Planar Quasiconformal Mappings; Deformations and Interactions

Total Page:16

File Type:pdf, Size:1020Kb

Planar Quasiconformal Mappings; Deformations and Interactions Planar quasiconformal mappings; deformations and interactions Kari Astala To F.W. Gehring on his 70th birthday 1 Introduction The theory of quasiconformal mappings divides traditionally into two branches, the mappings in the plane and the case of higher dimensions. Basically, this is not due to the history of the topic but rather since planar quasiconformal mappings admit flexible methods (so far) not available in space. In this expository paper we wish to describe some recent trends and activities in quasiconformal theory peculiar to the plane. It is obvious, though, that not all topics can be covered no matter which point of view is taken; many important advances and connections must necessarily be bypassed. Therefore we concentrate on a specific theme, a property that singles out the difference between mappings in plane and in space: Planar quasiconformal mappings admit an effective deformation theory, any such mapping can be deformed to the identity, within the family of all quasiconformal mappings. Our goal is to discuss different aspects of these parametric representations of planar qua- siconformal mappings, with a few occasional words of the background of the topics, to see how and why these mappings can be deformed, and above all which are the results and con- sequences obtained, which are the questions still open. In particular, in many respects it is the deformation properties which are the "cause" of the many connections and applications of quasiconformality. Thus we try to give a particular attention to discuss how this field of ideas interacts with other parts of mathematical analysis. It will be no surprise to see the strong influence of Gehring also within these contexts. By the analytic definition a homeomorphism f :Ω ! Ω0 between planar domains Ω and Ω0 1;2 is called K-quasiconformal if it is contained in the Sobolev class Wloc (Ω) and its directional derivatives satisfy maxαj@αf(x)j ≤ Kminαj@αf(x)j a:e: x 2 Ω: A corresponding definition works of course for higher dimensional mappings as well, but what is special to the mappings in plane is the well known Beltrami differential equation, that the inequality is equivalent to @f(z) = µ(z) @f(z); (1) K−1 where µ is the complex dilatation or the Beltrami coefficient of f with kµk1 = K+1 < 1. This is a first order linear differential equation, solvable by singular integrals. Note also that even if the recent work of Donaldson and Sullivan [12], developed also by Iwaniec and Martin [20], found a counterpart of (1) in even dimensional spaces this machinery still lacks the required properties for finding parametric representations in space. However, in the planar case (1) always does admit quasiconformal solutions: by the measurable Riemann 1 1 1;2 mapping theorem for any µ 2 L (C) with kµk1 < 1 we have a homeomorphic Wloc -solution of (1), see [2] [9] [21], unique up to postcomposing with a M¨obiustransformation. This result is the starting point of the deformation theory for planar quasiconformal map- pings. It turns out that using (1) we can actually construct two natural and similar but in many respects different ways of parametrizing the mappings. To be more precise, the basic idea for the first deformation is that we may embedd our mapping f to a flow fftgt2R of a (nonautonomous) vector field v. In fact, by careful choices obtain a geodesic in the Teichm¨ullermetric, −1 log K(ft ◦ fs ) = jt − sj; f0 = id: For this approach, to control the flow we must find the infinitesimal generating vector field v for which @tft(x) = v(t; ft(x)); ft0 = f: (2) Thus the questions here are how to find v, what are its properties and, especially, how to characterize vector fields generating quasiconformal flows. The second approach arises in connection with deformations of complex analytic structures, where the holomorphicity of the parametrization is an essential feature required. Then one must abandon the vector field setting and more directly use the properties of the equation (1). Namely, if µ 2 L1(C) has a compact support, let f = f µ be the solution of (1) with f(z) = z + O(jzj−1). Then, if µ varies holomorphically; so does f µ(z); (3) thus giving us holomorphic ways of deforming f to the identity. Indeed, as is well known this holomorphic dependence follows with the help of the classical Beurling operator, the two dimensional counterpart of the Hilbert transform, 1 Z !(ζ) dm(ζ) S!(z) = − : π (ζ − z)2 C It defines an isometry in L2(C), and moreover, what makes it useful for the Beltrami equation is that S transforms @-derivatives to @-derivatives, S(@u) = @u; 8u 2 W 1;2(C): (4) Using this identity for u(z) = f(z) − z and combining with the equation @f(z) = µ(z)@f(z) we have @f(z) = 1 + (I − Sµ)−1S(µ): (5) 2 Since as an operator on L (C), the norm of g 7! S(µg) is at most kµk1kSkL2 < 1, the expression is well defined and shows that @f(z) depends holomorphically on µ. Since by the Cauchy formula 1 Z µ∂f(ζ) dm(ζ) f(z) = z − ; π (ζ − z) C we have the same conclusion for f(z). 2 Flows of quasiconformal mappings The parametric representations in terms of flows of quasiconformal mappings was originally derived by Shah Dao-Shing [32]. However Gehring and Reich [17] were the first who sys- tematically developed and utilized the method. Their goal was to find a new approach to 2 quasiconformal distortion problems. We return to these applications below after first analysing the basic properties and later consequences of their construction. To build up the flow, let us agree first on the notation fµ for the mapping with complex dilatation µ, normalized so that 0, 1 and 1 are fixed. To embedd a given K-quasiconformal mapping f to a continuous flow, we may assume that f = fµ and define then a further family of dilatations by µ(z) t µ (z) = tanh( tanh−1 jµ(z)j);T = log K: t jµ(z)j T According to the measurable Riemann mapping theorem for each t 2 R there exists a unique normalized mapping ft = fµt . Geometrically, µt(z) is chosen to be the point on the radius through µ(z) such that in the unit disk the hyperbolic distance from 0 to µt(z) is equal to t=T times the distance between 0 and µ(z). From the analytic view, or by the transformation formulas of the complex dilatation for a composition of two quasiconformal mappings [21], the µt(z)'s are constructed precisely so that we have 1 log K(f ◦ f −1) = jt − sj: t s T 1 We have made the above choice T for the speed of the flow for later purposes, c.f. (6), (7). Sometimes it is also convenient to have a finite version of the flow. That is, considering t only a discrete set of parameter values and noticing that K(fµt ) ≤ e we have the factorization theorem: For any > 0, all quasiconformal mappings in C can be written as a composition of a finite number of (1 + )-quasiconformal mappings. The generating vectorfield can next be given simply as −1 v(z; t) = @tft ◦ ft (z): (6) In other words, the equation @tft(x) = v(t; ft(x)), fT = f is satisfied directly by the definition. Less obvious from here is to see any specific properties and, in particular, any explicit character- izations of vector fields that generate quasiconformal flows. However, by a clever combination of (1) and (2) Gehring and Reich found two interesting identities 1 µ(z) @zft @zv ◦ ft = (7) 2 jµ(z)j @zft and @tJft = 2 Re[@v ◦ ft]Jft (8) for the vector field v and the Jacobian determinant Jft of ft. The first of these answer imme- diately our question. Corollary 2.1 The vector field v generating the above flow of normalized quasiconformal map- pings fftgt2R satisfies 1 k@vk ≤ : 1 2 Indeed, as shown later by Reimann [28] and Semenov [33] the corollary admits a direct converse. 1;1 Theorem 2.2 Suppose we have a continuous vector field v 2 Wloc (C) such that v = O(jzj log jzj) as jzj ! 1 and k@vk1 ≤ C < 1: Then the flow fgtgt2R of v consists of quasiconformal mappings with Ct K(gt) ≤ e : 3 The exact value of the constant C counts, of course, only for the speed of the flow. Remarkably this direction holds in any dimensions ([28]),([33]). We only need to reinterpret the boundedness of j@v(z)j in terms of the differential operator 1 1 ΣV (x) = [DV (x)t + DV (x)] − (T rDV (x))I; x 2 Rn; 2 n in the case of vectorfields in Rn. As shown by Ahlfors [3], in two dimensions jΣV (z)j = j@V (z)j. To indicate the usefulness of this approach let us recall another closely related argument of Reimann [30]: Suppose we have a quasiconformal mapping f on the sphere S2 which defines an isomorphisms between two groups of M¨obiustransformations G and Γ, i.e. f −1 ◦ Γ ◦ f = G . Since the M¨obiusgroups themselves extend canonically to the unit ball B3 we may ask if there is a way to extend the isomorphism as well. For this [30] shows first that the generating vector field v constructed above from the (two dimensional) mapping f is G-invariant in an appropriate sense. Moreover, in all dimensions an M¨obius-equivariant extension of vector fields is easily found by a Poisson-type integraloperator.
Recommended publications
  • The Riemann Mapping Theorem Christopher J. Bishop
    The Riemann Mapping Theorem Christopher J. Bishop C.J. Bishop, Mathematics Department, SUNY at Stony Brook, Stony Brook, NY 11794-3651 E-mail address: [email protected] 1991 Mathematics Subject Classification. Primary: 30C35, Secondary: 30C85, 30C62 Key words and phrases. numerical conformal mappings, Schwarz-Christoffel formula, hyperbolic 3-manifolds, Sullivan’s theorem, convex hulls, quasiconformal mappings, quasisymmetric mappings, medial axis, CRDT algorithm The author is partially supported by NSF Grant DMS 04-05578. Abstract. These are informal notes based on lectures I am giving in MAT 626 (Topics in Complex Analysis: the Riemann mapping theorem) during Fall 2008 at Stony Brook. We will start with brief introduction to conformal mapping focusing on the Schwarz-Christoffel formula and how to compute the unknown parameters. In later chapters we will fill in some of the details of results and proofs in geometric function theory and survey various numerical methods for computing conformal maps, including a method of my own using ideas from hyperbolic and computational geometry. Contents Chapter 1. Introduction to conformal mapping 1 1. Conformal and holomorphic maps 1 2. M¨obius transformations 16 3. The Schwarz-Christoffel Formula 20 4. Crowding 27 5. Power series of Schwarz-Christoffel maps 29 6. Harmonic measure and Brownian motion 39 7. The quasiconformal distance between polygons 48 8. Schwarz-Christoffel iterations and Davis’s method 56 Chapter 2. The Riemann mapping theorem 67 1. The hyperbolic metric 67 2. Schwarz’s lemma 69 3. The Poisson integral formula 71 4. A proof of Riemann’s theorem 73 5. Koebe’s method 74 6.
    [Show full text]
  • Schiffer Operators and Calculation of a Determinant Line in Conformal Field
    New York Journal of Mathematics New York J. Math. 27 (2021) 253{271. Schiffer operators and calculation of a determinant line in conformal field theory David Radnell, Eric Schippers, Mohammad Shirazi and Wolfgang Staubach Abstract. We consider an operator associated to compact Riemann surfaces endowed with a conformal map, f, from the unit disk into the surface, which arises in conformal field theory. This operator projects holomorphic functions on the surface minus the image of the conformal map onto the set of functions h so that the Fourier series h ◦ f has only negative powers. We give an explicit characterization of the cok- ernel, kernel, and determinant line of this operator in terms of natural operators in function theory. Contents 1. Introduction 253 2. Preliminaries 255 3. Jump decomposition and the Faber and Grunsky operators 260 4. Applications to the determinant line bundle 265 References 270 1. Introduction In conformal field theory the central charge, or conformal anomaly, plays an important role and is encoded geometrically by the determinant line bundle of a certain family of operators π(R;f) over the rigged moduli space, see Y.-Z. Huang [3] or G. Segal [15]. Here R is a Riemann surface and f is a conformal map into that surface (see below). In this paper, we give a simple and explicit description of the cokernel of π(R;f), in the case of surfaces with one boundary curve. Furthermore, we explicitly relate the operator π(R;f) to classical operators of function theory: the Faber, Grunsky, and Schiffer operators. In particular, we give an explicit formula for its inverse in terms of Received May 20, 2020.
    [Show full text]
  • Meromorphic Functions with Two Completely Invariant Domains
    MEROMORPHIC FUNCTIONS WITH TWO COMPLETELY INVARIANT DOMAINS WALTER BERGWEILER AND ALEXANDRE EREMENKO Dedicated to the memory of Professor I. N. Baker Abstract. We show that if a meromorphic function has two completely invari- ant Fatou components and only finitely many critical and asymptotic values, then its Julia set is a Jordan curve. However, even if both domains are attract- ing basins, the Julia set need not be a quasicircle. We also show that all critical and asymptotic values are contained in the two completely invariant compo- nents. This need not be the case for functions with infinitely many critical and asymptotic values. 1. Introduction and main result Let f be a meromorphic function in the complex plane C. We always assume that f is not fractional linear or constant. For the definitions and main facts of the theory of iteration of meromorphic functions we refer to a series of papers by Baker, Kotus and L¨u [2, 3, 4, 5], who started the subject, and to the survey article [8]. For the dynamics of rational functions we refer to the books [7, 11, 20, 24]. 1 AcomponentD of the set of normality is called completely invariant if f − (D)= D. There is an unproved conjecture (see [4, p. 608], [8, Question 6]) that a mero- morphic function can have at most two completely invariant domains. For rational functions this fact easily follows from Fatou’s investigations [14], and it was first explicitly stated by Brolin [10, 8]. Moreover, if a rational function has two com- pletely invariant domains, then§ their common boundary is a Jordan curve on the Riemann sphere, and each of the domains coincides with with the basin of attrac- tion of an attracting or superattracting fixed point, or of an attracting petal of a neutral fixed point with multiplier 1; see [14, p.
    [Show full text]
  • Bending Laminations on Convex Hulls of Anti-De Sitter Quasicircles 3
    BENDING LAMINATIONS ON CONVEX HULLS OF ANTI-DE SITTER QUASICIRCLES LOUIS MERLIN AND JEAN-MARC SCHLENKER 2 Abstract. Let λ− and λ+ be two bounded measured laminations on the hyperbolic disk H , which “strongly fill” (definition below). We consider the left earthquakes along λ− and λ+, considered as maps from the universal Teichm¨uller space T to itself, and we prove that the composition of those left earthquakes has a fixed point. The proof uses anti-de Sitter geometry. Given a quasi-symmetric homeomorphism u : RP1 → RP1, the boundary of the convex hull in AdS3 of its graph in RP1 × RP1 ≃ ∂AdS3 is the disjoint union of two embedded copies of the hyperbolic plane, pleated along measured geodesic laminations. Our main result is that any pair of bounded measured laminations that “strongly fill” can be obtained in this manner. Contents 1. Introduction and main results 1 1.1. Quasi-symmetric homeomorphisms 1 1.2. Measured laminations in the hyperbolic plane 2 1.3. Fixed points of compositions of earthquakes 2 1.4. The anti-de Sitter space and its boundary 3 1.5. Quasicircles in ∂AdS3 3 1.6. Bending laminations on the boundary of the convex hull 3 1.7. Related results 3 1.8. Examples and limitations 4 Acknowledgement 4 2. Backbground material. 4 2.1. Cross-ratios 4 2.2. Measured laminations. 4 2.3. On the geometry of the AdS3-space. 5 2.4. The Rhombus 7 2.5. The width of acausal meridians 7 3. Ideal polyhedra and approximation of laminations. 7 3.1. Ideal polyhedra with prescribed dihedral angles 7 3.2.
    [Show full text]
  • DIMENSION of QUASICIRCLES a Homeomorphism Φ of Planar
    DIMENSION OF QUASICIRCLES STANISLAV SMIRNOV Abstract. We introduce canonical antisymmetric quasiconformal maps, which minimize the quasiconformality constant among maps sending the unit circle to a given quasicircle. As an application we prove Astala’s conjecture that the Hausdorff dimension of a k-quasicircle is at most 1 + k2. A homeomorphism φ of planar domains is called k-quasiconformal, if it belongs locally to 1 the Sobolev class W2 and its Beltrami coefficient ¯ µφ(z):=∂φ(z)/∂φ(z), has bounded L∞ norm: kφk≤k. Equivalently one can demand that for almost every point z in the domain of definition directional derivatives satisfy max |∂αf(z)|≤K min |∂αf(z)| , α α then the constants of quasiconformality are related by K − 1 1+k k = ∈ [0, 1[,K= ∈ [1, ∞[. K +1 1 − k Quasiconformal maps send infinitesimal circles to ellipsis with eccentricities bounded by K, and they constitute an important generalization of conformal maps, which one obtains for K = 1 (or k = 0). While conformal maps preserve the Hausdorff dimension, quasiconformal maps can alter it. Understanding this phenomenon was a major challenge until the work [1] of Astala, where he obtained sharp estimates for the area and dimension distortion in terms of the quasiconformality constants. Particularly, the image of a set of Hausdorff dimension 1 under k-quasiconformal map can have dimension at most 1 + k, which can be attained for certain Cantor-type sets (like the Garnett-Ivanov square fractal). Nevertheless this work left open the question of the dimensional distortion of the subsets of smooth curves. Indeed, Astala’s [1] implies that a k-quasicircle (that is an image of a circle under a k-quasiconformal map) has the Hausdorff dimension at most 1+k, being an image of a 1-dimensional set.
    [Show full text]
  • THE IMPACT of RIEMANN's MAPPING THEOREM in the World
    THE IMPACT OF RIEMANN'S MAPPING THEOREM GRANT OWEN In the world of mathematics, scholars and academics have long sought to understand the work of Bernhard Riemann. Born in a humble Ger- man home, Riemann became one of the great mathematical minds of the 19th century. Evidence of his genius is reflected in the greater mathematical community by their naming 72 different mathematical terms after him. His contributions range from mathematical topics such as trigonometric series, birational geometry of algebraic curves, and differential equations to fields in physics and philosophy [3]. One of his contributions to mathematics, the Riemann Mapping Theorem, is among his most famous and widely studied theorems. This theorem played a role in the advancement of several other topics, including Rie- mann surfaces, topology, and geometry. As a result of its widespread application, it is worth studying not only the theorem itself, but how Riemann derived it and its impact on the work of mathematicians since its publication in 1851 [3]. Before we begin to discover how he derived his famous mapping the- orem, it is important to understand how Riemann's upbringing and education prepared him to make such a contribution in the world of mathematics. Prior to enrolling in university, Riemann was educated at home by his father and a tutor before enrolling in high school. While in school, Riemann did well in all subjects, which strengthened his knowl- edge of philosophy later in life, but was exceptional in mathematics. He enrolled at the University of G¨ottingen,where he learned from some of the best mathematicians in the world at that time.
    [Show full text]
  • The Measurable Riemann Mapping Theorem
    CHAPTER 1 The Measurable Riemann Mapping Theorem 1.1. Conformal structures on Riemann surfaces Throughout, “smooth” will always mean C 1. All surfaces are assumed to be smooth, oriented and without boundary. All diffeomorphisms are assumed to be smooth and orientation-preserving. It will be convenient for our purposes to do local computations involving metrics in the complex-variable notation. Let X be a Riemann surface and z x iy be a holomorphic local coordinate on X. The pair .x; y/ can be thoughtD C of as local coordinates for the underlying smooth surface. In these coordinates, a smooth Riemannian metric has the local form E dx2 2F dx dy G dy2; C C where E; F; G are smooth functions of .x; y/ satisfying E > 0, G > 0 and EG F 2 > 0. The associated inner product on each tangent space is given by @ @ @ @ a b ; c d Eac F .ad bc/ Gbd @x C @y @x C @y D C C C Äc a b ; D d where the symmetric positive definite matrix ÄEF (1.1) D FG represents in the basis @ ; @ . In particular, the length of a tangent vector is f @x @y g given by @ @ p a b Ea2 2F ab Gb2: @x C @y D C C Define two local sections of the complexified cotangent bundle T X C by ˝ dz dx i dy WD C dz dx i dy: N WD 2 1 The Measurable Riemann Mapping Theorem These form a basis for each complexified cotangent space. The local sections @ 1 Â @ @ Ã i @z WD 2 @x @y @ 1 Â @ @ Ã i @z WD 2 @x C @y N of the complexified tangent bundle TX C will form the dual basis at each point.
    [Show full text]
  • The Riemann Mapping Theorem
    University of Toronto – MAT334H1-F – LEC0101 Complex Variables 18 – The Riemann mapping theorem Jean-Baptiste Campesato December 2nd, 2020 and December 4th, 2020 Theorem 1 (The Riemann mapping theorem). Let 푈 ⊊ ℂ be a simply connected open subset which is not ℂ. −1 Then there exists a biholomorphism 푓 ∶ 푈 → 퐷1(0) (i.e. 푓 is holomorphic, bijective and 푓 is holomorphic). We say that 푈 and 퐷1(0) are conformally equivalent. Remark 2. Note that if 푓 ∶ 푈 → 푉 is bijective and holomorphic then 푓 −1 is holomorphic too. Indeed, we proved that if 푓 is injective and holomorphic then 푓 ′ never vanishes (Nov 30). Then we can conclude using the inverse function theorem. Note that this remark is false for ℝ-differentiability: define 푓 ∶ ℝ → ℝ by 푓(푥) = 푥3 then 푓 ′(0) = 0 and 푓 −1(푥) = √3 푥 is not differentiable at 0. Remark 3. The theorem is false if 푈 = ℂ. Indeed, by Liouville’s theorem, if 푓 ∶ ℂ → 퐷1(0) is holomorphic then it is constant (as a bounded entire function), so it can’t be bijective. The Riemann mapping theorem states that up to biholomorphic transformations, the unit disk is a model for open simply connected sets which are not ℂ. Otherwise stated, up to a biholomorphic transformation, there are only two open simply connected sets: 퐷1(0) and ℂ. Formally: Corollary 4. Let 푈, 푉 ⊊ ℂ be two simply connected open subsets, none of which is ℂ. Then there exists a biholomorphism 푓 ∶ 푈 → 푉 (i.e. 푓 is holomorphic, bijective and 푓 −1 is holomorphic).
    [Show full text]
  • Math 311: Complex Analysis — Automorphism Groups Lecture
    MATH 311: COMPLEX ANALYSIS | AUTOMORPHISM GROUPS LECTURE 1. Introduction Rather than study individual examples of conformal mappings one at a time, we now want to study families of conformal mappings. Ensembles of conformal mappings naturally carry group structures. 2. Automorphisms of the Plane The automorphism group of the complex plane is Aut(C) = fanalytic bijections f : C −! Cg: Any automorphism of the plane must be conformal, for if f 0(z) = 0 for some z then f takes the value f(z) with multiplicity n > 1, and so by the Local Mapping Theorem it is n-to-1 near z, impossible since f is an automorphism. By a problem on the midterm, we know the form of such automorphisms: they are f(z) = az + b; a; b 2 C; a 6= 0: This description of such functions one at a time loses track of the group structure. If f(z) = az + b and g(z) = a0z + b0 then (f ◦ g)(z) = aa0z + (ab0 + b); f −1(z) = a−1z − a−1b: But these formulas are not very illuminating. For a better picture of the automor- phism group, represent each automorphism by a 2-by-2 complex matrix, a b (1) f(z) = ax + b ! : 0 1 Then the matrix calculations a b a0 b0 aa0 ab0 + b = ; 0 1 0 1 0 1 −1 a b a−1 −a−1b = 0 1 0 1 naturally encode the formulas for composing and inverting automorphisms of the plane. With this in mind, define the parabolic group of 2-by-2 complex matrices, a b P = : a; b 2 ; a 6= 0 : 0 1 C Then the correspondence (1) is a natural group isomorphism, Aut(C) =∼ P: 1 2 MATH 311: COMPLEX ANALYSIS | AUTOMORPHISM GROUPS LECTURE Two subgroups of the parabolic subgroup are its Levi component a 0 M = : a 2 ; a 6= 0 ; 0 1 C describing the dilations f(z) = ax, and its unipotent radical 1 b N = : b 2 ; 0 1 C describing the translations f(z) = z + b.
    [Show full text]
  • Faber and Grunsky Operators Corresponding to Bordered Riemann Surfaces
    CONFORMAL GEOMETRY AND DYNAMICS An Electronic Journal of the American Mathematical Society Volume 24, Pages 177–201 (September 16, 2020) https://doi.org/10.1090/ecgd/355 FABER AND GRUNSKY OPERATORS CORRESPONDING TO BORDERED RIEMANN SURFACES MOHAMMAD SHIRAZI Abstract. Let R be a compact Riemann surface of finite genus g > 0and let Σ be the subsurface obtained by removing n ≥ 1 simply connected regions + + Ω1 ,...,Ωn from R with non-overlapping closures. Fix a biholomorphism fk + from the unit disc onto Ωk for each k and let f =(f1,...,fn). We assign a Faber and a Grunsky operator to R and f when all the boundary curves of Σ are quasicircles in R. We show that the Faber operator is a bounded isomorphism and the norm of the Grunsky operator is strictly less than one for this choice of boundary curves. A characterization of the pull-back of the holomorphic Dirichlet space of Σ in terms of the graph of the Grunsky operator is provided. 1. Introduction and main results Let C denote the complex plane, let D denote the open unit disc in C, and let C = C ∪{∞}denote the Riemann sphere. Let also D− = {z ∈ C : |z| > 1}∪{∞}, and let S1 be the unit circle in C.Weuseg for the genus of a Riemann surface to distinguish it from its Green’s function g. Let Σ be a bordered Riemann surface (in the sense of L. V. Ahlfors and L. Sario [1, II. 3A]) obtained from a compact Riemann surface R of genus g ≥ 0from ≥ + + which n 1 simply connected domains Ω1 ,...,Ωn , with non-overlapping closures removed.
    [Show full text]
  • On the Scaling Ratios for Siegel Disks
    Commun. Math. Phys. 333, 931–957 (2015) Communications in Digital Object Identifier (DOI) 10.1007/s00220-014-2181-z Mathematical Physics On the Scaling Ratios for Siegel Disks Denis Gaidashev Department of Mathematics, Uppsala University, Uppsala, Sweden. E-mail: [email protected] Received: 10 September 2013 / Accepted: 17 July 2014 Published online: 21 October 2014 – © Springer-Verlag Berlin Heidelberg 2014 Abstract: The boundary of the Siegel disk of a quadratic polynomial with an irrationally indifferent fixed point and the rotation number whose continued fraction expansion is preperiodic has been observed to be self-similar with a certain scaling ratio. The restriction of the dynamics of the quadratic polynomial to the boundary of the Siegel disk is known to be quasisymmetrically conjugate to the rigid rotation with the same rotation number. The geometry of this self-similarity is universal for a large class of holomorphic maps. A renormalization explanation of this universality has been proposed in the literature. In this paper we provide an estimate on the quasisymmetric constant of the conjugacy, and use it to prove bounds on the scaling ratio λ of the form αγ ≤|λ|≤Cδs, where s is the period of the continued fraction, and α ∈ (0, 1) depends on the rotation number in an explicit way, while C > 1, δ ∈ (0, 1) and γ ∈ (0, 1) depend only on the maximum of the integers in the continued fraction expansion of the rotation number. Contents 1. Introduction ................................. 932 2. Preliminaries ................................. 936 3. Statement of Results ............................. 944 4. An Upper Bound on the Scaling Ratio .................... 945 5.
    [Show full text]
  • G(*)
    transactions of the american mathematical society Volume 282. Number 2. April 1984 JORDAN DOMAINS AND THE UNIVERSALTEICHMÜLLER SPACE BY Abstract. Let L denote the lower half plane and let B(L) denote the Banach space of analytic functions/in L with ||/||; < oo, where ||/||/ is the suprenum over z G L of the values |/(z)|(Im z)2. The universal Teichmüller space, T, is the subset of B(L) consisting of the Schwarzian derivatives of conformai mappings of L which have quasiconformal extensions to the extended plane. We denote by J the set {Sf-. /is conformai in L and f(L) is a Jordan domain} , which is a subset of B(L) contained in the Schwarzian space S. In showing S — T ^ 0, Gehring actually proves S — J ¥= 0. We give an example which demonstrates that J — T ¥= 0. 1. Introduction. If D is a simply connected domain of hyperbolic type in C, then the hyperbolic metric in D is given by Pfl(z)=^% ^D, l-|g(*)l where g is any conformai mapping of D onto the unit disk A = {z: \z |< 1}. If /is a locally univalent meromorphic function in D, the Schwarzian derivative of /is given by /")' WZ")2 5/ \f I 2 \ /' at finite points of D which are not poles of /. The definition of Sf is extended to all of D by means of inversions. We let B(D) denote the Banach space of Schwarzian derivatives of all such functions / in a fixed domain D for which the norm ||5/||û = sup|5/(z)|pD(z)-2 zSD is finite.
    [Show full text]