Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

The Theory and Practice of Conformal Geometry
The Theory and Practice of Conformal Geometry
The Theory and Practice of Conformal Geometry
Ebook441 pages4 hours

The Theory and Practice of Conformal Geometry

Rating: 0 out of 5 stars

()

Read preview

About this ebook

In this original text, prolific mathematics author Steven G. Krantz addresses conformal geometry, a subject that has occupied him for four decades and for which he helped to develop some of the modern theory. This book takes readers with a basic grounding in complex variable theory to the forefront of some of the current approaches to the topic. "Along the way," the author notes in his Preface, "the reader will be exposed to some beautiful function theory and also some of the rudiments of geometry and analysis that make this subject so vibrant and lively."
More up-to-date and accessible to advanced undergraduates than most of the other books available in this specific field, the treatment discusses the history of this active and popular branch of mathematics as well as recent developments. Topics include the Riemann mapping theorem, invariant metrics, normal families, automorphism groups, the Schwarz lemma, harmonic measure, extremal length, analytic capacity, and invariant geometry. A helpful Bibliography and Index complete the text.
LanguageEnglish
Release dateMar 17, 2016
ISBN9780486810324
The Theory and Practice of Conformal Geometry

Read more from Steven G. Krantz

Related to The Theory and Practice of Conformal Geometry

Related ebooks

Mathematics For You

View More

Related articles

Reviews for The Theory and Practice of Conformal Geometry

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    The Theory and Practice of Conformal Geometry - Steven G. Krantz

    Geometry

    Chapter 1

    The Riemann Mapping Theorem

    Prologue: There is hardly a more profound theorem from nineteenth century complex analysis than the Riemann Mapping theorem. Even to conceive of such a theorem is virtually miraculous. Although Riemann’s original proof was flawed, it pointed in the right direction. Certainly a great deal of modern complex function theory has been inspired by the Riemann Mapping theorem (RMT).

    Throughout this book, we shall use the term domain to mean a connected, open set. While the Riemann Mapping theorem gives us a complex-analytic classification of simply connected planar domains, a theory (in fact several theories) has developed for multiply connected domains. This includes the Ahlfors map, the canonical representation, and the uniformization theorem. We treat all of these in the present chapter. Although we do not treat the topic here, Riemann surface theory is an outgrowth of the study of conformal mappings.

    Perhaps the most important modern concept in this circle of ideas is Teichmüller theory, which creates a moduli space for Riemann surfaces. It is beyond the scope of the present book, but it provides a pointer for further reading.

    1.0Introduction

    Capsule: It is natural to think of the Riemann Mapping theorem in the context of simply connected domains. However, from the point of view of analysis, it is more convenient to have a different formulation of the topological condition. In this section we introduce the notion of holomorphic simple connectivity: A domain U is holomorphically simply connected if any holomorphic function on U has a holomorphic antiderivative.

    It is easy to verify that any topologically simply connected domain is holomorphically simply connected. So we certainly suffer no loss of generality to use this substitute idea. It also streamlines our treatment.

    In thinking about the topology of the plane, it is natural to ask which planar open sets are homeomorphic to the open unit disc. The startling answer is that the Riemann Mapping theorem tells us that any connected, simply connected open set (except the plane) is not only homeomorphic to the disc but conformally equivalent to the disc. One can verify separately, by hand, that the entire plane is also homeomorphic to the disc (but certainly not conformally equivalent).

    Riemann’s astonishing theorem has many different proofs, and we shall consider some of them here. Some of the proofs are existence proofs, and some constructive. Some are geometric and some are analytic. The book [BIS] covers ideas connected to the Riemann Mapping theorem comprehensively.

    We end this section with a formal enunciation of the Riemann mapping theorem:

    Theorem (RMT): Let U be any simply connected domain that is not conformal to the entire plane. Then U is conformally equivalent to the unit disc.

    1.1The Proof of the Analytic Form of the Riemann Mapping Theorem

    Capsule: We actually prove the Riemann Mapping theorem in an appendix. In this section, we set up the proof. We develop the idea of holomorphic simple connectivity, and establish some of its properties. We discuss some of the significance of this important result.

    Classical arguments, which may be found in any complex analysis text (see for example [GRK1]), show that topological simple connectivity implies an analytic form of simple connectivity that we now define.

    Definition 1.1 A connected open set U is holomorphically simply connected if, for each holomorphic function f : U , there is a holomorphic antiderivative F—that is, a function satisfying F′(z) = f(z) on U.

    Example 1.2 Certainly open discs and open rectangles are holomorphically simply connected. One constructs F from f with a simple line integral.

    Notice that, on a holomorphically simply connected domain, a nonvanishing, holomorphic function f will have a logarithm—for we can just take an antiderivative of f′/f. Then it follows that such an f will have a square root.

    Let D be the unit disc. Let U . Fix a point P U and set

    We shall prove the following three assertions:

    (1) is nonempty.

    (2)such that

    (3)If g , then g maps U onto the unit disc D.

    These three assertions taken together imply the Riemann Mapping theorem (see the discussion below). The proof of assertion (1) is by direct construction. Statement (2) is proved with a normal families argument. Statement (3) is the least obvious and will require some work: If the conclusion of (3) .

    The proof of the Riemann Mapping theorem (as we know it today, not developed by Riemann but rather by Carathéodory and Koebe and others) is quite standard and can be found in most any text. For completeness we include it in an appendix to this section.

    Riemann’s original proof of his theorem solved a somewhat different extremal problem related to the Dirichlet problem. His argument was flawed because he, in fact, assumed that the extremal problem had a solution. He did not prove it. The proof was rescued some years later by refining the Dirichlet principle used by Riemann. Our modern proof sidesteps those difficulties, and uses Montel’s theorem to show that the extremal problem has a solution.

    There are a number of other approaches to proving the Riemann Mapping theorem. We shall begin the considerations of this chapter by presenting a quite modern proof based on ideas of Thurston. In fact, Thurston introduced the profound new idea of circle packing, and Rodin and Sullivan found a way to prove the Riemann Mapping theorem using circle packing. We present their proof here.

    A proof that relies on the solution of the Dirichlet problem follows next. This is quite similar in spirit to Riemann’s original approach to the theorem. Then we turn to Ahlfors’s generalization of Riemann’s theorem.

    APPENDIX: Traditional Proof of the Riemann Mapping Theorem

    Proof of (1): If U is bounded, then this assertion is easy: If we simply let a = 1/(2 sup{|z| : z U}) and b = –aP, then the function f(z) = az + b . If U is unbounded, we must work a bit harder. Since U , there is a point Q U. The function ϕ(z) = z Q is nonvanishing on U, and U is holomorphically simply connected. Therefore there exists a holomorphic function h such that h² = ϕ. Notice that h must be one-to-one since ϕ is. Also there cannot be two distinct points z1, z2 ∈ U such that h(z1) = –h(z2) [otherwise ϕ(z1) = ϕ(z2)]. Now h is a nonconstant holomorphic function; hence an open mapping. Thus the image of h contains a disc D(b, r). But then the image of h must be disjoint from the disc D(–b, r). We may therefore define the holomorphic function

    Since |h(z) – (–b)| ≥ r for z U, it follows that f maps U to D. Since h is one-to-one, so is f. Composing f with a suitable automorphism of D (a Möbius transformation), we obtain a function that is not only one-to-one and holomorphic with image in the disc, but also maps P .

    Proof of (2): so that

    . We want to claim that {fj} is a normal family. But of course each |fj| is bounded by 1, so Montel’s theorem applies.

    converge normally to a function f0, with

    We want to show that f0 is one-to-one into D. The argument principle, specifically Hurwitz’s theorem, will now yield this conclusion:

    Fix a point b U. Consider the holomorphic functions gj(z) ≡ fj(z) – fj(b) on the open set U\{b}. Each fj is one-to-one; hence each gj is nowhere vanishing on U \ {b}. Hurwitz’s theorem guarantees that either the limit function f0(z) – f0(b, it must hold that h′(P) ≠ 0 because if h′ (P) were equal to zero, then that h . Thus the function fand f0 cannot be constant. The only possible conclusion is that f0(z) – f0(b) is nowhere zero on U \ {b}. Since this statement holds for each b U, we conclude that f0 is one-to-one.

    Proof of (3): and suppose that there is a point R D such that the image of g does not contain R. Set

    Here we have composed g with a transformation that preserves the disc and is one-to-one. Note that ϕ is nonvanishing.

    The holomorphic simple connectivity of U guarantees the existence of a holomorphic function ψ : U such that ψ² = ϕ. Now ψ since it is nonvanishing. We repair this by composing with another Möbius transformation. Define

    Then ρ(P) = 0, ρ maps U into the disc, and ρ . Now we will calculate the derivative of ρ at P and show that it is actually larger in modulus than the derivative of g at P.

    We have

    Also,

    But g(P) = 0; hence

    We conclude that

    However 1 + |R| > 1 (since R , that

    Thus, if the mapping g of statement (3) at the beginning of the section were not onto, then it could not have property (2), of maximizing the absolute value of the derivative at P.

    We have completed the proofs of each of the three assertions and hence of the analytic form of the Riemann Mapping theorem. For completeness, we again enunciate the result.

    Prelude: This is the celebrated Riemann Mapping theorem (RMT). Our three steps converge now rather nicely to the proof of this result.

    THEOREM 1.3 Let U be a planar domain which is not the entire plane and which is analytically simply connected. Then there is a conformal map φ : U D from U to the unit disc.

    Certainly statements (2) and (3) taken together give us such a mapping which is both one-to-one and onto.

    The proof of statement (3) may have seemed unmotivated. Let us have another look at it. Let

    and

    Then, by our construction, with h = μS τ–1:

    Now the chain rule tells us that

    Since h(0) = 0, and since h is not a conformal equivalence of the disc to itself, the Schwarz lemma tells us that |h′(0)| must be less than 1. But this says that |g′(P)| < |ρ′(P)|, giving us the required contradiction.

    1.2A New Proof of the RMT

    Capsule: In a lecture given in the 1980s, W. Thurston proposed the idea that conformality can be understood in terms of circle packing. This idea captured the imagination of a number of mathematicians, and a new subject was quickly born.

    B. Rodin and D. Sullivan were fairly quickly able to create a circle-packing proof of the Riemann Mapping theorem. This result really put Thurston’s idea on the map, and also placed into focus some of the key techniques that would be widely used in this new study.

    In this section, we explain the proof of Rodin and Sullivan.

    There are many proofs of the Riemann Mapping theorem. Some of them are quite analytical—see, for instance, [BUR, pp. 293 ff.] Others are more geometric. We present one of those here.

    In fact, this proof stems from a lecture given by W. P. Thurston in 1985. The basic idea is that we can internally approximate a simply connected planar domain with a circle packing, and then produce a corresponding packing in the unit disc. The natural mapping of the nerves of the packings then gives an approximate conformal mapping. In the limit, the Riemann mapping is produced.

    Circle packing has given rise to an entire new approach to complex function theory. The elegant book [STE] is an homage to these new developments. B. Rodin and D. Sullivan [ROS] determined how to make Thurston’s ideas concrete, and how to provide a rigorous circle-packing proof of the Riemann Mapping theorem. That is the proof that we present here.

    1.2.1Some Definitions

    Let U be a planar domain. A circle packing in U is a collection of closed discs contained in U and having disjoint interiors. See Figures 1.1, 1.2, and 1.3 for a variety of circle packings.

    The nerve of the circle packing is the embedded 1-complex whose vertices are the centers of the discs and whose edges are the geodesic segments joining the centers of tangent discs and passing through the point of tangency. We shall restrict our attention in this discussion to a circle packing whose nerve is the 1-skeleton of a triangulation of some open, connected subset in the plane or the Riemann sphere. In particular, a circle packing of the sphere is one whose associated triangulation is a triangulation of the sphere. It is easy to see that, in a circle packing of the unit disc D, the carrier of the associated triangulation is a proper submanifold of the unit disc.

    A finite sequence of circles from a circle packing is called a chain if each circle except the last is tangent to its successor. The chain is a cycle if the first and last circles are tangent.

    Figure 1.1: A circle packing.

    Figure 1.2: A square and a hexagonal circle packing.

    Now let c be a circle in a circle packing. The flower centered at c is the closed set consisting of c and its interior, together with all circles tangent to c and their interiors, and also the interiors of all the triangular interstices formed by these circles.

    We shall not prove the next result, but instead refer the reader to [AND1], [AND2], [THU, Chapter 13], and [ROS, Appendix 2].

    Figure 1.3: Another circle packing.

    Prelude: This theorem is geometrically central to the theory. It tells us that any triangulation that we will encounter comes from a circle packing. It also gives an important uniqueness statement.

    THEOREM 1.4 Any triangulation of the sphere is isomorphic to the triangulation associated to some circle packing. The isomorphism can be required to preserve the orientation of the sphere and then this circle packing is unique up to Möbius transformations.

    A topological annulus A in the plane has a modulus mod A that can be defined, without reference to conformal mapping, as the infimum of the L² norms of all Borel measurable functions ρ on the plane such that 1 ≤ ∫ ρ(z) |dz| along all degree one curves in A. This is closely related to the idea of extremal length; see [AHL1] and also our Chapter 7.

    Definition 1.5 An orientation-preserving homeomorphism f between two planar domains is called K–quasiconformal, 1 ≤ K < ∞, if

    for every annulus A in the domain of f.

    Intuitively, a K-quasiconformal mapping does not distort the picture by more than a factor of K. Some useful facts about quasiconformal mappings are these:

    (1)K-quasiconformality is a local property (see [AHL3, page 22]).

    (2)A 1-quasiconformal map is conformal and conversely.

    Further, we need the fact that a simplicial homeomorphism is K-quasiconformal for K depending only on the shapes of the triangles involved.

    1.2.2A Sketch of Thurston’s Idea

    Let U be a simply connected, proper subdomain of the plane. Now efficiently fill U with small circles from a regular, hexagonal circle packing. Surround these circles by a Jordan curve. Use Andreeev’s theorem to produce a combinatorially equivalent packing of the unit disc, with the unit circle corresponding to the Jordan curve. What we hope is that the correspondence between the circles of the two packings will approximate the Riemann mapping.

    1.2.3Details of the Proof

    LEMMA 1.6 (THE RING LEMMA) There is a constant r depending only on k such that, if k circles surround the unit disc (i.e., they form a cycle externally tangent to the unit disc), then each circle has radius at least r. See Figure 1.4.

    Figure 1.4: The ring lemma.

    Proof of the Lemma: Fix k. There is a uniform lower bound for the radius of the largest outer circle c1 (this occurs when all the k outer circles have equal radius). A circle c2 adjacent to c1 also has a uniform lower bound for its radius because, if c2 were extremely small, then a chain of k – 1 circles starting from c2 could not escape from the crevice between c1 and the unit circle. Repeat this reasoning for the circle c3 adjacent to c2, and so forth.

    LEMMA 1.7 (THE LENGTH-AREA LEMMA) Let c be a circle in a circle packing of the unit disc. Let S1, S2,…, Sk be k disjoint chains which separate c from the origin and from a point of the boundary of the disc. Denote the combinatorial lengths of these chains by n1, n2,… nk. Then

    Proof: Suppose that the chain Sj consists of circles of radii rjℓ. Then, by the Schwarz inequality,

    Let sj = ∑ℓrjℓ be the geometric length of Sj. We find that

    and

    Therefore

    satisfies

    Since s is greater than the diameter of c, this last inequality proves the lemma.

    LEMMA 1.8 (THE HEXAGONAL PACKING LEMMA) There is a sequence of real numbers sj, decreasing to zero, with the following property. Let c0 be a circle in a finite packing P of circles in the plane, and suppose that the packing P around c0 is combinatorially equivalent to n generations of the regular hexagonal circle packing about one of its circles. Then the ratio of radii of any two circles in the flower around c differs from unity by less than sn.

    COROLLARY 1.9 A circle packing in the plane with the hexagonal pattern is the regular hexagonal packing.

    Proof of the lemma: Suppose that, for each n = 1, 2,…, we have a circle packing Pn which is combinatorially equivalent to n generations of the regular hexagonal circle packing centered around the unit circle c0. We apply the Ring Lemma to the Pns. This shows that the radii of circles of generation k in Pk, Pk+1,… are uniformly bounded away from zero and infinity. Therefore we can select a subsequence of the Pjs so that the generation one circles converge geometrically. A further subsequence can be selected so that the generation two circles converge geometrically, and so forth.

    In this fashion, a limit infinite packing of a planar domain is obtained. This packing has the combinatorics of the regular hexagonal packing, and the domain is an increasing sequence of discs and so is connected and simply connected.

    If the lemma were false, then we could select the subsequences so that, in the limit packing, one of the six circles around c0 would have radius different from 1. This contradicts the uniqueness of packings in the plane with the pattern of the hexagonal packing.

    REMARK 1.10 It appears that we are using the corollary to prove the lemma, but an independent proof of the corollary can be given. See Appendix 1 in [ROS].

    1.2.4Convergence of Circle Packings to the Riemann Mapping

    Let U be a simply connected, bounded domain in the plane with two distinguished points z0, z1 ∈ U> 0, consider the regular hexagonal packing H . Let c0 be a circle whose flower contains zof circles from H , starting with c0, such that the flowers of the circles in the chain are contained in U. The circles which appear in such chains are called inner circles. The set of inner circles is denoted by I .

    The circles in H which are not inner circles but which are tangent to inner circles will be called border circles. The set B of border circles can be cyclically ordered to form a cycle called the border. The border has the property that the linear polygon obtained by joining the centers in order is a Jordan curve surrounding the inner circles. The inner circles I and border circles B form a packing C whose nerve is the 1-skeleton of a triangulation T .

    Complete T of the sphere by adding a vertex at ∞ plus disjoint (except for the point at ∞) Jordan arcs from ∞ to the

    Enjoying the preview?
    Page 1 of 1