Covering Spaces for Domains in the Complex Plane

Hjörtur Björnsson

FacultyFaculty of of Physical Physical Sciences Sciences UniversityUniversity of of Iceland Iceland 20202020

COVERING SPACES FOR DOMAINS IN THE COMPLEX PLANE

Hjörtur Björnsson

60 ECTS thesis submitted in partial fulfillment of a Magister Scientiarum degree in Mathematics

Advisor Reynir Axelsson

Faculty Representative Ragnar Sigurðsson

M.Sc. committee Reynir Axelsson Jón Ingólfur Magnússon

Faculty of Physical Sciences School of Engineering and Natural Sciences University of Iceland Reykjavík, May 2020 Covering Spaces for Domains in the Complex Plane 60 ECTS thesis submitted in partial fulfillment of a M.Sc. degree in Mathematics

Copyright c 2020 Hjörtur Björnsson All rights reserved

Faculty of Physical Sciences School of Engineering and Natural Sciences University of Iceland Dunhagi 5 IS-107, Reykjavík, Reykjavík Iceland

Telephone: 525 4000

Bibliographic information: Hjörtur Björnsson, 2020, Covering Spaces for Domains in the Complex Plane, M.Sc. thesis, Faculty of Physical Sciences, University of Iceland.

Printing: Háskólaprent, Fálkagata 2, 107 Reykjavík Reykjavík, Iceland, May 2020 For Stefania

Abstract

In this thesis we present a modern version of a proof published by Tibor Radó in 1922, which shows that the universal covering space for a domain G in C\{0, 1} is bi- holomorphically equivalent to the . Radó uses multivalued functions which we replace by mappings on appropriate covering spaces. We use Riemann domains for this purpose, which are essentially Riemann surfaces with global coordinates.

Útdráttur

Í þessari ritgerð fjöllum við um nútímalega útgáfu af sönnun sem Tibor Radó gaf út árið 1922. Hún sýnir að allsherjarþekjurúm sérhvers svæðis G í C \{0, 1} er fágað einsmóta einingarhringskífunni. Radó notaðist við marggild föll í sinni grein, en við skoðum í staðinn varpanir á viðeigandi þekjurúmum. Við notumst við Riemann-svæði, en þau eru sértilfelli af almennum Riemann-flötum, þar sem við höfum víðfeðm hnitaföll.

v

Contents

Þakkarorð ix

1. Introduction1 1.1. Background ...... 2 1.1.1. The Uniformization Theorem ...... 2 1.1.2. Parametrization of Algebraic Curves ...... 3 1.1.3. Classification of Riemann Surfaces ...... 4 1.2. Proof of the ...... 6 1.3. Picard’s Little Theorem ...... 8

2. Preliminaries9 2.1. Topological Preliminaries ...... 9 2.1.1. Local Homeomorphisms and Sections ...... 9 2.1.2. Chains and Step Polygons ...... 11 2.2. Riemann Domains ...... 19

3. Uniformization Theorem for Domains in C 25 3.1. The Set Σ(G, y0) ...... 26 3.2. Constructing a Surjective Function ...... 29

A. Appendix 33 A.1. Path Homotopy and the Fundamental Group ...... 33 A.2. Covering Spaces ...... 35

Bibliography 37

vii

Þakkarorð

Ég vil þakka Reyni Axelssyni og Jón Ingólfi Magnússyni, fyrir alla þá hjálp sem þeir hafa gefið mér við vinnslu þessarar ritgerðar. Ég vil einnig sérstaklega þakka Sigurði Frey Hafstein fyrir þann mikla stuðning og þolinmæði sem hann hefur sýnt mér í gegnum árin. Síðast en ekki síst vil ég þakka Stefaniu Crotti, grazie di cuore.

ix

1. Introduction

In 1922 Radó published a proof of the famous Riemann mapping theorem [10], by L.Fejér and F.Riesz.

Theorem 1.0.1 (Riemann mapping theorem). Let G ⊂ C be a simply connected domain which is not all of C. Then there exists a biholomorphic map f : G → E, where E is the open unit disk {z ∈ C | |z| < 1}. Furthermore we can, for any p ∈ G, choose f in such a way that f(p) = 0, and the map f is then uniquely determined up to a rotation.

Their approach was a new one: they showed that the Riemann mapping theorem could be obtained as a solution of an extremal problem. The argument was very short, only about one full page. Many modern textbooks that contain a proof of the Riemann mapping theorem use similar methods, although modernized. The proof is presented in section 1.2.

In his paper, Radó goes on to prove the following theorem:

Theorem 1.0.2. Let G be a domain in C, such that the complement of G contains at least two points. Then there exists a holomorphic covering map p: E → G.

Radó uses a similar idea as Fejér and Riesz used to prove Theorem 1.0.1. He starts by considering a certain class G of multivalued holomorphic functions from G to E which is defined in such a way the inverse of any function f ∈ G is a well defined single valued from the image of f in E to G. By picking a specific solution to an extremal problem Radó manages to construct a multivalued holomorphic surjection f : G → E, whose inverse is a well defined holomorphic covering map f −1 : E → G.

In this thesis we present a more modern version of Radó’s proof. Instead of multival- ued functions on G, we consider well defined holomorphic functions on appropriate covering spaces of G, satisfying the condition that they are injective with respect to the covering map (see Definition 2.2.4). This condition guarantees that a certain

1 1. Introduction

‘inverse’ of these functions is a well defined holomorphic map (see Lemma 2.2.5). Using the same idea as Radó, we prove Theorem 3.0.1, which says that for any domain G in C, there exists a holomorphic covering map p: Y → G, where Y is biholomorphically equivalent to either C or E. The end result is an argument pleas- antly similar to a method used to prove the Riemann mapping theorem in many modern textbooks, albeit much longer.

In section 1.1 we give a brief exposition on the Uniformization theorem and its consequences, and how it relates to our thesis. In section 1.2 we present Radó’s original argument to prove the Riemann mapping theorem, and we end this chapter by showing in section 1.3 how Picard’s little theorem is a simple corollary of the main result of this thesis. Chapter 2 presents some topological preliminaries and our main tool, the Riemann domains, and finally, in chapter 3 we have the main proof of our thesis. We have also included appendix A which contains some basic definitions and lemmas on homotopy and covering spaces.

1.1. Background

The main theorem of this thesis, Theorem 3.0.1, mentioned in the introduction, can be viewed as a special case of the famous Uniformization theorem. The theorem, stated in the next section deals with general Riemann surfaces, but in this thesis we are working with domains of C. We give an overview of the theorem and its consequences with simplified steps to reduce unnecessary complexities that would be outside the scope of this thesis.

1.1.1. The Uniformization Theorem

The Uniformization theorem is simple to state, but it took over a century for math- ematicians to formulate it and give a convincing proof. It wasn’t until the year 1907 that Henri Poincaré and Paul Koebe gave what is considered the first proofs of the theorem. The book "Uniformization of Riemann Surfaces: Revisiting a hundred- year-old theorem"[3] gives a detailed account of the very interesting history of the Uniformization theorem and presents proofs of various versions of it, up to the final form stated here below.

Theorem 1.1.1 (Uniformization theorem). Every simply connected Riemann sur- face is biholomorphically equivalent to the , the complex plane, or the unit disc.

2 1.1. Background

The modern version of the Uniformization theorem is a classification result, but historically uniformization is about parametrization. The theorem arises from the study of algebraic curves and differential equations. In the next two subsections we give a brief description of parametrization of algebraic curves as a motivation, and classification of Riemann surfaces by the Uniformization theorem.

1.1.2. Parametrization of Algebraic Curves

To explain the connection between the Uniformization theorem and parametrization, we give a simplified description of a certain class of algebraic curves called elliptic curves. An elliptic curve is a cubic polynomial of the form y2 = x3 + ax + b where a,b are complex numbers such that 4a3 + 27b2 6= 0. We want to parametrize this elliptic curve, and for the purpose let us assume that there exists a function z 7→ (x(z), y(z)) = (f(z), f 0(z)). Then f must satisfy the differential equation (f 0)2 = f 3 + af + b. We can ‘solve’ this differential equation formally with the integrals Z df Z dx z = = √ . pf 3 + af + b x3 + ax + b This type of integral is know as an elliptic integral, since it arises when evaluating the arc length of ellipses. The integral depends not only on which branch of the square root we use, but also on the path of integration. In this sense we can regard z as multivalued function of x. The key idea to obtain the parametrization is that we look at the inverse of this multivalued function, which, as turns out, gives us x as a single valued periodic function of z, and therefore we have parametrized the elliptic curve. The above calculations can be made precise [6, Ch. 10.6].

This idea brings us to the topic of doubly periodic meromorphic functions. One example is the Weierstrass ℘ function which is meromorphic on C with periods 1 and i, that is to say ℘(z + 1) = ℘(z) and ℘(z + i) = ℘(z) for all z, and has double poles at every point on the lattice Λ = {m + in | m, n ∈ Z}, see [6, Ch. 10.6]. This function is given by the series 1 X  1 1  ℘(z) = + − , z2 (z − m − ni)2 (m + ni)2 m,n∈Z (m,n)6=(0,0) and there are complex valued constants C1,C2 ∈ C such that 0 2 3 (℘ (z)) = 4 (℘(z)) − C1℘(z) − C2

3 1. Introduction for all z ∈ C.

The map (℘, ℘0): C → S then gives us a parametrization of the elliptic curve S 2 3 where S is the curve given by y = 4x − C1x − C2 in the complex projective plane, and we get an identification of C/Λ and S.

Now the constants C1 and C2 above depend on the period lattice Λ, and by choosing a different period lattice we obtain a parametrization for a different elliptic curve.

We can think of these elliptic curves where we allow x and y to take complex values as complex tori. Now if Γ and Γ0 are two lattices in C, then the complex tori C/Γ and C/Γ0 are topologically homeomorphic, but there is no biholomorphic equivalence between them unless there exists an α ∈ C \{0} such that Γ = αΓ0. However the universal covering space for a complex torus is C, so we get a meromorphic parametrization of the elliptic curve from the complex plane C.

We have uniformized the parameter, in the sense that every elliptic curve can be parametrized by a function on C, even though the underlying complex tori are not biholomorphically equivalent.

Every algebraic curve defines by desingularization a compact , see [4, Thm. 8.9], and every compact Riemann surface can be classified topologically by its topological genus, and therefore we can classify algebraic curves by the same genus. It can be shown that curves of genus 0 are birationally equivalent to the Riemann sphere, and are therefore rational curves. Curves of genus 1 are equivalent to a complex torus and are elliptic. All other curves, which are of genus ≥ 2, can be parametrized by a function on the unit disk E by the Uniformization theorem, and are called hyperbolic.

1.1.3. Classification of Riemann Surfaces

Let X be a Riemann surface and p: X˜ → X be its universal covering. Then G = Deck(X/X˜ ) is defined as the subgroup of the holomorphic automorphism group Aut(X˜), consisting of all f such that p ◦ f = p. Then the group G acts without fixed points and discretely on X˜, that is if σ ∈ G \{id} then σx 6= x for all x ∈ X˜ and the orbit Gx = {σx | σ ∈ G} is a discrete subset of X˜ for every x ∈ X˜. This can be seen from the fact that a covering transformation is uniquely determined by the image of any point, and the orbit Gx is equal to the fiber p−1[p(x)], furthermore we have that G is isomorphic to the fundamental group π1(X, x) for any x ∈ X.

Since X˜ is simply connected, then by the Uniformization theorem it is biholomor- phically equivalent to the Riemann sphere C, the complex plane C or open unit disk

4 1.1. Background

E, and we get a classification of all Riemann surfaces, by their universal covering spaces.

By Theorem [6, Thm. 6.3.5] every automorphism of C is a Möbius transformation of the form az + b φ(z) = a, b, c, d ∈ , ad − bc 6= 0. cz + d C

Every Möbius transformation of this form φ: C → C has at least one fixed point. So, if X˜ is biholomorphically equivalent to the Riemann sphere C, then G = Deck(X/X˜ ) = {id} and X is simply connected. Therefore p: X˜ → X is a bi- holomorphism and X is biholomorphically equivalent to the Riemann sphere C.

Every automorphism of C is of the form:

φ(z) = az + b, a ∈ C \{0}, b ∈ C, and if a 6= 1 this transformation has a fixed point. Thus a subgroup G of Aut(C) that acts discretely and without fixed points consists only of translations φ(z) = z + b. Let Γ be the orbit of zero under G. Then Γ is a discrete additive subgroup of C, and G is the subgroup of all translations z 7→ z + b with b ∈ Γ. Then by [4, Thm. 27.11], one of the following is true:

(i) G = {id}.

(ii) G a cyclic group of all translations of the form z 7→ z + nb, for all n ∈ Z, where b 6= 0 is a fixed complex number.

(iii) G consists of all translations of the form z 7→ z + nb1 + mb2, for all n, m ∈ Z, where b1, b2 are two fixed complex numbers, linearly independent over R.

If X˜ is biholomorphically equivalent to C then G = Deck(X/X˜ ) is one of the three types of groups above. For (i), we get that X is biholomorphic to C, in case (ii) we 2πi  get that the covering is isomorphic to C → C \{0}, z 7→ exp b . Finally for the third case X is biholomorphic to a complex torus C/Γ.

The set of automorphisms of the unit disk E are transformations of the form [6, Thm. 6.2.3]: z − a φ(z) = ω , a, ω ∈ , |a| < 1, |ω| = 1. 1 − az C The unit disk E is biholomorphic to the upper half plane, via the (inverse) Cayley z−i transform f(z) = z+i , which maps H = {z ∈ C | Im(z) > 0} to the unit disk E. Therefore we can work in H instead of E and Aut(H) can be described as the set

5 1. Introduction of all transformations of the form [6, Ch. 10.5]: az + b φ(z) = a, b, c, d ∈ , ad − bc > 0. cz + d R The structure of subgroups in Aut(H) that act discretely and without fixed points is very rich, so we will not go into details here. One example of such a group is the subgroup of translations of the form φ(z) = z + nb, with n integer and b some fixed real number.

If X is a Riemann surface and its universal covering space p: X˜ → X is not biholo- morphic to C or C, then X˜ is biholomorphic to E, and there exists a biholomorphic covering map π : E → X.

Riemann surfaces are called elliptic, parabolic or hyperbolic when their universal covering is isomorphic to C, C, or E respectively, see [4, thm 27.10]. From the preceding discussion we see that the Riemann sphere C is the only elliptic Riemann surface. The complex plane C, the punctured plane C\{0} and all complex tori C/Γ are parabolic, and every other Riemann surface is hyperbolic. Calling the Riemann sphere elliptic can cause some confusion as, by the classification of the algebraic curves, complex tori are sometimes called elliptic curves.

For the purpose of this thesis we are working with domains G in C \{0, 1}, which by the main theorem of this thesis, Theorem 3.0.1, can be covered by the unit disk E and are therefore hyperbolic.

1.2. Proof of the Riemann Mapping Theorem

In his paper Radó only deals with the case when G is bounded. The general case then follows with a standard argument which is written in detail in many textbooks, see e.g. [1, Thm. VI.5.1]. We thus assume that G is bounded, fix a point p ∈ G, and define the class

F = {g ∈ O(G) | g is bounded and injective, g(p) = 0 and g0(p) = 1}.

By translating G we may take p as the origin 0. The set F is not empty, as the function G → G, z 7→ z, is in F. Consider the uniform norm given by g 7→ |g|G = supz∈G |g(z)|. By definition of F, we have |g|G < ∞ for every g ∈ F. Define the constant ρ = inf{|g|G | g ∈ F}, then there exists a sequence of functions (ψn) in F such that |ψn|G converges to ρ. The sequence (ψn) is bounded on G, so by Montel’s theorem, see [1, Thm. IV.1.5] or Theorem 2.2.8, there exists a subsequence, also

6 1.2. Proof of the Riemann Mapping Theorem

denoted by (ψn), that converges uniformly on compact subsets of G to a holomorphic function ψ.

0 As ψk(0) = 0 and ψk(0) = 1 for all k ∈ N, and since the sequence converges uniformly on compact sets, it follows that ψ(0) = 0 and ψ0(0) = 1. Therefore ψ is not constant on G and |ψ|G = ρ > 0. Since ψ is holomorphic, and a limit of injective holomorphic functions, then ψ is also injective, and therefore ψ ∈ F. In fact, if ψ is not injective, then there are two distinct points x, y ∈ G such that ψ(x) = ψ(y), and by Hurwitz’s theorem, see Theorem 2.2.10, there are distinct points xk, yk ∈ G such that ψk(xk) = ψk(yk) for k large enough, which is a contradiction as fk is injective.

Since |ψ|G = ρ then the image ψ(G) is contained in ρE, where E is the open unit disk. The only thing left to show is that ψ : G → ρE is a surjective function.

Assume that ψ is not surjective onto ρE, then there is a complex number reiφ ∈ ρE, with 0 < r < ρ, which is not included in the image of ψ. We use what Radó calls the Carathéodory-Koebe square root transformation to construct a new function F ∈ F such that |F |G < ρ, which is a contradiction. Therefore ψ must be a biholomorphic map ψ : G → ρE.

r iφ z−w Let β = ρ < 1 and set w = βe . Denote by hw : E → E, the function hw(z) = zw−1 ,  1  then hw(0) = w, hw(w) = 0, and the map z 7→ hw ρ ψ(z) does not take the value 0. Since G is a simply connected domain there exists a l : G → E such that l2(z) = 1 √ 1 i 2 φ hw( ρ ψ(z)) and l(0) = w1 where w1 = βe . Using one more transformation z−w1 hw : → , hw (z) = , we can define a map g : G → by 1 E E 1 zw1−1 E

1 i 2 φ g(z) = e hw1 (l(z)).

We see, by definitions of the transformations hw and hw1 , that g(0) = 0, and β + 1 1 g0(0) = √ ψ0(0). 2 β ρ Set √ 2 β F (z) = ρg(z), β + 1 0 0 then F (0) = 0 and F (0) = ψ (0) = 1, and since all of the maps ψ, hw, l, hw1 and√ g 2 β are injective and bounded, we have that F ∈ F. Since |g|G ≤ 1, then |F |G ≤ ρ, √ β+1 2 β but since 0 < β < 1 we have that β+1 < 1, and therefore |F |G < ρ, which is a contradiction.

By scaling the function ψ obtained above we get a biholomorphic map f : G → E, f = ψ/ρ. Full calculations of the derivatives above, and more discussion on the properties of the maps hw can be found in Definition 3.1.4 and Lemma 3.1.5.

7 1. Introduction

Now it only remains to prove uniqueness. Let f be the biholomorphic map f : G → E constructed above, and let g be any biholomorphic map g : G → E such that g(0) = 0. Then λ = f ◦ g−1 : E → E is an automorphism of E such that λ(0) = 0, and by Schwarz’s lemma, see Lemma 2.2.9, f ◦ g−1(z) = az with |a| = 1. Therefore f(z) = ag(z), and g is just a rotation of f.

1.3. Picard’s Little Theorem

The main result of this paper, Theorem 3.0.2 is a proof of the fact that the universal covering space of a domain G in C, of which the complement contains at least two points, is the unit disk E. This result is just a very special case of the general Uniformization theorem for abstract Riemann surfaces, and the full general theorem needs much more theory to be proven.

One of the corollaries of this special case is Picard’s little theorem, and we obtain a very short and simple proof.

Theorem 1.3.1 (Picard’s little theorem). If a function f : C → C is holomorphic and not constant, then the image of f is either the whole of C or C minus a single point.

Let G denote the image of f and assume that the complement C \ G contains at least two distinct points, then we can lift f to a holomorphic function g : C → E, since p: E → G is the universal covering space of G, by Theorem 3.0.1, but g is a bounded entire function and therefore constant, so f = p ◦ g must be constant.

8 2. Preliminaries

In this chapter we present some well known topological lemmas and constructions, with the intention of proving Theorems 2.1.17 and 2.1.18. They state that given a space G and a universal cover p : Y → G, then a certain subspace of Y is the universal cover for a specific subspace of G. We also define Riemann domains, which serve as our primary tools in the proof of the Uniformization theorem for domains in C (see Lemma 3.0.1). We present Hurwitz’s theorem in 2.2.10 and Montel’s theorem for Riemann domains in 2.2.8, as they are used in later chapters.

2.1. Topological Preliminaries

2.1.1. Local Homeomorphisms and Sections

Definition 2.1.1. Let U, V be Hausdorff spaces, p: U → V be a continuous map, and W be a subset of V . Then a section of p over W is a continuous map σ : W → U such that (p ◦ σ)(x) = x for all x in W .

The following are well known lemmas on sections of local homeomorphisms, for a proof see [5, VII, prop 1.4-6].

Lemma 2.1.2. Let U and V be as in Definition 2.1.1, and let p: U → V be a local homeomorphism

(i) For any point v ∈ V and u ∈ p−1[v] there exists a neighbourhood W around v and a section σ : W → U such that σ(v) = u.

(ii) A section σ : W → U of p induces a homeomorphism W → σ(W ), x 7→ σ(x).

(iii) Let W be a connected subset of V and σ1, σ2 : W → U be two sections of p. If σ1 and σ2 agree at a single point in W then they agree on the whole of W , that is to say σ1 = σ2.

9 2. Preliminaries

Lemma 2.1.3. Let Y,Z,G be Hausdorff spaces and p: Y → G and q : Z → G be two local homeomorphisms and σ be a section p. Furthermore assume that f : Y → Z is a continuous map such that p = q◦f, that is to say the following diagram commutes.

f Y > Z

p q

∨ < G

Then f ◦ σ is a section of q.

Proof. Assume that the section σ is defined on a subset U of G. Then

q ◦ (f ◦ σ) = (q ◦ f) ◦ σ = p ◦ σ = idU so f ◦ σ is a section of q.

Lemma 2.1.4. Let Y,Z,G be connected and locally connected Hausdorff spaces. Let p: Y → G be a covering map, f : Y → Z surjective, and q : Z → G be a local homeomorphism such that p = q ◦ f; then f and q are covering maps.

Proof. Fix a point x ∈ G and let U be a connected neighbourhood of x evenly covered by p. Then there exists a set I and a family of sections σi : U → Y , i ∈ I such that −1 [ p [U] = σi(U) i∈I where the sets on the right hand side are pairwise disjoint. For each i ∈ I the map f ◦ σi : U → Z is a section of q, by Lemma 2.1.3. Put Vi = f ◦ σi(U), for all i ∈ I. Now for any i, j in the set I, we have that either Vi = Vj or Vi ∩ Vj = ∅, since two sections that agree at a point agree on the whole of U, by Lemma 2.1.2(iii). Therefore there exists a set J ⊂ I, such that [ [ f ◦ σi(U) = f ◦ σi(U) i∈I i∈J where the sets on the right hand side are all pairwise disjoint. It remains to show that −1 [ q [U] = f ◦ σi(U). i∈J

10 2.1. Topological Preliminaries

Let z ∈ q−1[U], since f is surjective there exists a y ∈ Y with f(y) = z. But then p(y) ∈ U so y ∈ σi(U) for some i ∈ I, therefore [ [ z ∈ f ◦ σi(U) ⊂ f ◦ σj(U) = f ◦ σj(U). j∈I j∈J Then q is a covering map and it is well known that if f is a surjective continuous fiber preserving map between covering maps, then f is also a covering map, see for example [5, Ch. VII, prop. 4.3].

2.1.2. Chains and Step Polygons

The goal of this section is to prove Lemma 2.1.16 and Theorem 2.1.18, and for that purpose we introduce the notion of chains and step polygons. The topic is covered in depth in in [1, Ch. I] and [11, Ch. 12].

Definition 2.1.5. A path in an open domain U ⊂ C is a continuous function τ from the interval [0, 1] ⊂ R to U. A closed path τ in U is a path such that τ(0) = τ(1). We denote the image of the path τ : [0, 1] → U by |τ|, |τ| = {τ(x) | x ∈ [0, 1]}.

Let U be a domain in C, we denote by C1(U) the free abelian group generated by all paths in U.

Definition 2.1.6. A chain γ in a domain U of C is an element γ in the free abelian group C1(U). We can write γ as the finite sum

γ = a1τ1 + ··· + anτn, ai ∈ Z where each of the τi, 1 ≤ i ≤ n is a path in U.

Denote by C0(U) the free abelian group generated by all points u ∈ U. For each path τ in U we define the boundary ∂0(τ) ∈ C0(U), by ∂0(τ) = τ(1) − τ(0). Note that the binary operation here is subtraction in the group C0(U) not the normal subtraction in C. A path τ is closed if, and only if, ∂0(τ) = 0 ∈ C0(U).

By extending ∂0 linearly, a well defined group homomorphism ∂ : C1 → C0 is obtained. It is given by

n ! n X X ∂(γ) = ∂ ajτj = aj∂0(τj) ∈ C0(U). j=1 j=1

11 2. Preliminaries where aj ∈ Z and each τj is a path in U.

We say a chain γ is closed if its boundary ∂(γ) is 0 ∈ C0(U). In other words, let the Pn closed chain γ be given by the sum j=1 ajτj. Then each starting point and end point of paths τj in γ appear equally often and with opposite signs, counted with multiplicity.

Pn We define the support of the chain γ = j=1 ajτj by

n ! n X [ supp(γ) = supp ajτj = |τj|,

j=1 j=1,aj 6=0 which is simply the union of the images of all the closed path τj, and it is a compact set.

For any point p 6∈ supp(γ) we can define the winding number of a closed chain γ around p, 1 Z dz I(γ, p) = 2πi γ z − p see e.g. [1, Def. 1.14] or [11, Ch. 12.4]. The winding number I(γ, p) is an integer for all p ∈ C \ supp(γ) and the mapping C \ supp(γ) → Z, p 7→ I(γ, p), is constant on every connected component of C \ supp(γ).

We can furthermore define the interior and exterior of the closed chain γ as the open sets

I(γ) = {z ∈ C \ supp(γ) | I(γ, z) 6= 0},E(γ) = {z ∈ C \ supp(γ) | I(γ, z) = 0} respectively.

Definition 2.1.7. Let p1, p2 be points in C, then the segment [p1, p2] is defined as the continuous map [0, 1] → C, t 7→ (1−t)p1 +tp2. A step polygon is a closed chain γ composed of k segments, [p1, p2]+[p2, p3]+···+[pk, p1], where each of the segments is parallel to either the real or the imaginary axis in C, each of the points p1, p2, . . . , pk are pairwise distinct, and the images of the segments are pairwise disjoint except for the starting and ending points of the segments.

A step polygon γ = [p1, p2]+...+[pk, p1] can be regarded as a closed path τ : [0, 1] → C by dividing the interval [0, 1] into k equal intervals [ti, ti+1] with t1 = 0 and tk+1 = 1, and mapping the intervals [ti, ti+1] onto the segment [pi, pi+1] and [tk, tk+1] onto [pk, p1]. The closed path τ restricted to the half open interval [0, 1) is then injective.

We have the following three lemmas for step polygons, for detailed proofs see [11, Ch. 12.4], and [1, Thm. I.2.7].

12 2.1. Topological Preliminaries

Lemma 2.1.8. Every step polygon γ divides C into exactly two domains, C \ supp(γ) = I(γ) ∪ E(γ), and the winding number of γ is either 1 everywhere or −1 everywhere on the domain I(γ), and 0 on E(γ).

Lemma 2.1.9. Let U be an open subset of C and K connected compact subset of U. Then there exists a closed chain τ = τ1 + . . . τN in U \ K, where each τi is a step polygon, such that K ⊂ I(τ) ⊂ U.

The previous lemma is telling us that we can ‘go around’ the compact set K, in U. The idea for the proof is simple√ and we go through some of the details here. Pick a δ > 0 such that d(K, ∂U) > 2δ, where d(A, B) = inf{|a − b| | a ∈ A, b ∈ B}. We can now tile C with closed squares where each side has length δ and is parallel to either the real or the imaginary axis. Since K is a compact set, it intersects only finitely many of these squares. Let us denote them by R1,R2,...,Rk, each of those squares intersects K so, by our choice of δ, they are all included in U, and since Sk j j they cover K we have K ⊂ j=1 R ⊂ U. Let σ(R ) be the sum of the line segments j j j j Rr,Ru,Rl ,Rd where they denote the right, upper, left and lower sides of the square j j Pk j R traversed counter-clockwise around R . By considering the sum γ = j=1 σ(R ) we see that γ is a closed chain. Let c be any point in K, then c intersects some square Rl. If c is included in the interior of Rl then I(σ(Rl), c) = 1 and I(σ(Rk), c) = 0 for k 6= l, therefore I(γ, c) = 1.

If K intersects a line segment of σ(Rj) for some j, then that line segment is a common side for another square Rn, and it appears in σ(Rn) with opposite orientation. By removing sides of opposite orientation we obtain another chain τ = τ1 +τ2 +···+τN such that supp(τ) ⊂ U \ K, which is equivalent to γ, and such that each of the τi is a step polygon, see [11, Thm. 12.4.1].

Equivalence here means that Z Z f(z)dz = f(z)dz γ τ for every continuous function defined on supp(γ) ∪ supp(τ). This tells us that for a constant c ∈ K included in the interior of Rl, like above, we have that I(τ, c) = I(γ, c) = 1. By continuity of the winding number map on the set C \ supp(τ), which contains the connected set K, we get that the winding number I(τ, c) is defined for all c ∈ K and is equal to 1. Therefore we have that K ⊂ I(τ) and since the interior of τ is contained in the union of all squares which intersect K we have that K ⊂ I(τ) ⊂ U.

Lemmas 2.1.8 and 2.1.9 give us what R.Remmert calls the Circuit theorem

13 2. Preliminaries

Theorem 2.1.10 (Circuit theorem). Let K be a connected compact set in an open domain U. Then there exists a step polygon γ in U \ K such that I(γ, p) = 1 for all p ∈ K.

Proof. Let τ be like in Lemma 2.1.9. Fix a c ∈ K, we have that

N X 1 = I(τ, c) = I(τj, c) j=1 then by Lemma 2.1.8 therefore there exists an 1 ≤ m ≤ N, such that I(τm, c) = 1. Set γ = τm, and γ is precisely a step polygon in U ⊂ K such that K ⊂ I(γ).

By Lemma 2.1.8 every step polygon divides C into two domains, the interior and the exterior. Let V = E(γ) ∪ {∞}, and let us regard it as a subset of the Riemann sphere C = C∪{∞}. Since the complement of V , C\V = E(γ)∪supp γ is connected then V is simply connected. Furthermore if the interior is non-empty, then by the Riemann mapping theorem the set V can be mapped biholomorphically onto the open unit disk E. Proof of the preceding statements can be found in [1, Thm. VI.5.2-4], as that book defines simply connected subsets of the Riemann sphere as subsets such that their complement is connected, and then proves that is equivalent to the topological definition. A natural question is, can we extend the map V → E to the boundary?

The following definition introduces the notion of a simple boundary point, which by Theorem 2.1.12 is precisely the condition for the map in the previous paragraph to have an extension to the boundary.

Definition 2.1.11. [1, Def. VI.5.7] Let U be a domain in C. We say that z ∈ ∂U is a simple boundary point of U if, for every sequence (zn) ∈ U such that lim zn = z, n→∞ there exists a path γ : [0, 1] → C and a strictly increasing sequence (tn) ∈ [0, 1[, such that:

(i) lim tn = 1. n→∞

(ii) γ(tn) = zn for all n.

(iii) γ(t) ∈ U for all t ∈ [0, 1[.

Theorem 2.1.12. Let U be a bounded simply connected domain with every boundary point simple and f : U → E be a biholomorphic mapping. Then we can extend f to a homeomorphism f : U → E.

14 2.1. Topological Preliminaries

A proof of 2.1.12 can be found in [1, Thm. VI.5.10]. Let γ be a step polygon, and consider the exterior E(γ) of γ, we want to show that every point on the boundary of E(γ) is a simple boundary point. We start with a lemma:

Lemma 2.1.13. Let P be the convex set

P = {z ∈ C | Im(z) > 0, |z| < 1}. Then 0 ∈ ∂P is a simple boundary point of P .

Proof. Let (zn) be a sequence in P such that lim zn = 0 as n → ∞. We define a continuous function α: [0, ∞) → P by mapping every interval [n − 1, n] to the straight line segment [zn, zn+1], that is for every t ∈ [n − 1, n], t 7→ (n − t)zn + (t − n + 1)zn+1. It is clear that α(t) ∈ P for all t ∈ [0, ∞[ since P is convex, and limt→∞ α(t) = 0. Then the function ( α t  t ∈ [0, 1), τ : [0, 1] → C τ(t) = 1−t 0 t = 1, satisfies the conditions in Definition 2.1.11.

Lemma 2.1.14. Let γ be a step polygon with I(γ) not empty. Then both E(γ) and I(γ) are domains with every boundary point simple.

Proof. Let x be a point on ∂E(γ) = supp(γ) then x lies on some segment in γ. Assume that x is an endpoint of a segment, then there are segments [p, x] and [x, q] in γ for some points p, q ∈ C, such that p, q and x are all distinct. Then there exists a δ > 0 small enough that the open disk B(x, δ) around x intersects supp(γ) only on the segments [p, x] and [x, q], and p, q 6∈ B(x, δ). The segments [p, x] and [x, q] are both straight and meet in the centre x of B(x, δ) and both endpoints p and q are outside the open disk B(x, δ), the open set B(x, δ) \ supp(γ) divides into exactly two components U and V . Since

U ∪ V = (C \ supp(γ)) ∩ B(x, δ) = (E(γ) ∪ I(γ)) ∩ B(x, δ), then one of the components U and V is contained in the interior of γ and the other in exterior. Assume that U is a subset of the exterior. By rotating and translating U, such that x goes to 0, we can see that U is one of the three following sets  kπ  P = reiθ | 0 < r < δ, 0 < θ < , k = 1, 2, 3. k 2

3π There exists a well defined angle on P3, i.e. a continuous map φ: P3 →]0, 2 [, so the |z| 2iφ(z)/k maps ψk(z): Pk → P , ψk(z) = δ e , where P is the set in Lemma 2.1.13, are

15 2. Preliminaries

homeomorphisms. Furthermore for any sequence (zn) in Pk such that lim zn = 0, then limn→∞ ψk(zn) = 0.

Let (zn) be any sequence in E(γ) such that zn → x, then the sequence is eventually in U. We therefore get a corresponding sequence (ψk(zn)) in P which tends to 0, but by Lemma 2.1.13, the point 0 is a simple boundary point of P . Then there exists a path τ in P , and a strictly increasing sequence (tn) ∈ [0, 1[ such that lim tn = 1, τ(tn) = ψk(zn) for all n large enough, and τ(t) ∈ P for all [0, 1[.

−1 −1 −1 Therefore, (ψk) ◦ τ is a path in C such that (ψk) ◦ τ([0, 1[) ⊂ E(γ), and (ψk) ◦ τ(tn) = zn for all n large enough. Since there exists a constant N > 0 such that zn ∈ B1 for all n > N, and since E(γ) is path connected, we can find a path τ1 : [0, tN+1] → E(γ) such that τ1(tn) = zn for all n ≤ N + 1. Then ( τ (t) t ∈ [0, t ] α(t) = 1 N+1 −1 (ψk) ◦ τ(t) t ∈ [tN+1, 1] is a path satisfying the conditions in Definition 2.1.11. Therefore x is a simple boundary point of E(γ).

If x is not an endpoint of a segment, then x is on some [p, q] in γ. Let δ > 0 be small enough such that the open disk B(x, δ) intersects supp(γ) only on the segment [p, q] in γ and p, q∈ / B(x, δ). In the same way as before B(x, δ) \ supp(γ) divides into two components U and V , and we can assume U = E(γ) ∩ B(x, δ). By rotating and translating U we can see that U is the set P2 defined above. Using the same arguments as when x is an endpoint of a segment we see that x is a simple boundary point of E(γ).

The argument above works in the same way if we use the component of B(x, δ) \ supp(γ) that lies on the interior of γ instead.

Now we are ready to show the following lemma.

Lemma 2.1.15. Let γ be a step polygon, then the domains I(γ) and V˜ = E(γ) ∪ {∞} are both biholomorphic to E, furthermore those biholomorphisms can be extended to homeomorphism of I(γ) ∪ supp(γ) and E(γ) ∪ supp(γ) ∪ {∞} onto the closed unit disk E.

Proof. Since I(γ) is simply connected and bounded, there exists a biholomorphism F : I(γ) → E, by the Riemann mapping theorem. By Lemma 2.1.14, every boundary point of I(γ) is simple so F extends to a homeomorphism F : I(γ) ∪ supp(γ) → E.

16 2.1. Topological Preliminaries

By translation we can assume that 0 ∈ I(γ). Then there exists a constant r > 0 1 such that the closed disk B(0, r) is contained in I(γ). The map f(z) = z maps 1 the domain E(γ) ∪ {∞} biholomorphically onto a set W contained in B(0, r ). By the remark preceding Definition 2.1.11 the set E(γ) ∪ {∞} is simply connected so W is as well. The boundary of E(γ) ∪ {∞} is supp(γ), and every boundary point is a simple boundary point of E(γ) ∪ {∞}, by Lemma 2.1.14. We therefore get a mapping of supp(γ) onto the boundary of W . Now for any point z ∈ ∂W , and any sequence (zn) ∈ W \{0} with lim zn = z, we get a corresponding sequence 0 0 0 0 zn = f(zn) in E(γ) ∪ {∞} such that lim zn = z with z = f(z) ∈ supp(γ), and since every boundary point of E(γ) ∪ {∞} is a simple boundary point we get a map β : [0, 1] → C, with β[0, 1[⊂ E(γ) ∪ {∞} that satisfies the conditions in Definition 2.1.11. But then the f ◦ β : [0, 1] → C satisfies the conditions of Definition 2.1.11 for z being a simple boundary point of W .

Now W is a simply connected bounded domain of C so there exists a biholomorphism G: W → E by the Riemann mapping theorem. Since every boundary point of W is a simple boundary point, the map G extends to a homeomorphism G: W → E. But then G ◦ f : E(γ) ∪ supp(γ) ∪ {∞} → E is a homeomorphism that extends the biholomorphic map

G ◦ f : E(γ) ∪ {∞} → E.

The next lemma gives us the main result of this section.

Lemma 2.1.16. Let U be a domain in C, W be a simply connected domain in C, a ∈ U ∩ W and γ : [0, 1] → C be a closed path in U ∩ W at a. Let h be a homotopy from γ to the constant path at a, such that h(t, 1) = γ(t), h(t, 0) = a, h(1, s) = h(0, s) = a and h(t, s) ∈ U for all t, s. Then there exists a homotopy k, from γ to the constant path at a, fulfilling the same conditions as h, such that k(t, s) ∈ U ∩ W for all t, s ∈ [0, 1]

Proof. The image of the closed loop |γ| is a compact set contained in U ∩ W . By Lemma 2.1.10 there exists a step polygon τ in (U ∩ W ) \ |γ|, such that supp(τ) ⊂ U ∩ W , and |γ| ⊂ I(τ). Since W is simply connected I(τ) ⊂ W , see [11, Thm. 8.2.6]. By Lemma 2.1.8 the set C \ supp(τ) is divided up into two components I(τ) and E(τ) neither of which is empty, and they share a common boundary, ∂I(τ) = ∂E(τ) = supp(τ). Denote by X the set E(τ)∪supp(τ), then ∂X = supp(τ), and by Lemma 2.1.15 we get a homeomorphism f : X → E \{0}.

17 2. Preliminaries

The map r : \{0} → ∂ , r(z) = z , satisfies r| = id , so it is a retraction of E E |z| ∂E ∂E E \{0} to ∂E.

Then the map f −1 ◦ r ◦ f : X → ∂X is also a retraction. Since ∂X = supp(τ) we get a well defined continuous map F : C → I(τ) ( z z ∈ I(τ) = supp(τ) ∪ I(τ), F (z) = (f −1 ◦ r ◦ f)(z) z ∈ X = supp(τ) ∪ E(τ).

Consequently, the map k = F ◦ h is a homotopy between γ and the constant path at a such that k(t, s) ∈ U ∩ W for all (t, s) ∈ I × I. Note that k(t, s) = h(t, s) for any (t, s) such that h(t, s) ∈ I(τ) ∩ W ∩ U.

As a direct consequence we get the following theorem:

Theorem 2.1.17. Let U, W and a be as in Lemma 2.1.16. Then the map

π1(U ∩ W, a) → π1(U, a) induced by the inclusion U ∩ W,→ U is injective.

Proof. Pick an element [γ] ∈ π1(U ∩ W, a) that maps to the identity element in π1(U, a), represented by some closed loop γ in U ∩ W at a. Since [γ] maps to the identity element in π1(U, a), then γ is homotopic to the constant path at a in U. But by the previous Lemma 2.1.16 then there exists a homotopy to the constant path a in the set U ∩ W , so [γ] is the identity element. Therefore the map is injective.

The intuitive idea behind the previous statement is that for U a subset of the plane, we can not introduce more ‘holes’ by intersecting with a simply connected domain W . It is worth pointing out that this is a special property of the subsets of the plane as Theorem 2.1.17 is not true for arbitrary surfaces. A simple example is the following, let U be the open subset C \{0} of the Riemann sphere, and let W be an open disk around 0 which does not include ∞. Then U is simply connected, but U ∩ W is not, so the map in Theorem 2.1.17 is not injective.

Theorem 2.1.18. Let U be a domain in C and p: Y → U be a universal cover for U. Pick a base point a ∈ U and a point y ∈ Y such that p(y) = a. Fix some r > 0 and denote by U 0 the connected component of U ∩ B(a, r) which contains a. Furthermore, let Y 0 be the connected component of p−1[U 0] which contains y. Then Y 0 is a universal covering of U 0, and p0(y) = a, where p0 is the restriction of p to the set Y 0.

18 2.2. Riemann Domains

Proof. Consider the following commutative diagram, where the horizontal arrows are inclusions.

(Y 0, y) > (Y, y)

p0 p (2.1) ∨ ∨ (U 0, a) > (U, a)

If we apply π1, we arrive at another commutative diagram.

0 π1(Y , y) > π1(Y, y)

0 π1(p ) π1(p) (2.2) ∨ ∨ 0 π1(U , a) > π1(U, a)

0 0 The spaces Y and Y are covering spaces, therefore the maps π1(p ) and π1(p) are 0 0 injective. This can be seen by considering an element [γ] ∈ π1(Y , y) that π1(p ) maps to the identity element. Then there exists a homotopy of the closed loop p0(γ) at a to the constant path at a. This homotopy can be lifted, by the lifting property of covering spaces, to a homotopy of the closed loop γ at y in Y 0 to the constant path at y.

0 Theorem 2.1.17 tells us that the map π1(U , a) → π1(U, a) is injective, so the map 0 0 π1(Y , y) → π1(U , a) → π1(U, a) is also injective. The space Y is simply connected, 0 so π1(Y, y) is trivial. Since the diagram commutes, then π1(Y , y) must also be trivial. Therefore Y 0 is simply connected so it is a universal covering space for U 0.

2.2. Riemann Domains

Here we present the definition of a Riemann domain. We are mostly interested in the case where a Riemann domain is a covering space, which have the important property of being second countable, and we will use them to capture the multivalued functions presented in Radó’s original proof [10]. We can prove that these covering Riemann domains are second countable without appealing to a famous theorem of

19 2. Preliminaries

Radó, see [4, Thm. 23.3], and, being second countable, allows us to get a special case of Montel’s theorem.

Definition 2.2.1. (i) A Riemann domain over C is a connected Hausdorff space X with a local homeomorphism p: X → C.

(ii) Let p: X → C be a Riemann domain over C. A function f : X → C is said to be holomorphic if f ◦ σ is holomorphic for every section σ of p.

If U is an open domain in C and p: U → C is the inclusion p(z) = z, then p: U → C is a Riemann domain and the holomorphic functions on the Riemann domain p: U → C are precisely the usual holomorphic functions on the open domain U in C.

Additionally, let p: X → C be a Riemann domain, and let G be a subset of C such that p(X) ⊂ G, then by shrinking the codomain of the function p we can talk about a Riemann domain p: X → G over G.

Definition 2.2.2. A morphism f : X → Y between two Riemann domains, p: X → C and q : Y → C, is a continuous map such that p = q ◦ f.

Lemma 2.2.3. Let p: X → C and q : Y → C be two Riemann domains over C. Let f : X → Y be a morphism and σ : U → X be a section of p where U is open set. Then the map σ(U) → f(σ(U)), z 7→ f(z), is a homeomorphism.

Proof. By Lemma 2.1.3 the map f ◦ σ is a section of q. But, by Lemma 2.1.2 (ii), the map U → f(σ(U)), z 7→ f(σ(z)) is a homeomorphism. Consequently the map σ(U) → f(σ(U)), z 7→ ((f ◦ σ) ◦ p)(z) = f(z) is a homeomorphism.

Definition 2.2.4. Let p: X → C be a Riemann domain. We say that a holomorphic function f : X → C is injective with respect to p (or over p) if for any x, y ∈ X such that f(x) = f(y) then p(x) = p(y).

The condition that a holomorphic function is injective with respect to p is precisely what we need to define a local homeomorphism q : f(X) → C such that f : X → f(X) is a morphism of Riemann domains.

Lemma 2.2.5. Let p: X → C be a Riemann domain and f : X → C be holomorphic. The map f is injective with respect to p if and only if there is a continuous mapping

20 2.2. Riemann Domains q : f(X) → C with q ◦ f = p. Then q is uniquely defined, f(X) is open in C, q is a holomorphic function, and q : f(X) → C is a Riemann domain.

Proof. Assume there exists a q : f(X) → C such that q◦f = p, and let x, y ∈ X such that f(x) = f(y). Then p(x) = q(f(x)) = q(f(y)) = p(y) and f is injective with respect to p. Now assume that f is injective with respect to p and let y be a point in f[X]. Then p is constant on f −1[y], therefore q : f(X) → C, y 7→ p(x) where p is any point in f −1[y], is a well defined map such that q ◦ f = p, and furthermore is unique.

To prove the rest, let y ∈ f(X). We want to show that there exists a neighbourhood −1 Y around y such that q|Y is a homeomorphism. Pick a point x ∈ f [y] and let σ : U → X be a section of p where U is a neighbourhood of p(x) and σ(p(x)) = x. By definition the map f ◦ σ : U → f(X) is holomorphic and (f ◦ σ)(p(x)) = y. Let z1, z2 ∈ U such that f ◦ σ(z1) = f ◦ σ(z2), since f is injective with respect to p, we have that z1 = p(σ(z1)) = p(σ(z2)) = z2 so the holomorphic map f ◦ σ is injective. An injective holomorphic map is a homeomorphism onto its image, so f [σ(U)] is a neighbourhood of y homeomorphic to U, and since q ◦ f = p we have that q ◦ (f ◦ σ) = idU , and the restriction q : f [σ(U)] → U is a biholomorphism. Therefore q is a locally biholomorphic map and q : f(X) → C is a Riemann domain.

Lemma 2.2.6. Let p: X → C be a Riemann domain over C and f : X → C be holomorphic and injective over p. Let a, b ∈ X be such that f(a) = f(b), U ⊂ C be a connected neighbourhood of w : = p(a) = p(b), and σa, σb : U → X be sections of p such that σa(w) = a and σb(w) = b. Then f ◦ σa = f ◦ σb.

Proof. By Lemma 2.2.5 then q : f(X) → C is a Riemann domain and f : X → f(X) is a morphism of Riemann domains. Then f ◦ σa and f ◦ σb are two sections of q which agree at a point so by Lemma 2.1.2 they agree on the whole of U.

We can use Lemma 2.2.6 to define derivatives of holomorphic functions on Riemann domains. Let p: X → G be a Riemann domain and f : X → C be a holomorphic function which is injective with respect to p. Pick points g0 ∈ G and x0 ∈ X such that p(x0) = g0. If σ1 and σ2 are two sections of p on a neighbourhood U of g0, such that σ1(g0) = x0 = σ2(g0), then by Lemma 2.2.6 the functions (f ◦ σ1): U → C 0 and (f ◦ σ2): U → C are equal. Therefore the ordinary derivatives (f ◦ σ1) (g0) and 0 (f ◦ σ2) (g0) are equal and independent of a choice of a section. From the preceding discussion we get the following definition:

21 2. Preliminaries

Definition 2.2.7. Let p: X → G be a Riemann domain, f : X → C be holomorphic and injective with respect to p, g0 ∈ G and x0 ∈ X. Then the derivative of f at x0, is defined by 0 0 f (x0) = (f ◦ σ) (g0) where σ is a section of p on some neighbourhood U of g0 such that σ(g0) = x0.

Lemma 2.2.5 tells us that for a Riemann domain p: X → G, with G open subset of C, any holomorphic map f : X → C which is injective with respect to p induces a local biholomorphism q : f(X) → C such that f(X) with this map is a Riemann domain. We will denote this map by f ∗, and this is the unique local biholomorphism induced by f such that the following diagram commutes.

f X > f(X)

p f ∗

∨ < G

Insisting that the map p: X → G is a covering map, instead of being just a local homeomorphism, we have by Lemma 2.1.4 that f and f ∗ are covering maps as well. It is well known that π1(G, a) is countable ([7, Thm. 8.11]), if G is an open domain −1 of C. Fix an x0 ∈ X such that p(x0) = a. For any x ∈ p [a], there exists a path γ : [0, 1] → X with γ(0) = x0 and γ(1) = x. Then [p ◦ γ] ∈ π1(G, a), and −1 we get a surjective map π1(G, a) → p [a], [β] 7→ β(1) where β is the lifting of β −1 with β(0) = x0. But then the fibers p [g], g ∈ G are countable, since π1(G, g) is countable, and therefore X is second countable. We then get the following theorem:

Theorem 2.2.8 (Montel’s theorem for Riemann domains). If (fn) is a sequence of holomorphic functions on a Riemann domain p: X → U, with p a covering map, and (fn) is uniformly bounded on each compact subset of X, then there exists a subse- quence of (fn) which converges uniformly on compact subsets of X to a holomorphic function on X.

This is a special case of more general theorem of Montel for arbitrary Riemann surfaces [9], but the general case relies on Radó’s theorem to show that arbitrary Riemann surfaces are second countable.

For the sake of completeness we present here two results from classical that will be used in later sections.

22 2.2. Riemann Domains

Lemma 2.2.9 (Schwarz’s lemma). Let f : E → C be a holomorphic map such that f(0) = 0 and |f(z)| ≤ 1. Then |f(z)| ≤ |z| for all z ∈ E and |f 0(0)| ≤ 1. Moreover if |f 0(0)| = 1 or if |f(z)| = |z| for some z 6= 0, then there exists a constant a ∈ C with |a| = 1 such that f(z) = az.

Theorem 2.2.10 (Hurwitz’s theorem). Let (fk) be a sequence of holomorphic func- tions on a domain U ⊂ C that converges uniformly on compact subsets of U to a holomorphic function f on U that is not identically zero. If f has a zero of order m at z0, then for every small enough ρ > 0 there exists a natural number N ∈ N such that fn has exactly m zeroes, counted with multiplicity, in the disk defined by |z − z0| < ρ, for all n > N. Furthermore these zeroes converge to z0 as n → ∞.

For proofs of Lemma 2.2.9 and Theorem 2.2.10 see [2, Thm. VI.2.1] and [2, Thm. VII.2.5].

23

3. Uniformization Theorem for Domains in C

In this chapter we present the main results of this thesis, proving the following theorem:

Theorem 3.0.1 (Uniformization theorem for domains in C). Let G be a domain in C, then there exists a holomorphic covering map p : X → G, where X is either C or the unit disk E.

Note that if G = C or G = C∗ := C\{0}, then G can be covered by id : C → C or by exp : C → C∗, respectively. We only have to consider the case when the complement of G contains at least two points. That is to say, we prove the following:

Theorem 3.0.2. Let G be any domain in C \{0, 1}, and p: Y → G be a universal covering space for G. Then Y is isomorphic to the open unit disk E.

The preceding theorems give us a classification of domains, the plane C is a universal covering space for itself or the punctured plane C \{z}, for any z ∈ C. All other domains in C have the unit disk E as a universal cover.

The strategy used to prove 3.0.2 is similar to the method presented by Radó in [10], shown in section 1.2, to prove the Riemann mapping theorem. The natural way to extend this method for a general domain G is to consider a family of functions fn : G → E and construct a surjective function G → E by solving an extremal problem as in the proof for the Riemann mapping theorem. Since we are trying to construct a covering map p : E → G, we will have to deal with some multivalued functions. This is precisely what Radó does in his paper [10].

Here we look at a universal covering space p : Y → G, and work with maps fn : Y → E, injective with respect to p, and use the same kind of extremal ar- gument as Radó does in his paper, to construct a surjective map f : Y → E, which is injective with respect to p. In light of Lemma 2.2.5 and Lemma 2.1.4, the map

25 3. Uniformization Theorem for Domains in C f : Y → E is a covering map, but Y and E are both simply connected so they are biholomorphic, that is to say, G can be holomorphically covered by E.

3.1. The Set Σ(G, y0)

As we said above, we are only considering the case when the complement of G con- tains at least two points. Without loss of generality we can assume that G ⊂ C \{0, 1}.

Let p: Y → G be a universal covering space of G, g0 ∈ G and y0 ∈ Y such that p(y0) = g0. The space Y is then a Riemann domain. We are interested in the space of holomorphic functions on the Riemann domain Y , which are injective with respect to p.

Definition 3.1.1. Let G be a domain of C\{0, 1}, p: Y → G be a universal covering space of G, and pick points g0 ∈ G and y0 in Y such that p(y0) = g0. We denote by Σ(G, y0) the set of holomorphic functions f : Y → E injective with respect to p such 0 0 that f(y0) = 0, and such that the derivative f (y0) = (f ◦σ) (g0), see Definition 2.2.7, is a real number in the interval ]0, ∞[, where σ is some section p with σ(g0) = y0. Furthermore we define λ(G, y0) as

0 λ(G, y0) = sup {f (y0) | f ∈ Σ(G, y0)} .

We now have to show that the set Σ(G, y0) is not empty. Fortunately, when G is bounded it is easy to find a function f ∈ Σ(G, y0).

Lemma 3.1.2. If G is bounded then Σ(G, y0) is non-empty.

Proof. Let C be a large enough positive real number such that f(G) ⊂ E, where z−g0 0 f(z) = C . Then f (g0) > 0, and if p: Y → G is a universal cover such that p(y0) = g0, we see that the function f ◦ p: Y → E is in Σ(G, y0).

Lemma 3.1.3. If Σ(G, y0) is not empty then λ(G, y0) is finite.

Proof. Since p is a local homeomorphism there exists a section σ : U → Y , with U open disk of radius ρ centered at g0, such that σ(g0) = y0. Now for any function f ∈ Σ(G, y0), we have that h = f ◦ σ is a holomorphic function from U to E, and 0 0 1 h (g0) > 0. Schwarz’s lemma gives us that h (g0) ≤ ρ , and since f was arbitrary, it 1 follows that λ(G, y0) ≤ ρ .

26 3.1. The Set Σ(G, y0)

Definition 3.1.4. For any point w ∈ E, w 6= 0, we define the map hw : E → E by z − w h (z) = . w zw − 1

The map hw is the unique automorphism of E such that 0 7→ w and w 7→ 0. This can be seen by letting g be any automorphism of E such that g(0) = w and g(w) = 0. −1 Then the map ψ = g ◦ hw is an automorphism of E and ψ(0) = 0 and ψ(w) = w. By the Schwarz lemma we have that ψ : E → E is the identity map, and therefore g = hw.

Taking the derivative of the map hw we get (zw − 1) − w(z − w) |w|2 − 1 h0 (z) = = , w (zw − 1)2 (zw − 1)2

0 2 0 1 so hw(0) = |w| − 1 and hw(w) = |w|2−1 .

0 The next lemma shows that if we can find a map f ∈ Σ(G, y0) with f (y0) = λ(G, y0), then it must be surjective onto E. In light of Lemma 2.1.4 then a surjective morphism f : Y → E is a covering map. The method used in the proof is what Radó refers to as the Carathéodory-Koebe square root transformation.

Lemma 3.1.5. If f ∈ Σ(G, y0) is not surjective onto E, then there exists a mapping 0 0 F ∈ Σ(G, y0) such that F (g0) > f (g0).

Proof. Let w ∈ E be a value that f does not take, and write w = βeiφ, 0 < β < 1.

Let hw : E → E be the automorphism in Definition 3.1.4, by composing hw and f we get a map hw ◦ f : Y → E \{0}. Since Y is simply connected, there exists a function 2 √ i 1 φ r : Y → E \{0} such that r = hw ◦ f and r(y0) = w1, where w1 = βe 2 .

Define a function F : Y → E by

1 i 2 φ F (y) = e [hw1 (r(y))] .

To show that F is injective with respect to p assume that there exist c1, c2 ∈ Y such that F (c1) = F (c2). Then hw1 (r(c1)) = hw1 (r(c2)) and since hw1 is bijective, r(c1) = r(c2). By definition of r,

2 2 hw(f(c1)) = (r(c1)) = (r(c2)) = hw(f(c2))

27 3. Uniformization Theorem for Domains in C

and since hw is bijective, f(c1) = f(c2). But f is injective with respect to p so p(c1) = p(c2), and therefore F is injective with respect to p.

We see that F (y0) = 0, and by the chain rule 0 0 i 1 φ 0 0 i 1 φ 0 0 i 1 φ r (y0) F (y ) = e 2 h (r(y ))r (y ) = e 2 h (w )r (y ) = e 2 . 0 w1 0 0 w1 1 0 2 |w1| − 1 0 To evaluate r (y0) we note that 0 0 0 0 0 2r(y0)r (y0) = hw(f(y0))f (y0) = hw(0)f (y0) so 2 0 0 (|w| − 1)f (y0) r (y0) = . 2w1 We gather all the terms and we have 2 1 1 |w| − 1 1 0 i 2 φ 0 F (y0) = e 2 f (y0) 2 |w1| − 1 w1 2 i 1 φ 1 β − 1 1 0 = e 2 f (y0) √ i 1 φ 2 β − 1 βe 2 β + 1 = √ f 0(y ) 2 β 0

0 Then F (y0) is a real number and greater then 0, so F ∈ Σ(G, y0). Since 0 < β < 1, β√+1 0 0 it follows that 2 β > 1, which implies F (y0) > f (y0).

0 As said before the previous Lemma 3.1.5 shows that if f ∈ Σ(G, y0) and f (y0) = λ(G, y0), then f : Y → E is a surjective function. Since f induces a locally biholo- morphic map f ∗ : E → G, such that p = f ∗ ◦ f, Lemma 2.1.4 tells us that both f and f ∗ are covering maps.

0 4 1 Lemma 3.1.6. If f ∈ Σ(G, y0) and f (y0) > 5 λ(G, y0), then 4 E ⊂ f(Y ).

Proof. If f is a surjective function onto E there is nothing to prove. Assume f is not surjective and let w = βeiφ be a value which f does not take. Define F as in the proof for Lemma 3.1.5. We get the following inequalities β + 1 β + 1 4 λ(G, y ) ≥ F 0(y ) = √ f 0(y ) > √ λ(G, y ) 0 0 2 β 0 2 β 5 0 so that β + 1 5 √ < . 2 β 4 1 1 Then β > 4 and w 6∈ 4 E.

28 3.2. Constructing a Surjective Function

Let f be as in Lemma 3.1.6 and fix ρ > 0 so U = B(g0, ρ) is evenly covered by p. Then f ∗ : f(Y ) → G is a function such that for any section σ : U → Y of p, we have ∗ ∗ 0 ∗ 0 the equality f ◦ (f ◦ σ) = idU . Therefore (f ) (0) is a real number and (f ) (0) > 0.

3.2. Constructing a Surjective Function

In this section we are going to show that there exists a function f ∈ Σ(G, y0) with 0 f (y0) = λ(G, y0) for any domain G in C\{0, 1}, and the condition on the derivative shows that it is a surjective function.

Let Gn be the connected component of G ∩ B(g0, n) which contains g0, for n = 1, 2,.... Let g be any point in G, since G is a domain there exists a path γ : [0, 1] → G such that γ(0) = g0 and γ(1) = g. The image of γ is a compact set |γ| in C so there exists an N > 0 such that |γ| ⊂ B(g0,N). Then g is in GN , and since g was arbitrary we see that ∞ [ G = Gn. 1

Let p: Y → G be a universal cover of G and let y0 ∈ Y be such that p(y0) = g0. −1 Denote by Yn the connected component of p [Gn] that contains y0. We have by Lemma 2.1.18 that Yn is a universal cover of Gn and

Y1 ⊂ Y2 ⊂ ... ⊂ Y.

Using a similar argument as before we can show that the union of all the Yn is Y . Let y ∈ Y be an arbitrary point. Then there exists a path γ : [0, 1] → Y such that γ(0) = y0 and γ(1) = y. Then p ◦ γ is a path in G from g0 to p(y), and the image of p ◦ γ is compact so there exists an N > 0 such that |p ◦ γ| ⊂ GN . But since YN is a universal covering space of GN , there exists a lifting γ˜ of p ◦ γ to YN such that

γ˜(0) = y0 and p|YN ◦ γ˜ = p ◦ γ. Since YN ⊂ Y then γ˜ is also lifting in Y , but by the ˜ lifting property γ˜ and γ are equal and therefore y = γ(1) = γ(1) ∈ YN and we get:

∞ [ Yn = Y. 1

Since every Gn is bounded, Lemma 3.1.2 tells us that Σ(Gn, y0) is not empty, for any n ≥ 1. Let n, m ∈ N such that n ≥ m. Since Ym ⊂ Yn we can restrict any function f ∈ Σ(Gn, y0) to the subspace Ym, and we get a function f|Ym ∈ Σ(Gm, y0). Therefore we have that

29 3. Uniformization Theorem for Domains in C

λ(Gm, y0) ≥ λ(Gn, y0) ≥ λ(G, y0) n ≥ m.

For the final piece of the puzzle we need to obtain a lower bound on λ(G, y0), and for that purpose we use Schottky’s theorem.

Theorem 3.2.1 (Schottky’s theorem). Let a ∈ C \{0, 1}, then there exists a con- stant M > 0 such that for any holomorphic function f : E → C\{0, 1} with f(0) = a, 1 |f(z)| < M for all z ∈ 2 E.

Schottky’s theorem can be proved using the covering space of the set C \{0, 1} which is precisely what we are trying to establish. There exist however proofs, using elementary methods, so we avoid circular reasoning. See for example [11, Ch. 10.3] for a proof using Bloch’s theorem.

Using this theorem, and the same notation we have used through out this section, we get the following lemma:

Lemma 3.2.2. There exists a constant C > 0 such that λ(Gn, y0) ≥ C for all n ≥ 1.

Proof. Let 0 < ρ < 1 such that the open disk U = B(g0, ρ) is evenly covered by p: Y → G, and U ⊂ G1. Then there exists a section σ : U → Y of p, such that σ(g0) = y0 and the image of σ is contained in Y1, and therefore in Yn for all n ≥ 1.

Fix n0 ≥ 1, since Gn0 is bounded the set Σ(Gn0 , y0) is not empty and therefore 1 λ(Gn0 , y0) > 0. Let f ∈ Σ(Gn0 , y0) be as in Lemma 3.1.6, then 4 E ⊂ f(Yn0 ). ∗ According to Lemma 2.2.5, f induces a local biholomorphism f : f(Yn0 ) → G. By ∗ 1 ∗ restricting f we get a map h: 4 E → G, z 7→ f (z), and since G ⊂ C \{0, 1}, Schottky’s theorem tells us there exists a constant M > 0 such that |h(z)| < M for 1 all z ∈ 8 E, and this constant M is independent of n0. Schwarz’s lemma 2.2.9, tells us that |h0(0)| ≤ 8M, but h is just a restriction of f ∗ so |(f ∗)0(0)| ≤ 8M. By the remarks after Lemma 3.1.6, (f ∗)0(0) is a positive real number so (f ∗)0(0) < 8M.

∗ Since σ(U) ⊂ Yn, for all n we get that f ◦ (f ◦ σ) = idU . Therefore, finally, by the chain rule: 1 1 f 0(y ) = > . 0 (f ∗)0(0) 8M

It is important to note that the lower bound obtained in Lemma 3.2.2 depends only on the map p: Y → G, the constant g0, and the open disk U used in the proof. This

30 3.2. Constructing a Surjective Function constant is independent of which n ≥ 1 is used. Next we will show simultaneously 0 that there exists a function F ∈ Σ(G, y0), F : Y → E, and that F (y0) = λ(G, y0), which implies that F : Y → E is a covering map.

0 Theorem 3.2.3. There exists a function F ∈ Σ(G, y0) such that F (y0) = λ(G, y0).

Proof. Since Σ(Gn, y0) 6= ∅ by 3.1.2, we can choose a sequence (Fn) such that Fn ∈ Σ(Gn, y0), and 1 λ(G , y ) − F 0 (y ) ≤ for all n = 1, 2,... n 0 n 0 n

The sequence of functions Fn : Yn → E is uniformly bounded, but have different domains, so we can not apply Montel’s theorem directly. By restricting all of the functions (Fn) to Y1 we get a sequence of functions Fn|Y1 : Y1 → E, which is uniformly bounded. By Montel’s theorem we get a subsequence (F1,n) of (Fn) such that the restrictions to Y1 converge uniformly on compact subsets of Y1 to a holomorphic function F : Y1 → E. In the same way as before we can restrict the functions in the sequence (F1,n) to the set Y2, and by Montel’s theorem we get a subsequence (F2,n) such that the restrictions to Y2 converge uniformly on compact subsets of Y2 to a holomorphic function Y2 → E. By continuing this process we obtain for any i > 1 a subsequence (Fi,n) of (Fi−1,n), which converges uniformly on compact subsets of Yi to a holomorphic function Yi → E.

Consider the ‘diagonal’ sequence (fn) defined by fn = Fn,n for all n ≥ 1. We get a corresponding increasing sequence m1, m2,... of natural numbers such that fn : Ymn → E. The sequence (fn) converges uniformly on compact subsets of Y to a holomorphic function f : Y → E. Indeed, let K be a compact subset of Y , by definition of (Yn) there exists a constant N such that K ⊂ YN , and therefore the domains of (fn) eventually contain YN . By restricting the functions to the set YN , we get that these restrictions converge uniformly on compact subsets of YN , and therefore on K. Since K was arbitrary we have that (fn) converges uniformly on compact subsets of Y to a holomorphic function f : Y → E.

By Lemma 3.2.2 there exists a constant C > 0 such that λ(Gn, y0) ≥ C for all n ≥ 1, and it follows that

0 0 f (y0) = lim fn(y0) = lim λ(Gn, y0) ≥ C. n→∞ n→∞ Therefore the holomorphic function f is not constant on Y . To conclude that f ∈ Σ(G, y0) we need to show that f is injective with respect to p. Assume there are two points z1, z2 ∈ Y with p(z1) 6= p(z2) such that f(z1) = f(z2). Let N > 0 be such that p(z1), p(z2) ∈ Gn ⊂ Gn+1 for all n ≥ N. Let V,W ⊂ YN

31 3. Uniformization Theorem for Domains in C

be neighbourhoods of z1 and z2, respectively, such that p induces homeomorphisms V → p(V ) , W → p(W ), with p(V ) ∩ p(W ) = ∅, and denote by σV and σW the inverses of these homeomorphisms, respectively. The functions fn ◦ σV and fn ◦ σW converge to the functions f ◦ σV and f ◦ σW neither of which is identically zero. By Hurwitz’s theorem 2.2.10 there are points αk ∈ p(V ) and βk ∈ p(W ) with fk(σV (αk)) = fk(σW (βk)) for all k sufficiently large, but p(V ) ∩ p(W ) = ∅ which is a contradiction since the functions fn are injective with respect to p|Yn .

0 The only thing left is to show that f (y0) = λ(G, y0). Since f is in the set Σ(G, y0) we have the following inequality

0 0 λ(G, y0) ≥ f (y0) = lim fn(y0) n→∞

= lim λ(Gn, yn) ≥ λ(G, y0). n→∞

0 So λ(G, y0) = f (y0) > C.

In conclusion, by the last Theorem 3.2.3 there exists a surjective holomorphic func- tion f : Y → E which is injective with respect to p: Y → G. So f(Y ) = E is a Riemann domain, f : Y → E a morphism of Riemann domains and we have the following commutative diagram.

f Y > E

p f ∗

∨ < G

Since f is surjective, we have by Lemma 2.1.4 that f ∗ : E → G is a covering map. And since E is simply connected, f ∗ : E → G is a universal covering space of G.

We have now proved Theorem 3.0.1 and have obtained the main results of our thesis, a Uniformization theorem for domains in C.

32 A. Appendix

In this chapter we are presenting theorems and definition concerning the fundamen- tal group and covering spaces. These are all well known results but stated here for the sake of completeness. The definitions are taken from [8] and the proofs of the theorems can be found in [8] or [4].

A.1. Path Homotopy and the Fundamental Group

Definition A.1.1. Let f0, f1 : X → Y be continuous maps and H ⊂ X.

(i) We say that f0 is homotopic to f1, denoted f0 ≈ f1, if there is a continuous map F : X × I → Y such that

F (x, 0) = f0(x), and F (x, 1) = f1(x),

We call F a homotopy between f0 and f1.

(ii) We say that f0 is homotopic to f1 relative to H, f0 ≈ f1(rel H), if f0 ≈ f1 and there exists a homotopy between them with the further condition that:

F (x, s) = f0(x) for all x ∈ H, s ∈ I.

Both homotopy and relative homotopy are equivalence relations on the set of func- tions continuous functions X → Y , and we denote by [f] the equivalence class of a continuous function f : X → Y .

We have a special case for paths.

Definition A.1.2. Two paths f0, f1 : I → X, where I is the interval [0, 1], are said to be path homotopic if they are homotopic rel{0, 1}.

33 A. Appendix

Definition A.1.3. If f, g : I → X are paths such that f(1) = g(0) then we can define the product h = f ∗ g by

( 1 f(2t) for t ∈ [0, 2 ], h(t) = 1 g(2t − 1) for t ∈ [ 2 , 1].

The function h is well defined and a continuous path in X. The product also induces a well defined operation on the path homotopy equivalence classes which are also denoted by [·]: [f] ∗ [g] = [f ∗ g].

The operation is associative on path homotopy classes. Let cx : I → X denote the constant path at x, that is cx(s) = x for all s ∈ I. This acts as an identity, that is if f is a path with x0 = f(0) and x1 = f(1), then [cx0 ] ∗ [f] = [f] and −1 [f] ∗ [cx1 ] = [f]. For any path f, we can define a path f (s) = f(1 − s) for s ∈ [0, 1] −1 −1 −1 and [f] ∗ [f ] = [cf(0)], that is [f ] = [f] . A closed loop at x ∈ X is a path f : I → X such that f(0) = f(1) = x. With the operation ∗ the classes of closed loops at x will form a group.

Definition A.1.4. Let x ∈ X, the fundamental group of X at x (or relative to the base point x) is the group of path homotopy classes of closed loops at x with ∗ acting as the group operation. The fundamental group of X at x is denoted by π1(X, x).

Here we are mostly interested in the fundamental group of domains in C (or R2), and it is well known that those fundamental groups are countable. It is simple to get a more general theorem which we state here. For a proof see [7, Thm. 8.11].

Theorem A.1.5. The fundamental group of a second countable manifold is count- able.

Definition A.1.6. Let X be a topological space and x ∈ X. We say X is simply connected if it is path connected and π1(X, x) is the trivial group.

Definition A.1.7. Let h:(X, x0) → (Y, y0) be a continuous map. Define

h∗ : π1(X, x0) → π1(Y, y0), by h∗([f]) = [h ◦ f].

The map h∗ is called the homomorphism induced by h relative to the base point x0. It is a well defined group homomorphism.

34 A.2. Covering Spaces

A.2. Covering Spaces

Definition A.2.1. Let p: X → Y be a continuous surjective map. The open set U of Y is said to be evenly covered by p if the inverse image can be written as the union of pairwise disjoint open sets Vx in X such that p induces a homeomorphism of Vx onto U, for each x.

Definition A.2.2. Let p: X → Y be a continuous and surjective map, X and Y be Hausdorff spaces. If every point y of Y has a neighbourhood U that is evenly covered by p, then p is called a covering map, and X is said to be a covering space, or a cover, of Y .

Theorem A.2.3. Let p: X → Y be a covering map and Y0 be a subspace of Y . −1 Then the induced map p0 : X0 → Y0, where X0 = p [Y0], is a covering map.

Definition A.2.4. Let p: X → Y and f : H → Y be continuous maps. We say that f˜ is a lifting of f, if f˜: H → X is a continuous map such that p ◦ f˜ = f.

Definition A.2.5. A continuous map p: X → Y is said to have the curve lifting property if the following condition holds. For every curve u: [0, 1] → Y and for every point x0 ∈ X with p(x0) = u(0), there exists a lifting u˜: [0, 1] → X of u such that u˜(0) = x0.

Lemma A.2.6. Let p: X → Y be a covering map. Then it has the curve lifting property.

Lemma A.2.7. (Homotopy lifting lemma) Let p: X → Y be a covering map and let p(x0) = y0. Let the map F : I × I → Y be continuous, with F (0, 0) = y0. There ˜ ˜ is a unique lifting of F to a continuous map F : I × I → Y such that F (0, 0) = x0. If F is a path homotopy, then F˜ is a path homotopy.

Let p: X → Y be a covering map, and let α, β : [0, 1] → Y be closed curves at y0 ∈ Y . ˜ ˜ Let x0 ∈ X such that p(x0) = y0, and α,˜ β be liftings to X with α˜(0) = β(0) = x0. We see from Lemma A.2.7 that if α, β are path homotopic then α˜(1) = β˜(1). If X is simply connected we have that α,β are path homotopic if and only if α˜(1) = β˜(1).

The Homotopy lifting lemma has the following stronger form:

35 A. Appendix

Lemma A.2.8. (General lifting lemma) Let X,Y,H be connected and locally path connected spaces. Let p: X → Y be a covering map with p(x0) = y0 and f : H → Y ˜ be a continuous map with f(h0) = y0. The map f can be lifted to a map f : H → X ˜ such that f(h0) = x0 if and only if

f∗(π1(H, h0)) ⊂ p∗(π1(X, x0)).

Furthermore, if such a lifting exits, it is unique.

We see that Lemma A.2.8 implies Lemma A.2.7 since I is simply connected.

Definition A.2.9. (Universal covering) Suppose X and Y are connected topological spaces and p: X → Y is a covering map, p is called the universal covering of Y if it satisfies the following property: For every covering map q : Z → Y with Z connected, and every choice of points x0 ∈ X, z0 ∈ Z with p(x0) = q(z0), there exists exactly one continuous fiber-preserving mapping f : X → Z such that f(x0) = z0.

Let Y be connected and locally simply connected. If p: X → Y is a covering map and X is simply connected then p is a universal covering of Y .

Theorem A.2.10. (Existence of the Universal covering) Suppose Y is a connected manifold. Then there exists a connected, simply connected manifold X and a cov- ering map p: X → Y . (This implies that X is a universal covering, by the remark above)

Theorem A.2.11. Suppose Y is a manifold, X is a Hausdorff space and p: X → Y is a local homeomorphism with the curve lifting property. Then p is a covering map.

36 Bibliography

[1] Reynir Axelsson. Tvinnfallagreining II. Háskólaprent, 2005.

[2] John B Conway. Functions of one complex Variable I, volume 159. Springer, 1978.

[3] Henri Paul de Saint-Gervais. Uniformization of Riemann surfaces. European Mathematical Society, 2016.

[4] Otto Forster. Lectures on Riemann surfaces, volume 81. Springer, 2012.

[5] Claude Godbillon. Éléments de topologie algébrique. Hermann Paris, 1971.

[6] Robert Everist Greene and Steven George Krantz. Function theory of one complex variable, volume 40. American Mathematical Soc., 2006.

[7] John Lee. Introduction to topological manifolds, volume 202. Springer, 2010.

[8] James R Munkres. Topology. Prentice Hall, 2000.

[9] Terrence Napier and Mohan Ramachandran. An introduction to Riemann sur- faces. Springer, 2011.

[10] T Radó. Über die Fundamentalabbildungen schlichter Gebiete. Acta Litt. Sci. Univ. Hung, 1:240–251, 1922.

[11] Reinhold Remmert. Classical topics in complex function theory, volume 172. Springer, 2013.

37