<<

arXiv:2108.12487v1 [math.DG] 27 Aug 2021 nw sthe as known tec point each at 1,Polm2] 1 .58) being 578]): p. [1, 22], Problem [16, eidti usini h lsia rbe funiformiza of problem classical the is question Behind nwrt hsclbae rbe sgvnb h Uniformiza the by (see given theorem is Koebe’s problem celebrated this to answer imn sphere Riemann S 04,57K20. 30F45, uesrae ohNs monster. Ness Loch surface, flute e 47019. mes e AAOIIYO EOTITTGTFUESURFACES FLUTE TIGHT ZERO-TWIST OF PARABOLICITY LXARDNO SALMRLS AIORMRZMALUENDAS RAMÍREZ CAMILO MORALES, ISRAEL ARREDONDO, ALEX fteReansphere Riemann the of ntecneto ipycnetdReansrae ehv t have we surfaces Riemann connected simply of context the In udmna pnqeto ntecasfiaino Riemann of classification the in question open fundamental A 2000 e od n phrases. and words Key atal upre yUiesddNcoa eClmi,S Colombia, de Nacional Universidad by supported Partially N NFRIAINO H OHNS MONSTER NESS LOCH THE OF UNIFORMIZATION AND oemrhct h ohNs Monster. Ness Loch the to homeomorphic a yei n only and if oemrhct h ohNs ose.Mr rcsl,w as we precisely, sequence More each Monster. to Ness Loch the to homeomorphic iegs scneuneo h eetwr fBsain Ha Basmajian, of work recent the Šari and of consequence As diverges. rv htteFcsa group Fuchsian the that prove esl,w soit oec euneo oiiera numb real positive of sequence each numb real to positive of associate sequence we a them versely, of one each to associate H ( A ahmtc ujc Classification. Subject Mathematics hte imn ufc uprsaGensfunction? Green’s a supports S surface Riemann a whether x n 2 BSTRACT n / ≤ nadto,w rsn nucutbefml fhyperbolic of uncountable an present we addition, In ) n Γ ∈ x b N n 0 sioerct eotittgtflt surface flute tight zero-twist a to isometric is yeproblem type p ≤ ,w banta h eotitflt surface flute zero-twist the that obtain we c, ´ oso-reFcsa group Fuchsian torsion-free a ∈ c C nti ae esuytezr-ws uesraeadwe and surface flute zero-twist the study we paper this In . n ˆ S ≤ n h onaédisc Poincaré the and h map the , d P n e.g. y ≤ x aaoi ye ece-ile aaees eotitt zero-twist parameters, Fenchel-Nielsen type, Parabolic n = b e C diverges. n 1 .588]). p. [1, ( is: n ooopi functions holomorphic and ≤ y n I ) a n NTRODUCTION n ∈ f + Z 1 Γ where , salcluiomzn aibeat variable uniformizing local a is uhingroup Fuchsian a , x sfis idi n nyi h series the if only and if kind first is 1 S imn ufc,fidaldomains all find surface, Riemann a y 01,5N5 71,3F0 30F25, 30F20, 57N16, 57N05, 20H10, n Γ ∆ x = uhta h ovxcr of core convex the that such di re’ ucin,while functions, Green’s admit ( a n , b G n , y c n uhthat such S d aiae,PoetHer- Project Manizales, ede S , d x x oevr we Moreover, . n , f so parabolic of is e : n ) S e ∈ r.Con- ers. H surfaces ers → inPoincaré- tion 2 R in(see tion sociate kobian / 5 G P S x and y x uhthat such is = surfaces, n p a the hat The . e.g. ight 2 the complex plane C does not. In general, Riemann surfaces are classified accordingly to: Definition 0.1 ([21, Section 2] [12, p.164]). We call the Riemann surface S elliptic if and only if S is compact (equivalently, closed). An open Riemann surface S is said to be parabolic type or parabolic, if S does not carry a negative non-constant subharmonic function. All open Riemann surface which are not parabolic will be called hyperbolic. Recall that all compact Riemann surface admits Green’s Function [15, p. 19]. To digress from this definition (see [24], [25], [28]) we recall that a Dirichlet problem on the surface S deals with the construction of harmonic functions u on a region Ω, such that coincides with the function f on ∂Ω. For this purpose, we are required to construct a continuous function u on Ω¯ = Ω ∂Ω, which coincides with f on ∂Ω and is harmonic on Ω. By the maximum∪ principle if such a function exists, then, it is uniquely deter- mined. This uniqueness imply that S does not admit a negative nonconstant subharmonic function. Equivalently, such a function exists if and only if there is not Green’s function on S (see [3, p. 204]). The type problem becomes equivalent to answering the question which open Riemann surfaces are of parabolic type? The interest in solving this problem has captured the attention of the math- ematical community for almost a hundred years. The first partial results to this question can be found in theorems of section 6 in [3] (see p. 204). After these results, numerous known characterizations of parabolic surfaces, that is equivalent to the type problem have been produced from potential the- ory, theory of functions, dynamics and geometry of surfaces, among others. This is a short list of known characterizations of Riemann surfaces of par- abolic type: if the Riemann surface S is the quotient of the Poincaré unit disk ∆ by a Fuchsian group Γ, then S is parabolic if and only if 1 γ(0) −k k The series 1+ γ(0) diverges [20, Theorem 5.2.1], • γ Γ k k ∈   Geodesic flowP on the unit tangent bundle of S is ergodic [20, Theo- • rem 8.3.4]; Mostow rigidity holds for G [2], [8], [27]; • The group Γ has the Bowen’s property [8], [13]. • Parabolicity of zero-twist tight flute surfaces. Recall that a flute surface is the unique infinite-type surface, up to homeomorphism, of zero and with space of ends homeomorphic to the ordinal ω + 1 (§1.1). The class conformed by all the Riemann flute surfaces is one of the simplest families where we can ask about the parabolicity problem. However, even in this class the problem of parabolicity is widely open (see e.g. [19], [11]). 3

A better manage family in the class of Riemann flute surfaces are the so- called tight flute surfaces. Such surfaces are hyperbolic surfaces obtained by starting with a geodesic pair of pants P0 with two punctures and one boundary geodesic and consecutively gluing geodesic pair of pants Pn (n 1) with one cusp and two boundary geodesics in an infinite chain (§1.5). A≥ tight flute surface is determined by its Fenchel-Nielsen parameters (§1.2) ( l , t ) where l and t are the length and twist parameter of the boundary { n n} n n closed geodesic αn of the surface obtained after gluing n pair of pants. This surface is denoted by S = S ( ln, tn )n N. If tn = 0 for all n N, then we say { } ∈ ∈ that S = S ( ln, 0 )n N is a zero-twist flute surface. In this article,{ } we∈ aim to present a new way of describing zero-twist flute surfaces; this description differs from the Fenchel-Nielsen parameters. Our punctual contribution is as follows: = Theorem 0.1. Given a zero-twist flute surface S S ( ln, 0 )n N0 there ex- = { } ∈ ists a unique sequence x (xn)n N0 of positive real numbers and a Fuchsian ∈ group Γx (completely determined by x) generated by the Möbius transfor- mations

2sn 1 sn 1 1 + − z 2sn 1 1 + − sn sn 1 ! − − sn sn 1 ! (1) gn(z):= − − − − , 2 2sn 1 z + 1 + − − sn sn 1 sn sn 1 ! − − − − n 2 with n N0 and sn = xi, such that the convex core of the surface H /Γx ∈ i=0 is isometric to S. Moreover,P Γx is of first kind if and only if the series xn H2 Γ diverges. In this case, / x coincides, up to isometry, with S andP it is complete.

Our construction above is inspired on Basmajian work in [9]. In his work, he gave a realization theorem showing that ( l , s ) are the Fenchel- { n} { n} Nielsen coordinates of a tight flute surface if and only if tn > ln2 for all n n i+1 odd, and t < ln2 for all n even, where t := = ( 1) l , see [9, Theo- n n i 1 − i rem 2]. Observe that these conditions does not dependP on twist parameters. In contrast with our construction of a zero-twist tight flute surface in terms N = R 0 of the sequence x (xn)n N0 + , we have that it does not depend of any additional condition on∈ the∈ sequence. Additionally, we can recover the Fenchel-Nielsen parameters of S = S ( l , 0 ) using { n }

sn 1 + sn + 2 √sn 1 sn (2) ln = ln − − . sn 1 + sn 2 √sn 1 sn ! − − − 4

N0 Let R+ be the space conformed by all sequences of positive real numbers indexed by the set N0 endowed with the compact-open topology, and denote = = by the set of all zero-twist flute surfaces, that is, : S S ( ln, 0 )n N0 . F F { N0 { } ∈ } Accordingly to Theorem 0.1, there is a bijective map σ : R+ which N0 → F associate to each sequence in R+ a unique zero-twist flute surface. We de- N0 note by S the image of x R+ under σ. x ∈ Recently, A. Basmajian, H. Hakobyan and D. Šaric´ in [11, Theorems 1.1 and 1.2] gave sufficient conditions on Fenchel-Nielsen parameters of a surface to determine parabolicity. In the particular case of zero-twist flute surfaces, they were able to characterize parabolicty, see [11, Theorem 1.5]. Using this characterization we have the following result.

Corollary 0.1. A zero-twist flute surface S x is of parabolic type if and only if xn diverges. P Theorem 0.1 gives an explicit set of generators for Γx. We have that if N x 1, 2 0 , then Γ is a subgroup of PSL(2, Z). This proves the following: ∈{ } x Corollary 0.2. The set of all zero-twist tight flute surfaces uniformized by subgroups of PSL(2, Z) contains a homeomorphic copy of the Cantor set.

An uncountable family of hyperbolic Loch Ness Monsters. Recall that the Loch Ness monster surface is the only surface (up to homeomorphism) which has infinite genus and only one end (§1.1). In [7] the authors intro- duced a Fuchsian group uniformizing the Loch Ness monster . We general- ize their construction and find an uncountable family of hyperbolic surfaces such that each one of these surface is homeomorphic to the Loch Ness Mon- ster. Denote by the set of all sequences y := (yn)n Z of elements yn = (a , b , c , d , eN) R5 satisfying ∈ n n n n n ∈

(3) a < b < c < d < e and e a + n n n n n n ≤ n 1 Let y := (yn)n Z . For each n Z, let fn and gn be the Möbius ∈ ∈ N ∈ transformations mapping σn ontoσ ˜ n and ρn ontoρ ˜n, respectively, where 2 σn, ρn,σ ˜ n andρ ˜n are the half-circles in the hyperbolic plane H depicted in Figure 1. Define (4) G := f , g : n Z Isom+(H2). y h n n ∈ i ≤ The result we establish is the following: 5

gn fn

σ˜ n σn ρ˜n ρn

an bn cn dn en

FIGURE 1. Half-circles σn, ρn, σ˜ n and ρ˜n.

Theorem 0.2. For each y , Gy is a Fuchsian group such that S y := 2 ∈ N H /Gy is topologically equivalent to the Loch Ness Monster. Moreover, if Gy is of first kind, then en = an+1 for all n Z, lim en = and lim an = . n n ∈ →∞ ∞ →−∞ −∞ Organization of the paper. In Section 1 we compile some classic results and concepts of the theory of surfaces necessary for the development of this work. Section 2 is dedicated to the proof of Theorem 0.1 and Corollary 0.1. In Section 3 we give the proof of Theorem 0.2.

1. PRELIMINARIES 1.1. Topological surfaces. A topological surface S is a topological space, connected, Hausdorff, second countable and locally homeomorphic to R2. In this text we are only concerned with orientable surfaces. Topological orientable surfaces are classified, up to homeomorphisms, by their genus g(S ) N0 , and a pair of nested topological spaces Ends (S ) Ends(S )∈ homeomorphic∪ {∞} to a pair of nested closed subsets of the Cantor∞ ⊆ space. The spaces Ends(S ) and Ends (S ) are called the end space and the non-planar ends space of S (or, ends∞ accumulated by genus), respectively. Moreover, any pair of nested closed subsets of the Cantor set can be realized as the space of ends of a connected, orientable topological surface. For more details, we refer the reader to [23]. Theorem 1.1 (Classification of topological surfaces, [18, §7], [23, Theorem 1]). Two orientable surfaces S 1 and S 2 having the same genus are topolog- ical equivalent if and only if there exists a homeomorphism f : Ends(S ) 1 → Ends(S 2) such that f (Ends (S 1)) = Ends (S 2). ∞ ∞ A topological surface is of infinite-type if it has fundamental group in- finitely generated. Now, we recall the definitions of the two infinite-type surfaces discussed in this text, these surfaces are the simplest infinite-type surface that we can find. We call to the isolated elements of the end space punctures of the surface. 6

Definition 1.1 ([9, p. 423]). The flute surface is the unique topological surface, up to homeomorphisms, of genus zero and with ends space homeo- morphic to the ordinal number ω + 1, see Figure 2......

...

FIGURE 2. A flute surface.

Definition 1.2 ([29, Definition 2]). The Loch Ness Monster is the unique infinite-type surface, up to homeomorphism, of infinite genus with exactly one end, see Figure 3. From the historical point of view as shown in [6], this is due to A. Phillips and D. Sullivan [22].

...

FIGURE 3. The Loch Ness monster.

In section 3 we use the following result to prove the first part of Theorem 0.2. Lemma 1.1 ([26, §5.1., p. 320]). The surface S has exactly one end if and only if for all compact subset K S there is a compact subset K′ S such ⊂ ⊂ that K K′ and S K′ is connected. ⊂ \ We say that a simple close curve on a surface S is essential if it is not isotopic to either a disk or a punctured disk. Recall that a topological pair of pants is a topological surface homeomorphic to a three punctured sphere. Definition 1.3. We say that a collection of pairwise disjoint essential curves in a surface S is a pair of pants decomposition of S if it decomposes the surface into a disjoint union of pair of pants. 7

Observe that any topological surface admits a pair of pants decomposi- tion [5]. 1.2. Fenchel-Nielsen parameters. We think that the reader is familiar with the basic tools of hyperbolic geometry, which can be found in [17], [12]. In this subsection we recall the definition of the Fenchel-Nielsen parameters of a hyperbolic surface. We refer the reader to see [14, Chapter 10] for the definition of the Fenchel-Nielsen parameters in the context of finite-type surfaces. We also recommend consulting [4] where the authors studied the Fenchel-Nielsen parameters for surfaces of infinite-type. A geodesic pair of pants is a complete surface S (respect to its hyperbolic structure) of finite hyperbolic area such that its interior is homeomorphic to a pair of pants and with at least one boundary a closed geodesic, see Figure 4. A tight pair of pants is a geodesic pair of pants that has exactly one punc- ture, see Figure 4-b. Similarly, a collection of pairwise disjoint of essential geodesic curves in a surface S is geometric pair of pants decomposition of S if it decompose the surface into geodesic pair of pants. In [4, Theorem 4.5] the authors gave sufficient conditions under which a topological pair of pants decomposition of a surface is straightened to a geometric pair of pants decomposition.

a) b) c)

FIGURE 4. All possible representations of geodesics pair of pants. In particular b) represents a tight pair of pants.

1.2.1. Length and twist parameters. Let P be a geodesic pair of pants and fix α a boundary curve of P. We choose a marked point x in α in the following way: Let β be either a boundary component (different from α) or a puncture of P. Let γ be the orthogeodesics between α and β. Then the marked point x in α is defined as the intersection point of α with γ. In Figure 5 is shown such marked point x in α which is obtained from drawing the orthogeodesic γ connecting the boundary component α with β. The hyperbolic length of the geodesic α is denoted by l(α). Let P′ be another geodesic pair of pants with a boundary geodesic α′ and suppose that l(α) = l(α′). We identify α and α′ by an isometry to obtain 8

β β

x γ x γ • P • P α α

FIGURE 5. The point x α is the intersection of α with the orthogeodesic γ. ∈ a complete hyperbolic surface from the two pairs of pants. The isometric identification is determined by the relative position of the marked points 1 1 x α and x′ α′ which is recorded by the twist parameter t(α) , . ∈ ∈ ∈ − 2 2 Namely, if x = x′ then t(α) = 0. If x , x′ then α r x, x′ consistsh of two { } arcs. Define t(α) as the length of the shorter arc contained in α r x, x′ . If t(α) = 1 then| we| define t(α) = 1 . If t(α) < 1 then we orient α{with} the | | 2 2 | | 2 orientation induced from P. If the shorter of the two arcs of α r x, x′ is the { } arc (x, x′) from x to x′ then t(α) = t(α) ; otherwise t(α) = t(α) . | | −| | Let P and P′ be two geodesic pair of pants with α and α′ boundary geo- desic curve of P and P′, respectively. Let γ and γ′ be the geodesic in P and P′, respectively, that define the marked points x in α and x′ in α′ as above. Suppose that α and α′ have the same hyperbolic length, that is, l(α) = l(α′). Then we identify α and α′, by gluing with an isometry T, to obtain a com- plete surface S from P and P′. This isometric identification is determined by the pair l(α), t(α) , such that the length parameter l(α) is the hyperbolic { } length of α, and the twist parameter t(α) 1 , 1 corresponds to the rel- ∈ − 2 2 ative position of the marked points x in α and x′ iin α′, which is given as follows:

If x = T(x′), then t(α) = 0. ∗ 1 If x , T(x′), then t(α) is equal to the hyperbolic length of the ∗ | | ≤ 2 shorter arc σ having endpoints x and T(x′) contained in α x, T(x′) , = 1 = 1 \{1 } divided by l(α). If t(α) 2 , then t(α) 2 . If t(α) < 2 then we orient α with the orientation| | induced from P. If σ| is the| arc from x to T(x′) then t(α) = t(α) ; otherwise t(α) = t(α) . See Figure 6. a). | | −| |

Remark 1.1. If the twist parameter t(α) = 0, then γ γ′ is a geodesic in S that is orthogonal to the closed geodesic α, see Figure∪ 6. b). Definition 1.4. Let S be a surface obtained by gluing countably many ge- odesic pair of pants in the way described above. These gluing generate a 9

σ T(x′) x′ x x x′ • • γ′ γ • γ • γ′ α α′ α α′

P P′ P P′

a). Twist parameter. b). If t(α) = 0, then γ γ′ is orthogonal to α. ∪ FIGURE 6 geometric pair of pants decomposition α : i N ofS. Letl and t be the { i ∈ } i i length and twist parameter respectively of αi. The collection of pairs

(5) ( li, ti )i N, { } ∈ is called the Fenchel-Nielsen parameters of S. As the hyperbolic metric on S is uniquely determined by its Fenchel-Nielsen parameters ( li, ti )i N, then { } ∈ S := S ( li, ti )i N. { } ∈ The surface S might not be complete in the induced hyperbolic metric. V. Álvarez and J. M. Rodríguez in [5] showed that the boundary of the metric completion of S consists of simple closed geodesics and bi-infinite simple geodesics. Moreover, they proved that by attaching funnels to the closed geodesics and attaching geodesic half-planes to the bi-infinite geodesics of the boundary of the metric completion of S , we obtain a surface Sˆ homeo- morphic to S with a geodesically complete hyperbolic metric such that the -inclusion i : S ֒ Sˆ is an isometric embedding. Conversely, any geodesi cally complete surface→ is obtained by attaching funnels and half-planes to the convex core of the surface, see also [10]. We are in a position to define the so-called tight flute surfaces. Definition 1.5 ([9, p. 423]). A tight flute surface S is a surface obtained by starting with a geodesic pair of pants P0 with two punctures and then consecutively gluing tight pairs of pants P ,n 1, as is shown in Figure 7. n ≥

If S is a tight flute surface, we denote by αn the closed geodesic of the boundary of the surface obtained after gluing n geodesic pairs of pants and let ln and tn be the length and twist parameters associated to αn. In terms of Fenchel-Nielsen parameters this surface is denoted by S := S ( ln, tn )n N. { } ∈ 10

α2 α3 ... P0 α1 P1 α2 P2 P3

FIGURE 7. A tight flute surface.

Definition 1.6. If all twist parameters of a tight flute surface are zero, that is S := S ( ln, 0 )n N, then we called it a zero-twist tight flute surface. { } ∈ We have that the problem of parabolicity is fully understood for zero- twist flute surfaces.

Theorem 1.2 ([11, Theorem 1.5]). A zero-twist tight flute surface S ( ln, 0 )n N is parabolic if and only if one of the following holds: { } ∈

(1) The surface S ( ln, 0 )n N is complete. { } ∈ ∞ ln (2) The series e− 2 = . n=1 ∞ P 2. PROOF OF THEOREM 0.1 = We begin by associating to each zero-twist flute surface S S ( ln, 0 )n N0 { } ∈ a sequence x of positive real numbers and a Fuchsian group Γx such that 2 the convex core of H /Γx is isometric to S . We do this by constructing explicitly a hyperbolic polygon by cutting S along an infinite collection P of bi-infinite geodesics which serves as fundamental domain for Γx. = Let S S ( ln, 0 )n N0 be a zero-twist flute surface, which come from { } ∈ gluing the geodesic pairs of pants of the family Pn n N0 (see Figure 8). For each n N, let α be the closed geodesic in S ,{ which} ∈ comes from gluing ∈ n the geodesic pair of pants Pn 1 and Pn. Let 0 and s0 be the punctures of P0 and for n 1, let s be the unique− puncture of P . Now we describe how ≥ n n to obtain an ideal hyperbolic polygon by removing a collection γn n N0 of geodesics in S having endpoints onP the punctures of S . Let γ{ } ∈P 0 ⊂ 0 be the geodesic connecting the two punctures of P0. Let also γn be the geodesic connecting the punctures sn 1 and sn, which is orthogonal to the − close geodesic αn 1, see Remark 1.1. Finally, we draw on S the geodesic − ray β orthogonal to each αn and having one endpoint in 0. In Figure 8 is shown S with the closed geodesics αn, the bi-infinite geodesics γn and the geodesic ray β. 11

s0 s1 s2

γ2 γ0 γ1 γ3

P0 α1 P1 α2 P2 α3 ... 0 β

FIGURE 8. A zero-twist tight flute surface.

Now, we cut S along the geodesics γn. Then S turns into an ideal hy- perbolic polygon with infinitely many sides and infinitely many ideal vertices. More precisely,P the edges and vertices of are given as follows: + P For each n 0, let γ and γ− be the sides of coming from cutting ∗ ≥ n n P S along γn. + For each n 0, let s and s− be the ideal vertices of coming from ∗ ≥ n n P the puncture sn, see Figure 9.

s0− s1− s2−

γ1− γ2− γ0− β 0 + γ0 + + γ1 γ2 + + + s0 s1 s2

FIGURE 9. Ideal hyperbolic polygon obtained by cutting P S ( ln, 0 )n N0 along the geodesics γn. { } ∈

+ Remark 2.1. For each n 0, the sides γ and γ− of the ideal hyperbolic ≥ n n polygon are identified by a Möbius transformation gn. If we identify P+ the sides γn and γn− using gn, then we recover the zero-twist tight surface

S ( ln, 0 )n N0 . { } ∈ 12

The ideal hyperbolic polygon can be thought in the hyperbolic plane H2 satisfying the following propertiesP (see Figure 10): Its vertices are on the real axis. Up to take a real translation, we can ∗ assume that the vertex 0 of is equal to the complex number zero and that the geodesic ray β coincidesP with the imaginary axis. The + collection of vertices (sn )n N0 defines a strictly increasing sequence ∈ of positive real numbers. The collection of vertices (sn−)n N0 defines a strictly decreasing sequence of negative real numbers.∈ Given that + has a reflexive symmetry fixing the ray β, then sn and sn− are P = + symmetric with respect to the imaginary axis, it means, sn− sn , for all n 0. − ≥ N + For each n 0, the edges γn and γn− of are half-circles having the ∗ ∈ + + P same radius and endpoints sn 1 and sn ; and sn− 1 and sn−, respectively. Thus, the intersection of any− two of those− edges is either empty or they meet at the same point in the real line. Moreover, if we choose γ one of these half-circles, then the other half-circles are in the exterior of γ.

α2 α1

β + γ− γ 1> α0 + < 1 + γ2− γ0− γ0 γ2 ... > > < < ... + + + s2− s1− s0− 0 s0 s1 s2

FIGURE 10. Realization of the ideal hyperbolic polygon in H2. P

= To the zero-twist flute surface S we associated the sequence x (xn)n N0 of positive real numbers that satisfies ∈ n + = (6) sn xi. Xi=0

Thus, to the sequence x we associated the group Γx given by (7) Γ = g : n N . x h n ∈ 0i Now we prove that Γx satisfies the properties described in Theorem 0.1. N = + H2 For each n 0, let sn : sn . On the hyperbolic plane , we draw the half- + ∈ circles γ0 and γ0− having endpoints 0 and s0, and 0 and s0, respectively. For + − each n 1, we draw the half-circles γn and γn− having endpoints sn 1 and ≥ − 13 sn, and sn 1 and sn, respectively. See Figure 10. We remark that the half- −+ − − circle γn and γn− are symmetric with respect to the imaginary axis. Then the Möbius transformation gn is given by

2sn 1 sn 1 1 + − z 2sn 1 1 + − sn sn 1 ! − − sn sn 1 ! (8) gn(z):= − − − − , 2 2sn 1 z + 1 + − − sn sn 1 sn sn 1 ! − − − − + which send the half-circle γ onto the half-circle γ−, for each n 0. n n ≥ Observe that g0 is parabolic and, for each n 1, gn is hyperbolic with trace equal to ≥

2sn 1 2 1 + − > 2. sn sn 1 ! − − Γ As x is obtained by sided paring the sides of the convex ideal polygon , then it is a non elementary group and its set of generators is conformed P by non elliptic elements. Thus, Γx is a Fuchsian group (see [12, Theorem 8.3.1, p. 198]), with fundamental domain for its action on H2. In what follows we describe more precisely. P Recall that if γ is aP half-circle in H2 having as center the point z R and radius r > 0, then H2 γ has two connected components. The connected∈ component of H2 γ equal\ to the set w H2 : z w > r is called the exterior of γ and it\ is denoted by Ext(γ{). To∈ the complement| − | } of the closure in H2 of Ext(γ) we called the interior of γ and we denote it by Int(γ), see Figure 11.

Ext(γ) γ Int(γ)

z r z z + r −

FIGURE 11. Exterior and interior of a half-circle γ in H2.

The limit lim sn is well defined because the sequence (sn)n N0 is the par- n ∈ tial sums of x→∞. Let a be such limit, which must be either a positive real number or infinity. If a is infinity, then is equal to the closure (in H2) of P + 2 D := Ext(γ ) Ext(γ−) H . n ∩ n ⊂ n\N0 ∈  14

Otherwise, if a is a positive real number, then we denote by γ the half-circle in H2 having a and a as ends points. In this case is given by − P D Int(γ) H2. ∩ ⊂ Notice that D is a fundamental domain for Γx and D/Γx is a complete hy- perbolic surface. Therefore, if a is infinity then /Γ = D/Γ is isometric P x x to S and Γx is a Fuchsian group of first kind. Otherwise, if a is finite then /Γx is isometric to S and it is not a complete hyperbolic surface because Pthe projection points of γ are limit points which are not in the surface. How- ever, /Γx is the convex core of D/Γx. Here finished the proof of Theorem 0.1. P 

3. PROOF OF THEOREM 0.2 From the introduction we recall that denotes the set of all sequences N 5 y := (yn)n Z of elements yn = (an, bn, cn, dn, en) R satisfying ∈ ∈

(9) a < b < c < d < e and e a + n n n n n n ≤ n 1 Let y := (yn)n Z . For each n Z, let fn and gn be the Möbius ∈ ∈ N ∈ transformations mapping σn ontoσ ˜ n and ρn ontoρ ˜n, respectively, where 2 σn, ρn,σ ˜ n andρ ˜n are the half-circles in the hyperbolic plane H depicted in Figure 1. Let (10) G := f , g : n Z Isom+(H2). y h n n ∈ i ≤ To start with the proof of Theorem 0.2, the following lemma is required. Lemma 3.1. Let σ and σ˜ denote the half-circles having endpoints a and b; c and d respectively, where a < b < c < d (see Figure 12). Then the Möbius transformation f which sends the half-circle σ onto the half-circle σ˜ is hyperbolic. Proof. By hypothesis, the half-circles σ andσ ˜ are disjoint, and each one of the half-circles is contained in the exterior of the other half-circle. Explic- = + b a ˜ = + d c itly, they have centers at the real numbers O a −2 and O c −2 = b a = d c respectively, and radius r −2 andr ˜ −2 , respectively. Then the Möbius transformation f is given by rr˜ (11) f (z) = − + O˜. z O − Let us note that the fixed points of f are (r + r˜) √δ (12) z = − ± , 2 15

f

σ˜ σ r r˜

a O b c O˜ d

FIGURE 12. Half-circles σ and σ˜ . where δ = (O+O˜)2 4(OO˜ +rr˜) = (O O˜)2 4rr˜. Since 0 < r +r˜ < O˜ O , then − − − | − | 0 (r r˜)2 = (r + r˜)2 4rr˜ < (O˜ O)2 4rr˜ = δ. ≤ − − − − It implies that √δ is a real number. Therefore, f is hyperbolic. 

The group Gy is Fuchsian. By Lemma 3.1, Gy is generated by hyperbolic elements. So, by Theorem 8.2.1 in [12], Gy is a Fuchsian group. By con- struction, the collection of half-circles = σn, σ˜ n,ρn, ρ˜n : n Z are pairwise disjoint, and each one of the half-circlesC { in is contained∈ in} the exterior of each one of the other half-circles in . FromC Theorem 3.3.5 in C [17], a fundamental region D(Gy) for the Fuchsian group Gy is given by

(13) D(G ) = (Ext(σ ) Ext(˜σ ) Ext(ρ ) Ext(˜ρ )) H2. y n ∩ n ∩ n ∩ n ⊂ \n Z ∈ Given that the intersection of any two different elements belonged to is either empty or at infinity, the last means, they meet at the same point inC 2 the real line, then Gy acts freely and properly discontinuously on whole H . 2 Hence, the quotient space S := H /Gy is a complete Riemann surface.  In order to prove that S is topologically equivalent to the Loch Ness mon- ster, we must verify that S has only one end and infinite genus. The proof uses the same ideas appearing in [6]. The surface S has only one end. By Lemma 1.1 is enough to prove that for any compact subset K of S , there exists a compact subset K′ of S such that K K′ and the space S K′ is connected. ⊂ \ 1 Let K be a compact subset of S . Observe that K˜ := π− (K) is a compact subset of D(Gy) where π : D(Gy) S = D(Gy)/Gy is the quotient map. Thus, K˜ is a compact subset of H2.→ Therefore, there exist closed intervals Ix and Iy in the real and imaginary axes, respectively, such that, πx(K˜ ) Ix 2 ⊂ and πy(K˜ ) Iy, where πx : H R is the standard projection map on the real axis and⊂ π : H2 (0, + →) is the projection map on the imaginary y → ∞ 16 axis. By construction of the fundamental domain for Gy we have that K˜ ′ := D(Γ) I I is a compact subset of D(G ) such that K˜ K˜ ′. Thus, ∩ x × y y ⊂ K′ := π(K˜ ′) is a compact subset of S which contains K. Finally, we can verify that S r K′ is connected. The surface S has infinite genus. For each n Z, we define the strip ∈ S˜ := D(G ) z H2 : a < Re(z) < e H2. n y ∩{ ∈ n n} ⊂ Observe that S˜ / f , g defines, under the projection map π, a subsurface n h n ni S n of S topologically equivalent to a torus minus a disk. By construction, any for two different integers m , n, the subsurfaces S n and S m are disjoint because the strips S˜ S˜ = . This shows that S has infinite genus. n ∩ m ∅ If Gy is of first kind, then en = an+1 for all n Z, lim en = and lim an = n n ∈ →∞ ∞ →−∞ . Suppose that Gy if of first kind but that the conclusion is false. In any −∞case it is possible to find an interval I R not contained in the limit set of ⊂ G which leads to a contradiction. Indeed, if e , a + for some n Z then y n n 1 ∈ I = (en, an+1) is such interval. If lim en = r < then I = (a, ). Finally, if n →∞ ∞ ∞ lim an = R > , then I = ( , R). n →−∞ −∞ −∞ 4. SOME OPEN QUESTIONS We conclude with some more open questions: Problem 4.1. Determine the points y of , which defines Fuchsian groups N Γy of first kind. Problem 4.2. Which points y of define Fuchasian groups uniformizing Loch Ness monsters of parabolic type?N

N0 Problem 4.3. Find a sequence x R+ such that the zero twist tight flute surface associated be of parabolic∈ type but not complete.

ACKNOWLEDGEMENTS Camilo Ramírez Maluendas was partially supported by UNIVERSIDAD NACIONAL DE COLOMBIA, SEDE MANIZALES. He has dedicated this work to his beautiful family: Marbella and Emilio, in appreciation of their love and support.

REFERENCES [1] William Abikoff, The uniformization theorem, Amer. Math. Monthly 88 (1981),no. 8, 574-592. [2] Stephen Agard, A geometric proof of Mostow’s rigidity theorem for groups of diver- gence type, Acta Math. 151 (1983), 231-252. 17

[3] Lars V. Ahlfors and Leo Sario, Riemann surfaces, Princeton Mathematical Series, No. 26, Princeton University Press, Princeton, N.J., 1960. [4] Daniele and Liu Alessandrini Lixin and Papadopoulos, On Fenchel-Nielsen coordi- nates on Teichmüller spaces of surfaces of infinite type, Ann. Acad. Sci. Fenn. Math. 36 (2011), no. 2, 621-659. [5] Venancio and Rodríguez Álvarez José M., Structure theorems for Riemann and topo- logical surfaces, Journal of the London Mathematical Society 69 (2004), no. 1, 153- 168. [6] John A. and Ramírez Maluendas Arredondo Camilo, On the infinite Loch Ness mon- ster, Comment. Math. Univ. Carolin. 58 (2017), no. 4, 465-479. [7] John A. and Ramírez Maluendas Arredondo Camilo, On Infinitely generated Fuch- sian groups of the Loch Ness monster, the Cantor tree and the Blooming Cantor tree, Comp. Man. 7 (2020), no. 1, 73-92. [8] Kari and Zinsmeister Astala Michel, Mostow rigidity and Fuchsian groups, C. R. Acad. Sci. Paris Ser. I Math. 311 (1990), 301-306. [9] Basmajian, Hyperbolic Structures for Surfaces of Infinite Type, Transactions of the American Mathematical Society 336 (1993), no. 1, 421-444. [10] Ara and Šaric´ Basmajian Dragomir, Geodesically complete hyperbolic structures, Mathematical Proceedings of the Cambridge Philosophical Society, 2019, pp. 219- 242. [11] Ara and Hakobyan Basmajian Hrant and Šaric,´ The type problem for Riemann sur- faces via Fenchel-Nielsen parameters, arXiv preprint arXiv:2011.03166 (2020). [12] Alan F. Beardon, The Geometry of Discrete Groups, 1st, Graduate Texts in Mathe- matics, vol. 91, Springer-Verlag, New York, 1983. [13] Christopher J. Bishop, Divergence Groups Have the Bowen Property, Annals of Mathematics 154 (2001), no. 1, 205–217. [14] Benson Farb and Dan Margalit, A primer on mapping class groups, Princeton Math- ematical Series, vol. 49, Princeton University Press, Princeton, NJ, 2012. [15] Hershel M. and Kra Farkas Irwin, Riemann Surfaces, 2nd ed., Graduate Texts in Mathematics, vol. 71, Springer-Verlag, New York, 1992. [16] David Hilbert, Mathematical problems, Bull. Amer. Math. Soc. (N.S.) 37 (2000), no. 4, 407-436. Reprinted from Bull. Amer. Math. Soc. 8 (1902), 437-479. [17] Svetlana Katok, Fuchsian groups, Chicago Lectures in Mathematics, University of Chicago Press, Chicago, IL, 1992. [18] Béla Kerékjártó, Vorlesungen über Topologie I, Mathematics: Theory & Applications, Springer, Berlín, 1923. [19] Katsuhiko and Rodríguez Matsuzaki José M, Planar Riemann surfaces with uni- formly distributed cusps: parabolicity and hyperbolicity, Mathematische Nachrichten 290 (2017), no. 7, 1097–1112. [20] Peter J. Nicholls, The Ergodic Theory of Discrete Groups, London Mathematical So- ciety Lectures Note Series, vol. 143, Cambridge Univertisy Press, New York, 1989. [21] Makoto Ohtsuka, Dirichlet problems on Riemann surfaces and conformal mappings, Nagoya Mathematical Journal 3 (1951), 91-137. [22] Anthony and Sullivan Phillips Dennis, Geometry of leaves, Topology 20 (1981),no. 2, 209-218. [23] Ian Richards, On the classification of noncompact surfaces, Trans. Amer. Math. Soc. 106 (1963), 259-269. 18

[24] L. Sario and M. Nakai, Classification theory of Riemann surfaces, Die Grundlehren der mathematischen Wissenschaften, Band 164, Springer-Verlag, New York-Berlin, 1970. [25] Wilhelm Schlag, A course in complex analysis and Riemann surfaces, Graduate Stud- ies in Mathematics, vol. 154, American Mathematical Society, Providence, RI, 2014. [26] Ernst Specker, Die erste Cohomologiegruppe von Überlagerungen und Homotopie- Eigenschaften dreidimensionaler Mannigfaltigkeiten, Comment. Math. Helv. 23 (1949), 303-333 (German). [27] Pekka Tukia, Differentiability and rigidity of Möbius groups, Invent. Math. 82 (1985), 557-578. [28] M. Tsuji, Potential theory in modern function theory, Chelsea Publishing Co., New York, 1975. Reprinting of the 1959 original. [29] Ferrán Valdez, Infinite genus surfaces and irrational polygonal billiards, Geom. Ded- icata 143 (2009), 143-154.

JOHN A. ARREDONDO, DEPARTAMENTO DE MATEMÁTICAS, FUNDACIÓN UNIVERSITARIA KONRAD LORENZ, BOGOTÁ, COLOMBIA. Email address: [email protected]

ISRAEL MORALES, INSTITUTODE MATEMÁTICAS UNAM UNIDAD OAXACA,OAXACADE JUÁREZ,OAX- ACA,MÉXICO. Email address: [email protected]

CAMILO RAMÍREZ MALUENDAS, DEPARTAMENTO DE MATEMÁTICAS Y ESTADÍSTICA, UNIVERSIDAD NACIONALDE COLOMBIA, SEDE MANIZALES,MANIZALES 170004, COLOMBIA. Email address: [email protected]