj:=Aut(Πj)\Gamma_{j}:=\operatorname{Aut}(\Pi_{j}), where 𝕌S2=\mathbb{U}_{S_{2}}=\mathbb{H} or \mathbb{C}. Fix a point p0Σgh(𝔻¯)p_{0}\in\Sigma_{g}\setminus h(\bar{\mathbb{D}}), and choose z0Π11(η1(p0))z_{0}\in\Pi_{1}^{-1}(\eta_{1}(p_{0})). Let S~1\tilde{S}^{\prime}_{1} denote the component of Π11(S1)\Pi_{1}^{-1}(S^{\prime}_{1}) containing z0z_{0}, and let Γ1\Gamma^{\prime}_{1} be the subgroup of Γ1\Gamma_{1} consisting of covering transformations leaving S~1\tilde{S}^{\prime}_{1} invariant. The restriction Π1:=Π1|S~1\Pi^{\prime}_{1}:=\Pi_{1}|_{\tilde{S}^{\prime}_{1}} of Π1\Pi_{1} to S~1\tilde{S}^{\prime}_{1} is a holomorphic covering map of S1S^{\prime}_{1}, and Γ1\Gamma^{\prime}_{1} acts on S~1\tilde{S}^{\prime}_{1} as the covering transformation group Aut(Π1)\operatorname{Aut}(\Pi^{\prime}_{1}). The embedding η1\eta_{1} induces an injective group homomorphism (η1)(\eta_{1})_{*} of the fundamental group π1(Σ˙g,p0)\pi_{1}(\dot{\Sigma}_{g},p_{0}) of Σ˙g\dot{\Sigma}_{g} with base point p0p_{0} into π1(S1,η1(p0))\pi_{1}(S_{1},\eta_{1}(p_{0})). The latter fundamental group is canonically isomorphic to Γ1\Gamma_{1} with Γ1\Gamma^{\prime}_{1} corresponding to (η1)(π1(Σ˙g,p0))(\eta_{1})_{*}(\pi_{1}(\dot{\Sigma}_{g},p_{0})). Let Δ1\Delta_{1} be the subset of Γ1\Gamma_{1} assigned to the elements of π1(S1,η1(p0))\pi_{1}(S_{1},\eta_{1}(p_{0})) represented by loops freely homotopic to loops in D1D_{1}.

Any loop in S~1\tilde{S}^{\prime}_{1} with initial point z0z_{0} is trivial in 𝕌S1\mathbb{U}_{S_{1}} and hence is projected to a loop trivial in S1S_{1}. Thus the induced group homomorphism (F(Π1×idI))(F^{\prime}\circ(\Pi^{\prime}_{1}\times\operatorname{id}_{I}))_{*} maps the fundamental group π1(S~1×I,(z0,0))\pi_{1}(\tilde{S}^{\prime}_{1}\times I,(z_{0},0)) to the trivial subgroup of π1(S2,f0η1(p0))\pi_{1}(S_{2},f_{0}\circ\eta_{1}(p_{0})). Therefore, there is F~Cont(S~1×I,𝕌S2)\tilde{F}^{\prime}\in\operatorname{Cont}(\tilde{S}^{\prime}_{1}\times I,\mathbb{U}_{S_{2}}) such that Π2F~=F(Π1×idI)\Pi_{2}\circ\tilde{F}^{\prime}=F^{\prime}\circ(\Pi^{\prime}_{1}\times\operatorname{id}_{I}). Set f~t=F~(,t)\tilde{f}^{\prime}_{t}=\tilde{F}^{\prime}(\,\cdot\,,t) for tIt\in I. Since Γ2\Gamma_{2} is discrete, there is a group homomorphism ρ:Γ1Γ2\rho^{\prime}:\Gamma^{\prime}_{1}\to\Gamma_{2} such that f~tγ=ρ(γ)f~t\tilde{f}^{\prime}_{t}\circ\gamma=\rho^{\prime}(\gamma)\circ\tilde{f}^{\prime}_{t} for tIt\in I. Note that ρ\rho^{\prime} does not depend on tt.

For k=0,1k=0,1 there is f~kCont(𝕌S1,𝕌S2)\tilde{f}_{k}\in\operatorname{Cont}(\mathbb{U}_{S_{1}},\mathbb{U}_{S_{2}}) together with a group homomorphism ρk:Γ1Γ2\rho_{k}:\Gamma_{1}\to\Gamma_{2} such that fkΠ1=Π2f~kf_{k}\circ\Pi_{1}=\Pi_{2}\circ\tilde{f}_{k} and f~kγ=ρk(γ)f~k\tilde{f}_{k}\circ\gamma=\rho_{k}(\gamma)\circ\tilde{f}_{k} for γΓ1\gamma\in\Gamma_{1}. Since fkf_{k} is identical with F(,k)F^{\prime}(\,\cdot\,,k) on S1S^{\prime}_{1}, we can choose f~k\tilde{f}_{k} so that f~k(z0)=f~k(z0)\tilde{f}_{k}(z_{0})=\tilde{f}^{\prime}_{k}(z_{0}). Then ρk\rho_{k} coincides with ρ\rho^{\prime} on Γ1\Gamma^{\prime}_{1}. Observe that Δ1Kerρk\Delta_{1}\subset\operatorname{Ker}\rho_{k} as S2S_{2} is closed. Since Γ1\Gamma_{1} is generated by Γ1\Gamma^{\prime}_{1} and Δ1\Delta_{1}, we infer that ρ0=ρ1\rho_{0}=\rho_{1} and hence f0f1f_{0}\simeq f_{1} (see, for example, Ahlfors [2, Lemma on p.119] or Bourque [12, Lemma 2.9]). This completes the proof. ∎

In Teichmüller theory closed Riemann surfaces of genus gg have been marked with sense-preserving homeomorphisms of Σg\Sigma_{g}, not Σ˙g\dot{\Sigma}_{g}. Our definition leads us to essentially the same space of marked closed Riemann surfaces. In fact, if η0\eta_{0} is a gg-handle mark of a closed Riemann surface SS of genus gg, then it follows from Proposition 3.7 that there is ηHomeo+(Σg,S)\eta\in\operatorname{Homeo}^{+}(\Sigma_{g},S) such that (S,η˙)(S,\dot{\eta}) is equivalent to (S,η0)(S,\eta_{0}), where η˙=η|Σ˙g\dot{\eta}=\eta|_{\dot{\Sigma}_{g}}. Moreover, let η˙\dot{\eta}^{\prime} be a gg-handle mark of a closed Riemann surface SS^{\prime} of genus gg, and assume that it is extended to a homeomorphism η\eta^{\prime} of Σg\Sigma_{g} onto SS^{\prime}. If (S,η˙)(S^{\prime},\dot{\eta}^{\prime}) is equivalent to (S,η˙)(S,\dot{\eta}), then κη˙η˙\kappa\circ\dot{\eta}\simeq\dot{\eta}^{\prime} for some κCHomeo(S,S)\kappa\in\operatorname{CHomeo}(S,S^{\prime}). The inclusion mapping ι:Σ˙gΣg\iota:\dot{\Sigma}_{g}\to\Sigma_{g} is a gg-handle mark of Σg\Sigma_{g}. Since (κη)ιηι(\kappa\circ\eta)\circ\iota\simeq\eta^{\prime}\circ\iota, Proposition 3.8 implies that κηη\kappa\circ\eta\simeq\eta^{\prime}.

Definition 3.9 (continuous mapping).

Let 𝑺j{\bm{S}}_{j}, j=1,2j=1,2, be marked Riemann surfaces of genus gg, and consider all triples (f,η1,η2)(f,\eta_{1},\eta_{2}), where (Sj,ηj)𝑺j(S_{j},\eta_{j})\in{\bm{S}}_{j} and fCont(S1,S2)f\in\operatorname{Cont}(S_{1},S_{2}). Two such triples (f,η1,η2)(f,\eta_{1},\eta_{2}) and (f,η1,η2)(f^{\prime},\eta^{\prime}_{1},\eta^{\prime}_{2}), where (Sj,ηj)𝑺j(S^{\prime}_{j},\eta^{\prime}_{j})\in{\bm{S}}_{j} and fCont(S1,S2)f^{\prime}\in\operatorname{Cont}(S^{\prime}_{1},S^{\prime}_{2}), are said to be equivalent to each other if fκ1=κ2ff^{\prime}\circ\kappa_{1}=\kappa_{2}\circ f for some κjCHomeo(Sj,Sj)\kappa_{j}\in\operatorname{CHomeo}(S_{j},S^{\prime}_{j}) with κjηjηj\kappa_{j}\circ\eta_{j}\simeq\eta^{\prime}_{j}. We call each equivalence class 𝒇=[f,η1,η2]{\bm{f}}=[f,\eta_{1},\eta_{2}] a continuous mapping of 𝑺1{\bm{S}}_{1} into 𝑺2{\bm{S}}_{2} and use the notation 𝒇:𝑺1𝑺2{\bm{f}}:{\bm{S}}_{1}\to{\bm{S}}_{2}.

If ff has some conformally invariant properties in addition, then 𝒇\bm{f} is said to possess the same properties. For example, if ff is a quasiconformal embedding of S1S_{1} into S2S_{2}, then 𝒇\bm{f} is called a quasiconformal embedding of 𝑺1{\bm{S}}_{1} into 𝑺2{\bm{S}}_{2}.

To compose continuous mappings of a marked Riemann surface into another we need to place restrictions on the mappings. In practice they target continuous mappings of marked tori.

Definition 3.10 (conformally compatible continuous mapping).

Let S1S_{1} and S2S_{2} be Riemann surfaces of positive genus. Then fCont(S1,S2)f\in\operatorname{Cont}(S_{1},S_{2}) is said to be conformally compatible if for any κ1Aut0(S1)\kappa_{1}\in\operatorname{Aut}_{0}(S_{1}) there is κ2Aut0(S2)\kappa_{2}\in\operatorname{Aut}_{0}(S_{2}) such that fκ1=κ2ff\circ\kappa_{1}=\kappa_{2}\circ f.

If S1S_{1} is not a torus, then every continuous mapping of S1S_{1} into S2S_{2} is conformally compatible as Aut0(S1)={idS1}\operatorname{Aut}_{0}(S_{1})=\{\operatorname{id}_{S_{1}}\}. If S1S_{1} is a torus while S2S_{2} is not, then every conformally compatible continuous mapping of S1S_{1} into S2S_{2} is constant because Aut0(S1)=Auttr(S1)\operatorname{Aut}_{0}(S_{1})=\operatorname{Aut}_{\mathrm{tr}}(S_{1}) acts transitively on S1S_{1} while Aut0(S2)\operatorname{Aut}_{0}(S_{2}) is trivial. To determine conformally compatible continuous mappings of S1S_{1} into S2S_{2} in the case where both S1S_{1} and S2S_{2} are tori, take holomorphic universal covering mappings Πj:Sj\Pi_{j}:\mathbb{C}\to S_{j}, j=1,2j=1,2. Suppose that fCont(S1,S2)f\in\operatorname{Cont}(S_{1},S_{2}) is conformally compatible. Let f~Cont(,)\tilde{f}\in\operatorname{Cont}(\mathbb{C},\mathbb{C}) be a lift of ff. Thus we have Π2f~=fΠ1\Pi_{2}\circ\tilde{f}=f\circ\Pi_{1}. For any τ1Auttr()\tau_{1}\in\operatorname{Aut}_{\mathrm{tr}}(\mathbb{C}) there is κ1Auttr(S1)=Aut0(S1)\kappa_{1}\in\operatorname{Aut}_{\mathrm{tr}}(S_{1})=\operatorname{Aut}_{0}(S_{1}) such that κ1Π1=Π1τ1\kappa_{1}\circ\Pi_{1}=\Pi_{1}\circ\tau_{1}. Since ff is conformally compatible, we find κ2Aut0(S2)=Auttr(S2)\kappa_{2}\in\operatorname{Aut}_{0}(S_{2})=\operatorname{Aut}_{\mathrm{tr}}(S_{2}) for which fκ1=κ2ff\circ\kappa_{1}=\kappa_{2}\circ f, which leads us to

Π2f~τ1=fΠ1τ1=fκ1Π1=κ2fΠ1=κ2Π2f~=Π2τ2f~\Pi_{2}\circ\tilde{f}\circ\tau_{1}=f\circ\Pi_{1}\circ\tau_{1}=f\circ\kappa_{1}\circ\Pi_{1}=\kappa_{2}\circ f\circ\Pi_{1}=\kappa_{2}\circ\Pi_{2}\circ\tilde{f}=\Pi_{2}\circ\tau_{2}\circ\tilde{f}

for some τ2Auttr()\tau_{2}\in\operatorname{Aut}_{\mathrm{tr}}(\mathbb{C}). Therefore, we have f~(z+a)=f~(z)+l(a)\tilde{f}(z+a)=\tilde{f}(z)+l(a) for z,az,a\in\mathbb{C}, where l(a)=f~(a)f~(0)l(a)=\tilde{f}(a)-\tilde{f}(0). Since l(a1+a2)=l(a1)+l(a2)l(a_{1}+a_{2})=l(a_{1})+l(a_{2}), it follows that l(ca)=cl(a)l(ca)=cl(a) for cc\in\mathbb{Q} and hence for cc\in\mathbb{R} by continuity. Consequently, ll is an \mathbb{R}-linear mapping of \mathbb{C} into itself, and f~\tilde{f} is an \mathbb{R}-affine transformation. For any γ\gamma in the covering transformation group Aut(Π1)\operatorname{Aut}(\Pi_{1}) of Π1\Pi_{1} there is ρ(γ)Aut(Π2)\rho(\gamma)\in\operatorname{Aut}(\Pi_{2}) such that f~γ=ρ(γ)f~\tilde{f}\circ\gamma=\rho(\gamma)\circ\tilde{f}. Conversely, if f~:\tilde{f}:\mathbb{C}\to\mathbb{C} is an \mathbb{R}-affine mapping possessing the last property, then it induces a conformally compatible element fCont(S1,S2)f\in\operatorname{Cont}(S_{1},S_{2}) such that Π2f~=fΠ1\Pi_{2}\circ\tilde{f}=f\circ\Pi_{1}. In particular, the inverse of a conformally compatible homeomorphism of a torus onto another is conformally compatible. Note that whether S1S_{1} is a torus or not, conformal embeddings are conformally compatible.

Let 𝑺j{\bm{S}}_{j}, j=1,2j=1,2, be marked Riemann surfaces of genus gg, and let 𝒇\bm{f} be a continuous mapping of 𝑺1{\bm{S}}_{1} into 𝑺2{\bm{S}}_{2}. Take representatives (f,η1,η2),(f,η1,η2)𝒇(f,\eta_{1},\eta_{2}),(f^{\prime},\eta^{\prime}_{1},\eta^{\prime}_{2})\in{\bm{f}}, where (Sj,ηj),(Sj,ηj)𝑺j(S_{j},\eta_{j}),(S^{\prime}_{j},\eta^{\prime}_{j})\in{\bm{S}}_{j}. If ff is conformally compatible, then so is ff^{\prime}. Hence we can speak of a conformally compatible continuous mapping of a marked Riemann surface into another.

Definition 3.11 (composition).

Let 𝒇1:𝑺1𝑺2{\bm{f}}_{1}:{\bm{S}}_{1}\to{\bm{S}}_{2} and 𝒇2:𝑺2𝑺3{\bm{f}}_{2}:{\bm{S}}_{2}\to{\bm{S}}_{3} be conformally compatible continuous mappings. Take (f1,η1,η2)𝒇1(f_{1},\eta_{1},\eta_{2})\in{\bm{f}}_{1} and (f2,η2,η3)𝒇2(f_{2},\eta^{\prime}_{2},\eta_{3})\in{\bm{f}}_{2}, where (Sj,ηj)𝑺j(S_{j},\eta_{j})\in{\bm{S}}_{j}, j=1,2,3j=1,2,3, and (S2,η2)𝑺2(S^{\prime}_{2},\eta^{\prime}_{2})\in{\bm{S}}_{2}. Define a continuous mapping 𝒇2𝒇1{\bm{f}}_{2}\circ{\bm{f}}_{1} of 𝑺1{\bm{S}}_{1} into 𝑺3{\bm{S}}_{3} to be the equivalence class [f2κ22f1,η1,η3][f_{2}\circ\kappa_{22^{\prime}}\circ f_{1},\eta_{1},\eta_{3}], where κ22CHomeo(S2,S2)\kappa_{22^{\prime}}\in\operatorname{CHomeo}(S_{2},S^{\prime}_{2}) with κ22η2η2\kappa_{22^{\prime}}\circ\eta_{2}\simeq\eta^{\prime}_{2}.

It is easy to verify that 𝒇2𝒇1{\bm{f}}_{2}\circ{\bm{f}}_{1} is well-defined, that is, that (f2κ22f1,η1,η3)(f_{2}\circ\kappa_{22^{\prime}}\circ f_{1},\eta_{1},\eta_{3}) represents a unique continuous mapping of 𝑺1{\bm{S}}_{1} into 𝑺3{\bm{S}}_{3}. Actually, 𝒇2𝒇1{\bm{f}}_{2}\circ{\bm{f}}_{1} is well-defined if 𝒇2{\bm{f}}_{2} is conformally compatible. We do not need to require 𝒇1{\bm{f}}_{1} to be conformally compatible.

Compositions of conformally compatible continuous mappings of marked Riemann surfaces are again conformally compatible. Also, the binary operation \circ is associative.

In general, for a marked Riemann surface 𝑺\bm{S} of genus gg let 𝟏𝑺{\bm{1}}_{\bm{S}} denote the conformal automorphism of 𝑺\bm{S} represented by (idS,η,η)(\operatorname{id}_{S},\eta,\eta), where (S,η)𝑺(S,\eta)\in{\bm{S}}. Then 𝟏𝑺2𝒇=𝒇𝟏𝑺1=𝒇\mathbf{1}_{{\bm{S}}_{2}}\circ{\bm{f}}={\bm{f}}\circ\mathbf{1}_{{\bm{S}}_{1}}={\bm{f}} for any conformally compatible continuous mappings 𝒇\bm{f} of 𝑺1{\bm{S}}_{1} into 𝑺2{\bm{S}}_{2}.

We are now ready to define the category 𝔉g\mathfrak{F}_{g}. Its class ob(𝔉g)\operatorname{ob}(\mathfrak{F}_{g}) of objects consists of all marked Riemann surfaces of genus gg. The hom-set Cont(𝑺,𝑹)\operatorname{Cont}({\bm{S}},{\bm{R}}) of morphisms from 𝑺\bm{S} to 𝑹\bm{R} is composed of all conformally compatible continuous mappings of 𝑺\bm{S} into 𝑹\bm{R}. The Teichmüller space 𝔗g\mathfrak{T}_{g} of genus gg is, by definition, the full subcategory of 𝔉g\mathfrak{F}_{g} whose class of objects is exactly the set of marked closed Riemann surfaces of genus gg. For the sake of simplicity we abbreviate ob(𝔉g)\operatorname{ob}(\mathfrak{F}_{g}) and ob(𝔗g)\operatorname{ob}(\mathfrak{T}_{g}) to 𝔉g\mathfrak{F}_{g} and 𝔗g\mathfrak{T}_{g}, respectively.

Remark.

The set Cont(𝑺,𝑹)\operatorname{Cont}({\bm{S}},{\bm{R}}) is not the set of continuous mappings of 𝑺\bm{S} into 𝑹\bm{R}. Every element of Cont(𝑺,𝑹)\operatorname{Cont}({\bm{S}},{\bm{R}}) should be conformally compatible.

Denote by TEmb(𝑺,𝑹)\operatorname{TEmb}({\bm{S}},{\bm{R}}) (resp. QCEmb(𝑺,𝑹)\operatorname{QCEmb}({\bm{S}},{\bm{R}}), CEmb(𝑺,𝑹)\operatorname{CEmb}({\bm{S}},{\bm{R}})) the set of conformally compatible topological (resp. quasiconformal, conformal) embeddings of 𝑺\bm{S} into 𝑹\bm{R}. The set of sense-preserving elements in TEmb(𝑺,𝑹)\operatorname{TEmb}({\bm{S}},{\bm{R}}) is denoted by TEmb+(𝑺,𝑹)\operatorname{TEmb}^{+}({\bm{S}},{\bm{R}}). Also, let Homeo(𝑺,𝑹)\operatorname{Homeo}({\bm{S}},{\bm{R}}) denote the set of surjective elements in TEmb(𝑺,𝑹)\operatorname{TEmb}({\bm{S}},{\bm{R}}), and set

Homeo+(𝑺,𝑹)=Homeo(𝑺,𝑹)TEmb+(𝑺,𝑹).\operatorname{Homeo}^{+}({\bm{S}},{\bm{R}})=\operatorname{Homeo}({\bm{S}},{\bm{R}})\cap\operatorname{TEmb}^{+}({\bm{S}},{\bm{R}}).

Replacing “Emb” with “Homeo” in the notations for the other classes above, we obtain notations for the subset of homeomorphisms of 𝑺\bm{S} onto 𝑹\bm{R} in the classes under consideration. Thus QCHomeo(𝑺,𝑹)\operatorname{QCHomeo}({\bm{S}},{\bm{R}}) means the set of conformally compatible quasiconformal homeomorphisms of 𝑺\bm{S} onto 𝑹\bm{R}.

Definition 3.12 (homotopically consistent embedding).

A sense-preserving topological embedding 𝒇:𝑺1𝑺2{\bm{f}}:{\bm{S}}_{1}\to{\bm{S}}_{2} is called homotopically consistent if fη1η2f\circ\eta_{1}\simeq\eta_{2}, where (Sj,ηj)𝑺j(S_{j},\eta_{j})\in{\bm{S}}_{j}, j=1,2j=1,2, and (f,η1,η2)𝒇(f,\eta_{1},\eta_{2})\in{\bm{f}}.

This definition does not depend on a particular choice of representatives. If 𝒇1TEmb+(𝑺1,𝑺2){\bm{f}}_{1}\in\operatorname{TEmb}^{+}({\bm{S}}_{1},{\bm{S}}_{2}) and 𝒇2TEmb+(𝑺2,𝑺3){\bm{f}}_{2}\in\operatorname{TEmb}^{+}({\bm{S}}_{2},{\bm{S}}_{3}) are homotopically consistent, then so is 𝒇2𝒇1{\bm{f}}_{2}\circ{\bm{f}}_{1}. If X(𝑺,𝑹)\mathrm{X}({\bm{S}},{\bm{R}}) is a subset of TEmb+(𝑺,𝑹)\operatorname{TEmb}^{+}({\bm{S}},{\bm{R}}), then we denote by Xhc(𝑺,𝑹)\mathrm{X}_{\mathrm{hc}}({\bm{S}},{\bm{R}}) the subset of homotopically consistent elements in X(𝑺,𝑹)\mathrm{X}({\bm{S}},{\bm{R}}). For example, QCEmbhc(𝑺,𝑹)\operatorname{QCEmb}_{\mathrm{hc}}({\bm{S}},{\bm{R}}) means the set of homotopically consistent and conformally compatible quasiconformal embeddings of 𝑺\bm{S} into 𝑹\bm{R}.

Let 𝒇0=[f0,η10,η20]{\bm{f}}_{0}=[f_{0},\eta_{10},\eta_{20}] and 𝒇1=[f1,η11,η21]{\bm{f}}_{1}=[f_{1},\eta_{11},\eta_{21}] be continuous mappings of 𝑺1{\bm{S}}_{1} into 𝑺2{\bm{S}}_{2}. There are κjCHomeo(Sj0,Sj1)\kappa_{j}\in\operatorname{CHomeo}(S_{j0},S_{j1}), j=1,2j=1,2, with κjηj0ηj1\kappa_{j}\circ\eta_{j0}\simeq\eta_{j1}, where 𝑺j=[Sj0,ηj0]=[Sj1,ηj1]{\bm{S}}_{j}=[S_{j0},\eta_{j0}]=[S_{j1},\eta_{j1}]. If f1κ1κ2f0f_{1}\circ\kappa_{1}\simeq\kappa_{2}\circ f_{0}, then we say that 𝒇1{\bm{f}}_{1} is homotopic to 𝒇0{\bm{f}}_{0} and write 𝒇1𝒇0{\bm{f}}_{1}\simeq{\bm{f}}_{0}. Again, this definition does not depend on a particular choice of representatives.

A continuous mapping of 𝑺1{\bm{S}}_{1} into 𝑺2{\bm{S}}_{2} is said to be homotopically consistent if it is homotopic to a homotopically consistent topological embedding of 𝑺1{\bm{S}}_{1} into 𝑺2{\bm{S}}_{2}. If 𝒇jCont(𝑺j,𝑺j+1){\bm{f}}_{j}\in\operatorname{Cont}({\bm{S}}_{j},{\bm{S}}_{j+1}), j=1,2j=1,2, are homotopically consistent, then so is 𝒇2𝒇1{\bm{f}}_{2}\circ{\bm{f}}_{1}. We use the notation Conthc(𝑺,𝑹)\operatorname{Cont}_{\mathrm{hc}}({\bm{S}},{\bm{R}}) to express the set of homotopically consistent elements in Cont(𝑺,𝑹)\operatorname{Cont}({\bm{S}},{\bm{R}}).

Definition 3.13 (continuation).

By a continuation of 𝑺𝔉g{\bm{S}}\in\mathfrak{F}_{g} we mean a pair (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) where 𝑹𝔉g{\bm{R}}\in\mathfrak{F}_{g} and 𝜾CEmbhc(𝑺,𝑹){\bm{\iota}}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{S}},{\bm{R}}).

A continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) is called closed (resp. open, compact) if 𝑹\bm{R} is closed (resp. open, compact). We say that (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) is dense if so is (R,ι)(R,\iota), where (S,η)𝑺(S,\eta)\in{\bm{S}}, (R,θ)𝑹(R,\theta)\in{\bm{R}} and (ι,η,θ)𝜾(\iota,\eta,\theta)\in{\bm{\iota}}. If (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) and (𝑹,𝜾)({\bm{R}}^{\prime},{\bm{\iota}}^{\prime}) are continuations of 𝑺\bm{S} and 𝑹\bm{R}, respectively, then (𝑹,𝜾𝜾)({\bm{R}}^{\prime},{\bm{\iota}}^{\prime}\circ{\bm{\iota}}) is a continuation of 𝑺\bm{S}.

Let 𝑺=[S,η]𝔉g{\bm{S}}=[S,\eta]\in\mathfrak{F}_{g}. If (R,ι)(R,\iota) is a genus-preserving continuation of SS, then, setting 𝑹=[R,ιη]{\bm{R}}=[R,\iota\circ\eta] and 𝜾=[ι,η,ιη]{\bm{\iota}}=[\iota,\eta,\iota\circ\eta], we obtain a continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) of 𝑺\bm{S}. In particular, if SS is a subsurface of RR and ι:SR\iota:S\to R is the inclusion mapping, then we call (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) the inclusion continuation of 𝑺\bm{S}. If this is the case, then for 𝒇Cont(𝑹,𝑹){\bm{f}}\in\operatorname{Cont}({\bm{R}},{\bm{R}}^{\prime}) we call 𝒇𝜾{\bm{f}}\circ{\bm{\iota}} the restriction of 𝐟\bm{f} to 𝐒\bm{S} and denote it by 𝒇|𝑺{\bm{f}}|_{\bm{S}}.

Let 𝑺=[S,η]{\bm{S}}=[S,\eta] be a marked finite open Riemann surface of genus gg, and let (S˘,ι˘)(\breve{S},\breve{\iota}) be a natural compact continuation of SS. Setting 𝑺˘=[S˘,ι˘η]\breve{\bm{S}}=[\breve{S},\breve{\iota}\circ\eta] and 𝜾˘=[ι˘,η,ι˘η]\breve{\bm{\iota}}=[\breve{\iota},\eta,\breve{\iota}\circ\eta], we obtain a continuation (𝑺˘,𝜾˘)(\breve{\bm{S}},\breve{\bm{\iota}}) of 𝑺\bm{S}, which is referred to as the natural compact continuation of 𝑺\bm{S}.

Let 𝑺1,𝑺2𝔉g{\bm{S}}_{1},{\bm{S}}_{2}\in\mathfrak{F}_{g}. For quasiconformal embeddings 𝒇\bm{f} of 𝑺1{\bm{S}}_{1} into 𝑺2{\bm{S}}_{2} we can speak of their maximal dilatations K(𝒇)K({\bm{f}}). If 𝑺1{\bm{S}}_{1} belongs to 𝔗g\mathfrak{T}_{g}, then so does 𝑺2{\bm{S}}_{2}, and their Teichmüller distance dT(𝑺1,𝑺2)d_{T}({\bm{S}}_{1},{\bm{S}}_{2}) is defined to be inf(logK(𝒇))/2\inf(\log K({\bm{f}}))/2, where the infimum is taken over all homotopically consistent quasiconformal homeomorphisms 𝒇\bm{f} of 𝑺1{\bm{S}}_{1} onto 𝑺2{\bm{S}}_{2}. The complete metric space 𝔗g\mathfrak{T}_{g} is in fact known to be a complex manifold homeomorphic to 2dg\mathbb{R}^{2d_{g}} and biholomorphic to a bounded domain in dg\mathbb{C}^{d_{g}}, where dg=max{g,3g3}d_{g}=\max\{g,3g-3\}.

Now, for 𝑹0=[R0,θ0]𝔉g{\bm{R}}_{0}=[R_{0},\theta_{0}]\in\mathfrak{F}_{g} we are concerned with the set 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) of 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g} such that CEmbhc(𝑹0,𝑹)\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}})\neq\varnothing. It follows from Bochner [10, Satz V] that 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is nonempty.

Example 3.14.

Under the notations in Example 3.6 the correspondence τ𝑻τ\tau\mapsto{\bm{T}}_{\tau} is a biholomorphism of \mathbb{H} onto 𝔗1\mathfrak{T}_{1}. If we identify 𝔗1\mathfrak{T}_{1} with \mathbb{H} through the biholomorphism, then the Teichmüller distance on 𝔗1\mathfrak{T}_{1} coincides with the distance on \mathbb{H} induced by the hyperbolic metric |dz|/(2Imz)|dz|/(2\operatorname{Im}z). Now, let 𝑹0{\bm{R}}_{0} be a marked open Riemann surface of genus one. Then 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is a closed disk or a singleton in 𝔗1\mathfrak{T}_{1} by [56, Theorem 5]. It degenerates to a singleton if and only if R0OADR_{0}\in O_{AD} (see [56, Theorem 6]). Note that a finite open Riemann surface belongs to OADO_{AD} if and only if it is analytically finite.

4 Self-weldings with positive quadratic differentials

We introduce an operation called a self-welding, which brings us genus-preserving closed continuations of finite open Riemann surfaces. Though the procedure is plain, the harvest is rich. The current and next sections are devoted to developing an elementary theory of self-weldings.

Let φ=φ(z)dz2\varphi=\varphi(z)\,dz^{2} be a quadratic differential on a Riemann surface SS; it is an assignment of a function φ(z)\varphi(z) of zz to each local coordinate zz so that φ(z)(dz)2\varphi(z)\,(dz)^{2} is invariant under coordinate changes. Then |φ|:=|φ(z)|dxdy|\varphi|:=|\varphi(z)|\,dx\wedge dy, z=x+iyz=x+iy, defines a 2-form on SS. If φ\varphi is measurable, that is, φ(z)\varphi(z) is measurable for each zz, then set

φE=E|φ|\|\varphi\|_{E}=\iint_{E}|\varphi|

for measurable subsets EE of SS. If φE\|\varphi\|_{E} is finite, then φ\varphi is said to be integrable over EE. Also, |φ|:=|φ(z)||dz|\sqrt{|\varphi|\,}:=\sqrt{|\varphi(z)|\,}\,|dz| and |Imφ|:=|Im(φ(z)dz)|=|Im(φ(z))dx+Re(φ(z))dy||\operatorname{Im}\sqrt{\varphi\,}|:=|\operatorname{Im}(\sqrt{\varphi(z)\,}\,dz)|=|\operatorname{Im}(\sqrt{\varphi(z)\,})\,dx+\operatorname{Re}(\sqrt{\varphi(z)\,})\,dy| are invariant under coordinate changes. The φ\varphi-length Lφ(c)L_{\varphi}(c) and the φ\varphi-height Hφ(c)H_{\varphi}(c) of a curve cc on SS are defined by

Lφ(c)=c|φ|andHφ(c)=c|Imφ|,L_{\varphi}(c)=\int_{c}\sqrt{|\varphi|\,}\quad\text{and}\quad H_{\varphi}(c)=\int_{c}|\operatorname{Im}\sqrt{\varphi\,}|,

respectively, provided that the integrals are meaningful. Otherwise, we just set Lφ(c)=L_{\varphi}(c)=\infty or Hφ(c)=H_{\varphi}(c)=\infty. Note that φE\|\varphi\|_{E} is the area of EE with respect to the area element corresponding to the length element |φ|\sqrt{|\varphi|\,}.

Let (R,ι)(R,\iota) be a continuation of a Riemann surface SS, and let ψ\psi be a quadratic differential on RR. For pSp\in S take local coordinates zz and ww around pp and ι(p)\iota(p), respectively, and consider ι\iota as a holomorphic function w=ι(z)w=\iota(z). Set φ(z)=ψ(ι(z))ι(z)2\varphi(z)=\psi(\iota(z))\iota^{\prime}(z)^{2}, where ψ(w)\psi(w) is the function assigned to the local coordinate ww by the quadratic differential ψ\psi. Then assigning φ(z)\varphi(z) to zz defines a quadratic differential on SS called the pull-back of ψ\psi by ι\iota to be denoted by ιψ\iota^{*}\psi.

Let ιCEmb(S,R)\iota\in\operatorname{CEmb}(S,R). Considering ι\iota as a conformal homeomorphism SS onto ι(S)\iota(S), we can speak of the pull-back of a quadratic differential φ\varphi on SS by ι1\iota^{-1}. It is a quadratic differential on ι(S)\iota(S), which will be denoted by ιφ\iota_{*}\varphi.

A quadratic differential φ=φ(z)dz2\varphi=\varphi(z)\,dz^{2} is called meromorphic (resp. holomorphic) if φ(z)\varphi(z) is meromorphic (resp. holomorphic) for each local coordinate zz. If SS is a bordered Riemann surface and φ\varphi is a meromorphic (resp. holomorphic) quadratic differential on SS, then for any open continuation (R,ι)(R,\iota) of SS there is a meromorphic (resp. holomorphic) quadratic differential ψ\psi on a neighborhood of ι(S)\iota(S) such that the pull-back ιψ\iota^{*}\psi coincides with φ\varphi on SS.

Let M(S)M(S) be the complex vector space of meromorphic quadratic differentials on SS. Denote by A(S)A(S) the subspace composed of holomorphic quadratic differentials on SS. For nonzero φM(S)\varphi\in M(S) the algebraic degree of φ\varphi at pSp\in S is denoted by ordpφ\operatorname{ord}_{p}\varphi. If S\partial S\neq\varnothing and pSp\in\partial S, then, taking (R,ι)(R,\iota) and ψ\psi as in the preceding paragraph, we understand that ordpφ=ordι(p)ψ\operatorname{ord}_{p}\varphi=\operatorname{ord}_{\iota(p)}\psi. By a critical point of φ\varphi we mean a point pp for which ordpφ0\operatorname{ord}_{p}\varphi\neq 0. It is a zero of φ\varphi of order ordpφ\operatorname{ord}_{p}\varphi if ordpφ>0\operatorname{ord}_{p}\varphi>0 while it is a pole of φ\varphi of order ordpφ-\operatorname{ord}_{p}\varphi if ordpφ<0\operatorname{ord}_{p}\varphi<0. Zeros and poles on the border S\partial S are referred to as border zeros and border poles, respectively. If n:=ordpφ1n:=\operatorname{ord}_{p}\varphi\geqq-1, then there is a local coordinate ζ\zeta around pp with ζ(p)=0\zeta(p)=0 such that φ=(n/2+1)2ζndζ2\varphi=(n/2+1)^{2}\zeta^{n}\,d\zeta^{2} (see Strebel [65, Theorem 6.1]). Such a local coordinate ζ\zeta will be called a natural parameter of φ\varphi around pp. If φU<+\|\varphi\|_{U}<+\infty for an open set UU, then ordpφ1\operatorname{ord}_{p}\varphi\geqq-1 for all pUp\in U, that is, all poles of φ\varphi in UU are simple.

A meromorphic quadratic differential φ=φ(z)dz2\varphi=\varphi(z)\,dz^{2} is said to be positive along a smooth curve a:ISa:I\to S, where II is an interval in \mathbb{R} which may be closed, half-closed, open or infinite, if φ((za)(t))(za)(t)2\varphi((z\circ a)(t))(z\circ a)^{\prime}(t)^{2} is finite and positive whenever a(t)a(t) is in the domain of a local coordinate zz. A horizontal trajectory of φ\varphi is, by definition, a maximal smooth curve along which φ\varphi is positive. A horizontal trajectory can be a loop called a closed horizontal trajectory. Note that any horizontal trajectory of φ\varphi contains no critical points of φ\varphi. By a horizontal arc we mean a simple subarc, not a loop, of a horizontal trajectory of φ\varphi. For example, if ζ\zeta is a natural parameter of φ\varphi around pSp\in S^{\circ} with n:=ordpφ1n:=\operatorname{ord}_{p}\varphi\geqq-1, then for sufficiently small δ>0\delta>0 the arcs (0,δ)tζ1(te2kπi/(n+2))(0,\delta)\ni t\mapsto\zeta^{-1}(te^{2k\pi i/(n+2)}), k=0,1,,n+1k=0,1,\ldots,n+1, are horizontal arcs of φ\varphi emanating from pp though pp does not lie on the arcs. If pp is a noncritical point of φ\varphi, that is, if n=0n=0, then (δ,δ)tζ1(t)(-\delta,\delta)\ni t\mapsto\zeta^{-1}(t) is a horizontal arc of φ\varphi passing through pp.

Suppose that SS is a bordered Riemann surface. A nonzero element φ\varphi of M(S)M(S) is called positive if S\partial S consists of horizontal trajectories and, possibly, zeros of φ\varphi; thus φ\varphi is holomorphic on S\partial S. The set of positive quadratic differentials in M(S)M(S) is denoted by M+(S)M_{+}(S). Define A+(S)=A(S)M+(S)A_{+}(S)=A(S)\cap M_{+}(S). If S^\hat{S} denotes the double of SS (for the definition, see [3, II.3E]), then the reflection principle enables us to extend every element in M+(S)M_{+}(S) (resp. A+(S)A_{+}(S)) to a quadratic differential in M(S^)M(\hat{S}) (resp. A(S^)A(\hat{S})). Note that the orders of border zeros of elements in M+(S)M_{+}(S) are even. For the sake of convenience we set M+(S)=M(S){0}M_{+}(S)=M(S)\setminus\{0\} and A+(S)=A(S){0}A_{+}(S)=A(S)\setminus\{0\} for closed Riemann surfaces SS.

Let R0R_{0} be a finite open Riemann surface, and let (R˘0,ι˘0)(\breve{R}_{0},\breve{\iota}_{0}) be a natural compact continuation of R0R_{0}. Then R˘0\breve{R}_{0} is a compact Riemann surface. Set M+(R0)=ι˘0M+(R˘0)M_{+}(R_{0})=\breve{\iota}_{0}^{*}M_{+}(\breve{R}_{0}) and A+(R0)=ι˘0A+(R˘0)A_{+}(R_{0})=\breve{\iota}_{0}^{*}A_{+}(\breve{R}_{0}). These spaces do not depend of a particular choice of (R˘0,ι˘0)(\breve{R}_{0},\breve{\iota}_{0}), and are determined solely by R0R_{0}.

Proposition 4.1.

Let R0R_{0} be a finite open Riemann surface, and let φ\varphi be a nonzero meromorphic quadratic differential on R0R_{0}. If there is a sequence {φn}\{\varphi_{n}\} in A+(R0)A_{+}(R_{0}) such that φnφU0\|\varphi_{n}-\varphi\|_{U}\to 0 as nn\to\infty for some open subset UU of R0R_{0}, then φA+(R0)\varphi\in A_{+}(R_{0}) and φnφR00\|\varphi_{n}-\varphi\|_{R_{0}}\to 0 as nn\to\infty.

Proof.

Let (R˘0,ι˘0)(\breve{R}_{0},\breve{\iota}_{0}) be a natural compact continuation of R0R_{0}. If R˘0\breve{R}_{0} is bordered, then let RR denote its double. Otherwise, set R=R˘0R=\breve{R}_{0}. In either case RR is a closed Riemann surface. Identifying R0R_{0} with ι˘(R0)\breve{\iota}(R_{0}), we consider R0R_{0} as a subdomain of RR, and regard A+(R0)A_{+}(R_{0}) as a subset of A(R)A(R). Both R\|\cdot\|_{R} and U\|\cdot\|_{U} are norms on A(R)A(R); note that ψU=0\|\psi\|_{U}=0 implies ψ=0\psi=0 by the identity theorem. Since A(R)A(R) is of finite dimension, there is M>0M>0 such that ψUψRMψU\|\psi\|_{U}\leqq\|\psi\|_{R}\leqq M\|\psi\|_{U} for all ψA(R)\psi\in A(R).

Now, since {φn}\{\varphi_{n}\} is a Cauchy sequence in the Banach space (A(R),U)(A(R),\|\cdot\|_{U}), it is also a Cauchy sequence in (A(R),R)(A(R),\|\cdot\|_{R}) and hence there is ψA(R)\psi\in A(R) for which φnψR0\|\varphi_{n}-\psi\|_{R}\to 0 as nn\to\infty. The identity theorem implies that ψ=φ\psi=\varphi on R0R_{0}. Thus ψ\psi is nonzero and belongs to A+(R0)A_{+}(R_{0}). ∎

Definition 4.2 (self-welding).

Let SS be a compact bordered Riemann surface. A self-welding of SS is a pair R,ι\langle R,\iota\rangle, where

  • (i)

    RR is a compact Riemann surface,

  • (ii)

    ιCont(S,R)\iota\in\operatorname{Cont}(S,R) and ι|SCEmb(S,R)\iota|_{S^{\circ}}\in\operatorname{CEmb}(S^{\circ},R) with ι(S)=R\iota(S)=R, and

  • (iii)

    there are φM+(S)\varphi\in M_{+}(S) and ψM+(R)\psi\in M_{+}(R) with (ι|S)ψ=φ(\iota|_{S^{\circ}})^{*}\psi=\varphi on SS^{\circ} such that Rι(S)R\setminus\iota(S^{\circ}) consists of finitely many horizontal arcs and, possibly, critical points of ψ\psi.

The quadratic differentials φ\varphi and ψ\psi are referred to as a welder of the welding R,ι\langle R,\iota\rangle and the co-welder of the welder φ\varphi, respectively. If SS and RR are of the same genus, then R,ι\langle R,\iota\rangle is said to be genus-preserving. If RR is closed, then R,ι\langle R,\iota\rangle is called closed.

Two self-weldings R1,ι1\langle R_{1},\iota_{1}\rangle and R2,ι2\langle R_{2},\iota_{2}\rangle of SS are defined to be equivalent to each other if there is κCHomeo(R1,R2)\kappa\in\operatorname{CHomeo}(R_{1},R_{2}) such that ι2=κι1\iota_{2}=\kappa\circ\iota_{1}. Here, we do not force the self-weldings to have a common welder.

Remark.

A positive meromorphic quadratic differential on SS can induce two or more inequivalent self-weldings of SS (see Examples 4.3 and 5.2). Also, a self-welding of SS can have linearly independent welders (see Examples 4.3 and 5.5).

Example 4.3.

Consider the meromorphic functions ιj\iota_{j}, j=0,1j=0,1, on 𝔻¯\bar{\mathbb{D}} defined by ιj(z)=z+(1)j/z\iota_{j}(z)=z+(-1)^{j}/z. Then ^,ιj\langle\hat{\mathbb{C}},\iota_{j}\rangle, j=0,1j=0,1, are genus-preserving closed self-weldings of 𝔻¯\bar{\mathbb{D}}, which are inequivalent to each other. The meromorphic quadratic differential φ0:=(11/z2)2dz2\varphi_{0}:=(1-1/z^{2})^{2}\,dz^{2} on 𝔻¯\bar{\mathbb{D}} is a common welder of the self-weldings. The co-welders of φ0\varphi_{0} are ψ0:=dw2\psi_{0}:=dw^{2} for ^,ι0\langle\hat{\mathbb{C}},\iota_{0}\rangle and ψ1:=w2dw2/(w2+4)2\psi_{1}:=w^{2}\,dw^{2}/(w^{2}+4)^{2} for ^,ι1\langle\hat{\mathbb{C}},\iota_{1}\rangle. Note that φ0\varphi_{0} has double zeros at ±1𝔻\pm 1\in\partial\mathbb{D} while ψ0\psi_{0} has no critical points on ι0(𝔻)=[2,2]\iota_{0}(\partial\mathbb{D})=[-2,2]. Also, φ1:=(1+1/z2)2dz2\varphi_{1}:=-(1+1/z^{2})^{2}\,dz^{2} is a welder of ^,ι1\langle\hat{\mathbb{C}},\iota_{1}\rangle. Its co-welder is ψ0-\psi_{0}.

Definition 4.4 (self-welding continuation).

Let R0R_{0} be a nonanalytically finite open Riemann surface. A compact continuation (R,ι)(R,\iota) of R0R_{0} is called a self-welding continuation of R0R_{0} if there are a natural compact continuation (R˘0,ι˘0)(\breve{R}_{0},\breve{\iota}_{0}) of R0R_{0} and a self-welding R,κ\langle R,\kappa\rangle of R˘0\breve{R}_{0} such that ι=κι˘0\iota=\kappa\circ\breve{\iota}_{0}.

If φ˘M+(R˘0)\breve{\varphi}\in M_{+}(\breve{R}_{0}) is a welder of the self-welding R,κ\langle R,\kappa\rangle, then φ:=ι˘φ˘M+(R0)\varphi:=\breve{\iota}^{*}\breve{\varphi}\in M_{+}(R_{0}) is said to be a welder of the continuation (R,ι)(R,\iota). The co-welder of φ˘\breve{\varphi} is also referred to as the co-welder of φ\varphi.

Let SS be a compact bordered Riemann surface. If R,ι\langle R,\iota\rangle is a self-welding of SS, then ι1(R)\iota^{-1}(\partial R) is included in S\partial S. Moreover, (R,ι|S)(R,\iota|_{S^{\circ}}) is a self-welding continuation of SS^{\circ}. If the self-welding is closed, then so is the induced continuation.

Let R0R_{0} be a nonanalytically finite open Riemann surface, and let (R˘0,ι˘0)(\breve{R}_{0},\breve{\iota}_{0}) be a natural compact continuation of R0R_{0}. Equivalent self-weldings of R˘0\breve{R}_{0} induce equivalent continuations of R0R_{0}. Conversely, let (R,ι)(R,\iota) be a self-welding continuation of R0R_{0}, and let (R,ι)(R^{\prime},\iota^{\prime}) be a continuation of R0R_{0} equivalent to (R,ι)(R,\iota). Then there is a self-welding R,κ\langle R,\kappa\rangle of R˘0\breve{R}_{0} such that ι=κι˘0\iota=\kappa\circ\breve{\iota}_{0}. Also, there is κCHomeo(R,R)\kappa^{\prime}\in\operatorname{CHomeo}(R,R^{\prime}) such that ι=κι\iota^{\prime}=\kappa^{\prime}\circ\iota. Thus R,κκ\langle R^{\prime},\kappa^{\prime}\circ\kappa\rangle is a self-welding of R˘0\breve{R}_{0} equivalent to R,κ\langle R,\kappa\rangle and induces the continuation (R,ι)(R^{\prime},\iota^{\prime}) of R0R_{0}.

Proposition 4.5.

Let SS be a compact bordered Riemann surface, and let (R,ι)(R,\iota) be a dense compact continuation of SS^{\circ}. Suppose that there are φM+(S)\varphi\in M_{+}(S) and ψM+(R)\psi\in M_{+}(R) such that φ=ιψ\varphi=\iota^{*}\psi. Then ι\iota is extended to ι¯Cont(S,R)\bar{\iota}\in\operatorname{Cont}(S,R) with ι¯(S)=R\bar{\iota}(S)=R for which R,ι¯\langle R,\bar{\iota}\rangle is a self-welding of SS with welder φ\varphi.

Proof.

Take an arbitrary point p0p_{0} on the border S\partial S, and let ζ\zeta be a natural parameter of φ\varphi around p0p_{0}. We choose ζ\zeta so that it maps a neighborhood UU of p0p_{0} conformally onto a half-disk {ζ¯|ζ|<r}\{\zeta\in\bar{\mathbb{H}}\mid|\zeta|<r\}. If ordp0φ=2n\operatorname{ord}_{p_{0}}\varphi=2n, then φ=(n+1)2ζ2ndζ2\varphi=(n+1)^{2}\,\zeta^{2n}\,d\zeta^{2} on UU; recall that the order of φ\varphi at p0p_{0} is even. Set Uk={pUkπ/(n+1)<argζ(p)<(k+1)π/(n+1)}U_{k}=\{p\in U\mid k\pi/(n+1)<\arg\zeta(p)<(k+1)\pi/(n+1)\} for k=0,1,,nk=0,1,\dots,n. Thus a single-valued branch of the integral Φ\Phi of φ\sqrt{\varphi\,} on USU\cap S^{\circ} is represented as Φ(ζ)=ζn+1\Phi(\zeta)=\zeta^{n+1} and maps each UkU_{k} onto a half disk of radius rn+1r^{n+1} centered at 0. Let {pν}\{p_{\nu}\} be a sequence in a sector UkU_{k} converging to p0p_{0}. Since RR is compact, a subsequence of {ι(pν)}\{\iota(p_{\nu})\}, to be denoted again by {ι(pν)}\{\iota(p_{\nu})\}, converges to a point q0q_{0} in RR. As φ=ιψ\varphi=\iota^{*}\psi, the function Ψ\Psi on ι(US)\iota(U\cap S^{\circ}) defined by Ψι=Φ\Psi\circ\iota=\Phi is a single-valued branch of the integral of ψ\sqrt{\psi\,}, and ι\iota maps horizontal arcs of φ\varphi in UkU_{k} onto those of ψ\psi in ι(Uk)\iota(U_{k}). Since Ψ(ι(pν))=Φ(pν)0\Psi(\iota(p_{\nu}))=\Phi(p_{\nu})\to 0 as ν0\nu\to 0, we know that m:=ordq0ψ1m:=\operatorname{ord}_{q_{0}}\psi\geqq-1. Let ω\omega be a natural parameter of ψ\psi around q0q_{0} so that ψ=(m/2+1)2ωmdω2\psi=(m/2+1)^{2}\omega^{m}\,d\omega^{2}, and set Vl={q0<|ω(p)|<r(2n+2)/(m+2),2lπ/(m+2)<argω(p)<2(l+1)π/(m+2)}V_{l}=\{q\mid 0<|\omega(p)|<r^{(2n+2)/(m+2)},2l\pi/(m+2)<\arg\omega(p)<2(l+1)\pi/(m+2)\} for l=0,1,,m+1l=0,1,\dots,m+1. Then ι\iota maps UkU_{k} conformally onto some VlV_{l}. It follows that ι\iota extends continuously to USU\cap\partial S so that ι\iota maps each of the two components of (US){p0}(U\cap\partial S)\setminus\{p_{0}\} onto a horizontal arc of ψ\psi.

We have shown that ι\iota is extended to a continuous mapping ι¯\bar{\iota} of SS into RR so that ι¯(S)\bar{\iota}(\partial S) consists of finitely many horizontal arcs and critical points of ψ\psi. Since ι¯(S)\bar{\iota}(S^{\circ}) is dense in RR and ι¯(S)\bar{\iota}(S) is compact, we infer that ι¯\bar{\iota} is surjective and that Rι¯(S)R\setminus\bar{\iota}(S^{\circ}) is included in ι¯(S)\bar{\iota}(\partial S). Consequently, R,ι¯\langle R,\bar{\iota}\rangle is a self-welding of SS with welder φ\varphi. ∎

Corollary 4.6.

Let R0R_{0} be a nonanalytically finite open Riemann surface, and let (R,ι)(R,\iota) be a dense compact continuation of R0R_{0}. If there are φM+(R0)\varphi\in M_{+}(R_{0}) and ψM+(R)\psi\in M_{+}(R) such that φ=ιψ\varphi=\iota^{*}\psi, then (R,ι)(R,\iota) is a self-welding continuation of R0R_{0} with welder φ\varphi.

Proof.

Let (R˘0,ι˘0)(\breve{R}_{0},\breve{\iota}_{0}) be a natural compact continuation of R0R_{0}. Since punctures are removable singularities for conformal embeddings into compact Riemann surfaces, there is κCEmb((R˘0),R)\kappa\in\operatorname{CEmb}((\breve{R}_{0})^{\circ},R) such that ι=κι˘0\iota=\kappa\circ\breve{\iota}_{0}. Clearly, (R,κ)(R,\kappa) is a dense compact continuation of (R˘0)(\breve{R}_{0})^{\circ}. Moreover, φ˘=κψ\breve{\varphi}=\kappa^{*}\psi on (R˘0)(\breve{R}_{0})^{\circ} for some φ˘M+(R˘0)\breve{\varphi}\in M_{+}(\breve{R}_{0}), for, ι˘0κψ=ιψ=φM+(R0)\breve{\iota}_{0}^{*}\kappa^{*}\psi=\iota^{*}\psi=\varphi\in M_{+}(R_{0}). It follows from Proposition 4.5 that κ\kappa is extended to κ¯Cont(R˘0,R)\bar{\kappa}\in\operatorname{Cont}(\breve{R}_{0},R) with κ¯(R˘0)=R\bar{\kappa}(\breve{R}_{0})=R for which R,κ¯\langle R,\bar{\kappa}\rangle is a self-welding of R˘0\breve{R}_{0} with welder φ˘\breve{\varphi}. As ι=κ¯ι˘0\iota=\bar{\kappa}\circ\breve{\iota}_{0} and φ=ι˘0φ˘\varphi=\breve{\iota}_{0}^{*}\breve{\varphi}, we obtain the corollary. ∎

Corollary 4.7.

Let R0R_{0} be a nonanalytically finite open Riemann surface, and let (R,ι)(R,\iota) be a compact continuation of R0R_{0}. If there is ψM+(R)\psi\in M_{+}(R) such that Rι(R0)R\setminus\iota(R_{0}) consists of finitely many horizontal arcs of ψ\psi together with finitely many points, then (R,ι)(R,\iota) is a self-welding continuation of R0R_{0} with welder ιψ\iota^{*}\psi.

Proof.

Let (R˘0,ι˘0)(\breve{R}_{0},\breve{\iota}_{0}) and κ\kappa be as in the proof of the preceding corollary. Thus (R,κ)(R,\kappa) is a dense compact continuation of (R˘0)(\breve{R}_{0})^{\circ} with ι=κι˘0\iota=\kappa\circ\breve{\iota}_{0}. Since Rκ((R˘0))R\setminus\kappa((\breve{R}_{0})^{\circ}) consists of finitely many horizontal arcs of ψ\psi together with finitely many points, it follows from a theorem of Caratédory that κ\kappa is extended to a continuous mapping κ¯\bar{\kappa} of R˘0\breve{R}_{0} onto RR, which is holomorphic on R˘0\partial\breve{R}_{0} off finitely points by the reflection principle, and hence the pull-back κψ\kappa^{*}\psi is extended to a quadratic differential φ˘\breve{\varphi} in M+(R˘0)M_{+}(\breve{R}_{0}). Corollary 4.6 shows that (R,ι)(R,\iota) is a self-welding continuation of RR with welder ιψ=ι˘0φ˘M+(R0)\iota^{*}\psi=\breve{\iota}_{0}^{*}\breve{\varphi}\in M_{+}(R_{0}). ∎

We now introduce a constructive procedure to obtain self-weldings of a compact bordered Riemann surface SS. It is different from the welding procedure introduced in Ahlfors-Sario [3, II.3C–D] because ι\iota is not required to be holomorphic on the border S\partial S. In addition, we need to take into account quadratic differentials on SS and the corresponding quadratic differentials on the resulting Riemann surfaces. In the following, for a curve c:IXc:I\to X on a topological space XX, where II is an interval on \mathbb{R}, by abuse of notation we sometimes use the same letter cc to denote its image c(I)c(I). Let φM+(S)\varphi\in M_{+}(S), and let a1a_{1} and a2a_{2} be simple arcs, not loops, of the same φ\varphi-length, say, LL, on the border S\partial S. They may pass through zeros of φ\varphi. Suppose that each of the arcs contains its endpoints and that the arcs are nonoverlapping, that is, a1a_{1}^{\circ} and a2a_{2}^{\circ} are disjoint, where cc^{\circ} stands for the part obtained from a simple arc cc by deleting its endpoints; we allow a1a_{1} and a2a_{2} to have a common endpoint. Then we can identify the arcs to obtain a new Riemann surface so that φ\varphi induces a meromorphic quadratic differential on the new surface.

Specifically, parametrize aka_{k}, k=1,2k=1,2, with φ\varphi-length parameter. Thus the φ\varphi-length of the subarc ak|[0,s]a_{k}|_{[0,s]} of the arc ak:[0,L]Sa_{k}:[0,L]\to\partial S is identical with ss. There are two natural ways of identifying a1a_{1} and a2a_{2}. One is to identify a1(s)a_{1}(s) with a2(s)a_{2}(s) for s[0,L]s\in[0,L], and the other is to identify a1(s)a_{1}(s) with a2(Ls)a_{2}(L-s) for s[0,L]s\in[0,L]. Exactly one of the identifications leads us to an orientable topological surface RR, possibly with border. For the sake of definiteness we assume that the first identification has been chosen. The genus of RR is the same as that of SS if and only if a1a_{1} and a2a_{2} lie on the same component of S\partial S. Let ι:SR\iota:S\to R be the natural continuous mapping. Then ιa1\iota\circ a_{1} is identical with ιa2\iota\circ a_{2}, which defines a simple arc a:[0,L]Ra:[0,L]\to R. Note that aa^{\circ} lies in the interior RR^{\circ} though the endpoints of aa may or may not be on the border R\partial R.

We endow RR^{\circ} with a conformal structure as follows. We begin with adopting local parameters at each point of RR^{\circ} off aa so that ι\iota is a conformal homeomorphism of SS^{\circ} onto ι(S)\iota(S^{\circ}). We then apply an extension of the welding procedure described in [3, II.3C–D] to choose local parameters at each point on aRa\cap R^{\circ}. To be more precise, taking s0(0,L)s_{0}\in(0,L), let ζk=ξk+iηk\zeta_{k}=\xi_{k}+i\eta_{k} be a natural parameter of φ\varphi around pk:=ak(s0)p_{k}:=a_{k}(s_{0}), where SS is considered as a subsurface of its double S^\hat{S}. A small neighborhood UkU_{k} of pkp_{k} in S^\hat{S} is mapped conformally onto a neighborhood of 0 in \mathbb{C} by ζk\zeta_{k}. We may suppose that SS^{\circ} lies on the left and right of a1a_{1} and a2a_{2}, respectively, so that ζ1(U1S)\zeta_{1}(U_{1}\cap S^{\circ})\subset\mathbb{H} while ζ2(U2S)¯=\zeta_{2}(U_{2}\cap S^{\circ})\cap\bar{\mathbb{H}}=\varnothing. Note that

s=sgn(ξkak)(s)|(ξkak)(s)|nk+1+s0s=\operatorname{sgn}(\xi_{k}\circ a_{k})(s)\cdot|(\xi_{k}\circ a_{k})(s)|^{n_{k}+1}+s_{0}

for ss sufficiently near s0s_{0}, where 2nk=ordpkφ2n_{k}=\operatorname{ord}_{p_{k}}\varphi with nonnegative integer nkn_{k}. Take a neighborhood UU of a(s0)=ι(p1)=ι(p2)a(s_{0})=\iota(p_{1})=\iota(p_{2}) included in ι((U1U2)S)\iota((U_{1}\cup U_{2})\cap S), and define a mapping z:Uz:U\to\mathbb{C} by

zι(p)={ζ1(p)2(n1+1)/(n1+n2+2)if pU1S,ζ2(p)2(n2+1)/(n1+n2+2)if pU2Sz\circ\iota(p)=\begin{cases}\zeta_{1}(p)^{2(n_{1}+1)/(n_{1}+n_{2}+2)}&\text{if $p\in U_{1}\cap S$},\\ \zeta_{2}(p)^{2(n_{2}+1)/(n_{1}+n_{2}+2)}&\text{if $p\in U_{2}\cap S$}\end{cases}

with za(s)>0z\circ a(s)>0 for s>s0s>s_{0}. Then zz is a well-defined element of TEmb(U,)\operatorname{TEmb}(U,\mathbb{C}), which is holomorphic off aa. With the aid of these parameters zz we make R{a(0),a(l)}R^{\circ}\setminus\{a(0),a(l)\} a Riemann surface. If the endpoints of aa lie on the border R\partial R, then we are done. If a(0)Ra(0)\in R^{\circ}, then a1(0)=a2(0)a_{1}(0)=a_{2}(0). Let ζ\zeta be a natural parameter of φ\varphi around the common endpoint. Then the function zz defined by zι(p)=ζ(p)2z\circ\iota(p)=\zeta(p)^{2} is homeomorphic on a neighborhood of a(0)a(0) and holomorphic on a punctured neighborhood of a(0)a(0). Thus zz works as an analytic local parameter around a(0)a(0). We deal with the case where a(L)Ra(L)\in R^{\circ} in a similar manner. We have thus endowed RR^{\circ} with conformal structure.

If (R˘,ι˘)(\breve{R},\breve{\iota}) is a natural compact continuation of RR^{\circ}, then ι˘\breve{\iota} is extended to a homeomorphism of RR onto R˘\breve{R}. We introduce a conformal structure on RR so that the homeomorphism is actually a conformal homeomorphism of RR onto R˘\breve{R}. We have thus obtained a compact Riemann surface RR.

The continuous mapping ι\iota is holomorphic on S(a1a2)S\setminus(a_{1}\cup a_{2}). Let pp be a point on a1a2a_{1}\cup a_{2}. If pp lies on a1a2a_{1}^{\circ}\cup a_{2}^{\circ}, then ι\iota is holomorphic at pp if and only if ordpφ=ordqφ\operatorname{ord}_{p}\varphi=\operatorname{ord}_{q}\varphi, where qq is the other point than pp projected to ι(p)\iota(p). If pp is an endpoint of a1a_{1} or a2a_{2}, then ι\iota is holomorphic at pp if and only if pp is a common endpoint of a1a_{1} and a2a_{2}, or equivalently, ι(p)\iota(p) lies in RR^{\circ}.

The quadratic differential ιφ\iota_{*}\varphi on ι(S)\iota(S^{\circ}) induced by ι\iota from φ\varphi is extended to a meromorphic quadratic differential ψ\psi on RR. Observe that

(4.1) orda(s)ψ={orda1(s)φ+orda2(s)φ2if 0<s<L,orda1(s)φ+orda2(s)φ41if “s=0 or L” and a(s)R,orda1(s)φ+orda2(s)φ+2if “s=0 or L” and a(s)R.\operatorname{ord}_{a(s)}\psi=\begin{cases}\dfrac{\,\operatorname{ord}_{a_{1}(s)}\varphi+\operatorname{ord}_{a_{2}(s)}\varphi\,}{2}&\text{if $0<s<L$},\\[8.61108pt] \dfrac{\,\operatorname{ord}_{a_{1}(s)}\varphi+\operatorname{ord}_{a_{2}(s)}\varphi\,}{4}-1&\text{if ``$s=0$ or $L$" and $a(s)\in R^{\circ}$},\\[8.61108pt] \operatorname{ord}_{a_{1}(s)}\varphi+\operatorname{ord}_{a_{2}(s)}\varphi+2&\text{if ``$s=0$ or $L$" and $a(s)\in\partial R$}.\end{cases}

In particular, ψ\psi has a pole on aa if and only if a1a_{1} and a2a_{2} has a common endpoint at which φ\varphi does not vanish. If this is the case, then the pole is simple and lies in the interior RR^{\circ}. In any case, ψ\psi is holomorphic and positive on R\partial R and hence belongs to M+(R)M_{+}(R). Clearly, the simple arc aa is composed of finitely many horizontal arcs and, possibly, critical points of ψ\psi. Therefore, R,ι\langle R,\iota\rangle is a self-welding of SS with welder φ\varphi, and ψ\psi is the co-welder of φ\varphi. We say that R,ι\langle R,\iota\rangle is the self-welding of SS with welder φ\varphi along (a1,a2)(a_{1},a_{2}).

More generally, let aka_{k}, k=1,,2Nk=1,\ldots,2N, be nonoverlapping simple arcs on S\partial S with Lj:=Lφ(a2j1)=Lφ(a2j)L_{j}:=L_{\varphi}(a_{2j-1})=L_{\varphi}(a_{2j}) for j=1,,Nj=1,\ldots,N. Identifying a2j1a_{2j-1} with a2ja_{2j} as above for j=1,,Nj=1,\ldots,N leads us to

  • (i)

    a compact Riemann surface RR,

  • (ii)

    a continuous mapping ι\iota of SS onto RR with ι|SCEmb(S,R)\iota|_{S^{\circ}}\in\operatorname{CEmb}(S^{\circ},R) such that ιa2j1(s)=ιa2j(s)R\iota\circ a_{2j-1}(s)=\iota\circ a_{2j}(s)\in R^{\circ} or ιa2j1(s)=ιa2j(Ljs)R\iota\circ a_{2j-1}(s)=\iota\circ a_{2j}(L_{j}-s)\in R^{\circ} for s(0,Lj)s\in(0,L_{j}), and

  • (iii)

    a quadratic differential ψ\psi in M+(R)M_{+}(R) with φ=(ι|S)ψ\varphi=(\iota|_{S^{\circ}})^{*}\psi on SS^{\circ}.

By subdividing aka_{k}’s if necessary, it is no harm to assume from the outset that each pair (a2j1,a2j)(a_{2j-1},a_{2j}) induces a simple arc, not a loop, on RR. The pair R,ι\langle R,\iota\rangle will be called the self-welding of SS with welder φ\varphi along (a2j1,a2j)(a_{2j-1},a_{2j}), j=1,,Nj=1,\ldots,N.

Remark.

For more general procedures and applications of weldings we refer the reader to Bishop [8, 9], Hamilton [21], Ishida [26], Maitani [33] and Nishikawa-Maitani [44], Oikawa [46], Semmes [54] and Williams [66].

Let R1,ι1\langle R_{1},\iota_{1}\rangle be the self-welding of SS with welder φ\varphi along (a2j1,a2j)(a_{2j-1},a_{2j}), j=1,,N1j=1,\ldots,N-1, and let ψ1M+(R1)\psi_{1}\in M_{+}(R_{1}) be the co-welder of φ\varphi. Then the image arcs (ι1)a2N1(\iota_{1})_{*}a_{2N-1} and (ι1)a2N(\iota_{1})_{*}a_{2N} are nonoverlapping simple arcs of the same ψ1\psi_{1}-length on the border R1\partial R_{1}. Hence we can construct the self-welding R,ι\langle R^{\prime},\iota^{\prime}\rangle of R1R_{1} with welder ψ1\psi_{1} along ((ι1)a2N1,(ι1)a2N)((\iota_{1})_{*}a_{2N-1},(\iota_{1})_{*}a_{2N}). It is easy to verify that R,ιι1\langle R^{\prime},\iota^{\prime}\circ\iota_{1}\rangle is a self-welding of SS equivalent to R,ι\langle R,\iota\rangle. Therefore, the self-welding R,ι\langle R,\iota\rangle is also obtained by consecutive applications of self-welding procedures along one pair of arcs.

Example 4.8.

The quadratic differential φ=(1/z2)dz2\varphi=(-1/z^{2})\,dz^{2} belongs to M+(𝔻¯)M_{+}(\bar{\mathbb{D}}), and the arcs a1,,a8a_{1},\ldots,a_{8} on 𝔻\partial\mathbb{D} defined by aj(t)=e(t+j1)πi/4a_{j}(t)=e^{(t+j-1)\pi i/4}, t[0,1]t\in[0,1], are of the same φ\varphi-length. Let R,ι\langle R,\iota\rangle be the self-welding of 𝔻¯\bar{\mathbb{D}} with welder φ\varphi along (a1,a6)(a_{1},a_{6}), (a2,a5)(a_{2},a_{5}), (a3,a8)(a_{3},a_{8}) and (a4,a7)(a_{4},a_{7}). Then RR is a torus. Hence the self-welding is closed but not genus-preserving.

Any self-welding of SS is a self-welding of SS along finitely many pairs of simple arcs on S\partial S. To see this, let R,ι\langle R,\iota\rangle be an arbitrary self-welding of SS with welder φ\varphi, and denote by ψ\psi the co-welder of φ\varphi. Then Rι(S)R\setminus\iota(S^{\circ}) consists of finitely many nonoverlapping simple arcs e1,,eNe_{1},\ldots,e_{N}. Note that every eje_{j} consists of horizontal arcs and, possibly, critical points of ψ\psi. Observe that there are nonoverlapping simple arcs a1,,a2Na_{1},\ldots,a_{2N} on S\partial S such that ιa2j1=ιa2j=ej\iota_{*}a_{2j-1}=\iota_{*}a_{2j}=e_{j} for each jj. Then R,ι\langle R,\iota\rangle is the self-welding of SS with welder φ\varphi along (a2j1,a2j)(a_{2j-1},a_{2j}), j=1,,Nj=1,\ldots,N.

Example 4.9.

Let ^,ιj\langle\hat{\mathbb{C}},\iota_{j}\rangle be the self-weldings of 𝔻¯\bar{\mathbb{D}} in Example 4.3. The arcs a01a_{01} and a02a_{02} on 𝔻\partial\mathbb{D} defined by a0k(t)=e(t+k1)πia_{0k}(t)=e^{(t+k-1)\pi i}, t[0,1]t\in[0,1], are of the same φ0\varphi_{0}-length, where φ0=(11/z2)2dz2\varphi_{0}=(1-1/z^{2})^{2}\,dz^{2}. Then ^,ι0\langle\hat{\mathbb{C}},\iota_{0}\rangle is the self-welding of 𝔻¯\bar{\mathbb{D}} with welder φ0\varphi_{0} along (a01,a02)(a_{01},a_{02}). If a1k:=ia0ka_{1k}:=ia_{0k}, then ^,ι1\langle\hat{\mathbb{C}},\iota_{1}\rangle is the self-welding of 𝔻¯\bar{\mathbb{D}} with welder φ0\varphi_{0} along (a11,a12)(a_{11},a_{12}).

5 Genus-preserving closed regular self-weldings

Let SS be a compact bordered Riemann surface. We introduce important classes of self-weldings of SS.

Definition 5.1 (full self-welding).

Let CC be a union of connected components of S\partial S. A genus-preserving self-welding R,ι\langle R,\iota\rangle of SS is called CC-full if ι(C)R\iota(C)\subset R^{\circ} while ι(SC)=R\iota(\partial S\setminus C)=\partial R.

As is easily verified, a self-welding of SS is S\partial S-full if and only if it is genus-preserving and closed. Note that the self-welding of SS with welder φ\varphi along (a2j1,a2j)(a_{2j-1},a_{2j}), j=1,,Nj=1,\ldots,N, is CC-full if and only if the arcs aka_{k}, k=1,,2Nk=1,\ldots,2N, exhaust CC and a2j1a_{2j-1} and a2ja_{2j} lie on the same component of CC without separating any other pairs for j=1,,Nj=1,\ldots,N.

Example 5.2.

Each φM+(S)\varphi\in M_{+}(S) induces a CC-full self-welding of SS. To obtain an example let C1,,CNC_{1},\dots,C_{N} be the connected components of CC. Divide each CjC_{j} into two subarcs a2j1a_{2j-1} and a2ja_{2j} of the same φ\varphi-length. Then the self-welding of SS with welder φ\varphi along (a2j1,a2j)(a_{2j-1},a_{2j}), j=1,,Nj=1,\ldots,N, is CC-full.

Let R,ι\langle R,\iota\rangle be a CC-full self-welding of SS with welder φ\varphi, and set Gι=ι(C)G_{\iota}=\iota(C). We consider GιG_{\iota} as a graph, called the weld graph of R,ι\langle R,\iota\rangle, by declaring the points pGιp\in G_{\iota} with cardι1(p)2\operatorname{card}\iota^{-1}(p)\neq 2 to be the vertices of GιG_{\iota}. Each component of GιG_{\iota} is in fact a tree. Note that each component of GιG_{\iota} contains more than one vertex. If vv is a vertex of GιG_{\iota}, then

(5.1) cardι1(v)=degGιvordvψ+2,\operatorname{card}\iota^{-1}(v)=\deg_{G_{\iota}}v\leqq\operatorname{ord}_{v}\psi+2,

where ψ\psi is the co-welder of φ\varphi. An end-vertex of GιG_{\iota} is a vertex vv with degGιv=1\deg_{G_{\iota}}v=1. It is characterized as a point on GιG_{\iota} whose preimage by ι\iota is a singleton (see (5.1)).

Definition 5.3 (regular self-welding).

A self-welding R,ι\langle R,\iota\rangle of SS is called φ\varphi-regular if φ\varphi is a welder of the self-welding and the co-welder of φ\varphi is holomorphic on RR. A self-welding of SS is said to be regular if it is φ\varphi-regular for some welder φ\varphi of the self-welding.

Remark.

If a self-welding R,ι\langle R,\iota\rangle of SS is φ\varphi-regular, then φ\varphi necessarily belongs to A+(S)A_{+}(S). Note that if ψ\psi is the co-welder of φ\varphi, then ψR=φS\|\psi\|_{R}=\|\varphi\|_{S}.

The next example is one of our motivations for introducing the notion of a genus-preserving closed regular self-welding. It also shows that an element of A+(S)A_{+}(S) can be a common welder of infinitely many inequivalent genus-preserving closed regular self-weldings of SS.

Example 5.4.

Let ω=ω(z)dz\omega=\omega(z)\,dz be a nonzero holomorphic semiexact 1-form on SS whose imaginary part vanishes along S\partial S (for the definition of a semiexact 1-form see [3, V.5B]). Let (R,ι)(R,\iota) be a hydrodynamic continuation of SS^{\circ} with respect to ω\omega introduced by Shiba-Shibata [64]. Thus RR is a closed Riemann surface of the same genus as SS, and ι\iota is a conformal embedding of SS^{\circ} into RR such that ω=ις\omega=\iota^{*}\varsigma for some holomorphic 1-form ς\varsigma on RR for which Rι(S)R\setminus\iota(S^{\circ}) consists of finitely many arcs along which Imς\operatorname{Im}\varsigma vanishes. Corollary 4.7 implies that ι\iota is extended to a continuous mapping ι¯\bar{\iota} of SS onto RR such that R,ι¯\langle R,\bar{\iota}\rangle is a genus-preserving closed ω2\omega^{2}-regular self-welding of SS; note that ω2\omega^{2} is a positive holomorphic quadratic differential on SS. As is remarked in [55, Section 26], for some SS there is ω\omega inducing uncountably many hydrodynamic continuations of SS^{\circ}, or uncountably many inequivalent genus-preserving closed ω2\omega^{2}-regular self-weldings of SS.

If φ\varphi is a welder of a self-welding of SS, then so is rφr\varphi for any positive real number rr. The next example shows that linearly independent quadratic differentials can induce the same genus-preserving closed regular self-welding.

Example 5.5.

Let RR be a closed Riemann surface of genus g>1g>1, and assume that it is symmetric in the sense that it admits an anti-conformal involution JJ whose set FF of fixed points is composed of finitely many analytic simple loops on RR. Take a simple arc cc on FF, and set R0=RcR_{0}=R\setminus c. The subsurface R0R_{0} is a nonanalytically finite open Riemann surface of genus gg. Let (R˘0,ι˘0)(\breve{R}_{0},\breve{\iota}_{0}) be a natural compact continuation of R0R_{0}. The element of CEmb(ι˘0(R0),R)\operatorname{CEmb}(\breve{\iota}_{0}(R_{0}),R) defined by ι˘0(p)p\breve{\iota}_{0}(p)\mapsto p, pR0p\in R_{0}, is extended to a continuous mapping ι\iota of R˘0\breve{R}_{0} onto RR. Let Ω\Omega be the vector space over \mathbb{R} of holomorphic 1-forms ω\omega on RR such that Imω=0\operatorname{Im}\omega=0 along FF. Recall that dimΩ=g\dim_{\mathbb{R}}\Omega=g. For any nonzero ωΩ\omega\in\Omega the pull-back ι(ω2)\iota^{*}(\omega^{2}) of its square belongs to A+(R˘0)A_{+}(\breve{R}_{0}) and the genus-preserving closed self-welding R,ι\langle R,\iota\rangle of R˘0\breve{R}_{0} is ι(ω2)\iota^{*}(\omega^{2})-regular.

Let CC be a union of connected components of S\partial S. Each positive meromorphic quadratic differential on SS defines infinitely many CC-full inequivalent self-weldings (see Example 5.2). We now ask which quadratic differentials φ\varphi on SS make CC-full φ\varphi-regular self-weldings.

Proposition 5.6.

Let φA+(S)\varphi\in A_{+}(S), and let CC be a union of connected components of S\partial S. A CC-full self-welding R,ι\langle R,\iota\rangle of SS is φ\varphi-regular if and only if every end-vertex of its weld graph is the image of a border zero of φ\varphi by ι\iota.

Proof.

Let R,ι\langle R,\iota\rangle be a CC-full self-welding of SS with welder φ\varphi, and let ψ\psi be the co-welder of φ\varphi. If pp is a point of S\partial S for which ι(p)\iota(p) is an end-vertex of the weld graph GιG_{\iota}, then

(5.2) ordι(p)ψ=ordpφ21\operatorname{ord}_{\iota(p)}\psi=\frac{\,\operatorname{ord}_{p}\varphi\,}{2}-1

by (4.1). If the self-welding is φ\varphi-regular, then ψ\psi is holomorphic on RR. Therefore, ordι(p)ψ0\operatorname{ord}_{\iota(p)}\psi\geqq 0 so that ordpφ2\operatorname{ord}_{p}\varphi\geqq 2 by (5.2). Hence pp is a border zero of φ\varphi.

If the self-welding is not φ\varphi-regular, then ψ\psi has a pole qq on RR. Since φA+(S)\varphi\in A_{+}(S), the co-welder ψ\psi is holomorphic on ι(S)\iota(S^{\circ}) so that the pole qq lies on GιG_{\iota}, which must be an end-vertex of GιG_{\iota} by (5.1). Another application of (5.2) shows that the point in ι1(q)\iota^{-1}(q) is not a zero of φ\varphi. This completes the proof. ∎

The following corollary is a generalization of Shiba-Shibata [64, Lemma 3].

Corollary 5.7.

Let φA+(S)\varphi\in A_{+}(S), and let CC be a union of connected components of S\partial S. If there is a CC-full φ\varphi-regular self-welding of SS, then each component of CC contains two or more zeros of φ\varphi.

Proof.

If R,ι\langle R,\iota\rangle is a CC-full φ\varphi-regular self-welding of SS, then each component of CC contains at least two points p1p_{1} and p2p_{2} whose images by ι\iota are end-vertices of the weld graph. Since the self-welding is φ\varphi-regular, p1p_{1} and p2p_{2} must be zeros of φ\varphi by Proposition 5.6. ∎

Definition 5.8 (border length condition).

Let CC be a union of connected components of S\partial S. A positive meromorphic quadratic differential φ\varphi on SS is said to satisfy the border length condition on CC if

Lφ(a)1 2Lφ(C)L_{\varphi}(a)\leqq\frac{1}{\,2\,}L_{\varphi}(C^{\prime})

for any component CC^{\prime} of CC and any horizontal trajectory aa of φ\varphi included in CC^{\prime}.

Denote by ML(S,C)M_{L}(S,C) the set of positive meromorphic quadratic differentials on SS satisfying the border length condition on CC, and set AL(S,C)=ML(S,C)A(S)A_{L}(S,C)=M_{L}(S,C)\cap A(S). We abbreviate ML(S,S)M_{L}(S,\partial S) and AL(S,S)A_{L}(S,\partial S) to ML(S)M_{L}(S) and AL(S)A_{L}(S), respectively.

Theorem 5.9.

Let SS be a compact bordered Riemann surface, and let CC be a union of connected components of S\partial S. For φA+(S)\varphi\in A_{+}(S) there exists a CC-full φ\varphi-regular self-welding of SS if and only if φ\varphi satisfies the border length condition on CC.

Corollary 5.10.

For φA+(S)\varphi\in A_{+}(S) there is a genus-preserving closed φ\varphi-regular self-welding of SS if and only if φAL(S)\varphi\in A_{L}(S).

For the proof Theorem 5.9 we make use of the following lemma. A topological space is said to be I-shaped (resp. Y-shaped ) if it is homeomorphic to [0,1][0,1] (resp. {zz3[0,1]}\{z\in\mathbb{C}\mid z^{3}\in[0,1]\}). For example, each component of the weld graph of the self-welding in Example 5.2 is I-shaped.

Lemma 5.11.

Let φM+(S)\varphi\in M_{+}(S). Let pjp_{j}, j=1,2,3j=1,2,3, be points on a component CC of S\partial S which divide CC into three arcs of φ\varphi-lengths less than Lφ(C)/2L_{\varphi}(C)/2. Then there is a CC-full self-welding R,ι\langle R,\iota\rangle of SS with welder φ\varphi such that its weld graph is a Y-shaped tree with end-vertices ι(pj)\iota(p_{j}), j=1,2,3j=1,2,3. If R,ι\langle R^{\prime},\iota^{\prime}\rangle is a CC-full self-welding of SS whose weld graph is Y-shaped with end-vertices ι(pj)\iota^{\prime}(p_{j}), j=1,2,3j=1,2,3, then R,ι\langle R^{\prime},\iota^{\prime}\rangle is equivalent to R,ι\langle R,\iota\rangle.

Proof.

We denote by cjc_{j}, j=1,2,3j=1,2,3, the arcs on CC obtained from CC by cutting CC at pjp_{j}, j=1,2,3j=1,2,3, where we label them so that pjcjp_{j}\not\in c_{j}. By assumption we have

Lφ(c1)+Lφ(c2)+Lφ(c3)=Lφ(C)L_{\varphi}(c_{1})+L_{\varphi}(c_{2})+L_{\varphi}(c_{3})=L_{\varphi}(C)

and

Lφ(cj)<Lφ(C)2,j=1,2,3.L_{\varphi}(c_{j})<\frac{\,L_{\varphi}(C)\,}{2},\qquad j=1,2,3.

Take a point qjq_{j} on cjc_{j}^{\circ} to divide cjc_{j} into two arcs, and name the resulting six arcs aj,aja_{j},a^{\prime}_{j}, j=1,2,3j=1,2,3, as follows:

c1=a2a3,c2=a3a1,c3=a1a2,\displaystyle c_{1}=a_{2}\cup a^{\prime}_{3},\quad c_{2}=a_{3}\cup a^{\prime}_{1},\quad c_{3}=a_{1}\cup a^{\prime}_{2},
ajaj={pj},j=1,2,3.\displaystyle a_{j}\cap a^{\prime}_{j}=\{p_{j}\},\quad j=1,2,3.

If we choose qjq_{j} so that

Lφ(aj)=Lφ(C)2Lφ(cj),j=1,2,3,L_{\varphi}(a_{j})=\frac{\,L_{\varphi}(C)\,}{2}-L_{\varphi}(c_{j}),\quad j=1,2,3,

then Lφ(aj)=Lφ(aj)L_{\varphi}(a_{j})=L_{\varphi}(a^{\prime}_{j}) for j=1,2,3j=1,2,3 and hence we can make a CC-full self-welding R,ι\langle R,\iota\rangle of SS with welder φ\varphi along (aj,aj)(a_{j},a^{\prime}_{j}), j=1,2,3j=1,2,3. Its weld graph GιG_{\iota} is a Y-shaped tree and the end-vertices of GιG_{\iota} are exactly ι(pj)\iota(p_{j}), j=1,2,3j=1,2,3.

Let R,ι\langle R^{\prime},\iota^{\prime}\rangle be another CC-full self-welding of SS with welder φ\varphi whose weld graph GιG_{\iota^{\prime}} is Y-shaped with end-vertices vj:=ι(pj)v^{\prime}_{j}:=\iota^{\prime}(p_{j}), j=1,2,3j=1,2,3. Since GιG_{\iota^{\prime}} is Y-shaped, it has one more vertex v4v^{\prime}_{4} with degGιv4=3\deg_{G_{\iota^{\prime}}}v^{\prime}_{4}=3 and three edges eje^{\prime}_{j}, j=1,2,3j=1,2,3, where vjejv^{\prime}_{j}\in e^{\prime}_{j}. As

Lψ(e1)+Lψ(e2)=Lφ(c3),Lψ(e2)+Lψ(e3)=Lφ(c1),Lψ(e3)+Lψ(e1)=Lφ(c2),L_{\psi^{\prime}}(e^{\prime}_{1})+L_{\psi^{\prime}}(e^{\prime}_{2})=L_{\varphi}(c_{3}),\quad L_{\psi^{\prime}}(e^{\prime}_{2})+L_{\psi^{\prime}}(e^{\prime}_{3})=L_{\varphi}(c_{1}),\quad L_{\psi^{\prime}}(e^{\prime}_{3})+L_{\psi^{\prime}}(e^{\prime}_{1})=L_{\varphi}(c_{2}),

where ψM+(R)\psi^{\prime}\in M_{+}(R^{\prime}) is the co-welder of φ\varphi, the lengths Lψ(ej)L_{\psi^{\prime}}(e^{\prime}_{j}), j=1,2,3j=1,2,3, are uniquely determined, and hence (ι)1(v4)(\iota^{\prime})^{-1}(v^{\prime}_{4}) depends only on φ\varphi. Consequently, R,ι\langle R^{\prime},\iota^{\prime}\rangle is equivalent to R,ι\langle R,\iota\rangle. ∎

Remark.

The three points qjq_{j}, j=1,2,3j=1,2,3, in the above proof are projected to the same point on RR. It is a zero of the co-welder of φ\varphi even if none of qjq_{j} is a zero of φ\varphi.

Proof of Theorem 5.9.

Let R,ι\langle R,\iota\rangle be an arbitrary CC-full self-welding of SS with welder φ\varphi. Take nonoverlapping arcs aka_{k}, k=1,,2Nk=1,\ldots,2N, on CC so that R,ι\langle R,\iota\rangle is the self-welding of SS with welder φ\varphi along (a2j1,a2j)(a_{2j-1},a_{2j}), j=1,,Nj=1,\dots,N. By subdividing the arcs if necessary we may assume that every zero of φ\varphi on CC is an endpoint of some aka_{k}. Then the closure of each horizontal trajectory of φ\varphi on CC is a union of aka_{k}’s. Denote by ψ\psi the co-welder of φ\varphi.

If φAL(S,C)\varphi\not\in A_{L}(S,C), then some component CC^{\prime} of CC includes a horizontal trajectory aa of φ\varphi with Lφ(a)>Lφ(C)/2L_{\varphi}(a)>L_{\varphi}(C^{\prime})/2. Since a2j1a_{2j-1} and a2ja_{2j} are of the same φ\varphi-length and a2ja_{2j} lies on CC^{\prime} if a2j1a_{2j-1} does, some pair, which may be assumed to be a1a_{1} and a2a_{2}, lie on aa. If they have a common endpoint on aa, then it is not a zero of φ\varphi but is mapped to an end-vertex of the weld graph GιG_{\iota}, and hence the self-welding R,ι\langle R,\iota\rangle is not φ\varphi-regular by Proposition 5.6. Otherwise, the closure of C(a1a2)C^{\prime}\setminus(a_{1}^{\circ}\cup a_{2}^{\circ}) has a component c0c_{0} included in aa. Let R1,ι1\langle R_{1},\iota_{1}\rangle be the self-welding of SS with welder φ\varphi along (a1,a2)(a_{1},a_{2}). Observe that ι1(c0)\iota_{1}(c_{0}) is a component of the border R1\partial R_{1} and contains exactly one zero of the co-welder ψ1A+(R1)\psi_{1}\in A_{+}(R_{1}) of φ\varphi. It follows from Corollary 5.7 that the self-welding R,ι\langle R,\iota^{\prime}\rangle of R1R_{1} with welder ψ1\psi_{1} along ((ι1)a2j1,(ι1)a2j)((\iota_{1})_{*}a_{2j-1},(\iota_{1})_{*}a_{2j}), 2jN2\leqq j\leqq N, is not ψ1\psi_{1}-regular. Therefore ψ\psi has a pole on RR as it is the co-welder of ψ1\psi_{1}. This means that the self-welding R,ι\langle R,\iota\rangle of SS is not φ\varphi-regular, either. We have thus proved that if some CC-full self-welding of SS is φ\varphi-regular, then φ\varphi satisfies the border length condition on CC.

To prove the converse suppose that φAL(S,C)\varphi\in A_{L}(S,C). Let CC^{\prime} be a component of CC. We will take one or three pairs of arcs of equal φ\varphi-lengths on CC^{\prime} to obtain a CC^{\prime}-full φ\varphi-regular self-welding R1,ι1\langle R_{1},\iota_{1}\rangle of SS. Then the co-welder ψ1\psi_{1} satisfies the border length condition on C1:=ι1(CC)C_{1}:=\iota_{1}(C\setminus C^{\prime}), or ψ1AL(R1,C1)\psi_{1}\in A_{L}(R_{1},C_{1}). Since the number of components of C1C_{1} is smaller by one than that of CC, repeating this process leads us to a CC-full φ\varphi-regular self-welding of SS.

Let ZZ be the set of zeros of φ\varphi on CC^{\prime}. The border length condition implies that cardZ2\operatorname{card}Z\geqq 2. If there are two points in ZZ which divide CC^{\prime} into two arcs a1a_{1} and a2a_{2} of the same φ\varphi-length, then the self-welding of SS with welder φ\varphi along (a1,a2)(a_{1},a_{2}) is φ\varphi-regular by Proposition 5.6. Otherwise, ZZ contains at least three points. Two distinct points p,qZp,q\in Z divide CC^{\prime} into two arcs. Let pq¯\overline{pq} denote the one with shorter φ\varphi-length. Thus pq¯\overline{pq} is the simple arc on CC^{\prime} joining pp and qq whose φ\varphi-length Lφ(pq¯)L_{\varphi}(\overline{pq}) is less than Lφ(C)/2L_{\varphi}(C^{\prime})/2. Choose two distinct points p1,p2Zp_{1},p_{2}\in Z so that Lφ(p1p2¯)L_{\varphi}(\overline{p_{1}p_{2}}) is the largest among Lφ(pq¯)L_{\varphi}(\overline{pq}), where p,qZp,q\in Z with pqp\neq q. The border length condition implies that ZZ contains a point p3p_{3} that does not lie on p1p2¯\overline{p_{1}p_{2}}. Since Lφ(pkpk+1¯)Lφ(p1p2¯)<Lφ(C)/2L_{\varphi}(\overline{p_{k}p_{k+1}})\leqq L_{\varphi}(\overline{p_{1}p_{2}})<L_{\varphi}(C^{\prime})/2 for k=1,2,3k=1,2,3, where p4=p1p_{4}=p_{1}, it follows from Lemma 5.11 that there is a CC-full self-welding R1,ι1\langle R_{1},\iota_{1}\rangle of SS with welder φ\varphi whose weld graph is Y-shaped with end-vertices ι1(pj)\iota_{1}(p_{j}), j=1,2,3j=1,2,3. Thus the self-welding is φ\varphi-regular by Proposition 5.6, for, the points pjp_{j}, j=1,2,3j=1,2,3, are zeros of φ\varphi. This completes the proof. ∎

Example 5.12.

We consider the case of genus one. Let R,ι\langle R,\iota\rangle be a genus-preserving closed φ\varphi-regular self-welding of SS, and let ψ\psi be the co-welder of φ\varphi. Since RR is a torus, ψ=ς2\psi=\varsigma^{2} for some holomorphic 11-form ς\varsigma on RR. Set ω=ις\omega=\iota^{*}\varsigma. Since φ=ω2\varphi=\omega^{2} is positive, the imaginary part of ω\omega vanishes along S\partial S. Moreover, ω\omega is semiexact as Cω=ιCς=0\int_{C}\omega=\int_{\iota_{*}C}\varsigma=0 for all components CC of S\partial S. Therefore, (R,ι|S)(R,\iota|_{S^{\circ}}) is a hydrodynamic continuation of SS^{\circ} with respect to ω\omega. As ς\varsigma is free of zeros, it follows from (4.1) that ω\omega has exactly two zeros on each component of S\partial S, and these points are projected to end-vertices of the weld graph GιG_{\iota} by Proposition 5.6. Since GιG_{\iota} has no other end-vertices, each component of GιG_{\iota} is I-shaped.

Let R0R_{0} be a finite open Riemann surface, and let (R˘0,ι˘0)(\breve{R}_{0},\breve{\iota}_{0}) denote a natural compact continuation of R0R_{0}. Assuming that R0R_{0} is nonanalytically finite, let (R,ι)(R,\iota) be a genus-preserving closed self-welding continuation of R0R_{0}. For φA+(R0)\varphi\in A_{+}(R_{0}) the self-welding continuation is said to be φ\varphi-regular if φ\varphi is a welder of (R,ι)(R,\iota) and the co-welder of φ\varphi is holomorphic on RR. It is called regular if it is φ\varphi-regular for some φ\varphi. By Corollary 5.10 there is a genus-preserving closed φ\varphi-regular self-welding continuation of R0R_{0} if and only if φ\varphi belongs to AL(R0):=ι˘0AL(R˘0)A_{L}(R_{0}):=\breve{\iota}_{0}^{*}A_{L}(\breve{R}_{0}), which is a proper subset of A+(R0)A_{+}(R_{0}). In the case where R0R_{0} is analytically finite, we define AL(R0)=ι˘0A+(R˘0)A_{L}(R_{0})=\breve{\iota}_{0}^{*}A_{+}(\breve{R}_{0}). In either case AL(R0)A_{L}(R_{0}) is closed in A+(R0)A_{+}(R_{0}).

As will be remarked in §8, a closed continuation (R,ι)(R,\iota) of a finite open Riemann surface R0R_{0} is a regular self-welding continuation if and only if ι\iota is a Teichmüller conformal embedding. If this is the case, then an welder of (R,ι)(R,\iota) and its co-welder are initial and terminal differentials of ι\iota, respectively.

6 An extremal property of closed regular self-welding continuations

We begin this section with introducing spaces of quadratic differentials on marked Riemann surfaces of genus gg. Let 𝑺𝔉g{\bm{S}}\in\mathfrak{F}_{g}, and consider all pairs (φ,η)(\varphi,\eta), where (S,η)(S,\eta) represents 𝑺\bm{S} and φ\varphi is a quadratic differential on SS. Two such pairs (φ1,η1)(\varphi_{1},\eta_{1}) and (φ2,η2)(\varphi_{2},\eta_{2}), where (Sj,ηj)𝑺(S_{j},\eta_{j})\in{\bm{S}}, are defined to be equivalent to each other if φ1=κφ2\varphi_{1}=\kappa^{*}\varphi_{2} for some κCHomeo(S1,S2)\kappa\in\operatorname{CHomeo}(S_{1},S_{2}) with κη1η2\kappa\circ\eta_{1}\simeq\eta_{2}. Each equivalence class 𝝋=[φ,η]{\bm{\varphi}}=[\varphi,\eta] is called a quadratic differential on 𝑺\bm{S}. If φ\varphi possesses some conformally invariant properties, then we say that 𝝋\bm{\varphi} has the same properties. For example, 𝝋\bm{\varphi} is called meromorphic if φ\varphi is. If φ\varphi is measurable on SS, then we set 𝝋𝑺=φS\|{\bm{\varphi}}\|_{\bm{S}}=\|\varphi\|_{S}. If 𝝋𝑺\|{\bm{\varphi}}\|_{\bm{S}} is finite, then 𝝋\bm{\varphi} is called integrable. Also, 𝝋\bm{\varphi} is said to coincide with 𝝍\bm{\psi} almost everywhere on 𝑺\bm{S} if φ=ψ\varphi=\psi almost everywhere on SS for some (φ,η)𝝋(\varphi,\eta)\in{\bm{\varphi}} and (ψ,η)𝝍(\psi,\eta)\in{\bm{\psi}}, where (S,η)𝑺(S,\eta)\in{\bm{S}}. If 𝑺\bm{S} is not a marked torus, then for quadratic differentials 𝝋\bm{\varphi} and 𝝍\bm{\psi} on 𝑺\bm{S} and complex numbers cc we define 𝝋+𝝍{\bm{\varphi}}+{\bm{\psi}} and c𝝋c{\bm{\varphi}} in the obvious manner:

𝝋+𝝍=[φ+ψ,η]andc𝝋=[cφ,η],{\bm{\varphi}}+{\bm{\psi}}=[\varphi+\psi,\eta]\quad\text{and}\quad c{\bm{\varphi}}=[c\varphi,\eta],

where (S,η)𝑺(S,\eta)\in{\bm{S}} and φ\varphi and ψ\psi are quadratic differentials on SS. If 𝑺\bm{S} is a marked torus, then the addition operator ++ is available only if φ\varphi or ψ\psi is invariant under conformal automorphisms homotopic to the identity, that is, φ\varphi or ψ\psi is holomorphic on SS.

Let 𝜾=[ι,η1,η2]CEmb(𝑺1,𝑺2){\bm{\iota}}=[\iota,\eta_{1},\eta_{2}]\in\operatorname{CEmb}({\bm{S}}_{1},{\bm{S}}_{2}), where 𝑺j=[Sj,ηj]{\bm{S}}_{j}=[S_{j},\eta_{j}], j=1,2j=1,2, and let 𝝋=[φ,η2]{\bm{\varphi}}=[\varphi,\eta_{2}] be a quadratic differential on 𝑺2{\bm{S}}_{2}. If 𝑺2{\bm{S}}_{2} is not a marked torus, then the pull-back 𝜾𝝋{\bm{\iota}}^{*}{\bm{\varphi}} of 𝝋\bm{\varphi} is defined by 𝜾𝝋=[ιφ,η1]{\bm{\iota}}^{*}{\bm{\varphi}}=[\iota^{*}\varphi,\eta_{1}]. This definition does not depend on a particular choice of representatives. In the case where 𝑺2{\bm{S}}_{2} is a marked torus, the pull-back operator 𝜾{\bm{\iota}}^{*} applies only to holomorphic quadratic differentials on 𝑺2{\bm{S}}_{2}.

The sets of meromorphic and holomorphic quadratic differentials on 𝑺\bm{S} are denoted by M(𝑺)M({\bm{S}}) and A(𝑺)A({\bm{S}}), respectively. We define M+(𝑺)M_{+}({\bm{S}}), A+(𝑺)A_{+}({\bm{S}}), ML(𝑺)M_{L}({\bm{S}}) and AL(𝑺)A_{L}({\bm{S}}) to be the subsets consisting of those 𝝋=[φ,η]{\bm{\varphi}}=[\varphi,\eta] with φ\varphi belonging to M+(S)M_{+}(S), A+(S)A_{+}(S), ML(S)M_{L}(S) and AL(S)A_{L}(S), respectively, provided that those spaces are meaningful. If 𝑺\bm{S} is closed, that is, if 𝑺𝔗g{\bm{S}}\in\mathfrak{T}_{g}, then A(𝑺)A({\bm{S}}) is a complex Banach space of dimension dg=max{g,3g3}d_{g}=\max\{g,3g-3\} with norm 𝑺\|\cdot\|_{\bm{S}}.

Now, let 𝑺=[S,η]\bm{S}=[S,\eta] be a marked compact bordered Riemann surface of genus gg, and let R,ι\langle R,\iota\rangle be a genus-preserving self-welding of SS. Set θ=ιη\theta=\iota\circ\eta, which is a gg-handle mark of RR as η(Σ˙g)S\eta(\dot{\Sigma}_{g})\subset S^{\circ}, to obtain 𝑹:=[R,θ]𝔉g{\bm{R}}:=[R,\theta]\in\mathfrak{F}_{g} and 𝜾:=[ι,η,θ]Conthc(𝑺,𝑹){\bm{\iota}}:=[\iota,\eta,\theta]\in\operatorname{Cont}_{\mathrm{hc}}({\bm{S}},{\bm{R}}). If R,ι\langle R^{\prime},\iota^{\prime}\rangle is a self-welding of SS equivalent to R,ι\langle R,\iota\rangle and θ\theta^{\prime} is a gg-handle mark of RR^{\prime} with ιηθ\iota^{\prime}\circ\eta\simeq\theta^{\prime}, then (R,θ)𝑹(R^{\prime},\theta^{\prime})\in{\bm{R}} and (ι,η,θ)𝜾(\iota^{\prime},\eta,\theta^{\prime})\in{\bm{\iota}}. Also, if (S,η)𝑺(S^{\prime},\eta^{\prime})\in{\bm{S}} and [κ,η,η]CHomeohc(𝑺,𝑺)[\kappa,\eta,\eta^{\prime}]\in\operatorname{CHomeo}_{\mathrm{hc}}({\bm{S}},{\bm{S}}^{\prime}), then R,ικ1\langle R,\iota\circ\kappa^{-1}\rangle is a self-welding of SS^{\prime} with κι1θη\kappa\circ\iota^{-1}\circ\theta\simeq\eta^{\prime} and (ικ1,η,θ)(\iota\circ\kappa^{-1},\eta^{\prime},\theta) represents 𝜾\bm{\iota}. Thus 𝑹,𝜾\langle{\bm{R}},{\bm{\iota}}\rangle is a well-defined pair, which we call a self-welding of 𝑺\bm{S}. If φM+(S)\varphi\in M_{+}(S) is a welder of R,ι\langle R,\iota\rangle and ψM+(R)\psi\in M_{+}(R) is its co-welder, then we call 𝝋:=[φ,η]M+(𝑺){\bm{\varphi}}:=[\varphi,\eta]\in M_{+}({\bm{S}}) and 𝝍:=[ψ,θ]M+(𝑹){\bm{\psi}}:=[\psi,\theta]\in M_{+}({\bm{R}}) a welder of 𝑹,𝜾\langle{\bm{R}},{\bm{\iota}}\rangle and the co-welder of 𝝋\bm{\varphi}, respectively. If 𝝍A+(𝑹){\bm{\psi}}\in A_{+}({\bm{R}}), then 𝑹,𝜾\langle{\bm{R}},{\bm{\iota}}\rangle is said to be 𝝋\bm{\varphi}-regular. If this is the case, then 𝝋AL(𝑺){\bm{\varphi}}\in A_{L}({\bm{S}}) and 𝝋=𝜾𝝍{\bm{\varphi}}={\bm{\iota}}^{*}{\bm{\psi}}. A self-welding 𝑹,𝜾\langle{\bm{R}},{\bm{\iota}}\rangle is called regular if it is 𝝋\bm{\varphi}-regular for some 𝝋AL(𝑹){\bm{\varphi}}\in A_{L}({\bm{R}}). Also, if R,ι\langle R,\iota\rangle has some additional properties, then we say that 𝑹,𝜾\langle{\bm{R}},{\bm{\iota}}\rangle possesses the same properties. For example, If R,ι\langle R,\iota\rangle is closed, then so is 𝑹,𝜾\langle{\bm{R}},{\bm{\iota}}\rangle.

Let us return to our investigations on 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). We are exclusively concerned with the case where 𝑹0{\bm{R}}_{0} is a marked finite open Riemann surface of positive genus gg. Let (𝑹˘0,𝜾˘0)(\breve{\bm{R}}_{0},\breve{\bm{\iota}}_{0}) denote the natural compact continuation of 𝑹0{\bm{R}}_{0}. Since punctures are removable singularities for conformal embeddings of a Riemann surface into closed Riemann surfaces, we have 𝔐((𝑹˘0))=𝔐(𝑹0)\mathfrak{M}((\breve{\bm{R}}_{0})^{\circ})=\mathfrak{M}({\bm{R}}_{0}). Without causing any trouble we sometimes assume, if necessary, that 𝑹0{\bm{R}}_{0} is the interior of a marked compact bordered Riemann surface.

Suppose that 𝑹0{\bm{R}}_{0} is nonanalytically finite. Then 𝑹˘0\breve{\bm{R}}_{0} is a marked compact bordered Riemann surface. A continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) of 𝑹0{\bm{R}}_{0} is said to be a self-welding continuation of 𝑹0{\bm{R}}_{0} if 𝜾=𝜿𝜾˘0{\bm{\iota}}={\bm{\kappa}}\circ\breve{\bm{\iota}}_{0} for some self-welding 𝑹,𝜿\langle{\bm{R}},{\bm{\kappa}}\rangle of 𝑹˘0\breve{\bm{R}}_{0}. The pull-back 𝝋:=𝜾˘0𝝋˘{\bm{\varphi}}:=\breve{\bm{\iota}}_{0}^{*}\breve{\bm{\varphi}} of a welder 𝝋˘\breve{\bm{\varphi}} of the self-welding 𝑹,𝜿\langle{\bm{R}},{\bm{\kappa}}\rangle is referred to as a welder of the self-welding continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}), and the co-welder of 𝝋˘\breve{\bm{\varphi}} is also called the co-welder of 𝝋\bm{\varphi}. If the self-welding 𝑹,𝜿\langle{\bm{R}},{\bm{\kappa}}\rangle is 𝝋˘\breve{\bm{\varphi}}-regular, then the continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) is said to be 𝝋\bm{\varphi}-regular. A self-welding continuation is called regular if it is 𝝋\bm{\varphi}-regular for some welder 𝝋\bm{\varphi}.

Proposition 6.1.

Let 𝐑0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface, and let (𝐑,𝛊)({\bm{R}},{\bm{\iota}}) be a dense compact continuation of 𝐑0{\bm{R}}_{0}. If there are 𝛗A+(𝐑0){\bm{\varphi}}\in A_{+}({\bm{R}}_{0}) and 𝛙A+(𝐑){\bm{\psi}}\in A_{+}({\bm{R}}) such that 𝛗=𝛊𝛙{\bm{\varphi}}={\bm{\iota}}^{*}{\bm{\psi}}, then (𝐑,𝛊)({\bm{R}},\bm{\iota}) is a self-welding continuation of 𝐑0{\bm{R}}_{0} with welder 𝛗{\bm{\varphi}}.

This is an immediate consequence of Corollary 4.6. Recall that the pull-back 𝜾𝝍{\bm{\iota}}^{*}{\bm{\psi}} is well-defined for 𝝍A+(𝑹){\bm{\psi}}\in A_{+}({\bm{R}}) even if 𝑹\bm{R} is a marked torus.

Let 𝑹0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface of genus gg. Note that Theorem 1.3 follows at once from Corollary 5.10. Let 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}). For 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}) let CEmb𝝋(𝑹0,𝑹)\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0},{\bm{R}}) be the set of 𝜾CEmbhc(𝑹0,𝑹){\bm{\iota}}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) such that (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) is a closed 𝝋{\bm{\varphi}}-regular self-welding continuation of 𝑹0{\bm{R}}_{0}. It may be an empty set. Let 𝔐𝝋(𝑹0)\mathfrak{M}_{\bm{\varphi}}({\bm{R}}_{0}) denote the set of 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}) for which CEmb𝝋(𝑹0,𝑹)\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0},{\bm{R}})\neq\varnothing. We are interested in the set

𝔐L(𝑹0):=𝝋AL(𝑹0)𝔐𝝋(𝑹0).\mathfrak{M}_{L}({\bm{R}}_{0}):=\bigcup_{{\bm{\varphi}}\in A_{L}({\bm{R}}_{0})}\mathfrak{M}_{\bm{\varphi}}({\bm{R}}_{0}).

If 𝑹0{\bm{R}}_{0} is analytically finite, then 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is exactly the singleton {𝑹˘0}\{\breve{\bm{R}}_{0}\}. In this case we set 𝔐L(𝑹0)=𝔐(𝑹0)\mathfrak{M}_{L}({\bm{R}}_{0})=\mathfrak{M}({\bm{R}}_{0}) for the sake of convenience.

For the investigation of 𝔐L(𝑹0)\mathfrak{M}_{L}({\bm{R}}_{0}) we recall the definition and some properties of measured foliations on surfaces. Let 𝒮(Σg)\mathscr{S}(\Sigma_{g}) denote the set of free homotopy classes of homotopically nontrivial simple loops on Σg\Sigma_{g}. The set of nonnegative functions on 𝒮(Σg)\mathscr{S}(\Sigma_{g}) is identified with the product space +𝒮(Σg)\mathbb{R}_{+}^{\mathscr{S}(\Sigma_{g})}, where +=[0,+)\mathbb{R}_{+}=[0,+\infty). We endow it with the topology of pointwise convergence. Following [4], we define (Σg)\mathscr{MF}(\Sigma_{g}) to be the closure of the set of functions of the form 𝒮(Σg)γri(α,γ)+\mathscr{S}(\Sigma_{g})\ni\gamma\mapsto ri(\alpha,\gamma)\in\mathbb{R}_{+} with r+r\in\mathbb{R}_{+} and α𝒮(Σg)\alpha\in\mathscr{S}(\Sigma_{g}), where i(α,γ)i(\alpha,\gamma) denotes the geometric intersection number of α\alpha and γ\gamma, that is, i(α,γ)i(\alpha,\gamma) is the minimum of the numbers of common points of loops in α\alpha and γ\gamma. Every element of (Σg)\mathscr{MF}(\Sigma_{g}) is called a measured foliation on Σg\Sigma_{g}. If (Σg)\mathcal{F}\in\mathscr{MF}(\Sigma_{g}) and r+r\in\mathbb{R}_{+}, then r(Σg)r\mathcal{F}\in\mathscr{MF}(\Sigma_{g}).

Important examples of measured foliations are those induced by holomorphic quadratic differentials defined as follows. Let 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g}. Choose a closed Riemann surface RR of genus gg together with θHomeo+(Σg,R)\theta\in\operatorname{Homeo}^{+}(\Sigma_{g},R) so that (R,θ˙)𝑹(R,\dot{\theta})\in{\bm{R}}, where θ˙=θ|Σ˙g\dot{\theta}=\theta|_{\dot{\Sigma}_{g}}. For measurable quadratic differentials 𝝍=[ψ,θ˙]{\bm{\psi}}=[\psi,\dot{\theta}] on 𝑹\bm{R} define a mapping 𝑹(𝝍):𝒮(Σg)+\mathcal{H}_{\bm{R}}({\bm{\psi}}):\mathscr{S}(\Sigma_{g})\to\mathbb{R}_{+} by

𝑹(𝝍)(γ)=infcγHψ(θc)=infcγθc|Imψ|.\mathcal{H}_{\bm{R}}({\bm{\psi}})(\gamma)=\inf_{c\in\gamma}H_{\psi}(\theta_{*}c)=\inf_{c\in\gamma}\int_{\theta_{*}c}|\operatorname{Im}\sqrt{\psi\,}|.

This definition does not depend on a particular choice of representatives even if g=1g=1. In the case where 𝝍A(𝑹){\bm{\psi}}\in A({\bm{R}}), the mapping 𝑹(𝝍)\mathcal{H}_{\bm{R}}({\bm{\psi}}) is a measured foliation on Σg\Sigma_{g} called the horizontal foliation of 𝝍\bm{\psi}. Note that 𝑹(r𝝍)=r𝑹(𝝍)\mathcal{H}_{\bm{R}}(r{\bm{\psi}})=\sqrt{r\,}\mathcal{H}_{\bm{R}}({\bm{\psi}}) for r+r\in\mathbb{R}_{+}.

Proposition 6.2 (Hubbard-Masur [24]).

For any 𝐑𝔗g{\bm{R}}\in\mathfrak{T}_{g} the correspondence

𝑹:𝝍𝑹(𝝍)\mathcal{H}_{\bm{R}}:{\bm{\psi}}\mapsto\mathcal{H}_{\bm{R}}({\bm{\psi}})

is a homeomorphism of A(𝐑)A({\bm{R}}) onto (Σg)\mathscr{MF}(\Sigma_{g}).

For the proof see also Gardiner [16, Theorem 6]. The inverse of the homeomorphism 𝑹\mathcal{H}_{\bm{R}} will be denoted by 𝑸𝑹:(Σg)A(𝑹){\bm{Q}}_{\bm{R}}:\mathscr{MF}(\Sigma_{g})\to A({\bm{R}}). Thus 𝑸𝑹(){\bm{Q}}_{\bm{R}}(\mathcal{F}) stands for the holomorphic quadratic differential on 𝑹\bm{R} such that 𝑹(𝑸𝑹())=\mathcal{H}_{\bm{R}}({\bm{Q}}_{\bm{R}}(\mathcal{F}))=\mathcal{F}. In fact, the correspondence (𝑹,)𝑸𝑹()({\bm{R}},\mathcal{F})\mapsto{\bm{Q}}_{\bm{R}}(\mathcal{F}) defines a homeomorphism of 𝔗g×(Σg)\mathfrak{T}_{g}\times\mathscr{MF}(\Sigma_{g}) onto the complex vector bundle of holomorphic quadratic differentials over 𝔗g\mathfrak{T}_{g}. The bundle is canonically identified with the cotangent bundle T𝔗gT^{*}\mathfrak{T}_{g} of the Teichmüller space through the bilinear form (μ,ψ)ReRμψ(\mu,\psi)\mapsto\operatorname{Re}\int_{R}\mu\psi on the space of pairs of bounded measurable (1,1)(-1,1)-forms μ\mu on RR and holomorphic quadratic differentials ψ\psi on RR. Note that 𝑸𝑹(r)=r2𝑸𝑹(){\bm{Q}}_{\bm{R}}(r\mathcal{F})=r^{2}{\bm{Q}}_{\bm{R}}(\mathcal{F}) for r+r\in\mathbb{R}_{+}.

For (Σg)\mathcal{F}\in\mathscr{MF}(\Sigma_{g}) and 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g} the extremal length Ext(𝑹,)\operatorname{Ext}({\bm{R}},\mathcal{F}) of \mathcal{F} on 𝑹\bm{R} is defined by

Ext(𝑹,)=𝑸𝑹()𝑹.\operatorname{Ext}({\bm{R}},\mathcal{F})=\|{\bm{Q}}_{\bm{R}}(\mathcal{F})\|_{\bm{R}}.

Set Ext(𝑹)=Ext(𝑹,)\operatorname{Ext}_{\mathcal{F}}({\bm{R}})=\operatorname{Ext}({\bm{R}},\mathcal{F}) to obtain a nonnegative function Ext\operatorname{Ext}_{\mathcal{F}} on 𝔗g\mathfrak{T}_{g}.

Theorem 6.3.

Let 𝐑0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface. Let (𝐑,𝛊)({\bm{R}},{\bm{\iota}}) be a closed 𝛗\bm{\varphi}-regular self-welding continuation of 𝐑0{\bm{R}}_{0}, and let 𝛙\bm{\psi} be the co-welder of 𝛗\bm{\varphi}. Set =𝐑(𝛙)\mathcal{F}=\mathcal{H}_{\bm{R}}({\bm{\psi}}). Then

Ext(𝑹)Ext(𝑹)\operatorname{Ext}_{\mathcal{F}}({\bm{R}}^{\prime})\leqq\operatorname{Ext}_{\mathcal{F}}({\bm{R}})

for all 𝐑𝔐(𝐑0){\bm{R}}^{\prime}\in\mathfrak{M}({\bm{R}}_{0}).

In other words, the function Ext\operatorname{Ext}_{\mathcal{F}} attains its maximum on 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) at 𝑹\bm{R}. To prove Theorem 6.3 we first show the following proposition and lemma. The lemma will be also applied when we investigate the maximal sets for measured foliations on 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) as well as uniqueness of conformal embeddings (see §§9 and 12).

In general, let 𝝋=[φ,θ˙]{\bm{\varphi}}=[\varphi,\dot{\theta}] be a measurable quadratic differential on 𝑹{\bm{R}}. Define a mapping 𝑹(𝝋)\mathcal{H}^{\prime}_{\bm{R}}({\bm{\varphi}}) of 𝒮(Σg)\mathscr{S}(\Sigma_{g}) into +\mathbb{R}_{+} by 𝑹(𝝋)(γ)=infcHφ(θc)\mathcal{H}^{\prime}_{\bm{R}}({\bm{\varphi}})(\gamma)=\inf_{c}H_{\varphi}(\theta_{*}c), where the infimum is taken over all cγc\in\gamma for which θc\theta_{*}c is a piecewise analytic simple loop on RR.

Proposition 6.4.

Let 𝐑𝔗g{\bm{R}}\in\mathfrak{T}_{g} and 𝛙A(𝐑){\bm{\psi}}\in A({\bm{R}}), and let 𝛗\bm{\varphi} be an integrable quadratic differential on 𝐑\bm{R}. If 𝐑(𝛙)(γ)𝐑(𝛗)(γ)\mathcal{H}_{\bm{R}}({\bm{\psi}})(\gamma)\leqq\mathcal{H}^{\prime}_{\bm{R}}({\bm{\varphi}})(\gamma) for all γ𝒮(Σg)\gamma\in\mathscr{S}(\Sigma_{g}), then

𝝍𝑹𝝋𝑹.\|{\bm{\psi}}\|_{\bm{R}}\leqq\|{\bm{\varphi}}\|_{\bm{R}}.

The sign of equality occurs if and only if 𝛗=𝛙{\bm{\varphi}}={\bm{\psi}} almost everywhere on 𝐑\bm{R}.

Proof.

Let (R,θ˙)𝑹(R,\dot{\theta})\in{\bm{R}} and (φ,θ˙)𝝋(\varphi,\dot{\theta})\in{\bm{\varphi}}. Though φ\varphi is not continuous, we can apply the arguments in the proof of the second minimal norm property [17, Theorem 9 in §2.6]. In fact, to estimate the integrals over spiral domains DD of ψ\psi, where (ψ,θ˙)𝝍(\psi,\dot{\theta})\in{\bm{\psi}}, we choose a horizontal arc aa of ψ\psi in DD so that |φ|\sqrt{|\varphi|\,} is integrable on aa as in [18, Lemma 4 in §12.7], where the roles of ψ\psi and φ\varphi are interchanged and the letter α\alpha is used for aa. Then the reasoning in [17] works without any further modifications. ∎

Let SS^{\prime} be a subsurface of a Riemann surface SS, and let φ\varphi^{\prime} be a quadratic differential on SS^{\prime}. By the zero-extension of φ\varphi^{\prime} to SS we mean the quadratic differential φ\varphi on SS defined by φ=φ\varphi=\varphi^{\prime} on SS^{\prime} and φ=0\varphi=0 on SSS^{\prime}\setminus S. If (𝑺,𝜾)=([S,η],[ι,η,η])({\bm{S}},{\bm{\iota}})=([S,\eta],[\iota,\eta^{\prime},\eta]) is a continuation of 𝑺=[S,η]{\bm{S}}^{\prime}=[S^{\prime},\eta^{\prime}] and 𝝋=[φ,η]{\bm{\varphi}}^{\prime}=[\varphi^{\prime},\eta^{\prime}] is a quadratic differential on 𝑺{\bm{S}}^{\prime}, then the (𝑺,𝜾)({\bm{S}},{\bm{\iota}})-zero-extension of 𝝋{\bm{\varphi}}^{\prime} means the quadratic differential [φ,η][\varphi,\eta] on 𝑺\bm{S}, where φ\varphi is the zero-extension of the quadratic differential ιφ\iota_{*}\varphi^{\prime} on ι(S)\iota(S^{\prime}) to SS.

Lemma 6.5.

Let 𝐑0{\bm{R}}_{0}, (𝐑,𝛊)({\bm{R}},{\bm{\iota}}), 𝛗\bm{\varphi} and \mathcal{F} be as in Theorem 6.3, and let (𝐑,𝛊)({\bm{R}}^{\prime},{\bm{\iota}}^{\prime}) be a closed continuation of 𝐑0{\bm{R}}_{0}. If 𝛗{\bm{\varphi}}^{\prime} is the (𝐑,𝛊)({\bm{R}}^{\prime},{\bm{\iota}}^{\prime})-zero-extension of 𝛗\bm{\varphi}, then for any γ𝒮(Σg)\gamma\in\mathscr{S}(\Sigma_{g}) the inequality

(6.1) 𝑹(𝝍)(γ)𝑹(𝝋)(γ)\mathcal{H}_{{\bm{R}}^{\prime}}({\bm{\psi}}^{\prime})(\gamma)\leqq\mathcal{H}^{\prime}_{{\bm{R}}^{\prime}}({\bm{\varphi}^{\prime}})(\gamma)

holds, where 𝛙=𝐐𝐑(){\bm{\psi}}^{\prime}={\bm{Q}}_{{\bm{R}}^{\prime}}(\mathcal{F}).

Proof.

Take representatives (R0,θ0)(R_{0},\theta_{0}), (R,θ˙)(R,\dot{\theta}), (ι,θ0,θ˙)(\iota,\theta_{0},\dot{\theta}) and (φ,θ0)(\varphi,\theta_{0}) of 𝑹0{\bm{R}}_{0}, 𝑹\bm{R}, 𝜾\bm{\iota} and 𝝋\bm{\varphi}, respectively, where θ˙\dot{\theta} stands for the restriction of θHomeo+(Σg,R)\theta\in\operatorname{Homeo}^{+}(\Sigma_{g},R) to Σ˙g\dot{\Sigma}_{g}. Let (𝑹˘0,𝜾˘0)=([R˘0,θ˘0],[ι˘0,θ0,θ˘0])(\breve{\bm{R}}_{0},\breve{\bm{\iota}}_{0})=([\breve{R}_{0},\breve{\theta}_{0}],[\breve{\iota}_{0},\theta_{0},\breve{\theta}_{0}]) be the natural compact continuation of 𝑹0{\bm{R}}_{0}. There is a closed 𝝋˘\breve{\bm{\varphi}}-regular self-welding 𝑹,𝜾˘\langle{\bm{R}},\breve{\bm{\iota}}\rangle of 𝑹˘0\breve{\bm{R}}_{0} such that 𝜾=𝜾˘𝜾˘0{\bm{\iota}}=\breve{\bm{\iota}}\circ\breve{\bm{\iota}}_{0} and 𝝋=𝜾˘0𝝋˘{\bm{\varphi}}=\breve{\bm{\iota}}_{0}^{*}\breve{\bm{\varphi}} with 𝝋˘=[φ˘,θ˘0]AL(𝑹˘0)\breve{\bm{\varphi}}=[\breve{\varphi},\breve{\theta}_{0}]\in A_{L}(\breve{\bm{R}}_{0}). We may assume that R0R_{0} has no punctures so that ι˘0(R0)=(R˘0)\breve{\iota}_{0}(R_{0})=(\breve{R}_{0})^{\circ}. Let Φ˘\breve{\Phi} be the integral of φ˘\sqrt{\breve{\varphi}\,}, which is a multi-valued function and may have finitely many singular points on R˘0\breve{R}_{0}. Let C˘1,,C˘n0\breve{C}_{1},\ldots,\breve{C}_{n_{0}} be the components of R˘0\partial\breve{R}_{0}. For each C˘k\breve{C}_{k} choose a doubly connected closed neighborhood U˘k\breve{U}_{k} in R˘0\breve{R}_{0} of C˘k\breve{C}_{k} so that

  • (i)

    U˘1,,U˘n0\breve{U}_{1},\ldots,\breve{U}_{n_{0}} are mutually disjoint,

  • (ii)

    each U˘k\breve{U}_{k} is divided into finitely many simply connected closed domains D˘1(k),,D˘mk(k)\breve{D}_{1}^{(k)},\ldots,\breve{D}_{m_{k}}^{(k)},

  • (iii)

    a branch of Φ˘\breve{\Phi} maps each D˘j(k)\breve{D}_{j}^{(k)} homeomorphically onto a closed rectangle in \mathbb{C} with sides parallel to the real and imaginary axes, and

  • (iv)

    the vertical sides of the rectangles Φ˘(D˘j(k))\breve{\Phi}(\breve{D}_{j}^{(k)}), 1jmk1\leqq j\leqq m_{k}, 1kn01\leqq k\leqq n_{0}, are of the same length.

Note that (R˘0)U˘k(\breve{R}_{0})^{\circ}\cap\partial\breve{U}_{k} is composed of finitely many horizontal and vertical arcs of φ˘\breve{\varphi}. Set D˘0=R˘0kU˘k\breve{D}_{0}=\breve{R}_{0}\setminus\bigcup_{k}\breve{U}_{k} and D=ι˘(D˘0)D=\breve{\iota}(\breve{D}_{0}), where 𝜾˘=[ι˘,θ˘0,θ]\breve{\bm{\iota}}=[\breve{\iota},\breve{\theta}_{0},\theta]. Observe that ι˘(U˘k)\breve{\iota}(\breve{U}_{k}) is a component of RDR\setminus D, which is a topological closed disk on RR, even though ι˘\breve{\iota} is not injective on U˘k\breve{U}_{k}. Also, each ι˘(D˘j(k))\breve{\iota}(\breve{D}_{j}^{(k)}) meets the weld graph Gι˘G_{\breve{\iota}} since D˘j(k)C˘k\breve{D}_{j}^{(k)}\cap\breve{C}_{k}\neq\varnothing.

Choose a closed Riemann surface RR^{\prime} of genus gg together with θHomeo+(Σg,R)\theta^{\prime}\in\operatorname{Homeo}^{+}(\Sigma_{g},R^{\prime}) so that (R,θ˙)𝑹(R^{\prime},\dot{\theta}^{\prime})\in{\bm{R}}^{\prime}, where θ˙=θ|Σ˙g\dot{\theta}^{\prime}=\theta^{\prime}|_{\dot{\Sigma}_{g}}, and take a representative (ι,θ0,θ˙)𝜾(\iota^{\prime},\theta_{0},\dot{\theta}^{\prime})\in{\bm{\iota}}^{\prime}. Let γ\gamma be an arbitrary element of 𝒮(Σg)\mathscr{S}(\Sigma_{g}), and let cθγc^{\prime}\in\theta^{\prime}_{*}\gamma be a piecewise analytic simple loop on RR^{\prime}. We may assume that the initial (and terminal) point of cc^{\prime} is in the domain D:=ιι˘01(D˘0)D^{\prime}:=\iota^{\prime}\circ\breve{\iota}_{0}^{-1}(\breve{D}_{0}). Divide cc^{\prime} into subarcs to obtain c=c1d1cm1dm1cmc^{\prime}=c^{\prime}_{1}d^{\prime}_{1}\cdots c^{\prime}_{m-1}d^{\prime}_{m-1}c^{\prime}_{m} so that c1,,cmc^{\prime}_{1},\ldots,c^{\prime}_{m} lie in D¯\bar{D}^{\prime} while d1,,dm1d^{\prime}_{1},\ldots,d^{\prime}_{m-1} lie in RDR^{\prime}\setminus D^{\prime}. For j=1,,mj=1,\ldots,m let cjc_{j} be the image arc of cjc^{\prime}_{j} by ι(ι)1\iota\circ(\iota^{\prime})^{-1}. They are piecewise analytic simple arcs on D¯\bar{D}. Note that the terminal point qjq_{j} of cjc_{j} and the initial point pj+1p_{j+1} of cj+1c_{j+1} lie on the same component, say ι˘(U˘k)\breve{\iota}(\breve{U}_{k}), of RDR\setminus D. We choose a piecewise analytic simple arc djd_{j} joining qjq_{j} with pj+1p_{j+1} within ι˘(U˘k)\breve{\iota}(\breve{U}_{k}) as follows. Let 𝝍=[ψ,θ˙]A(𝑹){\bm{\psi}}=[\psi,\dot{\theta}]\in A({\bm{R}}) be the co-welder of 𝝋˘\breve{\bm{\varphi}}, and let ε>0\varepsilon>0. Take D˘ν(k)\breve{D}_{\nu}^{(k)} and D˘μ(k)\breve{D}_{\mu}^{(k)} so that qjq_{j} and pj+1p_{j+1} belong to ι˘(D˘ν(k))\breve{\iota}(\partial\breve{D}_{\nu}^{(k)}) and ι˘(D˘μ(k))\breve{\iota}(\partial\breve{D}_{\mu}^{(k)}), respectively. We then let dj′′d^{\prime\prime}_{j} be a simple arc on ι˘(U˘k)\breve{\iota}(\breve{U}_{k}) joining qjq_{j} with pj+1p_{j+1} composed of horizontal and vertical arcs, and possibly, zeros of ψ\psi, where the horizontal and vertical arcs should lie on ι˘(D˘ν(k)D˘μ(k))Gι˘\breve{\iota}(\partial\breve{D}_{\nu}^{(k)}\cup\partial\breve{D}_{\mu}^{(k)})\cup G_{\breve{\iota}} and ι˘(D˘ν(k)D˘μ(k))\breve{\iota}(\breve{D}_{\nu}^{(k)}\cup\breve{D}_{\mu}^{(k)}), respectively. We require that each dj′′d^{\prime\prime}_{j} should include at most two vertical arcs, which implies Hψ(dj′′)Hφ(dj)H_{\psi}(d^{\prime\prime}_{j})\leqq H_{\varphi^{\prime}}(d^{\prime}_{j}). Some of dj′′d^{\prime\prime}_{j}’s may have common points. We modify them by slightly shifting their horizontal and vertical arcs and going around the zeros of ψ\psi to obtain simple arcs djd_{j} without changing the endpoints so that Hψ(dj)Hψ(dj′′)+ε/mH_{\psi}(d_{j})\leqq H_{\psi}(d^{\prime\prime}_{j})+\varepsilon/m and that d1,,dm1d_{1},\ldots,d_{m-1} are mutually disjoint. The simple loop c:=c1d1c2d2dm1cmc:=c_{1}d_{1}c_{2}d_{2}\cdots d_{m-1}c_{m} belongs to the homotopy class θγ\theta_{*}\gamma, and satisfies Hψ(c)Hφ(c)+εH_{\psi}(c)\leqq H_{\varphi^{\prime}}(c^{\prime})+\varepsilon and hence 𝑹(𝝍)(γ)Hφ(c)+ε\mathcal{H}_{\bm{R}}({\bm{\psi}})(\gamma)\leqq H_{\varphi^{\prime}}(c^{\prime})+\varepsilon, which leads us to 𝑹(𝝍)(γ)𝑹(𝝋)(γ)\mathcal{H}_{\bm{R}}({\bm{\psi}})(\gamma)\leqq\mathcal{H}^{\prime}_{{\bm{R}}^{\prime}}({\bm{\varphi}^{\prime}})(\gamma). Since 𝑹(𝝍)(γ)=(γ)=𝑹(𝝍)(γ)\mathcal{H}_{{\bm{R}}^{\prime}}({\bm{\psi}}^{\prime})(\gamma)=\mathcal{F}(\gamma)=\mathcal{H}_{\bm{R}}({\bm{\psi}})(\gamma), we obtain (6.1). ∎

Proof of Theorem 6.3.

Let (𝑹,𝜾)({\bm{R}}^{\prime},{\bm{\iota}}^{\prime}) be an arbitrary closed continuation of 𝑹0{\bm{R}}_{0}, and let 𝝋{\bm{\varphi}}^{\prime} be the (𝑹,𝜾)({\bm{R}}^{\prime},{\bm{\iota}}^{\prime})-zero-extension of 𝝋\bm{\varphi}. Set 𝝍=𝑸𝑹(){\bm{\psi}}^{\prime}={\bm{Q}}_{{\bm{R}}^{\prime}}(\mathcal{F}). We apply Lemma 6.5 and Proposition 6.4 to obtain 𝝍𝑹𝝋𝑹=𝝋𝑹0=𝝍𝑹\|{\bm{\psi}}^{\prime}\|_{{\bm{R}}^{\prime}}\leqq\|{\bm{\varphi}}^{\prime}\|_{{\bm{R}}^{\prime}}=\|{\bm{\varphi}}\|_{{\bm{R}}_{0}}=\|{\bm{\psi}}\|_{\bm{R}}. This completes the proof as Ext(𝑹)=𝝍𝑹\operatorname{Ext}_{\mathcal{F}}({\bm{R}}^{\prime})=\|{\bm{\psi}}^{\prime}\|_{{\bm{R}}^{\prime}} and Ext(𝑹)=𝝍𝑹\operatorname{Ext}_{\mathcal{F}}({\bm{R}})=\|{\bm{\psi}}\|_{\bm{R}}. ∎

Let 𝑹j{\bm{R}}_{j}, j=1,2j=1,2, be distinct points of 𝔗g\mathfrak{T}_{g}. Then dT(𝑹1,𝑹2)=(logK(𝒉))/2d_{T}({\bm{R}}_{1},{\bm{R}}_{2})=(\log K({\bm{h}}))/2, where 𝒉\bm{h} is the Teichmüller quasiconformal homeomorphism of 𝑹1{\bm{R}}_{1} onto 𝑹2{\bm{R}}_{2}. Thus 𝒉\bm{h} belongs to QCHomeohc(𝑹1,𝑹2)\operatorname{QCHomeo}_{\mathrm{hc}}({\bm{R}}_{1},{\bm{R}}_{2}) and there are nonzero 𝝍jA(𝑹j){\bm{\psi}}_{j}\in A({\bm{R}}_{j}), j=1,2j=1,2, such that for some (h,θ1,θ2)𝒉(h,\theta_{1},\theta_{2})\in{\bm{h}} and (ψj,θj)𝝍j(\psi_{j},\theta_{j})\in{\bm{\psi}}_{j} with (Rj,θj)𝑹j(R_{j},\theta_{j})\in{\bm{R}}_{j}

  • (i)

    the Beltrami differential of hh is exactly k|ψ1|/ψ1k|\psi_{1}|/\psi_{1}, where k=(K(𝒉)1)/(K(𝒉)+1)k=(K({\bm{h}})-1)/(K({\bm{h}})+1),

  • (ii)

    hh maps every noncritical point of ψ1\psi_{1} to a noncritical point of ψ2\psi_{2}, and

  • (iii)

    hh is represented as

    ζ2=K(𝒉)Reζ1+iImζ1=(K(𝒉)+1)ζ1+(K(𝒉)1)ζ¯12\zeta_{2}=K({\bm{h}})\operatorname{Re}\zeta_{1}+i\operatorname{Im}\zeta_{1}=\frac{\,(K({\bm{h}})+1)\zeta_{1}+(K({\bm{h}})-1)\bar{\zeta}_{1}\,}{2}

    with respect to some natural parameter ζ1\zeta_{1} (resp. ζ2\zeta_{2}) of ψ1\psi_{1} (resp. ψ2\psi_{2}) around any noncritical point pp (resp. h(p)h(p)).

In other words, hh is a uniform stretching along horizontal trajectories of ψ1\psi_{1}. The quadratic differentials 𝝍1{\bm{\psi}}_{1} and 𝝍2{\bm{\psi}}_{2} are called initial and terminal quadratic differentials of 𝒉\bm{h}, respectively. Note that 𝒉1{\bm{h}}^{-1} is also the Teichmüller quasiconformal homeomorphism of 𝑹2{\bm{R}}_{2} onto 𝑹1{\bm{R}}_{1}, whose initial and terminal quadratic differentials are 𝝍2-{\bm{\psi}}_{2} and K2𝝍1-K^{2}{\bm{\psi}}_{1}, respectively.

Let 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g} and 𝝍A(𝑹){𝟎}{\bm{\psi}}\in A({\bm{R}})\setminus\{{\bm{0}}\}. For each t>0t>0 there uniquely exists 𝒓(t)𝔗g{\bm{r}}(t)\in\mathfrak{T}_{g} with dT(𝒓(t),𝑹)=td_{T}({\bm{r}}(t),{\bm{R}})=t such that 𝝍\bm{\psi} is an initial quadratic differential of the Teichmüller quasiconformal homeomorphism of 𝑹\bm{R} onto 𝒓(t){\bm{r}}(t). Set 𝒓(0)=𝑹{\bm{r}}(0)={\bm{R}}. Then the mapping 𝒓:+𝔗g{\bm{r}}:\mathbb{R}_{+}\to\mathfrak{T}_{g} is a (simple) ray emanating from 𝑹\bm{R}, called a Teichmüller geodesic ray. Note that

(6.2) dT(𝒓(t1),𝒓(t2))=|t1t2|d_{T}({\bm{r}}(t_{1}),{\bm{r}}(t_{2}))=|t_{1}-t_{2}|

for t1,t2+t_{1},t_{2}\in\mathbb{R}_{+}. For t1,t2[0,+]t_{1},t_{2}\in[0,+\infty] with t1<t2t_{1}<t_{2} we denote by 𝒓(t1,t2){\bm{r}}(t_{1},t_{2}) the image of the interval (t1,t2)(t_{1},t_{2}) by 𝒓\bm{r}:

(6.3) 𝒓(t1,t2)={𝒓(t)t1<t<t2}.{\bm{r}}(t_{1},t_{2})=\{{\bm{r}}(t)\mid t_{1}<t<t_{2}\}.

If we need to refer to the initial point 𝑹\bm{R} and the quadratic differential 𝝍\bm{\psi} we use the notation 𝒓𝑹[𝝍]{\bm{r}}_{\bm{R}}[{\bm{\psi}}] for 𝒓\bm{r}. Thus 𝝍\bm{\psi} is an initial quadratic differential of the Teichmüller quasiconformal homeomorphism 𝒉t{\bm{h}}_{t} of 𝑹\bm{R} onto 𝒓𝑹[𝝍](t){\bm{r}}_{\bm{R}}[{\bm{\psi}}](t). Let 𝝍t{\bm{\psi}}_{t} be the corresponding terminal quadratic differential of 𝒉t{\bm{h}}_{t}. If we set =𝑹(𝝍)\mathcal{F}=\mathcal{H}_{{\bm{R}}}({\bm{\psi}}), then we have 𝑸𝒓𝑹[𝝍](t)()=𝝍t{\bm{Q}}_{{\bm{r}}_{\bm{R}}[{\bm{\psi}}](t)}(\mathcal{F})={\bm{\psi}}_{t} as 𝒉t{\bm{h}}_{t} is homotopically consistent. It follows that

(6.4) Ext(𝒓𝑹[𝝍](t))=e2tExt(𝑹),=𝑹(𝝍).\operatorname{Ext}_{\mathcal{F}}({\bm{r}}_{\bm{R}}[{\bm{\psi}}](t))=e^{2t}\operatorname{Ext}_{\mathcal{F}}({\bm{R}}),\quad\mathcal{F}=\mathcal{H}_{{\bm{R}}}({\bm{\psi}}).

A Teichmüller geodesic ray is uniquely determined by the initial point 𝑹\bm{R} and a point 𝑹𝑹{\bm{R}}^{\prime}\neq{\bm{R}} that the ray passes through. We write 𝒓𝑹[𝑹]{\bm{r}}_{\bm{R}}[{\bm{R}}^{\prime}] to denote such a ray. Note that we parametrize each Teichmüller geodesic ray with respect to the distance from its initial point. Therefore, 𝒓𝑹[r𝝍]=𝒓𝑹[𝝍]{\bm{r}}_{\bm{R}}[r{\bm{\psi}}]={\bm{r}}_{\bm{R}}[{\bm{\psi}}] for r>0r>0.

As an application of Theorem 6.3 we show the following proposition. It proves a half of Theorem 1.4 (i).

Proposition 6.6.

Let 𝐑0{\bm{R}}_{0} be a marked finite open Riemann surface. Then 𝔐L(𝐑0)\mathfrak{M}_{L}({\bm{R}}_{0}) is included in the boundary 𝔐(𝐑0)\partial\mathfrak{M}({\bm{R}}_{0}).

Proof.

We have only to consider the case where 𝑹0{\bm{R}}_{0} is nonanalytically finite. Let (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) be a closed 𝝋\bm{\varphi}-regular self-welding continuation of 𝑹0{\bm{R}}_{0}, and denote by 𝝍\bm{\psi} the co-welder of 𝝋\bm{\varphi}. We know that 𝑹\bm{R} belongs to 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). If =𝑹(𝝍)\mathcal{F}=\mathcal{H}_{\bm{R}}({\bm{\psi}}), then it follows from (6.4) that Ext(𝒓𝑹[𝝍](t))=e2tExt(𝑹)>Ext(𝑹)\operatorname{Ext}_{\mathcal{F}}({\bm{r}}_{\bm{R}}[{\bm{\psi}}](t))=e^{2t}\operatorname{Ext}_{\mathcal{F}}({\bm{R}})>\operatorname{Ext}_{\mathcal{F}}({\bm{R}}) for t>0t>0. Theorem 6.3 then assures us that 𝒓𝑹[𝝍](t){\bm{r}}_{\bm{R}}[{\bm{\psi}}](t) lies outside of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). Since 𝒓𝑹[𝝍](t){\bm{r}}_{\bm{R}}[{\bm{\psi}}](t) tends to 𝑹\bm{R} as t0t\to 0, we conclude that 𝑹\bm{R} is certainly on the boundary of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). ∎

Remark.

We could apply Kahn-Pilgrim-Thurston [30] to prove Proposition 6.6. Let 𝑹\bm{R}, 𝝋\bm{\varphi} and 𝝍\bm{\psi} be as in the proof of the proposition. Approximating 𝝋\bm{\varphi} with Jenkins-Strebel quadratic differentials in A+(𝑹0)A_{+}({\bm{R}}_{0}), we know that the stretch factor of homotopically consistent topological embeddings of 𝑹0{\bm{R}}_{0} into 𝒓𝑹[𝝍](t){\bm{r}}_{\bm{R}}[{\bm{\psi}}](t) is exactly e2te^{2t}, which implies that 𝒓𝑹[𝝍](t)𝔐(𝑹0){\bm{r}}_{\bm{R}}[{\bm{\psi}}](t)\not\in\mathfrak{M}({\bm{R}}_{0}) for t>0t>0 by [30, Theorem 1].

Example 6.7.

We consider the case of genus one, and use the notations in Examples 2.3 and 3.6. The holomorphic quadratic differential dz2dz^{2} on \mathbb{C} is projected to a holomorphic quadratic differential ψτ\psi_{\tau} on TτT_{\tau} through the natural projection Πτ:Tτ=/Γτ\Pi_{\tau}:\mathbb{C}\to T_{\tau}=\mathbb{C}/\Gamma_{\tau}. Set 𝝍τ=[ψτ,η˙τ]A(𝑻τ){\bm{\psi}}_{\tau}=[\psi_{\tau},\dot{\eta}_{\tau}]\in A({\bm{T}}_{\tau}). Recall that Σ1=T1\Sigma_{1}=T_{\sqrt{-1}}. If =𝑻1(𝝍1)(Σ1)\mathcal{F}=\mathcal{H}_{{\bm{T}}_{\sqrt{-1}}}({\bm{\psi}}_{\sqrt{-1}})\in\mathscr{MF}(\Sigma_{1}), then 𝑸𝑻τ()=𝝍τ/(Imτ)2{\bm{Q}}_{{\bm{T}}_{\tau}}(\mathcal{F})={\bm{\psi}}_{\tau}/(\operatorname{Im}\tau)^{2} so that Ext(𝑻τ)=1/Imτ\operatorname{Ext}_{\mathcal{F}}({\bm{T}}_{\tau})=1/\operatorname{Im}\tau. Now, let 𝑹0=[R0,θ0]{\bm{R}}_{0}=[R_{0},\theta_{0}] be a marked finite open Riemann surface of genus one, and let (𝑹˘0,𝜾˘0)(\breve{\bm{R}}_{0},\breve{\bm{\iota}}_{0}) be the natural compact continuation. Let (R˘0,θ˘0)𝑹˘0(\breve{R}_{0},\breve{\theta}_{0})\in\breve{\bm{R}}_{0} and A0=(θ˘0)A1A_{0}=(\breve{\theta}_{0})_{*}A_{\sqrt{-1}}. For tt\in\mathbb{R} there exists a (unique) holomorphic semiexact 1-form ωt\omega_{t} on R˘0\breve{R}_{0} with A0ωt=1\int_{A_{0}}\omega_{t}=1 such that the imaginary part of eitπ/2ωte^{-it\pi/2}\omega_{t} vanishes along the border R˘0\partial\breve{R}_{0}. Set 𝝋t=[eitπωt2,θ˘0]{\bm{\varphi}}_{t}=[e^{-it\pi}\omega_{t}^{2},\breve{\theta}_{0}], which is an element of AL(𝑹˘0)A_{L}(\breve{\bm{R}}_{0}). Let 𝑻τ(t),𝜾t\langle{\bm{T}}_{\tau(t)},{\bm{\iota}}_{t}\rangle be a closed 𝝋t{\bm{\varphi}}_{t}-regular self-welding of 𝑹˘0\breve{\bm{R}}_{0}. Then 𝝍τ(t){\bm{\psi}}_{\tau(t)} is the co-welder of the welder 𝝋t{\bm{\varphi}}_{t}. Restricting ourselves to the case t=0t=0, we conclude from Theorem 6.3 that ImτImτ(0)\operatorname{Im}\tau\geqq\operatorname{Im}\tau(0) if 𝑻τ𝔐(𝑹0){\bm{T}}_{\tau}\in\mathfrak{M}({\bm{R}}_{0}). This is nothing but the first inequality of [56, Theorem 2 (I)]. We remark that Ext(𝑻τ(0))\operatorname{Ext}_{\mathcal{F}}({\bm{T}}_{\tau(0)}) is equal to the extremal length of the weak homology class of (θ0)A1(\theta_{0})_{*}A_{\sqrt{-1}} by [34, Proposition 1]. Thus we have obtained an alternative proof of [35, Lemma 1].

7 Sequences of continuations

The present section is of preparatory character. In this article we need to consider sequences of continuations of different Riemann surfaces. To deal with their convergence properties we pass to universal covering Riemann surfaces. As an application, we show, in the next section, that the boundary points obtained through closed regular self-welding continuations of 𝑹0{\bm{R}}_{0} actually exhaust 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}).

Let χ\chi be a 11-handle mark of a Riemann surface SS of positive genus, and let Πχ:𝕌SS\Pi_{\chi}:\mathbb{U}_{S}\to S be a holomorphic universal covering map with covering transformation group Γχ:=Aut(Πχ)\Gamma_{\chi}:=\operatorname{Aut}(\Pi_{\chi}). We normalize Πχ\Pi_{\chi} and Γχ\Gamma_{\chi} as follows. Set A=χA1A=\chi_{*}A_{\sqrt{-1}} and B=χB1B=\chi_{*}B_{\sqrt{-1}}.

If SS is not a torus, then we may assume that 𝕌S¯\mathbb{H}\subset\mathbb{U}_{S}\subset\bar{\mathbb{H}}. Then Γχ\Gamma_{\chi} is a torsion-free Fuchsian group keeping 𝕌S\mathbb{U}_{S} invariant. We can choose Πχ\Pi_{\chi} and Γχ\Gamma_{\chi} so that Γχ\Gamma_{\chi} contains hyperbolic transformations γχ(A)\gamma_{\chi}^{(A)} and γχ(B)\gamma_{\chi}^{(B)} such that the repelling and attracting fixed points of γχ(A)\gamma_{\chi}^{(A)} (resp. γχ(B)\gamma_{\chi}^{(B)}) are 0 and \infty (resp. 11 and some negative real number), respectively, and that the axes of γχ(A)\gamma_{\chi}^{(A)} and γχ(B)\gamma_{\chi}^{(B)} oriented from the repelling fixed points to the attracting fixed points are projected onto closed hyperbolic geodesics AχA_{\chi} and BχB_{\chi} on SS^{\circ} freely homotopic to AA and BB, respectively.

If SS is a torus, then we may set 𝕌S=\mathbb{U}_{S}=\mathbb{C}. Let ωχ\omega_{\chi} be the unique holomorphic 1-form on SS for which Aωχ=1\int_{A}\omega_{\chi}=1. Choose Πχ\Pi_{\chi} so that Πχωχ=dz\Pi_{\chi}^{*}\omega_{\chi}=dz. If we set τ=Bωχ\tau=\int_{B}\omega_{\chi}, then, using the notations in Example 2.3, we have S=TτS=T_{\tau} and Γχ=Γτ\Gamma_{\chi}=\Gamma_{\tau}. Denote by γχ(A)\gamma_{\chi}^{(A)} and γχ(B)\gamma_{\chi}^{(B)} the transformations in Γχ\Gamma_{\chi} defined by γχ(A):zz+1\gamma_{\chi}^{(A)}:z\mapsto z+1 and γχ(B):zz+τ\gamma_{\chi}^{(B)}:z\mapsto z+\tau.

In either case, Γχ\Gamma_{\chi} is uniquely determined and is referred to as the universal χ\chi-covering transformation group. Also, (γχ(A),γχ(B))(\gamma_{\chi}^{(A)},\gamma_{\chi}^{(B)}) is said to be the standard χ\chi-pair in Γχ\Gamma_{\chi}. We call Πχ\Pi_{\chi} a holomorphic universal χ\chi-covering map. Unless SS is a torus, it is uniquely decided.

Let χ\chi^{\prime} be another 11-handle mark of SS. Set A=χA1A^{\prime}=\chi^{\prime}_{*}A_{\sqrt{-1}} and B=χB1B^{\prime}=\chi^{\prime}_{*}B_{\sqrt{-1}}, and let γ(A)\gamma^{(A^{\prime})} and γ(B)\gamma^{(B^{\prime})} be covering transformations in Γχ\Gamma_{\chi} corresponding to AA^{\prime} and BB^{\prime}. This means, in the case where SS is not a torus, that γ(A)\gamma^{(A^{\prime})} and γ(B)\gamma^{(B^{\prime})} are hyperbolic transformations and that their axes have one point in common and are projected to the closed hyperbolic geodesics AχA_{\chi^{\prime}} and BχB_{\chi^{\prime}} freely homotopic to AA^{\prime} and BB^{\prime}. Let δ\delta be the unique element in Aut()\operatorname{Aut}(\mathbb{H}) that maps 0, \infty and 11 to the repelling fixed point of γ(A)\gamma^{(A^{\prime})}, the attracting fixed point of γ(A)\gamma^{(A^{\prime})} and the repelling fixed point of γ(B)\gamma^{(B^{\prime})}, respectively. Then Πχδ\Pi_{\chi}\circ\delta is the holomorphic universal χ\chi^{\prime}-covering map. If SS is a torus, then γ(A)\gamma^{(A^{\prime})} is of the form γA(z)=z+a\gamma_{A^{\prime}}(z)=z+a^{\prime} with a0a^{\prime}\neq 0. Taking δAut()\delta\in\operatorname{Aut}(\mathbb{C}) defined by δ(z)=az\delta(z)=a^{\prime}z, we obtain a holomorphic universal χ\chi^{\prime}-covering map Πχδ:S\Pi_{\chi}\circ\delta:\mathbb{C}\to S. In either case, δ1Γχδ\delta^{-1}\circ\Gamma_{\chi}\circ\delta is the universal χ\chi^{\prime}-covering transformation group and the standard χ\chi^{\prime}-pair is (δ1γ(A)δ,δ1γ(B)δ)(\delta^{-1}\circ\gamma^{(A^{\prime})}\circ\delta,\delta^{-1}\circ\gamma^{(B^{\prime})}\circ\delta).

Now, let S1S_{1} and S2S_{2} be Riemann surfaces of positive genera, and let fTEmb+(S1,S2)f\in\operatorname{TEmb}^{+}(S_{1},S_{2}). If χ1\chi_{1} is a 11-handle mark of S1S_{1}, then χ2:=fχ1\chi_{2}:=f\circ\chi_{1} is a 11-handle mark of S2S_{2}, and ff is lifted to a locally homeomorphic mapping f~\tilde{f} of 𝕌S1\mathbb{U}_{S_{1}} into 𝕌S2\mathbb{U}_{S_{2}} such that

Πχ2f~=fΠχ1.\Pi_{\chi_{2}}\circ\tilde{f}=f\circ\Pi_{\chi_{1}}.

Note that f~\tilde{f} is not necessarily injective. There is a group homomorphism ρ:Γχ1Γχ2\rho:\Gamma_{\chi_{1}}\to\Gamma_{\chi_{2}} such that

f~γ=ρ(γ)f~,γΓχ1.\tilde{f}\circ\gamma=\rho(\gamma)\circ\tilde{f},\quad\gamma\in\Gamma_{\chi_{1}}.

We can choose f~\tilde{f} so that if (γχ(A),γχ(B))(\gamma_{\chi}^{(A)},\gamma_{\chi}^{(B)}) is the standard χ1\chi_{1}-pair in Γχ1\Gamma_{\chi_{1}}, then (ρ(γχ(A)),ρ(γχ(B)))(\rho(\gamma_{\chi}^{(A)}),\rho(\gamma_{\chi}^{(B)})) is the standard χ2\chi_{2}-pair in Γχ2\Gamma_{\chi_{2}}. If S2S_{2} is not a torus, then nor is S1S_{1}, and those requirements determine f~\tilde{f} and ρ\rho uniquely. If S2S_{2} is a torus but S1S_{1} is not, then we adjust Πχ2\Pi_{\chi_{2}} in addition so that f~(i)=0\tilde{f}(i)=0, where i=1i=\sqrt{-1\,}. Again, f~\tilde{f} and ρ\rho are uniquely determined. In the case where S1S_{1} is a torus, so is S2S_{2} and ff is a member of Homeo+(S1,S2)\operatorname{Homeo}^{+}(S_{1},S_{2}). Moreover, ρ\rho is uniquely decided and we can choose Πχj\Pi_{\chi_{j}} and f~\tilde{f} so that f~(0)=0\tilde{f}(0)=0 though we cannot assert the uniqueness of f~\tilde{f}. In any case we say that f~\tilde{f} and ρ\rho are a χ1\chi_{1}-lift of ff and the χ1\chi_{1}-homomorphism induced by ff, respectively, where we should call f~\tilde{f} the χ1\chi_{1}-lift of ff unless S1S_{1} is a torus. Note that if fQCHomeo(S1,S2)f\in\operatorname{QCHomeo}(S_{1},S_{2}), then f~\tilde{f} is extended to a quasiconformal homeomorphism of ^\hat{\mathbb{C}} onto itself fixing 0, 11 and \infty.

Let f0,f1TEmb+(S1,S2)f_{0},f_{1}\in\operatorname{TEmb}^{+}(S_{1},S_{2}). Then f0f1f_{0}\simeq f_{1} if and only if f0f_{0} and f1f_{1} induce the same χ1\chi_{1}-homomorphism.

Next, we define 11-handle marks of marked Riemann surfaces of genus gg. Let 𝑺𝔉g{\bm{S}}\in\mathfrak{F}_{g}. Consider all pairs (χ,η)(\chi,\eta), where (S,η)(S,\eta) represents 𝑺\bm{S} and χ\chi is a 11-handle mark of SS. Two such pairs (χ1,η1)(\chi_{1},\eta_{1}) and (χ2,η2)(\chi_{2},\eta_{2}) are defined to be equivalent to each other if κχ1χ2\kappa\circ\chi_{1}\simeq\chi_{2} for some κCHomeo(S1,S2)\kappa\in\operatorname{CHomeo}(S_{1},S_{2}) with κη1η2\kappa\circ\eta_{1}\simeq\eta_{2}, where (Sj,ηj)𝑺(S_{j},\eta_{j})\in{\bm{S}}, j=1,2j=1,2. Each equivalence class 𝝌=[χ,η]{\bm{\chi}}=[\chi,\eta] is called a 11-handle mark of 𝑺\bm{S}. All representatives (S,η)𝑺(S,\eta)\in{\bm{S}} and (χ,η)𝝌(\chi,\eta)\in{\bm{\chi}} have the universal χ\chi-covering transformation group and the standard χ\chi-pair in common. Thus we can speak of the universal 𝛘\bm{\chi}-covering transformation group Γ𝝌\Gamma_{\bm{\chi}} and the standard 𝛘\bm{\chi}-pair in Γ𝝌\Gamma_{\bm{\chi}}.

Let 𝑺j=[Sj,ηj]𝔉g{\bm{S}}_{j}=[S_{j},\eta_{j}]\in\mathfrak{F}_{g} for j=1,2j=1,2, and take 𝒇=[f,η1,η2]TEmb+(𝑺1,𝑺2){\bm{f}}=[f,\eta_{1},\eta_{2}]\in\operatorname{TEmb}^{+}({\bm{S}}_{1},{\bm{S}}_{2}). If 𝝌1=[χ1,η1]{\bm{\chi}}_{1}=[\chi_{1},\eta_{1}] is a 11-handle mark of 𝑺1{\bm{S}}_{1}, then we denote by 𝒇𝝌1{\bm{f}}\circ{\bm{\chi}}_{1} the 11-handle mark of 𝑺2{\bm{S}}_{2} represented by (fχ1,η2)(f\circ\chi_{1},\eta_{2}). If 𝑺1{\bm{S}}_{1} is not a marked torus, then the χ1\chi_{1}-lift f~\tilde{f} of ff and the χ1\chi_{1}-homomorphism ρ\rho induced by ff are independent of a particular choice of representatives (Sj,ηj)(S_{j},\eta_{j}), (χ1,η1)(\chi_{1},\eta_{1}) and (f,η1,η2)(f,\eta_{1},\eta_{2}). We call f~\tilde{f} and ρ\rho the 𝝌1\bm{\chi}_{1}-lift of 𝒇\bm{f} and the 𝝌1{\bm{\chi}}_{1}-homomorphism induced by 𝒇\bm{f}, respectively. If 𝑺1{\bm{S}}_{1} is a marked torus, then the 𝝌1{\bm{\chi}}_{1}-homomorphism induced by 𝒇\bm{f} is well-defined while there are infinitely many χ1\chi_{1}-lifts of ff. Any one of them is referred to as a 𝝌1{\bm{\chi}}_{1}-lift of 𝒇\bm{f}.

Let 𝑺\bm{S} be a marked Riemann surface of genus gg without border. Then 𝕌S\mathbb{U}_{S}, (S,η)𝑺(S,\eta)\in{\bm{S}}, are identical with one another and it is no harm to denote it by 𝕌𝑺\mathbb{U}_{\bm{S}}. Thus 𝕌𝑺=\mathbb{U}_{\bm{S}}=\mathbb{C} if 𝑺\bm{S} is a marked torus while 𝕌𝑺=\mathbb{U}_{\bm{S}}=\mathbb{H} otherwise.

Definition 7.1 (convergence of sequence of homeomorphisms).

Let 𝑺\bm{S}, 𝑺n{\bm{S}}_{n}, nn\in\mathbb{N}, be marked Riemann surfaces of genus gg without border homeomorphic to one another, and let 𝒉nHomeo+(𝑺,𝑺n){\bm{h}}_{n}\in\operatorname{Homeo}^{+}({\bm{S}},{\bm{S}}_{n}). In the case where 𝑺\bm{S} is not a marked torus, the sequence {𝒉n}\{{\bm{h}}_{n}\} is said to converge to 𝟏𝐒{\bm{1}}_{\bm{S}} if for some 11-handle mark 𝝌\bm{\chi} of 𝑺\bm{S} the sequence of 𝝌\bm{\chi}-lifts of 𝒉n{\bm{h}}_{n} converges to id\operatorname{id}_{\mathbb{H}} locally uniformly on \mathbb{H}. If 𝑺\bm{S} is a marked torus, then for {𝒉n}\{{\bm{h}}_{n}\} to converge to 𝟏𝑺{\bm{1}}_{\bm{S}} we require some sequence of 𝝌\bm{\chi}-lifts of 𝒉n{\bm{h}}_{n} to converge to id\operatorname{id}_{\mathbb{C}} locally uniformly on \mathbb{C}.

We claim that the above definition does not depend on 𝝌\bm{\chi}. Set 𝝌n=𝒉n𝝌{\bm{\chi}}_{n}={\bm{h}}_{n}\circ{\bm{\chi}}, and let h~n:𝕌𝕌\tilde{h}_{n}:\mathbb{U}\to\mathbb{U} be a 𝝌\bm{\chi}-lift of 𝒉n{\bm{h}}_{n}, where 𝕌=𝕌𝑺=𝕌𝑺n\mathbb{U}=\mathbb{U}_{\bm{S}}=\mathbb{U}_{{\bm{S}}_{n}}. Suppose that {h~n}\{\tilde{h}_{n}\} converges to id𝕌\operatorname{id}_{\mathbb{U}} locally uniformly on 𝕌\mathbb{U}. As h~nHomeo+(𝕌,𝕌)\tilde{h}_{n}\in\operatorname{Homeo}^{+}(\mathbb{U},\mathbb{U}), the 𝝌\bm{\chi}-homomorphism ρn:Γ𝝌Γ𝝌n\rho_{n}:\Gamma_{\bm{\chi}}\to\Gamma_{{\bm{\chi}}_{n}} induced by 𝒉n{\bm{h}}_{n} is given by ρn(γ)=h~nγh~n1\rho_{n}(\gamma)=\tilde{h}_{n}\circ\gamma\circ\tilde{h}_{n}^{-1}, γΓ𝝌\gamma\in\Gamma_{\bm{\chi}}. Since {h~n1}\{\tilde{h}_{n}^{-1}\} converges to id𝕌\operatorname{id}_{\mathbb{U}} locally uniformly on 𝕌\mathbb{U} (see Lemma 7.2 below), it follows that ρn(γ)γ\rho_{n}(\gamma)\to\gamma as nn\to\infty for γΓ𝝌\gamma\in\Gamma_{\bm{\chi}}. Now, let 𝝌=[χ,η]{\bm{\chi}}^{\prime}=[\chi^{\prime},\eta] be another 11-handle mark of 𝑺=[S,η]{\bm{S}}=[S,\eta], and set 𝝌n=𝒉n𝝌{\bm{\chi}}^{\prime}_{n}={\bm{h}}_{n}\circ{\bm{\chi}}^{\prime}. Take γ(A),γ(B)Γ𝝌\gamma^{(A^{\prime})},\gamma^{(B^{\prime})}\in\Gamma_{\bm{\chi}} corresponding to A:=χA1A^{\prime}:=\chi^{\prime}_{*}A_{\sqrt{-1}} and B:=χB1B^{\prime}:=\chi^{\prime}_{*}B_{\sqrt{-1}}. Since ρn(γ(A))γ(A)\rho_{n}(\gamma^{(A^{\prime})})\to\gamma^{(A^{\prime})} and ρn(γ(B))γ(B)\rho_{n}(\gamma^{(B^{\prime})})\to\gamma^{(B^{\prime})} as nn\to\infty, we can choose δ,δnAut(𝕌)\delta,\delta_{n}\in\operatorname{Aut}(\mathbb{U}) with Γ𝝌=δ1Γ𝝌δ\Gamma_{{\bm{\chi}}^{\prime}}=\delta^{-1}\circ\Gamma_{\bm{\chi}}\circ\delta and Γ𝝌n=δn1Γ𝝌nδn\Gamma_{{\bm{\chi}}^{\prime}_{n}}=\delta_{n}^{-1}\circ\Gamma_{{\bm{\chi}}_{n}}\circ\delta_{n} so that δnδ\delta_{n}\to\delta. Consequently, the sequence {δn1h~nδ}\{\delta_{n}^{-1}\circ\tilde{h}_{n}\circ\delta\} tends to id𝕌\operatorname{id}_{\mathbb{U}} locally uniformly on 𝕌\mathbb{U}. Since δn1h~nδ\delta_{n}^{-1}\circ\tilde{h}_{n}\circ\delta is a 𝝌{\bm{\chi}}^{\prime}-lift of 𝒉n{\bm{h}}_{n}, our claim has been established.

The following lemma, which is well-known among those who are familiar with descriptive set theory, follows from Arens [5, Theorem 3]. We include a direct proof for the sake of convenience.

Lemma 7.2.

Let XX and YY be locally compact metric spaces homeomorphic to each other. If a sequence {hn}\{h_{n}\} in Homeo(X,Y)\operatorname{Homeo}(X,Y) converges to hHomeo(X,Y)h\in\operatorname{Homeo}(X,Y) uniformly on every compact subset of XX, then {hn1}\{h_{n}^{-1}\} converges to h1h^{-1} uniformly on every compact subset of YY.

Proof.

Let LL be an arbitrary compact subset of YY, and let ε>0\varepsilon>0. Take a compact subset UU of XX including the ε\varepsilon-neighborhood of K:=h1(L)K:=h^{-1}(L). Since h1h^{-1} is uniformly continuous on V:=h(U)V:=h(U), there is δ>0\delta>0 such that

dX(h1(y1),h1(y2))<εd_{X}(h^{-1}(y_{1}),h^{-1}(y_{2}))<\varepsilon

whenever y1,y2Vy_{1},y_{2}\in V and dY(y1,y2)<δd_{Y}(y_{1},y_{2})<\delta, or equivalently,

dY(h(x1),h(x2))δd_{Y}(h(x_{1}),h(x_{2}))\geqq\delta

whenever x1,x2Ux_{1},x_{2}\in U and dX(x1,x2)εd_{X}(x_{1},x_{2})\geqq\varepsilon, where dXd_{X} and dYd_{Y} denote the distance functions on XX and YY, respectively. Replacing δ\delta with a smaller one if necessary, we may assume that the δ\delta-neighborhood of LL is included in VV.

Since {hn}\{h_{n}\} converges to hh uniformly on UU, we can find NN\in\mathbb{N} for which

dY(hn(x),h(x))<δ 3d_{Y}(h_{n}(x),h(x))<\frac{\delta}{\,3\,}

for n>Nn>N and xUx\in U. Note that if x1,x2Ux_{1},x_{2}\in U and dX(x1,x2)εd_{X}(x_{1},x_{2})\geqq\varepsilon, then

dY(hn(x1),hn(x2))\displaystyle d_{Y}(h_{n}(x_{1}),h_{n}(x_{2})) dY(h(x1),h(x2))dY(hn(x1),h(x1))dY(hn(x2),h(x2))\displaystyle\geqq d_{Y}(h(x_{1}),h(x_{2}))-d_{Y}(h_{n}(x_{1}),h(x_{1}))-d_{Y}(h_{n}(x_{2}),h(x_{2}))
δδ 3δ 3=δ3\displaystyle\geqq\delta-\frac{\delta}{\,3\,}-\frac{\delta}{\,3\,}=\frac{\delta}{3}

for n>Nn>N. In other words,

dX(hn1(y1),hn1(y2))<εd_{X}(h_{n}^{-1}(y_{1}),h_{n}^{-1}(y_{2}))<\varepsilon

for n>Nn>N if y1,y2Vy_{1},y_{2}\in V and dY(y1,y2)<δ/3d_{Y}(y_{1},y_{2})<\delta/3.

We now claim that

dX(hn1(y),h1(y))<εd_{X}(h_{n}^{-1}(y),h^{-1}(y))<\varepsilon

for n>Nn>N and yLy\in L, which will prove the lemma. To verify the claim take yLy\in L and n>Nn>N arbitrarily, and set x=h1(y)x=h^{-1}(y) and yn=hn(x)y_{n}=h_{n}(x). Then xx is a point of KUK\subset U so that yny_{n} belongs to the (δ/3)(\delta/3)-neighborhood of LL and hence to VV, for, dY(y,yn)=dY(h(x),hn(x))<δ/3d_{Y}(y,y_{n})=d_{Y}(h(x),h_{n}(x))<\delta/3. We then have

dX(hn1(y),h1(y))=dX(hn1(y),hn1(yn))<ε,d_{X}(h_{n}^{-1}(y),h^{-1}(y))=d_{X}(h_{n}^{-1}(y),h_{n}^{-1}(y_{n}))<\varepsilon,

as claimed. ∎

Definition 7.3 (convergence of sequence of topological embeddings).

Let 𝑺\bm{S} and 𝑺{\bm{S}}^{\prime} be marked Riemann surfaces of genus gg without border, and let 𝒇TEmb+(𝑺,𝑺){\bm{f}}\in\operatorname{TEmb}^{+}({\bm{S}},{\bm{S}}^{\prime}). Also, let {𝑺n}\{{\bm{S}}_{n}\} and {𝑺n}\{{\bm{S}}^{\prime}_{n}\} be sequences in 𝔉g\mathfrak{F}_{g}, where 𝑺n{\bm{S}}_{n} and 𝑺n{\bm{S}}^{\prime}_{n} are supposed to be homeomorphic to 𝑺\bm{S} and 𝑺{\bm{S}}^{\prime}, respectively, and let 𝒇nTEmb+(𝑺n,𝑺n){\bm{f}}_{n}\in\operatorname{TEmb}^{+}({\bm{S}}_{n},{\bm{S}}^{\prime}_{n}). Then we say that {𝒇n}\{{\bm{f}}_{n}\} converges to 𝒇\bm{f} if there are 𝒉nHomeo+(𝑺,𝑺n){\bm{h}}_{n}\in\operatorname{Homeo}^{+}({\bm{S}},{\bm{S}}_{n}) and 𝒉nHomeo+(𝑺,𝑺n){\bm{h}}^{\prime}_{n}\in\operatorname{Homeo}^{+}({\bm{S}}^{\prime},{\bm{S}}^{\prime}_{n}) together with a 11-handle mark 𝝌\bm{\chi} of 𝑺\bm{S} such that

  • (i)

    {𝒉n}\{{\bm{h}}_{n}\} and {𝒉n}\{{\bm{h}}^{\prime}_{n}\} converge to 𝟏𝑺{\bm{1}}_{\bm{S}} and 𝟏𝑺{\bm{1}}_{{\bm{S}}^{\prime}}, respectively, and

  • (ii)

    the sequence of 𝒉n𝝌{\bm{h}}_{n}\circ{\bm{\chi}}-lifts of 𝒇n{\bm{f}}_{n} converges to the 𝝌\bm{\chi}-lift of 𝒇\bm{f} locally uniformly on \mathbb{H},

provided that 𝑺\bm{S} is not a marked torus. In the case where 𝑺\bm{S} is a marked torus, we replace condition (ii) with

  • (ii)

    some sequence of 𝒉n𝝌{\bm{h}}_{n}\circ{\bm{\chi}}-lifts of 𝒇n{\bm{f}}_{n} converges to a 𝝌\bm{\chi}-lift of 𝒇\bm{f} locally uniformly on \mathbb{C}.

Lemma 7.4.

Let 𝐒\bm{S} and 𝐒{\bm{S}}^{\prime} be marked Riemann surfaces of genus gg without border. Let {𝐒n}\{{\bm{S}}_{n}\} be a sequence in 𝔉g\mathfrak{F}_{g} and let (𝐒n,𝛊n)({\bm{S}}^{\prime}_{n},{\bm{\iota}}_{n}) be a continuation of 𝐒n{\bm{S}}_{n}. Assume that there are 𝐡nHomeo+(𝐒,𝐒n){\bm{h}}_{n}\in\operatorname{Homeo}^{+}({\bm{S}},{\bm{S}}_{n}) and 𝐡nHomeo+(𝐒,𝐒n){\bm{h}}^{\prime}_{n}\in\operatorname{Homeo}^{+}({\bm{S}}^{\prime},{\bm{S}}^{\prime}_{n}) such that (𝐡n)1𝛊n𝐡n({\bm{h}}^{\prime}_{n})^{-1}\circ{\bm{\iota}}_{n}\circ{\bm{h}}_{n}, nn\in\mathbb{N}, are mutually homotopic. If {𝐡n}\{{\bm{h}}_{n}\} and {𝐡n}\{{\bm{h}}^{\prime}_{n}\} converge to 𝟏𝐒{\bm{1}}_{\bm{S}} and 𝟏𝐒{\bm{1}}_{{\bm{S}}^{\prime}}, respectively, then there is a continuation (𝐒,𝛊)({\bm{S}}^{\prime},{\bm{\iota}}) of 𝐒\bm{S} such that a subsequence of {𝛊n}\{{\bm{\iota}}_{n}\} converges to 𝛊{\bm{\iota}}.

Proof.

Let 𝝌\bm{\chi} be a 11-handle mark of 𝑺\bm{S}, and set 𝝌n=𝒉n𝝌{\bm{\chi}}_{n}={\bm{h}}_{n}\circ{\bm{\chi}}, 𝝌n=𝜾n𝝌n{\bm{\chi}}^{\prime}_{n}={\bm{\iota}}_{n}\circ{\bm{\chi}}_{n} and 𝝌=(𝒉1)1𝝌1{\bm{\chi}}^{\prime}=({\bm{h}}^{\prime}_{1})^{-1}\circ{\bm{\chi}}^{\prime}_{1}. Let h~n\tilde{h}_{n} and h~n\tilde{h}^{\prime}_{n} be 𝝌\bm{\chi}- and 𝝌{\bm{\chi}}^{\prime}-lifts of 𝒉n{\bm{h}}_{n} and 𝒉n{\bm{h}}^{\prime}_{n}, respectively. By assumption we can choose {h~n}\{\tilde{h}_{n}\} and {h~n}\{\tilde{h}^{\prime}_{n}\} so that they converge locally uniformly to id𝕌𝑺\operatorname{id}_{\mathbb{U}_{\bm{S}}} and id𝕌𝑺\operatorname{id}_{\mathbb{U}_{{\bm{S}}^{\prime}}}, respectively.

If 𝑺\bm{S} is closed, then so are 𝑺{\bm{S}}^{\prime}, 𝑺n{\bm{S}}_{n} and 𝑺n{\bm{S}}^{\prime}_{n}, and hence 𝜾nCHomeohc(𝑺n,𝑺n){\bm{\iota}}_{n}\in\operatorname{CHomeo}_{\mathrm{hc}}({\bm{S}}_{n},{\bm{S}}^{\prime}_{n}). Thus 𝑺n{\bm{S}}_{n} coincides with 𝑺n{\bm{S}}^{\prime}_{n}, and id𝕌𝑺\operatorname{id}_{\mathbb{U}_{\bm{S}}} is a 𝝌n{\bm{\chi}}_{n}-lift of 𝜾n{\bm{\iota}}_{n}. The convergence properties of {h~n}\{\tilde{h}_{n}\} and {h~n}\{\tilde{h}^{\prime}_{n}\} imply that {𝑺n}\{{\bm{S}}_{n}\} and {𝑺n}\{{\bm{S}}^{\prime}_{n}\} converge to 𝑺\bm{S} and 𝑺{\bm{S}}^{\prime} in 𝔗g\mathfrak{T}_{g}, respectively. In particular, 𝑺\bm{S} is identical with 𝑺{\bm{S}}^{\prime}, and {𝜾n}\{{\bm{\iota}}_{n}\} converges to 𝟏𝑺{\bm{1}}_{\bm{S}}.

Suppose now that 𝑺\bm{S} is open. Then 𝕌𝑺=\mathbb{U}_{\bm{S}}=\mathbb{H}. Set 𝕌=𝕌𝑺\mathbb{U}=\mathbb{U}_{{\bm{S}}^{\prime}} for the sake of simplicity. If g=1g=1 and 𝑺𝔗1{\bm{S}}^{\prime}\in\mathfrak{T}_{1}, then 𝕌=\mathbb{U}=\mathbb{C}. Otherwise, 𝕌=\mathbb{U}=\mathbb{H}.

Take (S,η)𝑺(S,\eta)\in{\bm{S}} and (Sn,ηn)𝑺n(S_{n},\eta_{n})\in{\bm{S}}_{n}, and let (hn,η,ηn)𝒉n(h_{n},\eta,\eta_{n})\in{\bm{h}}_{n}, (χ,η)𝝌(\chi,\eta)\in{\bm{\chi}} and (χn,ηn)𝝌n(\chi_{n},\eta_{n})\in{\bm{\chi}}_{n}. Then we have

Πχnh~n=hnΠχ,\Pi_{\chi_{n}}\circ\tilde{h}_{n}=h_{n}\circ\Pi_{\chi},

and the χ\chi-homomorphism of Γχ\Gamma_{\chi} onto Γχn\Gamma_{\chi_{n}} induced by hnh_{n} is given by

γh~nγh~n1,γΓχ.\gamma\mapsto\tilde{h}_{n}\circ\gamma\circ\tilde{h}_{n}^{-1},\quad\gamma\in\Gamma_{\chi}.

Since {h~n}\{\tilde{h}_{n}\} converges to id\operatorname{id}_{\mathbb{H}} locally uniformly in \mathbb{H}, so does {h~n1}\{\tilde{h}_{n}^{-1}\} by Lemma 7.2.

Similarly, take (S,η)𝑺(S^{\prime},\eta^{\prime})\in{\bm{S}}^{\prime} and (Sn,ηn)𝑺n(S^{\prime}_{n},\eta^{\prime}_{n})\in{\bm{S}}^{\prime}_{n}, and let (hn,η,ηn)𝒉n(h^{\prime}_{n},\eta^{\prime},\eta^{\prime}_{n})\in{\bm{h}}^{\prime}_{n}, (χ,η)𝝌(\chi^{\prime},\eta^{\prime})\in{\bm{\chi}}^{\prime} and (χn,ηn)𝝌n(\chi^{\prime}_{n},\eta^{\prime}_{n})\in{\bm{\chi}}^{\prime}_{n}. If SS^{\prime} is not a torus, then Πχ\Pi_{\chi^{\prime}} and Πχn\Pi_{\chi^{\prime}_{n}} are uniquely determined and satisfy

(7.1) Πχnh~n=hnΠχ.\Pi_{\chi^{\prime}_{n}}\circ\tilde{h}^{\prime}_{n}=h^{\prime}_{n}\circ\Pi_{\chi^{\prime}}.

If SS^{\prime} is a torus, then we can adjust Πχ\Pi_{\chi^{\prime}}, Πχn\Pi_{\chi_{n}} and h~n\tilde{h}^{\prime}_{n} to obtain (7.1) for all nn. The χn\chi^{\prime}_{n}-homomorphism of Γχ\Gamma_{\chi^{\prime}} onto Γχn\Gamma_{\chi^{\prime}_{n}} induced by hnh^{\prime}_{n} is given by

γh~nγ(h~n)1,γΓχ.\gamma\mapsto\tilde{h}^{\prime}_{n}\circ\gamma\circ(\tilde{h}^{\prime}_{n})^{-1},\quad\gamma\in\Gamma_{\chi^{\prime}}.

By Lemma 7.2 again, the sequences {h~n}\{\tilde{h}^{\prime}_{n}\} and {(h~n)1}\{(\tilde{h}^{\prime}_{n})^{-1}\} converge to id𝕌\operatorname{id}_{\mathbb{U}} locally uniformly on 𝕌\mathbb{U}. In the case where 𝕌=\mathbb{U}=\mathbb{C}, we have Γχ=Γτ\Gamma_{\chi^{\prime}}=\Gamma_{\tau^{\prime}} and Γχn=Γτn\Gamma_{\chi^{\prime}_{n}}=\Gamma_{\tau^{\prime}_{n}} for some τ,τn\tau^{\prime},\tau^{\prime}_{n}\in\mathbb{H} (for the notations see Example 2.3). Note that τnτ\tau^{\prime}_{n}\to\tau^{\prime} as nn\to\infty.

Let ι~n\tilde{\iota}_{n} and ρn\rho_{n} be the χn\chi_{n}-lift of ιn\iota_{n} and the χn\chi_{n}-homomorphism induced by ιn\iota_{n}, respectively, where (ιn,ηn,ηn)𝜾n(\iota_{n},\eta_{n},\eta^{\prime}_{n})\in{\bm{\iota}}_{n}. Then we have

Πχnι~n=ιnΠχn\Pi_{\chi^{\prime}_{n}}\circ\tilde{\iota}_{n}=\iota_{n}\circ\Pi_{\chi_{n}}

and

ι~nγ=ρn(γ)ι~n,γΓχn.\tilde{\iota}_{n}\circ\gamma=\rho_{n}(\gamma)\circ\tilde{\iota}_{n},\quad\gamma\in\Gamma_{\chi_{n}}.

Observe that (h~n)1ι~nh~n(\tilde{h}^{\prime}_{n})^{-1}\circ\tilde{\iota}_{n}\circ\tilde{h}_{n} is the χ\chi-lift of (hn)1ιnhn(h^{\prime}_{n})^{-1}\circ\iota_{n}\circ h_{n}. Since (hn)1ιnhnTEmb+(S,S)(h^{\prime}_{n})^{-1}\circ\iota_{n}\circ h_{n}\in\operatorname{TEmb}^{+}(S,S^{\prime}), nn\in\mathbb{N}, are homotopic to one another, they induce the same χ\chi-homomorphism ρ:ΓχΓχ\rho:\Gamma_{\chi}\to\Gamma_{\chi^{\prime}}^{\prime}. It follows that

(7.2) ι~n(h~nγh~n1)=(h~nρ(γ)(h~n)1)ι~n,γΓχ,\tilde{\iota}_{n}\circ\bigl{(}\tilde{h}_{n}\circ\gamma\circ\tilde{h}_{n}^{-1}\bigr{)}=\bigl{(}\tilde{h}^{\prime}_{n}\circ\rho(\gamma)\circ(\tilde{h}^{\prime}_{n})^{-1}\bigr{)}\circ\tilde{\iota}_{n},\quad\gamma\in\Gamma_{\chi},

for all nn.

h~nι~n𝕌h~n𝕌ΠχΠχnΠχnΠχShnSnιnSnhnS\begin{CD}\mathbb{H}@>{\tilde{h}_{n}}>{}>\mathbb{H}@>{\tilde{\iota}_{n}}>{}>\mathbb{U}@<{\tilde{h}^{\prime}_{n}}<{}<\mathbb{U}\\ @V{{\Pi_{\chi}}}V{}V@V{\Pi_{\chi_{n}}}V{}V@V{\Pi_{\chi^{\prime}_{n}}}V{}V@V{\Pi_{\chi^{\prime}}}V{}V\\ S@>{h_{n}}>{}>S_{n}@>{\iota_{n}}>{}>S^{\prime}_{n}@<{h^{\prime}_{n}}<{}<S^{\prime}\end{CD}

We claim that {ι~n}\{\tilde{\iota}_{n}\} has a subsequence converging locally uniformly on \mathbb{H} to ι~\tilde{\iota}, which is a holomorphic function or the constant \infty. This is clear if 𝕌=\mathbb{U}=\mathbb{H}. If 𝕌=\mathbb{U}=\mathbb{C}, then there is a constant c~n\tilde{c}_{n} such that ι~nc~n\tilde{\iota}_{n}-\tilde{c}_{n} omits 0, for, ιn\iota_{n} is not surjective. Then it also omits 11, which is in the Γχn\Gamma_{\chi^{\prime}_{n}}-orbit of 0. Therefore, {ι~nc~n}\{\tilde{\iota}_{n}-\tilde{c}_{n}\} forms a normal family. Since {c~n}\{\tilde{c}_{n}\} can be chosen to be bounded, we conclude that {ι~n}\{\tilde{\iota}_{n}\} includes a desired subsequence.

For typographical reason we assume that {ι~n}\{\tilde{\iota}_{n}\} converges to ι~\tilde{\iota} locally uniformly on \mathbb{H}. It then follows from (7.2) that

(7.3) ι~γ=ρ(γ)ι~,γΓχ.\tilde{\iota}\circ\gamma=\rho(\gamma)\circ\tilde{\iota},\quad\gamma\in\Gamma_{\chi}.

If ι~\tilde{\iota} were a constant function, then the constant w0^w_{0}\in\hat{\mathbb{C}} should satisfy

ρ(γ)(w0)=ρ(γ)ι~(z)=ι~γ(z)=w0\rho(\gamma)(w_{0})=\rho(\gamma)\circ\tilde{\iota}(z)=\tilde{\iota}\circ\gamma(z)=w_{0}

for zz\in\mathbb{H} and γΓχ\gamma\in\Gamma_{\chi}. Thus w0w_{0} is a common fixed point of elements of ρ(Γχ)\rho(\Gamma_{\chi}), which implies that ρ(Γχ)\rho(\Gamma_{\chi}) is abelian (see, for example, Beardon [6, Theorem 5.1.2]). This is impossible if 𝕌=\mathbb{U}=\mathbb{H}. If 𝕌=\mathbb{U}=\mathbb{C}, then w0=0w_{0}=0 as ι~n(i)=0\tilde{\iota}_{n}(i)=0 for all nn. However, no finite point is fixed by any nontrivial elements of Γχ\Gamma_{\chi^{\prime}} and we again reach a contradiction. Therefore, ι~\tilde{\iota} is nonconstant and holomorphic, and hence is locally univalent as well as ι~n\tilde{\iota}_{n}.

It follows from (7.3) that ι~\tilde{\iota} induces a holomorphic and locally homeomorphic mapping ι:SS\iota:S\to S^{\prime} satisfying Πχι~=ιΠχ\Pi_{\chi^{\prime}}\circ\tilde{\iota}=\iota\circ\Pi_{\chi}. We claim that it is injective. Otherwise, we could find two distinct points p1,p2Sp_{1},p_{2}\in S mapped to the same point q0q_{0} of SS^{\prime} by ι\iota. Let zjΠχ1(pj)z_{j}\in\Pi_{\chi}^{-1}(p_{j}), j=1,2j=1,2, and set wj=ι~(zj)w_{j}=\tilde{\iota}(z_{j}). Since

Πχ(wj)=Πχι~(zj)=ιΠχ(zj)=ι(pj)=q0,\Pi_{\chi^{\prime}}(w_{j})=\Pi_{\chi^{\prime}}\circ\tilde{\iota}(z_{j})=\iota\circ\Pi_{\chi}(z_{j})=\iota(p_{j})=q_{0},

we have w2=γ(w1)w_{2}=\gamma^{\prime}(w_{1}) for some γΓχ\gamma^{\prime}\in\Gamma_{\chi^{\prime}}. Take a relatively compact neighborhood UjU_{j} of zjz_{j} such that Πχ(U1)Πχ(U2)=\Pi_{\chi}(U_{1})\cap\Pi_{\chi}(U_{2})=\varnothing and γι~(U1)=ι~(U2)\gamma^{\prime}\circ\tilde{\iota}(U_{1})=\tilde{\iota}(U_{2}). Since {ι~n}\{\tilde{\iota}_{n}\} (resp. {h~n}\{\tilde{h}_{n}\}, {h~n}\{\tilde{h}^{\prime}_{n}\}) converges to ι~\tilde{\iota} (resp. id\operatorname{id}_{\mathbb{H}}, id𝕌\operatorname{id}_{\mathbb{U}}) locally uniformly on \mathbb{H} (resp. \mathbb{H}, 𝕌\mathbb{U}), there are neighborhoods VjV_{j} of zjz_{j} with V¯jUj\bar{V}_{j}\subset U_{j} and elements γn\gamma^{\prime}_{n} in Γχn\Gamma_{\chi^{\prime}_{n}} such that Vjh~n(Uj)V_{j}\subset\tilde{h}_{n}(U_{j}) and γnι~n(V1)ι~n(V2)\gamma^{\prime}_{n}\circ\tilde{\iota}_{n}(V_{1})\cap\tilde{\iota}_{n}(V_{2})\neq\varnothing for sufficiently large nn. Choose zjnVjz_{jn}\in V_{j} so that γnι~n(z1n)=ι~n(z2n)\gamma^{\prime}_{n}\circ\tilde{\iota}_{n}(z_{1n})=\tilde{\iota}_{n}(z_{2n}). Then we obtain

ιnΠχn(z1n)=Πχnι~n(z1n)=Πχnγnι~n(z1n)=Πχnι~n(z2n)=ιnΠχn(z2n),\iota_{n}\circ\Pi_{\chi_{n}}(z_{1n})=\Pi_{\chi^{\prime}_{n}}\circ\tilde{\iota}_{n}(z_{1n})=\Pi_{\chi^{\prime}_{n}}\circ\gamma^{\prime}_{n}\circ\tilde{\iota}_{n}(z_{1n})=\Pi_{\chi^{\prime}_{n}}\circ\tilde{\iota}_{n}(z_{2n})=\iota_{n}\circ\Pi_{\chi_{n}}(z_{2n}),

which implies that Πχn(z1n)=Πχn(z2n)\Pi_{\chi_{n}}(z_{1n})=\Pi_{\chi_{n}}(z_{2n}), or, z2n=h~nγnh~n1(z1n)z_{2n}=\tilde{h}_{n}\circ\gamma_{n}\circ\tilde{h}_{n}^{-1}(z_{1n}) for some γnΓχ\gamma_{n}\in\Gamma_{\chi}, for, ιn\iota_{n} is injective. Then γnh~n1(z1n)=h~n1(z2n)γn(U1)U2=\gamma_{n}\circ\tilde{h}_{n}^{-1}(z_{1n})=\tilde{h}_{n}^{-1}(z_{2n})\in\gamma_{n}(U_{1})\cap U_{2}=\varnothing, which is absurd. This proves that ι\iota is injective, as claimed. If we set 𝜾=[ι,η,η]{\bm{\iota}}=[\iota,\eta,\eta^{\prime}], then (𝑺,𝜾)({\bm{S}}^{\prime},{\bm{\iota}}) is a continuation of 𝑺\bm{S} and {𝜾n}\{{\bm{\iota}}_{n}\} converges to 𝜾{\bm{\iota}}. ∎

Proposition 7.5.

If 𝐑0{\bm{R}}_{0} is a marked open Riemann surface of genus gg, then any sequence {(𝐑n,𝛊n)}\{({\bm{R}}_{n},{\bm{\iota}}_{n})\} of closed continuations of 𝐑0{\bm{R}}_{0} contains a subsequence {(𝐑nk,𝛊nk)}\{({\bm{R}}_{n_{k}},{\bm{\iota}}_{n_{k}})\} such that {𝐑nk}\{{\bm{R}}_{n_{k}}\} and {𝛊nk}\{{\bm{\iota}}_{n_{k}}\} converge to some 𝐑𝔗g{\bm{R}}\in\mathfrak{T}_{g} and 𝛊CEmbhc(𝐑0,𝐑){\bm{\iota}}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}), respectively.

Proof.

Take representatives (Rn,θn)𝑹n(R_{n},\theta_{n})\in{\bm{R}}_{n}, n0n\geqq 0, and (ιn,θ0,θn)𝜾n(\iota_{n},\theta_{0},\theta_{n})\in{\bm{\iota}}_{n}, n>0n>0. Note that ιnθ0θn\iota_{n}\circ\theta_{0}\simeq\theta_{n}. If g>1g>1, then RnR_{n}, n0n\geqq 0, carry hyperbolic metrics. Since ιn\iota_{n} decreases the hyperbolic metrics, for any simple loop cc on Σ˙g\dot{\Sigma}_{g} the hyperbolic length of (θn)c(\theta_{n})_{*}c does not exceed that of (θ0)c(\theta_{0})_{*}c. With the aid of the Fenchel-Nielsen coordinates on 𝔗g\mathfrak{T}_{g} (see, for example, Abikoff [1, Chapter II §3]) we deduce that {𝑹n}\{{\bm{R}}_{n}\} is relatively compact in 𝔗g\mathfrak{T}_{g}. If g=1g=1, then choosing a point pnRnιn(R0)p_{n}\in R_{n}\setminus\iota_{n}(R_{0}) for n>0n>0 and making use of hyperbolic metrics on R0R_{0} and Rn{pn}R_{n}\setminus\{p_{n}\}, we reach the same conclusion. In either case {𝑹n}\{{\bm{R}}_{n}\} includes a convergent subsequence. We may assume that {𝑹n}\{{\bm{R}}_{n}\} itself converges to some 𝑹\bm{R} in 𝔗g\mathfrak{T}_{g}. If 𝒉n{\bm{h}}_{n} denotes the Teichmüller quasiconformal homeomorphism of 𝑹\bm{R} onto 𝑹n{\bm{R}}_{n}, then {𝒉n}\{{\bm{h}}_{n}\} converges to 𝟏𝑹{\bm{1}}_{\bm{R}}. We apply Lemma 7.4 to obtain desired subsequences {𝑹nk}\{{\bm{R}}_{n_{k}}\} and {𝜾nk}\{{\bm{\iota}}_{n_{k}}\} and a continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) of 𝑹0{\bm{R}}_{0}. ∎

Corollary 7.6 (Oikawa [45]).

If 𝐑0{\bm{R}}_{0} is a marked open Riemann surface of genus gg, then 𝔐(𝐑0)\mathfrak{M}({\bm{R}}_{0}) is compact.

Remark.

Note that 𝑹0{\bm{R}}_{0} is not assumed to be finite. An alternative proof of the corollary is found in [38, Proposition 5.3]. In [45] Oikawa also proved that 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is connected. Theorem 1.1 gives an alternative proof of this fact in the case where 𝑹0{\bm{R}}_{0} is finite. The connectedness of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) for general 𝑹0{\bm{R}}_{0} easily follows from this special case, as was shown in [45].

Proposition 7.7.

Let 𝐑0{\bm{R}}_{0} be a marked finite open Riemann surface of genus gg, and let {(𝐑n,𝛊n)}\{({\bm{R}}_{n},{\bm{\iota}}_{n})\} be a sequence of closed regular self-welding continuations of 𝐑0{\bm{R}}_{0}. If {𝐑n}\{{\bm{R}}_{n}\} converges to a point 𝐑\bm{R} in 𝔗g\mathfrak{T}_{g}, then there is a closed regular self-welding continuation (𝐑,𝛊)({\bm{R}},{\bm{\iota}}) of 𝐑0{\bm{R}}_{0} such that a subsequence of {𝛊n}\{{\bm{\iota}}_{n}\} converges to 𝛊\bm{\iota}.

Proof.

It follows from Proposition 7.5 that a subsequence of {𝜾n}\{{\bm{\iota}}_{n}\} converges to some 𝜾CEmbhc(𝑹0,𝑹){\bm{\iota}}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}). We may assume that {𝜾n}\{{\bm{\iota}}_{n}\} converges to 𝜾\bm{\iota}. We need to show that (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) is a regular self-welding continuation of 𝑹0{\bm{R}}_{0}.

We know that =𝕌𝑹0\mathbb{H}=\mathbb{U}_{{\bm{R}}_{0}}. Set 𝕌=𝕌𝑹=𝕌𝑹n\mathbb{U}=\mathbb{U}_{{\bm{R}}}=\mathbb{U}_{{\bm{R}}_{n}} for n1n\geqq 1. Fix a 11-handle mark 𝝌0{\bm{\chi}}_{0} of 𝑹0{\bm{R}}_{0}, and set 𝝌n=𝜾n𝝌0{\bm{\chi}}_{n}={\bm{\iota}}_{n}\circ{\bm{\chi}}_{0}. Let 𝝋nAL(𝑹0){\bm{\varphi}}_{n}\in A_{L}({\bm{R}}_{0}) and 𝝍nA(𝑹n){\bm{\psi}}_{n}\in A({\bm{R}}_{n}) be a welder of the self-welding continuation (𝑹n,𝜾n)({\bm{R}}_{n},{\bm{\iota}}_{n}) of 𝑹0{\bm{R}}_{0} and its co-welder, respectively. They induce automorphic 22-forms φ~n\tilde{\varphi}_{n} and ψ~n\tilde{\psi}_{n} for Γ𝝌0\Gamma_{{\bm{\chi}}_{0}} and Γ𝝌n\Gamma_{{\bm{\chi}}_{n}}, respectively. We adopt the normalization condition 𝝋n𝑹0=1\|{\bm{\varphi}}_{n}\|_{{\bm{R}}_{0}}=1. As 𝝍n𝑹n=𝝋n𝑹0=1\|{\bm{\psi}}_{n}\|_{{\bm{R}}_{n}}=\|{\bm{\varphi}}_{n}\|_{{\bm{R}}_{0}}=1, by taking subsequences if necessary, we may assume that {φ~n}\{\tilde{\varphi}_{n}\} and {ψ~n}\{\tilde{\psi}_{n}\} converge locally uniformly to φ~A()\tilde{\varphi}\in A(\mathbb{H}) and ψ~A(𝕌)\tilde{\psi}\in A(\mathbb{U}), respectively. Proposition 4.1 implies that φ~\tilde{\varphi} is projected to some 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}) with 𝝋𝑹0=1\|{\bm{\varphi}}\|_{{\bm{R}}_{0}}=1. Also, ψ~\tilde{\psi} is projected to some 𝝍A(𝑹){\bm{\psi}}\in A({\bm{R}}) with 𝝍𝑹=1\|{\bm{\psi}}\|_{\bm{R}}=1.

Denote by ι~n\tilde{\iota}_{n} and ι~\tilde{\iota} the 𝝌0{\bm{\chi}}_{0}-lifts of 𝜾n{\bm{\iota}}_{n} and 𝜾\bm{\iota}, respectively. We know that {ι~n}\{\tilde{\iota}_{n}\} converges to ι~\tilde{\iota} locally uniformly on \mathbb{H}. Owing to ι~nψ~n=φ~n\tilde{\iota}_{n}^{*}\tilde{\psi}_{n}=\tilde{\varphi}_{n}, we have ι~ψ~=φ~\tilde{\iota}^{*}\tilde{\psi}=\tilde{\varphi}, or equivalently, 𝜾𝝍=𝝋{\bm{\iota}}^{*}{\bm{\psi}}={\bm{\varphi}}. The continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) of 𝑹0{\bm{R}}_{0} is dense as 𝜾𝝍𝑹0=𝝋𝑹0=1=𝝍𝑹\|{\bm{\iota}}^{*}{\bm{\psi}}\|_{{\bm{R}}_{0}}=\|{\bm{\varphi}}\|_{{\bm{R}}_{0}}=1=\|{\bm{\psi}}\|_{\bm{R}}. Proposition 6.1 thus implies that (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) is a closed 𝝋\bm{\varphi}-regular self-welding continuation of 𝑹0{\bm{R}}_{0}. ∎

Remark.

We could apply Bourque [12, Lemma 5.4] to obtain Proposition 7.7. It should be noted that our proof is direct and does not require any results on extremal quasiconformal embeddings.

8 Continuations to Riemann surfaces on 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0})

The aim of the present section is to prove the following proposition. It is claimed in Kahn-Pilgrim-Thurston [30, Proposition 1.7 and Remark 1.5] without proof.

Proposition 8.1.

Let 𝐑0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface of genus gg. Then for any 𝐑𝔐(𝐑0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}) there is 𝛊CEmbhc(𝐑0,𝐑)\bm{\iota}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) such that (𝐑,𝛊)({\bm{R}},{\bm{\iota}}) is a closed regular self-welding continuation of 𝐑0{\bm{R}}_{0}.

Propositions 6.6 and 8.1 prove Theorem 1.4 (i). To prove Proposition 8.1 we employ extremal quasiconformal embeddings of finite open Riemann surfaces into closed Riemann surfaces. Following Ioffe [25] (see also Kahn-Pilgrim-Thurston [30, Definition 4.1] and Bourque [12, Definition 3.4]) we make the following definition.

Definition 8.2 (Teichmüller quasiconformal embedding).

Let R0R_{0} be a finite open Riemann surface of genus gg without border, and let RR be a closed Riemann surface of the same genus. Then fQCEmb(R0,R)f\in\operatorname{QCEmb}(R_{0},R) is called a Teichmüller quasiconformal embedding if

  • (i)

    its Beltrami differential μf\mu_{f} is of the form k|φ|/φk|\varphi|/\varphi for some φA+(R0)\varphi\in A_{+}(R_{0}) and k[0,1)k\in[0,1),

  • (ii)

    there is ψA(R)\psi\in A(R) such that ff maps every noncritical point of φ\varphi to that of ψ\psi and is represented as

    ω=K(f)Reζ+iImζ=(K(f)+1)ζ+(K(f)1)ζ¯2\omega=K(f)\operatorname{Re}\zeta+i\operatorname{Im}\zeta=\frac{\,(K(f)+1)\zeta+(K(f)-1)\bar{\zeta}\,}{2}

    with respect to some natural parameter ζ\zeta (resp. ω\omega) of φ\varphi (resp. ψ\psi) around any noncritical point pp (resp. f(p)f(p)), and

  • (iii)

    the complement Rf(R0)R\setminus f(R_{0}) consists of finitely many horizontal arcs and, possibly, zeroes of ψ\psi together with finitely many isolated points.

The quadratic differentials φ\varphi and ψ\psi are referred to as initial and terminal quadratic differentials of ff, respectively. Note that ff maps each horizontal arc of φ\varphi onto a horizontal arc of ψ\psi and is a uniform stretching along horizontal trajectories of φ\varphi. The Beltrami differential of the inverse f1:f(R0)R0f^{-1}:f(R_{0})\to R_{0} is equal to k|ψ|/ψ-k|\psi|/\psi. If k=0k=0, then ff is conformal, in which case it is said to be a Teichmüller conformal embedding.

Remark.

Teichmüller conformal embeddings are called slit mappings in Bourque [12], and Teichmüller embeddings with dilatation 11 in Kahn-Pilgrim-Thurston [30].

Let (R,ι)(R,\iota) be a genus-preserving closed continuation of R0R_{0}. If it is a closed regular self-welding continuation, then ι\iota is a Teichmüller conformal embedding, and a welder of (R,ι)(R,\iota) and its co-welder are initial and terminal quadratic differentials of ι\iota, respectively. Conversely, if ι\iota is a Teichmüller conformal embedding, then (R,ι)(R,\iota) is a closed regular self-welding continuation of R0R_{0} by Corollary 4.7. More generally, we have the following lemma.

Lemma 8.3.

Let R0R_{0} be a nonanalytically finite open Riemann surface of genus gg, and let RR be a closed Riemann surface of genus gg. If ff is a Teichmüller quasiconformal embedding of R0R_{0} into RR with initial quadratic differential φ\varphi, then there are a closed φ\varphi-regular self-welding continuation (R,ι)(R^{\prime},\iota) of R0R_{0} and a Teichmüller quasiconformal homeomorphism hh of RR^{\prime} onto RR such that

  • (i)

    the co-welder of φ\varphi is an initial quadratic differential of hh, and

  • (ii)

    f=hιf=h\circ\iota.

Proof.

Let ψA(R)\psi\in A(R) be the terminal quadratic differential of ff corresponding to φ\varphi. Thus near a noncritical point pR0p\in R_{0} of φ\varphi the mapping ff is represented as ω=KReζ+iImζ\omega=K\operatorname{Re}\zeta+i\operatorname{Im}\zeta with respect to natural parameters ζ\zeta and ω\omega of φ\varphi and ψ\psi around pp and f(p)f(p), respectively, where K=K(f)K=K(f). Let hh^{\prime} be a Teichmüller quasiconformal homeomorphism of RR onto a closed Riemann surface RR^{\prime} for which μh=μf1=k|ψ|/ψ\mu_{h^{\prime}}=\mu_{f^{-1}}=-k|\psi|/\psi, where k=(K1)/(K+1)k=(K-1)/(K+1), and set ι=hf\iota=h^{\prime}\circ f, which is a conformal embedding of R0R_{0} into RR^{\prime}. If ψA(R)\psi^{\prime}\in A(R^{\prime}) denotes the terminal quadratic differential of hh^{\prime} corresponding to ψ-\psi so that hh^{\prime} is expressed as ω=KImωiReω\omega^{\prime}=K\operatorname{Im}\omega-i\operatorname{Re}\omega with a natural parameter ω\omega^{\prime} of ψ\psi^{\prime} around ι(p)=h(f(p))\iota(p)=h^{\prime}(f(p)), then ι\iota is given by ω=iKζ\omega^{\prime}=-iK\zeta near pp, which implies that φ=ι(ψ/K2)\varphi=\iota^{*}(-\psi^{\prime}/K^{2}). Since Rι(R0)=Rh(f(R0))R\setminus\iota(R_{0})=R\setminus h^{\prime}(f(R_{0})) consists of finitely many horizontal arcs of ψ/K2-\psi^{\prime}/K^{2} together with finitely many points, from Corollary 4.7 we infer that (R,ι)(R^{\prime},\iota) is a closed φ\varphi-regular self-welding continuation of RR. Finally, h:=(h)1h:=(h^{\prime})^{-1} is a Teichmüller quasiconformal homeomorphism of RR^{\prime} onto RR with f=hιf=h\circ\iota, and ψ/K2-\psi^{\prime}/K^{2} is an initial quadratic differential of hh. ∎

Remark.

Bourque [12, Remark 3.8] gives a similar decomposition though the order of the factors is opposite. Our factorization fits with the following arguments.

The following corollary is an immediate consequence of Lemma 8.3 and Corollary 5.10.

Corollary 8.4.

Let R0R_{0} be a nonanalytically finite open Riemann surface of genus gg. Then initial quadratic differentials of Teichmüller quasiconformal embeddings of R0R_{0} into closed Riemann surfaces of genus gg belong to AL(R0)A_{L}(R_{0}).

Let 𝑹0{\bm{R}}_{0} be a marked finite open Riemann surface, and let 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g}. A Teichmüller quasiconformal embedding of 𝑹0{\bm{R}}_{0} into 𝑹\bm{R} is, by definition, an element 𝒇QCEmbhc(𝑹0,𝑹)\bm{f}\in\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) for which f:R0Rf:R_{0}\to R is a Teichmüller quasiconformal embedding for some (R0,θ0)𝑹0(R_{0},\theta_{0})\in{\bm{R}}_{0}, (R,θ)𝑹(R,\theta)\in{\bm{R}} and (f,θ0,θ)𝒇(f,\theta_{0},\theta)\in{\bm{f}}. This definition does depend on a particular choice of representatives of 𝑹0{\bm{R}}_{0}, 𝑹\bm{R} and 𝒇\bm{f}. Any initial and terminal quadratic differentials φ\varphi and ψ\psi of ff determine well-defined elements 𝝋=[φ,θ0]A+(𝑹0){\bm{\varphi}}=[\varphi,\theta_{0}]\in A_{+}({\bm{R}}_{0}) and 𝝍=[ψ,θ]A(𝑹){\bm{\psi}}=[\psi,\theta]\in A({\bm{R}}), which will be referred to as initial and terminal quadratic differentials of 𝒇\bm{f}, respectively.

The next lemma follows at once from Lemma 8.3 and Corollary 8.4.

Lemma 8.5.

Let 𝐑0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface of genus gg, and let 𝐑𝔗g{\bm{R}}\in\mathfrak{T}_{g}. If 𝐟\bm{f} is a Teichmüller quasiconformal embedding of 𝐑0{\bm{R}}_{0} into 𝐑\bm{R} with initial quadratic differential 𝛗\bm{\varphi}, then there are a closed 𝛗\bm{\varphi}-regular self-welding continuation (𝐑,𝛊)({\bm{R}}^{\prime},{\bm{\iota}}) of 𝐑0{\bm{R}}_{0} and a Teichmüller quasiconformal homeomorphism 𝐡\bm{h} of 𝐑{\bm{R}}^{\prime} onto 𝐑\bm{R} such that

  • (i)

    the co-welder of 𝝋\bm{\varphi} is an initial quadratic differential of 𝒉\bm{h}, and

  • (ii)

    𝒇=𝒉𝜾{\bm{f}}={\bm{h}}\circ{\bm{\iota}}.

Moreover, 𝛗\bm{\varphi} belongs to AL(𝐑0)A_{L}({\bm{R}}_{0}).

If 𝑹𝔗g𝔐(𝑹0){\bm{R}}\in\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}), then QCEmbhc(𝑹0,𝑹)\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}})\neq\varnothing even though CEmbhc(𝑹0,𝑹)=\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}})=\varnothing. A homotopically consistent quasiconformal embedding of 𝑹0{\bm{R}}_{0} into 𝑹\bm{R} is said to be extremal if it has the smallest maximal dilatation in QCEmbhc(𝑹0,𝑹)\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}).

Proposition 8.6 (Ioffe [25, Theorem 0.1], Bourque [12, Theorem 3.11]).

Assume that 𝐑0{\bm{R}}_{0} is a marked finite open Riemann surface of genus gg. If 𝐑{\bm{R}} belongs to 𝔗g𝔐(𝐑0)\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}), then QCEmbhc(𝐑0,𝐑)\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) has an extremal element, which is a Teichmüller quasiconformal embedding.

Remark.

In [25, Theorem 0.1] Ioffe claimed that QCEmbhc(𝑹0,𝑹)\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) has a unique extremal element. Bourque has disproved the uniqueness by counterexamples in [12, §3.3]. The error affects the uniqueness assertion in Earle-Marden [14].

With the aid of Lemma 8.5 and Proposition 7.7 it is easy to prove Proposition 8.1.

Proof of Proposition 8.1.

Let 𝑹𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}). Take a sequence {𝑹n}\{{\bm{R}}_{n}\} in 𝔗g𝔐(𝑹0)\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}) converging to 𝑹\bm{R}, and choose an extremal 𝒇nQCEmbhc(𝑹0,𝑹n){\bm{f}}_{n}\in\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}_{n}) for each nn. It is a Teichmüller quasiconformal embedding by Proposition 8.6. By Lemma 8.5 there are a closed regular self-welding continuation (𝑹n,𝜾n)({\bm{R}}^{\prime}_{n},{\bm{\iota}}^{\prime}_{n}) of 𝑹0{\bm{R}}_{0} and a Teichmüller quasiconformal homeomorphism 𝒉n{\bm{h}}^{\prime}_{n} of 𝑹n{\bm{R}}^{\prime}_{n} onto 𝑹n{\bm{R}}_{n} such that 𝒇n=𝒉n𝜾n{\bm{f}}_{n}={\bm{h}}^{\prime}_{n}\circ{\bm{\iota}}^{\prime}_{n}. Since 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}), we can find 𝜿CEmbhc(𝑹0,𝑹)\bm{\kappa}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}). If 𝒉n{\bm{h}}_{n} is the Teichmüller quasiconformal homeomorphism of 𝑹\bm{R} onto 𝑹n{\bm{R}}_{n}, then 𝒉n𝜿QCEmbhc(𝑹0,𝑹n){\bm{h}}_{n}\circ{\bm{\kappa}}\in\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}_{n}). The fact that 𝒇n{\bm{f}}_{n} is extremal yields

e2dT(𝑹n,𝑹)=K(𝒉n)=K(𝒇n)K(𝒉n𝜿)=K(𝒉n)=e2dT(𝑹n,𝑹)1e^{2d_{T}({\bm{R}}^{\prime}_{n},{\bm{R}})}=K({\bm{h}}^{\prime}_{n})=K({\bm{f}}_{n})\leqq K({\bm{h}}_{n}\circ{\bm{\kappa}})=K({\bm{h}}_{n})=e^{2d_{T}({\bm{R}}_{n},{\bm{R}})}\to 1

as nn\to\infty. Therefore, Proposition 7.7 guarantees the existence of a closed regular self-welding continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) of 𝑹0{\bm{R}}_{0}, as desired. ∎

In general, let 𝑹0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface of genus gg. Two elements 𝝋\bm{\varphi} and 𝝋{\bm{\varphi}}^{\prime} of AL(𝑹0)A_{L}({\bm{R}}_{0}) are said to be projectively equivalent to each other if 𝝋=r𝝋{\bm{\varphi}}^{\prime}=r{\bm{\varphi}} for some r>0r>0. The set of projective equivalence classes [𝝋][{\bm{\varphi}}] is denoted by PAL(𝑹0)PA_{L}({\bm{R}}_{0}). Now, Theorem 1.4 (i) asserts that any point 𝑹\bm{R} on the boundary 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) is obtained as a closed 𝝋{\bm{\varphi}}-regular self-welding continuation of 𝑹0{\bm{R}}_{0} for some 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}) and that each 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}) gives rise to a point 𝑹\bm{R} on 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) through a closed 𝝋\bm{\varphi}-regular self-welding continuation of 𝑹0{\bm{R}}_{0}. Recall that projectively equivalent elements of AL(𝑹0)A_{L}({\bm{R}}_{0}) induce the same closed regular self-welding continuation of 𝑹0{\bm{R}}_{0} (see the paragraph precedent to Example 5.5). Examples 5.4 and 5.5 show that the correspondence 𝔐(𝑹0)𝑹[𝝋]PAL(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0})\ni{\bm{R}}\leftrightarrow[{\bm{\varphi}}]\in PA_{L}({\bm{R}}_{0}) is, in general, many-to-many.

Example 8.7.

We consider the case of genus one, and continue to use the notations in Example 6.7. The mapping t𝑻τ(t)t\mapsto{\bm{T}}_{\tau(t)} is a bijection of (1,1](-1,1] onto 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) by [56, Theorem 5]. Thus the above mentioned correspondence between 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) and PAL(𝑹0)PA_{L}({\bm{R}}_{0}) is in fact a bijection in the case where g=1g=1 (see Example 10.6 below).

9 Ioffe rays

Let 𝑹0{\bm{R}}_{0} be a marked open Riemann surface of genus gg. For K1K\geqq 1 we denote by 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) the set of 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g} into which 𝑹0{\bm{R}}_{0} can be mapped by a homotopically consistent KK-quasiconformal embedding. In particular, we have 𝔐1(𝑹0)=𝔐(𝑹0)\mathfrak{M}_{1}({\bm{R}}_{0})=\mathfrak{M}({\bm{R}}_{0}) since 11-quasiconformal embeddings are conformal embeddings. If KKK\leqq K^{\prime}, then 𝔐K(𝑹0)𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0})\subset\mathfrak{M}_{K^{\prime}}({\bm{R}}_{0}). Also, we have K1𝔐K(𝑹0)=𝔗g\bigcup_{K\geqq 1}\mathfrak{M}_{K}({\bm{R}}_{0})=\mathfrak{T}_{g}. The next proposition claims that 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) is the closed (logK)/2(\log K)/2-neighborhood of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}).

Proposition 9.1.

Let 𝐑𝔗g\bm{R}\in\mathfrak{T}_{g} and K1K\geqq 1. Then 𝐑𝔐K(𝐑0){\bm{R}}\in\mathfrak{M}_{K}({\bm{R}}_{0}) if and only if

(9.1) dT(𝑹,𝔐(𝑹0))1 2logK.d_{T}({\bm{R}},\mathfrak{M}({\bm{R}}_{0}))\leqq\frac{1}{\,2\,}\log K.
Proof.

Let 𝑹=[R,θ]𝔗g{\bm{R}}=[R,\theta]\in\mathfrak{T}_{g}, and suppose that there is 𝒇=[f,θ0,θ]QCEmbhc(𝑹0,𝑹){\bm{f}}=[f,\theta_{0},\theta]\in\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) with K(𝒇)KK({\bm{f}})\leqq K, where 𝑹0=[R0,θ0]{\bm{R}}_{0}=[R_{0},\theta_{0}]. Take a quasiconformal homeomorphism hh of RR onto a closed Riemann surface RR^{\prime} such that μh=μf1\mu_{h}=\mu_{f^{-1}} on f(R0)f(R_{0}) and μh=0\mu_{h}=0 on Rf(R0)R\setminus f(R_{0}). Setting 𝑹=[R,hfθ0]{\bm{R}}^{\prime}=[R^{\prime},h\circ f\circ\theta_{0}] and 𝒉=[hf,θ0,hfθ0]{\bm{h}}^{\prime}=[h\circ f,\theta_{0},h\circ f\circ\theta_{0}], we obtain a closed continuation (𝑹,𝒉)({\bm{R}}^{\prime},{\bm{h}}^{\prime}) of 𝑹0{\bm{R}}_{0}. Thus 𝑹𝔐(𝑹0){\bm{R}}^{\prime}\in\mathfrak{M}({\bm{R}}_{0}) and hence

dT(𝑹,𝔐(𝑹0))dT(𝑹,𝑹)1 2logK(𝒉)=1 2logK(𝒇)1 2logK.d_{T}({\bm{R}},\mathfrak{M}({\bm{R}}_{0}))\leqq d_{T}({\bm{R}},{\bm{R}}^{\prime})\leqq\frac{1}{\,2\,}\log K({\bm{h}})=\frac{1}{\,2\,}\log K({\bm{f}})\leqq\frac{1}{\,2\,}\log K.

Conversely, assume that inequality (9.1) holds. Let 𝑹{\bm{R}}^{\prime} be a point of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) nearest to 𝑹\bm{R} (see Corollary 7.6), and take 𝜾CEmbhc(𝑹0,𝑹){\bm{\iota}}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}^{\prime}). Let 𝒉\bm{h} be the Teichmüller quasiconformal homeomorphism of 𝑹{\bm{R}}^{\prime} onto 𝑹\bm{R}. Then 𝒉𝜾{\bm{h}}\circ{\bm{\iota}} belongs to QCEmbhc(𝑹0,𝑹)\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}), and its maximal dilatation satisfies

logK(𝒉𝜾)logK(𝒉)=2dT(𝑹,𝑹)=2dT(𝑹,𝔐(𝑹0))logK.\log K({\bm{h}}\circ{\bm{\iota}})\leqq\log K({\bm{h}})=2d_{T}({\bm{R}},{\bm{R}}^{\prime})=2d_{T}({\bm{R}},\mathfrak{M}({\bm{R}}_{0}))\leqq\log K.

Consequently, 𝑹\bm{R} lies in 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}). ∎

For 𝑹𝔗g{\bm{R}}\in\mathfrak{T}_{g} and ρ>0\rho>0 let 𝔅ρ(𝑹)\mathfrak{B}_{\rho}({\bm{R}}) denote the open ball of radius ρ\rho centered at 𝑹\bm{R}:

𝔅ρ(𝑹)={𝑺𝔗gdT(𝑺,𝑹)<ρ}.\mathfrak{B}_{\rho}({\bm{R}})=\{{\bm{S}}\in\mathfrak{T}_{g}\mid d_{T}({\bm{S}},{\bm{R}})<\rho\}.

Its closure will be denoted by 𝔅¯ρ(𝑹)\bar{\mathfrak{B}}_{\rho}({\bm{R}}), which is a compact subset of 𝔗g\mathfrak{T}_{g}. For the sake of convenience we define 𝔅¯0(𝑹)={𝑹}\bar{\mathfrak{B}}_{0}({\bm{R}})=\{{\bm{R}}\}.

Corollary 9.2.

𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) is compact for all K1K\geqq 1.

Proof.

It follows from Corollary 7.6 that 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is included in 𝔅¯ρ(𝑹)\bar{\mathfrak{B}}_{\rho}({\bm{R}}) for some 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}) and ρ>0\rho>0. Proposition 9.1 then implies that 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) is a closed subset of the compact set 𝔅¯ρK(𝑹)\bar{\mathfrak{B}}_{\rho_{K}}({\bm{R}}), where ρK=ρ+(logK)/2\rho_{K}=\rho+(\log K)/2, and hence is compact. ∎

Definition 9.3 (Ioffe ray).

Let 𝑹0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface. An Ioffe ray of 𝑹0{\bm{R}}_{0} is, by definition, a Teichmüller geodesic ray of the form 𝒓𝑹[𝝍]{\bm{r}}_{\bm{R}}[{\bm{\psi}}], where 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}) and 𝝍A(𝑹){\bm{\psi}}\in A({\bm{R}}) such that for some 𝜾CEmbhc(𝑹0,𝑹){\bm{\iota}}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) and 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}) the continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) is a closed 𝝋\bm{\varphi}-regular self-welding continuation of 𝑹0{\bm{R}}_{0} and that 𝝍\bm{\psi} is the co-welder of 𝝋\bm{\varphi}. Let (𝑹0)\mathscr{I}({\bm{R}}_{0}) denote the set of Ioffe rays of 𝑹0{\bm{R}}_{0}.

If 𝑹0{\bm{R}}_{0} is a marked analytically finite open Riemann surface, then 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) consists of exactly one point, say 𝑹\bm{R}. We call each Teichmüller geodesic ray emanating from 𝑹\bm{R} an Ioffe ray of 𝑹0{\bm{R}}_{0}.

Remark.

By Theorem 1.4 (i) the initial point of each Ioffe ray of 𝑹0{\bm{R}}_{0} lies on the boundary 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) and any boundary point of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is the initial point of some Ioffe ray of 𝑹0{\bm{R}}_{0}. Note that two or more Ioffe rays of 𝑹0{\bm{R}}_{0} can have a common initial point. See the paragraph precedent to Example 8.7.

One of the purposes of the present section is to establish the following theorem. As an application, we give an alternative proof of Bourque [12, Theorem 3.13] in the case where the target is a closed Riemann surface of the same genus as the domain surface of quasiconformal embeddings; it should be noted that Theorem 9.4 follows easily from the theorem of Bourque. Geometric aspects of Ioffe rays should be emphasized.

Theorem 9.4.

Let 𝐑0{\bm{R}}_{0} be a marked finite open Riemann surface of genus gg. Then the correspondence

(𝑹0)×(0,+)(𝒓,t)𝒓(t)𝔗g𝔐(𝑹0)\mathscr{I}({\bm{R}}_{0})\times(0,+\infty)\ni({\bm{r}},t)\mapsto{\bm{r}}(t)\in\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0})

is bijective.

Biernacki [7] called a plane domain DD linearly accessible in the strict sense if the compliment D\mathbb{C}\setminus D can be expressed as a union of mutually disjoint half lines except that the endpoint of one half line can lie on another half line. Lewandowski [31, 32] showed that the class of linearly accessible domains in the strict sense is precisely the class of close-to-convex domains of Kaplan [29]. Following these terminologies, we may say that 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is linearly accessible in the strict sense or close-to-convex.

Theorem 9.4 implies in particular that each Ioffe ray never hits 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) again after departure. We prove the theorem step by step. We begin with the following proposition. Theorem 1.4 (ii) is a corollary to the proposition.

Proposition 9.5.

Let 𝐑0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface of genus gg, and let 𝐑𝔐(𝐑0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}). Take 𝛗AL(𝐑0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}) so that 𝐑\bm{R} is obtained via a closed 𝛗\bm{\varphi}-regular self-welding continuation of 𝐑0{\bm{R}}_{0} and let 𝛙A(𝐑){\bm{\psi}}\in A({\bm{R}}) be the co-welder of 𝛗\bm{\varphi}. Then any element of CEmbhc(𝐑0,𝐑)\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) defines a closed 𝛗\bm{\varphi}-regular self-welding continuation of 𝐑0{\bm{R}}_{0} for which 𝛙\bm{\psi} is the co-welder of 𝛗\bm{\varphi}.

Proof.

Let (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) be a closed 𝝋\bm{\varphi}-regular self-welding continuation of 𝑹0{\bm{R}}_{0} with 𝝋=𝜾𝝍{\bm{\varphi}}={\bm{\iota}}^{*}{\bm{\psi}}. For any closed continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}^{\prime}) of 𝑹0{\bm{R}}_{0} let 𝝋{\bm{\varphi}}^{\prime} be the (𝑹,𝜾)({\bm{R}},{\bm{\iota}}^{\prime})-zero-extension of 𝝋\bm{\varphi}. It then follows from Lemma 6.5 that for any γ𝒮(Σg)\gamma\in\mathscr{S}(\Sigma_{g}) the inequality 𝑹(𝝍)(γ)𝑹(𝝋)(γ)\mathcal{H}_{\bm{R}}({\bm{\psi}})(\gamma)\leqq{\mathcal{H}}^{\prime}_{\bm{R}}({\bm{\varphi}}^{\prime})(\gamma) holds. Proposition 6.4 then implies that 𝝍𝑹𝝋𝑹\|{\bm{\psi}}\|_{\bm{R}}\leqq\|{\bm{\varphi}}^{\prime}\|_{\bm{R}}. Actually, the sign of equality occurs because 𝝋𝑹=𝝋𝑹0=𝝍𝑹\|{\bm{\varphi}}^{\prime}\|_{\bm{R}}=\|{\bm{\varphi}}\|_{{\bm{R}}_{0}}=\|{\bm{\psi}}\|_{\bm{R}}. Another application of Proposition 6.4 gives us 𝝋=𝝍{\bm{\varphi}}^{\prime}={\bm{\psi}} almost everywhere on 𝑹\bm{R}. In particular, (𝑹,𝜾)({\bm{R}},{\bm{\iota}}^{\prime}) is a dense continuation of 𝑹0{\bm{R}}_{0}. Since 𝝋=(𝜾)𝝍{\bm{\varphi}}=({\bm{\iota}}^{\prime})^{*}{\bm{\psi}}, Proposition 6.1 implies that (𝑹,𝜾)({\bm{R}},{\bm{\iota}}^{\prime}) is a closed 𝝋\bm{\varphi}-regular self-welding continuation of 𝑹0{\bm{R}}_{0} for which 𝝍\bm{\psi} is the co-welder of 𝝋\bm{\varphi}. ∎

Remark.

One might surmise that Proposition 9.5 assures us that CEmbhc(𝑹0,𝑹)\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) is a singleton. This is not always the case, however, as Bourque [12, §3.3] pointed out.

Proposition 9.6.

Suppose that 𝐑0{\bm{R}}_{0} is a marked finite open Riemann surface of genus gg. Let 𝐒𝔗g𝔐(𝐑0){\bm{S}}\in\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}). If 𝐑\bm{R} is a point of 𝔐(𝐑0)\mathfrak{M}({\bm{R}}_{0}) nearest to 𝐒\bm{S}, then 𝐫𝐑[𝐒]{\bm{r}}_{\bm{R}}[{\bm{S}}] is an Ioffe ray of 𝐑0{\bm{R}}_{0}.

Proof.

We may assume that 𝑹0{\bm{R}}_{0} is nonanalytically finite. Take 𝜾CEmbhc(𝑹0,𝑹){\bm{\iota}}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}). Let 𝒉\bm{h} be the Teichmüller quasiconformal homeomorphism of 𝑹\bm{R} onto 𝑺\bm{S}, and set 𝒇=𝒉𝜾{\bm{f}}={\bm{h}}\circ{\bm{\iota}}. If 𝒇QCEmbhc(𝑹0,𝑺){\bm{f}}^{\prime}\in\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{S}}), then Proposition 9.1 yields that

logK(𝒇)2dT(𝑺,𝔐(𝑹0))=2dT(𝑺,𝑹)=logK(𝒉)logK(𝒇).\log K({\bm{f}}^{\prime})\geqq 2d_{T}({\bm{S}},\mathfrak{M}({\bm{R}}_{0}))=2d_{T}({\bm{S}},{\bm{R}})=\log K({\bm{h}})\geqq\log K({\bm{f}}).

This shows that 𝒇\bm{f} is extremal in QCEmbhc(𝑹0,𝑺)\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{S}}), and hence is a Teichmüller quasiconformal embedding by Proposition 8.6.

Let 𝝋0AL(𝑹0){\bm{\varphi}}_{0}\in A_{L}({\bm{R}}_{0}) and 𝝍A(𝑺){\bm{\psi}}\in A({\bm{S}}) be initial and terminal quadratic differentials of 𝒇\bm{f}, respectively (see Corollary 8.4). Since 𝑹\bm{R} is a boundary point of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}), Proposition 9.5 implies that (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) is a regular self-welding continuation of 𝑹0{\bm{R}}_{0}. To show that 𝝋0\bm{\varphi}_{0} is a welder of the continuation take representatives (R0,θ0)𝑹0(R_{0},\theta_{0})\in{\bm{R}}_{0}, (R,θ)𝑹(R,\theta)\in{\bm{R}}, (S,η)𝑺(S,\eta)\in{\bm{S}}, (ι,θ0,θ)𝜾(\iota,\theta_{0},\theta)\in{\bm{\iota}}, (f,θ0,η)𝒇(f,\theta_{0},\eta)\in{\bm{f}}, (h,θ,η)𝒉(h,\theta,\eta)\in{\bm{h}}, (φ0,θ0)𝝋0(\varphi_{0},\theta_{0})\in{\bm{\varphi}}_{0} and (ψ,η)𝝍(\psi,\eta)\in{\bm{\psi}} so that f=hιf=h\circ\iota. The last identity shows that an initial quadratic differential φA(R)\varphi\in A(R) of hh satisfies φ¯/φ=ιφ0¯/ιφ0\bar{\varphi}/\varphi=\overline{\iota_{*}\varphi_{0}}/\iota_{*}\varphi_{0}. Thus the meromorphic function (ιφ0)/φ(\iota_{*}\varphi_{0})/\varphi on ι(R0)\iota(R_{0}) is real and hence constant. Since hh is a uniform stretch along horizontal trajectories of both ιφ0\iota_{*}\varphi_{0} and φ\varphi, the constant must be positive and hence may be supposed to be 11. Then we obtain φ0=ιφ\varphi_{0}=\iota^{*}\varphi. As ι(R0)\iota(R_{0}) is dense in RR, in view of Proposition 6.1 we see that (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) is a closed 𝝋0{\bm{\varphi}}_{0}-regular self-welding continuation of 𝑹0{\bm{R}}_{0} and that 𝝋:=[φ,θ]{\bm{\varphi}}:=[\varphi,\theta] is the co-welder of 𝝋0{\bm{\varphi}}_{0}. It follows that 𝒓𝑹[𝝋]{\bm{r}}_{\bm{R}}[{\bm{\varphi}}] is an Ioffe ray of 𝑹0{\bm{R}}_{0}. Since 𝝋\bm{\varphi} is an initial quadratic differential of the Teichmüller quasiconformal homeomorphism 𝒉\bm{h} of 𝑹\bm{R} onto 𝑺\bm{S}, we infer that 𝒓𝑹[𝑺]=𝒓𝑹[𝝋]{\bm{r}}_{\bm{R}}[\bm{S}]={\bm{r}}_{\bm{R}}[{\bm{\varphi}}]. ∎

Lemma 9.7.

Let 𝔎\mathfrak{K} be a compact subset of 𝔗g\mathfrak{T}_{g} and let 𝐫\bm{r} be a Teichmüller geodesic ray with 𝐫(0)𝔎{\bm{r}}(0)\in\mathfrak{K}. If 𝔅ρ(𝐫(ρ))𝔎=\mathfrak{B}_{\rho}({\bm{r}}(\rho))\cap\mathfrak{K}=\varnothing for all ρ>0\rho>0, then 𝔅¯ρ(𝐫(ρ))𝔎={𝐫(0)}\bar{\mathfrak{B}}_{\rho}({\bm{r}}(\rho))\cap\mathfrak{K}=\{{\bm{r}}(0)\} for ρ>0\rho>0.

Proof.

It follows from (6.2) that 𝒓(0){\bm{r}}(0) belongs to 𝔅¯ρ(𝒓(ρ))𝔎\bar{\mathfrak{B}}_{\rho}({\bm{r}}(\rho))\cap\mathfrak{K}. Suppose that 𝑹\bm{R} also belongs to both 𝔅¯ρ(𝒓(ρ))\bar{\mathfrak{B}}_{\rho}({\bm{r}}(\rho)) and 𝔎\mathfrak{K}. Then

(9.2) dT(𝑹,𝒓(2ρ))dT(𝑹,𝒓(ρ))+dT(𝒓(ρ),𝒓(2ρ))ρ+ρ=2ρ,d_{T}({\bm{R}},{\bm{r}}(2\rho))\leqq d_{T}({\bm{R}},{\bm{r}}(\rho))+d_{T}({\bm{r}}(\rho),{\bm{r}}(2\rho))\leqq\rho+\rho=2\rho,

and hence 𝑹𝔅¯2ρ(𝒓(2ρ)){\bm{R}}\in\bar{\mathfrak{B}}_{2\rho}({\bm{r}}(2\rho)). Since 𝔅2ρ(𝒓(2ρ))𝔎=\mathfrak{B}_{2\rho}({\bm{r}}(2\rho))\cap\mathfrak{K}=\varnothing, we obtain 𝑹𝔅2ρ(𝒓(2ρ)){\bm{R}}\in\partial\mathfrak{B}_{2\rho}({\bm{r}}(2\rho)), or equivalently, dT(𝑹,𝒓(2ρ))=2ρd_{T}({\bm{R}},{\bm{r}}(2\rho))=2\rho. Therefore, the signs of equality occur in (9.2). Since the metric space (𝔗g,dT)(\mathfrak{T}_{g},d_{T}) is a straight space in the sense of Busemann (see Abikoff [1, Section (3.2)]), the three points 𝑹\bm{R}, 𝒓(ρ){\bm{r}}(\rho) and 𝒓(2ρ){\bm{r}}(2\rho) lie on the same Teichmüller geodesic line, which must include the Teichmüller geodesic ray 𝒓\bm{r}. It then follows from (9.2) that 𝑹=𝒓(0){\bm{R}}={\bm{r}}(0), which finishes the proof. ∎

Theorem 9.8.

Let 𝐑0{\bm{R}}_{0} be a marked finite open Riemann surface. If 𝐫(𝐑0){\bm{r}}\in\mathscr{I}({\bm{R}}_{0}), then

(9.3) 𝔅¯ρ(𝒓(ρ+τ))𝔐e2τ(𝑹0)={𝒓(τ)}\bar{\mathfrak{B}}_{\rho}({\bm{r}}(\rho+\tau))\cap\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0})=\{{\bm{r}}(\tau)\}

and

(9.4) dT(𝒓(t),𝔐e2τ(𝑹0))=max{tτ,0}d_{T}({\bm{r}}(t),\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}))=\max\{t-\tau,0\}

for all ρ>0\rho>0, τ0\tau\geqq 0 and t0t\geqq 0.

To prove Theorem 9.8 we need the following lemma, which implies that logExt\log\operatorname{Ext}_{\mathcal{F}} is a Lipschitz continuous function on 𝔗g\mathfrak{T}_{g}. For the proof see Gardiner [16, Lemma 4] or Gardiner-Lakic [18, Lemma 12.5].

Lemma 9.9.

For 𝐑1,𝐑2𝔗g{\bm{R}}_{1},{\bm{R}}_{2}\in\mathfrak{T}_{g} and (Σg){0}\mathcal{F}\in\mathscr{MF}(\Sigma_{g})\setminus\{0\},

logExt(𝑹1)2dT(𝑹1,𝑹2)logExt(𝑹2)logExt(𝑹1)+2dT(𝑹1,𝑹2).\log\operatorname{Ext}_{\mathcal{F}}({\bm{R}}_{1})-2d_{T}({\bm{R}}_{1},{\bm{R}}_{2})\leqq\log\operatorname{Ext}_{\mathcal{F}}({\bm{R}}_{2})\leqq\log\operatorname{Ext}_{\mathcal{F}}({\bm{R}}_{1})+2d_{T}({\bm{R}}_{1},{\bm{R}}_{2}).
Proof of Theorem 9.8.

We can write 𝒓=𝒓𝑹[𝝍]{\bm{r}}={\bm{r}}_{\bm{R}}[{\bm{\psi}}] with 𝝍A(𝑹){\bm{\psi}}\in A({\bm{R}}), where 𝝍\bm{\psi} is the co-welder of a welder of some closed regular self-welding continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) of 𝑹0{\bm{R}}_{0}. Set =𝑹(𝝍)\mathcal{F}=\mathcal{H}_{\bm{R}}({\bm{\psi}}).

To prove (9.3) let 𝑺𝔗g{\bm{S}}\in\mathfrak{T}_{g}. If 𝑹{\bm{R}}^{\prime} is a point of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) nearest to 𝑺\bm{S}, then Lemma 9.9 and Theorem 6.3 imply that

logExt(𝑺)logExt(𝑹)+2dT(𝑺,𝑹)logExt(𝑹)+2dT(𝑺,𝔐(𝑹0)).\log\operatorname{Ext}_{\mathcal{F}}({\bm{S}})\leqq\log\operatorname{Ext}_{\mathcal{F}}({\bm{R}}^{\prime})+2d_{T}({\bm{S}},{\bm{R}}^{\prime})\leqq\log\operatorname{Ext}_{\mathcal{F}}({\bm{R}})+2d_{T}({\bm{S}},\mathfrak{M}({\bm{R}}_{0})).

Consider the Teichmüller geodesic ray 𝒓{\bm{r}}^{\prime} defined by 𝒓(t)=𝒓(t+τ){\bm{r}}^{\prime}(t)={\bm{r}}(t+\tau), t0t\geqq 0. Then another application of Lemma 9.9 together with (6.4) yields that

logExt(𝑺)\displaystyle\log\operatorname{Ext}_{\mathcal{F}}({\bm{S}}) logExt(𝒓(ρ+τ))2dT(𝑺,𝒓(ρ+τ))\displaystyle\geqq\log\operatorname{Ext}_{\mathcal{F}}({\bm{r}}(\rho+\tau))-2d_{T}({\bm{S}},{\bm{r}}(\rho+\tau))
=logExt(𝑹)+2(ρ+τ)2dT(𝑺,𝒓(ρ)).\displaystyle=\log\operatorname{Ext}_{\mathcal{F}}({\bm{R}})+2(\rho+\tau)-2d_{T}({\bm{S}},{\bm{r}}^{\prime}(\rho)).

With the aid of Proposition 9.1 we thus obtain

dT(𝑺,𝔐e2τ(𝑹0))dT(𝑺,𝔐(𝑹0))τρdT(𝑺,𝒓(ρ)).d_{T}({\bm{S}},\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}))\geqq d_{T}({\bm{S}},\mathfrak{M}({\bm{R}}_{0}))-\tau\geqq\rho-d_{T}({\bm{S}},{\bm{r}}^{\prime}(\rho)).

Therefore, 𝔅ρ(𝒓(ρ))𝔐e2τ(𝑹0)=\mathfrak{B}_{\rho}({\bm{r}}^{\prime}(\rho))\cap\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0})=\varnothing for all ρ>0\rho>0. Since 𝔐e2τ(𝑹0)\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}) is compact by Corollary 9.2 and contains 𝒓(0)=𝒓(τ){\bm{r}}^{\prime}(0)={\bm{r}}(\tau), identity (9.3) follows at once from Lemma 9.7.

To show (9.4) we first remark that if 0tτ0\leqq t\leqq\tau, then

dT(𝒓(t),𝔐(𝑹0))dT(𝒓(t),𝒓(0))=tτd_{T}({\bm{r}}(t),\mathfrak{M}({\bm{R}}_{0}))\leqq d_{T}({\bm{r}}(t),{\bm{r}}(0))=t\leqq\tau

as 𝒓(0)=𝑹𝔐(𝑹0){\bm{r}}(0)={\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}). Therefore, by Proposition 9.1 we know that 𝒓(t)𝔐e2τ(𝑹0){\bm{r}}(t)\in\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}), or equivalently, dT(𝒓(t),𝔐e2τ(𝑹0))=0d_{T}({\bm{r}}(t),\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}))=0. If t>τt>\tau, then

𝔅¯tτ(𝒓(t))𝔐e2τ(𝑹0)={𝒓(τ)}\bar{\mathfrak{B}}_{t-\tau}({\bm{r}}(t))\cap\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0})=\{{\bm{r}}(\tau)\}

by (9.3), which implies that dT(𝒓(t),𝔐e2τ(𝑹0))=tτd_{T}({\bm{r}}(t),\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}))=t-\tau. ∎

Corollary 9.10.

If 𝐫(𝐑0){\bm{r}}\in\mathscr{I}({\bm{R}}_{0}) and t>τ0t>\tau\geqq 0, then 𝐫(t)𝔗g𝔐e2τ(𝐑0){\bm{r}}(t)\in\mathfrak{T}_{g}\setminus\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}).

Proof.

By (9.4) we have dT(𝒓(t),𝔐e2τ(𝑹0))>0d_{T}({\bm{r}}(t),\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}))>0. Hence 𝒓(t)𝔐e2τ(𝑹0){\bm{r}}(t)\not\in\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}). ∎

Corollary 9.11.

Let K1K\geqq 1 and 𝐑𝔗g{\bm{R}}\in\mathfrak{T}_{g}. Then a point of 𝔐K(𝐑0)\mathfrak{M}_{K}({\bm{R}}_{0}) nearest to 𝐑\bm{R} is uniquely determined.

Proof.

We have only to consider the case 𝑹𝔐K(𝑹0){\bm{R}}\not\in\mathfrak{M}_{K}({\bm{R}}_{0}). By Proposition 9.6 there is 𝒓(𝑹0){\bm{r}}\in\mathscr{I}({\bm{R}}_{0}) for which 𝒓(t0)=𝑹{\bm{r}}(t_{0})={\bm{R}} for some t00t_{0}\geqq 0. Since 𝑹\bm{R} does not belong to the compact set 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}), it follows from (9.4) that t0>(logK)/2t_{0}>(\log K)/2. Setting t1=t0(logK)/2t_{1}=t_{0}-(\log K)/2, we then infer from (9.3) that 𝔅¯t1(𝑹)𝔐K(𝑹0)\bar{\mathfrak{B}}_{t_{1}}({\bm{R}})\cap\mathfrak{M}_{K}({\bm{R}}_{0}) is a singleton, which proves the corollary. ∎

We are now ready to prove Theorem 9.4

Proof of Theorem 9.4.

If (𝒓,t)(𝑹0)×(0,+)({\bm{r}},t)\in\mathscr{I}({\bm{R}}_{0})\times(0,+\infty), then 𝒓(t)𝔗g𝔐(𝑹0){\bm{r}}(t)\in\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}) by Corollary 9.10. Thus (𝒓,t)𝒓(t)({\bm{r}},t)\mapsto{\bm{r}}(t) defines a mapping of (𝑹0)×(0,+)\mathscr{I}({\bm{R}}_{0})\times(0,+\infty) into 𝔗g𝔐(𝑹0)\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}).

Let 𝑺𝔗g𝔐(𝑹0){\bm{S}}\in\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}). As we already know that 𝑺\bm{S} lies on some Ioffe ray 𝒓\bm{r} (see Proposition 9.6), we need to show the uniqueness. Set 𝑹=𝒓(0){\bm{R}}={\bm{r}}(0) and ρ=dT(𝑺,𝔐(𝑹0))\rho=d_{T}({\bm{S}},\mathfrak{M}({\bm{R}}_{0})). If 𝒓(t0)=𝑺{\bm{r}}(t_{0})={\bm{S}}, then ρ=dT(𝒓(t0),𝔐(𝑹0))=t0\rho=d_{T}({\bm{r}}(t_{0}),\mathfrak{M}({\bm{R}}_{0}))=t_{0} by (9.4). Therefore, by (9.3) we have

{𝑹}={𝒓(0)}=𝔅¯ρ(𝒓(ρ))𝔐(𝑹0)=𝔅¯d(𝑺)(𝑺)𝔐(𝑹0),\{{\bm{R}}\}=\{{\bm{r}}(0)\}=\bar{\mathfrak{B}}_{\rho}({\bm{r}}(\rho))\cap\mathfrak{M}({\bm{R}}_{0})=\bar{\mathfrak{B}}_{d({\bm{S}})}({\bm{S}})\cap\mathfrak{M}({\bm{R}}_{0}),

where d(𝑺)=dT(𝑺,𝔐(𝑹0))d({\bm{S}})=d_{T}({\bm{S}},\mathfrak{M}({\bm{R}}_{0})). Hence 𝑹\bm{R} is determined solely by 𝑺\bm{S}. Consequently, 𝑺\bm{S} lies on exactly one Ioffe ray since 𝒓=𝒓𝑹[𝑺]{\bm{r}}={\bm{r}}_{\bm{R}}[{\bm{S}}]. ∎

The following proposition describes the boundary and the exterior of 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) in terms of Ioffe rays.

Proposition 9.12.

If 𝐑0{\bm{R}}_{0} is a marked finite open Riemann surface of genus gg, then

𝔐e2τ(𝑹0)\displaystyle\partial\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}) ={𝒓(τ)𝒓(𝑹0)},and\displaystyle=\{{\bm{r}}(\tau)\mid{\bm{r}}\in\mathscr{I}({\bm{R}}_{0})\},\quad\text{and}
𝔗g𝔐e2τ(𝑹0)\displaystyle\mathfrak{T}_{g}\setminus\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}) ={𝒓(t)𝒓(𝑹0),t>τ}.\displaystyle=\{{\bm{r}}(t)\mid{\bm{r}}\in\mathscr{I}({\bm{R}}_{0}),t>\tau\}.

for τ0\tau\geqq 0.

Proof.

By the definition of Ioffe rays the initial points of Ioffe rays lie on 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}), and Theorem 1.4 (i) implies that any point on 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) is the initial point of some Ioffe ray. Hence, with the aid of Theorem 9.4 we know that the proposition is valid for τ=0\tau=0.

Assume that τ>0\tau>0 and 𝒓(𝑹0){\bm{r}}\in\mathscr{I}({\bm{R}}_{0}). It follows from (9.3) that 𝒓(τ)𝔐e2τ(𝑹0){\bm{r}}(\tau)\in\partial\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}). Also, 𝒓(t)𝔗g𝔐e2τ(𝑹0){\bm{r}}(t)\in\mathfrak{T}_{g}\setminus\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}) for t>τt>\tau by Corollary 9.10. If 0t<τ0\leqq t<\tau, then 𝔅τt(𝒓(t))\mathfrak{B}_{\tau-t}({\bm{r}}(t)) is included in 𝔐e2τ(𝑹0)\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}) by Proposition 9.1 as dT(𝒓(t),𝔐(𝑹0))=td_{T}({\bm{r}}(t),\mathfrak{M}({\bm{R}}_{0}))=t by (9.4). Hence 𝒓(t){\bm{r}}(t) is an interior point of 𝔐e2τ(𝑹0)\mathfrak{M}_{e^{2\tau}}({\bm{R}}_{0}). Now, the proposition follows at once. ∎

We conclude the section with the following proposition due to Bourque. Recall that QCEmbhc(𝑹0,𝑹)\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) includes extremal elements for any 𝑹𝔗g𝔐(𝑹0){\bm{R}}\in\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}), which are Teichmüller quasiconformal embeddings. The proposition asserts that Teichmüller quasiconformal embeddings are extremal. Note that the uniqueness does not hold in general.

Proposition 9.13 (Bourque [12, Theorem 3.13]).

Let 𝐑0{\bm{R}}_{0} be a marked finite open Riemann surface of genus gg. Then for 𝐑𝔗gInt𝔐(𝐑0){\bm{R}}\in\mathfrak{T}_{g}\setminus\operatorname{Int}\mathfrak{M}({\bm{R}}_{0}) Teichmüller quasiconformal embeddings of 𝐑0{\bm{R}}_{0} into 𝐑\bm{R} are extremal in QCEmbhc(𝐑0,𝐑)\operatorname{QCEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) and have initial and terminal quadratic differentials in common.

Proof.

Let 𝑹𝔗gInt𝔐(𝑹0){\bm{R}}\in\mathfrak{T}_{g}\setminus\operatorname{Int}\mathfrak{M}({\bm{R}}_{0}), and take a Teichmüller quasiconformal embedding 𝒇\bm{f} of 𝑹0{\bm{R}}_{0} into 𝑹\bm{R} with initial quadratic differential 𝝋\bm{\varphi}. If 𝑹𝔐(𝑹){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}), then the proposition follows from Proposition 9.5.

If 𝑹𝔗g𝔐(𝑹0){\bm{R}}\in\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}), then by Lemma 8.5 we obtain a closed 𝝋\bm{\varphi}-regular self-welding continuation (𝑹,𝜾)({\bm{R}}^{\prime},{\bm{\iota}}) of 𝑹0{\bm{R}}_{0} and a Teichmüller quasiconformal homeomorphism 𝒉\bm{h} of 𝑹{\bm{R}}^{\prime} onto 𝑹\bm{R} such that the co-welder 𝝋{\bm{\varphi}}^{\prime} of 𝝋\bm{\varphi} is an initial quadratic differential of 𝒉\bm{h} and that 𝒇=𝒉𝜾{\bm{f}}={\bm{h}}\circ{\bm{\iota}}. Thus 𝑹\bm{R} lies on the Ioffe ray 𝒓:=𝒓𝑹[𝝋]{\bm{r}}^{\prime}:={\bm{r}}_{{\bm{R}}^{\prime}}[{\bm{\varphi}}^{\prime}], and if 𝑹=𝒓(t){\bm{R}}={\bm{r}}^{\prime}(t), then dT(𝑹,𝔐(𝑹0))=t=(logK(𝒉))/2=(logK(𝒇))/2d_{T}({\bm{R}},\mathfrak{M}({\bm{R}}_{0}))=t=(\log K({\bm{h}}))/2=(\log K({\bm{f}}))/2 by (9.4), and hence 𝒇\bm{f} is extremal by Proposition 9.12. Since 𝑹{\bm{R}}^{\prime} is uniquely determined by Corollary 9.11, it follows that 𝒉\bm{h} is also uniquely determined. Since 𝒇\bm{f} and 𝒉\bm{h} have a terminal quadratic differential in common, this completes the proof of the proposition. ∎

10 Uniqueness of closed regular self-weldings

In this section we study uniqueness of closed regular self-weldings of a compact bordered Riemann surface SS. In Example 5.4 we have remarked that some positive holomorphic quadratic differential on SS satisfying the border length condition on S\partial S can induce two or more inequivalent closed regular self-weldings of SS. With this in mind we make the following definition.

Definition 10.1 (exceptional quadratic differential).

Let CC be a union of connected components of S\partial S, and let φAL(S,C)\varphi\in A_{L}(S,C). If there are two CC-full φ\varphi-regular self-weldings of SS that are inequivalent to each other, then φ\varphi is called exceptional for CC.

It is then natural to ask which elements of AL(S,C)A_{L}(S,C) are exceptional for CC. To answer the question we introduce a subset AE(S,C)A_{E}(S,C) of AL(S,C)A_{L}(S,C) as follows.

Definition 10.2 (class AE(S,C)A_{E}(S,C)).

Let CC be a union of connected components of S\partial S. Denote by AE(S,C)A_{E}(S,C) the set of φAL(S,C)\varphi\in A_{L}(S,C) such that some component CC^{\prime} of CC includes four or more horizontal trajectories of φ\varphi and that

(10.1) Lφ(a)<Lφ(C)/2L_{\varphi}(a)<L_{\varphi}(C^{\prime})/2

for all horizontal trajectories aa of φ\varphi on CC^{\prime}. Set AE(S)=AE(S,S)A_{E}(S)=A_{E}(S,\partial S).

Clearly, AE(S)=CAE(S,C)A_{E}(S)=\bigcup_{C}A_{E}(S,C), where the union is taken over all components CC of S\partial S. The following theorem asserts in particular that for φAL(S)AE(S)\varphi\in A_{L}(S)\setminus A_{E}(S) there is exactly one closed φ\varphi-regular self-welding of SS up to equivalence.

Theorem 10.3.

Let SS be a compact bordered Riemann surface, and let CC be a union of components of S\partial S. Then an element φ\varphi of AL(S,C)A_{L}(S,C) is exceptional for CC if and only if it belongs to AE(S,C)A_{E}(S,C). If this is the case, then the family of equivalence classes of CC-full φ\varphi-regular self-weldings of SS has the cardinality of the continuum.

Proof.

We may suppose from the outset that CC is a connected component of S\partial S. Assume first that Lφ(a)=Lφ(C)/2L_{\varphi}(a)=L_{\varphi}(C)/2 for some horizontal trajectory aa of φ\varphi on CC. The endpoints of aa, which are zeros of φ\varphi, divide CC into two arcs a1a_{1} and a2a_{2} of the same φ\varphi-length, where we label them so that a1=aa_{1}^{\circ}=a. Let R,ι\langle R,\iota\rangle be a CC-full φ\varphi-regular self-welding of SS, and let ψ\psi denote the co-welder of φ\varphi. Since φ\varphi does not vanish on a1a_{1}^{\circ}, it follows from Proposition 5.6 that ιa1\iota_{*}a_{1} is a simple arc on the weld graph GιG_{\iota}. As the sum of the ψ\psi-lengths of edges of GιG_{\iota} is identical with Lφ(C)/2L_{\varphi}(C)/2, we deduce that ιa1\iota_{*}a_{1} exhausts GιG_{\iota}, or GιG_{\iota} is I-shaped. Therefore, R,ι\langle R,\iota\rangle is equivalent to the self-welding of SS with welder φ\varphi along (a1,a2)(a_{1},a_{2}). Hence φ\varphi is nonexceptional for CC.

In the rest of the proof we suppose that inequalities (10.1) with CC^{\prime} replaced with CC hold for all horizontal trajectories aa of φ\varphi on CC. Then CC carries at least three horizontal trajectories of φ\varphi and hence the set ZZ of zeros of φ\varphi on CC contains more than two points.

Assume that cardZ=3\operatorname{card}Z=3 so that Z={p1,p2,p3}Z=\{p_{1},p_{2},p_{3}\}, say. The points of ZZ divide CC into three arcs a1a_{1}, a2a_{2} and a3a_{3}, where they are labeled so that pkakp_{k}\not\in a_{k} for k=1,2,3k=1,2,3. Let R,ι\langle R,\iota\rangle be a CC-full φ\varphi-regular self-welding of SS, and let GιG_{\iota} be its weld graph. Since aka_{k}^{\circ} contains no zeros of φ\varphi and the self-welding is φ\varphi-regular, it follows from (10.1) that the points vk:=ι(pk)v_{k}:=\iota(p_{k}), k=1,2,3k=1,2,3, are the end-vertices of GιG_{\iota} by Proposition 5.6, and hence GιG_{\iota} is Y-shaped. Lemma 5.11 assures us that R,ι\langle R,\iota\rangle is determined up to equivalence. This means that φ\varphi is nonexceptional for CC.

Suppose next that cardZ>3\operatorname{card}Z>3. We consider two cases. If ZZ contains two points p1p_{1} and p2p_{2} that divide CC into two arcs a1a_{1} and a2a_{2} of the same φ\varphi-length Lφ(C)/2L_{\varphi}(C)/2, then the self-welding R,ι\langle R,\iota\rangle of SS with welder φ\varphi along (a1,a2)(a_{1},a_{2}) is CC-full and φ\varphi-regular, and its weld graph is I-shaped. On the other hand, inequality (10.1) implies that each aka_{k}^{\circ} contains a point qkq_{k} of ZZ. Take short subarcs ak1a_{k1} and ak2a_{k2} of aka_{k} emanating from qkq_{k} so that Lφ(a11)=Lφ(a12)=Lφ(a21)=Lφ(a22)L_{\varphi}(a_{11})=L_{\varphi}(a_{12})=L_{\varphi}(a_{21})=L_{\varphi}(a_{22}), and let R,ι\langle R^{\prime},\iota^{\prime}\rangle be the welding of SS with welder φ\varphi along (ak1,ak2)(a_{k1},a_{k2}), k=1,2k=1,2. The two points pk:=ι(pk)p^{\prime}_{k}:=\iota^{\prime}(p_{k}), k=1,2k=1,2, divide the component CC^{\prime} of R\partial R^{\prime} containing pkp^{\prime}_{k} into two arcs a1a^{\prime}_{1} and a2a^{\prime}_{2} of the same ψ\psi^{\prime}-length, where ψA+(R)\psi^{\prime}\in A_{+}(R^{\prime}) is the co-welder of φ\varphi. If R′′,ι′′\langle R^{\prime\prime},\iota^{\prime\prime}\rangle denotes the self-welding of RR^{\prime} with welder ψ\psi^{\prime} along (a1,a2)(a^{\prime}_{1},a^{\prime}_{2}), then R′′,ι′′ι\langle R^{\prime\prime},\iota^{\prime\prime}\circ\iota^{\prime}\rangle is a CC-full φ\varphi-regular self-welding of SS. Since its weld graph Gι′′ιG_{\iota^{\prime\prime}\circ\iota^{\prime}} has four end-vertices, the self-welding is not equivalent to R,ι\langle R,\iota\rangle. Thus φ\varphi is exceptional for CC.

Finally, assume that no two points of ZZ divide CC into arcs of the same φ\varphi-length, where cardZ>3\operatorname{card}Z>3. Then we can choose three points pkp_{k}, k=1,2,3k=1,2,3, of ZZ to divide CC into three arcs aka_{k}, k=1,2,3k=1,2,3, such that inequality (10.1) holds for a=aka=a_{k}, k=1,2,3k=1,2,3. We then apply Lemma 5.11 to obtain a CC-full φ\varphi-regular self-welding R,ι\langle R,\iota\rangle of SS whose weld graph GιG_{\iota} is Y-shaped. Let p4p_{4} be a point of ZZ other than pkp_{k}, k=1,2,3k=1,2,3. Take two simple arcs a41a_{41} and a42a_{42} of the same φ\varphi-length on CC emanating from p4p_{4}. We choose them so that neither a41a_{41} nor a42a_{42} contains any of pkp_{k}, k=1,2,3k=1,2,3. Let R,ι\langle R^{\prime},\iota^{\prime}\rangle be the self-welding of SS with welder φ\varphi along (a41,a42)(a_{41},a_{42}), and let ψA+(R)\psi^{\prime}\in A_{+}(R^{\prime}) be the co-welder of φ\varphi. If the φ\varphi-length of a41a_{41} is sufficiently small, then the three points pk:=ι(pk)p^{\prime}_{k}:=\iota^{\prime}(p_{k}), k=1,2,3k=1,2,3, divide the component CC^{\prime} of R\partial R^{\prime} containing these points into three arcs with ψ\psi^{\prime}-length less than Lψ(C)/2L_{\psi^{\prime}}(C^{\prime})/2. Another application of Lemma 5.11 yields a CC^{\prime}-full ψ\psi^{\prime}-regular self-welding R′′,ι′′\langle R^{\prime\prime},\iota^{\prime\prime}\rangle of SS whose weld graph Gι′′G_{\iota^{\prime\prime}} is Y-shaped. Then R′′,ι′′ι\langle R^{\prime\prime},\iota^{\prime\prime}\circ\iota^{\prime}\rangle is a CC-full φ\varphi-regular self-welding of SS. It is not equivalent to R,ι\langle R,\iota\rangle because the weld graph of R′′,ι′′ι\langle R^{\prime\prime},\iota^{\prime\prime}\circ\iota^{\prime}\rangle is not Y-shaped. Therefore, φ\varphi is exceptional for CC.

The last assertion of the theorem is clear form our constructions of inequivalent CC-full φ\varphi-regular self-weldings of SS. The proof is complete. ∎

The following is a corollary to the proof of the above theorem.

Corollary 10.4.

Let SS be a compact bordered Riemann surface, and let CC be a component of S\partial S. Let φAE(S,C)\varphi\in A_{E}(S,C).

  • (i)

    For any closed φ\varphi-regular self-welding R,ι\langle R,\iota\rangle of SS the image ι(C)\iota(C) contains a zero of the co-welder of φ\varphi.

  • (ii)

    For some closed φ\varphi-regular self-welding R,κ\langle R,\kappa\rangle of SS the image κ(C)\kappa(C) is neither I-shaped nor Y-shaped.

Example 10.5.

Let SS be a compact bordered Riemann surface of genus one. Since nonzero holomorphic quadratic differentials on a torus have no zeros, Corollary 10.4 implies that AE(S)=A_{E}(S)=\varnothing. Thus every element φ\varphi of AL(S)A_{L}(S) yields exactly one closed φ\varphi-regular self-welding of SS.

Remark.

Obstruct problems raised by Fehlmann and Gardiner in [15] are closely related to continuations of Riemann surfaces. The uniqueness theorem in [15] is false in general due to the existence of exceptional quadratic differentials as Sasai [49] pointed out. In [50] she gave a sufficient condition for positive quadratic differentials to be exceptional.

Now, let 𝑺=[S,η]{\bm{S}}=[S,\eta] be a marked compact bordered Riemann surface of positive genus. An element 𝝋=[φ,η]{\bm{\varphi}}=[\varphi,\eta] of AL(𝑺)A_{L}({\bm{S}}) is called exceptional if φAE(S)\varphi\in A_{E}(S). Let AE(𝑺)A_{E}({\bm{S}}) stand for the set of exceptional elements of AL(𝑺)A_{L}({\bm{S}}).

Let 𝑹0=[R0,θ0]{\bm{R}}_{0}=[R_{0},\theta_{0}] be a marked nonanalytically finite open Riemann surface of positive genus. Set AE(𝑹0)=𝜾˘0AE(𝑹˘0)A_{E}({\bm{R}}_{0})=\breve{\bm{\iota}}_{0}^{*}A_{E}(\breve{\bm{R}}_{0}), where (𝑹˘0,𝜾˘0)(\breve{\bm{R}}_{0},\breve{\bm{\iota}}_{0}) denotes the natural compact continuation of 𝑹0{\bm{R}}_{0}. Elements of AE(𝑹0)A_{E}({\bm{R}}_{0}) are called exceptional.

Example 10.6.

We consider the case of genus one. We know that AE(𝑹0)A_{E}({\bm{R}}_{0}) is empty (see Example 10.5). Thus we have a well-defined mapping of PAL(𝑹0)PA_{L}({\bm{R}}_{0}) onto 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}). To see that it is injective take 𝝋jAL(𝑹0){\bm{\varphi}}_{j}\in A_{L}({\bm{R}}_{0}), j=1,2j=1,2, and suppose that they induce homotopically consistent conformal embeddings 𝜾j{\bm{\iota}}_{j} of 𝑹0{\bm{R}}_{0} into the same marked torus 𝑹𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}). It then follows from Proposition 9.5 that 𝝋j=𝜾j𝝍j{\bm{\varphi}}_{j}={\bm{\iota}}_{j}^{*}{\bm{\psi}}_{j} for some 𝝍jA(𝑹){\bm{\psi}}_{j}\in A({\bm{R}}). Since A(𝑹)A({\bm{R}}) is one-dimensional, there is a nonzero complex number cc such that 𝝍2=c𝝍1{\bm{\psi}}_{2}=c{\bm{\psi}}_{1} and hence 𝝋2=c𝝋1{\bm{\varphi}}_{2}=c{\bm{\varphi}}_{1}. As 𝝋jA+(𝑹0){\bm{\varphi}}_{j}\in A_{+}({\bm{R}}_{0}), j=1,2j=1,2, we conclude that c>0c>0 so that 𝝋1{\bm{\varphi}}_{1} and 𝝋2{\bm{\varphi}}_{2} represent the same element of PAL(𝑹0)PA_{L}({\bm{R}}_{0}). We have shown that the correspondence between PAL(𝑹0)PA_{L}({\bm{R}}_{0}) onto 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) is bijective. In particular, since AE(𝑹0)=A_{E}({\bm{R}}_{0})=\varnothing, we deduce that for each 𝑹𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}) there exists exactly one homotopically consistent conformal embedding of 𝑹0{\bm{R}}_{0} into 𝑹\bm{R}. We have thus given alternative proofs of [56, Theorems 4 and 5].

If we set

CEmb𝝋(𝑹0)=𝑹𝔐(𝑹0)CEmb𝝋(𝑹0,𝑹),\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0})=\bigcup_{{\bm{R}}\in\mathfrak{M}({\bm{R}}_{0})}\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0},{\bm{R}}),

we see that 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}) is exceptional if and only if cardCEmb𝝋(𝑹0)>1\operatorname{card}\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0})>1. Also, set

CEmbL(𝑹0)=𝝋AL(𝑹0)CEmb𝝋(𝑹0)andCEmbE(𝑹0)=𝝋AE(𝑹0)CEmb𝝋(𝑹0).\operatorname{CEmb}_{L}({\bm{R}}_{0})=\bigcup_{{\bm{\varphi}}\in A_{L}({\bm{R}}_{0})}\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0})\quad\text{and}\quad\operatorname{CEmb}_{E}({\bm{R}}_{0})=\bigcup_{{\bm{\varphi}}\in A_{E}({\bm{R}}_{0})}\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0}).
Theorem 10.7.

Let 𝐑0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface of positive genus. If 𝛗AE(𝐑0){\bm{\varphi}}\in A_{E}({\bm{R}}_{0}), then some element in CEmb𝛗(𝐑0)\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0}) cannot be approximated by any sequence in CEmbL(𝐑0)CEmbE(𝐑0)\operatorname{CEmb}_{L}({\bm{R}}_{0})\setminus\operatorname{CEmb}_{E}({\bm{R}}_{0}).

Proof.

Let 𝑹0=[R0,θ0]{\bm{R}}_{0}=[R_{0},\theta_{0}] and 𝝋=[φ,θ0]{\bm{\varphi}}=[\varphi,\theta_{0}]. Take a natural compact continuation (R˘0,ι˘0)(\breve{R}_{0},\breve{\iota}_{0}) of R0R_{0} together with φ˘AE(R˘0)\breve{\varphi}\in A_{E}(\breve{R}_{0}) for which φ=ι˘0φ˘\varphi=\breve{\iota}_{0}^{*}\breve{\varphi}. By Corollary 10.4 (ii) the weld graph of some closed φ˘\breve{\varphi}-regular self-welding R,ι˘\langle R,\breve{\iota}\rangle of R˘0\breve{R}_{0} has a component that is not I-shaped or Y-shaped. On the other hand, if φ˘AL(R˘0)AE(R˘0)\breve{\varphi}^{\prime}\in A_{L}(\breve{R}_{0})\setminus A_{E}(\breve{R}_{0}), then each component of the weld graph of any closed φ˘\breve{\varphi}^{\prime}-regular self-welding of R˘0\breve{R}_{0} is I-shaped or Y-shaped by Theorem 10.3 and Lemma 5.11. Consequently, no sequences in CEmbL(𝑹0)CEmbE(𝑹0)\operatorname{CEmb}_{L}({\bm{R}}_{0})\setminus\operatorname{CEmb}_{E}({\bm{R}}_{0}) converge to 𝜾:=[ι˘ι˘0,θ0,ι˘ι˘0θ0]CEmb𝝋(𝑹0,𝑹){\bm{\iota}}:=[\breve{\iota}\circ\breve{\iota}_{0},\theta_{0},\breve{\iota}\circ\breve{\iota}_{0}\circ\theta_{0}]\in\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0},{\bm{R}}), where 𝑹=[R,ι˘ι˘0θ0]{\bm{R}}=[R,\breve{\iota}\circ\breve{\iota}_{0}\circ\theta_{0}]. ∎

As an application of the theorem, we prove Theorem 1.4 (iii). To this end we prepare two lemmas. Let RR be a closed Riemann surface of positive genus. For a nontrivial complex vector space VV of holomorphic 11-forms on RR set

ordpV={ordpωωV{0}}\operatorname{ord}_{p}V=\{\operatorname{ord}_{p}\omega\mid\omega\in V\setminus\{0\}\}

for pRp\in R, where ordpω\operatorname{ord}_{p}\omega denotes the order of ω\omega at pp. If ordpV{0,1,,d1}\operatorname{ord}_{p}V\neq\{0,1,\dots,d-1\}, where d=dimVd=\dim V, then pp is called a Weierstrass point for VV. Otherwise, pp is said to be a non-Weierstrass point.

Lemma 10.8.

Let VV be a nontrivial complex vector space of holomorphic 11-forms on a closed Riemann surface RR of genus g>1g>1. Then the set of Weierstrass points for VV is nonempty and finite.

Proof.

Let ωj=ωj(z)dz\omega_{j}=\omega_{j}(z)\,dz, j=1,2,,dj=1,2,\dots,d, be a basis of VV. If W(z)W(z) is the Wronskian of ωj(z)\omega_{j}(z), that is, W(z)=det(ωj(k1)(z))j,k=1,,dW(z)=\det(\omega_{j}^{(k-1)}(z))_{j,k=1,\dots,d}, then W=W(z)dzd(d+1)/2W=W(z)\,dz^{d(d+1)/2} is a holomorphic d(d+1)/2d(d+1)/2-differential on RR. Let pRp\in R and take a local coordinate zz around pp with z(p)=0z(p)=0. If ordpV={ν1,,νd}\operatorname{ord}_{p}V=\{\nu_{1},\dots,\nu_{d}\} and ν=jνjd(d1)/2\nu=\sum_{j}\nu_{j}-d(d-1)/2, then the power series expansion of WW is of the form W(z)=czν+W(z)=cz^{\nu}+\cdots with c0c\neq 0. In particular, WW is nonzero. Moreover, pp is a Weierstrass point for VV if and only if pp is a zero of WW. This proves the lemma. ∎

Lemma 10.9.

Let SS be a compact bordered Riemann surface of genus g3g\geqq 3, and let S^\hat{S} denote its double. Then there is a holomorphic 11-form ω\omega on S^\hat{S} such that

  • (i)

    Imω=0\operatorname{Im}\omega=0 along S\partial S,

  • (ii)

    Cω=0\int_{C}\omega=0 for each component CC of S\partial S, and

  • (iii)

    ω\omega has a zero on SS^{\circ} and four zeros on some component of S\partial S.

Proof.

Let VV be the vector space of holomorphic 11-forms ω\omega on S^\hat{S} satisfying condition (ii). Then dimV=2g\dim V=2g; note that Sω=0\int_{\partial S}\omega=0 for all holomorphic 11-forms ω\omega on S^\hat{S}. Also, let VV_{\mathbb{R}} be the real vector space of ωV\omega\in V possessing property (i). Let JJ be the anti-conformal involution of S^\hat{S} fixing S\partial S pointwise. Then ωJω¯\omega\mapsto\overline{J^{*}\omega} is an \mathbb{R}-linear isomorphism of VV onto itself and σ(ω):=(ω+Jω¯)/2\sigma(\omega):=(\omega+\overline{J^{*}\omega})/2 belongs to VV_{\mathbb{R}} for ωV\omega\in V.

Take a non-Weierstrass point p0Sp_{0}\in S^{\circ} for VV, and let Vp0V_{p_{0}} be the set of 11-forms in VV vanishing at p0p_{0}. Then dimVp0=2g1\dim V_{p_{0}}=2g-1. If V0V_{0} denotes the set of 11-forms in Vp0V_{p_{0}} vanishing at J(p0)J(p_{0}), then d:=dimV02g2d:=\dim V_{0}\geqq 2g-2. Note that for any ωV0\omega\in V_{0} the point p0p_{0} is a zero of σ(ω)\sigma(\omega).

Since g3g\geqq 3, we have d4d\geqq 4. Fix a component C0C_{0} of S\partial S. Let p1C0p_{1}\in C_{0} be a non-Weierstrass point for V0V_{0}, and set V1(k)={ωV0ordp1ωk or ω=0}V_{1}^{(k)}=\{\omega\in V_{0}\mid\operatorname{ord}_{p_{1}}\omega\geqq k\text{ or }\omega=0\} for k=1,2k=1,2. Then dimV1(k)=dk\dim V_{1}^{(k)}=d-k. Take a non-Weierstrass point p2C0{p1}p_{2}\in C_{0}\setminus\{p_{1}\} for both V1(1)V_{1}^{(1)} and V1(2)V_{1}^{(2)}, and set V2(k)={ωV1(1)ordp2ωk or ω=0}V_{2}^{(k)}=\{\omega\in V_{1}^{(1)}\mid\operatorname{ord}_{p_{2}}\omega\geqq k\text{ or }\omega=0\} for k=1,2k=1,2. Note that dimV2(k)=d1k\dim V_{2}^{(k)}=d-1-k and that dim(V1(2)V2(1))=d3\dim(V_{1}^{(2)}\cap V_{2}^{(1)})=d-3. Finally, choose a non-Weierstrass point p3C0{p1,p2}p_{3}\in C_{0}\setminus\{p_{1},p_{2}\} for V2(k)V_{2}^{(k)}, k=1,2k=1,2, and V2(1)V1(2)V_{2}^{(1)}\cap V_{1}^{(2)}, and define V3(k)={ωV2(1)ordp3ωk or ω=0}V_{3}^{(k)}=\{\omega\in V_{2}^{(1)}\mid\operatorname{ord}_{p_{3}}\omega\geqq k\text{ or }\omega=0\} for k=1,2k=1,2. Then dimV3(1)=d3>dimV\dim V_{3}^{(1)}=d-3>\dim V^{\prime} for V=V3(2)V^{\prime}=V_{3}^{(2)}, V2(2)V3(1)V_{2}^{(2)}\cap V_{3}^{(1)} and V2(1)V1(2)V3(1)V_{2}^{(1)}\cap V_{1}^{(2)}\cap V_{3}^{(1)}. Consequently, ordplω1=1\operatorname{ord}_{p_{l}}\omega_{1}=1, l=1,2,3l=1,2,3, for some nonzero ω1V0\omega_{1}\in V_{0}. Set ω=σ(cω1)\omega=\sigma(c\omega_{1}), where c𝔻c\in\partial\mathbb{D}. Then ω\omega satisfies (i) and (ii). We can choose cc so that ω\omega has simple zeros at plp_{l}, l=1,2,3l=1,2,3. Since Imω=0\operatorname{Im}\omega=0 along C0C_{0} and C0ω=0\int_{C_{0}}\omega=0, the 11-form ω\omega has one more zero of odd order on C0C_{0}. Hence ω\omega satisfies (i), (ii) and (iii). ∎

We are now ready to prove Theorem 1.4 (iii).

Proof of Theorem 1.4 (iii).

Let (𝑹˘0,𝜾˘0)(\breve{\bm{R}}_{0},\breve{\bm{\iota}}_{0}), where 𝑹˘0=[R˘0,θ˘0]\breve{\bm{R}}_{0}=[\breve{R}_{0},\breve{\theta}_{0}] and 𝜾˘0=[ι˘0,θ0,θ˘0]\breve{\bm{\iota}}_{0}=[\breve{\iota}_{0},\theta_{0},\breve{\theta}_{0}], be the natural compact continuation of 𝑹0=[R0,θ0]{\bm{R}}_{0}=[R_{0},\theta_{0}]. Let R^0\hat{R}_{0} be the double of R˘0\breve{R}_{0}. We apply Lemma 10.9 to obtain a holomorphic 11-form ω\omega on R^0\hat{R}_{0} that has a zero on (R˘0)(\breve{R}_{0})^{\circ} and four zeros on a component of R˘0\partial\breve{R}_{0}. Set 𝝋=[ι˘0(ω2),θ0]{\bm{\varphi}}=[\breve{\iota}_{0}^{*}(\omega^{2}),\theta_{0}]. Then, since 𝝋AE(𝑹0){\bm{\varphi}}\in A_{E}({\bm{R}}_{0}) by Theorem 10.3, Theorem 10.7 implies that there is a closed 𝝋\bm{\varphi}-regular self-welding continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) of 𝑹0{\bm{R}}_{0} such that there are no sequences {(𝑹n,𝜾n)}\{({\bm{R}}_{n},{\bm{\iota}}_{n})\} of closed regular self-welding continuations of 𝑹0{\bm{R}}_{0} for which {𝜾n}\{{\bm{\iota}}_{n}\} is a sequence in CEmbL(𝑹0)CEmbE(𝑹0)\operatorname{CEmb}_{L}({\bm{R}}_{0})\setminus\operatorname{CEmb}_{E}({\bm{R}}_{0}) converging to 𝜾\bm{\iota}.

We claim that 𝑹\bm{R} is an interior point 𝔐E(𝑹0)\mathfrak{M}_{E}({\bm{R}}_{0}) in 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}). If not, then some sequence {𝑹n}\{{\bm{R}}^{\prime}_{n}\} in 𝔐(𝑹0)𝔐E(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0})\setminus\mathfrak{M}_{E}({\bm{R}}_{0}) would converge to 𝑹\bm{R}. Take closed regular self-welding continuations (𝑹n,𝜾n)({\bm{R}}^{\prime}_{n},{\bm{\iota}}^{\prime}_{n}) of 𝑹0{\bm{R}}_{0}. By Proposition 7.7 there is a closed regular self-welding continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}^{\prime}) of 𝑹0{\bm{R}}_{0} such that a subsequence of {𝜾n}\{{\bm{\iota}}^{\prime}_{n}\} converges to 𝜾{\bm{\iota}}^{\prime}. Since ι˘0(ω2)\breve{\iota}_{0}^{*}(\omega^{2}) has a zero on R0R_{0}, it follows from Bourque [12, Remark 4.3] that 𝜾=𝜾{\bm{\iota}}^{\prime}={\bm{\iota}}, which is impossible by the choice of (𝑹,𝜾)({\bm{R}},{\bm{\iota}}). This completes the proof. ∎

Remark.

Let 𝝋\bm{\varphi} be as in the above proof. Then CEmbhc(𝑹0,𝑹)\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) is a singleton for any 𝑹𝔐𝝋(𝑹0){\bm{R}}\in\mathfrak{M}_{\bm{\varphi}}({\bm{R}}_{0}). Since cardCEmb𝝋(𝑹0)=20\operatorname{card}\operatorname{CEmb}_{\bm{\varphi}}({\bm{R}}_{0})=2^{\aleph_{0}}, we know that card𝔐𝝋(𝑹0)=20\operatorname{card}\mathfrak{M}_{\bm{\varphi}}({\bm{R}}_{0})=2^{\aleph_{0}}.

11 Fillings for marked open Riemann surfaces

The aim of the present section is to introduce some key tools used in the following sections. Let {𝑺t}t[0,1]\{{\bm{S}}_{t}\}_{t\in[0,1]} be a one-parameter family of marked open Riemann surfaces of genus gg. We say that 𝔐(𝑺t)\mathfrak{M}({\bm{S}}_{t}) shrinks continuously if

  • (i)

    𝔐(𝑺t1)𝔐(𝑺t2)\mathfrak{M}({\bm{S}}_{t_{1}})\supset\mathfrak{M}({\bm{S}}_{t_{2}}) for t1<t2t_{1}<t_{2}, and

  • (ii)

    for each ε>0\varepsilon>0 and t[0,1]t\in[0,1] there is δ>0\delta>0 such that 𝔐(𝑺t1)\mathfrak{M}({\bm{S}}_{t_{1}}) is included in the ε\varepsilon-neighborhood of 𝔐(𝑺t2)\mathfrak{M}({\bm{S}}_{t_{2}}), where t1=max{tδ,0}t_{1}=\max\{t-\delta,0\} and t2=min{t+δ,1}t_{2}=\min\{t+\delta,1\}.

If, in addition, 𝔐(𝑺1)\mathfrak{M}({\bm{S}}_{1}) is a singleton, then 𝔐(𝑺t)\mathfrak{M}({\bm{S}}_{t}) is said to shrink continuously to a point.

Lemma 11.1.

Let 𝐒j{\bm{S}}_{j}, j=1,2j=1,2, be marked open Riemann surfaces of genus gg. If there is a homotopically consistent KK-quasiconformal embedding of 𝐒1{\bm{S}}_{1} into 𝐒2{\bm{S}}_{2}, then

𝔐(𝑺2)𝔐K(𝑺1).\mathfrak{M}({\bm{S}}_{2})\subset\mathfrak{M}_{K}({\bm{S}}_{1}).
Proof.

Let 𝒇QCEmbhc(𝑺1,𝑺2){\bm{f}}\in\operatorname{QCEmb}_{\mathrm{hc}}({\bm{S}}_{1},{\bm{S}}_{2}) with K(𝒇)KK({\bm{f}})\leqq K. If (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) is a closed continuation of 𝑺2{\bm{S}}_{2}, then 𝜾𝒇QCEmbhc(𝑺1,𝑹){\bm{\iota}}\circ{\bm{f}}\in\operatorname{QCEmb}_{\mathrm{hc}}({\bm{S}}_{1},{\bm{R}}) with K(𝜾𝒇)KK({\bm{\iota}}\circ{\bm{f}})\leqq K. Hence 𝑹\bm{R} belongs to 𝔐K(𝑺1)\mathfrak{M}_{K}({\bm{S}}_{1}). ∎

Proposition 11.2.

Let {𝐒t}t[0,1]\{{\bm{S}}_{t}\}_{t\in[0,1]} be a one-parameter family of marked finite open Riemann surfaces of genus gg. If 𝔐(𝐒t)\mathfrak{M}({\bm{S}}_{t}) shrinks continuously, then

Int𝔐(𝑺t)=u(t,1]Int𝔐(𝑺u)\operatorname{Int}\mathfrak{M}({\bm{S}}_{t})=\bigcup_{u\in(t,1]}\operatorname{Int}\mathfrak{M}({\bm{S}}_{u})

for t[0,1)t\in[0,1).

Proof.

Fix t[0,1)t\in[0,1). Since Int𝔐(𝑺u)Int𝔐(𝑺t)\operatorname{Int}\mathfrak{M}({\bm{S}}_{u})\subset\operatorname{Int}\mathfrak{M}({\bm{S}}_{t}) for u(t,1]u\in(t,1], we know that

u(t,1]Int𝔐(𝑺u)Int𝔐(𝑺t).\bigcup_{u\in(t,1]}\operatorname{Int}\mathfrak{M}({\bm{S}}_{u})\subset\operatorname{Int}\mathfrak{M}({\bm{S}}_{t}).

To show the converse inclusion relation take an arbitrary interior point 𝑹\bm{R} of 𝔐(𝑺t)\mathfrak{M}({\bm{S}}_{t}). Then for some ε>0\varepsilon>0 the neighborhood 𝔅4ε(𝑹)\mathfrak{B}_{4\varepsilon}({\bm{R}}) of 𝑹\bm{R} is included in 𝔐(𝑺t)\mathfrak{M}({\bm{S}}_{t}). We can choose u(t,1]u\in(t,1] sufficiently near to tt so that 𝔐(𝑺t)𝔐e4ε(𝑺u)\mathfrak{M}({\bm{S}}_{t})\subset\mathfrak{M}_{e^{4\varepsilon}}({\bm{S}}_{u}). We claim that 𝑹\bm{R} is an interior point of 𝔐(𝑺u)\mathfrak{M}({\bm{S}}_{u}). If not, then 𝑹\bm{R} would lie on an Ioffe ray 𝒓\bm{r} of 𝑺u{\bm{S}}_{u} with 𝒓(τ)=𝑹{\bm{r}}(\tau)={\bm{R}}, where τ=dT(𝑹,𝔐(𝑺u))2ε\tau=d_{T}({\bm{R}},\mathfrak{M}({\bm{S}}_{u}))\leqq 2\varepsilon by Proposition 9.1. Since

dT(𝒓(3ε),𝑹)=dT(𝒓(3ε),𝒓(τ))=3ετ<4ε,d_{T}({\bm{r}}(3\varepsilon),{\bm{R}})=d_{T}({\bm{r}}(3\varepsilon),{\bm{r}}(\tau))=3\varepsilon-\tau<4\varepsilon,

we have 𝒓(3ε)𝔅4ε(𝑹)𝔐(𝑺t)𝔐e4ε(𝑺u){\bm{r}}(3\varepsilon)\in\mathfrak{B}_{4\varepsilon}({\bm{R}})\subset\mathfrak{M}({\bm{S}}_{t})\subset\mathfrak{M}_{e^{4\varepsilon}}({\bm{S}}_{u}) and hence dT(𝒓(3ε),𝔐(𝑺u))2εd_{T}({\bm{r}}(3\varepsilon),\mathfrak{M}({\bm{S}}_{u}))\leqq 2\varepsilon. This contradicts the identity dT(𝒓(3ε),𝔐(𝑺u))=3εd_{T}({\bm{r}}(3\varepsilon),\mathfrak{M}({\bm{S}}_{u}))=3\varepsilon, and the proof is complete. ∎

Let 𝑹0=[R0,θ0]𝔉g{\bm{R}}_{0}=[R_{0},\theta_{0}]\in\mathfrak{F}_{g} be open and nonanalytically finite. We construct two one-parameter families of continuations of 𝑹0{\bm{R}}_{0}. In the following examples (𝑹˘0,𝜾˘0)(\breve{\bm{R}}_{0},\breve{\bm{\iota}}_{0}) denotes the natural compact continuation of 𝑹0{\bm{R}}_{0}, where 𝑹˘0=[R˘0,θ˘0]\breve{\bm{R}}_{0}=[\breve{R}_{0},\breve{\theta}_{0}] and 𝜾˘0=[ι˘0,θ0,θ˘0]\breve{\bm{\iota}}_{0}=[\breve{\iota}_{0},\theta_{0},\breve{\theta}_{0}].

In general, for r>0r>0 let 𝔻r\mathbb{D}_{r} denote the open disk in \mathbb{C} of radius rr centered at 0. Its closure is denoted by 𝔻¯r\bar{\mathbb{D}}_{r}. Set 𝔻¯0={0}\bar{\mathbb{D}}_{0}=\{0\} for convenience. We abbreviate 𝔻1\mathbb{D}_{1} and 𝔻¯1\bar{\mathbb{D}}_{1} to 𝔻\mathbb{D} and 𝔻¯\bar{\mathbb{D}}, respectively.

Example 11.3 (circular filling).

Let C1,,Cn0C_{1},\ldots,C_{n_{0}} be the connected components of the border R˘0\partial\breve{R}_{0}. Take doubly connected domains UjU_{j}, j=1,,n0j=1,\ldots,n_{0}, on (R˘0)(\breve{R}_{0})^{\circ} such that CjC_{j} is a component of Uj\partial U_{j} and that UjU_{j} is mapped onto a fixed annulus 𝔻r0𝔻¯\mathbb{D}_{r_{0}}\setminus\bar{\mathbb{D}} by a conformal homeomorphism zjz_{j} with |zj|=1|z_{j}|=1 on CjC_{j}, where r0>1r_{0}>1. We attach 𝔻¯\bar{\mathbb{D}} to each CjC_{j} by identifying pCjp\in C_{j} with zj(p)𝔻z_{j}(p)\in\partial\mathbb{D} to obtain a closed Riemann surface WW of genus gg. We consider R˘0\breve{R}_{0} as a closed subdomain of WW, and denote by DjD_{j} the component of WR˘0W\setminus\breve{R}_{0} with Cj=DjC_{j}=\partial D_{j}. Then zjz_{j} is extended to a conformal mapping of UjD¯jU_{j}\cup\bar{D}_{j} onto 𝔻r0\mathbb{D}_{r_{0}}. Considering θ˘0\breve{\theta}_{0} as a gg-handle mark of WW as well, set 𝑾=[W,θ˘0]{\bm{W}}=[W,\breve{\theta}_{0}]. Then 𝑾𝔐(𝑹0){\bm{W}}\in\mathfrak{M}({\bm{R}}_{0}).

For each t[0,1]t\in[0,1] we construct a subsurface WtW_{t} of WW homeomorphic to (R˘0)(\breve{R}_{0})^{\circ} as follows:

Wt=Wj=1n0zj1(𝔻¯1t)W_{t}=W\setminus\bigcup_{j=1}^{n_{0}}z_{j}^{-1}(\bar{\mathbb{D}}_{1-t})

Define 𝑾t=[Wt,θ˘0]{\bm{W}}_{t}=[W_{t},\breve{\theta}_{0}] for t[0,1]t\in[0,1]. Regarding 𝜾˘0\breve{\bm{\iota}}_{0} as a homotopically consistent conformal embedding ϵt{\bm{\epsilon}}_{t} of 𝑹0{\bm{R}}_{0} into 𝑾t{\bm{W}}_{t}, we call {(𝑾t,ϵt)}t[0,1]\{({\bm{W}}_{t},{\bm{\epsilon}}_{t})\}_{t\in[0,1]} a circular filling for 𝑹0{\bm{R}}_{0}. Note that W0ι˘0(R0)W_{0}\setminus\breve{\iota}_{0}(R_{0}) is a finite set.

Example 11.4 (linear filling).

As in Example 5.2, take φ˘M+(R˘0)\breve{\varphi}\in M_{+}(\breve{R}_{0}), and divide each border component CjC_{j} into two subarcs a2j1a_{2j-1} and a2ja_{2j} of the same φ˘\breve{\varphi}-length Lj:=Lφ˘(Cj)/2L_{j}:=L_{\breve{\varphi}}(C_{j})/2. Let W,ϵ\langle W,\epsilon\rangle be the closed self-welding of R˘0\breve{R}_{0} with welder φ˘\breve{\varphi} along (a2j1,a2j)(a_{2j-1},a_{2j}), j=1,,n0j=1,\ldots,n_{0}. and let ψM(W)\psi\in M(W) be the co-welder of φ˘\breve{\varphi}. The arcs a2j1a_{2j-1} and a2ja_{2j} are projected to a simple arc aja^{\prime}_{j} on WW, which is parametrized with respect to ψ\psi-length. Set

Wt=Wj=1n0aj([0,(1t)Lj])W_{t}=W\setminus\bigcup_{j=1}^{n_{0}}a^{\prime}_{j}([0,(1-t)L_{j}])

for t[0,1]t\in[0,1]. Let ϵt\epsilon_{t} denote the conformal embedding of R0R_{0} into WtW_{t} induced by ϵι˘0\epsilon\circ\breve{\iota}_{0}. The continuation (Wt,ϵt)(W_{t},\epsilon_{t}) of R0R_{0} yields a continuation (𝑾t,ϵt)({\bm{W}}_{t},{\bm{\epsilon}}_{t}) of 𝑹0{\bm{R}}_{0}. We call {(𝑾t,ϵt)}t[0,1]\{({\bm{W}}_{t},{\bm{\epsilon}}_{t})\}_{t\in[0,1]} a linear filling for 𝑹0{\bm{R}}_{0}. Again, W0ι˘0(R0)W_{0}\setminus\breve{\iota}_{0}(R_{0}) is a finite set.

Fillings introduced in the above examples provide us with continuity methods:

Proposition 11.5.

Let 𝐑0{\bm{R}}_{0} be a marked nonanalytically finite open Riemann surface of genus gg. If {(𝐖t,ϵt)}t[0,1]\{({\bm{W}}_{t},{\bm{\epsilon}}_{t})\}_{t\in[0,1]} is a circular or linear filling for 𝐑0{\bm{R}}_{0}, then 𝔐(𝐖t)\mathfrak{M}({\bm{W}}_{t}) shrinks continuously to a point.

Proof.

It is clear that 𝑾1{\bm{W}}_{1} is analytically finite and hence 𝔐(𝑾1)\mathfrak{M}({\bm{W}}_{1}) is a singleton, say, {𝑾}\{{\bm{W}}\}. If t1<t2t_{1}<t_{2}, then CEmbhc(𝑾t1,𝑾t2)\operatorname{CEmb}_{\mathrm{hc}}({\bm{W}}_{t_{1}},{\bm{W}}_{t_{2}})\neq\varnothing, and hence 𝔐(𝑾t2)𝔐1(𝑾t1)=𝔐(𝑾t1)\mathfrak{M}({\bm{W}}_{t_{2}})\subset\mathfrak{M}_{1}({\bm{W}}_{t_{1}})=\mathfrak{M}({\bm{W}}_{t_{1}}) by Lemma 11.1.

To show that 𝔐(𝑾t)\mathfrak{M}({\bm{W}}_{t}) shrinks continuously take ε>0\varepsilon>0 and t[0,1]t\in[0,1]. By Proposition 9.1 we are required to find δ>0\delta>0 such that 𝔐(𝑾t1)𝔐e2ε(𝑾t2)\mathfrak{M}({\bm{W}}_{t_{1}})\subset\mathfrak{M}_{e^{2\varepsilon}}({\bm{W}}_{t_{2}}), where t1=max{tδ,0}t_{1}=\max\{t-\delta,0\} and t2=min{t+δ,1}t_{2}=\min\{t+\delta,1\}. For t[0,1)t\in[0,1) we can choose a desired δ\delta with the aid of Lemma 11.1. To verify the existence of δ\delta for t=1t=1 we have only to show that 𝔐(𝑾u)𝔅ε(𝑾)\mathfrak{M}({\bm{W}}_{u})\subset\mathfrak{B}_{\varepsilon}({\bm{W}}) for all uu sufficiently near 11. If there were no such δ\delta, then we could find a point 𝑹\bm{R} in u<1𝔐(𝑾u)𝔅ε(𝑾)\bigcap_{u<1}\mathfrak{M}({\bm{W}}_{u})\setminus\mathfrak{B}_{\varepsilon}({\bm{W}}) as {𝔐(𝑾u)𝔅ε(𝑾)}u<1\{\mathfrak{M}({\bm{W}}_{u})\setminus\mathfrak{B}_{\varepsilon}({\bm{W}})\}_{u<1} would be a family of compact sets with finite intersection property. Choose 𝜾uCEmbhc(𝑾u,𝑹){\bm{\iota}}_{u}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{W}}_{u},{\bm{R}}) and 𝒉uHomeohc+(𝑾1,𝑾u){\bm{h}}_{u}\in\operatorname{Homeo}^{+}_{\mathrm{hc}}({\bm{W}}_{1},{\bm{W}}_{u}) for each u(0,1)u\in(0,1) so that {𝒉u}u\{{\bm{h}}_{u}\}_{u} converges to 𝟏𝑾1{\bm{1}}_{{\bm{W}}_{1}} as u1u\to 1. We apply Lemma 7.4 to obtain a sequence {un}\{u_{n}\} converging to 11 for which the sequence {𝜾un}\{{\bm{\iota}}_{u_{n}}\} converges to a homotopically consistent conformal embedding of 𝑾1{\bm{W}}_{1} into 𝑹\bm{R}. Then 𝑹𝔐(𝑾1){\bm{R}}\in\mathfrak{M}({\bm{W}}_{1}), or 𝑹=𝑾𝔅ε(𝑾){\bm{R}}={\bm{W}}\in\mathfrak{B}_{\varepsilon}({\bm{W}}), which is absurd. This finishes the proof of the proposition. ∎

We give an application of our methods. Combinations of circular and linear fillings give varieties of conformal embeddings.

Proposition 11.6.

If 𝐑Int𝔐(𝐑0){\bm{R}}\in\operatorname{Int}\mathfrak{M}({\bm{R}}_{0}), then cardCEmbhc(𝐑0,𝐑)=20\operatorname{card}\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}})=2^{\aleph_{0}}.

Proof.

We use the notations in Example 11.4. Thus {(𝑾t,ϵt)}t[0,1]\{({\bm{W}}_{t},{\bm{\epsilon}}_{t})\}_{t\in[0,1]} is a linear filling for 𝑹0=[R0,θ0]{\bm{R}}_{0}=[R_{0},\theta_{0}]. For each t[0,1)t\in[0,1) take a circular filling {(𝑾t(u),ϵt(u))}u[0,1]\{({\bm{W}}_{t}^{(u)},{\bm{\epsilon}}_{t}^{(u)})\}_{u\in[0,1]} for 𝑾t{\bm{W}}_{t}. Let 𝑹=[R,θ]Int𝔐(𝑹0)=Int𝔐(𝑾0){\bm{R}}=[R,\theta]\in\operatorname{Int}\mathfrak{M}({\bm{R}}_{0})=\operatorname{Int}\mathfrak{M}({\bm{W}}_{0}). Since 𝔐(𝑾t)\mathfrak{M}({\bm{W}}_{t}) shrinks continuously by Proposition 11.5, it follows from Proposition 11.2 that 𝑹Int𝔐(𝑾δ){\bm{R}}\in\operatorname{Int}\mathfrak{M}({\bm{W}}_{\delta}) for some δ(0,1)\delta\in(0,1). Another application of Propositions 11.2 and 11.5 assures us that for each t(0,δ)t\in(0,\delta) there is ut(0,1)u_{t}\in(0,1) such that 𝑹Int𝔐(𝑾t(ut)){\bm{R}}\in\operatorname{Int}\mathfrak{M}({\bm{W}}_{t}^{(u_{t})}). Choose 𝜿tCEmbhc(𝑾t(ut),𝑹){\bm{\kappa}}_{t}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{W}}_{t}^{(u_{t})},{\bm{R}}) and set 𝜾t=𝜿tϵt(ut)ϵt{\bm{\iota}}_{t}={\bm{\kappa}}_{t}\circ{\bm{\epsilon}}_{t}^{(u_{t})}\circ{\bm{\epsilon}}_{t} to obtain continuations (𝑹,𝜾t)({\bm{R}},{\bm{\iota}}_{t}), t(0,δ)t\in(0,\delta), of 𝑹0{\bm{R}}_{0}. Observe that through ιt\iota_{t}, where (ιt,θ0,θ)𝜾t(\iota_{t},\theta_{0},\theta)\in{\bm{\iota}}_{t}, each pair of arcs a2j1(t):=a2j1([0,(1t)Lj])a_{2j-1}^{(t)}:=a_{2j-1}([0,(1-t)L_{j}]) and a2j(t):=a2j([0,(1t)Lj])a_{2j}^{(t)}:=a_{2j}([0,(1-t)L_{j}]) on CjC_{j} yields a simple arc on RR while Cj(a2j1(t)a2j(t))C_{j}\setminus(a_{2j-1}^{(t)}\cup a_{2j}^{(t)}) gives rise to a simple loop on RR. Consequently, the continuations (𝑹,𝜾t)({\bm{R}},{\bm{\iota}}_{t}), t(0,δ)t\in(0,\delta), of 𝑹0{\bm{R}}_{0} are distinct from one another. This completes the proof. ∎

Remark.

In the case of genus one alternative proofs of the proposition are found in [35, Corollary 2] and [52] based on area theorems in [59].

We conclude this section by giving one more application of circular fillings. We apply it in §14.

Proposition 11.7.

If {(𝐖t,ϵt)}t[0,1]\{({\bm{W}}_{t},{\bm{\epsilon}}_{t})\}_{t\in[0,1]} is a circular filling for 𝐑0{\bm{R}}_{0}, then

𝔐(𝑾u)Int𝔐(𝑾t)\mathfrak{M}({\bm{W}}_{u})\subset\operatorname{Int}\mathfrak{M}({\bm{W}}_{t})

for 0t<u10\leqq t<u\leqq 1.

Proof.

We use the notations in Example 11.3. Let (𝑹,𝜾)=([R,θ],[ι,θ0˘,θ])({\bm{R}},{\bm{\iota}})=([R,\theta],[\iota,\breve{\theta_{0}},\theta]) be a closed continuation of 𝑾u{\bm{W}}_{u}, where u(0,1]u\in(0,1]. If 0t<u0\leqq t<u, then WtW_{t} is a subsurface of WuW_{u}. Since WuWtW_{u}\setminus W_{t} is of positive area, so is Rι(Wt)R\setminus\iota(W_{t}). Proposition 9.5 thus implies that 𝑹\bm{R} cannot lie on the boundary of 𝔐(𝑾t)\mathfrak{M}({\bm{W}}_{t}). Hence 𝑹Int𝔐(𝑾t){\bm{R}}\in\operatorname{Int}\mathfrak{M}({\bm{W}}_{t}), as desired. ∎

12 Maximal sets for measured foliations

Let 𝑹0{\bm{R}}_{0} be a marked finite open Riemann surface of genus gg. In this section we give a homeomorphism of AL(𝑹0)A_{L}({\bm{R}}_{0}) onto (Σg){0}\mathscr{MF}(\Sigma_{g})\setminus\{0\} explicitly to establish Theorem 1.5. Then we prove Theorem 1.6.

Definition 12.1 (maximal set).

Let (Σg)\mathcal{F}\in\mathscr{MF}(\Sigma_{g}), and let 𝔎\mathfrak{K} be a compact subset of 𝔗g\mathfrak{T}_{g}. A point of 𝔎\mathfrak{K} where the restriction of Ext\operatorname{Ext}_{\mathcal{F}} to 𝔎\mathfrak{K} attains its maximum is referred to as a maximal point for \mathcal{F} on 𝔎\mathfrak{K}. We call the set of maximal points for \mathcal{F} on 𝔎\mathfrak{K} the maximal set for \mathcal{F} on 𝔎\mathfrak{K}.

Since 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is compact by Corollary 7.6, we can speak of maximal points and sets for \mathcal{F} on 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). Denote by 𝔐(𝑹0)\mathfrak{M}_{\mathcal{F}}({\bm{R}}_{0}) the maximal set for \mathcal{F} on 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). Then we have the following proposition.

Proposition 12.2.

Let 𝐑0𝔉g{\bm{R}}_{0}\in\mathfrak{F}_{g} be finite and open, and let (Σg){0}\mathcal{F}\in\mathscr{MF}(\Sigma_{g})\setminus\{0\}. Then any point 𝐑𝔐(𝐑0)\bm{R}\in\mathfrak{M}_{\mathcal{F}}({\bm{R}}_{0}) lies on 𝔐(𝐑0)\partial\mathfrak{M}({\bm{R}}_{0}), and 𝐫𝐑[𝛙]{\bm{r}}_{\bm{R}}[{\bm{\psi}}] is an Ioffe ray of 𝐑0{\bm{R}}_{0}, where 𝛙=𝐐𝐑(){\bm{\psi}}={\bm{Q}}_{{\bm{R}}}(\mathcal{F}). Moreover, the pull-back 𝛊𝛙{\bm{\iota}}^{*}{\bm{\psi}} of 𝛙\bm{\psi} by 𝛊CEmbhc(𝐑0,𝐑){\bm{\iota}}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) is determined solely by \mathcal{F} and does not depend on 𝐑\bm{R} or 𝛊\bm{\iota}.

Proof.

Fix a maximal point 𝑹\bm{R} for \mathcal{F} on 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}), and set 𝒓=𝒓𝑹[𝝍]{\bm{r}}={\bm{r}}_{\bm{R}}[{\bm{\psi}}], where 𝝍=𝑸𝑹(){\bm{\psi}}={\bm{Q}}_{\bm{R}}(\mathcal{F}). If 𝑺𝔅ρ(𝒓(ρ)){\bm{S}}\in\mathfrak{B}_{\rho}({\bm{r}}(\rho)) with ρ>0\rho>0, then Lemma 9.9 together with (6.4) shows that

logExt(𝑺)\displaystyle\log\operatorname{Ext}_{\mathcal{F}}({\bm{S}}) logExt(𝒓(ρ))2dT(𝑺,𝒓(ρ))\displaystyle\geqq\log\operatorname{Ext}_{\mathcal{F}}({\bm{r}}(\rho))-2d_{T}({\bm{S}},{\bm{r}}(\rho))
=logExt(𝒓(0))+2ρ2dT(𝑺,𝒓(ρ))>logExt(𝒓(0)),\displaystyle=\log\operatorname{Ext}_{\mathcal{F}}({\bm{r}}(0))+2\rho-2d_{T}({\bm{S}},{\bm{r}}(\rho))>\log\operatorname{Ext}_{\mathcal{F}}({\bm{r}}(0)),

which implies that 𝑺𝔐(𝑹0){\bm{S}}\not\in\mathfrak{M}({\bm{R}}_{0}), for, 𝑹=𝒓(0){\bm{R}}={\bm{r}}(0) is a maximal point for \mathcal{F} on 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). We have proved that 𝔅ρ(𝒓(ρ))𝔐(𝑹0)=\mathfrak{B}_{\rho}({\bm{r}}(\rho))\cap\mathfrak{M}({\bm{R}}_{0})=\varnothing for all ρ>0\rho>0. We then apply Lemma 9.7 to obtain 𝔅¯ρ(𝒓(ρ))𝔐(𝑹0)={𝑹}\bar{\mathfrak{B}}_{\rho}({\bm{r}}(\rho))\cap\mathfrak{M}({\bm{R}}_{0})=\{{\bm{R}}\}. Hence 𝑹\bm{R} is a boundary point of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). Furthermore, 𝑹\bm{R} is the nearest point of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) to 𝒓(ρ){\bm{r}}(\rho). Since 𝒓=𝒓𝑹[𝒓(ρ)]{\bm{r}}={\bm{r}}_{\bm{R}}[{\bm{r}}(\rho)], Proposition 9.6 shows that 𝒓\bm{r} is an Ioffe ray emanating from 𝑹\bm{R}.

To prove the last assertion of the theorem we may assume that 𝑹0{\bm{R}}_{0} is nonanalytically finite. Since 𝒓𝑹[𝝍]{\bm{r}}_{\bm{R}}[{\bm{\psi}}] is an Ioffe ray of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}), for some 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}) there is a 𝝋{\bm{\varphi}}-regular self-welding continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}) of 𝑹0{\bm{R}}_{0} such that 𝝍\bm{\psi} is the co-welder of 𝝋\bm{\varphi}. Let 𝑹{\bm{R}}^{\prime} be an arbitrary maximal point for \mathcal{F} on 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}), and choose 𝜾CEmbhc(𝑹0,𝑹){\bm{\iota}}^{\prime}\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}^{\prime}) arbitrarily. Denote by 𝝋{\bm{\varphi}}^{\prime} the (𝑹,𝜾)({\bm{R}}^{\prime},{\bm{\iota}}^{\prime})-zero-extension of 𝝋\bm{\varphi}. By Lemma 6.5 and Proposition 6.4 we obtain 𝝍𝑹𝝋𝑹\|{\bm{\psi}}^{\prime}\|_{{\bm{R}}^{\prime}}\leqq\|{\bm{\varphi}}^{\prime}\|_{{\bm{R}}^{\prime}}, where 𝝍=𝑸𝑹(){\bm{\psi}}^{\prime}={\bm{Q}}_{{\bm{R}}^{\prime}}(\mathcal{F}). Actually, the sign of equality occurs because

𝝋𝑹=𝝋𝑹0=𝝍𝑹=Ext(𝑹)=Ext(𝑹)=𝝍𝑹.\|{\bm{\varphi}}^{\prime}\|_{{\bm{R}}^{\prime}}=\|{\bm{\varphi}}\|_{{\bm{R}}_{0}}=\|{\bm{\psi}}\|_{\bm{R}}=\operatorname{Ext}_{\mathcal{F}}({\bm{R}})=\operatorname{Ext}_{\mathcal{F}}({\bm{R}}^{\prime})=\|{\bm{\psi}}^{\prime}\|_{{\bm{R}}^{\prime}}.

Another application of Proposition 6.4 gives us 𝝋=𝝍{\bm{\varphi}}^{\prime}={\bm{\psi}}^{\prime} almost everywhere on 𝑹{\bm{R}}^{\prime}, which implies that (𝜾)𝝍=𝝋({\bm{\iota}}^{\prime})^{*}{\bm{\psi}}^{\prime}={\bm{\varphi}}. This completes the proof. ∎

Let 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}). If 𝑹0=[R0,θ0]{\bm{R}}_{0}=[R_{0},\theta_{0}] is nonanalytically finite, then 𝝋\bm{\varphi} is a welder of some regular self-welding continuation (𝑹,𝜾)=([R,θ],[ι,θ0,θ])({\bm{R}},{\bm{\iota}})=([R,\theta],[\iota,\theta_{0},\theta]) of 𝑹0{\bm{R}}_{0} by Theorem 1.3. Let 𝝍=[ψ,θ]A(𝑹){\bm{\psi}}=[\psi,\theta]\in A({\bm{R}}) be the co-welder of 𝝋{\bm{\varphi}}, and set

𝑹0(𝝋)=𝑹(𝝍).\mathcal{H}_{{\bm{R}}_{0}}({\bm{\varphi}})=\mathcal{H}_{\bm{R}}({\bm{\psi}}).

The measured foliation 𝑹0(𝝋)\mathcal{H}_{{\bm{R}}_{0}}({\bm{\varphi}}) on Σg\Sigma_{g} does not depend on a particular choice of the self-welding (𝑹,𝜾)({\bm{R}},{\bm{\iota}}), for, Rι(R0)R\setminus\iota(R_{0}) consists of finitely many horizontal arcs of ψ\psi together with finitely many points and contributes nothing for evaluating the ψ\psi-heights Hψ(c)H_{\psi}(c) of curves cc on RR. We have thus obtained a mapping 𝑹0\mathcal{H}_{{\bm{R}}_{0}} of AL(𝑹0)A_{L}({\bm{R}}_{0}) into (Σg){0}\mathscr{MF}(\Sigma_{g})\setminus\{0\}, which is bijective by Proposition 12.2. We denote its inverse 𝑹01\mathcal{H}_{{\bm{R}}_{0}}^{-1} by 𝑸𝑹0{\bm{Q}}_{{\bm{R}}_{0}}.

If 𝑹0{\bm{R}}_{0} is analytically finite, then it has a unique closed continuation (𝑹,𝜾)({\bm{R}},{\bm{\iota}}), and 𝜾\bm{\iota} induces a bijection of AL(𝑹0)A_{L}({\bm{R}}_{0}) onto A(𝑹){𝟎}A({\bm{R}})\setminus\{{\bm{0}}\}. We then define 𝑹0=𝑹𝜾\mathcal{H}_{{\bm{R}}_{0}}=\mathcal{H}_{\bm{R}}\circ{\bm{\iota}}_{*}, which is a homeomorphism of AL(𝑹0)A_{L}({\bm{R}}_{0}) onto (Σg){0}\mathscr{MF}(\Sigma_{g})\setminus\{0\} by Proposition 6.2. Again, we set 𝑸𝑹0=𝑹01{\bm{Q}}_{{\bm{R}}_{0}}=\mathcal{H}_{{\bm{R}}_{0}}^{-1}.

Proof of Theorem 1.5.

We are required to show that 𝑹0\mathcal{H}_{{\bm{R}}_{0}} and 𝑸𝑹0{\bm{Q}}_{{\bm{R}}_{0}} are continuous. To prove that 𝑹0\mathcal{H}_{{\bm{R}}_{0}} is continuous let R^0\hat{R}_{0} be the double of R˘0\breve{R}_{0}, where (𝑹˘0,𝜾˘0)=([R˘0,θ˘0],[ι˘0,θ0,θ˘0])(\breve{\bm{R}}_{0},\breve{\bm{\iota}}_{0})=([\breve{R}_{0},\breve{\theta}_{0}],[\breve{\iota}_{0},\theta_{0},\breve{\theta}_{0}]) stands for the natural compact continuation of 𝑹0{\bm{R}}_{0}. If n0n_{0} is the number of components of R˘0\partial\breve{R}_{0}, then R^0\hat{R}_{0} is a closed Riemann surface of genus g^:=2g+n01\hat{g}:=2g+n_{0}-1. We consider R˘0\breve{R}_{0} as a subdomain of R^0\hat{R}_{0}, and denote by ι\iota the inclusion mapping of R˘0\breve{R}_{0} into R^0\hat{R}_{0}. Then ι^0:=ιι˘0\hat{\iota}_{0}:=\iota\circ\breve{\iota}_{0} is a conformal embedding of R0R_{0} into R^0\hat{R}_{0}. Fix θ^0Homeo+(Σg^,R^0)\hat{\theta}_{0}\in\operatorname{Homeo}^{+}(\Sigma_{\hat{g}},\hat{R}_{0}) to obtain a marked closed Riemann surface 𝑹^0=[R^0,θ^0|Σ˙g^]\hat{\bm{R}}_{0}=[\hat{R}_{0},\hat{\theta}_{0}|_{\dot{\Sigma}_{\hat{g}}}] of genus g^\hat{g}.

Let 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}), and take 𝝋˘=[φ˘,θ˘]\breve{\bm{\varphi}}=[\breve{\varphi},\breve{\theta}] in AL(𝑹˘0)A_{L}(\breve{\bm{R}}_{0}) so that 𝝋=𝜾˘0𝝋˘{\bm{\varphi}}=\breve{\bm{\iota}}_{0}^{*}\breve{\bm{\varphi}}. The reflection principle enables us to extend φ˘\breve{\varphi} to a holomorphic quadratic differential φ^\hat{\varphi} on R^0\hat{R}_{0}. Set =𝑹0(𝝋)\mathcal{F}=\mathcal{H}_{{\bm{R}}_{0}}({\bm{\varphi}}) and ^=𝑹^0(𝝋^)\hat{\mathcal{F}}=\mathcal{H}_{\hat{\bm{R}}_{0}}(\hat{\bm{\varphi}}), where 𝝋^=[φ^,θ^0|Σ˙g^]\hat{\bm{\varphi}}=[\hat{\varphi},\hat{\theta}_{0}|_{\dot{\Sigma}_{\hat{g}}}]. Let γ𝒮(Σg)\gamma\in\mathscr{S}(\Sigma_{g}). Choose a simple loop c0c_{0} in the homotopy class γ\gamma so that c0c_{0} stays in Σ˙g\dot{\Sigma}_{g}, and denote by γ^\hat{\gamma} the free homotopy class in 𝒮(Σg^)\mathscr{S}(\Sigma_{\hat{g}}) for which (ι^0θ0)c0(θ^0)γ^(\hat{\iota}_{0}\circ\theta_{0})_{*}c_{0}\in(\hat{\theta}_{0})_{*}\hat{\gamma}.

We then claim that ^(γ^)=(γ)\hat{\mathcal{F}}(\hat{\gamma})=\mathcal{F}(\gamma). To verify the claim take a point p˘0\breve{p}_{0} in (R˘0)(\breve{R}_{0})^{\circ}, and construct a covering Riemann surface SS of R^0\hat{R}_{0} corresponding to the subgroup ιπ1(R˘0,p˘0)\iota_{*}\pi_{1}(\breve{R}_{0},\breve{p}_{0}) of the fundamental group π1(R^0,p˘0)\pi_{1}(\hat{R}_{0},\breve{p}_{0}). Let Π:SR^0\Pi:S\to\hat{R}_{0} denote the holomorphic covering map. For some component S0S_{0} of Π1(R˘0)\Pi^{-1}(\breve{R}_{0}) we have Π|S0CHomeo(S0,R˘0)\Pi|_{S_{0}}\in\operatorname{CHomeo}(S_{0},\breve{R}_{0}). Observe that each component of SS0S\setminus S_{0} is a doubly connected planar domain on SS. Therefore, if c:[0,1]R^0c:[0,1]\to\hat{R}_{0} is a loop in (θ^0)γ^(\hat{\theta}_{0})_{*}\hat{\gamma} with c(0)(R˘0)c(0)\in(\breve{R}_{0})^{\circ} and leaves from R˘0\breve{R}_{0} across a component CC of the border R˘0\partial\breve{R}_{0}, that is, c(t0)Cc(t_{0})\in C and c((t0,t1))R˘0=c((t_{0},t_{1}))\cap\breve{R}_{0}=\varnothing for some t0,t1t_{0},t_{1} with 0<t0<t1<10<t_{0}<t_{1}<1, then cc returns to R˘0\breve{R}_{0} across the same component CC in such a way that the subarc of cc between the leaving point and the returning point is homotopic to an arc on CC through a homotopy fixing the endpoints, or more precisely, there is t2[t1,1)t_{2}\in[t_{1},1) together with HCont([t0,t2]×[0,1],R^0)H\in\operatorname{Cont}([t_{0},t_{2}]\times[0,1],\hat{R}_{0}) such that H(t,0)=c(t)H(t,0)=c(t), H(t,1)CH(t,1)\in C for t[t0,t2]t\in[t_{0},t_{2}] and H(tj,s)=c(tj)H(t_{j},s)=c(t_{j}), j=0,2j=0,2, for s[0,1]s\in[0,1]. In other words, the loop c:[0,1]R^0c^{\prime}:[0,1]\to\hat{R}_{0} defined by c(t)=H(t,1)c^{\prime}(t)=H(t,1) for t[t1,t2]t\in[t_{1},t_{2}] and c(t)=c(t)c^{\prime}(t)=c(t) otherwise is freely homotopic to cc on R^0\hat{R}_{0}. Because φ^\hat{\varphi} are positive along CC, we obtain Hφ^(c)Hφ^(c)H_{\hat{\varphi}}(c^{\prime})\leqq H_{\hat{\varphi}}(c), which proves the claim.

Now, let {𝝋n}\{{\bm{\varphi}}_{n}\} be a sequence in AL(𝑹0)A_{L}({\bm{R}}_{0}) converging to 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}), and set n=𝑹0(𝝋n)\mathcal{F}_{n}=\mathcal{H}_{{\bm{R}}_{0}}({\bm{\varphi}}_{n}) and =𝑹0(𝝋)\mathcal{F}=\mathcal{H}_{{\bm{R}}_{0}}({\bm{\varphi}}). The quadratic differentials 𝝋n{\bm{\varphi}}_{n} and 𝝋\bm{\varphi} induce quadratic differentials 𝝋^n\hat{\bm{\varphi}}_{n} and 𝝋^\hat{\bm{\varphi}} on 𝑹^0\hat{\bm{R}}_{0} and measured foliations ^n\hat{\mathcal{F}}_{n} and ^\hat{\mathcal{F}} on Σg^\Sigma_{\hat{g}}. Since {𝝋^n}\{\hat{\bm{\varphi}}_{n}\} converges to 𝝋^\hat{\bm{\varphi}}, it follows from Proposition 6.2 that n(γ)=^n(γ^)^(γ^)=(γ)\mathcal{F}_{n}(\gamma)=\hat{\mathcal{F}}_{n}(\hat{\gamma})\to\hat{\mathcal{F}}(\hat{\gamma})=\mathcal{F}(\gamma) as nn\to\infty for γ𝒮(Σg)\gamma\in\mathscr{S}(\Sigma_{g}), which means that {n}\{\mathcal{F}_{n}\} converges to \mathcal{F}. We have proved that 𝑹0\mathcal{H}_{{\bm{R}}_{0}} is continuous.

Next, we show that 𝑸𝑹0{\bm{Q}}_{{\bm{R}}_{0}} is also continuous. Fix 𝑺0𝔗g{\bm{S}}_{0}\in\mathfrak{T}_{g}. Since 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is compact by Corollary 7.6, there is d>0d>0 such that dT(𝑹,𝑺0)dd_{T}({\bm{R}},{\bm{S}}_{0})\leqq d for all 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}). It thus follows from Lemma 9.9 that

e2dExt(𝑺0)Ext(𝑹)e2dExt(𝑺0)e^{-2d}\operatorname{Ext}_{\mathcal{F}}({\bm{S}}_{0})\leqq\operatorname{Ext}_{\mathcal{F}}({\bm{R}})\leqq e^{2d}\operatorname{Ext}_{\mathcal{F}}({\bm{S}}_{0})

for 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}) and (Σg)\mathcal{F}\in\mathscr{MF}(\Sigma_{g}). Let {n}\{\mathcal{F}_{n}\} be a sequence in (Σg){0}\mathscr{MF}(\Sigma_{g})\setminus\{0\} converging to (Σg){0}\mathcal{F}\in\mathscr{MF}(\Sigma_{g})\setminus\{0\}, and set 𝝋n=𝑸𝑹0(n){\bm{\varphi}}_{n}={\bm{Q}}_{{\bm{R}}_{0}}(\mathcal{F}_{n}). Let (𝑹n,𝜾n)({\bm{R}}_{n},{\bm{\iota}}_{n}) be a closed 𝝋n{\bm{\varphi}}_{n}-regular self-welding continuation of 𝑹0{\bm{R}}_{0}. Since Proposition 6.2 implies that

Extn(𝑺0)=𝑸𝑺0(n)𝑺0𝑸𝑺0()𝑺0=Ext(𝑺0)0\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{S}}_{0})=\|{\bm{Q}}_{{\bm{S}}_{0}}(\mathcal{F}_{n})\|_{{\bm{S}}_{0}}\to\|{\bm{Q}}_{{\bm{S}}_{0}}(\mathcal{F})\|_{{\bm{S}}_{0}}=\operatorname{Ext}_{\mathcal{F}}({\bm{S}}_{0})\neq 0

as nn\to\infty, the sequence {𝝋n𝑹0}={Extn(𝑹n)}\{\|{\bm{\varphi}}_{n}\|_{{\bm{R}}_{0}}\}=\{\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{R}}_{n})\} is bounded and bounded away from 0. With the aid of Proposition 4.1 together with normal family arguments we infer that any subsequence of {𝝋n}\{{\bm{\varphi}}_{n}\} contains a subsequence converging to some 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}). Since 𝑹0\mathcal{H}_{{\bm{R}}_{0}} is continuous, we obtain 𝑹0(𝝋)=\mathcal{H}_{{\bm{R}}_{0}}({\bm{\varphi}})=\mathcal{F}, or 𝝋=𝑸𝑹0(){\bm{\varphi}}={\bm{Q}}_{{\bm{R}}_{0}}(\mathcal{F}). Consequently, 𝝋n𝝋{\bm{\varphi}}_{n}\to{\bm{\varphi}} as nn\to\infty, proving that 𝑸𝑹0{\bm{Q}}_{{\bm{R}}_{0}} is continuous. This completes the proof. ∎

Proof of Theorem 1.6.

We begin with the proof of (i). Let (Σg){0}\mathcal{F}\in\mathscr{MF}(\Sigma_{g})\setminus\{0\}, and set 𝝋=𝑸𝑹0(){\bm{\varphi}}={\bm{Q}}_{{\bm{R}}_{0}}(\mathcal{F}). Equality (1.1) follows from Proposition 12.2. Any point of 𝔐𝝋(𝑹0)\mathfrak{M}_{\bm{\varphi}}({\bm{R}}_{0}) is a maximal point for \mathcal{F} on 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) by Theorem 6.3, and such points exhaust the maximal set for \mathcal{F} on 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) by Proposition 12.2. This proves (i).

Assertion (ii) follows from the remark following the proof of Theorem 1.4 (iii) (see §10). We have only to note that 𝔐(𝑹0)=𝔐𝝋(𝑹0)\mathfrak{M}_{\mathcal{F}}({\bm{R}}_{0})=\mathfrak{M}_{\bm{\varphi}}({\bm{R}}_{0}) if =𝑹0(𝝋)\mathcal{F}=\mathcal{H}_{{\bm{R}}_{0}}({\bm{\varphi}}).

To prove (iii) let 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}). Then we infer from Theorem 6.3 that

Ext(𝑹)𝑸𝑹0()𝑹0=maxExt(𝔐(𝑹0))\operatorname{Ext}_{\mathcal{F}}({\bm{R}})\leqq\|{\bm{Q}}_{{\bm{R}}_{0}}(\mathcal{F})\|_{{\bm{R}}_{0}}=\max\operatorname{Ext}_{\mathcal{F}}(\partial\mathfrak{M}({\bm{R}}_{0}))

holds for all (Σg){0}\mathcal{F}\in\mathscr{MF}(\Sigma_{g})\setminus\{0\}.

To show the converse assume that 𝑹𝔗g𝔐(𝑹0){\bm{R}}\in\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}). Let 𝑹{\bm{R}}^{\prime} be the point of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) nearest to 𝑹\bm{R}. The point 𝑹\bm{R} lies on an Ioffe ray 𝒓𝑹[𝝍]{\bm{r}}_{{\bm{R}}^{\prime}}[{\bm{\psi}}], where 𝝍A(𝑹){\bm{\psi}}\in A({\bm{R}}^{\prime}) (see Proposition 9.6). Then 𝝍\bm{\psi} is the co-welder of a welder 𝝋AL(𝑹0){\bm{\varphi}}\in A_{L}({\bm{R}}_{0}) of some 𝝋\bm{\varphi}-regular self-welding continuation (𝑹,𝜾)({\bm{R}}^{\prime},{\bm{\iota}}) of 𝑹0{\bm{R}}_{0}. Set =𝑹0(𝝋)\mathcal{F}=\mathcal{H}_{{\bm{R}}_{0}}({\bm{\varphi}}). Since 𝝍=Q𝑹(){\bm{\psi}}=Q_{{\bm{R}}^{\prime}}(\mathcal{F}) and 𝝍𝑹=𝝋𝑹0\|{\bm{\psi}}\|_{{\bm{R}}^{\prime}}=\|{\bm{\varphi}}\|_{{\bm{R}}_{0}}, we apply (6.4) to obtain

Ext(𝑹)=e2dT(𝑹,𝑹)Ext(𝑹)=e2dT(𝑹,𝑹)𝝍𝑹>𝝍𝑹=maxExt(𝔐(𝑹0)).\operatorname{Ext}_{\mathcal{F}}({\bm{R}})=e^{2d_{T}({\bm{R}},{\bm{R}}^{\prime})}\operatorname{Ext}_{\mathcal{F}}({\bm{R}}^{\prime})=e^{2d_{T}({\bm{R}},{\bm{R}}^{\prime})}\|{\bm{\psi}}\|_{{\bm{R}}^{\prime}}>\|{\bm{\psi}}\|_{{\bm{R}}^{\prime}}=\max\operatorname{Ext}_{\mathcal{F}}(\partial\mathfrak{M}({\bm{R}}_{0})).

This completes the proof. ∎

Example 12.3.

We consider the case of genus one. We identify 𝔗1\mathfrak{T}_{1} with \mathbb{H} as in Example 3.14, and use the notations in Example 6.7. For t(1,1]t\in(-1,1] set t=𝑹0(𝜾˘0𝝋t)(Σ1)\mathcal{F}_{t}=\mathcal{H}_{{\bm{R}}_{0}}(\breve{\bm{\iota}}_{0}^{*}{\bm{\varphi}}_{t})\in\mathscr{MF}(\Sigma_{1}), and define ct=cot(tπ/2)c_{t}=-\cot(t\pi/2), where c0=c_{0}=\infty. Then the level curves of the function Extt\operatorname{Ext}_{\mathcal{F}_{t}} are the horocycles in \mathbb{H} with center at ctc_{t}, that is, the circles in \mathbb{H} tangent to {}\mathbb{R}\cup\{\infty\} at ctc_{t}. Theorem 1.6 (iii) implies that 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is the envelope of the family of horocycles Extt(𝑺)=maxt(𝔐(𝑹0))\operatorname{Ext}_{\mathcal{F}_{t}}({\bm{S}})=\max\mathcal{F}_{t}(\mathfrak{M}({\bm{R}}_{0})), t(1,1]t\in(-1,1]. Such a description of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) first appeared in [36, §5]. Since 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is a closed disk, three of the horocycles, or equivalently, three of the measured foliations t\mathcal{F}_{t} are sufficient to determine it; see [35, Theorem 1], where maxt(𝔐(𝑹0))\max\mathcal{F}_{t}(\mathfrak{M}({\bm{R}}_{0})) is given by the extremal length of a weak homology class of R0R_{0}.

13 Shape of 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) with K>1K>1

Let 𝑹0{\bm{R}}_{0} be a marked finite open Riemann surface of positive genus gg. The aim of the present section is to prove Theorem 1.2.

For 𝑺𝔗g{\bm{S}}\in\mathfrak{T}_{g} let 𝑯K[𝑹0](𝑺){\bm{H}}_{K}[{\bm{R}}_{0}]({\bm{S}}) denote the point of 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) nearest to 𝑺\bm{S}. By Corollary 9.11 this gives a well-defined mapping 𝑯K[𝑹0]{\bm{H}}_{K}[{\bm{R}}_{0}] of 𝔗g\mathfrak{T}_{g} onto 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}). It fixes 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) pointwise and maps 𝔗g𝔐K(𝑹0)\mathfrak{T}_{g}\setminus\mathfrak{M}_{K}({\bm{R}}_{0}) into 𝔐K(𝑹0)\partial\mathfrak{M}_{K}({\bm{R}}_{0}). We abbreviate 𝑯1[𝑹0]{\bm{H}}_{1}[{\bm{R}}_{0}] to 𝑯[𝑹0]{\bm{H}}[{\bm{R}}_{0}].

Proposition 13.1.

If 𝐑0{\bm{R}}_{0} is a marked finite open Riemann surface of positive genus gg, then 𝐇K[𝐑0]:𝔗g𝔐K(𝐑0){\bm{H}}_{K}[{\bm{R}}_{0}]:\mathfrak{T}_{g}\to\mathfrak{M}_{K}({\bm{R}}_{0}) is a retraction.

Proof.

Since 𝑯K[𝑹0]{\bm{H}}_{K}[{\bm{R}}_{0}] fixes 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) pointwise, we have only to show that it is continuous on 𝔗gInt𝔐K(𝑹0)\mathfrak{T}_{g}\setminus\operatorname{Int}\mathfrak{M}_{K}({\bm{R}}_{0}). If 𝑺𝔗gInt𝔐K(𝑹0){\bm{S}}\in\mathfrak{T}_{g}\setminus\operatorname{Int}\mathfrak{M}_{K}({\bm{R}}_{0}), then there is an Ioffe ray 𝒓\bm{r} of 𝑹0{\bm{R}}_{0} such that 𝒓(d0)=𝑺{\bm{r}}(d_{0})={\bm{S}} for some d0d_{0}; note that d0=dT(𝑺,𝔐(𝑹0))(logK)/2d_{0}=d_{T}({\bm{S}},\mathfrak{M}({\bm{R}}_{0}))\geqq(\log K)/2 by (9.4). Then 𝑯K[𝑹0](𝑺)=𝒓((logK)/2){\bm{H}}_{K}[{\bm{R}}_{0}]({\bm{S}})={\bm{r}}((\log K)/2) by (9.3) and Proposition 9.12.

Let {𝑺n}\{{\bm{S}}_{n}\} be a sequence in 𝔗gInt𝔐K(𝑹0)\mathfrak{T}_{g}\setminus\operatorname{Int}\mathfrak{M}_{K}({\bm{R}}_{0}) converging to 𝑺\bm{S}. Take 𝒓n(𝑹0){\bm{r}}_{n}\in\mathscr{I}({\bm{R}}_{0}) and tn(logK)/2t_{n}\geqq(\log K)/2 so that 𝒓n(tn)=𝑺n{\bm{r}}_{n}(t_{n})={\bm{S}}_{n}. Then 𝑯K[𝑹0](𝑺n)=𝒓n((logK)/2){\bm{H}}_{K}[{\bm{R}}_{0}]({\bm{S}}_{n})={\bm{r}}_{n}((\log K)/2). The sequence {tn}\{t_{n}\} converges to d0d_{0}, for,

limntn=limndT(𝑺n,𝔐(𝑹0))=dT(𝑺,𝔐(𝑹0))=d0.\lim_{n\to\infty}t_{n}=\lim_{n\to\infty}d_{T}({\bm{S}}_{n},\mathfrak{M}({\bm{R}}_{0}))=d_{T}({\bm{S}},\mathfrak{M}({\bm{R}}_{0}))=d_{0}.

Since 𝔐K(𝑹0)\partial\mathfrak{M}_{K}({\bm{R}}_{0}) is compact, a subsequence {𝑯K[𝑹0](𝑺nk)}\{{\bm{H}}_{K}[{\bm{R}}_{0}]({\bm{S}}_{n_{k}})\} converges to some point 𝑺{\bm{S}}^{\prime} on 𝔐K(𝑹0)\partial\mathfrak{M}_{K}({\bm{R}}_{0}). Then we obtain

dT(𝑺,𝑺)\displaystyle d_{T}({\bm{S}},{\bm{S}}^{\prime}) =limkdT(𝑺nk,𝑯K[𝑹0](𝑺nk))=limkdT(𝒓nk(tnk),𝒓nk((logK)/2))\displaystyle=\lim_{k\to\infty}d_{T}({\bm{S}}_{n_{k}},{\bm{H}}_{K}[{\bm{R}}_{0}]({\bm{S}}_{n_{k}}))=\lim_{k\to\infty}d_{T}({\bm{r}}_{n_{k}}(t_{n_{k}}),{\bm{r}}_{n_{k}}((\log K)/2))
=d0(logK)/2=dT(𝑺,𝔐K(𝑹0))\displaystyle=d_{0}-(\log K)/2=d_{T}({\bm{S}},\mathfrak{M}_{K}({\bm{R}}_{0}))

by (9.4). Hence 𝑺{\bm{S}}^{\prime} is the point of 𝔐K(𝑹0)\mathfrak{M}_{K}({\bm{R}}_{0}) nearest to 𝑺\bm{S}, or, 𝑯K[𝑹0](𝑺)=𝑺{\bm{H}}_{K}[{\bm{R}}_{0}]({\bm{S}})={\bm{S}}^{\prime}. This completes the proof. ∎

For K>K1K^{\prime}>K\geqq 1 let 𝑯KK[𝑹0]{\bm{H}}^{K^{\prime}}_{K}[{\bm{R}}_{0}] denote the restriction of 𝑯K[𝑹0]{\bm{H}}_{K}[{\bm{R}}_{0}] to 𝔐K(𝑹0)\partial\mathfrak{M}_{K^{\prime}}({\bm{R}}_{0}). We consider it as a mapping of 𝔐K(𝑹0)\partial\mathfrak{M}_{K^{\prime}}({\bm{R}}_{0}) into 𝔐K(𝑹0)\partial\mathfrak{M}_{K}({\bm{R}}_{0}).

Corollary 13.2.

If 𝐑0{\bm{R}}_{0} is a marked finite open Riemann surface of genus gg, then 𝐇KK[𝐑0]Homeo(𝔐K(𝐑0),𝔐K(𝐑0)){\bm{H}}^{K^{\prime}}_{K}[{\bm{R}}_{0}]\in\operatorname{Homeo}(\partial\mathfrak{M}_{K^{\prime}}({\bm{R}}_{0}),\partial\mathfrak{M}_{K}({\bm{R}}_{0})) for K>K>1K^{\prime}>K>1.

Proof.

Theorem 9.4 implies that 𝑯KK[𝑹0]{\bm{H}}^{K^{\prime}}_{K}[{\bm{R}}_{0}] is a bijection. Since 𝔐K(𝑹0)\partial\mathfrak{M}_{K^{\prime}}({\bm{R}}_{0}) is compact and 𝔐K(𝑹0)\partial\mathfrak{M}_{K}({\bm{R}}_{0}) is Hausdorff, the continuous mapping 𝑯KK[𝑹0]{\bm{H}}^{K^{\prime}}_{K}[{\bm{R}}_{0}] is in fact a homeomorphism of 𝔐K(𝑹0)\partial\mathfrak{M}_{K^{\prime}}({\bm{R}}_{0}) onto 𝔐K(𝑹0)\partial\mathfrak{M}_{K}({\bm{R}}_{0}). ∎

Remark.

The continuous mapping 𝑯1K[𝑹0]:𝔐K(𝑹0)𝔐(𝑹0){\bm{H}}^{K}_{1}[{\bm{R}}_{0}]:\partial\mathfrak{M}_{K}({\bm{R}}_{0})\to\partial\mathfrak{M}({\bm{R}}_{0}) is not necessarily injective (see the remark after Definition 9.3) though it is always surjective.

Proposition 13.3.

If 𝐑0{\bm{R}}_{0} is a marked finite open Riemann surface of genus gg, then for each K>1K>1 there is a homeomorphism of 𝔗g\mathfrak{T}_{g} onto 2dg\mathbb{R}^{2d_{g}} that maps the compact sets 𝔐K(𝐑0)\mathfrak{M}_{K^{\prime}}({\bm{R}}_{0}), KKK^{\prime}\geqq K, onto concentric closed balls, where dg=max{g,3g3}d_{g}=\max\{g,3g-3\}.

The proof of the proposition requires two lemmas. For the proof of the first lemma we need the following result of Earle. For 𝑹,𝑺𝔗g{\bm{R}},{\bm{S}}\in\mathfrak{T}_{g} with 𝑹𝑺{\bm{R}}\neq{\bm{S}} let 𝑸𝑹[𝑺]{\bm{Q}}_{\bm{R}}[{\bm{S}}] denote the element of A(𝑹)A({\bm{R}}) with 𝑸𝑹[𝑺]𝑹=1\|{\bm{Q}}_{\bm{R}}[{\bm{S}}]\|_{\bm{R}}=1 such that 𝒓𝑹[𝑸𝑹[𝑺]]=𝒓𝑹[𝑺]{\bm{r}}_{\bm{R}}[{\bm{Q}}_{\bm{R}}[{\bm{S}}]]={\bm{r}}_{\bm{R}}[{\bm{S}}].

Proposition 13.4 (Earle [13]).

The Teichmüller distance function dTd_{T} is of class C1C^{1} on 𝔗g×𝔗g\mathfrak{T}_{g}\times\mathfrak{T}_{g} off the diagonal. The differential of the function 𝐑dT(𝐑,𝐒){\bm{R}}\mapsto d_{T}({\bm{R}},{\bm{S}}) is 𝐐𝐑[𝐒]-{\bm{Q}}_{\bm{R}}[{\bm{S}}].

Thus if (R,θ)(R,\theta) and (ψ,θ)(\psi,\theta) represent 𝑹\bm{R} and 𝑸𝑹[𝑺]{\bm{Q}}_{\bm{R}}[{\bm{S}}], respectively, then the differential of dT(,𝑺)d_{T}(\,\cdot\,,{\bm{S}}) at 𝑹\bm{R} is the linear functional

μReRμψ\mu\mapsto-\operatorname{Re}\int_{R}\mu\psi

on the space of bounded measurable (1,1)(-1,1)-forms μ\mu on RR. Note that the inequality ReRμψμ-\operatorname{Re}\int_{R}\mu\psi\leqq\|\mu\|_{\infty} holds and that the sign of equality occurs if μ=|ψ|/ψ\mu=-|\psi|/\psi. Since the tangent vector to 𝒓𝑹[𝑸𝑹[𝑺]]{\bm{r}}_{\bm{R}}[{\bm{Q}}_{\bm{R}}[{\bm{S}}]] at the initial point is represented by |ψ|/ψ|\psi|/\psi, we see that the tangent vector varies continuously with 𝑹{\bm{R}} and 𝑺{\bm{S}}.

We now turn to our first lemma. Recall that the image of the interval (t1,t2)(t_{1},t_{2}) by a Teichmüller geodesic ray 𝒓\bm{r} is denoted by 𝒓(t1,t2){\bm{r}}(t_{1},t_{2}) (see (6.3)).

Lemma 13.5.

Let 𝐑\bm{R} and 𝐒\bm{S} be distinct points of 𝔗g\mathfrak{T}_{g}. Then for ρ>0\rho>0 there are neighborhoods 𝔘\mathfrak{U} and 𝔙\mathfrak{V} of 𝐑\bm{R} and 𝐒\bm{S}, respectively, such that for any 𝐑,𝐑′′𝔘{\bm{R}}^{\prime},{\bm{R}}^{\prime\prime}\in\mathfrak{U} and 𝐒𝔙{\bm{S}}^{\prime}\in\mathfrak{V} the inclusion relation

(13.1) 𝒓′′(τ′′,τ′′+ρ)𝔅ρ(𝒓(τ+ρ)){\bm{r}}^{\prime\prime}(\tau^{\prime\prime},\tau^{\prime\prime}+\rho)\subset\mathfrak{B}_{\rho}({\bm{r}}^{\prime}(\tau^{\prime}+\rho))

holds, where 𝐫=𝐫𝐑[𝐒]{\bm{r}}^{\prime}={\bm{r}}_{{\bm{R}}^{\prime}}[{\bm{S}}^{\prime}], 𝐫′′=𝐫𝐑′′[𝐒]{\bm{r}}^{\prime\prime}={\bm{r}}_{{\bm{R}}^{\prime\prime}}[{\bm{S}}^{\prime}], τ=dT(𝐑,𝐒)\tau^{\prime}=d_{T}({\bm{R}}^{\prime},{\bm{S}}^{\prime}) and τ′′=dT(𝐑′′,𝐒)\tau^{\prime\prime}=d_{T}({\bm{R}}^{\prime\prime},{\bm{S}}^{\prime}).

Proof.

Set 𝒓=𝒓𝑹[𝑺]{\bm{r}}={\bm{r}}_{{\bm{R}}}[{\bm{S}}] and τ=dT(𝑹,𝑺)\tau=d_{T}({\bm{R}},{\bm{S}}). Since the derivative of the function

tdT(𝒓(t),𝒓(τ+ρ))=|τ+ρt|t\mapsto d_{T}({\bm{r}}(t),{\bm{r}}(\tau+\rho))=|\tau+\rho-t|

is 1-1 on the interval (0,τ+ρ)(0,\tau+\rho), Proposition 13.4 implies that there are neighborhoods 𝔘\mathfrak{U} and 𝔙\mathfrak{V} of 𝑹\bm{R} and 𝑺\bm{S}, respectively, together with δ(0,ρ)\delta\in(0,\rho) such that if 𝑹,𝑹′′𝔘{\bm{R}}^{\prime},{\bm{R}}^{\prime\prime}\in\mathfrak{U} and 𝑺𝔙{\bm{S}}^{\prime}\in\mathfrak{V}, then the derivative of the function

tdT(𝒓′′(t),𝒓(τ+ρ))t\mapsto d_{T}({\bm{r}}^{\prime\prime}(t),{\bm{r}}^{\prime}(\tau^{\prime}+\rho))

is negative on the interval (τ′′δ,τ′′+δ)(\tau^{\prime\prime}-\delta,\tau^{\prime\prime}+\delta), where 𝒓=𝒓𝑹[𝑺]{\bm{r}}^{\prime}={\bm{r}}_{{\bm{R}}^{\prime}}[{\bm{S}}^{\prime}], 𝒓′′=𝒓𝑹′′[𝑺]{\bm{r}}^{\prime\prime}={\bm{r}}_{{\bm{R}}^{\prime\prime}}[{\bm{S}}^{\prime}], τ=dT(𝑹,𝑺)\tau^{\prime}=d_{T}({\bm{R}}^{\prime},{\bm{S}}^{\prime}) and τ′′=dT(𝑹′′,𝑺)\tau^{\prime\prime}=d_{T}({\bm{R}}^{\prime\prime},{\bm{S}}^{\prime}). Since 𝒓′′(τ′′)=𝑺=𝒓(τ){\bm{r}}^{\prime\prime}(\tau^{\prime\prime})={\bm{S}}^{\prime}={\bm{r}}^{\prime}(\tau^{\prime}), it follows that

dT(𝒓′′(t),𝒓(τ+ρ))<dT(𝒓′′(τ′′),𝒓(τ+ρ))=ρd_{T}({\bm{r}}^{\prime\prime}(t),{\bm{r}}^{\prime}(\tau^{\prime}+\rho))<d_{T}({\bm{r}}^{\prime\prime}(\tau^{\prime\prime}),{\bm{r}}^{\prime}(\tau^{\prime}+\rho))=\rho

for t(τ′′,τ′′+δ)t\in(\tau^{\prime\prime},\tau^{\prime\prime}+\delta), or equivalently, that

(13.2) 𝒓′′(τ′′,τ′′+δ)𝔅ρ(𝒓(τ+ρ)).{\bm{r}}^{\prime\prime}(\tau^{\prime\prime},\tau^{\prime\prime}+\delta)\subset\mathfrak{B}_{\rho}({\bm{r}}^{\prime}(\tau^{\prime}+\rho)).

Since

𝒓(τ+δ,τ+ρ)𝔅ρ(𝒓(τ+ρ)){\bm{r}}(\tau+\delta,\tau+\rho)\subset\mathfrak{B}_{\rho}({\bm{r}}(\tau+\rho))

and the Teichmüller distance between 𝒓(τ+δ,τ+ρ){\bm{r}}(\tau+\delta,\tau+\rho) and 𝔅ρ(𝒓(τ+ρ))\partial\mathfrak{B}_{\rho}({\bm{r}}(\tau+\rho)) is exactly δ>0\delta>0, we can replace the neighborhoods 𝔘\mathfrak{U} and 𝔙\mathfrak{V} with smaller ones to ensure that

(13.3) 𝒓′′(τ′′+δ,τ′′+ρ)𝔅ρ(𝒓(τ+ρ)){\bm{r}}^{\prime\prime}(\tau^{\prime\prime}+\delta,\tau^{\prime\prime}+\rho)\subset\mathfrak{B}_{\rho}({\bm{r}}^{\prime}(\tau^{\prime}+\rho))

for 𝑹,𝑹′′𝔘{\bm{R}}^{\prime},{\bm{R}}^{\prime\prime}\in\mathfrak{U} and 𝑺𝔙{\bm{S}}^{\prime}\in\mathfrak{V}. Now inclusion relation (13.1) is an immediate consequence of (13.2) and (13.3). ∎

Lemma 13.6.

Let 𝔎\mathfrak{K} and 𝔏\mathfrak{L} be disjoint compact sets in 𝔗g\mathfrak{T}_{g}, and let ρ>0\rho>0. Then there is ε0>0\varepsilon_{0}>0 such that for any 𝐑𝔎{\bm{R}}\in\mathfrak{K}, 𝐒𝔏{\bm{S}}\in\mathfrak{L} and 𝐑𝔎𝔅ε0(𝐑){\bm{R}}^{\prime}\in\mathfrak{K}\cap\mathfrak{B}_{\varepsilon_{0}}({\bm{R}}) the inclusion relation

(13.4) 𝒓(τ,τ+ρ)𝔅ρ(𝒓(τ+ρ)),{\bm{r}}^{\prime}(\tau^{\prime},\tau^{\prime}+\rho)\subset\mathfrak{B}_{\rho}({\bm{r}}(\tau+\rho)),

holds, where 𝐫=𝐫𝐑[𝐒]{\bm{r}}={\bm{r}}_{{\bm{R}}}[{\bm{S}}], 𝐫=𝐫𝐑[𝐒]{\bm{r}}^{\prime}={\bm{r}}_{{\bm{R}}^{\prime}}[{\bm{S}}], τ=dT(𝐑,𝐒)\tau=d_{T}({\bm{R}},{\bm{S}}) and τ=dT(𝐑,𝐒)\tau^{\prime}=d_{T}({\bm{R}}^{\prime},{\bm{S}}).

Proof.

For each 𝑹𝔎{\bm{R}}\in\mathfrak{K} we choose a neighborhood 𝔚𝑹\mathfrak{W}_{\bm{R}} of 𝑹\bm{R} as follows. For 𝑺𝔏{\bm{S}}\in\mathfrak{L} take neighborhoods 𝔘=𝔘𝑺\mathfrak{U}=\mathfrak{U}_{\bm{S}} of 𝑹\bm{R} and 𝔙=𝔙𝑺\mathfrak{V}=\mathfrak{V}_{\bm{S}} of 𝑺\bm{S} as in Lemma 13.5. As 𝔏\mathfrak{L} is compact, there are finitely many points 𝑺1,,𝑺ν𝔏{\bm{S}}_{1},\ldots,{\bm{S}}_{\nu}\in\mathfrak{L} such that the corresponding open sets 𝔙𝑺1,,𝔙𝑺ν\mathfrak{V}_{{\bm{S}}_{1}},\ldots,\mathfrak{V}_{{\bm{S}}_{\nu}} cover 𝔏\mathfrak{L}. Then we set 𝔚𝑹=k=1ν𝔘𝑺k\mathfrak{W}_{\bm{R}}=\bigcap_{k=1}^{\nu}\mathfrak{U}_{{\bm{S}}_{k}}. It is a neighborhood of 𝑹\bm{R} such that (13.1) holds for 𝑹,𝑹′′𝔚𝑹{\bm{R}}^{\prime},{\bm{R}}^{\prime\prime}\in\mathfrak{W}_{\bm{R}} and 𝑺𝔏{\bm{S}}^{\prime}\in\mathfrak{L}. Then any Lebesgue number ε0\varepsilon_{0} of the open covering {𝔚𝑹}𝑹𝔎\{\mathfrak{W}_{\bm{R}}\}_{{\bm{R}}\in\mathfrak{K}} of 𝔎\mathfrak{K} possesses the required properties. ∎

We are now ready to prove Proposition 13.3.

Proof of Proposition 13.3.

If 𝑹0{\bm{R}}_{0} is analytically finite, then 𝔐(𝑹0)={𝑾}\mathfrak{M}({\bm{R}}_{0})=\{{\bm{W}}\} for some 𝑾𝔗g{\bm{W}}\in\mathfrak{T}_{g} so that 𝔐K(𝑹0)=𝔅¯(logK)/2(𝑾)\mathfrak{M}_{K}({\bm{R}}_{0})=\bar{\mathfrak{B}}_{(\log K)/2}({\bm{W}}). Hence the proposition is valid in this case.

Assume now that 𝑹0{\bm{R}}_{0} is nonanalytically finite. Let {(𝑾t,ϵt)}t[0,1]\{({\bm{W}}_{t},{\bm{\epsilon}}_{t})\}_{t\in[0,1]} be a circular filling for 𝑹0{\bm{R}}_{0}. Then 𝔐(𝑾1)\mathfrak{M}({\bm{W}}_{1}) is a singleton, say, {𝑾}\{{\bm{W}}\}. Fix ρ>0\rho>0 satisfying

(13.5) 𝔐(𝑹0)𝔅ρ(𝑾),\mathfrak{M}({\bm{R}}_{0})\subset\mathfrak{B}_{\rho}({\bm{W}}),

and set 𝔏=𝔅¯4ρ(𝑾)𝔅2ρ(𝑾)\mathfrak{L}=\bar{\mathfrak{B}}_{4\rho}({\bm{W}})\setminus\mathfrak{B}_{2\rho}({\bm{W}}). It is a compact subset of 𝔗g\mathfrak{T}_{g} that does not meet 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). By Lemma 13.6 there is ε0>0\varepsilon_{0}>0 such that for any 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}({\bm{R}}_{0}), 𝑺𝔏{\bm{S}}\in\mathfrak{L} and 𝑹𝔐(𝑹0)𝔅ε0(𝑹){\bm{R}}^{\prime}\in\mathfrak{M}({\bm{R}}_{0})\cap\mathfrak{B}_{\varepsilon_{0}}({\bm{R}}) inclusion relation (13.4) holds.

We claim that for every t[0,1]t\in[0,1] there exists δ(t)>0\delta(t)>0 such that if tj[0,1]t_{j}\in[0,1] and |tjt|<δ(t)|t_{j}-t|<\delta(t) for j=1,2j=1,2, then

dT(𝑯[𝑾t1](𝑺),𝑯[𝑾t2](𝑺))<ε0d_{T}({\bm{H}}[{\bm{W}}_{t_{1}}]({\bm{S}}),{\bm{H}}[{\bm{W}}_{t_{2}}]({\bm{S}}))<\varepsilon_{0}

for all 𝑺𝔏{\bm{S}}\in\mathfrak{L}; recall that 𝑯[𝑾t](𝑺){\bm{H}}[{\bm{W}}_{t}]({\bm{S}}) stands for the point of 𝔐(𝑾t)\mathfrak{M}({\bm{W}}_{t}) nearest to 𝑺\bm{S}. If the claim were false, then there would exist sequences {tn},{tn}[0,1]\{t_{n}\},\{t^{\prime}_{n}\}\subset[0,1] and {𝑺n}𝔏\{{\bm{S}}_{n}\}\subset\mathfrak{L} such that

limntn=limntn=tanddT(𝑹n,𝑹n)ε0,\lim_{n\to\infty}t_{n}=\lim_{n\to\infty}t^{\prime}_{n}=t\quad\text{and}\quad d_{T}({\bm{R}}_{n},{\bm{R}}^{\prime}_{n})\geqq\varepsilon_{0},

where 𝑹n=𝑯[𝑾tn](𝑺n){\bm{R}}_{n}={\bm{H}}[{\bm{W}}_{t_{n}}]({\bm{S}}_{n}) and 𝑹n=𝑯[𝑾tn](𝑺n){\bm{R}}^{\prime}_{n}={\bm{H}}[{\bm{W}}_{t^{\prime}_{n}}]({\bm{S}}_{n}). By taking subsequences if necessary we may assume that the sequences {𝑺n}\{{\bm{S}}_{n}\}, {𝑹n}\{{\bm{R}}_{n}\} and {𝑹n}\{{\bm{R}}^{\prime}_{n}\} converge to 𝑺\bm{S}, 𝑹\bm{R} and 𝑹{\bm{R}}^{\prime} in 𝔗g\mathfrak{T}_{g}, respectively. Then 𝑺𝔏{\bm{S}}\in\mathfrak{L} and 𝑹,𝑹𝔐(𝑾t){\bm{R}},{\bm{R}}^{\prime}\in\mathfrak{M}({\bm{W}}_{t}) with

(13.6) dT(𝑹,𝑹)ε0.d_{T}({\bm{R}},{\bm{R}}^{\prime})\geqq\varepsilon_{0}.

For any ε>0\varepsilon>0 we have

dT(𝑺n,𝑺)<ε,dT(𝑹n,𝑹)<ε,𝔐(𝑾t)𝔐e2ε(𝑾tn)d_{T}({\bm{S}}_{n},{\bm{S}})<\varepsilon,\quad d_{T}({\bm{R}}_{n},{\bm{R}})<\varepsilon,\quad\mathfrak{M}({\bm{W}}_{t})\subset\mathfrak{M}_{e^{2\varepsilon}}({\bm{W}}_{t_{n}})

for sufficiently large nn. Set d=dT(𝑺,𝑹)d=d_{T}({\bm{S}},{\bm{R}}) and dn=dT(𝑺n,𝑹n)=dT(𝑺n,𝔐(𝑾tn))d_{n}=d_{T}({\bm{S}}_{n},{\bm{R}}_{n})=d_{T}({\bm{S}}_{n},\mathfrak{M}({\bm{W}}_{t_{n}})). Since

𝔅dn(𝑺n)𝔐(𝑾tn)=and𝔅d4ε(𝑺)𝔅dnε(𝑺n),\mathfrak{B}_{d_{n}}({\bm{S}}_{n})\cap\mathfrak{M}({\bm{W}}_{t_{n}})=\varnothing\qquad\text{and}\qquad\mathfrak{B}_{d-4\varepsilon}({\bm{S}})\subset\mathfrak{B}_{d_{n}-\varepsilon}({\bm{S}}_{n}),

we obtain 𝔅d4ε(𝑺)𝔐(𝑾t)=\mathfrak{B}_{d-4\varepsilon}({\bm{S}})\cap\mathfrak{M}({\bm{W}}_{t})=\varnothing, or equivalently,

dT(𝑺,𝑹)4εdT(𝑺,𝔐(𝑾t)).d_{T}({\bm{S}},{\bm{R}})-4\varepsilon\leqq d_{T}({\bm{S}},\mathfrak{M}({\bm{W}}_{t})).

As ε>0\varepsilon>0 is arbitrary, we know that dT(𝑺,𝑹)dT(𝑺,𝔐(𝑾t))d_{T}({\bm{S}},{\bm{R}})\leqq d_{T}({\bm{S}},\mathfrak{M}({\bm{W}}_{t})). Since 𝑹\bm{R} belongs to 𝔐(𝑾t)\mathfrak{M}({\bm{W}}_{t}), we conclude that 𝑯[𝑾t](𝑺)=𝑹{\bm{H}}[{\bm{W}}_{t}]({\bm{S}})={\bm{R}}. Similarly, we obtain 𝑯[𝑾t](𝑺)=𝑹{\bm{H}}[{\bm{W}}_{t}]({\bm{S}})={\bm{R}}^{\prime}, and hence 𝑹=𝑹{\bm{R}}={\bm{R}}^{\prime}, contradicting (13.6).

Let δ0>0\delta_{0}>0 be a Lebesgue number of the open covering {(tδ(t),t+δ(t))}t[0,1]\{(t-\delta(t),t+\delta(t))\}_{t\in[0,1]} of [0,1][0,1]. Take an integer N>1/δ0N>1/\delta_{0} and set σk=1k/N\sigma_{k}=1-k/N for k=0,1,,Nk=0,1,\ldots,N. Note that

(13.7) 𝔐(𝑾σk)𝔐(𝑾σk+1)𝔐(𝑹0)𝔅ρ(𝑾)Int𝔐e2ρ(𝑾σk).\mathfrak{M}({\bm{W}}_{\sigma_{k}})\subset\mathfrak{M}({\bm{W}}_{\sigma_{k+1}})\subset\mathfrak{M}({\bm{R}}_{0})\subset\mathfrak{B}_{\rho}({\bm{W}})\subset\operatorname{Int}\mathfrak{M}_{e^{2\rho}}({\bm{W}}_{\sigma_{k}}).

For 𝑺𝔗g𝔐(𝑹0){\bm{S}}\in\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}) and k=0,,Nk=0,\ldots,N denote by 𝒓k[𝑺]{\bm{r}}_{k}[{\bm{S}}] the Ioffe ray of 𝑾σk{\bm{W}}_{\sigma_{k}} passing through 𝑺\bm{S}, and set τk(𝑺)=dT(𝑺,𝔐(𝑾σk))\tau_{k}({\bm{S}})=d_{T}({\bm{S}},\mathfrak{M}({\bm{W}}_{\sigma_{k}})). Thus 𝒓k[𝑺](0)=𝑯[𝑾σk](𝑺){\bm{r}}_{k}[{\bm{S}}](0)={\bm{H}}[{\bm{W}}_{\sigma_{k}}]({\bm{S}}) and 𝒓k[𝑺](τk(𝑺))=𝑺{\bm{r}}_{k}[{\bm{S}}](\tau_{k}({\bm{S}}))={\bm{S}}. We then claim that

(13.8) 𝒓k+1[𝑺](τk+1(𝑺),τk+1(𝑺)+ρ)𝔐e6ρ(𝑾σk)={\bm{r}}_{k+1}[{\bm{S}}](\tau_{k+1}({\bm{S}}),\tau_{k+1}({\bm{S}})+\rho)\cap\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}})=\varnothing

for 𝑺𝔐e6ρ(𝑾σk){\bm{S}}\in\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) and k=0,,N1k=0,\ldots,N-1. To show the claim note first that inequality

(13.9) dT(𝑺,𝒓k+1[𝑺](3ρ))<ρd_{T}({\bm{S}},{\bm{r}}_{k+1}[{\bm{S}}](3\rho))<\rho

holds for 𝑺𝔐e6ρ(𝑾σk){\bm{S}}\in\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) due to (13.7). As τk+1(𝑺)<3ρ\tau_{k+1}({\bm{S}})<3\rho and

dT(𝑺,𝒓k+1[𝑺](3ρ))=dT(𝒓k+1[𝑺](τk+1(𝑺)),𝒓k+1[𝑺](3ρ))=3ρτk+1(𝑺),d_{T}({\bm{S}},{\bm{r}}_{k+1}[{\bm{S}}](3\rho))=d_{T}({\bm{r}}_{k+1}[{\bm{S}}](\tau_{k+1}({\bm{S}})),{\bm{r}}_{k+1}[{\bm{S}}](3\rho))=3\rho-\tau_{k+1}({\bm{S}}),

we have τk+1(𝑺)>2ρ\tau_{k+1}({\bm{S}})>2\rho by (13.9). Thus

(13.10) τk+1(𝑺)(2ρ,3ρ)\tau_{k+1}({\bm{S}})\in(2\rho,3\rho)

for 𝑺𝔐e6ρ(𝑾σk){\bm{S}}\in\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}). Since |σkσk+1|<δ0|\sigma_{k}-\sigma_{k+1}|<\delta_{0}, there exists tk[0,1]t_{k}\in[0,1] such that |σjtk|<δ(tk)|\sigma_{j}-t_{k}|<\delta(t_{k}) for j=k,k+1j=k,k+1. Hence

dT(𝒓k[𝑺](0),𝒓k+1[𝑺](0))=dT(𝑯[𝑾σk](𝑺),𝑯[𝑾σk+1](𝑺))<ε0,d_{T}({\bm{r}}_{k}[{\bm{S}}](0),{\bm{r}}_{k+1}[{\bm{S}}](0))=d_{T}({\bm{H}}[{\bm{W}}_{\sigma_{k}}]({\bm{S}}),{\bm{H}}[{\bm{W}}_{\sigma_{k+1}}]({\bm{S}}))<\varepsilon_{0},

which implies that

𝒓k+1[𝑺](τk+1(𝑺),τk+1(𝑺)+ρ)𝔅ρ(𝒓k[𝑺](τk(𝑺)+ρ))=𝔅ρ(𝒓k[𝑺](4ρ)).{\bm{r}}_{k+1}[{\bm{S}}](\tau_{k+1}({\bm{S}}),\tau_{k+1}({\bm{S}})+\rho)\subset\mathfrak{B}_{\rho}({\bm{r}}_{k}[{\bm{S}}](\tau_{k}({\bm{S}})+\rho))=\mathfrak{B}_{\rho}({\bm{r}}_{k}[{\bm{S}}](4\rho)).

Since 𝔅ρ(𝒓k[𝑺](4ρ))\mathfrak{B}_{\rho}({\bm{r}}_{k}[{\bm{S}}](4\rho)) is disjoint from 𝔐e6ρ(𝑾σk)\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) by Theorem 9.8, we obtain (13.8).

Now, we inductively show that for K>1K>1 there is a homeomorphism of 𝔗g\mathfrak{T}_{g} onto 2dg\mathbb{R}^{2d_{g}} that maps 𝔐K(𝑾σk)\mathfrak{M}_{K^{\prime}}({\bm{W}}_{\sigma_{k}}), KKK^{\prime}\geqq K, onto concentric closed balls. This is true for k=0k=0 since 𝔐(𝑾σ0)={𝑾}\mathfrak{M}({\bm{W}}_{\sigma_{0}})=\{{\bm{W}}\}.

To complete the induction arguments suppose that the assertion is true for some kk. Thus there is a homeomorphism of 𝔗g\mathfrak{T}_{g} onto 2dg\mathbb{R}^{2d_{g}} that maps 𝔐e6ρ(𝑾σk)\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) onto a 2dg2d_{g}-dimensional euclidean closed ball B¯\bar{B}. Define a mapping 𝒉:𝔐e6ρ(𝑾σk)𝔐e6ρ(𝑾σk+1){\bm{h}}:\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}})\to\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k+1}}) by 𝒉(𝑺)=𝒓k+1[𝑺](3ρ){\bm{h}}({\bm{S}})={\bm{r}}_{k+1}[{\bm{S}}](3\rho). It is surjective by Theorem 9.4. To show that it is injective suppose that 𝒉(𝑺)=𝒉(𝑺){\bm{h}}({\bm{S}})={\bm{h}}({\bm{S}}^{\prime}) for some 𝑺,𝑺𝔐e6ρ(𝑾σk){\bm{S}},{\bm{S}}^{\prime}\in\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}). By the definition of 𝒉\bm{h} the point 𝑺{\bm{S}}^{\prime} lies on the Ioffe ray of 𝑾σk+1{\bm{W}}_{\sigma_{k+1}} passing through 𝒉(𝑺)=𝒉(𝑺){\bm{h}}({\bm{S}}^{\prime})={\bm{h}}({\bm{S}}), that is, on 𝒓k+1[𝑺]=𝒓k+1[𝑺]{\bm{r}}_{k+1}[{\bm{S}}^{\prime}]={\bm{r}}_{k+1}[{\bm{S}}]. Thus 𝒓k+1[𝑺](τk+1(𝑺))=𝑺𝔐e6ρ(𝑾σk){\bm{r}}_{k+1}[{\bm{S}}](\tau_{k+1}({\bm{S}}^{\prime}))={\bm{S}}^{\prime}\in\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) and hence τk+1(𝑺)(τk+1(𝑺),τk+1(𝑺)+ρ)\tau_{k+1}({\bm{S}}^{\prime})\not\in(\tau_{k+1}({\bm{S}}),\tau_{k+1}({\bm{S}})+\rho) by (13.8). Since τk+1(𝑺),τk+1(𝑺)(2ρ,3ρ)\tau_{k+1}({\bm{S}}),\tau_{k+1}({\bm{S}}^{\prime})\in(2\rho,3\rho) by (13.10), we see that τk+1(𝑺)τk+1(𝑺)\tau_{k+1}({\bm{S}}^{\prime})\leqq\tau_{k+1}({\bm{S}}). Interchanging the roles of 𝑺\bm{S} and 𝑺{\bm{S}}^{\prime} we conclude that τk+1(𝑺)=τk+1(𝑺)\tau_{k+1}({\bm{S}}^{\prime})=\tau_{k+1}({\bm{S}}) and hence that 𝑺=𝒓k+1[𝑺](τk+1(𝑺))=𝒓k+1[𝑺](τk+1(𝑺))=𝑺{\bm{S}}^{\prime}={\bm{r}}_{k+1}[{\bm{S}}^{\prime}](\tau_{k+1}({\bm{S}}^{\prime}))={\bm{r}}_{k+1}[{\bm{S}}](\tau_{k+1}({\bm{S}}))={\bm{S}}, which proves that 𝒉{\bm{h}} is injective.

The bijection 𝒉{\bm{h}} is identical with (𝑯eρe6ρ[𝑾σk+1])1𝑯eρ[𝑾σk+1]({\bm{H}}^{e^{6\rho}}_{e^{\rho}}[{\bm{W}}_{\sigma_{k+1}}])^{-1}\circ{\bm{H}}_{e^{\rho}}[{\bm{W}}_{\sigma_{k+1}}] on 𝔐e6ρ(𝑾σk)\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) and hence continuous by Proposition 13.1 and Corollary 13.2. It is homeomorphic since 𝔐e6ρ(𝑾σk)\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) is compact and 𝔐e6ρ(𝑾σk+1)\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k+1}}) is Hausdorff. Thus 𝔐e6ρ(𝑾σk+1)\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k+1}}) is homeomorphic to the sphere B¯\partial\bar{B}.

We extend 𝒉\bm{h} to a homeomorphism of 𝔗g\mathfrak{T}_{g} onto itself step by step. We begin with extending it to a homeomorphism of 𝔐e6ρ(𝑾σk)Int𝔐e4ρ(𝑾σk)\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}})\setminus\operatorname{Int}\mathfrak{M}_{e^{4\rho}}({\bm{W}}_{\sigma_{k}}) onto 𝔐e6ρ(𝑾σk+1)Int𝔐e6ρ(𝑾σk)\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k+1}})\setminus\operatorname{Int}\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) as follows. Each point 𝑺\bm{S} of 𝔐e6ρ(𝑾σk)Int𝔐e4ρ(𝑾σk)\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}})\setminus\operatorname{Int}\mathfrak{M}_{e^{4\rho}}({\bm{W}}_{\sigma_{k}}) is of the form 𝒓k[𝑺](t){\bm{r}}_{k}[{\bm{S}}^{\prime}](t), where 𝑺𝔐e6ρ(𝑾σk){\bm{S}}^{\prime}\in\partial\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) and t[2ρ,3ρ]t\in[2\rho,3\rho]. Then define 𝒉(𝑺)=𝒓k+1[𝑺](t){\bm{h}}({\bm{S}})={\bm{r}}_{k+1}[{\bm{S}}^{\prime}](t^{\prime}), where

tτk+1(𝑺) 3ρτk+1(𝑺)=t2ρρ;\frac{t^{\prime}-\tau_{k+1}({\bm{S}}^{\prime})}{\,3\rho-\tau_{k+1}({\bm{S}}^{\prime})\,}=\frac{\,t-2\rho\,}{\rho};

note that 𝒉(𝑺){\bm{h}}({\bm{S}}) belongs to 𝔐e6ρ(𝑾σk+1)Int𝔐e6ρ(𝑾σk)\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k+1}})\setminus\operatorname{Int}\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) by (13.8). It is easy to extend 𝒉\bm{h} to a homeomorphism of 𝔐e6ρ(𝑾σk)\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) onto 𝔐e6ρ(𝑾σk+1)\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k+1}}) because 𝔐e4ρ(𝑾σk)\mathfrak{M}_{e^{4\rho}}({\bm{W}}_{\sigma_{k}}) and 𝔐e6ρ(𝑾σk)\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) are homeomorphic to the closed ball B¯\bar{B}.

Finally, for 𝑺𝔗g𝔐e6ρ(𝑾σk){\bm{S}}\in\mathfrak{T}_{g}\setminus\mathfrak{M}_{e^{6\rho}}({\bm{W}}_{\sigma_{k}}) take 𝑺𝔐e3ρ(𝑾σk){\bm{S}}^{\prime}\in\partial\mathfrak{M}_{e^{3\rho}}({\bm{W}}_{\sigma_{k}}) and t(3ρ,+)t\in(3\rho,+\infty) so that 𝑺=𝒓k[𝑺](t){\bm{S}}={\bm{r}}_{k}[{\bm{S}}^{\prime}](t). We then define 𝒉(𝑺)=𝒓k+1[𝑺](t){\bm{h}}({\bm{S}})={\bm{r}}_{k+1}[{\bm{S}}^{\prime}](t) to extend 𝒉\bm{h} to a homeomorphism of 𝔗g\mathfrak{T}_{g} onto itself with 𝒉(𝔐K(𝑾σk))=𝔐K(𝑾σk+1){\bm{h}}(\mathfrak{M}_{K^{\prime}}({\bm{W}}_{\sigma_{k}}))=\mathfrak{M}_{K^{\prime}}({\bm{W}}_{\sigma_{k+1}}) for Ke6ρK^{\prime}\geqq e^{6\rho}.

With the aid of the induction hypothesis we can construct from 𝒉1{\bm{h}}^{-1} a homeomorphism of 𝔗g\mathfrak{T}_{g} onto 2dg\mathbb{R}^{2d_{g}} that maps 𝔐K(𝑾σk+1)\mathfrak{M}_{K^{\prime}}({\bm{W}}_{\sigma_{k+1}}), Ke6ρK^{\prime}\geqq e^{6\rho}, onto concentric closed balls. Corollary 13.2 now guarantees the existence of desired homeomorphisms for 𝔐K(𝑾σk+1)\mathfrak{M}_{K}({\bm{W}}_{\sigma_{k+1}}), K>1K>1. ∎

We conclude this section with the following proposition. Theorem 1.2 follows at once from Propositions 13.3 and 13.7.

Proposition 13.7.

If 𝐑0{\bm{R}}_{0} is a marked finite open Riemann surface of genus gg, then the boundary 𝔐K(𝐑0)\partial\mathfrak{M}_{K}({\bm{R}}_{0}) is a C1C^{1}-submanifold of 𝔗g\mathfrak{T}_{g} for K>1K>1.

Proof.

Let 𝑹\bm{R} be an arbitrary point of 𝔐K(𝑹0)\partial\mathfrak{M}_{K}({\bm{R}}_{0}), and take the Ioffe ray 𝒓{\bm{r}} of 𝑹0{\bm{R}}_{0} passing through 𝑹\bm{R}. Set τ=(logK)/2\tau=(\log K)/2. Thus 𝒓(τ)=𝑹{\bm{r}}(\tau)={\bm{R}}. Note that 𝔅τ(𝒓(2τ))𝔐K(𝑹0)=\mathfrak{B}_{\tau}({\bm{r}}(2\tau))\cap\mathfrak{M}_{K}({\bm{R}}_{0})=\varnothing by (9.3) and that 𝔅τ/2(𝒓(τ/2))𝔐K(𝑹0)\mathfrak{B}_{\tau/2}({\bm{r}}(\tau/2))\subset\mathfrak{M}_{K}({\bm{R}}_{0}) by Proposition 9.1. Also, as 𝔗g\mathfrak{T}_{g} is a straight space in the sense of Busemann, 𝑹\bm{R} is the unique common point of the closed balls 𝔅¯τ(𝒓(2τ))\bar{\mathfrak{B}}_{\tau}({\bm{r}}(2\tau)) and 𝔅¯τ/2(𝒓(τ/2))\bar{\mathfrak{B}}_{\tau/2}({\bm{r}}(\tau/2)).

Take a C1C^{1} coordinate system (𝔘,x)(\mathfrak{U},x) centered at 𝑹\bm{R}. Thus 𝔘\mathfrak{U} is a connected open neighborhood of 𝑹\bm{R}, and xx is a C1C^{1} diffeomorphism of 𝔘\mathfrak{U} onto a domain of 2dg\mathbb{R}^{2d_{g}} with x(𝑹)=0x({\bm{R}})=0. Write x=(x1,,x2dg)=(x1,x)x=(x_{1},\ldots,x_{2d_{g}})=(x_{1},x^{\prime}). We choose the system so that (x1(𝒓(t)),x(𝒓(t)))=(tτ,0)(x_{1}({\bm{r}}(t)),x^{\prime}({\bm{r}}(t)))=(t-\tau,0) and x(𝔘)=(δ,δ)×Ux(\mathfrak{U})=(-\delta,\delta)\times U^{\prime} for some δ>0\delta>0 and some domain UU^{\prime} in 2dg1\mathbb{R}^{2d_{g}-1}. Also, we require that the points 𝑺𝔘{\bm{S}}\in\mathfrak{U} with x1(𝑺)=δx_{1}({\bm{S}})=\delta (resp. x1(𝑺)=δx_{1}({\bm{S}})=-\delta) should be contained in 𝔅τ(𝒓(2τ))\mathfrak{B}_{\tau}({\bm{r}}(2\tau)) (resp. 𝔅τ/2(𝒓(τ/2))\mathfrak{B}_{\tau/2}({\bm{r}}(\tau/2))). Thus for each ξU\xi^{\prime}\in U^{\prime} there is 𝑺ξ𝔐K(𝑹0)𝔘{\bm{S}}_{\xi^{\prime}}\in\partial\mathfrak{M}_{K}({\bm{R}}_{0})\cap\mathfrak{U} such that x(𝑺ξ)=ξx^{\prime}({\bm{S}}_{\xi^{\prime}})=\xi^{\prime}. We show that 𝑺ξ{\bm{S}}_{\xi^{\prime}} is uniquely determined if ξ\xi^{\prime} is sufficiently near 0. Observe that even if 𝑺ξ{\bm{S}}_{\xi^{\prime}} is not unique, it tends to 𝑹\bm{R} as ξ0\xi^{\prime}\to 0 because 𝑺ξ𝔅τ(𝒓(2τ))𝔅τ/2(𝒓(τ/2)){\bm{S}}_{\xi^{\prime}}\not\in\mathfrak{B}_{\tau}({\bm{r}}(2\tau))\cup\mathfrak{B}_{\tau/2}({\bm{r}}(\tau/2)). Since the partial derivatives of dT(,𝒓(2τ))d_{T}(\,\cdot\,,{\bm{r}}(2\tau)) and dT(,𝒓(τ/2))d_{T}(\,\cdot\,,{\bm{r}}(\tau/2)) at 𝑹\bm{R} with respect to x1x_{1} are 1-1 and 11, respectively, we may assume that the partial derivative of dT(,𝑺)d_{T}(\,\cdot\,,{\bm{S}}) with respect to x1x_{1} is negative on 𝔘\mathfrak{U} for any 𝑺{\bm{S}} in a neighborhood 𝔙1\mathfrak{V}_{1} of 𝒓(2τ){\bm{r}}(2\tau) and that the partial derivative of dT(,𝑺)d_{T}(\,\cdot\,,{\bm{S}}) with respect to x1x_{1} is positive on 𝔘\mathfrak{U} for any 𝑺{\bm{S}} in a neighborhood 𝔙2\mathfrak{V}_{2} of 𝒓(τ/2){\bm{r}}(\tau/2). Let 𝒓ξ{\bm{r}}_{\xi^{\prime}} denote the Ioffe ray of 𝑹0{\bm{R}}_{0} passing through 𝑺ξ{\bm{S}}_{\xi^{\prime}}. If ξ\xi^{\prime} is sufficiently near 0 so that 𝒓ξ(2τ)𝔙1{\bm{r}}_{\xi^{\prime}}(2\tau)\in\mathfrak{V}_{1} and 𝒓ξ(τ/2)𝔙2{\bm{r}}_{\xi^{\prime}}(\tau/2)\in\mathfrak{V}_{2}, then dT(𝑺,𝒓ξ(2τ))<dT(𝑺ξ,𝒓ξ(2τ))=τd_{T}({\bm{S}},{\bm{r}}_{\xi^{\prime}}(2\tau))<d_{T}({\bm{S}}_{\xi^{\prime}},{\bm{r}}_{\xi^{\prime}}(2\tau))=\tau for 𝑺𝔘{\bm{S}}\in\mathfrak{U} with x(𝑺)=ξx^{\prime}({\bm{S}})=\xi^{\prime} and x1(𝑺)>x1(𝑺ξ)x_{1}({\bm{S}})>x_{1}({\bm{S}}_{\xi^{\prime}}), and hence 𝑺𝔅τ(𝒓ξ(2τ))𝔗g𝔐K(𝑹0){\bm{S}}\in\mathfrak{B}_{\tau}({\bm{r}}_{\xi^{\prime}}(2\tau))\subset\mathfrak{T}_{g}\setminus\mathfrak{M}_{K}({\bm{R}}_{0}). Also, we have dT(𝑺,𝒓ξ(τ/2))<dT(𝑺ξ,𝒓ξ(τ/2))=τ/2d_{T}({\bm{S}},{\bm{r}}_{\xi^{\prime}}(\tau/2))<d_{T}({\bm{S}}_{\xi^{\prime}},{\bm{r}}_{\xi^{\prime}}(\tau/2))=\tau/2 for 𝑺𝔘{\bm{S}}\in\mathfrak{U} with x(𝑺)=ξx^{\prime}({\bm{S}})=\xi^{\prime} and x1(𝑺)<x1(𝑺ξ)x_{1}({\bm{S}})<x_{1}({\bm{S}}_{\xi^{\prime}}), and hence 𝑺𝔅τ/2(𝒓ξ(τ/2))Int𝔐K(𝑹0){\bm{S}}\in\mathfrak{B}_{\tau/2}({\bm{r}}_{\xi^{\prime}}(\tau/2))\subset\operatorname{Int}\mathfrak{M}_{K}({\bm{R}}_{0}). Consequently, 𝑺ξ{\bm{S}}_{\xi^{\prime}} is uniquely determined for ξ\xi^{\prime} sufficiently near 0. Replacing UU^{\prime} with a smaller one if necessary, we assume that 𝑺ξ{\bm{S}}_{\xi^{\prime}} is uniquely determined for ξU\xi^{\prime}\in U^{\prime}.

We have shown that there is a function f:Uf:U^{\prime}\to\mathbb{R} such that f(ξ)=x1(𝑺ξ)f(\xi^{\prime})=x_{1}({\bm{S}}_{\xi^{\prime}}). Since dT:𝔗g×𝔗g+d_{T}:\mathfrak{T}_{g}\times\mathfrak{T}_{g}\to\mathbb{R}_{+} is of class C1C^{1} off the diagonal by Proposition 13.4, the hypersurfaces x(𝔅τ(𝒓ξ(2τ))𝔘)x(\partial\mathfrak{B}_{\tau}({\bm{r}}_{\xi^{\prime}}(2\tau))\cap\mathfrak{U}) and x(𝔅τ/2(𝒓ξ(τ/2))𝔘)x(\partial\mathfrak{B}_{\tau/2}({\bm{r}}_{\xi^{\prime}}(\tau/2))\cap\mathfrak{U}) have a common tangent hyperplane at x(𝑺ξ)x({\bm{S}}_{\xi^{\prime}}). Since the graph of ff lies between these hypersurfaces, we infer that the hyperplane is also tangent to the graph. Consequently, ff is differentiable at ξ\xi^{\prime}. Since the hypersurfaces together with the common tangent hyperplanes move continuously with ξ\xi^{\prime}, we know that ff is of class C1C^{1}. The proof is complete. ∎

14 Geometric properties of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0})

The final section is devoted to establishing Theorem 1.1. Assuming that 𝑹0{\bm{R}}_{0} is nonanalytically finite, fix a circular filling {(𝑾t,ϵt)}t[0,1]\{({\bm{W}}_{t},{\bm{\epsilon}}_{t})\}_{t\in[0,1]} for 𝑹0{\bm{R}}_{0}. We use the same notations as in Example 11.3. Each WtW_{t} is the interior of a compact bordered Riemann surface of genus gg and includes R˘0\breve{R}_{0} for t(0,1)t\in(0,1). If t0(0,1)t_{0}\in(0,1) is sufficiently small, then every quadratic differential φ\varphi in AL(W0)A_{L}(W_{0}) is extended to a holomorphic quadratic differential on Wt0W_{t_{0}}, denoted again by φ\varphi. Thus we consider AL(W0)A_{L}(W_{0}) as a subset of A(Wt0)A(W_{t_{0}}). Let E(W0)E(W_{0}) be the set of quadratic differentials φ\varphi in AL(W0)A(Wt0)A_{L}(W_{0})\subset A(W_{t_{0}}) with φW0=1\|\varphi\|_{W_{0}}=1.

Lemma 14.1.

There exists l0>0l_{0}>0 such that each φE(W0)\varphi\in E(W_{0}) has a noncritical point pW0p\in\partial W_{0} together with a natural parameter ζ:U\zeta:U\to\mathbb{C} of φ\varphi around pp satisfying [2l0,2l0]×[tl0,l0]ζ(UWt)[-2l_{0},2l_{0}]\times[-tl_{0},l_{0}]\subset\zeta(U\cap W_{t}) for t(0,t0)t\in(0,t_{0}).

Proof.

For φE(W0)\varphi\in E(W_{0}) choose a noncritical point pW0p\in\partial W_{0}. There is a neighborhood VφV_{\varphi} of φ\varphi in E(W0)E(W_{0}) such that |(ψ/φ)|(p)||(\psi/\varphi)|(p)|, ψVφ\psi\in V_{\varphi}, are bounded and bounded away from 0; note that ψ/φ\psi/\varphi is a meromorphic function on Wt0W_{t_{0}}. Then there are a positive number lφl_{\varphi} and a neighborhood UU of pp such that a natural parameter ζψ:U\zeta_{\psi}:U\to\mathbb{C} of ψ\psi around pp satisfies [2lφ,2lφ]×[tlφ,lφ]ζψ(UWt)[-2l_{\varphi},2l_{\varphi}]\times[-tl_{\varphi},l_{\varphi}]\subset\zeta_{\psi}(U\cap W_{t}) for t(0,t0)t\in(0,t_{0}). Since E(W0)E(W_{0}) is compact, the open covering {Vφ}\{V_{\varphi}\} of E(W0)E(W_{0}) includes a finite subcovering {Vφ1,,Vφn}\{V_{\varphi_{1}},\ldots,V_{\varphi_{n}}\}. Then l0:=minjlφjl_{0}:=\min_{j}l_{\varphi_{j}} possesses the required properties. ∎

Lemma 14.2.

Let {tn}\{t_{n}\} be a sequence of positive numbers converging to 0. If a sequence {𝐒n}\{{\bm{S}}_{n}\} in 𝔗g𝔐(𝐑0)\mathfrak{T}_{g}\setminus\mathfrak{M}({\bm{R}}_{0}) converges to 𝐒𝔗g{\bm{S}}\in\mathfrak{T}_{g}, then 𝐇[𝐖tn](𝐒n)𝐇[𝐑0](𝐒){\bm{H}}[{\bm{W}}_{t_{n}}]({\bm{S}}_{n})\to{\bm{H}}[{\bm{R}}_{0}]({\bm{S}}) as nn\to\infty.

Proof.

Define 𝑹n=𝑯[𝑾tn](𝑺n){\bm{R}}_{n}={\bm{H}}[{\bm{W}}_{t_{n}}]({\bm{S}}_{n}), 𝑹n=𝑯[𝑹0](𝑺n){\bm{R}}^{\prime}_{n}={\bm{H}}[{\bm{R}}_{0}]({\bm{S}}_{n}) and 𝑹=𝑯[𝑹0](𝑺){\bm{R}}={\bm{H}}[{\bm{R}}_{0}]({\bm{S}}), and set ρn=dT(𝑺n,𝑹n)=dT(𝑺n,𝔐(𝑾tn))\rho_{n}=d_{T}({\bm{S}}_{n},{\bm{R}}_{n})=d_{T}({\bm{S}}_{n},\mathfrak{M}({\bm{W}}_{t_{n}})), ρn=dT(𝑺n,𝑹n)=dT(𝑺n,𝔐(𝑹0))\rho^{\prime}_{n}=d_{T}({\bm{S}}_{n},{\bm{R}}^{\prime}_{n})=d_{T}({\bm{S}}_{n},\mathfrak{M}({\bm{R}}_{0})) and ρ=dT(𝑺,𝑹)=dT(𝑺,𝔐(𝑹0))\rho=d_{T}({\bm{S}},{\bm{R}})=d_{T}({\bm{S}},\mathfrak{M}({\bm{R}}_{0})). Note that 𝑹n𝔐(𝑾tn)𝔐(𝑹0){\bm{R}}_{n}\in\partial\mathfrak{M}({\bm{W}}_{t_{n}})\subset\mathfrak{M}({\bm{R}}_{0}) and 𝑹n,𝑹𝔐(𝑹0){\bm{R}}^{\prime}_{n},{\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}) and that ρnρ\rho^{\prime}_{n}\to\rho as nn\to\infty.

For any ε>0\varepsilon>0 we have |ρnρ|<ε|\rho^{\prime}_{n}-\rho|<\varepsilon and 𝔐(𝑾tn)𝔐(𝑹0)𝔐e2ε(𝑾tn)\mathfrak{M}({\bm{W}}_{t_{n}})\subset\mathfrak{M}({\bm{R}}_{0})\subset\mathfrak{M}_{e^{2\varepsilon}}({\bm{W}}_{t_{n}}) for sufficiently large nn. Then dT(𝑹n,𝑯[𝑾tn](𝑹n))εd_{T}({\bm{R}}^{\prime}_{n},{\bm{H}}[{\bm{W}}_{t_{n}}]({\bm{R}}^{\prime}_{n}))\leqq\varepsilon and hence

ρε\displaystyle\rho-\varepsilon <ρnρndT(𝑺n,𝑯[𝑾tn](𝑹n))dT(𝑺n,𝑹n)+dT(𝑹n,𝑯[𝑾tn](𝑹n))\displaystyle<\rho^{\prime}_{n}\leqq\rho_{n}\leqq d_{T}({\bm{S}}_{n},{\bm{H}}[{\bm{W}}_{t_{n}}]({\bm{R}}^{\prime}_{n}))\leqq d_{T}({\bm{S}}_{n},{\bm{R}}^{\prime}_{n})+d_{T}({\bm{R}}^{\prime}_{n},{\bm{H}}[{\bm{W}}_{t_{n}}]({\bm{R}}^{\prime}_{n}))
ρn+ε<ρ+2ε.\displaystyle\leqq\rho^{\prime}_{n}+\varepsilon<\rho+2\varepsilon.

This proves that ρnρ\rho_{n}\to\rho as nn\to\infty.

We show that 𝑹\bm{R} is a unique accumulation point of {𝑹n}\{{\bm{R}}_{n}\}. Let {𝑹nk}\{{\bm{R}}_{n_{k}}\} be an arbitrary convergent subsequence of {𝑹n}\{{\bm{R}}_{n}\}. If 𝑹{\bm{R}}^{\prime} denotes its limit, then it belongs to 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). As

dT(𝑺,𝔐(𝑹0))=ρ=limkρnk=limkdT(𝑺nk,𝑹nk)=dT(𝑺,𝑹),d_{T}({\bm{S}},\mathfrak{M}({\bm{R}}_{0}))=\rho=\lim_{k\to\infty}\rho_{n_{k}}=\lim_{k\to\infty}d_{T}({\bm{S}}_{n_{k}},{\bm{R}}_{n_{k}})=d_{T}({\bm{S}},{\bm{R}}^{\prime}),

we know that 𝑹{\bm{R}}^{\prime} is the nearest point of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) to 𝑺\bm{S} and hence 𝑹=𝑯[𝑹0](𝑺)=𝑹{\bm{R}}^{\prime}={\bm{H}}[{\bm{R}}_{0}]({\bm{S}})={\bm{R}}. This completes the proof. ∎

Lemma 14.3.

There exist positive numbers cc and MM such that

maxExt(𝔐(𝑾t))maxExt(𝔐(𝑹0))t(c+Mt)maxExt(𝔐(𝑹0))\frac{\,\max\operatorname{Ext}_{\mathcal{F}}(\mathfrak{M}({\bm{W}}_{t}))-\max\operatorname{Ext}_{\mathcal{F}}(\mathfrak{M}({\bm{R}}_{0}))\,}{t}\leqq(-c+Mt)\max\operatorname{Ext}_{\mathcal{F}}(\mathfrak{M}({\bm{R}}_{0}))

for all 𝒮(Σg)\mathcal{F}\in\mathscr{MS}(\Sigma_{g}) and t(0,t0)t\in(0,t_{0}).

Proof.

Let 𝒮(Σg)\mathcal{F}\in\mathscr{MS}(\Sigma_{g}) with maxExt(𝔐(𝑹0))=1\max\operatorname{Ext}_{\mathcal{F}}(\mathfrak{M}({\bm{R}}_{0}))=1, and take 𝝋=[φ,θ˘0]AL(𝑾0){\bm{\varphi}}=[\varphi,\breve{\theta}_{0}]\in A_{L}({\bm{W}}_{0}) for which 𝑹0(ϵ0𝝋)=\mathcal{H}_{{\bm{R}}_{0}}({\bm{\epsilon}}_{0}^{*}{\bm{\varphi}})=\mathcal{F}. Note that φE(W0)\varphi\in E(W_{0}). Take a positive number l0l_{0} as in Lemma 14.1; there is a noncritical point pW0p\in\partial W_{0} of φ\varphi together with a natural parameter ζ:U\zeta:U\to\mathbb{C}, ζ=ξ+iη\zeta=\xi+i\eta, of φ\varphi around pp such that the closed rectangle EtE_{t} defined by |ξ|2l0|\xi|\leqq 2l_{0} and tl0ηl0-tl_{0}\leqq\eta\leqq l_{0} is included in UWtU\cap W_{t} for t(0,t0)t\in(0,t_{0}). Note that points in UU with η>0\eta>0 belong to W0W_{0}. In the following we identify UU with ζ(U)\zeta(U) in the obvious manner.

We divide EtE_{t} into three parts Et(1)E_{t}^{(1)}, Et(2)E_{t}^{(2)} and Et(3)E_{t}^{(3)} as follows:

Et(1)\displaystyle E_{t}^{(1)} =[l0,l0]×[tl0,l0],\displaystyle=[-l_{0},l_{0}]\times[-tl_{0},l_{0}],
Et(2)\displaystyle E_{t}^{(2)} ={ζEtl0<|ξ|2l0 and yt(ξ)ηl0}, and\displaystyle=\{\zeta\in E_{t}\mid l_{0}<|\xi|\leqq 2l_{0}\text{ and }y_{t}(\xi)\leqq\eta\leqq l_{0}\},\text{ and}
Et(3)\displaystyle E_{t}^{(3)} =Et(Et(1)Et(2)),\displaystyle=E_{t}\setminus(E_{t}^{(1)}\cup E_{t}^{(2)}),

where yt(ξ)=t(|ξ|2l0)y_{t}(\xi)=t(|\xi|-2l_{0}). Define a function vtv_{t} on EtE_{t} so that for y[0,l0]y\in[0,l_{0}] the function vtv_{t} assumes the value yy on the polygonal arc obtained by joining 2l0+iy-2l_{0}+iy, l0+i{(1+t)ytl0}-l_{0}+i\{(1+t)y-tl_{0}\}, l0+i{(1+t)ytl0}l_{0}+i\{(1+t)y-tl_{0}\} and 2l0+iy2l_{0}+iy successively. In other words, we define the function vtv_{t} by

vt(ζ)={η+tl0 1+tif ζEt(1),l0l0(ηl0)yt(ξ)l0if ζEt(2),0if ζEt(3).v_{t}(\zeta)=\begin{cases}\dfrac{\,\eta+tl_{0}\,}{\,1+t\,}&\text{if $\zeta\in E_{t}^{(1)}$},\\[8.61108pt] l_{0}-\dfrac{l_{0}(\eta-l_{0})}{\,y_{t}(\xi)-l_{0}\,}&\text{if $\zeta\in E_{t}^{(2)}$},\\[8.61108pt] 0&\text{if $\zeta\in E_{t}^{(3)}$}.\end{cases}

For t(0,t0)t\in(0,t_{0}) define a quadratic differential φt\varphi_{t} on WtW_{t} by

φt={{vtη(ζ)+ivtξ(ζ)}2dζ2on Et,φon W¯0Et,0on Wt(W¯0Et).\varphi_{t}=\begin{cases}\biggl{\{}\dfrac{\partial v_{t}}{\partial\eta}(\zeta)+i\dfrac{\partial v_{t}}{\partial\xi}(\zeta)\biggr{\}}^{2}\,d\zeta^{2}&\text{on $E_{t}$},\\ \varphi&\text{on $\bar{W}_{0}\setminus E_{t}$},\\ 0&\text{on $W_{t}\setminus(\bar{W}_{0}\cup E_{t})$}.\end{cases}

Since

φtEt\displaystyle\|\varphi_{t}\|_{E_{t}} =Et{vtξ(ζ)2+vtη(ζ)2}𝑑ξ𝑑η=2l02 1+t+ 2l02(t2+3)3tlog(1+t)\displaystyle=\iint_{E_{t}}\biggl{\{}\frac{\partial v_{t}}{\partial\xi}(\zeta)^{2}+\frac{\partial v_{t}}{\partial\eta}(\zeta)^{2}\biggr{\}}\,d\xi\,d\eta=\frac{2l_{0}^{2}}{\,1+t\,}+\frac{\,2l_{0}^{2}(t^{2}+3)\,}{3t}\log(1+t)
=φW0Et3l02t+O(t2)\displaystyle=\|\varphi\|_{W_{0}\cap E_{t}}-3l_{0}^{2}t+O(t^{2})

as t0t\to 0, there is a positive number MM such that

φtWt𝝋𝑾03l02t+Mt2\|\varphi_{t}\|_{W_{t}}\leqq\|{\bm{\varphi}}\|_{{\bm{W}}_{0}}-3l_{0}^{2}t+Mt^{2}

for all t(0,t0)t\in(0,t_{0}).

Let 𝑹𝔐(𝑹0){\bm{R}}\in\mathfrak{M}_{\mathcal{F}}({\bm{R}}_{0}). It is induced by a 𝝋\bm{\varphi}-regular self-welding continuation of 𝑾0{\bm{W}}_{0} by Theorem 1.6 (i). Let 𝝍A(𝑹){\bm{\psi}}\in A({\bm{R}}) be the co-welder of 𝝋\bm{\varphi}. Thus =𝑹(𝝍)\mathcal{F}=\mathcal{H}_{\bm{R}}({\bm{\psi}}) and 𝝋𝑾0=𝝍𝑹=Ext(𝑹)=1\|{\bm{\varphi}}\|_{{\bm{W}}_{0}}=\|{\bm{\psi}}\|_{\bm{R}}=\operatorname{Ext}_{\mathcal{F}}({\bm{R}})=1. For t(0,t0)t\in(0,t_{0}) let 𝑹~t=[R~t,θ~t]𝔐(𝑾t)\tilde{\bm{R}}_{t}=[\tilde{R}_{t},\tilde{\theta}_{t}]\in\mathfrak{M}_{\mathcal{F}}({\bm{W}}_{t}), and set 𝝍~t=𝑸𝑹~t()\tilde{\bm{\psi}}_{t}={\bm{Q}}_{\tilde{\bm{R}}_{t}}(\mathcal{F}). If 𝜾~t=[ι~t,θ˘0,θ~t]CEmbhc(𝑾t,𝑹~t)\tilde{\bm{\iota}}_{t}=[\tilde{\iota}_{t},\breve{\theta}_{0},\tilde{\theta}_{t}]\in\operatorname{CEmb}_{\mathrm{hc}}({\bm{W}}_{t},\tilde{\bm{R}}_{t}), then φt\varphi_{t} induces a quadratic differential on ι~t(Wt)\tilde{\iota}_{t}(W_{t}). Let φ~t\tilde{\varphi}_{t} denote its zero-extension to R~t\tilde{R}_{t}. Similarly, φ\varphi induces a holomorphic differential on ι~t(W0)\tilde{\iota}_{t}(W_{0}). Denote by φ~\tilde{\varphi} its zero-extension to R~t\tilde{R}_{t}. Since |Imφt|=|dvt||\operatorname{Im}\sqrt{\varphi_{t}\,}|=|dv_{t}| on EtE_{t}, for γ𝒮(Σg)\gamma\in\mathscr{S}(\Sigma_{g}) we have Hφ~t(c)=Hφ~(c)H_{\tilde{\varphi}_{t}}(c)=H_{\tilde{\varphi}}(c) for any piecewise analytic simple loop cc on R~t\tilde{R}_{t} with θ~tcγ\tilde{\theta}_{t}^{*}c\in\gamma. Hence 𝑹~t(𝝋~t)(γ)=𝑹~t(𝝋~)(γ)\mathcal{H}^{\prime}_{\tilde{\bm{R}}_{t}}(\tilde{\bm{\varphi}}_{t})(\gamma)=\mathcal{H}^{\prime}_{\tilde{\bm{R}}_{t}}(\tilde{\bm{\varphi}})(\gamma), where 𝝋~t=[φ~t,θ~t]\tilde{\bm{\varphi}}_{t}=[\tilde{\varphi}_{t},\tilde{\theta}_{t}] and 𝝋~=[φ~,θ~t]\tilde{\bm{\varphi}}=[\tilde{\varphi},\tilde{\theta}_{t}]. Since

𝑹~t(𝝍~t)(γ)=(γ)=𝑹(𝝍)(γ)𝑹~t(𝝋~)(γ)=𝑹~t(𝝋~t)(γ),\mathcal{H}_{\tilde{\bm{R}}_{t}}(\tilde{\bm{\psi}}_{t})(\gamma)=\mathcal{F}(\gamma)=\mathcal{H}_{\bm{R}}({\bm{\psi}})(\gamma)\leqq\mathcal{H}^{\prime}_{\tilde{\bm{R}}_{t}}(\tilde{\bm{\varphi}})(\gamma)=\mathcal{H}^{\prime}_{\tilde{\bm{R}}_{t}}(\tilde{\bm{\varphi}}_{t})(\gamma),

it follows from Proposition 6.4 that

Ext(𝑹~t)=𝝍~t𝑹~t𝝋~t𝑹~t=φtWt𝝋𝑾03l02t+Mt2=13l02t+Mt2.\operatorname{Ext}_{\mathcal{F}}(\tilde{\bm{R}}_{t})=\|\tilde{\bm{\psi}}_{t}\|_{\tilde{\bm{R}}_{t}}\leqq\|\tilde{\bm{\varphi}}_{t}\|_{\tilde{\bm{R}}_{t}}=\|\varphi_{t}\|_{W_{t}}\leqq\|{\bm{\varphi}}\|_{{\bm{W}}_{0}}-3l_{0}^{2}t+Mt^{2}=1-3l_{0}^{2}t+Mt^{2}.

Since 𝑹~t\tilde{\bm{R}}_{t} is a maximal point for \mathcal{F} on 𝔐(𝑾t)\mathfrak{M}({\bm{W}}_{t}), we obtain

maxExt(𝔐(𝑾t))13l02t+Mt2\max\operatorname{Ext}_{\mathcal{F}}(\mathfrak{M}({\bm{W}}_{t}))\leqq 1-3l_{0}^{2}t+Mt^{2}

for all t(0,t0)t\in(0,t_{0}), provided that maxExt(𝔐(𝑹0))=1\max\operatorname{Ext}_{\mathcal{F}}(\mathfrak{M}({\bm{R}}_{0}))=1. This proves the lemma. ∎

In general, let 𝔎\mathfrak{K} be a compact set of 𝔗g\mathfrak{T}_{g}. For 𝑺𝔎{\bm{S}}\in\mathfrak{K} denote by (Σg;𝔎,𝑺)\mathscr{MF}(\Sigma_{g};\mathfrak{K},{\bm{S}}) the set of measured foliations \mathcal{F} on Σg\Sigma_{g} such that 𝑺\bm{S} is a maximal point for \mathcal{F} on 𝔎\mathfrak{K}. For nonempty 𝔏𝔎\mathfrak{L}\subset\mathfrak{K} set (Σg;𝔎,𝔏)=𝑺𝔏(Σg;𝔎,𝑺)\mathscr{MF}(\Sigma_{g};\mathfrak{K},\mathfrak{L})=\bigcup_{{\bm{S}}\in\mathfrak{L}}\mathscr{MF}(\Sigma_{g};\mathfrak{K},{\bm{S}}).

Corollary 14.4.

There exist positive numbers cc and MM such that for all t(0,t0)t\in(0,t_{0}) and 𝐑𝔐(𝐑0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}) the inequality

Ext(𝑯[𝑾t](𝑹))Ext(𝑹)dT(𝑯[𝑾t](𝑹),𝑹)(c+Mt)Ext(𝑹)\frac{\,\operatorname{Ext}_{\mathcal{F}}({\bm{H}}[{\bm{W}}_{t}]({\bm{R}}))-\operatorname{Ext}_{\mathcal{F}}({\bm{R}})\,}{d_{T}({\bm{H}}[{\bm{W}}_{t}]({\bm{R}}),{\bm{R}})}\leqq(-c+Mt)\operatorname{Ext}_{\mathcal{F}}({\bm{R}})

holds for all (Σg;𝔐(𝐑0),𝐑)\mathcal{F}\in\mathscr{MF}(\Sigma_{g};\mathfrak{M}({\bm{R}}_{0}),{\bm{R}}).

Proof.

Note that 𝑯[𝑾t](𝑹)𝑹{\bm{H}}[{\bm{W}}_{t}]({\bm{R}})\neq{\bm{R}} by Proposition 11.7. Since there exists a homotopically consistent (1(log(1t))/logr0)(1-(\log(1-t))/\log r_{0})-quasiconformal homeomorphism of 𝑾t{\bm{W}}_{t} onto 𝑾0{\bm{W}}_{0} (for the definition of r0r_{0} see Example 11.3), Lemma 11.1 and Proposition 9.1 imply that there is c1>0c_{1}>0 such that

dT(𝑯[𝑾t](𝑹),𝑹)=dT(𝑹,𝔐(𝑾t))1 2log{1log(1t)logr0}<c1td_{T}({\bm{H}}[{\bm{W}}_{t}]({\bm{R}}),{\bm{R}})=d_{T}({\bm{R}},\mathfrak{M}({\bm{W}}_{t}))\leqq\frac{1}{\,2\,}\log\biggl{\{}1-\frac{\,\log(1-t)\,}{\log r_{0}}\biggr{\}}<c_{1}t

for all t(0,t0)t\in(0,t_{0}) and 𝑹𝔐(𝑾0)=𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{W}}_{0})=\partial\mathfrak{M}({\bm{R}}_{0}).

Let 𝑹𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}). If \mathcal{F} belongs to (Σg;𝔐(𝑹0),𝑹)\mathscr{MF}(\Sigma_{g};\mathfrak{M}({\bm{R}}_{0}),{\bm{R}}), then

Ext(𝑯[𝑾t](𝑹))Ext(𝑹)maxExt(𝔐(𝑾t))maxExt(𝔐(𝑹0))<0,\operatorname{Ext}_{\mathcal{F}}({\bm{H}}[{\bm{W}}_{t}]({\bm{R}}))-\operatorname{Ext}_{\mathcal{F}}({\bm{R}})\leqq\max\operatorname{Ext}_{\mathcal{F}}(\mathfrak{M}({\bm{W}}_{t}))-\max\operatorname{Ext}_{\mathcal{F}}(\mathfrak{M}({\bm{R}}_{0}))<0,

where the last inequality follows from the fact that 𝔐(𝑾t)Int𝔐(𝑹0)\mathfrak{M}({\bm{W}}_{t})\subset\operatorname{Int}\mathfrak{M}({\bm{R}}_{0}). Therefore, the corollary is an immediate consequence of Lemma 14.3. ∎

In general, let 𝑹=[R,θ]𝔗g{\bm{R}}=[R,\theta]\in\mathfrak{T}_{g}. For 𝝋=[φ,θ],𝝍=[ψ,θ]A(𝑹){𝟎}{\bm{\varphi}}=[\varphi,\theta],{\bm{\psi}}=[\psi,\theta]\in A({\bm{R}})\setminus\{{\bm{0}}\} define

D𝑹[𝝋](𝝍)=Rφ|ψ|ψ=Rφ(z)|ψ(z)|ψ(z)dzdz¯2i,D_{\bm{R}}[{\bm{\varphi}}]({\bm{\psi}})=\iint_{R}\frac{\,\varphi|\psi|\,}{\psi}=\iint_{R}\frac{\,\varphi(z)|\psi(z)|\,}{\psi(z)}\,\frac{\,dz\wedge d\bar{z}\,}{-2i},

where φ=φ(z)dz2\varphi=\varphi(z)\,dz^{2} and ψ=ψ(z)dz2\psi=\psi(z)\,dz^{2}. We need the following variational formula due to Gardiner, which implies that the extremal length function Ext\operatorname{Ext}_{\mathcal{F}} is continuously differentiable on 𝔗g\mathfrak{T}_{g}. For the proof see also Gardiner-Lakic [18, Theorem 12.5].

Proposition 14.5 (Gardiner [16, Theorem 8]).

Let 𝐑𝔗g{\bm{R}}\in\mathfrak{T}_{g} and 𝛗A(𝐑){𝟎}{\bm{\varphi}}\in A({\bm{R}})\setminus\{{\bm{0}}\} and set =𝐑(𝛗)\mathcal{F}=\mathcal{H}_{\bm{R}}({\bm{\varphi}}). Then for 𝛙A(𝐑){𝟎}{\bm{\psi}}\in A({\bm{R}})\setminus\{{\bm{0}}\}

logExt(𝒓𝑹[𝝍](t))=logExt(𝑹)+2t𝝋𝑹ReD𝑹[𝝋](𝝍)+o(t)\log\operatorname{Ext}_{\mathcal{F}}({\bm{r}}_{\bm{R}}[{\bm{\psi}}](t))=\log\operatorname{Ext}_{\mathcal{F}}(\bm{R})+\frac{2t}{\,\|{\bm{\varphi}}\|_{\bm{R}}\,}\operatorname{Re}D_{\bm{R}}[{\bm{\varphi}}]({\bm{\psi}})+o(t)

as t+0t\to+0.

In fact, the differential of logExt\log\operatorname{Ext}_{\mathcal{F}} at 𝑹\bm{R} is the linear functional

μ2𝝋𝑹ReRμφ\mu\mapsto\frac{2}{\,\|{\bm{\varphi}}\|_{\bm{R}}\,}\operatorname{Re}\int_{R}\mu\varphi

on the space of bounded measurable (1,1)(-1,1)-forms μ\mu on RR, where 𝑹=[R,θ]{\bm{R}}=[R,\theta] and 𝝋=[φ,θ]{\bm{\varphi}}=[\varphi,\theta]. The inequality (2/𝝋𝑹)ReRμφ2μ(2/\|{\bm{\varphi}}\|_{\bm{R}})\operatorname{Re}\int_{R}\mu\varphi\leqq 2\|\mu\|_{\infty} holds and the sign of equality occurs if μ=|φ|/φ\mu=|\varphi|/\varphi.

For 𝑹𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}) and t(0,1)t\in(0,1) we denote by 𝒓(t)[𝑹]{\bm{r}}^{(t)}[{\bm{R}}] the Ioffe ray of 𝑾t{\bm{W}}_{t} passing through 𝑹\bm{R}. Thus its initial point 𝒓(t)[𝑹](0){\bm{r}}^{(t)}[{\bm{R}}](0) coincides with 𝑯[𝑾t](𝑹){\bm{H}}[{\bm{W}}_{t}]({\bm{R}}). Also, set τ(t)(𝑹)=dT(𝑹,𝔐(𝑾t))\tau^{(t)}({\bm{R}})=d_{T}({\bm{R}},\mathfrak{M}({\bm{W}}_{t})) so that 𝒓(t)[𝑹](τ(t)(𝑹))=𝑹{\bm{r}}^{(t)}[{\bm{R}}](\tau^{(t)}({\bm{R}}))={\bm{R}}.

Lemma 14.6.

For each 𝐑𝔐(𝐑0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}) there exist a neighborhood 𝔘\mathfrak{U} of 𝐑\bm{R} and positive numbers τ0\tau_{0} and δ\delta such that for 𝐑𝔐(𝐑0)𝔘{\bm{R}}^{\prime}\in\partial\mathfrak{M}({\bm{R}}_{0})\cap\mathfrak{U}, t(0,τ0)t\in(0,\tau_{0}) and (Σg;𝔐(𝐑0),𝐑)\mathcal{F}\in\mathscr{MF}(\Sigma_{g};\mathfrak{M}({\bm{R}}_{0}),{\bm{R}}^{\prime}) the inequality

Ext(𝒓(t)[𝑹](s))>Ext(𝑹)\operatorname{Ext}_{\mathcal{F}}({\bm{r}}^{(t)}[{\bm{R}}^{\prime}](s))>\operatorname{Ext}_{\mathcal{F}}({\bm{R}}^{\prime})

holds whenever τ(t)(𝐑)<s<τ(t)(𝐑)+δ\tau^{(t)}({\bm{R}}^{\prime})<s<\tau^{(t)}({\bm{R}}^{\prime})+\delta.

Proof.

Let 𝑹𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}). If there were no required 𝔘\mathfrak{U}, τ0\tau_{0} or δ\delta, then there would exist a sequence {𝔘n}\{\mathfrak{U}_{n}\} of neighborhoods of 𝑹\bm{R} shrinking to {𝑹}\{{\bm{R}}\} and sequences {τn}\{\tau_{n}\} and {δn}\{\delta_{n}\} of positive numbers converging to 0 such that

(14.1) Extn(𝒓(τn)[𝑹n](sn))Extn(𝑹n)\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{r}}^{(\tau_{n})}[{\bm{R}}_{n}](s_{n}))\leqq\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{R}}_{n})

for some 𝑹n𝔐(𝑹0)𝔘n{\bm{R}}_{n}\in\partial\mathfrak{M}({\bm{R}}_{0})\cap\mathfrak{U}_{n}, some sns_{n} with

τ(τn)(𝑹n)<sn<τ(τn)(𝑹n)+δn<1\tau^{(\tau_{n})}({\bm{R}}_{n})<s_{n}<\tau^{(\tau_{n})}({\bm{R}}_{n})+\delta_{n}<1

and some measured foliation n\mathcal{F}_{n} in (Σg;𝔐(𝑹0),𝑹n)\mathscr{MF}(\Sigma_{g};\mathfrak{M}({\bm{R}}_{0}),{\bm{R}}_{n}) with Extn(𝑹n)=1\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{R}}_{n})=1.

Set 𝑺n=𝒓(τn)[𝑹n](sn){\bm{S}}_{n}={\bm{r}}^{(\tau_{n})}[{\bm{R}}_{n}](s_{n}) and 𝑺n=𝒓(τn)[𝑹n](1){\bm{S}}^{\prime}_{n}={\bm{r}}^{(\tau_{n})}[{\bm{R}}_{n}](1). Taking a subsequence if necessary, we may suppose that {𝑺n}\{{\bm{S}}^{\prime}_{n}\} and {n}\{\mathcal{F}_{n}\} converge to 𝑺𝔗g{\bm{S}}^{\prime}\in\mathfrak{T}_{g} and (Σg)\mathcal{F}\in\mathscr{MF}(\Sigma_{g}), respectively. Note that 𝑹n𝑹{\bm{R}}_{n}\to{\bm{R}} as nn\to\infty. Since Extn(𝑹)Extn(𝑹n)=1\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{R}}^{\prime})\leqq\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{R}}_{n})=1 for all 𝑹𝔐(𝑹0){\bm{R}}^{\prime}\in\mathfrak{M}({\bm{R}}_{0}), by letting nn\to\infty we obtain Ext(𝑹)Ext(𝑹)=1\operatorname{Ext}_{\mathcal{F}}({\bm{R}}^{\prime})\leqq\operatorname{Ext}_{\mathcal{F}}({\bm{R}})=1. In particular, 𝑹\bm{R} is a maximal point for \mathcal{F} on 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}).

As 𝑹n{\bm{R}}_{n} lies on the Teichmüller geodesic segment connecting 𝑺n=𝒓(τn)[𝑹n](1){\bm{S}}^{\prime}_{n}={\bm{r}}^{(\tau_{n})}[{\bm{R}}_{n}](1) and 𝒓(τn)[𝑹n](0){\bm{r}}^{(\tau_{n})}[{\bm{R}}_{n}](0), we have

dT(𝑺n,𝒓(τn)[𝑹n](0))dT(𝑺n,𝑹n).d_{T}({\bm{S}}^{\prime}_{n},{\bm{r}}^{(\tau_{n})}[{\bm{R}}_{n}](0))\geqq d_{T}({\bm{S}}^{\prime}_{n},{\bm{R}}_{n}).

Lemma 14.2 yields that 𝒓(τn)[𝑹n](0)=𝑯[𝑾τn](𝑺n)𝑯[𝑹0](𝑺){\bm{r}}^{(\tau_{n})}[{\bm{R}}_{n}](0)={\bm{H}}[{\bm{W}}_{\tau_{n}}]({\bm{S}}^{\prime}_{n})\to{\bm{H}}[{\bm{R}}_{0}]({\bm{S}}^{\prime}) as nn\to\infty. Taking limits of the both sides of the above inequality, we obtain

dT(𝑺,𝔐(𝑹0))=dT(𝑺,𝑯[𝑹0](𝑺))dT(𝑺,𝑹),d_{T}({\bm{S}}^{\prime},\mathfrak{M}({\bm{R}}_{0}))=d_{T}({\bm{S}}^{\prime},{\bm{H}}[{\bm{R}}_{0}]({\bm{S}}^{\prime}))\geqq d_{T}({\bm{S}}^{\prime},{\bm{R}}),

which means that 𝑹\bm{R} is a point of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) nearest to 𝑺{\bm{S}}^{\prime}. Therefore, 𝑯[𝑹0](𝑺){\bm{H}}[{\bm{R}}_{0}]({\bm{S}}^{\prime}) is identical with 𝑹\bm{R} by Corollary 9.11 and hence 𝒓:=𝒓𝑹[𝑺]{\bm{r}}:={\bm{r}}_{\bm{R}}[{\bm{S}}^{\prime}] is an Ioffe ray of 𝑹0{\bm{R}}_{0}.

If 𝝍=𝑸𝑹[𝑺]{\bm{\psi}}={\bm{Q}}_{\bm{R}}[{\bm{S}}^{\prime}] and 𝝍n=𝑸𝑹n[𝑺n]{\bm{\psi}}_{n}={\bm{Q}}_{{\bm{R}}_{n}}[{\bm{S}}^{\prime}_{n}], then if follows from Proposition 13.4 that {𝝍n}\{{\bm{\psi}}_{n}\} converges to 𝝍\bm{\psi} in the complex vector bundle of holomorphic quadratic differentials over 𝔗g\mathfrak{T}_{g}. Set 𝝋n=𝑸𝑹n(n){\bm{\varphi}}_{n}={\bm{Q}}_{{\bm{R}}_{n}}(\mathcal{F}_{n}) and σn=dT(𝑹n,𝑺n)\sigma_{n}=d_{T}({\bm{R}}_{n},{\bm{S}}_{n}). Since 𝑹n=𝒓𝑹n[𝝍n](0){\bm{R}}_{n}={\bm{r}}_{{\bm{R}}_{n}}[{\bm{\psi}}_{n}](0) and 𝑺n=𝒓𝑹n[𝝍n](σn){\bm{S}}_{n}={\bm{r}}_{{\bm{R}}_{n}}[{\bm{\psi}}_{n}](\sigma_{n}), Proposition 14.5 together with (14.1) implies

ReD𝑹n[𝝋n](𝝍n)=logExtn(𝑺n)logExtn(𝑹n)2σn+εnεn\operatorname{Re}D_{{\bm{R}}_{n}}[{\bm{\varphi}}_{n}]({\bm{\psi}}_{n})=\frac{\,\log\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{S}}_{n})-\log\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{R}}_{n})\,}{2\sigma_{n}}+\varepsilon_{n}\leqq\varepsilon_{n}

with εn0\varepsilon_{n}\to 0 as nn\to\infty and hence

(14.2) ReD𝑹[𝝋](𝝍)0,\operatorname{Re}D_{\bm{R}}[{\bm{\varphi}}]({\bm{\psi}})\leqq 0,

where 𝝋=𝑸𝑹(){\bm{\varphi}}={\bm{Q}}_{\bm{R}}(\mathcal{F}).

On the other hand, another application of Proposition 14.5 gives

ReD𝑹n[𝝋n](𝝍n)=logExtn(𝒓(τn)[𝑹n](0))logExtn(𝑹n)2τ(τn)(𝑹n)+εn\operatorname{Re}D_{{\bm{R}}_{n}}[{\bm{\varphi}}_{n}](-{\bm{\psi}}_{n})=\frac{\,\log\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{r}}^{(\tau_{n})}[{\bm{R}}_{n}](0))-\log\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{R}}_{n})\,}{2\tau^{(\tau_{n})}({\bm{R}}_{n})}+\varepsilon^{\prime}_{n}

with εn0\varepsilon^{\prime}_{n}\to 0 as nn\to\infty. Since Extn(𝑹n)=1\operatorname{Ext}_{\mathcal{F}_{n}}({\bm{R}}_{n})=1 and 𝒓(τn)[𝑹n](0)=𝑯[𝑾τ(n)](𝑹n){\bm{r}}^{(\tau_{n})}[{\bm{R}}_{n}](0)={\bm{H}}[{\bm{W}}_{\tau^{(n)}}]({\bm{R}}_{n}), it follows from Corollary 14.4 that

ReD𝑹n[𝝋n](𝝍n)1 2τ(τn)(𝑹n)log(1+(c+Mτn)τ(τn)(𝑹n))+εn\operatorname{Re}D_{{\bm{R}}_{n}}[{\bm{\varphi}}_{n}](-{\bm{\psi}}_{n})\leqq\frac{1}{\,2\tau^{(\tau_{n})}({\bm{R}}_{n})\,}\log(1+(-c+M\tau_{n})\tau^{(\tau_{n})}({\bm{R}}_{n}))+\varepsilon^{\prime}_{n}

for some positive constants cc and MM. Letting nn\to\infty, we obtain

ReD𝑹[𝝋](𝝍)c 2>0,\operatorname{Re}D_{{\bm{R}}}[{\bm{\varphi}}]({\bm{\psi}})\geqq\frac{c}{\,2\,}>0,

which contradicts (14.2). ∎

Proposition 14.7.

If 𝐑0{\bm{R}}_{0} is nonanalytically finite, then there is a homeomorphism of 𝔗g\mathfrak{T}_{g} onto 2dg\mathbb{R}^{2d_{g}} that maps 𝔐(𝐑0)\mathfrak{M}({\bm{R}}_{0}) onto a closed ball.

Proof.

Since 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) is compact, an application of Lemma 14.6 gives positive numbers τ0\tau_{0} and δ\delta such that for 𝑹𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}), t(0,τ0)t\in(0,\tau_{0}) and (Σg;𝔐(𝑹0),𝑹)\mathcal{F}\in\mathscr{MF}(\Sigma_{g};\mathfrak{M}({\bm{R}}_{0}),{\bm{R}}) the inequality

(14.3) Ext(𝒓(t)[𝑹](s))>Ext(𝑹)\operatorname{Ext}_{\mathcal{F}}({\bm{r}}^{(t)}[{\bm{R}}](s))>\operatorname{Ext}_{\mathcal{F}}({\bm{R}})

holds whenever τ(t)(𝑹)<s<τ(t)(𝑹)+δ\tau^{(t)}({\bm{R}})<s<\tau^{(t)}({\bm{R}})+\delta. Fix tt with 0<t<min{τ0,δ}0<t<\min\{\tau_{0},\delta\} so that 𝔐(𝑹0)𝔐e2δ(𝑾t)\mathfrak{M}({\bm{R}}_{0})\subset\mathfrak{M}_{e^{2\delta}}({\bm{W}}_{t}). By Proposition 11.7 there is σ>0\sigma>0 such that 𝔐e2σ(𝑾t)Int𝔐(𝑹0)\mathfrak{M}_{e^{2\sigma}}({\bm{W}}_{t})\subset\operatorname{Int}\mathfrak{M}({\bm{R}}_{0}). Inequality (14.3) implies that the Teichmüller geodesic arc 𝒓(t)[𝑹](τ(t)(𝑹),τ(t)(𝑹)+δ){\bm{r}}^{(t)}[{\bm{R}}](\tau^{(t)}({\bm{R}}),\tau^{(t)}({\bm{R}})+\delta) lies outside of 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) since 𝑹\bm{R} is a maximal point for \mathcal{F} on 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). Note that τ(t)(𝑹)δ\tau^{(t)}({\bm{R}})\leqq\delta as 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is included in 𝔐e2δ(𝑾t)\mathfrak{M}_{e^{2\delta}}({\bm{W}}_{t}). We thus deduce that the correspondence

𝔐(𝑹0)𝑹𝒓(t)[𝑹](σ)𝔐e2σ(𝑾t)\partial\mathfrak{M}({\bm{R}}_{0})\ni{\bm{R}}\mapsto{\bm{r}}^{(t)}[{\bm{R}}](\sigma)\in\partial\mathfrak{M}_{e^{2\sigma}}({\bm{W}}_{t})

is a continuous bijection, and hence is a homeomorphism as 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) is compact and 𝔐e2σ(𝑾t)\partial\mathfrak{M}_{e^{2\sigma}}({\bm{W}}_{t}) is Hausdorff. Observe that Int𝔐e2δ(𝑾t)𝔐e2σ(𝑾t)\operatorname{Int}\mathfrak{M}_{e^{2\delta}}({\bm{W}}_{t})\setminus\mathfrak{M}_{e^{2\sigma}}({\bm{W}}_{t}) is the disjoint union of Teichmüller geodesic arcs 𝒓(t)[𝑹](σ,δ){\bm{r}}^{(t)}[{\bm{R}}](\sigma,\delta), 𝑹𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}), and that 𝒓(t)[𝑹](σ,δ)𝔐(𝑹0)={𝑹}{\bm{r}}^{(t)}[{\bm{R}}](\sigma,\delta)\cap\partial\mathfrak{M}({\bm{R}}_{0})=\{{\bm{R}}\}. Since 𝑹τ(t)(𝑹){\bm{R}}\mapsto\tau^{(t)}({\bm{R}}) is continuous, there is a homeomorphism 𝒉\bm{h} of Int𝔐e2δ(𝑾t)𝔐e2σ(𝑾t)\operatorname{Int}\mathfrak{M}_{e^{2\delta}}({\bm{W}}_{t})\setminus\mathfrak{M}_{e^{2\sigma}}({\bm{W}}_{t}) onto itself for which 𝒉(𝑹)=𝒓(t)[𝑹]((σ+δ)/2){\bm{h}}({\bm{R}})={\bm{r}}^{(t)}[{\bm{R}}]((\sigma+\delta)/2) and 𝒉(𝒓(t)[𝑹](σ,δ))=𝒓(t)[𝑹](σ,δ){\bm{h}}({\bm{r}}^{(t)}[{\bm{R}}](\sigma,\delta))={\bm{r}}^{(t)}[{\bm{R}}](\sigma,\delta), 𝑹𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}). With the aid of Theorem 1.2 we can extend it to a homeomorphism 𝒉\bm{h} of 𝔗g\mathfrak{T}_{g} onto itself with 𝒉(𝔐(𝑹0))=𝔐eσ+δ(𝑾t){\bm{h}}(\mathfrak{M}({\bm{R}}_{0}))=\mathfrak{M}_{e^{\sigma+\delta}}({\bm{W}}_{t}), completing the proof. ∎

We have proved most assertions of Theorem 1.1. The following proposition will finish the proof.

Proposition 14.8.

If 𝐑0{\bm{R}}_{0} is nonanalytically finite, then 𝔐(𝐑0)\mathfrak{M}({\bm{R}}_{0}) is a closed Lipschitz domain satisfying an outer ball condition.

Proof.

The Teichmüller distance function dTd_{T} is of class C2C^{2} on 𝔗g×𝔗g\mathfrak{T}_{g}\times\mathfrak{T}_{g} off the diagonal by Rees [47]. Hence Theorem 9.8 shows that 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) satisfies an outer ball condition.

We verify that 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) is locally expressed as the graph of a Lipschitz function. For 𝑹𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}) let (𝑹)\mathscr{E}({\bm{R}}) denote the set of (Σg;𝔐(𝑹0),𝑹)\mathcal{F}\in\mathscr{MF}(\Sigma_{g};\mathfrak{M}({\bm{R}}_{0}),{\bm{R}}) such that Ext(𝑹)=1\operatorname{Ext}_{\mathcal{F}}({\bm{R}})=1. For 𝔈𝔐(𝑹0)\mathfrak{E}\subset\partial\mathfrak{M}({\bm{R}}_{0}) set (𝔈)=𝑹𝔈(𝑹)\mathscr{E}(\mathfrak{E})=\bigcup_{{\bm{R}}\in\mathfrak{E}}\mathscr{E}({\bm{R}}).

Now, take 𝑹𝔐(𝑹0)\bm{R}\in\partial\mathfrak{M}({\bm{R}}_{0}) arbitrarily. Choose a sequence {tn}\{t_{n}\} of positive numbers with tn0t_{n}\to 0 so that {𝑹n}:={𝒓(tn)[𝑹](1)}\{{\bm{R}}^{\prime}_{n}\}:=\{{\bm{r}}^{(t_{n})}[{\bm{R}}](1)\} converges to a point 𝑹𝔗g{\bm{R}}^{\prime}\in\mathfrak{T}_{g}. Then 𝒓𝑹[𝑹]{\bm{r}}_{\bm{R}}[{\bm{R}}^{\prime}] is an Ioffe ray of 𝑹0{\bm{R}}_{0} by Lemma 14.2. If (𝑹)\mathcal{F}\in\mathscr{E}({\bm{R}}), then Proposition 14.5 and Corollary 14.4 imply

ReD𝑹[𝑸𝑹()](𝑸𝑹[𝑹n])\displaystyle\operatorname{Re}D_{{\bm{R}}}[{\bm{Q}}_{\bm{R}}(\mathcal{F})](-{\bm{Q}}_{\bm{R}}[{\bm{R}}^{\prime}_{n}]) =logExt(𝒓(tn)[𝑹](0))2τ(tn)(𝑹)+εn\displaystyle=\frac{\,\log\operatorname{Ext}_{\mathcal{F}}({\bm{r}}^{(t_{n})}[{\bm{R}}](0))\,}{2\tau^{(t_{n})}({\bm{R}})}+\varepsilon_{n}
1 2τ(tn)(𝑹)log(1+(c+Mtn)τ(tn)(𝑹))+εn\displaystyle\leqq\frac{1}{\,2\tau^{(t_{n})}({\bm{R}})\,}\log(1+(-c+Mt_{n})\tau^{(t_{n})}({\bm{R}}))+\varepsilon_{n}

for some positive cc and MM independent of \mathcal{F}, where εn0\varepsilon_{n}\to 0 as nn\to\infty. Letting nn\to\infty, we obtain

ReD𝑹[𝑸𝑹()](𝑸𝑹[𝑹])c 2.\operatorname{Re}D_{\bm{R}}[{\bm{Q}}_{\bm{R}}(\mathcal{F})]({\bm{Q}}_{\bm{R}}[{\bm{R}}^{\prime}])\geqq\frac{c}{\,2\,}.

Set 𝒢=𝑹(𝑸𝑹[𝑹])\mathcal{G}=\mathcal{H}_{\bm{R}}({\bm{Q}}_{\bm{R}}[{\bm{R}}^{\prime}]). Since 𝒓𝑹[𝑸𝑹[𝑹]]=𝒓𝑹[𝑹]{\bm{r}}_{\bm{R}}[{\bm{Q}}_{\bm{R}}[{\bm{R}}^{\prime}]]={\bm{r}}_{\bm{R}}[{\bm{R}}^{\prime}] is an Ioffe ray of 𝑹0{\bm{R}}_{0}, we know that 𝒢(𝑹)\mathcal{G}\in\mathscr{E}({\bm{R}}); note that Ext𝒢(𝑹)=𝑸𝑹[𝑹]𝑹=1\operatorname{Ext}_{\mathcal{G}}({\bm{R}})=\|{\bm{Q}}_{\bm{R}}[{\bm{R}}^{\prime}]\|_{\bm{R}}=1. The above inequality can be expressed as

(14.4) ReD𝑹[𝑸𝑹()](𝑸𝑹(𝒢))c 2\operatorname{Re}D_{\bm{R}}[{\bm{Q}}_{\bm{R}}(\mathcal{F})]({\bm{Q}}_{\bm{R}}(\mathcal{G}))\geqq\frac{c}{\,2\,}

for (𝑹)\mathcal{F}\in\mathscr{E}({\bm{R}}).

We examine the behavior of the function F:=(logExt)/2F_{\mathcal{F}}:=(\log\operatorname{Ext}_{\mathcal{F}})/2 along 𝒓𝑺[𝝋]{\bm{r}}_{\bm{S}}[{\bm{\varphi}}]. Proposition 14.5 shows that

ddtF(𝒓𝑺[𝝋](t))=ReD𝒓𝑺[𝝋](t)[𝑸𝒓𝑺[𝝋](t)()](𝑸𝒓𝑺[𝝋](t)[𝒓𝑺[𝝋](1)]),0<t<1.\frac{d}{\,dt\,}F_{\mathcal{F}}({\bm{r}}_{\bm{S}}[{\bm{\varphi}}](t))=\operatorname{Re}D_{{\bm{r}}_{\bm{S}}[{\bm{\varphi}}](t)}[{\bm{Q}}_{{\bm{r}}_{\bm{S}}[{\bm{\varphi}}](t)}(\mathcal{F})]({\bm{Q}}_{{\bm{r}}_{\bm{S}}[{\bm{\varphi}}](t)}[{\bm{r}}_{\bm{S}}[{\bm{\varphi}}](1)]),\quad 0<t<1.

In general for 𝑺𝔗g{\bm{S}}\in\mathfrak{T}_{g}, 𝝍A(𝑺){0}{\bm{\psi}}\in A({\bm{S}})\setminus\{0\}, ε>0\varepsilon>0 and δ>0\delta>0 define

U𝑺(ε,𝝍)\displaystyle U_{\bm{S}}(\varepsilon,{\bm{\psi}}) ={𝑸𝑺[𝑺]𝑺𝔗g{𝑺} and dT(𝒓𝑺[𝑸𝑺[𝑺]](1),𝒓𝑺[𝝍](1))<ε}and\displaystyle=\{{\bm{Q}}_{\bm{S}}[{\bm{S}^{\prime}}]\mid{\bm{S}}^{\prime}\in\mathfrak{T}_{g}\setminus\{{\bm{S}}\}\text{ and }d_{T}({\bm{r}}_{\bm{S}}[{\bm{Q}}_{\bm{S}}[{\bm{S}^{\prime}}]](1),{\bm{r}}_{\bm{S}}[{\bm{\psi}}](1))<\varepsilon\}\quad\text{and}
𝑺(ε,δ)\displaystyle\mathfrak{C}_{\bm{S}}(\varepsilon,\delta) ={𝒓𝑺[𝝋](t)𝝋U𝑺(ε,𝑸𝑺(𝒢)) and 0<t<δ}.\displaystyle=\{{\bm{r}}_{\bm{S}}[-{\bm{\varphi}}](t)\mid{\bm{\varphi}}\in U_{\bm{S}}(\varepsilon,{\bm{Q}}_{\bm{S}}(\mathcal{G}))\text{ and }0<t<\delta\}.

We claim that there is a neighborhood 𝔘\mathfrak{U} of 𝑹\bm{R} together with positive numbers ε0\varepsilon_{0} and δ0\delta_{0} such that

(14.5) ddtF(𝒓𝑺[𝝋](t))>0,0<t<δ0,\frac{d}{\,dt\,}F_{\mathcal{F}}({\bm{r}}_{\bm{S}}[{\bm{\varphi}}](t))>0,\quad 0<t<\delta_{0},

for all 𝑺𝔘{\bm{S}}\in\mathfrak{U}, (𝔐(𝑹0)𝔘)\mathcal{F}\in\mathscr{E}(\partial\mathfrak{M}({\bm{R}}_{0})\cap\mathfrak{U}), 𝝋U𝑺(ε0,𝑸𝑺(𝒢)){\bm{\varphi}}\in U_{\bm{S}}(\varepsilon_{0},{\bm{Q}}_{\bm{S}}(\mathcal{G})). If not, then there would exist sequences {𝑺n}\{{\bm{S}}_{n}\}, {n}\{\mathcal{F}_{n}\}, {𝝋n}\{{\bm{\varphi}}_{n}\} and {tn}\{t_{n}\} such that

  • (i)

    𝑺n𝑹{\bm{S}}_{n}\to{\bm{R}} as nn\to\infty,

  • (ii)

    there are (𝑹)\mathcal{F}\in\mathscr{E}({\bm{R}}) and 𝑺n𝔐(𝑹0){\bm{S}}^{\prime}_{n}\in\partial\mathfrak{M}({\bm{R}}_{0}) with n(𝑺n)\mathcal{F}_{n}\in\mathscr{E}({\bm{S}}^{\prime}_{n}) such that n\mathcal{F}_{n}\to\mathcal{F} and 𝑺n𝑹{\bm{S}}^{\prime}_{n}\to{\bm{R}} as nn\to\infty,

  • (iii)

    𝝋nA(𝑺n){\bm{\varphi}}_{n}\in A({\bm{S}}_{n}) with 𝝋n𝑺n=1\|{\bm{\varphi}}_{n}\|_{{\bm{S}}_{n}}=1, and 𝝋n𝑸𝑹(𝒢){\bm{\varphi}}_{n}\to{\bm{Q}}_{\bm{R}}(\mathcal{G}) as nn\to\infty,

  • (iv)

    tn>0t_{n}>0 and 𝑺n′′:=𝒓𝑺n[𝝋n](tn)𝑹{\bm{S}}^{\prime\prime}_{n}:={\bm{r}}_{{\bm{S}}_{n}}[{\bm{\varphi}}_{n}](t_{n})\to{\bm{R}} as nn\to\infty, and

  • (v)

    ReD𝑺n′′[𝑸𝑺n′′(n)](𝑸𝑺n′′[𝒓𝑺n[𝝋n](1)])0\operatorname{Re}D_{{\bm{S}}^{\prime\prime}_{n}}[{\bm{Q}}_{{\bm{S}}^{\prime\prime}_{n}}(\mathcal{F}_{n})]({\bm{Q}}_{{\bm{S}}^{\prime\prime}_{n}}[{\bm{r}}_{{\bm{S}}_{n}}[{\bm{\varphi}}_{n}](1)])\leqq 0.

Letting nn\to\infty in the last inequality, we obtain ReD𝑹[𝑸𝑹()](𝑸𝑹(𝒢))0\operatorname{Re}D_{\bm{R}}[{\bm{Q}}_{\bm{R}}(\mathcal{F})]({\bm{Q}}_{\bm{R}}(\mathcal{G}))\leqq 0, contradicting (14.4).

Owing to Theorem 1.6 (iii) and Proposition 14.7 we can replace 𝔘\mathfrak{U} with a smaller one so that 𝔐(𝑹0)𝔘\mathfrak{M}({\bm{R}}_{0})\cap\mathfrak{U} is the set of 𝑺𝔘{\bm{S}}\in\mathfrak{U} such that (1.2) holds for all (𝔐(𝑹0)𝔘)\mathcal{F}\in\mathscr{E}(\partial\mathfrak{M}({\bm{R}}_{0})\cap\mathfrak{U}). Since F𝒢F_{\mathcal{G}} is a C1C^{1} function on 𝔗g\mathfrak{T}_{g} with nonvanishing derivatives, its level hypersurfaces are C1C^{1} submanifolds of 𝔗g\mathfrak{T}_{g}. Replacing 𝔘\mathfrak{U} with a smaller one if necessary, we take a C1C^{1} coordinate system (𝔘,x)(\mathfrak{U},x) centered at 𝑹{\bm{R}} so that if we write x=(x1,,x2dg)=(x1,x)x=(x_{1},\ldots,x_{2d_{g}})=(x_{1},x^{\prime}), then the level hypersurfaces F𝒢1(a)F_{\mathcal{G}}^{-1}(a) are represented as x1=ax_{1}=a and the Teichmüller geodesic rays 𝒓𝑺[𝑸𝑺(𝒢)]{\bm{r}}_{\bm{S}}[{\bm{Q}}_{\bm{S}}(\mathcal{G})] (resp. 𝒓𝑺[𝑸𝑺(𝒢)]{\bm{r}}_{\bm{S}}[-{\bm{Q}}_{\bm{S}}(\mathcal{G})]) with F𝒢(𝑺)=aF_{\mathcal{G}}({\bm{S}})=a are represented as (a+ξ,x(𝑺))(a+\xi,x^{\prime}({\bm{S}})), ξ0\xi\geqq 0 (resp. ξ0\xi\leqq 0). Note that if 𝑺\bm{S} is in a small neighborhood 𝔙\mathfrak{V} of 𝑹\bm{R} and ε0\varepsilon_{0} and δ0\delta_{0} are sufficiently small, then 𝑺(ε0,δ0)\mathfrak{C}_{\bm{S}}(\varepsilon_{0},\delta_{0}) is included in 𝔘\mathfrak{U} and x(𝑺(ε0,δ0))x(\mathfrak{C}_{\bm{S}}(\varepsilon_{0},\delta_{0})) contains a cone with vertex at x(𝑺)x({\bm{S}}) and axis parallel to the x1x_{1}-axis, where the cones can be chosen to be of the same shape regardless of 𝑺\bm{S}.

Suppose that 𝑺𝔐(𝑹0)𝔙{\bm{S}}\in\partial\mathfrak{M}({\bm{R}}_{0})\cap\mathfrak{V}. From (14.5) we infer that for (𝔐(𝑹0)𝔘)\mathcal{F}\in\mathscr{E}(\partial\mathfrak{M}({\bm{R}}_{0})\cap\mathfrak{U}) the function FF_{\mathcal{F}} is decreasing along the Teichmüller geodesic segment 𝒓𝑺[𝝋](t){\bm{r}}_{\bm{S}}[-{\bm{\varphi}}](t), 0<t<δ00<t<\delta_{0}, for 𝝋U𝑺(ε0,𝑸𝑺(𝒢)){\bm{\varphi}}\in U_{\bm{S}}(\varepsilon_{0},{\bm{Q}}_{\bm{S}}(\mathcal{G})) and hence that

Ext(𝒓𝑺[𝝋](t))Ext(𝒓𝑺[𝝋](0))=Ext(𝑺)maxExt(𝔐(𝑹0)).\operatorname{Ext}_{\mathcal{F}}({\bm{r}}_{\bm{S}}[-{\bm{\varphi}}](t))\leqq\operatorname{Ext}_{\mathcal{F}}({\bm{r}}_{\bm{S}}[-{\bm{\varphi}}](0))=\operatorname{Ext}_{\mathcal{F}}({\bm{S}})\leqq\max\operatorname{Ext}_{\mathcal{F}}(\partial\mathfrak{M}({\bm{R}}_{0})).

Thus 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) includes 𝑺(ε0,δ0)\mathfrak{C}_{\bm{S}}(\varepsilon_{0},\delta_{0}). In particular, 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) includes 𝑹(ε0,δ0)\mathfrak{C}_{\bm{R}}(\varepsilon_{0},\delta_{0}).

Take a neighborhood 𝔚\mathfrak{W} of 𝑹\bm{R} included in 𝔙\mathfrak{V} so that if 𝑺F𝒢1(0)𝔚{\bm{S}}\in F_{\mathcal{G}}^{-1}(0)\cap\mathfrak{W}, then the Teichmüller geodesic ray 𝒓𝑺[𝑸𝑺(𝒢)]{\bm{r}}_{\bm{S}}[-{\bm{Q}}_{\bm{S}}(\mathcal{G})] hits 𝑹(ε0,δ0)\mathfrak{C}_{\bm{R}}(\varepsilon_{0},\delta_{0}). Set 𝑴(𝑺)=𝒓𝑺[𝑸𝑺(𝒢)](σ𝑺)\bm{M}({\bm{S}})={\bm{r}}_{\bm{S}}[-{\bm{Q}}_{\bm{S}}(\mathcal{G})](\sigma_{\bm{S}}), where σ𝑺\sigma_{\bm{S}} is the minimum of nonnegative ss for which 𝒓𝑺[𝑸𝑺(𝒢)](s)𝔐(𝑹0){\bm{r}}_{\bm{S}}[-{\bm{Q}}_{\bm{S}}(\mathcal{G})](s)\in\mathfrak{M}({\bm{R}}_{0}). Then 𝑴{\bm{M}} is a mapping of F1(0)𝔚F^{-1}(0)\cap\mathfrak{W} into 𝔐(𝑹0)𝔘\partial\mathfrak{M}({\bm{R}}_{0})\cap\mathfrak{U}, and 𝑴(𝑺)(ε0,δ0)\mathfrak{C}_{{\bm{M}}({\bm{S}})}(\varepsilon_{0},\delta_{0}) is included in 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}). Consequently, the function x(𝑺)x1(𝑴(𝑺))x^{\prime}({\bm{S}})\mapsto x_{1}({\bm{M}}({\bm{S}})), 𝑺F𝒢1(0)𝔙{\bm{S}}\in F_{\mathcal{G}}^{-1}(0)\cap\mathfrak{V}, is a Lipschitz function on a neighborhood of 0 in the plane x1=0x_{1}=0, and x(𝔐(𝑹0))x(\partial\mathfrak{M}({\bm{R}}_{0})) is the graph of the function. This completes the proof. ∎

Theorem 14.9.

Suppose that 𝐑0{\bm{R}}_{0} is nonanalytically finite. Let 𝐑\bm{R} be a boundary point of 𝔐(𝐑0)\mathfrak{M}({\bm{R}}_{0}). If there are two Ioffe rays of 𝐑0{\bm{R}}_{0} emanating from 𝐑\bm{R}, then 𝔐(𝐑0)\partial\mathfrak{M}({\bm{R}}_{0}) is not smooth at 𝐑\bm{R}.

Proof.

By assumption there are linearly independent 𝝋jAL(𝑹){\bm{\varphi}}_{j}\in A_{L}({\bm{R}}), j=1,2j=1,2, such that 𝒓𝑹[𝝋j](𝑹0){\bm{r}}_{\bm{R}}[{\bm{\varphi}}_{j}]\in\mathscr{I}({\bm{R}}_{0}). Let 𝔈j\mathfrak{E}_{j} be the set of 𝑺𝔗g{\bm{S}}\in\mathfrak{T}_{g} for which Extj(𝑺)Extj(𝑹)\operatorname{Ext}_{\mathcal{F}_{j}}({\bm{S}})\leqq\operatorname{Ext}_{\mathcal{F}_{j}}({\bm{R}}), where j=𝑹(𝝋j)\mathcal{F}_{j}=\mathcal{H}_{\bm{R}}({\bm{\varphi}}_{j}). Then 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is included in 𝔈1𝔈2\mathfrak{E}_{1}\cap\mathfrak{E}_{2} by Theorem 6.3. Since 𝔈1\partial\mathfrak{E}_{1} and 𝔈2\partial\mathfrak{E}_{2} meet transversally at 𝑹\bm{R}, the boundary 𝔐(𝑹0)\partial\mathfrak{M}({\bm{R}}_{0}) cannot be smooth at 𝑹\bm{R}. ∎

See Example 5.5 for the existence of 𝑹0{\bm{R}}_{0} satisfying the assumptions of the theorem. In the case of genus one 𝔐(𝑹0)\mathfrak{M}({\bm{R}}_{0}) is a closed ball with respect to the Teichmüller distance provided that it is not a singleton. This is not always the case for g>1g>1 because balls with respect to the Teichmüller distance have smooth boundaries.

We conclude the paper with the following proposition. It supplements Kahn-Pilgrim-Thurston [30, Theorem 2]. Note that for 𝑹𝔐(𝑹0){\bm{R}}\in\partial\mathfrak{M}({\bm{R}}_{0}) every element of CEmbhc(𝑹0,𝑹)\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}) has a dense image.

Proposition 14.10.

Suppose that 𝐑0=[R0,θ0]{\bm{R}}_{0}=[R_{0},\theta_{0}] is open and nonanalytically finite. If 𝐑=[R,θ]Int𝔐(𝐑0){\bm{R}}=[R,\theta]\in\operatorname{Int}\mathfrak{M}({\bm{R}}_{0}), then there are continuations (𝐑,𝛊j)({\bm{R}},{\bm{\iota}}_{j}), j=1,2j=1,2, of 𝐑0{\bm{R}}_{0} such that Rι1(R0)R\setminus\iota_{1}({R}_{0}) is of positive area while Rι2(R0)R\setminus\iota_{2}({R}_{0}) has a vanishing area, where 𝛊j=[ιj,θ0,θ]{\bm{\iota}}_{j}=[\iota_{j},\theta_{0},\theta].

Proof.

Fix 𝑹Int𝔐(𝑹0){\bm{R}}\in\operatorname{Int}\mathfrak{M}({\bm{R}}_{0}). If {(𝑾t(1),ϵt(1))}\{({\bm{W}}^{(1)}_{t},{\bm{\epsilon}}^{(1)}_{t})\} is a circular filling for 𝑹0{\bm{R}}_{0}, then 𝑹\bm{R} belongs to 𝔐(𝑾t1(1))\mathfrak{M}({\bm{W}}^{(1)}_{t_{1}}) for some positive t1t_{1} by Proposition 11.2. Thus there is a homotopically consistent conformal embedding 𝜿1{\bm{\kappa}}_{1} of 𝑾t1(1){\bm{W}}^{(1)}_{t_{1}} into 𝑹\bm{R}. Then 𝜾1:=𝜿1ϵt1(1){\bm{\iota}}_{1}:={\bm{\kappa}}_{1}\circ{\bm{\epsilon}}^{(1)}_{t_{1}} is a homotopically consistent conformal embedding of 𝑹0{\bm{R}}_{0} into 𝑹\bm{R} and (𝑹,𝜾1)({\bm{R}},{\bm{\iota}}_{1}) is not a dense continuation of 𝑹0{\bm{R}}_{0}.

Next let {(𝑾t(2),ϵt(2))}t[0,1]\{({\bm{W}}^{(2)}_{t},{\bm{\epsilon}}^{(2)}_{t})\}_{t\in[0,1]} be a linear filling for 𝑹0{\bm{R}}_{0}. If 𝔐(𝑾1(2))={𝑹}\mathfrak{M}({\bm{W}}^{(2)}_{1})=\{{\bm{R}}\}, then ϵ1(2){\bm{\epsilon}}^{(2)}_{1} induces a required element 𝜾2{\bm{\iota}}_{2} of CEmbhc(𝑹0,𝑹)\operatorname{CEmb}_{\mathrm{hc}}({\bm{R}}_{0},{\bm{R}}). Otherwise, 𝑹𝔐(𝑾t(2)){\bm{R}}\not\in\mathfrak{M}({\bm{W}}^{(2)}_{t}) for some t<1t<1. Let t2t_{2} be the infimum of such tt. For each t(t2,1]t\in(t_{2},1] let 𝒓t{\bm{r}}_{t} denote the Ioffe ray of 𝑾t(2){\bm{W}}^{(2)}_{t} passing through 𝑹\bm{R}. Then 𝒓t(0)𝑹{\bm{r}}_{t}(0)\to{\bm{R}} as tt2t\to t_{2}. On the other hand, if we take a sequence {τn}\{\tau_{n}\} in (t2,1](t_{2},1] so that τnt2\tau_{n}\to t_{2} and 𝑺n:=𝒓τn(1)𝑺𝔗g{\bm{S}}_{n}:={\bm{r}}_{\tau_{n}}(1)\to{\bm{S}}\in\mathfrak{T}_{g} as nn\to\infty, then Lemma 14.2 implies that 𝒓τn(0)=𝑯[𝑾τn(2)](𝑺n)𝑯[𝑾t2(2)](𝑺){\bm{r}}_{\tau_{n}}(0)={\bm{H}}[{\bm{W}}^{(2)}_{\tau_{n}}]({\bm{S}}_{n})\to{\bm{H}}[{\bm{W}}^{(2)}_{t_{2}}]({\bm{S}}) as nn\to\infty. Hence 𝑹=𝑯[𝑾t2(2)](𝑺)𝔐(𝑾t2(2)){\bm{R}}={\bm{H}}[{\bm{W}}^{(2)}_{t_{2}}]({\bm{S}})\in\partial\mathfrak{M}({\bm{W}}^{(2)}_{t_{2}}). If 𝜿2{\bm{\kappa}}_{2} is a homotopically consistent conformal embedding of 𝑾t2(2){\bm{W}}^{(2)}_{t_{2}} into 𝑹\bm{R}, then 𝜾2:=𝜿2ϵt2(2){\bm{\iota}}_{2}:={\bm{\kappa}}_{2}\circ{\bm{\epsilon}}^{(2)}_{t_{2}} possesses the required properties. ∎

References

  • [1] W. Abikoff, The Real Analytic Theory of Teichmüller Space, Lecture Notes in Math. 820, Springer-Verlag, Berlin-Heidelberg-New, York, 1980.
  • [2] L. V. Ahlfors, Lectures on Quasiconformal Mappings, 2nd ed., Wadsworth, Belmont, 1987.
  • [3] L. V. Ahlfors and L. Sario, Riemann Surfaces, Princeton Univ. Press, Princeton, 1960.
  • [4] V. Alberge, H. Miyachi and K. Ohshika, Null-set compactifications of Teichmüller spaces, Handbook of Teichmüller Theory,Vol. VI, 71–94, IRMA Lect. Math. Theor. Phys. 27, Eur. Math. Soc., Zürich, 2016.
  • [5] R. Arens, Topologies for homeomorphism groups, Amer. J. Math. 68 (1946), 593-610.
  • [6] A. F. Beardon, The Geometry of Discrete Groups, Springer-Verlag, Berlin-Heidelberg-New, York, 1980.
  • [7] M. Biernacki, Sur la représentation conforme des domaines linéairement accessibles, Prace Mat. Fiz. 44 (1936), 293–314.
  • [8] C. Bishop, A counterexample in conformal welding concerning Hausdorff dimension, Michigan Math. J. 35 (1988), 151–159.
  • [9] C. Bishop, Conformal welding of rectifiable curves, Math. Scand. 67 (1990), 61–72.
  • [10] S. Bochner, Fortsetzung Riemannscher Flächen, Math. Ann. 98 (1928), 406–421.
  • [11] M. F. Bourque, The converse of the Schwarz lemma is false, Ann. Acad. Sci. Fenn. Math. 41 (2016), 235–241.
  • [12] M. F. Bourque, The holomorphic couch theorem, Invent. Math. 212 (2018), 319–406.
  • [13] C. J. Earle, The Teichmüller distance is differentiable, Duke Math. J. 44 (1977), 389–397.
  • [14] C. Earle and A. Marden, Conformal embeddings of Riemann surfaces, J. Anal. Math. 34 (1978), 194–203.
  • [15] R. Fehlmann and F. P. Gardiner, Extremal problems for quadratic differentials, Mich. Math. J. 43 (1995), 573–591.
  • [16] F. P. Gardiner, Measured foliations and the minimal norm property for quadratic differentials, Acta Math. 152 (1984), 57–76.
  • [17] F. P. Gardiner, Teichmüller theory and quadratic differentials, New York, John Wiley & Sons, 1987.
  • [18] F. P. Gardiner and N. Lakic, Quasiconformal Teichmüller theory, the American Mathematical Society, Providence, 2000.
  • [19] S. Hamano, Variation of Schiffer and hyperbolic spans under pseudoconvexity, Geometric complex analysis, Springer Proc. Math. Stat. 246, Springer, Singapore (2018), 171–178.
  • [20] S. Hamano, M. Shiba and H. Yamaguchi, Hyperbolic span and pseudoconvexity, Kyoto J. Math. 57 (2017), 165–183.
  • [21] D. H. Hamilton, Generalized conformal welding, Ann. Acad. Sci. Ser. AI Math. 16 (1991), 333–343.
  • [22] M. Heins, A problem concerning the continuation of Riemann surfaces, Ann. of Math. Studies 30 (1953), 55–62.
  • [23] R. Horiuchi and M. Shiba, Deformation of a torus by attaching the Riemann sphere, J. Reine Angew. Math. 456 (1994), 135–149.
  • [24] J. Hubbard and H. Masur, Quadratic differentials and foliations, Acta Math. 142 (1979), 221–274.
  • [25] M. S. Ioffe, Extremal quasiconformal embeddings of Riemann surfaces, Sib. Math. J. 16 (1975), 398–411.
  • [26] H. Ishida, Conformal sewings of slit regions, Proc. Japan Acad. 50 (1974), 616–618.
  • [27] M. Ito and M. Shiba, Area theorems for conformal mapping and Rankine ovoids, Computational methods and function theory 1997 (Nicosia), Ser. Approx. Decompos., 11, World Sci. Publ., River Edge (1999), 275–283.
  • [28] M. Ito and M. Shiba, Nonlinearity of energy of Rankine flows on a torus, Nonlinear Anal. 47 (2001), 5467–5477.
  • [29] W. Kaplan, Close-to-convex schlicht functions, Michigan Math. J. 1 (1952), 169–185.
  • [30] J. Kahn, K. M. Pilgrim and D. P. Thurston, Conformal surface embeddings and extremal length, Groups Geom. Dyn. 16 (2022) 403–435.
  • [31] Z. Lewandowski, Sur l’identité de certaines classes de fonctions univalentes. I, Ann. Univ. Mariae Curie-Skłodowska Sect. A 12 (1958), 131–146.
  • [32] Z. Lewandowski, Sur l’identité de certaines classes de fonctions univalentes. II, Ann. Univ. Mariae Curie-Skłodowska Sect. A 14 (1960), 19–46.
  • [33] F. Maitani, Conformal welding of annuli, Osaka J. Math. 35 (1998), 579–591.
  • [34] M. Masumoto, Estimates of the euclidean span for an open Riemann surface of genus one, Hiroshima Math. J. 22 (1992), 573–582.
  • [35] M. Masumoto, On the moduli set continuations of an open Riemann surface of genus one, J. Anal. Math. 63 (1994), 287–301.
  • [36] M. Masumoto, Holomorphic and conformal mappings of open Riemann surfaces of genus one, Nonlinear Anal. 30 (1997), 5419–5424.
  • [37] M. Masumoto, Extremal lengths of homology classes on Riemann surfaces, J. Reine Angew. Math. 508 (1999), 17–45.
  • [38] M. Masumoto, Hyperbolic lengths and conformal embeddings of Riemann surfaces, Israel J. Math. 116 (2000), 77–92.
  • [39] M. Masumoto, Once-holed tori embedded in Riemann surfaces, Math. Z. 257 (2007), 453–464.
  • [40] M. Masumoto, Corrections and complements to “Once-holed tori embedded in Riemann surfaces”, Math. Z. 267 (2011), 869–874.
  • [41] M. Masumoto, On critical extremal length for the existence of holomorphic mappings of once-holed tori, J. Inequal. Appl. 2013:282 (2013), 1–4.
  • [42] M. Masumoto, Holomorphic mappings of once-holed tori, J. Anal. Math. 129 (2016), 69–90.
  • [43] M. Masumoto, Holomorphic mappings of once-holed tori, II, J. Anal. Math. 139 (2019), 597–612.
  • [44] K. Nishikawa and F. Maitani, Moduli of domains obtained by a conformal welding, Kodai Math. J. 20 (1997), 161–171.
  • [45] K. Oikawa, On the prolongation of an open Riemann surface of finite genus, Kodai Math. Sem. Rep. 9 (1957), 34–41.
  • [46] K. Oikawa, Welding of polygons and the type of a Riemann surface, Kodai Math. Sem. Rep. 13 (1961), 37–52.
  • [47] M. Rees, Teichmüller distance for analytically finite surfaces is C2C^{2}, Proc. London Math. Soc. 85 (2002), 686–716.
  • [48] B. Rodin and L. Sario, Principal Functions, Van Nostrand, Princeton, 1968.
  • [49] R. Sasai, Generalized obstacle problem, J. Math. Kyoto Univ. 45 (2005), 183–203.
  • [50] R. Sasai, Non-uniqueness of obstacle problem on finite Riemann surface, Kodai Math. J. 29 (2006), 51–57.
  • [51] G. Schmieder and M. Shiba, Realisierungen des idealen Randes einer Riemannschen Fläche unter konformen Abschließungen, Arch. Math. (Basel) 68 (1997), 36–44.
  • [52] G. Schmieder and M. Shiba, One-parameter family of the ideal boundary and compact continuations of a Riemann surface, Analysis (Munich) 18 (1998), 125–130.
  • [53] G. Schmieder and M. Shiba, On the size of the ideal boundary of a finite Riemann surface, Ann. Univ. Mariae Curie-Skłodowska Sect. A 55 (2001), 175–180.
  • [54] S. Semmes, A counterexample in conformal welding concerning chord arc curves, Ark. Math. 24 (1985), 141–158.
  • [55] M. Shiba, The Riemann-Hurwitz relation, parallel slit covering map, and continuation of an open Riemann surface of finite genus, Hiroshima Math. J. 14 (1984), 371-399.
  • [56] M. Shiba, The moduli of compact continuations of an open Riemann surface of genus one, Trans. Amer. Math. Soc. 301 (1987), 299–311.
  • [57] M. Shiba, The period matrices of compact continuations of an open Riemann surface of finite genus, Holomorphic Functions and Moduli I, ed. by D. Drasin et al., Springer-Verlag, New York-Berlin-Heidelberg (1988), 237–246.
  • [58] M. Shiba, Conformal embeddings of an open Riemann surface into closed surfaces of the same genus, Analytic Function Theory of One Complex Variable, ed. by Y. Komatu et al., Longman Scientific & Technical, Essex (1989), 287–298.
  • [59] M. Shiba, The euclidean, hyperbolic, and spherical spans of an open Riemann surface of low genus and the related area theorems, Kodai Math. J. 16 (1993), 118–137.
  • [60] M. Shiba, Analytic continuation beyond the ideal boundary, Analytic extension formulas and their applications (Fukuoka, 1999/Kyoto, 2000), Kluwer Acad. Publ., Dordrecht (2001), 233–250.
  • [61] M. Shiba, Conformal mapping of Riemann surfaces and the classical theory of univalent functions, Publ. Inst. Math. (Beograd) (N.S.) 75 (89) (2004), 217–232.
  • [62] M. Shiba, Conformal embeddings of an open Riemann surface into another — a counter part of univalent function theory, Interdiscip. Inform. 25 (2019), 23–31.
  • [63] M. Shiba and K. Shibata, Singular hydrodynamical continuations of finite Riemann surfaces, J. Math. Kyoto Univ. 25 (1985), 745–755.
  • [64] M. Shiba and K. Shibata, Hydrodynamic continuations of an open Riemann surface of finite genus, Complex Variables Theory Appl. 8 (1987), 205–211.
  • [65] K. Strebel, Quadratic Differentials, Springer-Verlag, Berlin-Heidelberg-New York-Tokyo, 1984.
  • [66] G. B. Williams, Constructing conformal maps of triangulated surfaces, J. Math. Anal. Appl. 390 (2012), 113–120.
Masumoto: Department of Mathematics, Yamaguchi University, Yamaguchi 753-8512, Japan
E-mail: [email protected]

Shiba: Professor emeritus, Hiroshima University, Hiroshima 739-8511, Japan
E-mail: [email protected]