MATH 566 LECTURE NOTES 6: NORMAL FAMILIES AND THE THEOREMS OF PICARD

Similar documents
MATH 566 LECTURE NOTES 4: ISOLATED SINGULARITIES AND THE RESIDUE THEOREM

RIEMANN MAPPING THEOREM

Chapter 6: The metric space M(G) and normal families

MATH 722, COMPLEX ANALYSIS, SPRING 2009 PART 5

CONSEQUENCES OF POWER SERIES REPRESENTATION

An Introduction to Complex Analysis and Geometry John P. D Angelo, Pure and Applied Undergraduate Texts Volume 12, American Mathematical Society, 2010

4.6 Montel's Theorem. Robert Oeckl CA NOTES 7 17/11/2009 1

III. Consequences of Cauchy s Theorem

Complex Analysis Qualifying Exam Solutions

Analysis III. Exam 1

Part IB Complex Analysis

Course 214 Basic Properties of Holomorphic Functions Second Semester 2008

Houston Journal of Mathematics. c 2008 University of Houston Volume 34, No. 4, 2008

11 COMPLEX ANALYSIS IN C. 1.1 Holomorphic Functions

Complex Analysis Problems

MATH 566 LECTURE NOTES 1: HARMONIC FUNCTIONS

Riemann Mapping Theorem (4/10-4/15)

Theorem [Mean Value Theorem for Harmonic Functions] Let u be harmonic on D(z 0, R). Then for any r (0, R), u(z 0 ) = 1 z z 0 r

MA3111S COMPLEX ANALYSIS I

Solutions to Complex Analysis Prelims Ben Strasser

5.3 The Upper Half Plane

Complex Analysis Qual Sheet

The result above is known as the Riemann mapping theorem. We will prove it using basic theory of normal families. We start this lecture with that.

8 8 THE RIEMANN MAPPING THEOREM. 8.1 Simply Connected Surfaces

3. 4. Uniformly normal families and generalisations

Math 520a - Final take home exam - solutions

Holomorphic Dynamics Part 1. Holomorphic dynamics on the Riemann sphere

Math 220A - Fall Final Exam Solutions

MORE CONSEQUENCES OF CAUCHY S THEOREM

RANDOM HOLOMORPHIC ITERATIONS AND DEGENERATE SUBDOMAINS OF THE UNIT DISK

THE RIEMANN MAPPING THEOREM

Tools from Lebesgue integration

COMPLETELY INVARIANT JULIA SETS OF POLYNOMIAL SEMIGROUPS

NATIONAL UNIVERSITY OF SINGAPORE Department of Mathematics MA4247 Complex Analysis II Lecture Notes Part II

REAL ANALYSIS ANALYSIS NOTES. 0: Some basics. Notes by Eamon Quinlan. Liminfs and Limsups

FINAL EXAM MATH 220A, UCSD, AUTUMN 14. You have three hours.

Hartogs Theorem: separate analyticity implies joint Paul Garrett garrett/

Part IB. Further Analysis. Year

C4.8 Complex Analysis: conformal maps and geometry. D. Belyaev

PICARD S THEOREM STEFAN FRIEDL

Qualifying Exam Complex Analysis (Math 530) January 2019

Part III. 10 Topological Space Basics. Topological Spaces

is holomorphic. In other words, a holomorphic function is a collection of compatible holomorphic functions on all charts.

Complex Analysis Important Concepts

COMPLEX ANALYSIS Notes Lent 2006

z b k P k p k (z), (z a) f (n 1) (a) 2 (n 1)! (z a)n 1 +f n (z)(z a) n, where f n (z) = 1 C

POWER SERIES AND ANALYTIC CONTINUATION

Lecture 7 Local properties of analytic functions Part 1 MATH-GA Complex Variables

Complex Analysis Math 185A, Winter 2010 Final: Solutions

Math 209B Homework 2

b 0 + b 1 z b d z d

Complex Analysis for F2

NOTES ON EXISTENCE AND UNIQUENESS THEOREMS FOR ODES

Complex Analysis Homework 9: Solutions

Conformal Mappings. Chapter Schwarz Lemma

RIEMANN SURFACES. max(0, deg x f)x.

9. Series representation for analytic functions

The Arzelà-Ascoli Theorem

MORE NOTES FOR MATH 823, FALL 2007

Definition We say that a topological manifold X is C p if there is an atlas such that the transition functions are C p.

INTRODUCTION TO REAL ANALYTIC GEOMETRY

Continuous Functions on Metric Spaces

Part II. Riemann Surfaces. Year

WEIERSTRASS THEOREMS AND RINGS OF HOLOMORPHIC FUNCTIONS

Complex Analysis Slide 9: Power Series

Complex Analysis, Stein and Shakarchi Meromorphic Functions and the Logarithm

LAURENT SERIES AND SINGULARITIES

MATH COMPLEX ANALYSIS. Contents

Functional Analysis. Franck Sueur Metric spaces Definitions Completeness Compactness Separability...

Supplement. The Extended Complex Plane

Chapter 30 MSMYP1 Further Complex Variable Theory

Notes on Complex Analysis

Analysis Comprehensive Exam, January 2011 Instructions: Do as many problems as you can. You should attempt to answer completely some questions in both

Riemann sphere and rational maps

VII.5. The Weierstrass Factorization Theorem

Accumulation constants of iterated function systems with Bloch target domains

3 Fatou and Julia sets

Complex Analysis review notes for weeks 1-6

Journal of Inequalities in Pure and Applied Mathematics

Kwack s theorem and its applications

MATH8811: COMPLEX ANALYSIS

A FEW ELEMENTARY FACTS ABOUT ELLIPTIC CURVES

MA30056: Complex Analysis. Exercise Sheet 7: Applications and Sequences of Complex Functions

COMPLEX ANALYSIS Spring 2014

From the definition of a surface, each point has a neighbourhood U and a homeomorphism. U : ϕ U(U U ) ϕ U (U U )

Solutions to practice problems for the final

4 Uniform convergence

An introduction to holomorphic dynamics in one complex variable Informal notes Marco Abate

Siegel Discs in Complex Dynamics

7.2 Conformal mappings

LECTURE 15: COMPLETENESS AND CONVEXITY

Chapter 4: Open mapping theorem, removable singularities

Problem Set 5. 2 n k. Then a nk (x) = 1+( 1)k

Considering our result for the sum and product of analytic functions, this means that for (a 0, a 1,..., a N ) C N+1, the polynomial.

Math 259: Introduction to Analytic Number Theory Functions of finite order: product formula and logarithmic derivative

Complex Variables Notes for Math 703. Updated Fall Anton R. Schep

Some Background Material

Bernstein s analytic continuation of complex powers

arxiv: v2 [math.ds] 9 Jun 2013

f (n) (z 0 ) Theorem [Morera s Theorem] Suppose f is continuous on a domain U, and satisfies that for any closed curve γ in U, γ

Transcription:

MATH 566 LECTURE NOTES 6: NORMAL FAMILIES AND THE THEOREMS OF PICARD TSOGTGEREL GANTUMUR 1. Introduction Suppose that we want to solve the equation f(z) = β where f is a nonconstant entire function and β C. We know that if f is a polynomial, this equation has a solution for all β C. On the other hand, the equation exp z = β has no solution if β = 0. It turns out that in some sense this is the worst that can happen, in that f(z) = β is solvable for all β from C with a possible exception of one point. Theorem 1 (Picard s little theorem). If f O(C) and f(c) U where C \ U contains at least two points, then f is constant. This is a dramatic strengthening of Liouville s theorem, as Liouville s theorem can be thought of as the preceding theorem where the condition on U is replaced by the condition U = D. We will actually prove an even stronger result known as Picard s big theorem. Definition 2. Let S and M be Riemann surfaces, and let f O(S \ {α}, M) with α S. We say β M is an omitted value of f at α, if there is an open neighbourhood U S of α such that f(u) β. So in order for β not to be an omitted value, there must exist a sequence {z n } S with z n α such that f(z n ) = β for all n. Theorem 3 (Picard s big theorem). If f O(S \ {α}, Ĉ) omits 3 values from Ĉ at α then f extends to f O(S, Ĉ). In certain sense, this theorem strengthens Riemann s removable singularity theorem the same way the little theorem strengthens Liouville s theorem. Another way to formulate the big theorem is to say that a holomorphic function can omit at most one value at its essential singularity. We will give a proof at the end of these notes. For now, let us see how the little theorem follows from the big theorem. It suffices to treat the case where f is not a polynomial, since a polynomial of degree n takes every value in C exactly n times, counting multiplicities. It turns out that we can derive a result that is much stronger than the little theorem, namely that not only entire transcendental (i.e., non-polynomial) functions assume every value in C with at most one exception, but each of those values are taken infinitely many times. Corollary 4. If f O(C) is not a polynomial, then it assumes every value in C infinitely many times with at most one exception. Indeed, since f has an essential singularity at Ĉ, it cannot be extended to a meromorphic function on Ĉ. By the big theorem, at Ĉ, f can omit at most 2 values in Ĉ, or, since is already omitted, at most 1 value in C. For all other values β in C, we have a sequence {z n } C with z n such that f(z n ) = β for all n. In order to prove the big theorem, we will develop in the following sections a very precise theory on compactness properties of families of meromorphic functions. Date: December 7, 2010. 1

2 TSOGTGEREL GANTUMUR 2. Marty s theorem Obviously, a proof of Picard s big theorem requires us in one way or another to consider holomorphic functions having values in Ĉ (otherwise known as meromorphic functions). When working with Ĉ one has to deal with the point Ĉ from time to time, and as a result discussions become spotted with applications of special transformations such as 1/z, potentially obscuring the clarity of the arguments and giving the false impression that was somehow special. This can be dealt with by introducing general charts on Ĉ, but then its complexity becomes the same as considering a general Riemann surface. So for the sake of both clarity and generality we will be working with a general Riemann surface M instead of the Riemann sphere Ĉ. We equip M with a (metric space) metric ρ that is compatible with its topology. This metric induces a way to measure distances between functions: for f, g : U M defined on some set U C, we define ρ U (f, g) = sup ρ(f(z), g(z)). z U With Ω C an open connected set, we say that the sequence {f n } of functions f n : Ω M converges to f : Ω M uniformly on Ω if ρ Ω (f n, f) 0 as n, and that {f n } converges to f locally uniformly on Ω if for any compact K Ω, ρ K (f n, f) 0 as n. For each coordinate chart (U, ϕ) of M, we require the metric ρ to satisfy ρ(z, w) = C z ϕ(z) ϕ(w) + o( ϕ(z) ϕ(w) ), z, w U, where the constant C z > 0 depends continuously on z. Note that C z is in general different for different coordinate charts. If f : Ω M is a holomorphic function, then for z Ω we have ρ(f(z), f(z + h)) = C f(z) ϕ(f(z + h)) ϕ(f(z)) + o( ϕ(f(z + h)) ϕ(f(z)) ) and therefore the following limit exists = C f(z) (ϕ f) (z) h + o( h ), f ρ(f(z), f(z + h)) (z) = lim. h 0 h We call f : Ω R + the metric derivative of f. It is clear that f has the local expression which in particular shows that f is a continuous function. f (z) = C f(z) (ϕ f) (z), (1) Example 5. For the Riemann sphere Ĉ, an example of ρ satisfying the above assumptions is the cordial distance between α, β Ĉ considered as points on S2 R 3. Another is the geodesic distance between two points α, β S 2 with respect to the round metric on S 2. The following is a generalization of the Weierstrass convergence theorem. Theorem 6. Let {f n } O(Ω, M) be a sequence that converges locally uniformly on Ω to a function f : Ω M. Then f O(Ω, M), and {f n} converges locally uniformly. Proof. We first prove that f is continuous. From the triangle inequality for ρ, we have ρ(f(z), f(w)) ρ(f(z), f n (z)) + ρ(f n (z), f n (w)) + ρ(f n (w), f(w)), for z, w Ω and n N. For any compact set K Ω, one can choose n so large that ρ K (f n, f) is arbitrarily small, and then since f n is continuous we conclude that f is continuous. By continuity of f, for any z Ω there is an open neighbourhood U Ω such that f(u) is entirely in a single coordinate chart of M. Then for any open disk D U centred at z such that D U, the sets f n (D) will eventually be in the same coordinate chart, so that ρ D (f n, f) 0 implies ϕ f n ϕ f D 0, where ϕ is the coordinate map. Consequently,

NORMAL FAMILIES 3 we have ϕ f O(D), and since D is an open disk centred at an arbitrary point z Ω, we get f O(Ω, M). At the same time, the uniform convergence of ϕ f n to ϕ f on D implies locally uniform convergence of (ϕ f n ), which by (1) means that fn converges locally uniformly on Ω. We recall here a version of the Arzelà-Ascoli theorem. Theorem 7 (Arzelà-Ascoli). Assume that M is a compact Riemann surface. Let f n : Ω M be a sequence that is equicontinuous on compact subsets of Ω. Then there is a subsequence of {f n } that converges locally uniformly. In complex analysis, a very important notion is that of relative compactness in the topology associated to locally uniform convergence, so important that it has a name. Definition 8. A family F O(Ω, M) is called normal if every sequence in F has a subsequence that converges locally uniformly on Ω. Theorem 9 (Marty). Assume that M is a compact Riemann surface. Then F O(Ω, M) is normal if and only if sup f K < for any compact set K Ω. f F Proof. Firstly, if there is a compact set K Ω such that sup f K =, then there is a f F sequence {f n } F with fn K as n. Then by Theorem 6 a subsequence of {f n } cannot converge, meaning that the family F cannot be normal. As for the other direction, let f F. Also let K Ω be a compact set, let α, β K, and let h be such that β = α + nh for some large integer n. Then from the triangle inequality and the definition of f, we have n 1 ρ(f(α), f(β)) ρ(f(α + jh), f(α + (j + 1)h)) j=0 n 1 ( ) = f (α + jh) h + o( h ) j=0 n 1 = f (α + jh) h + o(1), j=0 and by sending n and taking into account that f is continuous, we infer ρ(f(α), f(β)) f (z) dz f K β α. [αβ] Hence F is equicontinuous on compact subsets, and an application of the Arzelà-Ascoli theorem finishes the proof. We end this section with a general version of Vitali s theorem, which will not used here, but is interesting in its own right. Theorem 10 (Vitali). If {f n } O(Ω, M) is normal, and converges pointwise on a set that has an accumulation point in Ω, then {f n } converges locally uniformly on Ω. Proof. Suppose the opposite. Then there are two subsequences {g m } and {h m } of {f n }, a sequence of points {z m } K from a compact set K Ω, and a number δ > 0, such that ρ(g m (z m ), h m (z m )) δ for all m. The sequences {g m } and {h m } are both normal, so up to subsequences they converge respectively to functions g O(Ω, M) and h O(Ω, M). Without loss of generality, we may assume z m z K, so that ρ(g(z), h(z)) δ.

4 TSOGTGEREL GANTUMUR On the other hand, {f n } converges pointwise on a set that has an accumulation point in Ω, forcing g = h on that set. Then the identity theorem concludes that g = h on Ω. 3. Zalcman s lemma The following remarkable result gives a precise characterization of loss of normality. Lemma 11 (Zalcman). Suppose that F O(Ω, M) is not normal. Then there exist z n Ω, with z n z Ω, ε n > 0, with ε n 0, and f n F such that the sequence of functions ζ f n (z n +ε n ζ) converges locally uniformly on C to a nonconstant function g O(C, M) with g C g (0) = 1. Proof. By Marty s theorem, there exist a sequence of points w n in a compact set K Ω, and a sequence of functions f n F such that f n(w n ). Without loss of generality we may assume that D Ω, w n 0, and f n(w n ) > 0. Let z n D and consider the disk centred at z n of radius 1 z n. We blow up this disk D 1 zn (z n ) so that it becomes a disk of radius R n with R n as n. More precisely, we define h n (ζ) = f n (z n + ε n ζ), ζ D Rn, where ε n = 1 z n R n. Note that z n and R n are assumed to satisfy z n D and R n ; otherwise they are arbitrary. This freedom will be used shortly. Fix some large R > 0, and for all n large enough so that R n > R, let us look at the normality of the sequence of functions h n. We have h n(ζ) = ε n f n(z n + ε n ζ), ζ D Rn, since ρ(h n (ζ), h n (ζ + h)) = ρ(f n (z n + ε n ζ), f n (z n + ε n ζ + ε n h)) = f n(z n + ε n ζ) ε n h + o(ε n h ). Trying to bound h n(ζ) from above for ζ D R, we get ε n f n(z n + ε n ζ) = (1 z n )f n(z n + ε n ζ)/r n (1 z n + ε n ζ + ε n ζ )f n(z n + ε n ζ)/r n (1 z n + ε n ζ )f n(z n + ε n ζ)/r n + (R/R n )ε n f n(z n + ε n ζ), where we have used the definition of ε n in the first line, the triangle inequality in the second, and ζ < R in the third. Since R/R n < 1, we can write this as (1 R/R n )h n(ζ) (1 z n + ε n ζ )f n(z n + ε n ζ)/r n max z 1 (1 z )f n(z)/r n. We have 1 R/R n 1 as n, therefore by choosing R n so that the right hand side of the preceding inequality stays bounded, we can ensure the boundedness of h n on D R uniformly in n. In particular, the choice is convenient, as this implies h n(ζ) R n = max z 1 f n(z)(1 z ), 1 1 R/R n 1 as n.

NORMAL FAMILIES 5 We have R n because f n is unbounded near 0. Moreover, we have h n(0) = ε n fn(z n ) = (1 z n )fn(z n ) = f n(z n )(1 z n ) R n max f n(z)(1 z ) z 1 So if we could choose z n D to be a maximum of z f n(z)(1 z ), that is, f n(z n )(1 z n ) = R n, then we would have h n(0) = 1 for all sufficiently large n. But since z fn(z)(1 z ) is continuous, is positive at z = w n and vanishes for z = 1, it achieves its maximum in D. The outcome of the preceding paragraph is that for any R > 0, there is n R such that the sequence H 0 = {h n : n n R } O(D R, M) is normal. So there is a subsequence H 1 = {h 11, h 12,...} of H 0 such that h 1k g 1 O(D R, M) as k. But a tail of H 1 is normal in O(D 2R, M), so there is a subsequence H 2 = {h 21, h 22,...} O(D 2R, M) of H 1 such that h 2k g 2 O(D 2R, M) as k. Continuing this process, we get a nested sequence H 1 H 2... of sequences H m = {h m1, h m2,...} O(D mr, M), with the property that h mk g m O(D mr, M) as k. From the nestedness, g m and g m+1 agree on D mr, so that the sequence g 1, g 2,... defines a function g O(C, M). Then the diagonal sequence {h kk }, which is obviously a subsequence of the original sequence {h n }, is defined eventually on any disk D R, and converges uniformly there to g. Finally, note that we have g (0) = 1 and g C 1 by construction. 4. Montel s theorem We have encountered a version of Montel s theorem during the proof of the Riemann mapping theorem, which is usually called the thesis version of Montel s theorem. The following strengthened version relates to the thesis version the same way Picard s little theorem relates to Liouville s theorem. Theorem 12 (Montel). If F O(Ω, Ĉ) omits 3 values, then F is normal. Proof. Without loss of generality, we may assume that Ω = D, and that F omits 0, 1, and. In particular, F O(D) and each f F admits a holomorphic n-th root for any n N. Let us collect all 2 n -th roots of elements of F and form the family F n = {g O(D) : g 2n = f pointwise for some f F}. It is obvious that F n omits 0,, and the 2 n -th roots of unity. Anticipating a contradiction, suppose that F is not normal. Then F n is not normal, because convergence of a sequence implies convergence of the sequence composed of 2 n -th power of the elements from the original sequence. Let g n O(C, Ĉ) be the limit function from Zalcman s lemma applied to F n. We have gn gn(0) = 1 and g n omits 0,, and the 2 n -th roots of unity. In particular, {g n } is normal by Marty s theorem, and passing to a subsequence, the limit function g = lim g n O(C, Ĉ) omits 0,, and the 2n -th roots of unity for all n. Moreover, g is not constant since g (0) = 1. It follows from the open mapping theorem that g omits D, hence either g(c) D or g(c) C \ D. Finally, Liouville s theorem applied to either g or 1/g implies that g is constant, reaching a contradiction. Now we are ready prove Picard s big theorem, which we rephrase here for convenience. Theorem 13. If f O(D ) omits 2 values in C, then f extends to f O(D, Ĉ).

6 TSOGTGEREL GANTUMUR Proof. With a positive sequence ε n 0, let g n (z) = f(ε n z) for z D. Then {g n } omits 3 values in Ĉ, so it is normal. Passing to a subsequence, let g = lim g n O(D, Ĉ). So either g is a meromorphic function on D, or g. If g, then there is a circle D r that does not pass through any pole of g, i.e., such that g Dr < M for some M > 0. This means that g n Dr < M for all large n, or in other words, that f(z) < M for z = ε n r for all large n. By the maximum principle, f < M on the annulus of inner and outer radii ε n+1 r and ε n r respectively, and this is true for all large n, hence 0 is a removable singularity of f. For the case g, we have f(z) as z 0, so 0 is a pole of f.