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

Only $11.99/month after trial. Cancel anytime.

Conformal Mapping on Riemann Surfaces
Conformal Mapping on Riemann Surfaces
Conformal Mapping on Riemann Surfaces
Ebook524 pages4 hours

Conformal Mapping on Riemann Surfaces

Rating: 3 out of 5 stars

3/5

()

Read preview

About this ebook

The subject matter loosely called "Riemann surface theory" has been the starting point for the development of topology, functional analysis, modern algebra, and any one of a dozen recent branches of mathematics; it is one of the most valuable bodies of knowledge within mathematics for a student to learn.
Professor Cohn's lucid and insightful book presents an ideal coverage of the subject in five parts. Part I is a review of complex analysis analytic behavior, the Riemann sphere, geometric constructions, and presents (as a review) a microcosm of the course. The Riemann manifold is introduced in Part II and is examined in terms of intuitive physical and topological technique in Part III. In Part IV the author shows how to define real functions on manifolds analogously with the algebraic and analytic points of view outlined here. The exposition returns in Part V to the use of a single complex variable z. As the text is richly endowed with problem material — 344 exercises — the book is perfect for self-study as well as classroom use.
Harvey Cohn is well-known in the mathematics profession for his pedagogically superior texts, and the present book will be of great interest not only to pure and applied mathematicians, but also engineers and physicists. Dr. Cohn is currently Distinguished Professor of Mathematics at the City University of New York Graduate Center.

LanguageEnglish
Release dateMay 5, 2014
ISBN9780486153292
Conformal Mapping on Riemann Surfaces

Related to Conformal Mapping on Riemann Surfaces

Titles in the series (100)

View More

Related ebooks

Mathematics For You

View More

Related articles

Reviews for Conformal Mapping on Riemann Surfaces

Rating: 3 out of 5 stars
3/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Conformal Mapping on Riemann Surfaces - Harvey Cohn

    SURFACES

    PART I

    REVIEW OF COMPLEX ANALYSIS

    INTRODUCTORY SURVEY

    We first undertake a critical review of introductory complex function theory for the purpose of examining the ideal case of the Riemann sphere (or the ordinary plane with a point at infinity adjoined). We make every effort to think of this two-dimensional plane (or sphere) as being represented by a single complex variable, say z.

    Here we find that, in a certain sense, those functions of z which belong in a natural way to the sphere are rational. This is significant from two seemingly opposite viewpoints:

    (a) Algebraic. The rational functions f(z) can first of all be handled formally, through high-school-level identities in z, without any apparent concern for the magnitudes of z or f(z) or for analysis (limiting operations, differentiation, integration, etc.).

    (b) Analytic. The rational functions f(z) are now determined by their (limiting) magnitude where they become infinite. This so-called singular behavior determines the functions completely, as though normal (analytic) behavior were unimportant.

    We find that to fully appreciate this condition aesthetically, we must explore the z plane as a two-dimensional (real) xy plane, as is done in Part III. Such considerations were first introduced as conformal mapping based on strong ties with applications; but the proper arena for such investigations is the Riemann manifold, which is introduced in Part II and is examined in terms of intuitive physical and topological techniques in Part III. In Part IV we show how to define real functions on manifolds analogously with the algebraic and analytic points of view outlined here. In Part V we return to the use of a single complex variable z.

    Thus Part I presents (as a review) the course in microcosm.

    CHAPTER 1

    ANALYTIC BEHAVIOR

    At first we restrict our interest to functions which are well behaved where denned. All well-behaved functions are equally well behaved under this restriction, for assuming differentiability in a neighborhood of a point, we find a valid power series to exist.

    DIFFERENTIATION AND INTEGRATION

    1-1 Analyticity

    Let us start with the concept of a real function u(x) defined in an interval I: (a < x < b). There is a certain interest in degrees of regularity of the function u(x) as represented by the smoothness of the curve or its tangent, etc. Thus we consider restrictions of u(x) to classes C, C′, C″, …, C(∞), A as follows:

    C: u(x) is continuous for all x in I .

    C′: u ′( x ) exists and is continuous for all x in I .

    C″: u ″( x ) exists and is continuous for all x in I .

    C(∞): u ( n ) ( x ) exists and is continuous for all x in I regardless of n .

    A: u ( x ) is representable by a convergent Taylor series about each x in I .

    It is clear, a priori, that each restriction implies the preceding ones, and Exercise 1 will indicate that each restriction is satisfied by fewer functions than the preceding one.

    (or i)? We write the independent variable z in terms of two independent real variables x and y, as

    and we write the dependent variable symbolically as w(z); and in terms of two real functions u(x,y) and v(x,y) we write w(z) as

    Thus in real terms a complex function is a two-dimensional surface [x,y,u(x,y),v(x,y)] in a real four-dimensional (x,y,u,v) space. We could speak of smoothness of u(x,y) and v(x,y) in terms of tangents to this four-dimensional surface, but this type of smoothness has no interest for us. We are considering the stronger condition of differentiability of w in terms of z.

    We call a function w(z) analytic in a regionif the derivative w′(z) exists‡ for each point of the region. The function is then called analytic at each point of the region. We do not define analyticity at a point except by defining it in a surrounding region. (Synonyms for analytic are regular and holomorphic.)

    THEOREM 1-1   An analytic function in has derivatives of all orders in and is representable as well by a convergent power series expansion at each point of the region within any circular disk centered at that point and lying in the region.

    This theorem is practically the contents of an entire course in complex variables! Its importance is evident since it essentially vitiates all regularity distinctions once the derivative is presupposed. The proof of this theorem is effected by the rather ingenious well-known integration procedure which leads to the existence of power series.

    Strangely enough, despite the fact that Theorem 1-1 does not mention integration, it has been proved classically§ only by integration. Perhaps then, rather than become overly impressed by this matter of semantics, we should do better to conclude that, from a more mature point of view, differentiation and integration are not so very different but are part of, say, some generalized study of linear operators.

    EXERCISES

    1 Show how to form real functions u ( x ) with all degrees of regularity desired above. { Hint: Show U t ( x ) = [exp (– 1/ x ² )]/ x t vanishes as the real variable x → 0 for any positive integer t ; show U 0 ( x ) [with U 0 (0) = 0] serves as a function with all orders of derivative but no power series representation at the origin.}

    2 Find a function w ( z ) = u ( x,y ) + iv ( x,y ) with a complex derivative at every point on the real axis but which is not analytic in any neighborhood of a real point. [ Hint: Use U 0 ( y ).]

    3 = g ( z whenever f ( z ) is analytic at z 0 .

    1-2 Integration on Curves and Chains

    We introduce integration by defining a smooth curve segment by differentiate functions parametrized by arc length s

    Here dx² + dy² = ds² and l is the length in usual elementary fashion. A piecewise smooth curve (or more simply a curveare two curve segments with

    then if x1(l1) = x2(0) and y1(l1) = y, writing the sum (purely formally) as

    It is clear how any number of segments

    to denoteparametrized in reverse fashion (with l s replacing s if l is its orientationbut one which returns to its origin.

    as x + iy = z(s). If f(z, then we define the integral by the Riemann sum as follows:

    has length l and 0 = s0 < S1 < · · · < sm = l is a partition of l with maximum interval (sj Sj–1) = δ and Zj = x(si + iy(si); and ζj is an arbitrary point on the curve segment from zj–1 to zj. The existence of the limit is shown in elementary texts.

    Now when we consider the integral (1-6), we become aware of the fact that retracing the path of integration nullifies the integral; for example,

    We therefore introduce a purely abstract concept to make manifest this almost trivial property.

    We consider the formal additive group of finite linear combinations of curves with integral coefficients. This group contains expressions such as

    . We say that two such formal expressions constitute the same chain are called 1-chains because of the dimensionality of curves.

    The purpose of all this formalism is the following idea: We can write

    and we can think of the above expression as being bilinear (or linear separately) in the differentialf(z) dz . The essence of the bilinearity is our ability to write

    and at the same time

    .

    as either a curve or a chain as the occasion requires.

    The important integral estimate acquires a special symmetry when we use the bilinear form (1-7)

    where l|f| is the supremum of |f(z)| for z designates a curve (not a chain).

    EXERCISES

    1 Show that two piecewise smooth curves can have an infinitude of isolated points of intersection. In fact this is true even for convex segments (segments where d ² y / dx ² is positive on each curve). [ Hint: For –1 ≤ x ≤ 1 consider the curves y 1 ( x ) = 20 x ² + x ⁵ sin (1/ x )( x ≠ 0), y 1 (0) = 0, and y 2 ( x ) = 20 x ² .] Also verify that they satisfy the smoothness condition C″.

    2 Prove the integral estimate (1-9) by direct appeal to the definition (1-6).

    1-3 Cauchy Integral Theorem

    We now think of curves in terms of the boundary of a region. We define a closed curve as one for which x(0) = x(l), y(0) = y(l) where l . We define a simple curve as a curve with no self-intersections except in the case of a closed curve which meets itself at its end points. Symbolically, if l is the length and if 0 ≤ s1 < s2 ≤ l, then (x(s1), y(s1)) = (x(s2), y(s2)) only (possibly) when s1 = 0, s2 = l.

    ) if the set of boundary points|. This is a concept we shall surely have to enlarge, but for the time being, it enables us to express theorems on integration.

    THEOREM 1-2   (Cauchy-Goursat) Let f(z) be analytic in a region which contains a simple closed curve and a subregion for which is the boundary; then

    This theorem was proved by Cauchy (1825) under a (really innocuous) variant of the concept of analyticity, requiring that f′(zbut also be continuous there. Goursat (1900) proved the stronger result by a proof which also had the advantage of keeping the argument in terms of complex quantities (avoiding the explicit use of f = u + iv, dz = dx + i dy, etc.).

    Although the sequence of proofs is somewhat interrupted, this is a good place to insert certain corollaries to Theorem 1-2.

    COROLLARY 1   If z1 and z2 are two points interior to can be defined independently of the curve joining z1 to z2 as long as the curve lies in .

    COROLLARY 2   (Morera’s theorem, 1886) If f(z) is a continuous function of z in and if

    about any closed curve in , then f(z) represents an analytic function in .

    Let us recall the proof of this last result. On the basis of Corollary 1, we can define the unique function

    where a and z . Then F′(z) exists [= f(z. Hence F(z) is analytic, and (by Theorem 1-1) so is its derivative f(z), which completes the proof. The naïvety of this device is in the characteristic vein of classical complex variable theory.

    COROLLARY 3   If fn(z) is a sequence of analytic functions converging uniformly to a limit function f(z) in a region as n → ∞, then f(z) is analytic in .

    , we apply Morera’s theorem to the following result (based on uniform convergence and Theorem 1-2):

    Of course, if f(z. The best known counterexample, of course, is

    increases by 2π with the parametrization. To see (1-14) as a triviality, it suffices to look upon the integrand dz/z asd log z and write

    meaning that if z is parametrized as z(s) = x(s) + iy(s) by a continuous variation of s from 0 to l (the length),

    Here we note that d log z is always single-valued as a differential despite‡ the fact that the indeterminacy of log z can be any number of the type 2πim for m a positive or negative integer.

    Equation (1-14) is important enough to be sometimes stated as a theorem. For our purposes, it is superseded by Cauchy’s residue theorem (see (simple or not) and any point a , we note that it enables us to define the following function:

    with respect to a. All we can say for now is that when is closed, the values of N;a) are integers (because the indeterminacy of the argument is in multiples of 2π).

    The expression N;aand a(although the term functional is traditionally more appropriate). Clearly, N depends continuously on a unless a (none passing through a) were parametrized continuously by parameter t, then N;a) would vary continuously with tis closed, N takes on discrete values 0, ±1, ±2, …. Thus the winding number for a closed curve about a point a remains constant, as and a vary continuously, as long as a is never on .

    Later on we shall see that the logarithm function essentially characterizes the geometry of the plane (rather than conversely)! Indeed, we generalize the plane to other geometrical entities and the logarithm function to so-called abelian integrals, but only for the plane will we be able to speak of anything like a winding number.

    EXERCISES

    1 Let f ( z ) = x iy . Test Morera’s theorem. Also show directly that f ′(0) does not exist.

    2 ( a ) (Fundamental theorem of calculus.) In Morera’s theorem show F ′( z ) = f ( z ) by taking the limit of

    . (bis restricted to an arbitrary triangle.

    3 Show that the definition of a region can be taken equivalently as an open set which is arcwise connected (such that any two points are connected by an arc lying in the set).

    TOPOLOGICAL CONSIDERATIONS

    1-4 Jordan Curve Theorem

    We have asserted that Theorem 1-1 on differentiation depends on Theorem 1-2 on integration. Equally remarkable is the fact that Theorem 1-2 (as well as its relation to Theorem 1-1) depends on another concept not yet stated, namely, the idea that a boundary of a region (although defined as a set of boundary points) actually, in some sense, encloses a region.

    Let a smooth be parametrized as z(s) = x(s) + iy(slies to the left for the given parametrization if for each point z(s, which is given by

    . More precisely, this means that the point

    if for every s there exists a real number Δ(s) > 0, such that

    .

    is called "n-tuply-connected (simply-connected" for n is called a disk for n = 1, an annulus, for n = 2.

    k (1 ≤ k nlies to the leftk ) is given formally as a chain

    .

    THEOREM 1-3   (Jordan curve theorem, 1887) Every simple closed curve in the z plane determines a finite region for which

    If we say that is positively oriented.

    COROLLARY 1   The complement of a simple closed curve consists of two disjoint regions, the finite region and an infinite region .

    COROLLARY 2   If is an n-tuply-connected region with boundary set consisting of the disjoint simple closed curves , then with proper choice of signs

    Thus in (the Cauchy-Goursat) .

    A proof of the Jordan curve theorem is a standard part of topology and will not be reproduced here. We consider only piecewise smooth curves so that the full strength of topology is not called into play.

    The relationship between the Jordan curve theorem and complex variable theory is essentially established through the winding number N;a) defined in (1-16).

    THEOREM 1-4   If for a simple closed curve , then

    (A sketch of the proof of Theorem 1-4 using the Jordan curve theorem is given as Exercises 2 to 4.)

    Thus Theorem 1-2 can be restated in terms of N;a).

    THEOREM 1-5   Let f(z) be analytic in a region . Consider a simple closed curve in such that N;a) = 0 for all a . Then

    This statement of the Cauchy integral theorem is the work of Ahlfors (1953).

    EXERCISES

    Assume the Jordan curve theorem (Theorem 1-3) for the following exercises.

    1 Deduce the corollaries of Theorem 1-3 . [ Hint: t t t by means of the ∪, ∩, and c (complement) symbols.

    2 Prove that N ; a and a point a sufficiently far from the origin, such that for every Z , | Z a | > l )/2π. ( Hint: Use the integral estimate of Sec. 1-2 above.) Prove that N ; a ) = 0 for a |.

    3 Prove that N ; a ) changes by 1 as the variable point a . For convenience, let point a along a normal at point A . [ Hint: containing A ).] Then, writing

    N;a) = N;a) + N;a)

    show that N;a) must change by 1 and N2; a) must change by 0, as a at A. [Hint: The hard part is to show N;a) changes by 1; but consider two positions of a .]

    4 Complete the proof of Theorem 1-4 .

    5 Prove Theorem 1-5 .

    1-5 Other Manifolds

    To appreciate the role of the Jordan curve theorem, it is desirable to imagine a suitable situation where it does not hold.

    We introduce the concept of an n-dimensional manifold as a connected point set in, say, k(≥n) dimensional euclidean space which "looks like n-dimensional euclidean space at each point." Technically speaking, this means that nto create a one-to-one correspondence with a neighborhood of the origin in n-dimensional euclidean space such that the correspondence and its inverse are continuous.† For example, a region of a plane is trivially a two-dimensional manifold because it is open, and a unit circle in the plane is a one-dimensional manifold. A dihedral angle formed by half planes in three-dimensional euclidean space is capable of being opened flat and hence is a two-dimensional manifold.

    To think of some nonmanifolds, take the union of two intersecting lines, say the real and imaginary axes (positive and negative). Here, a neighborhood of their point of intersection is a set consisting of four intersecting rays. It is not possible to construct a homeomorphism of it with a neighborhood of the origin† in one-dimensional space, e.g., with –ε ≤ x . It consists of two manifolds, the (open) disk |z| < 1 and the circle |zis, of course, not open; e.g., the points on |z| = 1 do not have the (full) neighborhoods in the closed disk as required by the definition.

    We use the generic term variety to describe a geometric set which is not a manifold. The term algebraic variety is used if the set is described by an algebraic system of equations (such as the union of the x and y axes which is described by xy = 0).

    We do not write any more formal definitions at this point since analytic in Theorem 1-3 essentially because of two specifications:

    (a(orientability).

    (b(separation).

    There are two well-known manifolds which show how Jordan’s theorem might fail on either specification.

    MÖBIUS STRIP

    given by coordinates (in the xy plane)

    We identify the points (0, –y) and (l,+yas a thin paper strip (see Fig. 1-1a) and the ends (x = 0 and x = l) are tied after a 180° twist [identifying (0, –y) with (l,+y)].

    Fig. 1-1 Möbius strip

    by going to the right of x = 1 (see long arrow).

    The curve (in the xy plane again)

    contains only points where y > 0, but these points still connect with points where y < 0 after we go from x = l to x = 0 (as shown by the long arrow in Fig. 1-1b).

    . We shall ultimately see that this situation can be completely avoided in complex variable theory.

    TORUS

    Consider the surface of revolution of, say, the circle

    about the ζ axis in (real) (ξ, η, ζ) space. Then, these loci

    : azimuthal circles (on ζ = const)

    constitute simple closed curves which still do not serve as boundaries of any region.

    (see Fig. 1-2b).

    Fig. 1-2 Torusjoins them to the right-hand ε neighborhood. The torus (a(so that there are two replicas of each) and the torus is flattened out (b).

    and then imagine it opened up and laid out flat on the rectangle measuring, say, 2πa by 2πb [allowing intuitively for some slight stretching (see Fig. 1-2aappears as both the upper and lower side of the rectangle identified (see long arrow in Fig. 1-2bappears as both the right- and left-hand side of the rectangle identified. The points B and D show the identification; the point A is the same at all four corners. In other words, the torus can be laid out as the rectangle and opposite sides identified without twisting; the Möbius strip involves twisting (in some intuitive fashion).

    1-6 Homologous Chains

    The term topological will be used rather loosely to mean pertaining to properties of sets, functions, etc., which are invariant under one-to-one continuous mappings with continuous inverses. (Such mappings are called bicontinuous, topological, homeomorphic, etc.) This interpretation of the term† is in close keeping with analysis.‡

    Fig. 1-3 Illustration for remarks on homology.

    Let us view a very familiar situation in the light of topological concepts by referring to Fig. 1-3. Here a function f(z. Then we would say

    (as shown in Fig. 1-3b), and in this manner we would take

    (by the cancellation of retraced portions in usual elementary fashion).

    . It is quite convenient to formalize this.

    } such that

    is bounded by a finite number of simple closed curves.

    ].

    .

    We next define† formal chains of regions or 2-chains:

    }. We consider chains to be equivalent , as in Fig. 1-3b, ignoring the common boundary. (This is a situation analogous to the ignoring of points of intersections when we assert the equivalence of the 1-chains referred to in Sec. 1-2.)

    denote a 2-chain, then we can define

    as the boundary is unaffected by equivalence operations; this amounts to realizing that (in Fig. 1-3bon the left.)

    ,

    Now Cauchy’s integral theorem can be restated as

    THEOREM 1-2 (bis) Let f(z) be analytic in and let be a 2-chain in . Then vanishes or

    and therefore the additive property (1-27c) gives us (1-28).

    are homologous or

    . We can then rewrite the Cauchy integral theorem as follows:

    ) can be filled in by a region. Thus, returning to Fig. 1-3a, etc.

    The ideas introduced in Sec. 1-2 and here do not reach fruition until analogs are developed for differentials on a manifold in Chaps. 7 and 8. But proceeding in a purely formal way, let us observe that the boundary operator ∂ has a natural application to 1-chains:

    Enjoying the preview?
    Page 1 of 1