Closed continuations of Riemann surfaces
Abstract. Any open Riemann surface of finite genus can be conformally embedded into a closed Riemann surface of the same genus, that is, is realized as a subdomain of a closed Riemann surface of genus . We are concerned with the set of such closed Riemann surfaces. We formulate the problem in the Teichmüller space setting to investigate geometric properties of . We show, among other things, that is a closed Lipschitz domain homeomorphic to a closed ball provided that is nonanalytically finite.
Keywords: Riemann surface, conformal embedding, Teichmüller space, quadratic differential, measured foliation, extremal length
MSC2020: 30Fxx, 32G15
1 Introduction
Let be an open Riemann surface of finite genus . If , then the general uniformization theorem assures us that is conformally equivalent to a domain on the Riemann sphere . In 1928 Bochner generalized it to the cases of higher genera in his study [10] on continuations of Riemann surfaces, showing that can be realized as a subdomain of a closed Riemann surface of the same genus . 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 with a domain conformally equivalent to provided that . It is then natural to ask into which closed Riemann surfaces of genus the Riemann surface can be conformally embedded.
Heins [22] tackled the problem for , and proved in 1953 that the set of closed Riemann surfaces of genus one which are continuations of 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 into which can be mapped by a homotopically consistent conformal embedding is compact and connected in the Teichmüller space of genus . 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 of marked Riemann surfaces of finite genus and conformally compatible continuous mappings in §3. Its class of objects consists of equivalence classes , where is a Riemann surface of genus , not necessarily closed, and is a sense-preserving homeomorphism of into , where is a fixed surface obtained from a closed oriented topological surface of genus by deleting one point (see Definition 3.5). A continuous mapping of into another marked Riemann surface is an equivalence class , where is a continuous mapping of into (see Definition 3.9). The composition is meaningful only for conformally compatible continuous mappings (see Definition 3.10). Also, a quadratic differential on is an equivalence class , where is a quadratic differential on (see §6).
The Teichmüller space of genus is defined to be the full-subcategory of whose class of objects is composed of all marked closed Riemann surfaces of genus . Equipped with the Teichmüller distance, is a metric space homeomorphic to , where . In fact, as is well-known, is a -dimensional complex manifold biholomorphic to a bounded domain in . It is quite new to understand the Teichmüller space as a full-subcategory of . 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 be a marked finite open Riemann surface of positive genus ; the fundamental group of is finitely generated, or equivalently, 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 is conformally equivalent to a Riemann surface obtained from a closed Riemann surface by removing finitely many points. For let be the set of homotopically consistent conformal embeddings of into (see Definition 3.12). By a closed continuation of we mean a pair , where and . We are concerned with the set of for which .
Theorem 1.1.
Let be a marked finite open Riemann surface of positive genus . Then is either a singleton or a closed Lipschitz domain homeomorphic to a closed ball in . The former case occurs if and only if is analytically finite.
The theorem is proved in the final section, where we show a stronger assertion that there is a homeomorphism of onto that maps onto a point or a closed ball. Being a closed Lipschitz domain, satisfies inner and outer cone conditions unless is analytically finite. Actually, it satisfies an outer ball condition. If , then its boundary is smooth and a sphere (or rather circle) with respect to the Teichmüller distance. On the other hand, if , then the boundary is nonsmooth for some and hence cannot be a sphere with respect to the Teichmüller distance.
For let denote the set of into which can be mapped by a homotopically consistent -quasiconformal embedding. Thus . If , then is the closed -neighborhood of .
Theorem 1.2.
Let be a marked finite open Riemann surface of positive genus . If , then is a closed domain homeomorphic to a closed ball in and has a -boundary.
Actually, there is a homeomorphism of onto such that , , 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 , which is also useful to construct various examples of closed continuations. Let denote the set of nonzero holomorphic quadratic differentials on with along the border . We use to identify arcs on the border. This operation gives rise to a continuous mapping of onto a marked closed Riemann surface holomorphic and injective on the interior of together with a meromorphic quadratic differential on . The pair is referred to as a closed self-welding of , and the quadratic differentials and are called a welder of and the co-welder of , respectively. Important are the self-weldings with holomorphic co-welders. Such a self-welding is said to be regular, or -regular if we need to refer to the welder .
If 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 , possibly, with finitely many points deleted. Each closed self-welding of yields a closed continuation of , called a closed self-welding continuation of . A welder of the self-welding is also referred to as a welder of the continuation. If is -regular, then is said to be -regular. If this is the case, then is a Teichmüller conformal embedding of into , and is an initial quadratic differential of .
Let be the space of nonzero holomorphic quadratic differentials on that can be extended to elements in . Denote by the set of such that for any horizontal trajectory of on its -length is at most the half of the -length of the component of including . Clearly, is a proper subset of .
Theorem 1.3.
Let be a marked nonanalytically finite open Riemann surface of positive genus . An element of induces a closed regular self-welding continuation of if and only if it belongs to .
The theorem shows that the procedure of closed regular self-welding provides us with an explicit method of obtaining all Teichmüller conformal embeddings of . Their importance is clarified in the next theorem.
For and let be the set of such that is a closed -regular self-welding continuation of ; it can be empty. Set . If , then is called exceptional. Let be the set of exceptional quadratic differentials in . Define and .
For let denote the set of for which . If , then is a singleton. Set and .
Theorem 1.4.
Let be a marked nonanalytically finite open Riemann surface of positive genus .
-
(i)
The boundary coincides with .
-
(ii)
If and , then .
-
(iii)
If , then has a nonempty interior with respect to the relative topology on .
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 is nowhere dense in (Theorem 10.3). Nevertheless, quadratic differentials in yield abundant boundary points of through closed regular self-weldings, as assertion (iii) claims. If , then . If , then there are examples of for which has a nonempty interior in .
Let be the space of measured foliations on . The horizontal foliation of a holomorphic quadratic differential on a marked closed Riemann surface determines an element of through . This defines a homeomorphism of the space of holomorphic quadratic differentials on onto by Hubbard-Masur [24].
Now, for , taking a closed -regular self-welding continuation of and denoting by the co-welder of , we set . Then is a well-defined mapping of into .
Theorem 1.5.
If is a marked nonanalytically finite open Riemann surface of positive genus , then is a homeomorphism of onto .
The extremal length function is a nonnegative continuous function on . Specifically, if , then is exactly , where stands for the -norm of a quadratic differential on a marked Riemann surface .
For let denote the set of points of at which the restriction of to attains its maximum. It is nonempty since is compact.
Theorem 1.6.
Let be a marked nonanalytically finite open Riemann surface of positive genus .
-
(i)
Let , and set . Then and
(1.1) -
(ii)
If , then there exists such that while for all .
-
(iii)
Let . Then if and only if
(1.2) for all .
Kahn-Pilgrim-Thurston [30, Theorem 1] gives a necessary and sufficient condition for to belong to in terms of stretch factors. A stretch factor of a topological embedding of into is defined with the extremal lengths of simple closed multi-curves on . 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 to obtain the stretch factors. Our theorem asserts that those in are sufficient.
Besides Theorems 1.3 – 1.6, the Teichmüller geodesic rays induced from the co-welders of elements of 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 , , with 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 and be topological spaces. Denote by the set of continuous mappings of into ; we avoid employing the usual notation so as not to confuse it with the set of conformal mappings in the case where and are Riemann surfaces. The subset of homeomorphisms of onto is denoted by . A topological embedding of into is, by definition, a mapping of into for which the correspondence defines an element of , where is endowed with the relative topology induced from the topology of . We denote by the set of topological embeddings of into . If and are oriented surfaces, then and denote the sets of sense-preserving elements in and , respectively.
For , , we write if is homotopic to . Also, for two loops and on we use the notation to mean that is freely homotopic to . This usage of is consistent with the previous one as we can regard and as continuous mappings of the unit circle into .
A surface means a connected -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 ; it is a bordered Riemann surface. Though the holomorphic mapping defined by is injective, it is not a conformal embedding since is not a homeomorphism of onto its image .
For Riemann surfaces and the sets of conformal embeddings of into and conformal homeomorphisms of onto are denoted by and , respectively. By a conformal automorphism of we mean a conformal homeomorphism of onto itself. Let be the group of conformal automorphisms of . Thus . Its subgroup consisting of satisfying is denoted by .
Example 2.2.
Conformal automorphisms of are of the form , where . Since is arcwise connected, we know that . The translations , , form a subgroup .
Example 2.3.
For each let denote the additive subgroup of generated by and . We endow the quotient group with conformal structure so that the natural projection is a holomorphic universal covering map. Then is a torus. As is well-known, each torus is conformally equivalent to some . For later use we introduce a few more notations. We denote by and the simple loops on defined by , , . For any the translation is a conformal automorphism of for which the image loops and are freely homotopic to and , respectively. If , then we have . The set of translations of is a subgroup of . In fact, is identical with (see Corollary 3.3 (ii)).
Definition 2.4 (continuation).
Let be a Riemann surface. A continuation of is, by definition, a pair where is a Riemann surface and is a conformal embedding of into .
Two continuations , , of are defined to be equivalent to each other if there is such that . A continuation of is called closed (resp. open, compact) if is closed (resp. open, compact). If is dense in , then we say that is a dense continuation of . A continuation of is said to be genus-preserving if each connected component of is topologically embedded into . In the case where is of finite genus, a continuation of is genus-preserving if and only if is of the same genus as . Observe that if is a bordered Riemann surface, then there is a genus-preserving open continuation of such that is a retract of . If is a continuation of and is a continuation of , then is a continuation of .
Remark.
Let be a quasiconformal embedding of a Riemann surface without border into a Riemann surface. If is a local parameter around a point and is a local parameter around , then with respect to these parameters is regarded as a complex function . The quantity does not depend on a particular choice of and defines a measurable -form on called the Beltrami differential of . Then is a measurable function on , whose -norm is less than one. The maximal dilatation of is defined by . Denote by and the sets of quasiconformal embeddings of into and quasiconformal homeomorphisms of onto , respectively.
Let , , where and are Riemann surfaces without border. If almost everywhere on , then . For any measurable -form on with there is a quasiconformal homeomorphism of onto a Riemann surface without border for which almost everywhere on .
It is convenient to define quasiconformal embeddings also on bordered Riemann surfaces. Let be a bordered Riemann surface. A topological embedding of into another Riemann surface is called quasiconformal if there are continuations and of and , respectively, where and are Riemann surfaces without border, together with such that . We can then speak of the maximal dilatation . Note that . Each component of the border of is mapped by onto a simple arc or loop of vanishing area.
Let be a real constant with . A quasiconformal embedding is called -quasiconformal if its maximal dilatation does not exceed . Conformal embeddings are -quasiconformal, and vice versa.
A subset of a Riemann surface is called a subsurface of if it carries its own conformal structure such that the inclusion mapping of into is a conformal embedding of into . Each component of the border of , if any, is an analytic curve on . Subdomains of , or nonempty connected open subsets of , are subsurfaces of .
Let be a bordered Riemann surface. We denote its border by , and set , called the interior of . Note that is a subsurface of and that , which is also a topological boundary of , is a 1-dimensional real analytic submanifold of . If denotes the inclusion mapping of into , then we call the canonical continuation of . In the case where is a Riemann surface without border, we define and for convenience, and call the interior of .
Definition 2.5 (finite Riemann surface).
A Riemann surface is called finite if its fundamental group is finitely generated.
A Riemann surface is said to be analytically finite if there is a closed continuation of such that is a finite set, that is, , where denotes the cardinal number of a set . 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 is an analytically finite Riemann surface and is a quasiconformal homeomorphism of onto another Riemann surface , then is also analytically finite.
Definition 2.6 (natural continuation).
Let be a finite open Riemann surface. A compact continuation of is called natural if contains at most finitely many points.
Clearly, is a dense continuation of . To obtain a natural compact continuation of we have only to consider the case where is nonanalytically finite. Then the universal covering Riemann surface of is conformally equivalent to the upper half plane . The covering transformation group is a torsion-free Fuchsian group of the second kind keeping invariant, and is conformally equivalent to . Let denote the region of discontinuity of , and set , where is the closure of in the extended complex plane . If is compact, then we set . Otherwise, has finitely many punctures and we let be the Riemann surface with the punctures filled in. In any case we have a conformal embedding of into to obtain a desired continuation . As is easily verified, two natural compact continuations of are equivalent to each other.
Example 2.7.
Let and be the open unit disk and the open square in the complex plane , respectively. They are subsurfaces of . Möbius transformations of the closure onto make a compact bordered Riemann surface so that the inclusion mapping of into is a conformal embedding. In other words, is a continuation of . Thus is a subsurface of . Take such that . Then is a natural compact continuation of . The embedding is extended to a homeomorphism of the topological closure onto . Its inverse defines a topological embedding of into , but is not a continuation of , for, is not analytic at four points on the border . Obviously, is a subsurface of though is not. Also, if , then is a natural compact continuation of .
3 Categorical definition of Teichmüller spaces
Let 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 . We are interested in the set of closed Riemann surfaces of the same genus as into which 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 of marked Riemann surfaces of genus and conformally compatible continuous mappings. Fix a closed oriented surface of positive genus , and remove one point from to obtain a once-punctured closed oriented surface . For the sake of definiteness, using the notations in Example 2.3, we set and .
Definition 3.1 (handle mark).
Let be a positive integer, and let be an oriented surface of genus or higher, possibly, with border. By a -handle mark of we mean an element of .
If is a -handle mark of , then . Thus is also considered as a -handle mark of .
Proposition 3.2.
Let be a -handle mark of a Riemann surface of positive genus, and assume that satisfies .
-
(i)
If is not a torus, then .
-
(ii)
If is a torus, then .
Proof.
(i) The interior carries a hyperbolic metric, that is, a complete conformal metric of constant curvature . Let and be the hyperbolic geodesic loops on that are freely homotopic to and , respectively (for the definition of and see Example 2.3). Then and due to . Since is isometric with respect to the hyperbolic metric, and are also hyperbolic geodesic loops and hence trace and in the same direction, respectively. Because and have exactly one point, say , in common, we know that , which implies coincides with on and . Therefore by the identity theorem.
(ii) We may assume that is identical with some in Example 2.3. The conformal automorphism of is lifted to a conformal automorphism of . The 1-form on is projected to a holomorphic 1-form on . Since is freely homotopic to , the pull-back of via has the same period along as :
As is a torus, we obtain and hence , or equivalently, , , for some constant . Consequently, belongs to . ∎
Corollary 3.3.
Let be a Riemann surface of positive genus.
-
(i)
If is not a torus, then is trivial.
-
(ii)
If is a torus, then .
Proof.
Choose a -handle mark of . If , then . Thus the corollary follows at once from Proposition 3.2. ∎
Corollary 3.4.
Let be a -handle mark of a Riemann surface of positive genus, and suppose that satisfy .
-
(i)
If is not a torus, then .
-
(ii)
If is a torus, then .
Definition 3.5 (marked Riemann surface).
Consider all pairs , where is a Riemann surface of genus and is a -handle mark of . We say that is equivalent to if there is such that . We call each equivalence class a marked Riemann surface of genus .
A marked Riemann surface is called a marked bordered Riemann surface or a marked Riemann surface with border if is a bordered Riemann surface. Otherwise, is referred to as a marked Riemann surface without border. If is closed (resp. open, compact), then is called closed (resp. open, compact). We also say that is a marked closed (resp. open, compact) Riemann surface. The meaning of a marked torus is obvious. If is finite (resp. analytically finite, nonanalytically finite), then is said to be finite (resp. analytically finite, nonanalytically finite). Since , regarding as a topological embedding of into , we obtain a marked Riemann surface without border, which will be called the interior of .
Remark.
Let . Suppose that satisfy for . Choose a -handle mark of to obtain a -handle mark of . Since , it follows from Corollary 3.4 that if is a marked torus and that otherwise.
Example 3.6.
We have obtained a torus for each in Example 2.3. The homeomorphism of onto itself induces and a marked torus , where .
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 be a subdomain of such that is compact, and set . For each with there are and such that , and for , and .
Proof.
For and set . Then defines a homeomorphism of onto itself with on . The correspondence
possesses the desired properties. ∎
Proposition 3.8.
Let and be oriented surfaces of genus possibly with border, and let and be topological embeddings of into . If is closed and for a -handle mark of , then .
Proof.
Take for which . The simple loop divides into two surfaces, one of which, say , is homeomorphic to while the other, say , is planar, that is, of genus zero. Since , there is , where , such that for .
Introducing a conformal structure into each , let be a holomorphic universal covering map with covering transformation group , where or . Fix a point , and choose . Let denote the component of containing , and let be the subgroup of consisting of covering transformations leaving invariant. The restriction of to is a holomorphic covering map of , and acts on as the covering transformation group . The embedding induces an injective group homomorphism of the fundamental group of with base point into . The latter fundamental group is canonically isomorphic to with corresponding to . Let be the subset of assigned to the elements of represented by loops freely homotopic to loops in .
Any loop in with initial point is trivial in and hence is projected to a loop trivial in . Thus the induced group homomorphism maps the fundamental group to the trivial subgroup of . Therefore, there is such that . Set for . Since is discrete, there is a group homomorphism such that for . Note that does not depend on .
For there is together with a group homomorphism such that and for . Since is identical with on , we can choose so that . Then coincides with on . Observe that as is closed. Since is generated by and , we infer that and hence (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 have been marked with sense-preserving homeomorphisms of , not . Our definition leads us to essentially the same space of marked closed Riemann surfaces. In fact, if is a -handle mark of a closed Riemann surface of genus , then it follows from Proposition 3.7 that there is such that is equivalent to , where . Moreover, let be a -handle mark of a closed Riemann surface of genus , and assume that it is extended to a homeomorphism of onto . If is equivalent to , then for some . The inclusion mapping is a -handle mark of . Since , Proposition 3.8 implies that .
Definition 3.9 (continuous mapping).
Let , , be marked Riemann surfaces of genus , and consider all triples , where and . Two such triples and , where and , are said to be equivalent to each other if for some with . We call each equivalence class a continuous mapping of into and use the notation .
If has some conformally invariant properties in addition, then is said to possess the same properties. For example, if is a quasiconformal embedding of into , then is called a quasiconformal embedding of into .
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 and be Riemann surfaces of positive genus. Then is said to be conformally compatible if for any there is such that .
If is not a torus, then every continuous mapping of into is conformally compatible as . If is a torus while is not, then every conformally compatible continuous mapping of into is constant because acts transitively on while is trivial. To determine conformally compatible continuous mappings of into in the case where both and are tori, take holomorphic universal covering mappings , . Suppose that is conformally compatible. Let be a lift of . Thus we have . For any there is such that . Since is conformally compatible, we find for which , which leads us to
for some . Therefore, we have for , where . Since , it follows that for and hence for by continuity. Consequently, is an -linear mapping of into itself, and is an -affine transformation. For any in the covering transformation group of there is such that . Conversely, if is an -affine mapping possessing the last property, then it induces a conformally compatible element such that . In particular, the inverse of a conformally compatible homeomorphism of a torus onto another is conformally compatible. Note that whether is a torus or not, conformal embeddings are conformally compatible.
Let , , be marked Riemann surfaces of genus , and let be a continuous mapping of into . Take representatives , where . If is conformally compatible, then so is . Hence we can speak of a conformally compatible continuous mapping of a marked Riemann surface into another.
Definition 3.11 (composition).
Let and be conformally compatible continuous mappings. Take and , where , , and . Define a continuous mapping of into to be the equivalence class , where with .
It is easy to verify that is well-defined, that is, that represents a unique continuous mapping of into . Actually, is well-defined if is conformally compatible. We do not need to require to be conformally compatible.
Compositions of conformally compatible continuous mappings of marked Riemann surfaces are again conformally compatible. Also, the binary operation is associative.
In general, for a marked Riemann surface of genus let denote the conformal automorphism of represented by , where . Then for any conformally compatible continuous mappings of into .
We are now ready to define the category . Its class of objects consists of all marked Riemann surfaces of genus . The hom-set of morphisms from to is composed of all conformally compatible continuous mappings of into . The Teichmüller space of genus is, by definition, the full subcategory of whose class of objects is exactly the set of marked closed Riemann surfaces of genus . For the sake of simplicity we abbreviate and to and , respectively.
Remark.
The set is not the set of continuous mappings of into . Every element of should be conformally compatible.
Denote by (resp. , ) the set of conformally compatible topological (resp. quasiconformal, conformal) embeddings of into . The set of sense-preserving elements in is denoted by . Also, let denote the set of surjective elements in , and set
Replacing “Emb” with “Homeo” in the notations for the other classes above, we obtain notations for the subset of homeomorphisms of onto in the classes under consideration. Thus means the set of conformally compatible quasiconformal homeomorphisms of onto .
Definition 3.12 (homotopically consistent embedding).
A sense-preserving topological embedding is called homotopically consistent if , where , , and .
This definition does not depend on a particular choice of representatives. If and are homotopically consistent, then so is . If is a subset of , then we denote by the subset of homotopically consistent elements in . For example, means the set of homotopically consistent and conformally compatible quasiconformal embeddings of into .
Let and be continuous mappings of into . There are , , with , where . If , then we say that is homotopic to and write . Again, this definition does not depend on a particular choice of representatives.
A continuous mapping of into is said to be homotopically consistent if it is homotopic to a homotopically consistent topological embedding of into . If , , are homotopically consistent, then so is . We use the notation to express the set of homotopically consistent elements in .
Definition 3.13 (continuation).
By a continuation of we mean a pair where and .
A continuation is called closed (resp. open, compact) if is closed (resp. open, compact). We say that is dense if so is , where , and . If and are continuations of and , respectively, then is a continuation of .
Let . If is a genus-preserving continuation of , then, setting and , we obtain a continuation of . In particular, if is a subsurface of and is the inclusion mapping, then we call the inclusion continuation of . If this is the case, then for we call the restriction of to and denote it by .
Let be a marked finite open Riemann surface of genus , and let be a natural compact continuation of . Setting and , we obtain a continuation of , which is referred to as the natural compact continuation of .
Let . For quasiconformal embeddings of into we can speak of their maximal dilatations . If belongs to , then so does , and their Teichmüller distance is defined to be , where the infimum is taken over all homotopically consistent quasiconformal homeomorphisms of onto . The complete metric space is in fact known to be a complex manifold homeomorphic to and biholomorphic to a bounded domain in , where .
Now, for we are concerned with the set of such that . It follows from Bochner [10, Satz V] that is nonempty.
Example 3.14.
Under the notations in Example 3.6 the correspondence is a biholomorphism of onto . If we identify with through the biholomorphism, then the Teichmüller distance on coincides with the distance on induced by the hyperbolic metric . Now, let be a marked open Riemann surface of genus one. Then is a closed disk or a singleton in by [56, Theorem 5]. It degenerates to a singleton if and only if (see [56, Theorem 6]). Note that a finite open Riemann surface belongs to 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 be a quadratic differential on a Riemann surface ; it is an assignment of a function of to each local coordinate so that is invariant under coordinate changes. Then , , defines a 2-form on . If is measurable, that is, is measurable for each , then set
for measurable subsets of . If is finite, then is said to be integrable over . Also, and are invariant under coordinate changes. The -length and the -height of a curve on are defined by
respectively, provided that the integrals are meaningful. Otherwise, we just set or . Note that is the area of with respect to the area element corresponding to the length element .
Let be a continuation of a Riemann surface , and let be a quadratic differential on . For take local coordinates and around and , respectively, and consider as a holomorphic function . Set , where is the function assigned to the local coordinate by the quadratic differential . Then assigning to defines a quadratic differential on called the pull-back of by to be denoted by .
Let . Considering as a conformal homeomorphism onto , we can speak of the pull-back of a quadratic differential on by . It is a quadratic differential on , which will be denoted by .
A quadratic differential is called meromorphic (resp. holomorphic) if is meromorphic (resp. holomorphic) for each local coordinate . If is a bordered Riemann surface and is a meromorphic (resp. holomorphic) quadratic differential on , then for any open continuation of there is a meromorphic (resp. holomorphic) quadratic differential on a neighborhood of such that the pull-back coincides with on .
Let be the complex vector space of meromorphic quadratic differentials on . Denote by the subspace composed of holomorphic quadratic differentials on . For nonzero the algebraic degree of at is denoted by . If and , then, taking and as in the preceding paragraph, we understand that . By a critical point of we mean a point for which . It is a zero of of order if while it is a pole of of order if . Zeros and poles on the border are referred to as border zeros and border poles, respectively. If , then there is a local coordinate around with such that (see Strebel [65, Theorem 6.1]). Such a local coordinate will be called a natural parameter of around . If for an open set , then for all , that is, all poles of in are simple.
A meromorphic quadratic differential is said to be positive along a smooth curve , where is an interval in which may be closed, half-closed, open or infinite, if is finite and positive whenever is in the domain of a local coordinate . A horizontal trajectory of is, by definition, a maximal smooth curve along which is positive. A horizontal trajectory can be a loop called a closed horizontal trajectory. Note that any horizontal trajectory of contains no critical points of . By a horizontal arc we mean a simple subarc, not a loop, of a horizontal trajectory of . For example, if is a natural parameter of around with , then for sufficiently small the arcs , , are horizontal arcs of emanating from though does not lie on the arcs. If is a noncritical point of , that is, if , then is a horizontal arc of passing through .
Suppose that is a bordered Riemann surface. A nonzero element of is called positive if consists of horizontal trajectories and, possibly, zeros of ; thus is holomorphic on . The set of positive quadratic differentials in is denoted by . Define . If denotes the double of (for the definition, see [3, II.3E]), then the reflection principle enables us to extend every element in (resp. ) to a quadratic differential in (resp. ). Note that the orders of border zeros of elements in are even. For the sake of convenience we set and for closed Riemann surfaces .
Let be a finite open Riemann surface, and let be a natural compact continuation of . Then is a compact Riemann surface. Set and . These spaces do not depend of a particular choice of , and are determined solely by .
Proposition 4.1.
Let be a finite open Riemann surface, and let be a nonzero meromorphic quadratic differential on . If there is a sequence in such that as for some open subset of , then and as .
Proof.
Let be a natural compact continuation of . If is bordered, then let denote its double. Otherwise, set . In either case is a closed Riemann surface. Identifying with , we consider as a subdomain of , and regard as a subset of . Both and are norms on ; note that implies by the identity theorem. Since is of finite dimension, there is such that for all .
Now, since is a Cauchy sequence in the Banach space , it is also a Cauchy sequence in and hence there is for which as . The identity theorem implies that on . Thus is nonzero and belongs to . ∎
Definition 4.2 (self-welding).
Let be a compact bordered Riemann surface. A self-welding of is a pair , where
-
(i)
is a compact Riemann surface,
-
(ii)
and with , and
-
(iii)
there are and with on such that consists of finitely many horizontal arcs and, possibly, critical points of .
The quadratic differentials and are referred to as a welder of the welding and the co-welder of the welder , respectively. If and are of the same genus, then is said to be genus-preserving. If is closed, then is called closed.
Two self-weldings and of are defined to be equivalent to each other if there is such that . Here, we do not force the self-weldings to have a common welder.
Remark.
Example 4.3.
Consider the meromorphic functions , , on defined by . Then , , are genus-preserving closed self-weldings of , which are inequivalent to each other. The meromorphic quadratic differential on is a common welder of the self-weldings. The co-welders of are for and for . Note that has double zeros at while has no critical points on . Also, is a welder of . Its co-welder is .
Definition 4.4 (self-welding continuation).
Let be a nonanalytically finite open Riemann surface. A compact continuation of is called a self-welding continuation of if there are a natural compact continuation of and a self-welding of such that .
If is a welder of the self-welding , then is said to be a welder of the continuation . The co-welder of is also referred to as the co-welder of .
Let be a compact bordered Riemann surface. If is a self-welding of , then is included in . Moreover, is a self-welding continuation of . If the self-welding is closed, then so is the induced continuation.
Let be a nonanalytically finite open Riemann surface, and let be a natural compact continuation of . Equivalent self-weldings of induce equivalent continuations of . Conversely, let be a self-welding continuation of , and let be a continuation of equivalent to . Then there is a self-welding of such that . Also, there is such that . Thus is a self-welding of equivalent to and induces the continuation of .
Proposition 4.5.
Let be a compact bordered Riemann surface, and let be a dense compact continuation of . Suppose that there are and such that . Then is extended to with for which is a self-welding of with welder .
Proof.
Take an arbitrary point on the border , and let be a natural parameter of around . We choose so that it maps a neighborhood of conformally onto a half-disk . If , then on ; recall that the order of at is even. Set for . Thus a single-valued branch of the integral of on is represented as and maps each onto a half disk of radius centered at . Let be a sequence in a sector converging to . Since is compact, a subsequence of , to be denoted again by , converges to a point in . As , the function on defined by is a single-valued branch of the integral of , and maps horizontal arcs of in onto those of in . Since as , we know that . Let be a natural parameter of around so that , and set for . Then maps conformally onto some . It follows that extends continuously to so that maps each of the two components of onto a horizontal arc of .
We have shown that is extended to a continuous mapping of into so that consists of finitely many horizontal arcs and critical points of . Since is dense in and is compact, we infer that is surjective and that is included in . Consequently, is a self-welding of with welder . ∎
Corollary 4.6.
Let be a nonanalytically finite open Riemann surface, and let be a dense compact continuation of . If there are and such that , then is a self-welding continuation of with welder .
Proof.
Let be a natural compact continuation of . Since punctures are removable singularities for conformal embeddings into compact Riemann surfaces, there is such that . Clearly, is a dense compact continuation of . Moreover, on for some , for, . It follows from Proposition 4.5 that is extended to with for which is a self-welding of with welder . As and , we obtain the corollary. ∎
Corollary 4.7.
Let be a nonanalytically finite open Riemann surface, and let be a compact continuation of . If there is such that consists of finitely many horizontal arcs of together with finitely many points, then is a self-welding continuation of with welder .
Proof.
Let and be as in the proof of the preceding corollary. Thus is a dense compact continuation of with . Since consists of finitely many horizontal arcs of together with finitely many points, it follows from a theorem of Caratédory that is extended to a continuous mapping of onto , which is holomorphic on off finitely points by the reflection principle, and hence the pull-back is extended to a quadratic differential in . Corollary 4.6 shows that is a self-welding continuation of with welder . ∎
We now introduce a constructive procedure to obtain self-weldings of a compact bordered Riemann surface . It is different from the welding procedure introduced in Ahlfors-Sario [3, II.3C–D] because is not required to be holomorphic on the border . In addition, we need to take into account quadratic differentials on and the corresponding quadratic differentials on the resulting Riemann surfaces. In the following, for a curve on a topological space , where is an interval on , by abuse of notation we sometimes use the same letter to denote its image . Let , and let and be simple arcs, not loops, of the same -length, say, , on the border . They may pass through zeros of . Suppose that each of the arcs contains its endpoints and that the arcs are nonoverlapping, that is, and are disjoint, where stands for the part obtained from a simple arc by deleting its endpoints; we allow and to have a common endpoint. Then we can identify the arcs to obtain a new Riemann surface so that induces a meromorphic quadratic differential on the new surface.
Specifically, parametrize , , with -length parameter. Thus the -length of the subarc of the arc is identical with . There are two natural ways of identifying and . One is to identify with for , and the other is to identify with for . Exactly one of the identifications leads us to an orientable topological surface , possibly with border. For the sake of definiteness we assume that the first identification has been chosen. The genus of is the same as that of if and only if and lie on the same component of . Let be the natural continuous mapping. Then is identical with , which defines a simple arc . Note that lies in the interior though the endpoints of may or may not be on the border .
We endow with a conformal structure as follows. We begin with adopting local parameters at each point of off so that is a conformal homeomorphism of onto . We then apply an extension of the welding procedure described in [3, II.3C–D] to choose local parameters at each point on . To be more precise, taking , let be a natural parameter of around , where is considered as a subsurface of its double . A small neighborhood of in is mapped conformally onto a neighborhood of in by . We may suppose that lies on the left and right of and , respectively, so that while . Note that
for sufficiently near , where with nonnegative integer . Take a neighborhood of included in , and define a mapping by
with for . Then is a well-defined element of , which is holomorphic off . With the aid of these parameters we make a Riemann surface. If the endpoints of lie on the border , then we are done. If , then . Let be a natural parameter of around the common endpoint. Then the function defined by is homeomorphic on a neighborhood of and holomorphic on a punctured neighborhood of . Thus works as an analytic local parameter around . We deal with the case where in a similar manner. We have thus endowed with conformal structure.
If is a natural compact continuation of , then is extended to a homeomorphism of onto . We introduce a conformal structure on so that the homeomorphism is actually a conformal homeomorphism of onto . We have thus obtained a compact Riemann surface .
The continuous mapping is holomorphic on . Let be a point on . If lies on , then is holomorphic at if and only if , where is the other point than projected to . If is an endpoint of or , then is holomorphic at if and only if is a common endpoint of and , or equivalently, lies in .
The quadratic differential on induced by from is extended to a meromorphic quadratic differential on . Observe that
(4.1) |
In particular, has a pole on if and only if and has a common endpoint at which does not vanish. If this is the case, then the pole is simple and lies in the interior . In any case, is holomorphic and positive on and hence belongs to . Clearly, the simple arc is composed of finitely many horizontal arcs and, possibly, critical points of . Therefore, is a self-welding of with welder , and is the co-welder of . We say that is the self-welding of with welder along .
More generally, let , , be nonoverlapping simple arcs on with for . Identifying with as above for leads us to
-
(i)
a compact Riemann surface ,
-
(ii)
a continuous mapping of onto with such that or for , and
-
(iii)
a quadratic differential in with on .
By subdividing ’s if necessary, it is no harm to assume from the outset that each pair induces a simple arc, not a loop, on . The pair will be called the self-welding of with welder along , .
Remark.
Let be the self-welding of with welder along , , and let be the co-welder of . Then the image arcs and are nonoverlapping simple arcs of the same -length on the border . Hence we can construct the self-welding of with welder along . It is easy to verify that is a self-welding of equivalent to . Therefore, the self-welding is also obtained by consecutive applications of self-welding procedures along one pair of arcs.
Example 4.8.
The quadratic differential belongs to , and the arcs on defined by , , are of the same -length. Let be the self-welding of with welder along , , and . Then is a torus. Hence the self-welding is closed but not genus-preserving.
Any self-welding of is a self-welding of along finitely many pairs of simple arcs on . To see this, let be an arbitrary self-welding of with welder , and denote by the co-welder of . Then consists of finitely many nonoverlapping simple arcs . Note that every consists of horizontal arcs and, possibly, critical points of . Observe that there are nonoverlapping simple arcs on such that for each . Then is the self-welding of with welder along , .
Example 4.9.
Let be the self-weldings of in Example 4.3. The arcs and on defined by , , are of the same -length, where . Then is the self-welding of with welder along . If , then is the self-welding of with welder along .
5 Genus-preserving closed regular self-weldings
Let be a compact bordered Riemann surface. We introduce important classes of self-weldings of .
Definition 5.1 (full self-welding).
Let be a union of connected components of . A genus-preserving self-welding of is called -full if while .
As is easily verified, a self-welding of is -full if and only if it is genus-preserving and closed. Note that the self-welding of with welder along , , is -full if and only if the arcs , , exhaust and and lie on the same component of without separating any other pairs for .
Example 5.2.
Each induces a -full self-welding of . To obtain an example let be the connected components of . Divide each into two subarcs and of the same -length. Then the self-welding of with welder along , , is -full.
Let be a -full self-welding of with welder , and set . We consider as a graph, called the weld graph of , by declaring the points with to be the vertices of . Each component of is in fact a tree. Note that each component of contains more than one vertex. If is a vertex of , then
(5.1) |
where is the co-welder of . An end-vertex of is a vertex with . It is characterized as a point on whose preimage by is a singleton (see (5.1)).
Definition 5.3 (regular self-welding).
A self-welding of is called -regular if is a welder of the self-welding and the co-welder of is holomorphic on . A self-welding of is said to be regular if it is -regular for some welder of the self-welding.
Remark.
If a self-welding of is -regular, then necessarily belongs to . Note that if is the co-welder of , then .
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 can be a common welder of infinitely many inequivalent genus-preserving closed regular self-weldings of .
Example 5.4.
Let be a nonzero holomorphic semiexact 1-form on whose imaginary part vanishes along (for the definition of a semiexact 1-form see [3, V.5B]). Let be a hydrodynamic continuation of with respect to introduced by Shiba-Shibata [64]. Thus is a closed Riemann surface of the same genus as , and is a conformal embedding of into such that for some holomorphic 1-form on for which consists of finitely many arcs along which vanishes. Corollary 4.7 implies that is extended to a continuous mapping of onto such that is a genus-preserving closed -regular self-welding of ; note that is a positive holomorphic quadratic differential on . As is remarked in [55, Section 26], for some there is inducing uncountably many hydrodynamic continuations of , or uncountably many inequivalent genus-preserving closed -regular self-weldings of .
If is a welder of a self-welding of , then so is for any positive real number . The next example shows that linearly independent quadratic differentials can induce the same genus-preserving closed regular self-welding.
Example 5.5.
Let be a closed Riemann surface of genus , and assume that it is symmetric in the sense that it admits an anti-conformal involution whose set of fixed points is composed of finitely many analytic simple loops on . Take a simple arc on , and set . The subsurface is a nonanalytically finite open Riemann surface of genus . Let be a natural compact continuation of . The element of defined by , , is extended to a continuous mapping of onto . Let be the vector space over of holomorphic 1-forms on such that along . Recall that . For any nonzero the pull-back of its square belongs to and the genus-preserving closed self-welding of is -regular.
Let be a union of connected components of . Each positive meromorphic quadratic differential on defines infinitely many -full inequivalent self-weldings (see Example 5.2). We now ask which quadratic differentials on make -full -regular self-weldings.
Proposition 5.6.
Let , and let be a union of connected components of . A -full self-welding of is -regular if and only if every end-vertex of its weld graph is the image of a border zero of by .
Proof.
The following corollary is a generalization of Shiba-Shibata [64, Lemma 3].
Corollary 5.7.
Let , and let be a union of connected components of . If there is a -full -regular self-welding of , then each component of contains two or more zeros of .
Proof.
If is a -full -regular self-welding of , then each component of contains at least two points and whose images by are end-vertices of the weld graph. Since the self-welding is -regular, and must be zeros of by Proposition 5.6. ∎
Definition 5.8 (border length condition).
Let be a union of connected components of . A positive meromorphic quadratic differential on is said to satisfy the border length condition on if
for any component of and any horizontal trajectory of included in .
Denote by the set of positive meromorphic quadratic differentials on satisfying the border length condition on , and set . We abbreviate and to and , respectively.
Theorem 5.9.
Let be a compact bordered Riemann surface, and let be a union of connected components of . For there exists a -full -regular self-welding of if and only if satisfies the border length condition on .
Corollary 5.10.
For there is a genus-preserving closed -regular self-welding of if and only if .
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 (resp. ). For example, each component of the weld graph of the self-welding in Example 5.2 is I-shaped.
Lemma 5.11.
Let . Let , , be points on a component of which divide into three arcs of -lengths less than . Then there is a -full self-welding of with welder such that its weld graph is a Y-shaped tree with end-vertices , . If is a -full self-welding of whose weld graph is Y-shaped with end-vertices , , then is equivalent to .
Proof.
We denote by , , the arcs on obtained from by cutting at , , where we label them so that . By assumption we have
and
Take a point on to divide into two arcs, and name the resulting six arcs , , as follows:
If we choose so that
then for and hence we can make a -full self-welding of with welder along , . Its weld graph is a Y-shaped tree and the end-vertices of are exactly , .
Let be another -full self-welding of with welder whose weld graph is Y-shaped with end-vertices , . Since is Y-shaped, it has one more vertex with and three edges , , where . As
where is the co-welder of , the lengths , , are uniquely determined, and hence depends only on . Consequently, is equivalent to . ∎
Remark.
The three points , , in the above proof are projected to the same point on . It is a zero of the co-welder of even if none of is a zero of .
Proof of Theorem 5.9.
Let be an arbitrary -full self-welding of with welder . Take nonoverlapping arcs , , on so that is the self-welding of with welder along , . By subdividing the arcs if necessary we may assume that every zero of on is an endpoint of some . Then the closure of each horizontal trajectory of on is a union of ’s. Denote by the co-welder of .
If , then some component of includes a horizontal trajectory of with . Since and are of the same -length and lies on if does, some pair, which may be assumed to be and , lie on . If they have a common endpoint on , then it is not a zero of but is mapped to an end-vertex of the weld graph , and hence the self-welding is not -regular by Proposition 5.6. Otherwise, the closure of has a component included in . Let be the self-welding of with welder along . Observe that is a component of the border and contains exactly one zero of the co-welder of . It follows from Corollary 5.7 that the self-welding of with welder along , , is not -regular. Therefore has a pole on as it is the co-welder of . This means that the self-welding of is not -regular, either. We have thus proved that if some -full self-welding of is -regular, then satisfies the border length condition on .
To prove the converse suppose that . Let be a component of . We will take one or three pairs of arcs of equal -lengths on to obtain a -full -regular self-welding of . Then the co-welder satisfies the border length condition on , or . Since the number of components of is smaller by one than that of , repeating this process leads us to a -full -regular self-welding of .
Let be the set of zeros of on . The border length condition implies that . If there are two points in which divide into two arcs and of the same -length, then the self-welding of with welder along is -regular by Proposition 5.6. Otherwise, contains at least three points. Two distinct points divide into two arcs. Let denote the one with shorter -length. Thus is the simple arc on joining and whose -length is less than . Choose two distinct points so that is the largest among , where with . The border length condition implies that contains a point that does not lie on . Since for , where , it follows from Lemma 5.11 that there is a -full self-welding of with welder whose weld graph is Y-shaped with end-vertices , . Thus the self-welding is -regular by Proposition 5.6, for, the points , , are zeros of . This completes the proof. ∎
Example 5.12.
We consider the case of genus one. Let be a genus-preserving closed -regular self-welding of , and let be the co-welder of . Since is a torus, for some holomorphic -form on . Set . Since is positive, the imaginary part of vanishes along . Moreover, is semiexact as for all components of . Therefore, is a hydrodynamic continuation of with respect to . As is free of zeros, it follows from (4.1) that has exactly two zeros on each component of , and these points are projected to end-vertices of the weld graph by Proposition 5.6. Since has no other end-vertices, each component of is I-shaped.
Let be a finite open Riemann surface, and let denote a natural compact continuation of . Assuming that is nonanalytically finite, let be a genus-preserving closed self-welding continuation of . For the self-welding continuation is said to be -regular if is a welder of and the co-welder of is holomorphic on . It is called regular if it is -regular for some . By Corollary 5.10 there is a genus-preserving closed -regular self-welding continuation of if and only if belongs to , which is a proper subset of . In the case where is analytically finite, we define . In either case is closed in .
As will be remarked in §8, a closed continuation of a finite open Riemann surface is a regular self-welding continuation if and only if is a Teichmüller conformal embedding. If this is the case, then an welder of and its co-welder are initial and terminal differentials of , 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 . Let , and consider all pairs , where represents and is a quadratic differential on . Two such pairs and , where , are defined to be equivalent to each other if for some with . Each equivalence class is called a quadratic differential on . If possesses some conformally invariant properties, then we say that has the same properties. For example, is called meromorphic if is. If is measurable on , then we set . If is finite, then is called integrable. Also, is said to coincide with almost everywhere on if almost everywhere on for some and , where . If is not a marked torus, then for quadratic differentials and on and complex numbers we define and in the obvious manner:
where and and are quadratic differentials on . If is a marked torus, then the addition operator is available only if or is invariant under conformal automorphisms homotopic to the identity, that is, or is holomorphic on .
Let , where , , and let be a quadratic differential on . If is not a marked torus, then the pull-back of is defined by . This definition does not depend on a particular choice of representatives. In the case where is a marked torus, the pull-back operator applies only to holomorphic quadratic differentials on .
The sets of meromorphic and holomorphic quadratic differentials on are denoted by and , respectively. We define , , and to be the subsets consisting of those with belonging to , , and , respectively, provided that those spaces are meaningful. If is closed, that is, if , then is a complex Banach space of dimension with norm .
Now, let be a marked compact bordered Riemann surface of genus , and let be a genus-preserving self-welding of . Set , which is a -handle mark of as , to obtain and . If is a self-welding of equivalent to and is a -handle mark of with , then and . Also, if and , then is a self-welding of with and represents . Thus is a well-defined pair, which we call a self-welding of . If is a welder of and is its co-welder, then we call and a welder of and the co-welder of , respectively. If , then is said to be -regular. If this is the case, then and . A self-welding is called regular if it is -regular for some . Also, if has some additional properties, then we say that possesses the same properties. For example, If is closed, then so is .
Let us return to our investigations on . We are exclusively concerned with the case where is a marked finite open Riemann surface of positive genus . Let denote the natural compact continuation of . Since punctures are removable singularities for conformal embeddings of a Riemann surface into closed Riemann surfaces, we have . Without causing any trouble we sometimes assume, if necessary, that is the interior of a marked compact bordered Riemann surface.
Suppose that is nonanalytically finite. Then is a marked compact bordered Riemann surface. A continuation of is said to be a self-welding continuation of if for some self-welding of . The pull-back of a welder of the self-welding is referred to as a welder of the self-welding continuation , and the co-welder of is also called the co-welder of . If the self-welding is -regular, then the continuation is said to be -regular. A self-welding continuation is called regular if it is -regular for some welder .
Proposition 6.1.
Let be a marked nonanalytically finite open Riemann surface, and let be a dense compact continuation of . If there are and such that , then is a self-welding continuation of with welder .
This is an immediate consequence of Corollary 4.6. Recall that the pull-back is well-defined for even if is a marked torus.
Let be a marked nonanalytically finite open Riemann surface of genus . Note that Theorem 1.3 follows at once from Corollary 5.10. Let . For let be the set of such that is a closed -regular self-welding continuation of . It may be an empty set. Let denote the set of for which . We are interested in the set
If is analytically finite, then is exactly the singleton . In this case we set for the sake of convenience.
For the investigation of we recall the definition and some properties of measured foliations on surfaces. Let denote the set of free homotopy classes of homotopically nontrivial simple loops on . The set of nonnegative functions on is identified with the product space , where . We endow it with the topology of pointwise convergence. Following [4], we define to be the closure of the set of functions of the form with and , where denotes the geometric intersection number of and , that is, is the minimum of the numbers of common points of loops in and . Every element of is called a measured foliation on . If and , then .
Important examples of measured foliations are those induced by holomorphic quadratic differentials defined as follows. Let . Choose a closed Riemann surface of genus together with so that , where . For measurable quadratic differentials on define a mapping by
This definition does not depend on a particular choice of representatives even if . In the case where , the mapping is a measured foliation on called the horizontal foliation of . Note that for .
Proposition 6.2 (Hubbard-Masur [24]).
For any the correspondence
is a homeomorphism of onto .
For the proof see also Gardiner [16, Theorem 6]. The inverse of the homeomorphism will be denoted by . Thus stands for the holomorphic quadratic differential on such that . In fact, the correspondence defines a homeomorphism of onto the complex vector bundle of holomorphic quadratic differentials over . The bundle is canonically identified with the cotangent bundle of the Teichmüller space through the bilinear form on the space of pairs of bounded measurable -forms on and holomorphic quadratic differentials on . Note that for .
For and the extremal length of on is defined by
Set to obtain a nonnegative function on .
Theorem 6.3.
Let be a marked nonanalytically finite open Riemann surface. Let be a closed -regular self-welding continuation of , and let be the co-welder of . Set . Then
for all .
In other words, the function attains its maximum on at . 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 as well as uniqueness of conformal embeddings (see §§9 and 12).
In general, let be a measurable quadratic differential on . Define a mapping of into by , where the infimum is taken over all for which is a piecewise analytic simple loop on .
Proposition 6.4.
Let and , and let be an integrable quadratic differential on . If for all , then
The sign of equality occurs if and only if almost everywhere on .
Proof.
Let and . Though 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 of , where , we choose a horizontal arc of in so that is integrable on as in [18, Lemma 4 in §12.7], where the roles of and are interchanged and the letter is used for . Then the reasoning in [17] works without any further modifications. ∎
Let be a subsurface of a Riemann surface , and let be a quadratic differential on . By the zero-extension of to we mean the quadratic differential on defined by on and on . If is a continuation of and is a quadratic differential on , then the -zero-extension of means the quadratic differential on , where is the zero-extension of the quadratic differential on to .
Lemma 6.5.
Let , , and be as in Theorem 6.3, and let be a closed continuation of . If is the -zero-extension of , then for any the inequality
(6.1) |
holds, where .
Proof.
Take representatives , , and of , , and , respectively, where stands for the restriction of to . Let be the natural compact continuation of . There is a closed -regular self-welding of such that and with . We may assume that has no punctures so that . Let be the integral of , which is a multi-valued function and may have finitely many singular points on . Let be the components of . For each choose a doubly connected closed neighborhood in of so that
-
(i)
are mutually disjoint,
-
(ii)
each is divided into finitely many simply connected closed domains ,
-
(iii)
a branch of maps each homeomorphically onto a closed rectangle in with sides parallel to the real and imaginary axes, and
-
(iv)
the vertical sides of the rectangles , , , are of the same length.
Note that is composed of finitely many horizontal and vertical arcs of . Set and , where . Observe that is a component of , which is a topological closed disk on , even though is not injective on . Also, each meets the weld graph since .
Choose a closed Riemann surface of genus together with so that , where , and take a representative . Let be an arbitrary element of , and let be a piecewise analytic simple loop on . We may assume that the initial (and terminal) point of is in the domain . Divide into subarcs to obtain so that lie in while lie in . For let be the image arc of by . They are piecewise analytic simple arcs on . Note that the terminal point of and the initial point of lie on the same component, say , of . We choose a piecewise analytic simple arc joining with within as follows. Let be the co-welder of , and let . Take and so that and belong to and , respectively. We then let be a simple arc on joining with composed of horizontal and vertical arcs, and possibly, zeros of , where the horizontal and vertical arcs should lie on and , respectively. We require that each should include at most two vertical arcs, which implies . Some of ’s may have common points. We modify them by slightly shifting their horizontal and vertical arcs and going around the zeros of to obtain simple arcs without changing the endpoints so that and that are mutually disjoint. The simple loop belongs to the homotopy class , and satisfies and hence , which leads us to . Since , we obtain (6.1). ∎
Proof of Theorem 6.3.
Let , , be distinct points of . Then , where is the Teichmüller quasiconformal homeomorphism of onto . Thus belongs to and there are nonzero , , such that for some and with
-
(i)
the Beltrami differential of is exactly , where ,
-
(ii)
maps every noncritical point of to a noncritical point of , and
-
(iii)
is represented as
with respect to some natural parameter (resp. ) of (resp. ) around any noncritical point (resp. ).
In other words, is a uniform stretching along horizontal trajectories of . The quadratic differentials and are called initial and terminal quadratic differentials of , respectively. Note that is also the Teichmüller quasiconformal homeomorphism of onto , whose initial and terminal quadratic differentials are and , respectively.
Let and . For each there uniquely exists with such that is an initial quadratic differential of the Teichmüller quasiconformal homeomorphism of onto . Set . Then the mapping is a (simple) ray emanating from , called a Teichmüller geodesic ray. Note that
(6.2) |
for . For with we denote by the image of the interval by :
(6.3) |
If we need to refer to the initial point and the quadratic differential we use the notation for . Thus is an initial quadratic differential of the Teichmüller quasiconformal homeomorphism of onto . Let be the corresponding terminal quadratic differential of . If we set , then we have as is homotopically consistent. It follows that
(6.4) |
A Teichmüller geodesic ray is uniquely determined by the initial point and a point that the ray passes through. We write to denote such a ray. Note that we parametrize each Teichmüller geodesic ray with respect to the distance from its initial point. Therefore, for .
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 be a marked finite open Riemann surface. Then is included in the boundary .
Proof.
We have only to consider the case where is nonanalytically finite. Let be a closed -regular self-welding continuation of , and denote by the co-welder of . We know that belongs to . If , then it follows from (6.4) that for . Theorem 6.3 then assures us that lies outside of . Since tends to as , we conclude that is certainly on the boundary of . ∎
Remark.
We could apply Kahn-Pilgrim-Thurston [30] to prove Proposition 6.6. Let , and be as in the proof of the proposition. Approximating with Jenkins-Strebel quadratic differentials in , we know that the stretch factor of homotopically consistent topological embeddings of into is exactly , which implies that for 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 on is projected to a holomorphic quadratic differential on through the natural projection . Set . Recall that . If , then so that . Now, let be a marked finite open Riemann surface of genus one, and let be the natural compact continuation. Let and . For there exists a (unique) holomorphic semiexact 1-form on with such that the imaginary part of vanishes along the border . Set , which is an element of . Let be a closed -regular self-welding of . Then is the co-welder of the welder . Restricting ourselves to the case , we conclude from Theorem 6.3 that if . This is nothing but the first inequality of [56, Theorem 2 (I)]. We remark that is equal to the extremal length of the weak homology class of 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 actually exhaust .
Let be a -handle mark of a Riemann surface of positive genus, and let be a holomorphic universal covering map with covering transformation group . We normalize and as follows. Set and .
If is not a torus, then we may assume that . Then is a torsion-free Fuchsian group keeping invariant. We can choose and so that contains hyperbolic transformations and such that the repelling and attracting fixed points of (resp. ) are and (resp. and some negative real number), respectively, and that the axes of and oriented from the repelling fixed points to the attracting fixed points are projected onto closed hyperbolic geodesics and on freely homotopic to and , respectively.
If is a torus, then we may set . Let be the unique holomorphic 1-form on for which . Choose so that . If we set , then, using the notations in Example 2.3, we have and . Denote by and the transformations in defined by and .
In either case, is uniquely determined and is referred to as the universal -covering transformation group. Also, is said to be the standard -pair in . We call a holomorphic universal -covering map. Unless is a torus, it is uniquely decided.
Let be another -handle mark of . Set and , and let and be covering transformations in corresponding to and . This means, in the case where is not a torus, that and are hyperbolic transformations and that their axes have one point in common and are projected to the closed hyperbolic geodesics and freely homotopic to and . Let be the unique element in that maps , and to the repelling fixed point of , the attracting fixed point of and the repelling fixed point of , respectively. Then is the holomorphic universal -covering map. If is a torus, then is of the form with . Taking defined by , we obtain a holomorphic universal -covering map . In either case, is the universal -covering transformation group and the standard -pair is .
Now, let and be Riemann surfaces of positive genera, and let . If is a -handle mark of , then is a -handle mark of , and is lifted to a locally homeomorphic mapping of into such that
Note that is not necessarily injective. There is a group homomorphism such that
We can choose so that if is the standard -pair in , then is the standard -pair in . If is not a torus, then nor is , and those requirements determine and uniquely. If is a torus but is not, then we adjust in addition so that , where . Again, and are uniquely determined. In the case where is a torus, so is and is a member of . Moreover, is uniquely decided and we can choose and so that though we cannot assert the uniqueness of . In any case we say that and are a -lift of and the -homomorphism induced by , respectively, where we should call the -lift of unless is a torus. Note that if , then is extended to a quasiconformal homeomorphism of onto itself fixing , and .
Let . Then if and only if and induce the same -homomorphism.
Next, we define -handle marks of marked Riemann surfaces of genus . Let . Consider all pairs , where represents and is a -handle mark of . Two such pairs and are defined to be equivalent to each other if for some with , where , . Each equivalence class is called a -handle mark of . All representatives and have the universal -covering transformation group and the standard -pair in common. Thus we can speak of the universal -covering transformation group and the standard -pair in .
Let for , and take . If is a -handle mark of , then we denote by the -handle mark of represented by . If is not a marked torus, then the -lift of and the -homomorphism induced by are independent of a particular choice of representatives , and . We call and the -lift of and the -homomorphism induced by , respectively. If is a marked torus, then the -homomorphism induced by is well-defined while there are infinitely many -lifts of . Any one of them is referred to as a -lift of .
Let be a marked Riemann surface of genus without border. Then , , are identical with one another and it is no harm to denote it by . Thus if is a marked torus while otherwise.
Definition 7.1 (convergence of sequence of homeomorphisms).
Let , , , be marked Riemann surfaces of genus without border homeomorphic to one another, and let . In the case where is not a marked torus, the sequence is said to converge to if for some -handle mark of the sequence of -lifts of converges to locally uniformly on . If is a marked torus, then for to converge to we require some sequence of -lifts of to converge to locally uniformly on .
We claim that the above definition does not depend on . Set , and let be a -lift of , where . Suppose that converges to locally uniformly on . As , the -homomorphism induced by is given by , . Since converges to locally uniformly on (see Lemma 7.2 below), it follows that as for . Now, let be another -handle mark of , and set . Take corresponding to and . Since and as , we can choose with and so that . Consequently, the sequence tends to locally uniformly on . Since is a -lift of , 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 and be locally compact metric spaces homeomorphic to each other. If a sequence in converges to uniformly on every compact subset of , then converges to uniformly on every compact subset of .
Proof.
Let be an arbitrary compact subset of , and let . Take a compact subset of including the -neighborhood of . Since is uniformly continuous on , there is such that
whenever and , or equivalently,
whenever and , where and denote the distance functions on and , respectively. Replacing with a smaller one if necessary, we may assume that the -neighborhood of is included in .
Since converges to uniformly on , we can find for which
for and . Note that if and , then
for . In other words,
for if and .
We now claim that
for and , which will prove the lemma. To verify the claim take and arbitrarily, and set and . Then is a point of so that belongs to the -neighborhood of and hence to , for, . We then have
as claimed. ∎
Definition 7.3 (convergence of sequence of topological embeddings).
Let and be marked Riemann surfaces of genus without border, and let . Also, let and be sequences in , where and are supposed to be homeomorphic to and , respectively, and let . Then we say that converges to if there are and together with a -handle mark of such that
-
(i)
and converge to and , respectively, and
-
(ii)
the sequence of -lifts of converges to the -lift of locally uniformly on ,
provided that is not a marked torus. In the case where is a marked torus, we replace condition (ii) with
-
(ii′)
some sequence of -lifts of converges to a -lift of locally uniformly on .
Lemma 7.4.
Let and be marked Riemann surfaces of genus without border. Let be a sequence in and let be a continuation of . Assume that there are and such that , , are mutually homotopic. If and converge to and , respectively, then there is a continuation of such that a subsequence of converges to .
Proof.
Let be a -handle mark of , and set , and . Let and be - and -lifts of and , respectively. By assumption we can choose and so that they converge locally uniformly to and , respectively.
If is closed, then so are , and , and hence . Thus coincides with , and is a -lift of . The convergence properties of and imply that and converge to and in , respectively. In particular, is identical with , and converges to .
Suppose now that is open. Then . Set for the sake of simplicity. If and , then . Otherwise, .
Take and , and let , and . Then we have
and the -homomorphism of onto induced by is given by
Since converges to locally uniformly in , so does by Lemma 7.2.
Similarly, take and , and let , and . If is not a torus, then and are uniquely determined and satisfy
(7.1) |
If is a torus, then we can adjust , and to obtain (7.1) for all . The -homomorphism of onto induced by is given by
By Lemma 7.2 again, the sequences and converge to locally uniformly on . In the case where , we have and for some (for the notations see Example 2.3). Note that as .
Let and be the -lift of and the -homomorphism induced by , respectively, where . Then we have
and
Observe that is the -lift of . Since , , are homotopic to one another, they induce the same -homomorphism . It follows that
(7.2) |
for all .
We claim that has a subsequence converging locally uniformly on to , which is a holomorphic function or the constant . This is clear if . If , then there is a constant such that omits , for, is not surjective. Then it also omits , which is in the -orbit of . Therefore, forms a normal family. Since can be chosen to be bounded, we conclude that includes a desired subsequence.
For typographical reason we assume that converges to locally uniformly on . It then follows from (7.2) that
(7.3) |
If were a constant function, then the constant should satisfy
for and . Thus is a common fixed point of elements of , which implies that is abelian (see, for example, Beardon [6, Theorem 5.1.2]). This is impossible if . If , then as for all . However, no finite point is fixed by any nontrivial elements of and we again reach a contradiction. Therefore, is nonconstant and holomorphic, and hence is locally univalent as well as .
It follows from (7.3) that induces a holomorphic and locally homeomorphic mapping satisfying . We claim that it is injective. Otherwise, we could find two distinct points mapped to the same point of by . Let , , and set . Since
we have for some . Take a relatively compact neighborhood of such that and . Since (resp. , ) converges to (resp. , ) locally uniformly on (resp. , ), there are neighborhoods of with and elements in such that and for sufficiently large . Choose so that . Then we obtain
which implies that , or, for some , for, is injective. Then , which is absurd. This proves that is injective, as claimed. If we set , then is a continuation of and converges to . ∎
Proposition 7.5.
If is a marked open Riemann surface of genus , then any sequence of closed continuations of contains a subsequence such that and converge to some and , respectively.
Proof.
Take representatives , , and , . Note that . If , then , , carry hyperbolic metrics. Since decreases the hyperbolic metrics, for any simple loop on the hyperbolic length of does not exceed that of . With the aid of the Fenchel-Nielsen coordinates on (see, for example, Abikoff [1, Chapter II §3]) we deduce that is relatively compact in . If , then choosing a point for and making use of hyperbolic metrics on and , we reach the same conclusion. In either case includes a convergent subsequence. We may assume that itself converges to some in . If denotes the Teichmüller quasiconformal homeomorphism of onto , then converges to . We apply Lemma 7.4 to obtain desired subsequences and and a continuation of . ∎
Corollary 7.6 (Oikawa [45]).
If is a marked open Riemann surface of genus , then is compact.
Remark.
Note that 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 is connected. Theorem 1.1 gives an alternative proof of this fact in the case where is finite. The connectedness of for general easily follows from this special case, as was shown in [45].
Proposition 7.7.
Let be a marked finite open Riemann surface of genus , and let be a sequence of closed regular self-welding continuations of . If converges to a point in , then there is a closed regular self-welding continuation of such that a subsequence of converges to .
Proof.
It follows from Proposition 7.5 that a subsequence of converges to some . We may assume that converges to . We need to show that is a regular self-welding continuation of .
We know that . Set for . Fix a -handle mark of , and set . Let and be a welder of the self-welding continuation of and its co-welder, respectively. They induce automorphic -forms and for and , respectively. We adopt the normalization condition . As , by taking subsequences if necessary, we may assume that and converge locally uniformly to and , respectively. Proposition 4.1 implies that is projected to some with . Also, is projected to some with .
Denote by and the -lifts of and , respectively. We know that converges to locally uniformly on . Owing to , we have , or equivalently, . The continuation of is dense as . Proposition 6.1 thus implies that is a closed -regular self-welding continuation of . ∎
8 Continuations to Riemann surfaces on
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 be a marked nonanalytically finite open Riemann surface of genus . Then for any there is such that is a closed regular self-welding continuation of .
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 be a finite open Riemann surface of genus without border, and let be a closed Riemann surface of the same genus. Then is called a Teichmüller quasiconformal embedding if
-
(i)
its Beltrami differential is of the form for some and ,
-
(ii)
there is such that maps every noncritical point of to that of and is represented as
with respect to some natural parameter (resp. ) of (resp. ) around any noncritical point (resp. ), and
-
(iii)
the complement consists of finitely many horizontal arcs and, possibly, zeroes of together with finitely many isolated points.
The quadratic differentials and are referred to as initial and terminal quadratic differentials of , respectively. Note that maps each horizontal arc of onto a horizontal arc of and is a uniform stretching along horizontal trajectories of . The Beltrami differential of the inverse is equal to . If , then is conformal, in which case it is said to be a Teichmüller conformal embedding.
Remark.
Let be a genus-preserving closed continuation of . If it is a closed regular self-welding continuation, then is a Teichmüller conformal embedding, and a welder of and its co-welder are initial and terminal quadratic differentials of , respectively. Conversely, if is a Teichmüller conformal embedding, then is a closed regular self-welding continuation of by Corollary 4.7. More generally, we have the following lemma.
Lemma 8.3.
Let be a nonanalytically finite open Riemann surface of genus , and let be a closed Riemann surface of genus . If is a Teichmüller quasiconformal embedding of into with initial quadratic differential , then there are a closed -regular self-welding continuation of and a Teichmüller quasiconformal homeomorphism of onto such that
-
(i)
the co-welder of is an initial quadratic differential of , and
-
(ii)
.
Proof.
Let be the terminal quadratic differential of corresponding to . Thus near a noncritical point of the mapping is represented as with respect to natural parameters and of and around and , respectively, where . Let be a Teichmüller quasiconformal homeomorphism of onto a closed Riemann surface for which , where , and set , which is a conformal embedding of into . If denotes the terminal quadratic differential of corresponding to so that is expressed as with a natural parameter of around , then is given by near , which implies that . Since consists of finitely many horizontal arcs of together with finitely many points, from Corollary 4.7 we infer that is a closed -regular self-welding continuation of . Finally, is a Teichmüller quasiconformal homeomorphism of onto with , and is an initial quadratic differential of . ∎
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.
Corollary 8.4.
Let be a nonanalytically finite open Riemann surface of genus . Then initial quadratic differentials of Teichmüller quasiconformal embeddings of into closed Riemann surfaces of genus belong to .
Let be a marked finite open Riemann surface, and let . A Teichmüller quasiconformal embedding of into is, by definition, an element for which is a Teichmüller quasiconformal embedding for some , and . This definition does depend on a particular choice of representatives of , and . Any initial and terminal quadratic differentials and of determine well-defined elements and , which will be referred to as initial and terminal quadratic differentials of , respectively.
Lemma 8.5.
Let be a marked nonanalytically finite open Riemann surface of genus , and let . If is a Teichmüller quasiconformal embedding of into with initial quadratic differential , then there are a closed -regular self-welding continuation of and a Teichmüller quasiconformal homeomorphism of onto such that
-
(i)
the co-welder of is an initial quadratic differential of , and
-
(ii)
.
Moreover, belongs to .
If , then even though . A homotopically consistent quasiconformal embedding of into is said to be extremal if it has the smallest maximal dilatation in .
Proposition 8.6 (Ioffe [25, Theorem 0.1], Bourque [12, Theorem 3.11]).
Assume that is a marked finite open Riemann surface of genus . If belongs to , then has an extremal element, which is a Teichmüller quasiconformal embedding.
Remark.
Proof of Proposition 8.1.
Let . Take a sequence in converging to , and choose an extremal for each . It is a Teichmüller quasiconformal embedding by Proposition 8.6. By Lemma 8.5 there are a closed regular self-welding continuation of and a Teichmüller quasiconformal homeomorphism of onto such that . Since , we can find . If is the Teichmüller quasiconformal homeomorphism of onto , then . The fact that is extremal yields
as . Therefore, Proposition 7.7 guarantees the existence of a closed regular self-welding continuation of , as desired. ∎
In general, let be a marked nonanalytically finite open Riemann surface of genus . Two elements and of are said to be projectively equivalent to each other if for some . The set of projective equivalence classes is denoted by . Now, Theorem 1.4 (i) asserts that any point on the boundary is obtained as a closed -regular self-welding continuation of for some and that each gives rise to a point on through a closed -regular self-welding continuation of . Recall that projectively equivalent elements of induce the same closed regular self-welding continuation of (see the paragraph precedent to Example 5.5). Examples 5.4 and 5.5 show that the correspondence is, in general, many-to-many.
9 Ioffe rays
Let be a marked open Riemann surface of genus . For we denote by the set of into which can be mapped by a homotopically consistent -quasiconformal embedding. In particular, we have since -quasiconformal embeddings are conformal embeddings. If , then . Also, we have . The next proposition claims that is the closed -neighborhood of .
Proposition 9.1.
Let and . Then if and only if
(9.1) |
Proof.
Let , and suppose that there is with , where . Take a quasiconformal homeomorphism of onto a closed Riemann surface such that on and on . Setting and , we obtain a closed continuation of . Thus and hence
For and let denote the open ball of radius centered at :
Its closure will be denoted by , which is a compact subset of . For the sake of convenience we define .
Corollary 9.2.
is compact for all .
Proof.
Definition 9.3 (Ioffe ray).
Let be a marked nonanalytically finite open Riemann surface. An Ioffe ray of is, by definition, a Teichmüller geodesic ray of the form , where and such that for some and the continuation is a closed -regular self-welding continuation of and that is the co-welder of . Let denote the set of Ioffe rays of .
If is a marked analytically finite open Riemann surface, then consists of exactly one point, say . We call each Teichmüller geodesic ray emanating from an Ioffe ray of .
Remark.
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 be a marked finite open Riemann surface of genus . Then the correspondence
is bijective.
Biernacki [7] called a plane domain linearly accessible in the strict sense if the compliment 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 is linearly accessible in the strict sense or close-to-convex.
Theorem 9.4 implies in particular that each Ioffe ray never hits 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 be a marked nonanalytically finite open Riemann surface of genus , and let . Take so that is obtained via a closed -regular self-welding continuation of and let be the co-welder of . Then any element of defines a closed -regular self-welding continuation of for which is the co-welder of .
Proof.
Let be a closed -regular self-welding continuation of with . For any closed continuation of let be the -zero-extension of . It then follows from Lemma 6.5 that for any the inequality holds. Proposition 6.4 then implies that . Actually, the sign of equality occurs because . Another application of Proposition 6.4 gives us almost everywhere on . In particular, is a dense continuation of . Since , Proposition 6.1 implies that is a closed -regular self-welding continuation of for which is the co-welder of . ∎
Remark.
Proposition 9.6.
Suppose that is a marked finite open Riemann surface of genus . Let . If is a point of nearest to , then is an Ioffe ray of .
Proof.
We may assume that is nonanalytically finite. Take . Let be the Teichmüller quasiconformal homeomorphism of onto , and set . If , then Proposition 9.1 yields that
This shows that is extremal in , and hence is a Teichmüller quasiconformal embedding by Proposition 8.6.
Let and be initial and terminal quadratic differentials of , respectively (see Corollary 8.4). Since is a boundary point of , Proposition 9.5 implies that is a regular self-welding continuation of . To show that is a welder of the continuation take representatives , , , , , , and so that . The last identity shows that an initial quadratic differential of satisfies . Thus the meromorphic function on is real and hence constant. Since is a uniform stretch along horizontal trajectories of both and , the constant must be positive and hence may be supposed to be . Then we obtain . As is dense in , in view of Proposition 6.1 we see that is a closed -regular self-welding continuation of and that is the co-welder of . It follows that is an Ioffe ray of . Since is an initial quadratic differential of the Teichmüller quasiconformal homeomorphism of onto , we infer that . ∎
Lemma 9.7.
Let be a compact subset of and let be a Teichmüller geodesic ray with . If for all , then for .
Proof.
It follows from (6.2) that belongs to . Suppose that also belongs to both and . Then
(9.2) |
and hence . Since , we obtain , or equivalently, . Therefore, the signs of equality occur in (9.2). Since the metric space is a straight space in the sense of Busemann (see Abikoff [1, Section (3.2)]), the three points , and lie on the same Teichmüller geodesic line, which must include the Teichmüller geodesic ray . It then follows from (9.2) that , which finishes the proof. ∎
Theorem 9.8.
Let be a marked finite open Riemann surface. If , then
(9.3) |
and
(9.4) |
for all , and .
To prove Theorem 9.8 we need the following lemma, which implies that is a Lipschitz continuous function on . For the proof see Gardiner [16, Lemma 4] or Gardiner-Lakic [18, Lemma 12.5].
Lemma 9.9.
For and ,
Proof of Theorem 9.8.
We can write with , where is the co-welder of a welder of some closed regular self-welding continuation of . Set .
To prove (9.3) let . If is a point of nearest to , then Lemma 9.9 and Theorem 6.3 imply that
Consider the Teichmüller geodesic ray defined by , . Then another application of Lemma 9.9 together with (6.4) yields that
With the aid of Proposition 9.1 we thus obtain
Therefore, for all . Since is compact by Corollary 9.2 and contains , identity (9.3) follows at once from Lemma 9.7.
Corollary 9.10.
If and , then .
Proof.
By (9.4) we have . Hence . ∎
Corollary 9.11.
Let and . Then a point of nearest to is uniquely determined.
Proof.
We are now ready to prove Theorem 9.4
The following proposition describes the boundary and the exterior of in terms of Ioffe rays.
Proposition 9.12.
If is a marked finite open Riemann surface of genus , then
for .
Proof.
We conclude the section with the following proposition due to Bourque. Recall that includes extremal elements for any , 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 be a marked finite open Riemann surface of genus . Then for Teichmüller quasiconformal embeddings of into are extremal in and have initial and terminal quadratic differentials in common.
Proof.
Let , and take a Teichmüller quasiconformal embedding of into with initial quadratic differential . If , then the proposition follows from Proposition 9.5.
If , then by Lemma 8.5 we obtain a closed -regular self-welding continuation of and a Teichmüller quasiconformal homeomorphism of onto such that the co-welder of is an initial quadratic differential of and that . Thus lies on the Ioffe ray , and if , then by (9.4), and hence is extremal by Proposition 9.12. Since is uniquely determined by Corollary 9.11, it follows that is also uniquely determined. Since and 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 . In Example 5.4 we have remarked that some positive holomorphic quadratic differential on satisfying the border length condition on can induce two or more inequivalent closed regular self-weldings of . With this in mind we make the following definition.
Definition 10.1 (exceptional quadratic differential).
Let be a union of connected components of , and let . If there are two -full -regular self-weldings of that are inequivalent to each other, then is called exceptional for .
It is then natural to ask which elements of are exceptional for . To answer the question we introduce a subset of as follows.
Definition 10.2 (class ).
Let be a union of connected components of . Denote by the set of such that some component of includes four or more horizontal trajectories of and that
(10.1) |
for all horizontal trajectories of on . Set .
Clearly, , where the union is taken over all components of . The following theorem asserts in particular that for there is exactly one closed -regular self-welding of up to equivalence.
Theorem 10.3.
Let be a compact bordered Riemann surface, and let be a union of components of . Then an element of is exceptional for if and only if it belongs to . If this is the case, then the family of equivalence classes of -full -regular self-weldings of has the cardinality of the continuum.
Proof.
We may suppose from the outset that is a connected component of . Assume first that for some horizontal trajectory of on . The endpoints of , which are zeros of , divide into two arcs and of the same -length, where we label them so that . Let be a -full -regular self-welding of , and let denote the co-welder of . Since does not vanish on , it follows from Proposition 5.6 that is a simple arc on the weld graph . As the sum of the -lengths of edges of is identical with , we deduce that exhausts , or is I-shaped. Therefore, is equivalent to the self-welding of with welder along . Hence is nonexceptional for .
In the rest of the proof we suppose that inequalities (10.1) with replaced with hold for all horizontal trajectories of on . Then carries at least three horizontal trajectories of and hence the set of zeros of on contains more than two points.
Assume that so that , say. The points of divide into three arcs , and , where they are labeled so that for . Let be a -full -regular self-welding of , and let be its weld graph. Since contains no zeros of and the self-welding is -regular, it follows from (10.1) that the points , , are the end-vertices of by Proposition 5.6, and hence is Y-shaped. Lemma 5.11 assures us that is determined up to equivalence. This means that is nonexceptional for .
Suppose next that . We consider two cases. If contains two points and that divide into two arcs and of the same -length , then the self-welding of with welder along is -full and -regular, and its weld graph is I-shaped. On the other hand, inequality (10.1) implies that each contains a point of . Take short subarcs and of emanating from so that , and let be the welding of with welder along , . The two points , , divide the component of containing into two arcs and of the same -length, where is the co-welder of . If denotes the self-welding of with welder along , then is a -full -regular self-welding of . Since its weld graph has four end-vertices, the self-welding is not equivalent to . Thus is exceptional for .
Finally, assume that no two points of divide into arcs of the same -length, where . Then we can choose three points , , of to divide into three arcs , , such that inequality (10.1) holds for , . We then apply Lemma 5.11 to obtain a -full -regular self-welding of whose weld graph is Y-shaped. Let be a point of other than , . Take two simple arcs and of the same -length on emanating from . We choose them so that neither nor contains any of , . Let be the self-welding of with welder along , and let be the co-welder of . If the -length of is sufficiently small, then the three points , , divide the component of containing these points into three arcs with -length less than . Another application of Lemma 5.11 yields a -full -regular self-welding of whose weld graph is Y-shaped. Then is a -full -regular self-welding of . It is not equivalent to because the weld graph of is not Y-shaped. Therefore, is exceptional for .
The last assertion of the theorem is clear form our constructions of inequivalent -full -regular self-weldings of . The proof is complete. ∎
The following is a corollary to the proof of the above theorem.
Corollary 10.4.
Let be a compact bordered Riemann surface, and let be a component of . Let .
-
(i)
For any closed -regular self-welding of the image contains a zero of the co-welder of .
-
(ii)
For some closed -regular self-welding of the image is neither I-shaped nor Y-shaped.
Example 10.5.
Let 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 . Thus every element of yields exactly one closed -regular self-welding of .
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 be a marked compact bordered Riemann surface of positive genus. An element of is called exceptional if . Let stand for the set of exceptional elements of .
Let be a marked nonanalytically finite open Riemann surface of positive genus. Set , where denotes the natural compact continuation of . Elements of are called exceptional.
Example 10.6.
We consider the case of genus one. We know that is empty (see Example 10.5). Thus we have a well-defined mapping of onto . To see that it is injective take , , and suppose that they induce homotopically consistent conformal embeddings of into the same marked torus . It then follows from Proposition 9.5 that for some . Since is one-dimensional, there is a nonzero complex number such that and hence . As , , we conclude that so that and represent the same element of . We have shown that the correspondence between onto is bijective. In particular, since , we deduce that for each there exists exactly one homotopically consistent conformal embedding of into . We have thus given alternative proofs of [56, Theorems 4 and 5].
If we set
we see that is exceptional if and only if . Also, set
Theorem 10.7.
Let be a marked nonanalytically finite open Riemann surface of positive genus. If , then some element in cannot be approximated by any sequence in .
Proof.
Let and . Take a natural compact continuation of together with for which . By Corollary 10.4 (ii) the weld graph of some closed -regular self-welding of has a component that is not I-shaped or Y-shaped. On the other hand, if , then each component of the weld graph of any closed -regular self-welding of is I-shaped or Y-shaped by Theorem 10.3 and Lemma 5.11. Consequently, no sequences in converge to , where . ∎
As an application of the theorem, we prove Theorem 1.4 (iii). To this end we prepare two lemmas. Let be a closed Riemann surface of positive genus. For a nontrivial complex vector space of holomorphic -forms on set
for , where denotes the order of at . If , where , then is called a Weierstrass point for . Otherwise, is said to be a non-Weierstrass point.
Lemma 10.8.
Let be a nontrivial complex vector space of holomorphic -forms on a closed Riemann surface of genus . Then the set of Weierstrass points for is nonempty and finite.
Proof.
Let , , be a basis of . If is the Wronskian of , that is, , then is a holomorphic -differential on . Let and take a local coordinate around with . If and , then the power series expansion of is of the form with . In particular, is nonzero. Moreover, is a Weierstrass point for if and only if is a zero of . This proves the lemma. ∎
Lemma 10.9.
Let be a compact bordered Riemann surface of genus , and let denote its double. Then there is a holomorphic -form on such that
-
(i)
along ,
-
(ii)
for each component of , and
-
(iii)
has a zero on and four zeros on some component of .
Proof.
Let be the vector space of holomorphic -forms on satisfying condition (ii). Then ; note that for all holomorphic -forms on . Also, let be the real vector space of possessing property (i). Let be the anti-conformal involution of fixing pointwise. Then is an -linear isomorphism of onto itself and belongs to for .
Take a non-Weierstrass point for , and let be the set of -forms in vanishing at . Then . If denotes the set of -forms in vanishing at , then . Note that for any the point is a zero of .
Since , we have . Fix a component of . Let be a non-Weierstrass point for , and set for . Then . Take a non-Weierstrass point for both and , and set for . Note that and that . Finally, choose a non-Weierstrass point for , , and , and define for . Then for , and . Consequently, , , for some nonzero . Set , where . Then satisfies (i) and (ii). We can choose so that has simple zeros at , . Since along and , the -form has one more zero of odd order on . Hence satisfies (i), (ii) and (iii). ∎
We are now ready to prove Theorem 1.4 (iii).
Proof of Theorem 1.4 (iii).
Let , where and , be the natural compact continuation of . Let be the double of . We apply Lemma 10.9 to obtain a holomorphic -form on that has a zero on and four zeros on a component of . Set . Then, since by Theorem 10.3, Theorem 10.7 implies that there is a closed -regular self-welding continuation of such that there are no sequences of closed regular self-welding continuations of for which is a sequence in converging to .
We claim that is an interior point in . If not, then some sequence in would converge to . Take closed regular self-welding continuations of . By Proposition 7.7 there is a closed regular self-welding continuation of such that a subsequence of converges to . Since has a zero on , it follows from Bourque [12, Remark 4.3] that , which is impossible by the choice of . This completes the proof. ∎
Remark.
Let be as in the above proof. Then is a singleton for any . Since , we know that .
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 be a one-parameter family of marked open Riemann surfaces of genus . We say that shrinks continuously if
-
(i)
for , and
-
(ii)
for each and there is such that is included in the -neighborhood of , where and .
If, in addition, is a singleton, then is said to shrink continuously to a point.
Lemma 11.1.
Let , , be marked open Riemann surfaces of genus . If there is a homotopically consistent -quasiconformal embedding of into , then
Proof.
Let with . If is a closed continuation of , then with . Hence belongs to . ∎
Proposition 11.2.
Let be a one-parameter family of marked finite open Riemann surfaces of genus . If shrinks continuously, then
for .
Proof.
Fix . Since for , we know that
To show the converse inclusion relation take an arbitrary interior point of . Then for some the neighborhood of is included in . We can choose sufficiently near to so that . We claim that is an interior point of . If not, then would lie on an Ioffe ray of with , where by Proposition 9.1. Since
we have and hence . This contradicts the identity , and the proof is complete. ∎
Let be open and nonanalytically finite. We construct two one-parameter families of continuations of . In the following examples denotes the natural compact continuation of , where and .
In general, for let denote the open disk in of radius centered at . Its closure is denoted by . Set for convenience. We abbreviate and to and , respectively.
Example 11.3 (circular filling).
Let be the connected components of the border . Take doubly connected domains , , on such that is a component of and that is mapped onto a fixed annulus by a conformal homeomorphism with on , where . We attach to each by identifying with to obtain a closed Riemann surface of genus . We consider as a closed subdomain of , and denote by the component of with . Then is extended to a conformal mapping of onto . Considering as a -handle mark of as well, set . Then .
For each we construct a subsurface of homeomorphic to as follows:
Define for . Regarding as a homotopically consistent conformal embedding of into , we call a circular filling for . Note that is a finite set.
Example 11.4 (linear filling).
As in Example 5.2, take , and divide each border component into two subarcs and of the same -length . Let be the closed self-welding of with welder along , . and let be the co-welder of . The arcs and are projected to a simple arc on , which is parametrized with respect to -length. Set
for . Let denote the conformal embedding of into induced by . The continuation of yields a continuation of . We call a linear filling for . Again, is a finite set.
Fillings introduced in the above examples provide us with continuity methods:
Proposition 11.5.
Let be a marked nonanalytically finite open Riemann surface of genus . If is a circular or linear filling for , then shrinks continuously to a point.
Proof.
It is clear that is analytically finite and hence is a singleton, say, . If , then , and hence by Lemma 11.1.
To show that shrinks continuously take and . By Proposition 9.1 we are required to find such that , where and . For we can choose a desired with the aid of Lemma 11.1. To verify the existence of for we have only to show that for all sufficiently near . If there were no such , then we could find a point in as would be a family of compact sets with finite intersection property. Choose and for each so that converges to as . We apply Lemma 7.4 to obtain a sequence converging to for which the sequence converges to a homotopically consistent conformal embedding of into . Then , or , 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 , then .
Proof.
We use the notations in Example 11.4. Thus is a linear filling for . For each take a circular filling for . Let . Since shrinks continuously by Proposition 11.5, it follows from Proposition 11.2 that for some . Another application of Propositions 11.2 and 11.5 assures us that for each there is such that . Choose and set to obtain continuations , , of . Observe that through , where , each pair of arcs and on yields a simple arc on while gives rise to a simple loop on . Consequently, the continuations , , of are distinct from one another. This completes the proof. ∎
Remark.
We conclude this section by giving one more application of circular fillings. We apply it in §14.
Proposition 11.7.
If is a circular filling for , then
for .
12 Maximal sets for measured foliations
Let be a marked finite open Riemann surface of genus . In this section we give a homeomorphism of onto explicitly to establish Theorem 1.5. Then we prove Theorem 1.6.
Definition 12.1 (maximal set).
Let , and let be a compact subset of . A point of where the restriction of to attains its maximum is referred to as a maximal point for on . We call the set of maximal points for on the maximal set for on .
Since is compact by Corollary 7.6, we can speak of maximal points and sets for on . Denote by the maximal set for on . Then we have the following proposition.
Proposition 12.2.
Let be finite and open, and let . Then any point lies on , and is an Ioffe ray of , where . Moreover, the pull-back of by is determined solely by and does not depend on or .
Proof.
Fix a maximal point for on , and set , where . If with , then Lemma 9.9 together with (6.4) shows that
which implies that , for, is a maximal point for on . We have proved that for all . We then apply Lemma 9.7 to obtain . Hence is a boundary point of . Furthermore, is the nearest point of to . Since , Proposition 9.6 shows that is an Ioffe ray emanating from .
To prove the last assertion of the theorem we may assume that is nonanalytically finite. Since is an Ioffe ray of , for some there is a -regular self-welding continuation of such that is the co-welder of . Let be an arbitrary maximal point for on , and choose arbitrarily. Denote by the -zero-extension of . By Lemma 6.5 and Proposition 6.4 we obtain , where . Actually, the sign of equality occurs because
Another application of Proposition 6.4 gives us almost everywhere on , which implies that . This completes the proof. ∎
Let . If is nonanalytically finite, then is a welder of some regular self-welding continuation of by Theorem 1.3. Let be the co-welder of , and set
The measured foliation on does not depend on a particular choice of the self-welding , for, consists of finitely many horizontal arcs of together with finitely many points and contributes nothing for evaluating the -heights of curves on . We have thus obtained a mapping of into , which is bijective by Proposition 12.2. We denote its inverse by .
If is analytically finite, then it has a unique closed continuation , and induces a bijection of onto . We then define , which is a homeomorphism of onto by Proposition 6.2. Again, we set .
Proof of Theorem 1.5.
We are required to show that and are continuous. To prove that is continuous let be the double of , where stands for the natural compact continuation of . If is the number of components of , then is a closed Riemann surface of genus . We consider as a subdomain of , and denote by the inclusion mapping of into . Then is a conformal embedding of into . Fix to obtain a marked closed Riemann surface of genus .
Let , and take in so that . The reflection principle enables us to extend to a holomorphic quadratic differential on . Set and , where . Let . Choose a simple loop in the homotopy class so that stays in , and denote by the free homotopy class in for which .
We then claim that . To verify the claim take a point in , and construct a covering Riemann surface of corresponding to the subgroup of the fundamental group . Let denote the holomorphic covering map. For some component of we have . Observe that each component of is a doubly connected planar domain on . Therefore, if is a loop in with and leaves from across a component of the border , that is, and for some with , then returns to across the same component in such a way that the subarc of between the leaving point and the returning point is homotopic to an arc on through a homotopy fixing the endpoints, or more precisely, there is together with such that , for and , , for . In other words, the loop defined by for and otherwise is freely homotopic to on . Because are positive along , we obtain , which proves the claim.
Now, let be a sequence in converging to , and set and . The quadratic differentials and induce quadratic differentials and on and measured foliations and on . Since converges to , it follows from Proposition 6.2 that as for , which means that converges to . We have proved that is continuous.
Next, we show that is also continuous. Fix . Since is compact by Corollary 7.6, there is such that for all . It thus follows from Lemma 9.9 that
for and . Let be a sequence in converging to , and set . Let be a closed -regular self-welding continuation of . Since Proposition 6.2 implies that
as , the sequence is bounded and bounded away from . With the aid of Proposition 4.1 together with normal family arguments we infer that any subsequence of contains a subsequence converging to some . Since is continuous, we obtain , or . Consequently, as , proving that is continuous. This completes the proof. ∎
Proof of Theorem 1.6.
We begin with the proof of (i). Let , and set . Equality (1.1) follows from Proposition 12.2. Any point of is a maximal point for on by Theorem 6.3, and such points exhaust the maximal set for on by Proposition 12.2. This proves (i).
Example 12.3.
We consider the case of genus one. We identify with as in Example 3.14, and use the notations in Example 6.7. For set , and define , where . Then the level curves of the function are the horocycles in with center at , that is, the circles in tangent to at . Theorem 1.6 (iii) implies that is the envelope of the family of horocycles , . Such a description of first appeared in [36, §5]. Since is a closed disk, three of the horocycles, or equivalently, three of the measured foliations are sufficient to determine it; see [35, Theorem 1], where is given by the extremal length of a weak homology class of .
13 Shape of with
Let be a marked finite open Riemann surface of positive genus . The aim of the present section is to prove Theorem 1.2.
For let denote the point of nearest to . By Corollary 9.11 this gives a well-defined mapping of onto . It fixes pointwise and maps into . We abbreviate to .
Proposition 13.1.
If is a marked finite open Riemann surface of positive genus , then is a retraction.
Proof.
Since fixes pointwise, we have only to show that it is continuous on . If , then there is an Ioffe ray of such that for some ; note that by (9.4). Then by (9.3) and Proposition 9.12.
Let be a sequence in converging to . Take and so that . Then . The sequence converges to , for,
Since is compact, a subsequence converges to some point on . Then we obtain
by (9.4). Hence is the point of nearest to , or, . This completes the proof. ∎
For let denote the restriction of to . We consider it as a mapping of into .
Corollary 13.2.
If is a marked finite open Riemann surface of genus , then for .
Proof.
Theorem 9.4 implies that is a bijection. Since is compact and is Hausdorff, the continuous mapping is in fact a homeomorphism of onto . ∎
Remark.
The continuous mapping is not necessarily injective (see the remark after Definition 9.3) though it is always surjective.
Proposition 13.3.
If is a marked finite open Riemann surface of genus , then for each there is a homeomorphism of onto that maps the compact sets , , onto concentric closed balls, where .
The proof of the proposition requires two lemmas. For the proof of the first lemma we need the following result of Earle. For with let denote the element of with such that .
Proposition 13.4 (Earle [13]).
The Teichmüller distance function is of class on off the diagonal. The differential of the function is .
Thus if and represent and , respectively, then the differential of at is the linear functional
on the space of bounded measurable -forms on . Note that the inequality holds and that the sign of equality occurs if . Since the tangent vector to at the initial point is represented by , we see that the tangent vector varies continuously with and .
We now turn to our first lemma. Recall that the image of the interval by a Teichmüller geodesic ray is denoted by (see (6.3)).
Lemma 13.5.
Let and be distinct points of . Then for there are neighborhoods and of and , respectively, such that for any and the inclusion relation
(13.1) |
holds, where , , and .
Proof.
Set and . Since the derivative of the function
is on the interval , Proposition 13.4 implies that there are neighborhoods and of and , respectively, together with such that if and , then the derivative of the function
is negative on the interval , where , , and . Since , it follows that
for , or equivalently, that
(13.2) |
Since
and the Teichmüller distance between and is exactly , we can replace the neighborhoods and with smaller ones to ensure that
(13.3) |
for and . Now inclusion relation (13.1) is an immediate consequence of (13.2) and (13.3). ∎
Lemma 13.6.
Let and be disjoint compact sets in , and let . Then there is such that for any , and the inclusion relation
(13.4) |
holds, where , , and .
Proof.
For each we choose a neighborhood of as follows. For take neighborhoods of and of as in Lemma 13.5. As is compact, there are finitely many points such that the corresponding open sets cover . Then we set . It is a neighborhood of such that (13.1) holds for and . Then any Lebesgue number of the open covering of possesses the required properties. ∎
We are now ready to prove Proposition 13.3.
Proof of Proposition 13.3.
If is analytically finite, then for some so that . Hence the proposition is valid in this case.
Assume now that is nonanalytically finite. Let be a circular filling for . Then is a singleton, say, . Fix satisfying
(13.5) |
and set . It is a compact subset of that does not meet . By Lemma 13.6 there is such that for any , and inclusion relation (13.4) holds.
We claim that for every there exists such that if and for , then
for all ; recall that stands for the point of nearest to . If the claim were false, then there would exist sequences and such that
where and . By taking subsequences if necessary we may assume that the sequences , and converge to , and in , respectively. Then and with
(13.6) |
For any we have
for sufficiently large . Set and . Since
we obtain , or equivalently,
As is arbitrary, we know that . Since belongs to , we conclude that . Similarly, we obtain , and hence , contradicting (13.6).
Let be a Lebesgue number of the open covering of . Take an integer and set for . Note that
(13.7) |
For and denote by the Ioffe ray of passing through , and set . Thus and . We then claim that
(13.8) |
for and . To show the claim note first that inequality
(13.9) |
holds for due to (13.7). As and
we have by (13.9). Thus
(13.10) |
for . Since , there exists such that for . Hence
which implies that
Now, we inductively show that for there is a homeomorphism of onto that maps , , onto concentric closed balls. This is true for since .
To complete the induction arguments suppose that the assertion is true for some . Thus there is a homeomorphism of onto that maps onto a -dimensional euclidean closed ball . Define a mapping by . It is surjective by Theorem 9.4. To show that it is injective suppose that for some . By the definition of the point lies on the Ioffe ray of passing through , that is, on . Thus and hence by (13.8). Since by (13.10), we see that . Interchanging the roles of and we conclude that and hence that , which proves that is injective.
The bijection is identical with on and hence continuous by Proposition 13.1 and Corollary 13.2. It is homeomorphic since is compact and is Hausdorff. Thus is homeomorphic to the sphere .
We extend to a homeomorphism of onto itself step by step. We begin with extending it to a homeomorphism of onto as follows. Each point of is of the form , where and . Then define , where
note that belongs to by (13.8). It is easy to extend to a homeomorphism of onto because and are homeomorphic to the closed ball .
Finally, for take and so that . We then define to extend to a homeomorphism of onto itself with for .
With the aid of the induction hypothesis we can construct from a homeomorphism of onto that maps , , onto concentric closed balls. Corollary 13.2 now guarantees the existence of desired homeomorphisms for , . ∎
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 is a marked finite open Riemann surface of genus , then the boundary is a -submanifold of for .
Proof.
Let be an arbitrary point of , and take the Ioffe ray of passing through . Set . Thus . Note that by (9.3) and that by Proposition 9.1. Also, as is a straight space in the sense of Busemann, is the unique common point of the closed balls and .
Take a coordinate system centered at . Thus is a connected open neighborhood of , and is a diffeomorphism of onto a domain of with . Write . We choose the system so that and for some and some domain in . Also, we require that the points with (resp. ) should be contained in (resp. ). Thus for each there is such that . We show that is uniquely determined if is sufficiently near . Observe that even if is not unique, it tends to as because . Since the partial derivatives of and at with respect to are and , respectively, we may assume that the partial derivative of with respect to is negative on for any in a neighborhood of and that the partial derivative of with respect to is positive on for any in a neighborhood of . Let denote the Ioffe ray of passing through . If is sufficiently near so that and , then for with and , and hence . Also, we have for with and , and hence . Consequently, is uniquely determined for sufficiently near . Replacing with a smaller one if necessary, we assume that is uniquely determined for .
We have shown that there is a function such that . Since is of class off the diagonal by Proposition 13.4, the hypersurfaces and have a common tangent hyperplane at . Since the graph of lies between these hypersurfaces, we infer that the hyperplane is also tangent to the graph. Consequently, is differentiable at . Since the hypersurfaces together with the common tangent hyperplanes move continuously with , we know that is of class . The proof is complete. ∎
14 Geometric properties of
The final section is devoted to establishing Theorem 1.1. Assuming that is nonanalytically finite, fix a circular filling for . We use the same notations as in Example 11.3. Each is the interior of a compact bordered Riemann surface of genus and includes for . If is sufficiently small, then every quadratic differential in is extended to a holomorphic quadratic differential on , denoted again by . Thus we consider as a subset of . Let be the set of quadratic differentials in with .
Lemma 14.1.
There exists such that each has a noncritical point together with a natural parameter of around satisfying for .
Proof.
For choose a noncritical point . There is a neighborhood of in such that , , are bounded and bounded away from ; note that is a meromorphic function on . Then there are a positive number and a neighborhood of such that a natural parameter of around satisfies for . Since is compact, the open covering of includes a finite subcovering . Then possesses the required properties. ∎
Lemma 14.2.
Let be a sequence of positive numbers converging to . If a sequence in converges to , then as .
Proof.
Define , and , and set , and . Note that and and that as .
For any we have and for sufficiently large . Then and hence
This proves that as .
We show that is a unique accumulation point of . Let be an arbitrary convergent subsequence of . If denotes its limit, then it belongs to . As
we know that is the nearest point of to and hence . This completes the proof. ∎
Lemma 14.3.
There exist positive numbers and such that
for all and .
Proof.
Let with , and take for which . Note that . Take a positive number as in Lemma 14.1; there is a noncritical point of together with a natural parameter , , of around such that the closed rectangle defined by and is included in for . Note that points in with belong to . In the following we identify with in the obvious manner.
We divide into three parts , and as follows:
where . Define a function on so that for the function assumes the value on the polygonal arc obtained by joining , , and successively. In other words, we define the function by
For define a quadratic differential on by
Since
as , there is a positive number such that
for all .
Let . It is induced by a -regular self-welding continuation of by Theorem 1.6 (i). Let be the co-welder of . Thus and . For let , and set . If , then induces a quadratic differential on . Let denote its zero-extension to . Similarly, induces a holomorphic differential on . Denote by its zero-extension to . Since on , for we have for any piecewise analytic simple loop on with . Hence , where and . Since
it follows from Proposition 6.4 that
Since is a maximal point for on , we obtain
for all , provided that . This proves the lemma. ∎
In general, let be a compact set of . For denote by the set of measured foliations on such that is a maximal point for on . For nonempty set .
Corollary 14.4.
There exist positive numbers and such that for all and the inequality
holds for all .
Proof.
Note that by Proposition 11.7. Since there exists a homotopically consistent -quasiconformal homeomorphism of onto (for the definition of see Example 11.3), Lemma 11.1 and Proposition 9.1 imply that there is such that
for all and .
Let . If belongs to , then
where the last inequality follows from the fact that . Therefore, the corollary is an immediate consequence of Lemma 14.3. ∎
In general, let . For define
where and . We need the following variational formula due to Gardiner, which implies that the extremal length function is continuously differentiable on . For the proof see also Gardiner-Lakic [18, Theorem 12.5].
Proposition 14.5 (Gardiner [16, Theorem 8]).
Let and and set . Then for
as .
In fact, the differential of at is the linear functional
on the space of bounded measurable -forms on , where and . The inequality holds and the sign of equality occurs if .
For and we denote by the Ioffe ray of passing through . Thus its initial point coincides with . Also, set so that .
Lemma 14.6.
For each there exist a neighborhood of and positive numbers and such that for , and the inequality
holds whenever .
Proof.
Let . If there were no required , or , then there would exist a sequence of neighborhoods of shrinking to and sequences and of positive numbers converging to such that
(14.1) |
for some , some with
and some measured foliation in with .
Set and . Taking a subsequence if necessary, we may suppose that and converge to and , respectively. Note that as . Since for all , by letting we obtain . In particular, is a maximal point for on .
As lies on the Teichmüller geodesic segment connecting and , we have
Lemma 14.2 yields that as . Taking limits of the both sides of the above inequality, we obtain
which means that is a point of nearest to . Therefore, is identical with by Corollary 9.11 and hence is an Ioffe ray of .
Proposition 14.7.
If is nonanalytically finite, then there is a homeomorphism of onto that maps onto a closed ball.
Proof.
Since is compact, an application of Lemma 14.6 gives positive numbers and such that for , and the inequality
(14.3) |
holds whenever . Fix with so that . By Proposition 11.7 there is such that . Inequality (14.3) implies that the Teichmüller geodesic arc lies outside of since is a maximal point for on . Note that as is included in . We thus deduce that the correspondence
is a continuous bijection, and hence is a homeomorphism as is compact and is Hausdorff. Observe that is the disjoint union of Teichmüller geodesic arcs , , and that . Since is continuous, there is a homeomorphism of onto itself for which and , . With the aid of Theorem 1.2 we can extend it to a homeomorphism of onto itself with , completing the proof. ∎
We have proved most assertions of Theorem 1.1. The following proposition will finish the proof.
Proposition 14.8.
If is nonanalytically finite, then is a closed Lipschitz domain satisfying an outer ball condition.
Proof.
The Teichmüller distance function is of class on off the diagonal by Rees [47]. Hence Theorem 9.8 shows that satisfies an outer ball condition.
We verify that is locally expressed as the graph of a Lipschitz function. For let denote the set of such that . For set .
Now, take arbitrarily. Choose a sequence of positive numbers with so that converges to a point . Then is an Ioffe ray of by Lemma 14.2. If , then Proposition 14.5 and Corollary 14.4 imply
for some positive and independent of , where as . Letting , we obtain
Set . Since is an Ioffe ray of , we know that ; note that . The above inequality can be expressed as
(14.4) |
for .
We examine the behavior of the function along . Proposition 14.5 shows that
In general for , , and define
We claim that there is a neighborhood of together with positive numbers and such that
(14.5) |
for all , , . If not, then there would exist sequences , , and such that
-
(i)
as ,
-
(ii)
there are and with such that and as ,
-
(iii)
with , and as ,
-
(iv)
and as , and
-
(v)
.
Letting in the last inequality, we obtain , contradicting (14.4).
Owing to Theorem 1.6 (iii) and Proposition 14.7 we can replace with a smaller one so that is the set of such that (1.2) holds for all . Since is a function on with nonvanishing derivatives, its level hypersurfaces are submanifolds of . Replacing with a smaller one if necessary, we take a coordinate system centered at so that if we write , then the level hypersurfaces are represented as and the Teichmüller geodesic rays (resp. ) with are represented as , (resp. ). Note that if is in a small neighborhood of and and are sufficiently small, then is included in and contains a cone with vertex at and axis parallel to the -axis, where the cones can be chosen to be of the same shape regardless of .
Suppose that . From (14.5) we infer that for the function is decreasing along the Teichmüller geodesic segment , , for and hence that
Thus includes . In particular, includes .
Take a neighborhood of included in so that if , then the Teichmüller geodesic ray hits . Set , where is the minimum of nonnegative for which . Then is a mapping of into , and is included in . Consequently, the function , , is a Lipschitz function on a neighborhood of in the plane , and is the graph of the function. This completes the proof. ∎
Theorem 14.9.
Suppose that is nonanalytically finite. Let be a boundary point of . If there are two Ioffe rays of emanating from , then is not smooth at .
Proof.
By assumption there are linearly independent , , such that . Let be the set of for which , where . Then is included in by Theorem 6.3. Since and meet transversally at , the boundary cannot be smooth at . ∎
See Example 5.5 for the existence of satisfying the assumptions of the theorem. In the case of genus one is a closed ball with respect to the Teichmüller distance provided that it is not a singleton. This is not always the case for 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 every element of has a dense image.
Proposition 14.10.
Suppose that is open and nonanalytically finite. If , then there are continuations , , of such that is of positive area while has a vanishing area, where .
Proof.
Fix . If is a circular filling for , then belongs to for some positive by Proposition 11.2. Thus there is a homotopically consistent conformal embedding of into . Then is a homotopically consistent conformal embedding of into and is not a dense continuation of .
Next let be a linear filling for . If , then induces a required element of . Otherwise, for some . Let be the infimum of such . For each let denote the Ioffe ray of passing through . Then as . On the other hand, if we take a sequence in so that and as , then Lemma 14.2 implies that as . Hence . If is a homotopically consistent conformal embedding of into , then 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 , 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] |