Anda di halaman 1dari 109

Complex Analysis

Mathematics 113. Analysis I: Complex Function Theory

Complex Analysis
Mathematics 113. Analysis I: Complex Function Theory

For the students of Math 113

Harvard University, Spring 2013

FOREWORD

Welcome to Mathematics 113: Complex Function Theory! These are the compiled lecture
notes for the class, taught by Professor Andrew Cotton-Clay at Harvard University during
the spring of 2013. The course covers an introductory undergraduate-level sequence in complex analysis, starting from basics notions and working up to such results as the Riemann
mapping theorem or the prime number theorem. We hope that these lecture notes will be
useful to you in the future, either as memories of the class or as a handy reference.
These notes may not be accurate, and should not replace lecture attendance.

Instructor: Andy Cotton-Clay, acotton@math.harvard.edu


Course assistants: Felix Wong, fwong@college.harvard.edu, Anirudha Balasubramanian, balasubramanian@college.harvard.edu
Course websites: Professor Cotton-Clays website, http://math.harvard.edu/~acotton/
math113.html, Harvard course website http://courses.fas.harvard.edu/0405

Cover image: domain colored plot of the meromorphic function f (z) = (z1)(z+1)
.
(z+i)(zi)2
Source: K. Poelke and K. Polthier. Lifted Domain Coloring. Eurographics/ IEEE-VGTC Symposium on Visualization, (2009) 28:3. Credits: V.Y.
Downloaded from www.flickr.com/photos/syntopia/6811008466/sizes/o/in/photostream/.

TABLE OF CONTENTS

The contents of this compilation are arranged by lecture, and generally have the topics
covered in a linear manner. Additional notes are provided in the appendices.

Foreword

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Introduction

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Riemann Sphere, Complex-Differentiability, and Convergence . . . . .

Power Series and Cauchy-Riemann Equations . . . . . . . . . . . . . . . 11

Defining Your Favorite Functions . . . . . . . . . . . . . . . . . . . . . . . 15

The Closed Curve Theorem and Cauchys Integral Formula . . . . . . 18

Applications of Cauchys Integral Formula, Liouville Theorem, Mean


Value Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

Mean Value Theorem and Maximum Modulus Principle . . . . . . . . 28

Generalized Closed Curve Theorem and Moreras Theorem . . . . . . 31

Moreras Theorem, Singularities, and Laurent Expansions . . . . . . . 34

10

Meromorphic Functions and Residues . . . . . . . . . . . . . . . . . . . . 40

11

Winding Numbers and Cauchys Integral Theorem . . . . . . . . . . . . 42

12

The Argument Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

13

Integrals and Some Geometry . . . . . . . . . . . . . . . . . . . . . . . . . 49

14

Fourier Transform and Schwarz Reflection Principle . . . . . . . . . . . 52

15

M
obius Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

16

Automorphisms of D and H . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

17

Schwarz-Christoffel and Infinite Products . . . . . . . . . . . . . . . . . . 62

18

Riemann Mapping Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 65

19

Riemann Mapping Theorem and Infinite Products . . . . . . . . . . . . 69

20

Analytic Continuation of Gamma and Zeta . . . . . . . . . . . . . . . . . 72

21

Zeta function and Prime Number Theorem . . . . . . . . . . . . . . . . . 78

22

Prime Number Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

23

Elliptic Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

24

Weierstrasss Elliptic Function and an Overview of Elliptic Invariants


and Moduli Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

Some Things to Remember for the Midterm . . . . . . . . . . . . . . . . 94

On the Fourier Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

Afterword . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

CHAPTER

1
INTRODUCTION

Consider R but throw in i = 1 to get C = {a + bi : a, b R}. We will be looking at


functions f : C C which are complex-differentiable. Complex-differentiable, or holomorphic, functions are quite a bit different from real-differentiable functions.
We can think of the real world as rigid, and the complex world as flexible. Complex analysis
has applications in topology and geometry (in particular, consider complex 4-manifolds),
and of course, physics and everywhere else.
Anyway, write C = R2 with the identification (a, b) = a + bi. Addition and multiplication by a R is as for R. For multiplication in general, (a, b)(c, d) := (ac bd, ad + bc).
Claim. C is a field: in particular, its an additively commutative group, and multiplicatively, C {0} is a commutative group.
Proof. Should be immediate.

Claim. Every nonzero complex number z C has a multiplicative inverse.
b
a
Proof. In fact, z 1 := a2 +b
2 ( a2 +b2 )i works.

For z C, z = a + bi, we define the complex conjugate z := a bi = a + b(i).


Claim. z + w = z + w, zw = zw for z, w C.
Proof. Write it out.
We also define the norm |z| :=
norm:

a2 + b2 = zz. We can deduce a few properties of the

Claim. (multiplicative property) |z||w| = |zw| for z, w C.


Proof. It suffices to work with squares. |z|2 |w|2 = zzww = zwzw = |zw|2 .

Algebraically, can we do better than the complex numbers (e.g. can we throw in 1 + i,
etc.)? The Fundamental Theorem of Algebra tells us were done as soon as we throw in i;

that is, we get all the solutions to polynomials.


P
In particular, let p : C C, p(z) =
ak z k for ai C. If the degree is > 0, then
z C : p(z) = 0.
Consider the equation z 2 (1 + i) = 0. Lets work this out explicitly for square roots.
Given a + bi, we want x + iy s.t. (x + iy)2 = a + bi, which gives a system of equations:
(
(
x2 y 2 = a
a2 = (x2 y 2 )2 = x4 2x2 y 2 + y 4
=
= (x2 y 2 )2 +(2xy)2 = a2 +b2
2xy = b
(x2 + y 2 )2 = x4 + 2x2 y 2 + y 4
(

x2 = 21 (a + a2 + b2 ) 0

y 2 = 12 (a + a2 + b2 ) 0
!
p
p

a + a 2 + b2
a + a2 + b2
+ isgn(b)
2
2

p
= x + y = a2 + b2 =
2

= x + iy =

We can take this big general formula and use it to solve our specific cases.
For z C, z = a + bi, we define the real part Re(z) = a =
Im(z) = b = zz
2i .

z+z
2 ,

the imaginary part

The Geometry of the Complex Plane. We have a multiplicative norm |z|. Viewing z, w C as vectors, addition is visualized as a parallelogram, and multiplication is best
seen in polar coordinates. Polar coordinates can be obtained by r = |z| and = Arg(z), the
argument of z. So we get the pretty important identity
z = r cos + ir sin = r(cos + i sin ).
For multiplication, try z1 z2 , where z1 = r1 (cos 1 + i sin 1 ), z2 = r2 (cos 2 + i sin 2 ). Then
z1 z2 = r1 r2 (cos(1 + 2 ) + i sin(1 + 2 )).
So |z1 z2 | = |z1 ||z2 | and Arg(z1 z2 ) = Arg(z1 ) + Arg(z2 ) mod 2. Namely, multiplication by
z = r(cos + i sin ) dilates the complex plane by r and rotates it counterclockwise by .
Another view of the complex number is that z = r(cos + i sin ) acts on C
= R2 by

 
 

cos sin
r cos r sin
a b
r
=
=
.
sin
cos
r sin r cos
b a
Of course,
i sin(n)).

1
z

1
r (cos()

+ i sin()) =

1
r (cos

i sin ) =

1
|z|2 z.

So z n = rn (cos(n) +

Problem. Find all cube roots of 1.


Solution. Want z 3 = 1 = 1(cos 0 + i sin 0). So r3 = 1 = r = 1 and 3 = 0 mod 2 =
4
= 0, 2
3 , 3 . These give the cube roots of unity, and we can plot them on the unit circle
in C.

2
th
Notation. n = cos 2
root of unity. The nth roots of unity
n + i sin n , a primitive n
2k
are nk = cos 2k
n + i sin n , k = 0, ..., n 1.

The equation for nth roots of unity is z n 1 = 0 = (z 1)(z n1 + z n2 + ... + z + 1).


This gives a simple formula for 1 + cos + cos 2 + ... + cos(n 1). For z = cos + i sin ,
P
Pn1
1 + z + ... + z n1 = ( k=0 cos k + i sin k). Now multiply by z 1 = (cos 1) + i(sin )
to get
(cos n 1) + i sin n
(cos n 1)(cos 1) sin n sin
.
=
(cos 1) + i sin
(cos 1)2 + sin2
So

n1
X
k=0

cos k =

(cos n 1)(cos 1) i sin n sin


.
2 2 cos

This isnt the best formula, but were happy with it.

CHAPTER

RIEMANN SPHERE, COMPLEX-DIFFERENTIABILITY,


AND CONVERGENCE

Stereographic projection. Consider a sphere S 2 := {(u, v, w) : u2 + v 2 + w2 = 1} in R3 .


We have a bijection : S 2 N C, where N is the north pole (point at ), given by
:= C {}
stereographic projection. Indeed, we define the Riemann sphere C
= S2.

Diagram 1: Stereographic projection. Source: Wikipedia


Given (u, v, w) S 2 N and (x, y) R2 , the bijection is given by




v
2x
2y
x2 + y 2 1
u
,
, 1 : (x, y) 7
,
,
.
: (u, v, w) 7
1w 1w
x2 + y 2 + 1 x2 + y 2 + 1 x2 + y 2 + 1
Claim. Any circle on S 2 maps to a circle or line in C under and vice-versa.
Proof. I hope you paid attention in lecture.

Now consider the map (u, v, w) 7 (u, v, w), which swaps 0 C with S, the south
with N , the north pole. What does it do to z = x + iy? We see that
pole, and C
uiv
z 7 1+w .

Claim. This is z1 .
Proof. Trivial.
In fact,

1
z

is just rotating the sphere around.

Claim. The inversion

1
z

sends (circles and lines) to (circle and lines).

If we have a function f : C C and we think its nice at , how do we quantify


that? We simply use the inversion z1 . In particular, we look at f ( z1 ). Suppose f () = ;
then f (11 ) sends 0 to 0.
z
For example, consider a polynomial of degree n, f (z) = c0 + ... + cn z n and cn 6= 0. Take
1
zn
1
=
=
,
c0 + ... + cn z n
c0 z n + ... + cn
f ( z1 )
which is well-defined in a neighborhood of 0 C and clearly, sends 0 7 0.
Differentiation. Lets talk about something we all like, differentiation. :-)
Definition. Given f (z) : C open U C, z U , we say f is complex-differentiable
(or holomorphic) at z if
f (z + h) f (z)
f 0 (z) := lim
h0
h
exists for h C.
Proposition. If f, g are holomorphic, then f g and f + g is as well, with (f g)0 = f 0 g + f g 0
and (f + g)0 = f 0 + g 0 .
Proof. Also trivial.

Example. Consider complex polynomials in z, which are holomorphic: f (z) =
Pn1
Then f 0 (z) exists and f 0 (z) =
(k + 1)ck+1 z k .

ck z k .

Example. Let f (z) = z. By showing that the limits in the real and imaginary directions do not agree, show that this is not holomorphic.
Before starting anything else, we would like to review real-differentiable functions. There
are different severities of being real differentiable. Let f : R open U R:
1.
2.
3.
4.
5.

real differentiable: f 0 (x) exists x in domain


C 1 : f 0 (x) exists and is continuous
C k : f, f 0 , ..., f (k) exists and are continuous
C : (smooth) all derivatives exists and are continuous
P
real-analytic: f (x) =
ck xk is a convergent power series

The punchline is that homolorphic functions are always C and analytic! (This is something that should really be appreciated.) Indeed, we oftentimes interchange holomorphic,
C , and analytic in the complex world.

To show this, we will go back to convergence. Consider a sequence of functions {fk (z)}
defined on E C that is containable in a compact set. We say that it is pointwise convergence if limk fk (z) exists z; the pointwise limit is defined as f (z).
Suppose fk (z) is continuous. Its uniformly convergent if , N : n N : |fn (z)
f( z)| , z E.
Lemma. If {fk (z)} are continuous and converge uniformly on E, then pointwise limit
f (z) is continuous.
Proof. Use triangle inequality and a standard epsilon pushing argument.

Power series. Consider f (z) =

P
0

ck z k . We need a condition on |ck | for this to converge.

Definition. Given a sequence {ak }


1 , we define the limit supremum


lim sup ak = lim sup ak = lim ak
k

kn

The polynomial condition is then to look at lim sup |ck |1/k .


Theorem. If lim|ck |1/k = L for L = 0, 0 < L < , or L = , then:
P
1. If L = 0, then f (z) = ck z k converges for all z.
P
2. If 0 < L < , then
ck z k converges for |z| < R = L1 , diverges for |z| > R, and R is
called the radius of convergence. (We do not know for |z| = R.)
P
3. If L = ,
ck z k diverges z 6= 0.
Proof. L = 0:  > 0, N : k > N, |ck |1/k < . Let  = z1 |z|. Then |ck |1/k < 12 |z|,
P
k
ck z k converges. We can do
Re|ck | < 21k 1|z|k . We want |ck |1/k |z| < 21 , so |ck ||z|k 12 so
this with the M -test, by considering partial sums, etc.
0 < L < : lim|ck |1/k = L. Let |z| = R(1 2), then lim|z||ck |1/k = 1 2 for R = L1 .
P
So N : |z||ck |1/k < 1 n > N , and
ck z k converges because |ck |z k < (1 )k .
Likewise, if we suppose |z| = R(1 + 2). Then lim|ck |1/k |z| = 1 + 2. So for infinitely
many k, |ck |1/k |z| > 1+ and |ck z k | > 1. The last condition is proved in the same manner. 
Example.
Example.
lim(k + 1)

P
0

zk =

P
0

1/k

1
1z .

Deduce this from the theorem above.

(k + 1)z k =

1
(1z)2 .

We see this by taking ck = 1 and showing that

= 1.

P
Theorem. Suppose f (z) = 0 ck z k converges for |z| < R. Then:
P
k1
1.
converges for |z| < R.
1 kck z
P
0
2. f (z) exists and equals 1 kck z k1 .
Proof. (1) It suffices to check lim|ck |1/k R1 = lim(k + 1)1/l |ck |1/k
lim(k + 1)1/k |ck |1/k = lim(k + 1)1/k lim|ck |1/k .
(2) R = : subtract and show that limits 0.

1
R.

X
f (z + h) f (z) X
1X

kck z k1 =
ck [(z + h)k z k ]
kck z k1
h
h
9

Note that

1
1
ck [ (z + h)k z k kz k1 ].
h
h

Recall the binomial theorem,


(z + h)k =

 
k l kl
hz ,
l

which we can use to rewrite the expression above as


=

X
0

k  
X
k  
X
X
k kl l1
k kl l1
k1
ck [
z h
kz
]=
z h
l
l
0
2
l=1

X
k  
X
k kl l2
=h
z h .
l
0
2

We also have

X  

 

k kl l2 X k

z h
|z|k1 = (|z| + 1)k ,

l
l
P
P
so
ck (|z| + 1)k converges because R = for
ck z k .
R < : Let |z| = R 2, |h| < ; then |z + h| < R . Then

k  
X
X
k kl l2
f (z + h) f (z) X

kck z k1 = h
ck
z h .
h
l
0
2

Write

 
k
k(k 1)...(k l + 1)
=
.
l
l!


k
For l 2 we can bound kl k 2 l2
. So


X k 

k2

kl l2
z h 2 (|z| + |h|)k


|z|
l
2
and the equation above translates into
=

h X 2
k ck (|z| + |h|)k ,
|z|2

which converges and 0 because of the h in the front.

10

CHAPTER

POWER SERIES AND CAUCHY-RIEMANN EQUATIONS

P
Power series. Last time we were dealing with power series, f (z) = k=0 ck z k . We defined
1
the radius of convergence R = L , where L = lim|ck |1/k and either L = 0, 0 < L < , or
L = .
Inside the radius of convergence, f (z) is a convergent power series with some R. What
are the consequences of this?
Corollary. Inside its radius of convergence, a power series is -differentiable, with expected derivatives as convergent power series.
(k)

Corollary. If R > 0, then ck = f k!(0) .


Proof. Take f (k) (x). The corollary follows.

There are also several uniqueness properties:


P
Lemma. If f (z) =
ck z k is a convergent power series and f (zn ) = 0 for a sequence

{zn }n=1 with zn 0, zn 6= 0, then ck = 0 k and Re(f (z)) 0.


2
Proof. c0 = f (0) = limn f (zn ) = 0. Now form g1 (z) = f (z)
z = c1 + c2 z + c3 z + ...
(exercise: show that this holds for some radius of convergence). Check c1 = g1 (0) =
limn g1 (zn ) = limn f (z)

z = 0, and induct on gi .
P
P
Proposition. If f (z) =
ak z k , g(z) =
bk z k and these agree on some set accumulating at 0, then ak =Pbk k, i.e. f (z) = g(z).
Proof. Consider (ak bk )z k and apply the lemma. Show that lim|ak |1/k L and
lim|bk |1/k L = lim|ak bk |1/k L.

P
Note: To center the power series at w C, consider
ck (z w)k , which shifts the center
of the power series from 0 C to w and maintains the radius of convergence at R.
Complex-differentiability. We refer to complex-differentiability as either holomorphic
or complex-analytic (some texts, e.g. Ahlfors, simply like to use analytic). This is a very
nice property of functions that we will be exploring in the upcoming weeks.

11

Definition. f : C C is holomorphic at z C if
f (z + h) f (z)
h

lim

h0

exists for h C. Equivalently, a number f 0 (z) :


0 = lim

h0

f (z + h) f (z)
f 0 (z).
h

The Cauchy-Riemann equations. By considering real and imaginary parts of holomorphic functions, we get the celebrated Cauchy-Riemann equations.
Proposition. If f (z) : C C is holomorphic at z (alternatively, by R2
= C, we can
f
and
exist
and
regard f (x, y) : R2 R2 ), then f
x
y
i

f
f
=
.
x
y

(3.1)

Equivalently, for the decomposition f (x, y) = u(x, y) + iv(x, y) and


u
v
f
=
+i
x
x
x
f
u
v
=
+i ,
y
y
y
what we mean is that, by comparing real and imaginary parts of the equation


u
v
u
v
i
+i
=
+i
x
x
y
y,
we obtain the equalities
(

u
v
x = y
v
x
= u
y .

(3.2)

Proof. This is a bit tedious to type up, so I hope you paid attention in lecture. The proof
follows from a simple computation of the partials; see any textbook.

The CR conditions (1) or (2), along with the condition that the partials are continuous,
ascertains that f is itself holomorphic.
If the partials are not continuous, consider f (x, y) = xy(x+iy)
x2 +y 2 , z 6= 0, and 0 for z = 0.
Show that this is not differentiable at 0 but

f
x (0, 0)

f
y (0, 0)

= 0.

f
Proposition. If f (x, y) = u(x, y)+iv(x, y) has continuous partial derivatives and i f
x = y
(satisfies CR equations), then f is holomorphic.
Proof. Messy as well, but use the mean value theorem and write out the partials.


12

An equivalent (geometric) formulation of the CR equations. Consider the Jacobian


of f : R2 R2 :
!
u
x
v
x

Jf :=
and the rotation matrices

u
y
v
y

sin
cos

cos
sin


.

So multiplication by i is multiplication by the matrix (aside: see the relation to complex


structure)


0 1
.
1 0
The equivalent form is that Jf I = IJf ; e.g. the Jacobian matrix commutes with rotation
by /2. Check this:
!
!
!
!



u
v
v
u
u
u
u
u
x
y
0 1
0 1
x
y
y
x
x
y
=
=
=
.
v
v
v
v
u
v
v
1 0
1 0
x
u
x
y
y
x
y
x
y
An algebraic interpretation. Consider complex polynomials, p(z) =
at complex-valued polynomials of real variables x, y:
n
X

p(x, y) =

Pn

k=0 ck z

ck,l xk y l

. Look

(3.3)

m=0 k+l=m,
k,l0

(0, 0), and take a new basis as z = x + iy and z = x iy. Write the
with ck,l = kp
x l y
polynomial in z, z :
X X
ck,l z k z l .
(3.4)
p(z, z) =
m=0 k+l=m,
k,l0

We claim that theres a 1-1 correspondence between polynomials given by (3) and (4).

Indeed, x
, y
act on these polynomials:
1

=
z
2

i
x
y

1
=
z
2

+i
x
y

Claim. We have the equalities


k l
(z z ) = lz k z l1 ,
z
k l
(z z ) = kz k1 z l .
z
Proof. We simply check:
k l
1
i
(z z ) = [kz k1 z l + lz k z l1 ] + [kiz k1 z l + (i)lz k z l1 ] = lz k z l1 ,
z
2
2
13

and likewise for the second equality.

Claim. (equivalent form of CR equations) f (z) satisfies CR equations


Proof. A simple check.

z f

= 0.


PP
The conclusion is that p(z, z) = P P
ck,l z k z l is holomorphic iff ck,l = 0 for l > 0 (e.g.
0
no z s).P Alternatively, p(x, y) =
ck,l xk y l is holomorphic iff it can be written as
k
p(z) = ck z .
Lets give some basic hints that holomorphic functions behave a bit like power series:
Lemma. Suppose f (x, y) = u(x, y) + iv(x, y) is holomorphic on a disk of radius R and
u(x, y) is constant. Then f is constant.
Proof. ux = uy = 0, so by CR equations, ux = vy , uy = vx and vx = vy = 0. By the
mean value theorem, this shows that v is constant along horizontal and vertical lines in the
plane. So v is constant throughout the disk.

Lemma. If f = u + iv is holomorphic on a disk in C and ||f ||2 is constant, then f is
constant.
Proof. We have u2 + v 2 = c, so taking partials by x and y gives 2ux u + 2vx v = 0 and
2uy u + 2vy v = 0. Applying the CR equations, we get 2ux v 2vx u = 0. Adding equations
gives 2ux (u2 + v 2 ) = 0, and similarly 2vx (u2 + v 2 ) = 0. So either u2 + v 2 = 0 (f = 0), or all
partials = 0.
In the second case, given any open ball, f is constant in that open ball since its partials
= 0 in the ball. But the disk is connected, so f is constant everywhere.

14

CHAPTER

4
DEFINING YOUR FAVORITE FUNCTIONS

Today is about extending your favorite (real analytic) functions from R to C. When determining their complex analogues, we could (1) ask for the same fundamental properties; (2)
use their power series; (2) extend such that the result is holomorphic (actually, this is the
same as 2). Happily, these all agree!
1. ex
d x
e = ex , ... or (2) ex =
We can consider (1) e0 = 1, dx

P xn
n!

(1) Lets look for f (z) with f (x) = e and f (z1 + z2 ) = f (z1 )f (z2 ) and holomorphic.
Apparently, f (x + iy) = ex f (iy). Let f (iy) = A(y) + iB(y), so f (x + iy) = ex A(y) +
ex iB(y). Then use the Cauchy-Riemann equations to show that A(y) = cos y and
B(y) = sin y by the Fundamental Theorem of ODEs (namely, any linear ordinary
differential equation has an ez solution). So define ez = ex (cos y + i sin y), for z =
x + iy.
(2) Let
f (z) =

X
zn
.
n!
n=0

1 1/n
The radius of convergence is , since limn ( n!
)
= 0 (e.g. its an entire function).
Also, f (z) is holomorphic on all of C. We can easily verify the properties that f 0 (z) =
f (z), f (0) = 1, f (z)f (w) = f (z + w), etc.

2. sin z, cos z
2

z
z
z
z
For f (z) = ez , look at f (iz) = 1+(iz)+ (iz)
2! +... = (1 2! + 4! +...)+i(z 3! + 5! +...).
Then we set f (iy) =: cos y + i sin y.

We can verify the usual properties that z


cos z = sin z, z
sin z = cos z, cos 0 = 1,
iz
sin 0 = 0, cos(z) = cos z, sin(z) = sin z, e = cos z + i sin z, etc.

Likewise, we can express


cos z =

eiz + eiz
,
2
15

sin z =

eiz eiz
.
2i

We can likewise check the addition formulas for cos(z1 + z2 ) and sin(z1 + z2 ).
3. log z
w = log z z = ew . For w = a + bi, ew = ea (cos b + i sin b). Apparently, log z =
log r+i+i(2n), so the complex log is multi-valued. Write z = rei = r(cos +i sin ).
We define the principal branch of log z as follows. Consider C R0 . log z has imaginary part in (, ). Write log z = log |z| + iArg(z) for Arg(z) (, ); this is our
definition of log z for the principal branch, which has range < Im(z) < .
Claim. log z is holomorphic given a branch cut, and

z (log z)

= 1/z.

Proof. Consider
lim

h0

log(z + h) log z
.
h

Set w1 = log(z + h), w = log z. Then, noting that log is continuous given a branch
cut, we see that
lim

h0

w1 w
1
1
1
log(z + h) log z
= lim w1
=
ew1 ew = w = .
h0 e
h
ew
e
z
limw1 w w1 w


Intuitively, log z looks like a helix above C {0}; we call this the Riemann surface for
log z.
4. z , R
Define z = e log z = e(log z+2ni) , which is potentially multi-valued. e log z is the
principal branch; now examine e2ni . If
/ Q, e2ni takes infinitely many values.
k
2nki/l
If = l in reduced form, e
takes l values (e.g. eni = 1 for = 21 ).

rei = rei/2 ,
For example, when considering z, we still need a branch cut.
(, ).
Integration. Given f (z) holomorphic, can we find another function F (z) holomorphic with
0
RF (z) = f (z)? The answer is, generally, yes. We begin with integration along a contour,
f (z)dz for C a contour.
C
Definition. The image of (t) : [a, b] C is a contour curve if:
1. is continuously differentiable except at finitely many points
2. When differentiable, 0 (t) 6= 0 except at finitely many points
R
Rb
For C a contour, C f (z)dz = a f ((t)) 0 (t)dt. Call 1 (t) 2 (t) for 1 : [a, b] C and
2 : [c, d] C if (t) : [c, d] [a, b] s.t. 2 (t) = 1 ((t)).
Claim. This is well defined:
Z b

f (1 (t))10 (t)dt =

16

f (2 (t))20 (t)dt.

Lemma. (complex chain rule) Let g(t) = f ((t)), f (z) holomorphic.


0 (t)f 0 ((t)).
Proof. As usual, with limits or whatnot.

Then g 0 (t) =


Example. Compute
Z

z k dz

S1

for k Z, and S 1 the unit circle in C oriented counterclockwise.


Explanation. Let (t) = eit = cos t + i sin t. By chain rule, 0 (t) = ieit . Then
Z

z dz =
S1

ikt

it

ie dt = i

ei(k+1)t dt.

If k = 1, this is
Z
S1

1
dz = i
z

1dt = 2i.
0

If k 6= 1, this is
2

i(k+1)t

e
0

2
ei(k+1)t
= 0.
dt =
i(k + 1) 0

So z k for k Z, k 6= 1 has an antiderivative in C {0}:


log z, but its multivalued.

z k+1
k+1 .

1/z does not; wed want

Proposition. If f (z) holomorphic with F (z) holomorphic and F 0 (z) = f (z), then
Z
f (z)dz = F (end) F (start),
C

for C a path from start to end.


Proof. Use the complex chain rule.

Next time, well consider functions that are defined on all of C (which we call entire functions), well show F (z) : F 0 (z) = f (z). Along the way, well prove the closed curve theorem.
For motivation, we recall Greens theorem:

Z b
Z Z 
Q P
0

dxdy
(P, Q) r (t)dt =
x
y
a
D
for r(t) : [a, b] R2 a closed curve around a region D. This requires that Qx , Py be
continuous (or something like that). The theorem well prove does not need this requirement.

17

CHAPTER

THE CLOSED CURVE THEOREM AND CAUCHYS


INTEGRAL FORMULA

Today well be talking about the closed curve theorem and Cauchys integral theorem.
Recall from last lecture that, for a parameterization of the contour C, we had
Z
Z
f (z)dz = f ((t)) 0 (t)dt.
C

We also know that the path integral is determined by the antiderivatives value at the curve
endpoints:
Proposition. If F (z) is holomorphic, F 0 (z) = f (z), then
Z
f (z)dz = F (final) F (initial).
C

The closed curve theorem. We will first provide motivation for the closed curve theorem.
Given f and , consider splitting into real and imaginary parts:
f (z) = f (x, y) = u(x, y) + iv(x, y)

f ((t)) 0 (t)dt =
Z
=

(t) = a(t) + ib(t)


Z

+ i(ub + v a)]dt
(u + iv)(a + ib)dt
= [(ua v b)

(u, v) (a,
b)dt
+i

(v, u) (a,
b)df

(5.1)

If C is the closed, counterclockwise boundary of D, then by Greens theorem (which only


applies when all partials are continuous) we have that the above is equal to


Z Z 
Z Z 
v
u
u v

dxdy + i

dxdy = 0,
x y
x y
D
D

18

where the last equality follows from the Cauchy-Riemann equations. We can also note that,
in the language of vector fields, (1) is equivalent to
Z Z
Z Z
curl(f )dA +
div(f )dA,
D

where f is thought of as a vector field on R2 induced by f . The equivalent formulation of


the CR equations would then be curl(f ) = div(f ) = 0 as a vector field on R2 .
The conclusion above gives us a glimpse into the closed curve theorem, which, however,
does not require the first partials to be continuous. Keep in mind that our objective is the
result that f being holomorphic on the interior of a disk implies that f has a convergent
power series expansion on that disk: we will do this using the closed curve theorem and
friends.
Claim. The closed curve theorem (stated below) is true for linear functions
f (z) = a + bz.
R
Proof. Let F (z) = az + 12 bz 2 , so F 0 (z) = f (z). So we have C f (z)dz = F (final)
F (initial) = 0.

Theorem. (closed curve theorem) Suppose f (z) is holomorphic in an open disk D. Let C
be any closed contour curve in D. Then
Z
f (z)dz = 0.
C

Remark. Our strategy is in three steps: (1) show this for any polygonal curve; (2) show
that f (z) has an antiderivative on D; (3) conclude the result.
Proof for triangle. Let T be a triangle contour. The idea is that f being holomorphic
implies that f is well-approximated by linear functions
on small scales; see the claim above.
R
First, we subdivide the triangle. Suppose | T f (z)dz| = C. Then Ti with boundary i
R
such that | i f (z)dz| c/4. Call this T (1) with boundary (1) . Keep subdividing to get
R
(dropping the parentheses) T 1 , T 2 , T 3 , ... and 1 , 2 , 3 , .... such that | k f (z)dz| 4ck .
k
Note that
k=1 T = {point}; call the point z0 . f (z) is holomorphic at z0 , so we have


f (z) f (z0 )

0

lim
f (z0 ) = 0.
zz0
z z0
Thus
f (z) = f (z0 ) + f 0 (z0 )(z z0 ) + (z)(z z0 )
with (z) 0 as z z0 . Now consider
Z
Z
Z
f (z)dz =
[f (z0 ) + f 0 (z0 )(z z0 ) + (z)(z z0 )]dz =
k

(z)(z z0 )dz.

Given  > 0, N s.t. for k N , (z) <  for z T k . Let the perimeter of T be L and the
perimeter of T k be 2Lk . The max distance between two points in T k would then be 2Lk .
Thus

Z
2


c

L  L = L ,

(z)(z

z
)dz
0


k
k
k
k
4
2 2
4
k

19

where the first term after the equality is length of k , theR second is an upper bound on (z),
and the third is an upper bound on |zz0 |. We had c = | f (z)dz|, so c L2 = c = 0. 
Corollary. (extension to polygons) If P isRa polygon and f (z) is holomorphic on an open
neighborhood of P with boundary , then P f (z)dz = 0.
Proof. Divide P into triangles.

Theorem. (antiderivative theorem) If f (z) is holomorphic on an open disk D (or an
open polygonally simply connected region C), then F (z) holomorphic on D (or ) s.t.
F 0 (z) = f (z).
Definition. An open region C is polygonally connected (in topology, connected) if
z, w , piecewise linear curve in connecting z, w. A region is simply polygonally
connected (simply connected) if any closed polygonal curve is the boundary of a union of
polygons in .
Rz
Proof of antiderivative theorem. Let p D (or C). Let F (z) = p f (z)dz (the
line integral along any polygonal path from p to z, which is well-defined by the result on
triangles above and the p.s.c. property of ). We check that
Z z
1
F (z) F (z0 )
=
f (z)dz.
z z0
z z0 z0
Note

because




Z z
F (z) F (z0 )
1




f (z0 ) =
(f (z) f (z0 ))dz

z z0
z z0 z0
Z

1dz = f (z0 )(z z0 ).

f (z0 )dz = f (z0 )


z0

Then we have

z0



Z z
1

1

(f (z) f (z0 ))dz
|z z0 |,
z z0
|z

z0 |
z0

where the middle term after the inequality is the length of our curve and  is upper bound
for |f (z) f (z0 )| (by continuity). So the LHS 0 as  0.

Corollary. (closed curve theorem
for regions) If f (z) is holomorphic on an open disk
R
D (or p.c.r. region ), then C f (z)dz = 0 for C being any
R closed contour in D (or ).
Proof. Take F (z), the antiderivative of f (z), so that C f (z)dz = F (final) F (initial) =
0.

Cauchys integral theorem. Given that f (z) is holomorphic, lets look at
(
f (z)f (w)
z 6= w
zw
g(z) =
0
f (w)
z=w
for w C.
Claim. g(z) satisfies the closed curve theorem.
20

Proof. g(z) is continuous because f 0 (w) exists. The goal is to prove the above result
for triangles for g(z); then it will follow G(z) : G0 (z) = g(z), along with the closed curve
theorem for g(Z).
Consider w. If w is outside T , then g(z) holomorphic in an open neighborhood of T and
the old closed curve theorem applies.
R
If w = T , then we cut off small pieces {Ti } so i f (z)dz = 0 (because they do not
contain w). The length of i can be arbitrarily small at M , the upper bound on g(z) on T ,
so it goes to 0 in the limit. If w is in the interior, we also use the same trick.

Cauchy integral formula. This is one of the central results of complex analysis. Let
C be a circle surrounding w C, and f (z) holomorphic on the closed boundary C of the
open disk D. Then we have the formula
1
f (w) =
2i

f (z)
dz
zw

Proof. Suppose C is a circle centered at w. The circle is parameterized by (t) = w+Reit


for 0 t 2. We know from the closed curve theorem that
Z
f (z) f (w)
0=
dz,
zw
C
because the integrand is g(z) defined on an interior satisfying the closed curve theorem.
Now split this up:
Z
Z
Z
Z 2
f (z)
1
1
iReit dt
dz = f (w)
dz,
dz =
= 2i
Reit
C zw
C zw
C zw
0
Z
1
f (z)
dz.
= f (w) =
2i
zw

Theorem. Suppose f (z) is holomorphic on thePopen disk Dk (0) of radius R around 0 C.

k
1/k
Then there
a convergent power series
R1 such that
k=0 ck z with lim|ck |
P exists
k
f (z) = k=0 ck z .
Proof. From Cauchys integral formula, we have
Z
1
f (w)
f (z) =
dw.
2i
wz
< R, and write
Let |z| < R
1
1 1
=
wz
w1
| wz | =

|z|

z
w

1
w

1
wz

< 1 so this converges uniformly to


1
f (z) =
2i

Z
f (w)(


1+


z
z2
+ 2 + ...
w w

Then
for w C.

z
z2
1
+ 2 + 3 + ...)dz
w w
w

21

1
2i

Z
C

f (w)
dw +
w

i.e.

1
2i

Z
C




Z
1
f (w)
f (w)
dw
z
+
dw
z 2 + ...,
w2
2i C w3

1
Ck =
2i

f (w)
dw.
wk + 1

We check this as follows:


|Ck |

1
1

= M/Rk ,
2 RM
k+1
2
R

is the length of R and


where the terms of the right of the inequality come as follows: 2 R
M 1/k
1
1/k
M is an upper bound on f in DR (0). So lim|Ck |
lim R = R , and
f (k) (0)
1
=
k!
2i
for C containing 0.

Z
C

f (w)
dw
wk+1


Thus we have the result that f holomorphic = f analytic = f infinitely differentiable.

22

CHAPTER

APPLICATIONS OF CAUCHYS INTEGRAL FORMULA,


LIOUVILLE THEOREM, MEAN VALUE THEOREM

Applications of Cauchys integral formula. Recall that we have the Cauchy integral
formula:
Z
1
f (w)
f (z) =
dw,
2i C w z
P
for C a curve containing z. For f (w) = ck wk , note that we can write
Z
Z X
Z

1
f (w)
1
c0
1
ck wk1 dw =
dw =
dw = c0 .
f (0) =
2i C w
2i C
2i C w
k=1

Lets use this formula to compute some stuff. The general method is to decompose a closed
contour (over which the integral is zero) into a sum of directed edges, then evaluating the
integral over the other edges to give us results.
Claim.

Z
0

sin x
dx = .
x
2

Proof. Let CR denote the contour of an upper half-circle of radius R centered at the
origin. For r < R, we can use Cauchys integral theorem to write (noting that the integrand
ix
is equivalent to the imaginary part of ex :
Z
0=
r

eix
dx +
x
i

Now parameterize Cr with z = re

Z
CR

eiz
dz +
z

eix

Z
Cr

eiz
dz.
z

for [0, ] and likewise with z = Rei to see

Z iRei
Z r ix
Z irei
eix
e
e
e
i
0=
dx +
iRe d +
dx
irei d.
i
i
x
Re
x
re
r
0
R
0
Z R ix
Z
Z r ix
Z
i
i
e
e
=
dx +
ieiRe d +
dx
ieire d.
x
r
0
R x
0
Z

23

But note that the second integral 0 as R and the fourth integral i as r 0
(show this yourself), so were done as sinx x (the imaginary part of the remaining two integrals) is symmetric over the y-axis and its integral from 0 to
R would then be equal to the
imaginary part of our answer divided by 2. Namely, we have 0 sinx x dx = 2 as desired. 
Claim (added by Felix, for more practice)

sin(x2 )dx =

x2

2

ex dx =

e(y

dx

2
.
4

cos(x2 )dx =

Proof. We first show that


Z

as follows:

+z 2 )

rer drd = ,

dydz =
0

with y = r cos and z = r sin .


2
Using what we showed above, we integrate f (z) = ez on the contour defined by
C = {Reit : 0 < t < /4}. Observe that we can bound the integral by
/4

Z
Z


0 f (z)dz

eR

(cos2 (t)sin2 (t))

/4

eR

Rdt =

cos(2t)

Rdt,

R /4
2
but also (from brute force bounding) we know that R 0 eR cos(2t) dtR 0 as R
(). This is good, because we can invoke Cauchys theorem to see that C 0 f (z) for C 0 =
{reit : 0 < r < R, 0 < t < /4} = 1 + 2 + 3 (where 1 is the path from 0 to R, 2 is the
i/4
eighths-circle, and 3 is the line from
back to 0), a closed boundary, is zero. Letting
R Re
R , we saw from above that 2 f (z) = 0, so what we have left is (Ill drop the f (z)s
to make writing easier)
Z
Z
Z
Z
Z
Z

2
+
=
ez +
= /2 +
= 0 /2 =
.
1

So to finish, we evaluate
Z

Z
=

R
3

(ei/4 t)2 i/4

dt = e

i/4

it2

=e

i/4

Thus we see that


Z
Z

/2 = ei/4
cos(t2 )i sin(t2 )dt
0

cos(t2 ) i sin(t2 )dt.

cos(t2 )i sin(t2 )dt =

/2

=
2/2 + 2/2i

R
R
Equating real and imaginary parts, we see that 0 sin(x2 )dx = 0 cos(x2 )dx =
desired.
() Brute force bounding: For more exercise, well explicitly show that
Z
R

/4

eR

cos(2t)

24

dt 0

2
2

i.
4
4

2
4 ,

as

as R here. We have
Z /4
Z
2
R
eR cos(2t) dt = R
0

/4

eR

sin(2t)

/6

eR

dt = R

sin(2t)

/4

eR

dt + R

sin(2t)

dt.

/6

d
Then, since we can bound sin x (1/2)x for 0 x 1/3 ( dx
sin x 1/2 for x 1/3) and
sin x 1/2 for /6 /4, we see that

/6

eR

sin(2t)

/4

eR

dt + R

sin(2t)

ReR t dt +

/6

/6

dt

/4

ReR

/2

dt

/6

R2
2

1
ReR /2 + (1 e 6 ) 0,
12
R
as R . Integrals are fun stuff.

Now lets turn to other theorems we can prove using the Cauchy integral formula.
Theorem. (Liouville) A bounded entire function f (z) is constant.
Proof. Suppose |f (z)| M for all z. Consider f (a), f (b) for |a|, |b| < R. We would like
to show that f (a) = f (b). Let C0 (R) be the circle of radois R about 0. Then
Z

Z
1
f (z)
f (z)
|f (a) f (b)| =
dz
dz


2 C0 (R) z a
C0 (R) z b
and

f (z)
(b a)f (z)
f (z)

=
.
za zb
(z a)(z b)

For |b a| 2R, we can also get


Z

Z

Z

1
f (z)
f (z)
1
f (z)(b a)
1 2RM |b a|

dz
dz =
dz


1 2



2 C0 (R) z a
z

b
2
(z

a)(z

b)
2
C0 (R)
C0 (R)
4R
=

4M |b a|
0
R

as R .
Lets look at polynomials P (z) =


Pn

k=0 ck z

, cn 6= 0. What can we say about them?

Claim. limz P (z) = in the sense that M, r : |z| > R, |P (z)| > M .
Proof. Should be obvious.

Theorem.
(Fundamental Theorem of Algebra) Given a nonconstant polynomial P (z) =
Pn
k
c
z
,
c
n 6= 0, n 1, w C : p(w) = 0.
k=0 k
1
Proof. Suppose not. Then P (z)
=: f (z) is a bounded entire function. Choose R such
that
n1
X |ck |
1
>
R
|cn |
k=0

25

and R > 1 so that for |z| > R we have


1

P (z)

1
1
2 R|cn |

Note that, in obtaining this inequality, we noted that the highest power of any polynomial
dominates in the limit |z| . So f (z) is bounded by C for |z| > R, and its similarly
bounded on any compact region D0 (R); thus f (z) is constant by Liouvilles theorem.

Theorem. (Generalization of Liouvilles Theorem) If |f (z)| A + B|z|N and f (z) is
entire, then f (z) is a polynomial of degree N .
Proof. By induction. The base case is Liouvilles theorem. Now define
(
f (z)f ()
z 6=
z
g(z) =
0
f ()
z=
for any . Notice that g(z) is holomorphic, since we saw G(z) that is holomorphic with
G0 (z) = g(z), and we showed that holomorphic implies C so g(z) is also holomorphic.
Notice further that |g(z)| C + D|z|N 1 .
Addendum by Felix: We can see the last statement more clearly as follows. If to the
contrary |g(z)| > D|z|n1 , then taking limits as |z| we see that the growth rate
()
or f 0 () (growth order N 1).
(growth order N ) would not match with f (z)f
za
(Also, as said in lecture, we can see the last statement as follows: look inside a large
ball, then outside the large ball. Inside, its bounded by some constant; outside, because of
the induction hypothesis, it has the form D|z|N 1 .)
So g(z) is a polynomial of degree N 1 by induction and f (z) = g(z)(z ) is a
polynomial of degree N . This completes the inductive step.

Uniqueness theorem. It turns out that holomorphic functions are determined up to
their values in a certain region; this is the content of the uniqueness theorem.
Definition. A region in C is a polygonally connected open set.
Proposition. (uniqueness theorem) If f and g are holomorphic on a region D C with
f (z) = g(z) on a collection of points accumulating somewhere, then f (z) = g(z) on D.
Proof. Let A = {z D : infinitely many points w in any disk containing z s.t.
f (w) = g(w)}. Let B = D\A. We claim that both A and B are open.
To show that B is open, note that if z B, then a disk containing z with only finitely
many such ws. Let < limsuch w,w6=z |z|. Then every point in the open disk of radius is
in B.
To show that A is open, suppose z A. In D, a disk of radius r containing z. On
that disk, both f and g being holomorphic implies that both f and g have convergent power
series expansions centered at z of radius of convergence r. So by uniqueness for power
series, f = g on this disk, i.e. every point in this disk in in A.
So we have that A and B are both open, A B = D, and A B = . By elementary
topology, one must be the empty set.

It turns out that polynomials are the only entire functions that go to infinity with z in
with
the limit. Geometrically, this means that if we consider the Riemannian sphere C
26

C
entire, then f : 7 a for a finite implies that f (z) is bounded, i.e. a constant.
f :C
If f : 7 , then f (z) is a polynomial.
Proposition. If f (z) is entire and f (z) as z , then f (z) is a polynomial.
Proof. The zeroes of f all lie in a disk of some radius, call it R. There are only finitely
many zeroes, else they accumulate somewhere and then f = 0. Now consider
g(z) =

f (z)
.
(z z1 )m1 (z z2 )m2 ...(z zk )mk

P
Note that g(z) has no zeroes and is entire. lim f (z) = = |g(z)| |z|1m , where m = mi .
1
So g(z)
is entire and bounded, and by Liouvilles theorem it must be a constant. Thus f (z)
is a polynomial.

We take the time to state one last important result, the mean value theorem, without
proof:
Theorem. (mean value) If f (z) is holomorphic on D (R), then r : 0 < r < R, we
have
Z 2
1
f ( + rei )d.
f () =
2 0
Corollary. This is true as well for u = Re(f ), v = Im(f ) in place of f .
Proof. Use the integral formula.

27

CHAPTER

MEAN VALUE THEOREM AND MAXIMUM MODULUS


PRINCIPLE

Today, well be talking about the maximum modulus principle, the mean value theorem, and
some topological concerns. Recall from last time that we had the mean value theorem:
Theorem. (mean value) For f holomorphic in DR (z), and r < R,
Z 2
1
f (z) =
f (z + rei )d.
2 0
We also have the maximum modulus principle, which is one of the most useful theorems in
complex analysis:
Theorem. (maximum modulus) Let f be holomorphic in DR (z). If |f | takes a local
maximum value at z in the interior, then f is constant.
Proof. Use the mean value theorem. Alternatively, an equivalent formulation (Ahlfors)
is as follows: if f is holomorphic and nonconstant in DR (z), then its absolute value |f | has
no maximum in DR (z).
To show this, suppose that w = f (z) is any value in DR (z), so that the neighborhood
D (w) is contained in the image of DR (z). In this neighborhood, there are points of modulus
> |w| so |f (z)| is not the maximum of |f |.

Claim. If |f | is constant and holomorphic, then f is constant on a disk (or region).
Proof. Use the CR equations.

Corollary. (minimum modulus) If f is holomorphic and achieves a minimum of |f (z)|


in a disk, then f (z) = 0 there.
Proof. If f (z) 6= 0 at a minumum then 1/f is holomorphic in a neighborhood of z. Then
use the maximum modulus theorem.
Likewise, we can also use power series to prove this.

We went over another proof of the maximum modulus theorem in class using power series; it may be nice to go over this proof as well (which is, again, rather long); see page 87
28

in the textbook.
Corollary. If f is holomorphic in DR (z), then on Dr (z), r < R, the maximum value
of |f (z)| on Dr (z) is achieved on the boundary.
Proof. Consult the textbook.

The following corollary is more general than the above:
Corollary. If f is continuous on a closed bounded set E and holomorphic in its interior
E E, then the maximum of |f | is attained on E.
Proof. (Ahlfors) Since E is compact, |f (z)| has a maximum on E. Suppose that it is
attained at z0 ; if z0 E, were done. Else if z0 is an interior point, then |f (z0 )| is also
the maximum of |f (z)| in a disk |z z0 | < contained in E. But this is not possible unless
f (z) is constant in the component of the interior of E which contains z0 . It follows by
continuity that |f (z)| is equal to its maximum on the whole boundary of that component.
This boundary is not empty and is contained in the boundary of E, so the maximum is
always attained at a boundary point.

Proposition (anti-calculus) If |f (z)| achieves its max value on Dr (z) at w, then f 0 (w) 6=
0.
Proof. Sketched in class; see page 88 of the textbook.

We will use the maximum modulus principle to prove the following theorem:
Theorem. (open mapping) The map of an open set under a non-constant holomorphic
map f : U C, U C, is open. Note that f continuous at z means that  > 0, > 0 :
f (B (z)) B (f (z)).
Sketch of another proof. If f 0 (z) 6= 0, then


ux uy
Jf =
v x vy
is nonsingular. Then det Jf = ux vy uy vx = u2x + vx2 = |f 0 (z)|2 6= 0. The inverse function
theorem gives the result.
If f 0 (z) = 0, then consider f (B (0)) = Bf () (0). Write f (z) = z k (ck + ck+1 z + ...). When
f has a zero of order k at 0, i.e. f (0) = f 0 (0) = ... = f (k1) (0) = 0, it turns out that
f (z) = [g(z)]k , g(z) = Az + ... i.e. g 0 (0) 6= 0 (but we cannot prove this as of yet).

Our proof. See page 93 of the textbook.
Topology toward general closed curve theorem. We would like to take the time
to revisit some topological concerns about the complex plane. In particular, recall that we
defined the notions of polygonally connected and polygonally simply connected :
Definition. A set U C is polygonally connected if p, q U , there exists any piecewise linear path from p to q.
A polygonally connected set is connected, but a connected set may not be polygonally
connected (what is an example?). For open sets, they are equivalent:

29

Proposition. If U C is open, then U is connected iff its polygonally connected.


Definition. A set U C is polygonally simply connected (or, in the topological sense,
simply connected ) if for any closed polygonal curve is the boundary of a union of polygons
in U .
Definition. A region in C is an open connected (or polygonally connected) subset
of C.
If E(U ) denotes the space of directed edges (for purpose of integrating over) in the region U , then by composing elements
in E(U ) to form a closed contour we know from
R
Cauchys integral theorem that f (z)dz = 0 for f holomorphic.
The notion of connectedness goes into topology, involving such topics as homotopy and
homology. For those of you that are familiar with topology, a simply connected set has
trivial fundamantal group (e.g. every loop is homotopic to a constant). The Cauchy integral
formula also has a more general, homology version involving winding numbers.

30

CHAPTER

GENERALIZED CLOSED CURVE THEOREM AND


MORERAS THEOREM

Today well talk about the general closed curve theorem and Moreras theorem.
Theorem. (general closed curve theorem) If D is a region (open, polygonally connected
D is connected, then the closed curve theorem holds, i.e.
subset) in C and C
Z
f (z)dz = 0
C

for f holomorphic on D and C a contour curve in D.



with A B S and
Definition. A set S in C
= S 2 is connected if @A, B open in C
A B 6= .
The method for proving the general closed curve theorem is:
D is connected implies D p.s.c.
1. C
2. D p.s.c. implies closed curve theorem holds (see textbook for pictures).
To show this, well need to go back to last time: first let U be a region in C. Then
define
Pn
T (U ) := { i=1 ai Ti }, for Ti a triangle in U , ai Z, and any n.
Pn
E(U ) := { i=1 ai Ei }, for Ei an edge in U (recall that an edge means a pair of points
{P, Q} with P 6= Q).
Pn
P (U ) := { i=1 ai Pi }, for Pi a point in U .
T := E1 + E2 + E3 for T a triangle.
E = Q P ; : E(U ) P (U ) and : T (U ) E(U ).
e E(U ) is closed if e = 0, e.g. e = E1 + ... + En is closed iff Ei form loops.
Definition. e1 is homologous to e2 , denoted e1 e2 , if t : t = e1 e2 . This is an
equivalence relation.
31

Definition. U is p.s.c. if e is closed, e.g. e is null-homologous, which means e 0.


Proposition. If U is convex in C, then U is p.s.c.
Proof. Prove this yourself, by dividing the set into triangles.

Proposition. If U is p.s.c., then the closed curve theorem holds.


Proof. Note that the antiderivative theorem holds. Let e be a (not necessarily closed)
piecewise linear path from P to Q, e = Q P . Then
Z
X Z
f (z)dz =
ai
f (z)dz.
e

Define F (z) =

Rz
z0

Ei

f (z)dz. This makes sense because


Z
Z
f (z)dz =
f (z)dz.
e1

if e1 e2 , since then

e2

Z
f (z)dz =

e1 e2

Z
bj

f (z)dz

where is the boundary of the triangle and the integral is zero. So the antiderivative
theorem implies the closed curve theorem as before:
Z
f (z) = f (final) f (initial) = 0,
C

as desired.

Well introduce Moreras theorem since we have some time. First some topology:
Proposition. If U is an open subset in C, U is polygonally connected iff U is connected.
If C
D is connected, then D is p.s.c.
Claim. Let D be a region in C C.
Theorem. (Morera) If f (z) is continuous on a region D and
then f (z) is holomorphic.
Proof. See textbook.

f (z)dz = 0 for a triangle,




Moreras theorem is quite useful. Some of the uses are as follows:


1. Showing limn fn = f is holomorphic.
2. Showing that some functions defined by sums or integrals are holomorphic, e.g. the
Riemann zeta function and the gamma function:
(z) :=
Z
(z) :=
0

32

X
1
z
n
n=1

et tz1 dt

3. Showing functions that are continuous and holomorphic on all but some set (e.g. some
point) are holomorphic.
Proposition. Suppose {fn } is holomorphic in an open set D, and fn f uniformly on
compact subsets of D. Then f = limn fn is holomorphic.
Proof. It suffices to work in B () D. B/2 () is compact, so by assumption fn f
is uniform here. Note that fn f uniformly implies f is continuous. Also
Z
Z
Z
f (z)dz =
lim fn (z)dz = lim
fn (z)dz = 0
n

(Uniform convergence implies that we can swap and lim, as we all know from elementary
analysis.) By Moreras theorem, f is holomorphic.

The Riemann zeta function. The zeta function (z) is defined as
(z) =

X
1
,
nz
n=1

for Re(z) > 1. We claim that (z) is convergent for Re(z) > 1, and indeed, is uniformly
convergent for Re(z) > R > 1. (Hence by Moreras theorem (z) is holomorphic.)
Proof of claim. First note that n1z = e(log n)z . We first show uniform convergence. Let
fN (z) :=
We bound

n=N +1

N
N
X
X
1
=
e(log n)z .
z
n
n=1
n=1

|e(log n)z | for the whole region Re(z) R > 1:


|e(log n)z | = |e log n eRe(z) | R|e log n |

P
So were left with n=N +1
the tail end is small.

1
,
nR

R > 1. The integral test enures convergence, telling us that




33

CHAPTER

MORERAS THEOREM, SINGULARITIES, AND LAURENT


EXPANSIONS

Lets finish the topology stuff, talk about Moreras theorem, and then move on to singularies
and Laurent expansions.
= S2?
First, what does it mean to fix an open set in C
is
Definition. A neighborhood of is given by {} {z : |z| > R}. Then S C
open if:
(1) S C is open in C
(2) If S, then R : {z : |z| > R} S.
Lemma. If P is a closed polygonal path which doesnt cross itself (i.e. P is a piecewise linear
map
from
s.t. AB = CP

an interval that is injective except at the endpoints), then A, B open in C


and A B = .
Proof. Note R : P {|z| > R} = for z C P . Take L, a line segment from z to a
point in S = {|z| > R}, and count the number of times P crosses L to R and the number
of times P crosses R to L. This makes sense because P L is a finite union of edges of P
and points. Let z A if the count is odd; else z B.
Exercise. Show that z A or B is well-defined independent of the choice of L. Hint:
use wedges. Conclude that nearby points are in the same set.

Integration over other regions. We use C R0 to denote the complex plane with
(C R0 ) = {} R0 is connected.
a ray on the real axis missing. Note that C
Proof: any interval is connected. Thus we see that C R0 is polygonally simply connected
(PSC), and we can integrate along any closed curve in this region by applying the closed
curve theorem.
How about C = C {0} ? Applying the closed curve theorem for the difference contour

34

C2 C1 , we see that

Z
f (z)dz = 0
C2 C1

for f (z) holomorphic on C . An image of C2 (the outside counterclockwise contour) and C1


(the inside counterclockwise contour) is as follows:

Diagram 1: C1 and C2 in C .
Now lets pick off with Moreras theorem, which as you recall states
Theorem. (Morera) If f is continuous on an open set D and
ing the boundary of a triangle in D, then f (z) is holomorphic.

f (z)dz = 0 for be-

Corollary. If fn (z) is holomorphic on D open and fn f uniformly on compact subsets of D, then f (z) is holomorphic on D.
Proof. Proved last time.

The Riemann zeta function. The Riemann zeta function is a particular form of a
Dirichlet series or L-function given by
(z) =

X
1
.
nz
n=1

We claim that (z) is holomorphic on Re(z) > 1, and indeed we showed this last time. 
The gamma function. The gamma function, which is useful in probability and number
theory, is given by
Z

et tz1 dt.

(z) =
0

We claim that (z) is holomorphic for Re(z) > 0.


Remark. (z) is a natural extension of the usual factorial function in that (n + 1) = n!.
Show this yourself using integration byRparts and induction.

Proof of the claim. We show that 0 et tz dt is holomorphic in Re(z) > 1. To that


extent we work in 1 < R1 Re(z) R2 ; note that any compact subset of {z : Re(z) > 1}
is contained in one of these sets. Define
Z 1
Z n
fn (z) :=
et tz dt +
et tz dt,
1/n

which we will show converges uniformly. For t 1, we can bound


|et tz | et tRe(z) et tR2
because
tz = ez log t = e(Re(z)) log t eiIm(z) log z ,
35

the last term of which has norm 1. Then


Z

|fn (z)/2 f (z)/2|

et tR2 dt,

which converges for n = 1. So  > 0, N : n > N ,


Z
et tR2 dt < /2.
n

For t 1, |et tz | = et tRe(z) et tR1 and


Z
Z
Z
Z 1

1/n
1
1/n


t z
t z
t z
et tR1 dt
e t dt
e t dt
e t dt =


0
1/n
0
0
R
for R1 > 1. Let s = 1/t and ds = t12 dt, so that we get n e1/s s2R1 ds. We need
2R R1 < 1, i.e. 1 < R1 . Thus the sequence converge uniformly, so  > 0, N : n >

N, n e1/s s2R1 ds < /2.



Corollary. If f is continuous on D and f is holomorphic on D except at D, then
f is holomorphic on D.
Proof. Let = T . Then
Z
f (z)dz = 0

if 0
/ T . If T , we split the contour so that is located at the corners of the triangle.
If is at a corner, then
Z
Z
0 = lim
f (z)dz =
f (z)dz
l

by continuity of f .

Singularities of functions. (Newman and Bak 9) Singularities of functions will be


important when trying to access their behaviors around certain points so that we can define
things like Laurent series or perform residue calculus. Needless to say, an understanding of
the different types of singularities and how they pertain to classes of functions is important
for complex anaysis in general.
Definition. Given C, a deleted neighborhood D of is a neighborhood {}, e.g.
{z : 0 < |z | < }. f is said to have an isolated singularity at if f is defined and
holomorphic on a deleted neighborhood of .
Types of singularities. For D a deleted neighborhood of , we have the following types of
singularities:
1. Removable singularity: g(z) on D {} holomorphic with g(z) = f (z) on D.
2. Pole of order k: A(z), B(z) holomorphic on D {} s.t. A() 6= 0, B() = 0,
A(z)
f (z) = B(z)
and B(z) has a zero of order k at . Recall that we can expand B(z) into
a power series as B(z) = ck (z )k + ck+1 (z )k+1 .
3. Essential singularity: Neither of the above.

36

The following proposition allows us to recognize if something has a removable singularity.


Proposition. If
lim f (z)(z )

exists and is equal to 0, then f (z) has a removable singularity at .


Proof. Define h(z) as
(
f (z)(z ) z 6=
h(z) =
0
z = .
This is continuous, and holomorphic except at . So h(z) is holomorphic in a neighborhood
h(z)
of by the corollary to Moreras theorem. Moreover, we see that h() = 0, so g(z) = z
is holomorphic and equal to f (z) away from . (If h() = 0 and h is holomorphic, then
h(z)

z is holomorphic; one can prove this easily using power series.)
Likewise, the following proposition allows us to recognize a pole of f .
Proposition. Suppose f (z) is holomorphic in a deleted neighborhood of and n :
lim f (z)(z )n = 0.

Then letting k = 1 be the least such n, f (z) has a pole of order k at .


Proof. Again, let
(
f (z)(z )k+1 z 6=
h(z) =
0
z = .
h(z) is continuous on a neighborhood and holomorphic on a deleted neighborhood, so its
holomorphic on a neighborhood of again by the corollary to Moreras theorem. h() = 0
h(z)
g(z)
so g(z) = z
is holomorphic in D{} and g(z) = f (z)(z )k on D. Then f (z) = (z)
k.
(Note limz g(z) 6= 0 by assumption; also notice it exists).

For essential singularities, weve shown that it must be the case that limz f (z)(z )n
does not exist for all n. Notice as well the following:
If f is bounded in a deleted neighborhood, then f has a removable singularity.
If f < C1 |z|1N + C2 |z|N1 1 + ... + C as z 0, then f has a pole of order N .
We have a theorem for essential singularities, which tells us that the image of any deleted
neighborhood of an essential singularity under a holomorphic function is necessarily dense
in the complex plane:
Theorem. (Casorati-Weierstrass) If f has an essential singularity at and D is any
deleted neighborhood of , then f (D) = {f (z) : z D} is dense in C (i.e for any  > 0,
w C, B (w) f (D) 6= ).
Proof. Suppose not. Then B (w) with B (w) f (D) = . This means |f (z) w| > , or
1
1
1
|f (z)w| <  . So f (z)w is defined on D, a deleted neighborhood of , and bounded there.
1
Thus g(z) = f (z)w
is holomorphic on a neighborhood of .
1
= f (z) w, so that f (z) = wg(z)+1
is a ratio of two holomorphic
Now consider g(z)
g(z)
functions; we see that the singularity must be removable, or f (z) must have a pole.

37

Exercise left to the reader: prove the converse of the theorem.

Example. Consider f (z) = e1/z defined on C , which has an essential singularity at 0 C.


We claim that f (B (0) {0}) is C .
To see this, note that f (z) is the composition of z1 and then ez . Under z1 , B (0) {0}
gets inverted to fill C outside the ball. ez is 2i-periodic, so we can draw horizontal lines
in the complex plane at + 2n, n Z. Then the interior of any strip created by these
horizontal lines gets mapped under ez to C R0 . Our claim is that there are strips outside
B0 (1/) so that f (B (0) {0}) = C .


Diagram 2: The mappings z1 : B (0) {0} C (B (0) {0}) and


ez : {z : Im(z) 3} C R0 .
Laurent
expansions.
P
f (z) = k= ck z k .

Simply put, Laurent expansions are Taylor series of the form

Theorem. If
f (z) is holomorphic on the annulus A(R1 , R2 ) := {z : R1 < |z| < R2 },
P
then f (z) = k= ck z k converges on all of A(R1 , R2 ). (Taking R1 = 0 and R2 = is
also fine.)
We say
L.

k=

ak = L if

k=0

ak exists,

P1

ak =

l=1

al exists, and their sum is

Proposition. If
1
limk |ck |1/k
P

R2 and lim |ck |1/k R1 ,


k

then f (z) = k= ck z converges and is holomorphic on A(R1 , R2 ).


Proof. Write
 k

X
X
1
f (z) =
ck z k +
ck
.
z
k=0

The first sum is convergent for |z|


| z1

1
limk |ck |1/k

= |z|

k=1

limk |ck |1/k


limk |ck |1/k . The

. The second sum is convergent for

first sum is clearly holomorphic. The


P
second sum is holomorphic as a function of 1/z; e.g. g(z) = k=1 ck wk is holomorphic
1
1
for |w| R1 . The second sum is a composition of z and g, so by the chain rule its also
holomorphic. So f is holomorphic.


38

Example. The following Laurent series expansion converges on A(0, ):


e1/z = 1 +

1
1
1
+ 2 + 3 + ...
z
2z
6z


Example. Near z = 0 (pole of order 1), we have


1
1
1
1
=
= (1 2z + 3z 2 4z 3 + ...).
z(1 + z)2
z (1 + z)2
z
This is because
near z = 1.

z ( 1+z )

1
= (1+z)
2 and

1
1+z

= 1 z + z 2 z 3 + .... We can do the same




39

CHAPTER

10
MEROMORPHIC FUNCTIONS AND RESIDUES

Last time, we introduced the Laurent series for f (z) as f (z) =


defined on A (R1 , R2 ), the annular region centered at .

Theorem. If f (z) is holomorphic on A (R1 , R2 ), then f (z) =


the series converges on A (R1 , R2 ).
Proof. Do this yourself by considering the function
(
f (w)f (z)
w 6= z
wz
g(w) =
0
f (w)
w=z
and using the closed curve theorem.

ck (z

)k . This is

k= ck (z

)k and

Remark.P Laurent series are P


unique to every function. We define the principal part of
1
f (z) =
near as the sum k= ck (z )k and observe the following:
(1) is a removable singularity iff ck = 0 for k < 0.
(2) is a pole of order n iff ck = 0 for k < n, cn 6= 0.
(3) is an essential singularity iff ck 6= 0 for infinitely many negative k (why?).
Heres an idea: at a pole, we can make the function f (z) still be holomorphic if we ex so that f (z) = there and holomorphically so.
pand its range to C,
1
Proposition. If f (z) has a pole of order k at , then f (z)
is holomorphic near and
has a pole of order k.
1
1
Proof. Write f (z) = (z)
k g(z). Then g(z) holomorphic near , g(z) 6= 0, so f (z) =
(z)k
g(z) ,

g() 6= 0. This is holomorphic on a neighborhood at and has zero of order k. 

Definition. f is meromorphic on D if f (z) is holomorphic on D except at isolated singularities at which f has poles.
The proposition above can then be restated as: a function f being meromorphic in D
is holomorphic.
really means that f : D C
40

Theorem. Let f be meromorphic in C and at infinity, and suppose that limz f (z) =
for some w C. (The limit means M, R : |f (z)| > M when |z| > R.) Then f (z) is a
rational function.
C
is holomorphic, then f is a rational function.
Theorem. If f : C
Residues of functions. By introducing residues, we seek to deduce the residue theorem as a means for computing integrals more effciently. Given a holomorphic
Pfunction f (z)
in a deleted neighborhood of , by the Laurent expansion we have f (z) = ck (z )k .
We define the residue of f (z) at the point , Res(f ; ), as
Res(f ; ) = c1 .
From Cauchys integral formula, we notice that
Z
1
f (w)dw = c1 .
2i CR ()
Also recall the expression for ck :
ck =

1
2i

Z
CR (0)

f (w)
dw.
wk+1

We see from this that the limsup stuff shows that Laurent series are uniformly convergent
on A(r1 , r2 ) by comparison with geometric series.
On a ball of radius , we can compute the residue of f at by using the formula:
Z
Z
1
1
Res(f ; ) = c1 =
f (w )dw =
f (w)dw.
2i C (0)
2i C ()
There are a few ways to compute
residues in general:
R
(0) Compute the integral C () f (w)dw.
(1) Compute the Laurent series for f centered at .
(2) If f has a simple pole, then
c1 = lim (z )f (z).
z

Also, if f (z) = A(z)/B(z) for A a nonzero function, and B having a simple zero, then we
have:
Lemma. Res(f ; ) =

A()
B 0 () .

We finish by working out an example. Consider


Z
1
dx.
2
1 + x
We can compute this by first taking a semicircle contour of radius R, for which there is one
pole contained inside ( = i). If we compute the residue at this pole, we can use the residue
theorem to get the value of the integral (this is left to the reader as an exercise).
41

CHAPTER

11

WINDING NUMBERS AND CAUCHYS INTEGRAL


THEOREM

Today, well cover residues, winding numbers, Cauchys residue theorem, and possibly the
argument principle. This is in the textbook, 10.1, 10.2.
Last time, we defined resudies. For f holomorphic in a deleted neighborhood of , we
have
Z
1
Res(f ; ) =
f (z)dz = c1 ,
2i C ()
P
where c1 appears in the Laurent expansion for f around , f = cn z n . Note that f
is holomorphic in C () except for the point at .
1
Example. Res( 1+z
2 ; i)
Method 1: Find the Laurent expansion about i, i.e. in terms of z i.

1
i 1
1 1
1
=
=
+ (z i) + ....
1 + z2
(z + i)(z i)
2 z1 4 8
So c1 = i
2 .
Method 2: If f (z) has a simple pole at , i.e. if f (z) =
0, B 0 () 6= 0, then Res(f ; ) =

A()
B 0 () .

Set

1
Res( 1+z
2 ; i)

1
2i

A(z)
B(z)

for A() 6= 0, B() =

= i/2.

An application of this is as follows. Consider the integral


Z
1
dx = lim arctan(R) lim arctan(R) = /2 (/2) = .
2
R
R
1 + x
We can also do this by contour integration over the contours C1 , C2 , C3 , where C1 = [R, R],
C2 is the CCW semicircle connecting R to R, and C3 is the CCW circle omitting the point
at i. Then
Z
f (z)dz = 0
C1 +C2 C3

42

by the closed curve theorem, and by the estimation lemma we have


Z



1
1


dz R 2 =
0

2
R
R
C2 1 + z
as R . Also,
Z
C3

1
1
dz = 2iRes(
; i) = 2i(i/2) = .
1 + z2
1 + z2

So
Z

lim

1
dx =
1 + x2

after substituting for the values of the integrals over C1 , C2 , C3 .


Winding numbers. These intuitive give the number of times a path loops around a
point, and are also dealt with in topology.
Definition. Let : [0, 1] C be a contour curve with (t) 6= t. Then the winding number of about , denoted n(, ), is the value
Z
1
1
dz.
2i z
Theorem. n(, ) Z.
Proof. We have
Z

1
dz =
z

Z
0

0 (t)
dt.
(t)

d
dt

Note that the integrand on the RHS is log((t)), with a caveat that log is a multivalued
function. log z = log |z| + iArg(z) modulo 2i. Well show that
Z
0

0 (t)
dt = 2ik
(t)

for some k. This k = n(, ). Let F (s) =


0

Then F (s) =

0 (s)
(s) .

0 (t)
dt
1 (t)

Rs

(secretly, log((s))log((1)).

We claim that
F (s) =

(s)
.
(0)

To see this, let G(s) = ((s) )eF (s) . Then G0 (s) = 0 (s)eF (s) + 0 (s)eF (s) = 0,
so G(s) is a constant. Check that at s = 0, we get G(0) = ((0) )eF (0) , so indeed
eF (0) = (0)
(t) .
R 1 0 (t)
F (1)
Now eF (1) = (1)
= 1. So F (1) = 2ik
(t) . F (1) = 0 (t) dt and (1) = (0), so e
for k Z.

Corollary. (version of Jordan curve theorem) Let : [0, 1] C be a closed curve. Then
C image() is disconnected.
43

Lets call a simple closed curve one that doesnt intersect itself. For us, if has n(, ) = 0
or 1, then we call winding simple. The Jordan curve theorem shows that any closed
curve has what we call an interior and exterior. For the interior region B, we would have
n = 1. For the exterior region A, we would have n = 0.
Cauchys residue theorem. It turns out that we can evaluate any contour integral
of a holomorphic function by considering its residues and winding numbers.
Theorem. Let f be holomorphic on a region D with zero first homology (i.e. PSC, i.e.
closed curve theorem holds for holomorphic functions on D) except for isolated singularities
at 1 , ..., n D. Let be any contour curve in D. Then
Z
f (z)dz = 2i

n
X

n(, k )Res(f ; k ).

k=1

Proof. The gist is to repeatedly subtract the principal parts of f at each k . This proof is
a bit long to type up, so it is left as an exercise to the reader.


44

CHAPTER

12
THE ARGUMENT PRINCIPLE

Recall from last time that we had Cauchys residue theorem:


D connected (i.e. closed curve theorem applies), i.e.
Theorem. Let D be a region with C
simply connected (all loops contract to a point in D). Let f be holomorphic on D except
at 1 , ..., n , and let be a contour curve in D missing these points. Then
Z
f (z)dz = 2i

n
X

Res(f ; k )n(, k ).

k=1

The argument principle. This is essentially relating the number of zeroes and poles of f
to (# of times f winds around 0). Let be a regular contour curve, e.g. if C im(),
either n(, ) = 0 or n(, ) = 1. Intuitively, this means that the points are either inside
of or outside; the former is given by the set {z C im() : n(, z) = 1} and the latter
is given by the set {z C im() : n(, z) = 0}.
Theorem. If f is holomorphic in a simply connected region D, and a regular curve
in D, then the following are equal:
(1) (# of zeroes of f with multiplicity) (# of poles of f with multiplicity inside )
(2) The winding number around zero of f ((t)), i.e. n(f , 0)
R f 0 (z)
1
(3) 2i
dz
f (z)
Remark. This assumes that for f | 6= 0, there are no poles. The multiplicity is also called
the order.
Proof. Lets do (2) (3). For : [0, 1] C, we have
1
n(f ; 0) =
2i

Z
f

1
1
dz =
z
2i

0 (t)f 0 ((t))
1
dt =
f ((t))
2i

f 0 (z)
dz.
f (z)

The integrand is actually z


(log f ), which keeps track of how Arg(f ) changes and goes
around .
(3) (1). To do this we check the residues of f 0 /f . The only singularities of f 0 /f
occur at the zeroes or poles of f . Near a zero or pole of f , we have f (z) = (z )n g(z)

45

for g(z) holomorphic near , g() 6= 0. Also,


f 0 (z) = n(z )n1 g(z) + (z )n g 0 (z),
so

f 0 (z)
n
g 0 (z)
=
+
,
f (z)
z
g(z)

n
where we note that the residue at of z
is n, and g(z) 6= 0 implies that the term on the
right is holomorphic and is the residue at . Hence
Z 0
X
X
1
f
dz =
n(, ) order(f, )
n(, ) order+poles(f, )
2i f
pole of f
outside

pole of f
inside

And n(, ) 1 in all these sums.

Moreover, we can say that the zeroes of order n look like z n ; they wrap around 0 n
times CCW, and the poles of order n look like z1n ; they wrap around 0 n times CW.
The idea of computing the number of zeroes in a curve by computing an integral is quite
nice. Most of the time, however, we just use the following corollary (also dropping the idea
of poles for now, since they arent really used in it):
Corollary. (Rouche Theorem) (Assume f has no zeroes on .) Let f, g be holomorphic in
the unit disk (in ), and ||g(z)|| < ||f (z)|| on the unit circle (on ). Then # of zeroes of f
in the unit disk = # of zeroes of f + g in the unit disk, for regular.
Examples. The # of zeroes of 3z 8 (+iz 6 + 1) in the unit disk is the same as 3z 8 , which has
8 zeroes. iz 8 + 3z 6 + 1 has 6 zeroes in the unit circle. iz 8 + z 6 + 3 has no zeroes.

Proof of Rouches Theorem. First note that for any functions A and B,
A0
B0
(AB)0
=
+
AB
A
B
Now write


(f + g) = f

g
1+
f


,

and note | fg | < 1 on . The # of zeroes of f + g in is given by


1
2i

(f + g)0
1
dz =
f +g
2i

f0
1
dz +
f
2i

Z (1 + g )0
f

(1 + fg )

dz,

but this is just




g
= (# of zeroes with multiplicity of f in ) + n 1 +
;0 .
f
Note that 1 + fg
is contained inside B1 (1) {Re(z) > 0}, the upper half-plane. So the
winding number is zero, and the above is
Z
1
1
=
dz
2i z
46

where is the curve traced out by the fact that 1 +


given by the principal branch of log z in Re(z) > 0.

g
f

< 1, which has an antiderivative




......................................................................................
Generalized Cauchys integral formula. Let be regular, and z be inside . Then
Z
k!
f (w)
(k)
f (z) =
dw.
2i (w z)k+1
Proof. The residue of

f (w)
(wz)k+1

at zero is


f (k) (z)
f (w)
1
= (w z)k f (w) =
.
k+1
w z (w z)
k!


Corollary. Suppose fn is holomorphic and fn f uniformly on compact sets (hence


holomorphic by Moreras theorem). Then fn0 f 0 uniformly as well.
Proof. We have
Z
1
fn (w) f (w)
fn0 (z) f 0 (z) =
dw
2i (w z)2
for = Cr (), z Dr (). On Dr/2 (), we have

Z
1
1 fn (w) f (w)
4
0
0
|fk (z) f (z)| =
dw
2r 2 max |fn (w) f (w)|,
2 (w z)2
2
r Cr ()
where the r42 comes as an upper bound on
Now choose n large enough so that

1
|wz|2

because |w z| r/2 =

|fn (w) f (w)| <

1
|wz|2

4
r2 .

r
4

on Cr (). Then |fn (z) f (z)| <  for z Dr/2 (), n > N .
Now cover the compact set in euqation by disks of half the radius. Because the set is
compact, finitely many suffiice. Take the largest N among these.

Theorem. (Hurwitz) Suppose fn f holomorphic in D, and uniformly so on compact
sets. Suppose that none of the fn s are zero in D. Then either f = 0 on D or f has no
zeroes.
Proof. Suppose f 6= 0, and let f () = 0 for some D. is not the limit of zeroes of f
(other than 0), because then f = 0 by the uniqueness theorem. Hence there is some closed
disk D (),  > 0, with no zeroes of f in D () other than . We have fn0 f 0 uniformly,
and there are no zeroes on C () of f ; none of fn are zero, so f1n f1 uniformly on C ()
(show this yourself). Thus
fn0
f0

fn
f
uniformly on C (). It follows that
Z
Z
1
fn0
1
f0
lim
=
.
n 2i C
fn
2i C () f
()
47

So
lim # of zeroes of fn in D () = # of zeroes of f in D () = 1,

a contradiction, so we cant have this isolated zero of f .

48

CHAPTER

13
INTEGRALS AND SOME GEOMETRY

Recall the residue theorem: let k be the singularities of f . Then


Z
X
f (z)dz = 2i
Res(f ; k )n(, k ).

There are many applications of this. If P, Q are polynomials, Q(x) 6= 0 for real x, and
deg Q deg P + 2, then
Z

X
P (x)dx
dx = 2i
Res
Q(x)dx
k=1


P (z)
, k .
Q(z)

You can show this with the usual residue calculation (yes, the ones that we beat to death
in our review session...).
Example.
Z

1
dx
x6 + 1

The poles are where z = 1, e.g. at e for = 6 , 3


6 , ....
Recall that, for f (z) = A(z)/B(z) with B() = 0, A() 6= 0, if B 0 () 6= 0 then this is a
simple pole, and the residue is given by Res(f ; ) = A()/B 0 ().
So using this we get that the residues are
6

Res(f (z); ei/6 ) =

1
6e5i/6

1
6ei/2
1
Res(f (z); e5i/6 ) = i/6
6e
Res(f (z); ei/2 ) =

Then

x6

1
2i 5i/6
2
dx =
(e
+ ei/2 + ei/6 ) =
.
+1
6
3
49

Geometry of complex functions. Recall the open mapping theorem: if f is holomorphic and nonconstant, and if U is open in C, then f (U ) is open, i.e. given , : B (z)
f 1 (B (w)) = {x C : f (x) B (w)}, i.e. |z | < = |f (z) f (w)| < . Open means
that ,  : f (B (z)) B (w), so |f (z) | <  = : |z | < , f () = .
Proof. Let C. Wlog f () = 0; take C inside B (). Take C inside B (); min f (C)
exists and its not zero, so call it z . Let w B (0). Then for z C, |f (z) w|
|f (z)| |w| 2  = . For z = , |f () w| = | w| < . Hence min |f (z) w| for
z B/2 () is not achieved on the boundary. Thus, by the minimum modulus theorem,
a zero, e.g. f (z) w = 0 in B.
Theorem. (Schwarz Lemma) Let D be the unit disk. If f : D D is holomorphic
(extending continuously to the boundary) and f (0) = 0, then
(
|f (z)| |z|
|f 0 (0)| 1
with equality in either iff f (z) = ei z.
Proof. Define
(
g(z) =

f (z)
z
0

f (0)

z=
6 0
z = 0,

(z)|
which is holomorphic on the circle of radius ; |g(z)| = |f|z|
1 . By the maximum modulus
principle, we know in fact |g(z)| 1 for all z D. Let 1.
So if a max is achieved on the interior, then f (z) is constant, i.e. g(z) is constant
if |f (z)| = |z| for any z D1 (0), z 6= 0, or |f 0 (0)| = 1. This means g(z) = ei ; then
g(z) = f (z)/z (f 0 (0)), so f (z) = ei z.


Corollary. If a map from D D, f (0) = , has a maximum value of |f 0 (0)| among


maps D D, then f is surjective.
Proof. Consider h(z) = B (f (z)). h(0) = f 0 (0)B0 (0), h(z)D D, h(z) : 0 7 0, |h0 (0)|
1.
The conclusion is that
1
.
|f 0 (1)| 0
B () = 1 ||2
Notice that this is achieved for inverse of B , which is B .
B (B (z)) = z.
Also note that
B (z) =

z
,
1 z

B0 (z) =

1
,
1 ||2

and

so B0 (0) = 1 ||2 , B (0) = 0, |B (z)| 1 for |z| 1.


We used h0 (0) 1. We also know |h(z)| |z| and h(z) = B (f (z)); |B (f (z))| |z|.
50

We can therefore conclude that


|f (z)| |B (z)|
if f (0) = , f : D D.
Riemann mapping theorem. (casually stated) Given a simply connected region U C,
U 6= C, f : D U where D denotes the open unit disk, which is a holomorphic bijection
with holomorphic inverse (and essentially unique).

51

CHAPTER

14

FOURIER TRANSFORM AND SCHWARZ REFLECTION


PRINCIPLE

We will go over contour integration via Fourier transforms and infinite sums, 11, the
Schwarz reflection principle 7.2, and the Mobius transformations 13.2.
The Fourier Transform. Let f : R R or R C. The Fourier transform of f is
given by
Z
f(y) =
f (x)e2ixy dx.

The inversion formula is

f(y)e2iyx dy.

f (x) =

2iyx

This is writing f in terms of e


, and has applications in physics, probability theory (proving CLT via generating functions), etc. Look at Wikipedia for a summary of how useful it is.
Example. f (x) =

1
1+x2 .

f(y) =

1
e2iyx dx.
1 + x2

Assume y 0. Let C(R) = C1 (R) + C2 (R) denote the semicircle contour with C1 (R)
denoting the part on the real axis and C2 (R) the semicircle part. Then consider
Z
1
e2iyz dz = 2i Res(gy (z); i)
2
1
+
z
C1 (R)+C2 (R)
where gy (z) is the integrand. With a simple bounding argument, we note that
Z

gy (z) =
C1 (R)

52

1
e2iyx dx.
1 + x2

2y
Also, Res(gy (z); i) = e 2i . Thus f(y) = e2y for y 0. Now for y 0 we use the same
2y
contour reflected over the real axis. Then Res(gy (z); i) = e 2i =

1
e2iyx dx = 2i n(C, i) Res(gy (z); i) = e2y .
1 + x2

So
f(y) = e2|y| .

The applications of Fourier transformations abound. Frankly Prof. Cotton-Clay hasnt
used it much, but natural scientists do. A very nice source for getting to know the complexanalytic details of Fourier transforms is Stein and Shakarchi.
P
2
Example. Verify that n=1 n12 = 6 .
P
2
We can do this by showing n=,n6=0 n12 = 3 . This is left as an exercise to the reader
(or see the textbook). Note that this is (2), and the method that you use for computing
this generalizes to (2n), n Z+ ; this was first computed by Euler.

Conformal mappings. These are functions that preserve angles. The Schwarz reflection principle is useful as a tool here.
Theorem. (Schwarz reflection) Let H denote the upper half-plane, {z C : Im(z) > 0}.
Suppose that f is holomorphic in a region D H, that f is continuous on D L where
L = D R, and that f is real on L. Then
(
f (z) z D L
g(z) :=
f (z) z D
is holomorphic on D L D.
Proof. First, g is continuous because both parts agree on L. Second, g is holomorphic
on D, and on D:
lim

h0

g(z + h) g(z)
f (z + h) f (z)
= lim
= f 0 (z)
h0
h
h

exists, so g is holomorphic on D. Then use Moreras theorem to show that g(z) is holomorphic.

Mobius transformations. This answers the question: what are the injective, holomorphic
C?

maps C
C
is holomorphic (meromorphic on C), then f is a rational
Recall that if f : C
P (z)
function. Solving = Q(z) , we have P (z) Q(z) = 0, and the LHS is a polynomial of
degree max(deg P, deg Q). For f (z) to be injective, we need max deg = 1. So
f (z) =

az + b
,
cz + d

53

where (a, b) and (c, d) are linearly independent; equivalently, ad bc 6= 0. Maps of the form
above are called Mobius transformations. It is obvious that the inverse f 1 is given by
f 1 (w) =

dw + b
cw a

with ad bc 6= 0.
Proposition. If
f (z) =

az + b
,
cz + d

g(z) =

z +
,
z +

then letting


A
C

B
D


=

a
c

we have
f (g(z)) =

b
d




,

Az + B
.
Cz + D

Proof. This is just a simple check.

54

CHAPTER

15

MOBIUS
TRANSFORMATIONS

Today well talk about Mobius transformations 13.2, cross ratios, and automorphisms of
D and H. Recall that we defined a Mobius transformation to be a function f (z) of the
following form:
az + b
f (z) =
.
cz + d
i.e. the holomorphic maps C
C
with holomorphic
These are the automorphisms of C,
inverse. We showed that composition translates into matrix multiplication:
fM1 (fM2 (z)) = FM1 M2 (z),
where M s are the matrix forms of the transformation.
These 2 2 invertible matrices with complex coefficients are known as the general linear
group of order 2 over C:



a b
GL2 C =
: a, b, c, d C, ad bc 6= 0 .
c d
GL2 C has Lie group structure (it is an algebraic group endowed with manifold structure,
e.g. the group operations are C ). Let M = {mobius transformations}. Then, modding
out by complex scalars 6= 0, we have in fact that
M=

GL2 C
= P GL2 C,
C

where P GL2 C is the projective linear group of order 2 over C, which is also a Lie group.
We mod out by scalar multiples of C because




a b
a b
f
=f
,
c d
c d
has no other coincidences; e.g. multiplication by a complex scalar gives the same transformation. We now claim that the generators for M are as follows:
Generators for M:
55

(1) Dilations/rotations: z 7 az.


(2) Translations: z 7 z + b.
(3) Inversion: z 7 z1 .
Claim. These generate M.
Proof. We get the map z 7 az + b from dilation and translation. Suppose c 6= 0; we
want a map of the form z 7 az+b
cz+d . Well,
z 7(1,2) cz + d 7(3)

az + b
1
7(1,2)
.
cz + d
cz + d

Lets show that we can achieve the latter map. We can write





az + b
1
A + Bcz + Bd
ad bc
1
a
=A
+B =
=
+ ,
cz + d
cz + d
cz + d
c
cz + d
c
for which we want B =
f (z) = ad z + db by (1,2).

a
c

and A = b ac d = adbc
c . Note that setting c = 0 we have


Corollary. {Mobius transformations} : {circles and lines} {circles and lines}.


Proof. All of the generators do.

Lets examine the dimensions of the groups we have above. Note that GL2 C has 4 complex
dimensions, or 8 real dimensions (by the obvious identification C
= R2 ). This is because
there are 4 degrees of freedom in choosing elements in the matrix; another way would be to
show this is to note that GL2 C is open in Mat2 C under the continuous determinant map,
the latter of which has complex dimension 4.
Since we are eliminating one degree of freedom by imposing the condition det(A) 6= 0
for A M, we see that dimC M = 3, or dimR M = 6. Given this, we can say something
about the action on the 3 points {, 0, 1}:
Lemma. Given f M and f 6= id, f has at most 2 fixed points (points w C : f (w) = w)
in {, 0, 1}.
2
Proof. Let c 6= 0, so that f (z) = az+b
cz+d and setting f (z) = z we have az + b = cz + dz,
a
which maps 7 c 6= because c 6= 0. For c = 0, we have 7 , and az + b = z has
1 solution.

Proposition. There exists a unique Mobius transformation f (z) sending 1 , 2 , 3 to
, 0, 1 respectively, and f (z) is given by
f (z) =

(z 2 )(3 1 )
.
(z 1 )(3 2 )

Note that if any j = , we cross out both terms in which it appears.


Proof. The existence of f (z) is trivial. Uniqueness follows thus: if we have two transformations f, g satisfying the above properties, we consder g f 1 , which sends 7 , 0 7
0, 1 7 1. So three points are fixed, which implies g f 1 (z) = z = f (z) = g(z).

This suggests that dimC M = 3 is no mistake. Indeed, we can define the cross-ratio

56

[1 : 2 : 3 : 4 ] as the number w C where 4 is sent if 1 7 , 2 7 0, and


3 7 1 by a Mobius transformation. Note that this number is equal to
(4 2 )(3 1 )
.
(4 1 )(3 2 )
Proposiiton. The cross-ratio is preserved by Mobius transformations, i.e. if f (z) = az+b
cz+d
with ad bc 6= 0, then [1 : 2 : 3 : 4 ] = [f (1 ); f (2 ); f (3 ); f (4 )].
Proof. Let g be a Mobius transformation sending 1 7 , 2 7 0, 3 7 1. We have
[1 ; 2 ; 3 ; 4 ] = g(4 ). Then g f 1 sends f (1 ) 7 , f (2 ) 7 0, f (3 ) 7 1, and
g f 1 (f (4 )) = g(4 ).

Proposition. There exists a non-unique Mobius transformation sending z1 7 w1 , z2 7 w2
and z3 7 w3 , and it satisfies, for w = f (z),
(z z2 )(z3 z1 )
(w w2 )(w3 w1 )
=
.
(w w1 )(w3 w2 )
(z z1 )(z3 z2 )
Proof. Existence comes from the formula and the cross-ratios. Uniqueness follows as
before.

Lets use the above propositions above to find a Mobius transformation f sending D H
bijectively. If we visualize how f sends D C H C, we can also visualize how it maps
In particular, it should send 1 7 , 0 7 i, 1 7 0, i 7 1, 1 7 1. Let w = f (z).
C.
Then using the equation above we have
(z 0)(1 + 1)
2z
z + 1
(w i)(0 )
=
= iw + 1 =
= w = i
.
(w )(0 i)
(z + 1)(1 0)
z+1
z+1
Claim. This sends D to H.
Proof. Lets go with a more geometric proof. Note that f (i) = 1. We have a subclaim:
there exists a unique circle or line through the 3 points.
given 3 points in C,
Proof of subclaim. Send the 3 points to , 0, 1 by a Mobius transformation. Theres a
unique circle/line through these, namely R. We showed that Mobius transformations sends
circles and lines to circles and lines.

Now f sends the unit circle to the real line by the formula f (eit ) = tan(t/2), where
0 t < 2. We make another subclaim, namely that f sends points inside D to points
inside H.
Proof of subclaim. Use the following proposition on curves that are circles and omits 1
(a Pac-Man curve).

Proposition. Let f (z) be a Mobius transformation, C, a contour curve, and
/ the
interior of . Suppose n(f (); f ()) = 0. Then n(; ) = n(f (); f ()).
Proof. Under dilations,
Z
Z 1 0
1
dz
1
(t)dt
n(; ) =
=
dt.
2i z
2i 0 (t)
The equalities for the other generators follow similarly. In particular, note that for inversion
we want to show the formula n( 1 ; 1 ) = n(; ) + n( 1 ; 0), from which we can conclude the
57

proposition. This is a simple check using the definition of winding number and the chain rule.
Maps from regions into other regions. We know that log maps to a horizontal strip.
In particular, log(z) : H {z : 0 < Im(z) < }, so
1
log(z) : H R [0, 1].

We then see that


D f H g R [0, 1]
for f (z) = i 1z
1+z , g(z) =
strip given by

log(z). Thus we have an automorphism from the unit disc to a


1z
1
log(i
) : D R [0, 1].

1+z

58

CHAPTER

16
AUTOMORPHISMS OF D AND H

We were looking at H = {z C : Im(z) > 0} and D = {z C : |z| < 1}.


z+i
Claim. f (z) = i z1
z+1 maps D to H, and g(z) = z+i maps H to D. Each are one-toone, holomorphic, and inverses of each other.
Proof. Check the last part; we will show that f maps D to H. Note that, under f ,
1 7 0, 1 7 , i 7 i( i1
i+1 ) = 1. Hence, like last time, we see that the unit circle is
mapped to the real line, since the 3 points determine a circle or a line. (Also note 0 7 i,
i 7 1.) Apparently, the real axis is mapped to the imaginary axis. Now fill up D by
circular arcs through 1 and 1. These map to straight lines through 0 and , all H
because they intersect the upper half of the unit circle, which is the image of the imaginary
axis D.
Another proof is the use the argument principle. Notice that the winding of f ((t))
around C is the number of zeroes of f (z) = ... but this argument is sketchier, so we
leave it to you to fill in the gaps.


Heres an unnecessary but amusing claim:


Claim. Let (t) be a contour curve and 0,
/ im(). Then




1 1
1
,
= n(, ) + n
, 0 = n(, ) n(, 0).
n

Proof. We have
Z
n(1/, 0) =
1/

dz
=
z0

Z
0

1
d( t
)
1
(t)

Z
=
0

0 (t)
dt = n(, 0).
(t)

On the other hand, we have


Z 1 d( 1 )
Z 1
Z 1
Z 1
0 (t)/(t)2
0 (t)
0 (t)
(t)
n(1/, 1/) =
=
=
dt
=
dt
1
1
2
0 (t)
0 1/(t) 1/
0 (t) (t)
0 (t)[(t) ]
Z 1
Z 1
0 (t)
0 (t)
=
dt +
dt = n(1/, 0) + n(, ),
(t)
0
0 (t)
59

which gives the result.

Automorphisms of D and H. What are the injective, surjective, holomorphic maps


D D with holomorphic inverses? Any such map is in the automorphism group of D,
denoted Aut(D). Recall that the Schwarz lemma tells us that if f : D D is such that
f (0) = 0, we have |f (0)| 1 and |f (z)| |z| for all z D, with equality in either where
f (z) = ei z.
Corollary. If f : D D, f (0) = 0, and f is an automorphism of the disk (there exists a holomorphic inverse f 1 : D D), then f (z) = ei z.
Proof. |f (z)| |z| = |f 1 (z)| |z| so |z| |f (z)|, e.g. equality is obtained and
f (z) = ei z.

z
for || < 1. This maps D D and 0 7 , 7 0. Hence
Recall that B (z) = 1z
B (B (z)) maps D D, 0 7 0, 7 , and their composition is the identity map.

Theorem. The automorphisms of D are precisely the maps




z
i
f (z) = e
.
1 z
Proof. Suppose f is an automorphism of D with f () = 0. Then f (B (w)) takes 0
to 0 and f (B (w)) = ei w. Take w = B (z) and B (w) = z. Then f (z) = ei B (z) =
z
ei ( 1z
).

Corollary. The automorphisms of D are the Mobius transformations taking D to itself.
Corollary. The automorphisms of H are the Mobius transformations taking H to itself.
Theorem. The automorphisms of H are precisely the maps f (z) = az+b
cz+d with ad bc > 0
and a, b, c, d R.
Proof. Take x1 7 , x2 7 0, x3 7 1 in R {} via a Mobius transformation for
x1 , x2 , x3 all distinct (note that there is a unique such Mobius transformation). Write it
down:
(z x2 )(x3 x1 )
.
f (z) =
(z x1 )(x3 x2 )
The collection of these is the collection of maps f (z) =
f (i) H, f (i) =

(bd+ac)+i(adbc)
.
c2 +d2

az+b
cz+d ,

ad bc 6= 0. Now check when

We see that f maps into H iff ad bc > 0.

Here is a classification of the automorphisms of D (and H), up to conjugation (e.g. for


f, g : D D, g 1 f g : D D is allowed):
Identity: f (z) = z
Elliptic automorphisms: f (z) = ei z (rotation on D)
Parabolic: f (z) = z + b on H
Hyperbolic: f (z) = az on H, for a R
Claim. Every automorphism of D (or H) is conjugate to precisely one of these.
60

Proof. Assume f is not the identity. We claim that every automorphism of H (or D)
fixes either one point inside H, one point on R, two points on R, and no others. To see this,
note that every Mobius transformation fixes one or two points. If one is in H, then the full
Mobius transformation is given by the Schwarz reflection

zH
f (z)
g(z) = something in R {} z R {}

f (z)
z H.
So if we have one in H, then there can be no more. Either this happens, or one/two points
are on R.
If one fixed point is in H, we conjugate to D and see its at D; conjugate by B
to get it at 0, then get a rotation. In particular, let f : H H, f () = f (). Take our
F : D H, F 1 : H D. F 1 f F then fixes F 1 () = . Then taking B (which is
its own inverse), we have B F 1 f F B sends 0 7 0, D D, so it equals ei z.
If we fix one point on R, we conjugate to the disk, then conjugate by rotation so that the
fixed point on the unit circle is at 1. Then we conjugate back to H so that the fixed point
is at (of Mobius transformation); i.e. f (z) = az + b. We need b 6= 0, else 0 7 0. We also
have a > 0 by ad bc > 0. If a 6= 0, we fix another point so that a = 1 and f (z) = z + b.
If we fix two points in R, we again place one at and conjugate the other by A(z + 1) for
appropriate A; then f (z) = az, a > 0.

In fact, these are the symmetries of the hyperbolic plane (taking the line y = 0 to be a
sort of line at infinity). As an aside, the hyperbolic metric on H is given by
ds2 =

dx2 + dy 2
,
y2

for y > 0. The arc length of (t) in the hyperbolic metric is


(t) = (x(t), y(t)) : [a, b] H.

61

ds =

R b x0 (t)2 +y0 (t)2


a

y(t)

dt for

CHAPTER

17

SCHWARZ-CHRISTOFFEL AND INFINITE PRODUCTS

Today well be talking about Schwarz-Christoffel functions 13.3 and infinite products and
Eulers formula for sin(), 17.3.
Schwarz-Christoffel maps. These are holomorphic maps from H to a single polygon.
Well see next time how it follows from the Riemann mapping theorem. But first off we can
note that there are only a limited number of ways to do this.
Lets draw a polygon with vertices vi , interior
angles P
i , and exterior angles i
P
(i + i = ). Note that 1 < i < 1, and
i = 2 =
i = 2.
Proposition. Given angles 1 , ..., n and that a1 , ..., an1 are points on R, then
Z z
dz
f (z) =

n 1
1
0 (z a1 ) ...(z an1 )
gives a holomorphic map from H to a polygon with angles i , and has a continuous extension to R with f (ai ) = vi , f () = an .
Remarks on the function. This is a contour integral taken on H R. We can take a
branch of (za1j )j = (z aj )j which is positive for z real, > aj and extends downward,
e.g. is defined on C {aj ir}, r > 0. We can require 0 6= aj , but in fact this still makes
R1
sense if 0 is one of those: 0 x1 dx converges for < 1. Lets examine the argument of
Qn1
f 0 (z) = j=1 (z aj )j on R:
Claim. Arg(f 0 (z)) is constant on the connected segments of R {j }. Also, Arg(f 0 (z))
increases by j when preserving aj .
Proof. This follows from:
Claim. Let z be defined on C {negative imaginary axis}, and be and positive real
for z real and positive. Then Arg(x ) = 0 for x > 0. If x is real, Arg(x ) = , x < 0.
Proof. Write z = rei . Then z = r ei gives the desired branch, which is in particular our negative real axis. So if x is negative, i.e. = rei , then we get x = r ei ,
i.e. Arg(x ) = . We can conclude that on the boundary (R) of H, f (z) maps to
62

straight lines, turning by angles k to L.

R Q
Claim. This closes up, e.g. (x aj )j dx = 0.
Proof. Use a semicircle contour, but skip over the poles on the real line. Then use a
combination of the closed curve theorem and bounding on the arc.

Example. sin z maps ( 2 , 2 ) (0, ) to H. We can view the domain as a triangle with
exterior angles /2, /2, . So, in some sense, sin1 z : H this strip. We can then write
Z z
Z z
dz
dz

=
,
sin1 (z) =
1/2 (1 z)1/2
2
(1
+
z)
1

z
0
0
for some branch of the root function. (This is because

1
d
dx (sin

x) =

1
.)
1x2

Infinite
Q products. We want to segue into the question, when does the infinite product k=1 ak converge?
Q
P
Lemma.
ak converges
log(aP
k ) converges, for Re(ak ) > 0.
Qn iff
n
Proof. Let pn := k=1 ak , sn = k=1 log(ak ). Then log pn = sn and pn = esn . e and
log are continuous, so convergence for one gives convergence for another.

Q
P
Lemma. (1
|ak | converges.
P+ ak ) converges if
P
Proof. If |ak | converges, then past some point N , |ak | < 1/z. So we check: k=N log(ak )
converges.






1 1
a3
a4
a2
| log(1 + ak )| = ak k + k k + ... |ak | 1 + + + ... 2|ak |.
2
3
4
2 4
P
P
P
So N | log(1
ak )|
2|ak | converges, hence N log(1 + ak ) converges; finally, this
Q+

implies that N (1 + ak ) converges.



P
Proposition. Suppose fk (z) is holomorphic
on a region D, with |fk (z)| uniformly converQ
gent on compact subsets of D. Then k=1 (1 + fk (z)) is uniformly convergence on compact
subsets and hence convergent.
Proof. Use the proof of the lemma in uniform convergence along with Moreras theorem.

Example. (Euler, 1734) We have the infinite product formula for sin(z):


Y
z2
sin(z) = z
1 2 .
k
k=1

Theorem. (Weierstrass product) There exists an entire, holomorphic function with zeroes
only at {k }k=1 if ak . Furthermore, the order of zero is the number of times repeated
in the list.
Q
Q
Proof. First attempt to go about it: f (z) = (z ak ) = f (z) = (1 azkk ) converges
P 1
if
|ak | converges.

63

Claim. If

1
|ak |2

converges, then
f (z) =



Y 
z
1
ez/ak .
ak

P
Proof. This converges if
|(log(1 azk ) + azk )| does. Suppose
is possible for k > N since ak . Then we can bound


|z|2
1 1
|z|2
2 1 + + + ... 2
,
ak
2 4
|ak |2
P 1
i.e. suppose
|ak |2 converges.

|ak |
2

> |z| for |z| R; this

Back to the full proof of Weierstrasss theorem: Let




Y 
z
f (z) =
1
Ek (z) ,
ak
for
Ek (z) = e

z
ak

z
+ 2a
+...+
k

zk
kak
k

Check via a diagonal argument that this works; in particular, consult your textbook.

Proposition. (Eulers formula)


sin z = z

Y

z2
1 2
k

Proof. Take the quotient


z

Q
1

Q(z) =

z2
k2

.
sin z
This is an entire function with no zeroes. Well argue that this has growth at most Aez ;
lets claim that it follows that Q(z) is constant.
R z 0 (z)
We note that Q(z) is even. So, taking the log of Q(z), we have 0 QQ(z)
dz. This has
growth at most constant = |z|3/2 for large values. So log Q(z) is linear, i.e. Q(z) = AeBz
with B = 0 since Q is even, i.e. Q(z) is constant. Q(z) is constant because limz0 sin(z)
=1
z
Q
z2
and limz0 k=1 (1 k2 ) = 1.
Now lets talk about the growth. sin1z is bounded on a square with sides from N 1/2
to N + 1/2 (exercise), and likewise on the edges. To bound the rest...actually, well save
that for next time because were out of time.

64

CHAPTER

18
RIEMANN MAPPING THEOREM

Guest lecture: Joe Rabinoff


Note: These notes are particularly messy since the scribe wasnt paying too much attention,
so use in discretion.
Youre lucky that we cover the Riemann mapping theorem, which is one of the most interesting results youll learn in this class. The context is the following:
R
Let regions R1 , R2 ( C be open and simply connected, with R1 connected and C
connected. Then there exists : R1 R2 that is onto, one-to-one, and holomorphic.
Thus 1 : R2 R1 is also holomorphic. As far as complex analysis is concerned, these
regions are identically the same. Now lets introduce a bit of reductionism here. Assume
: R2 U = {x C : |x| < 1}, so that we have the following diagram:

R1
R2
i

U
We need R1 6= C: if : C U , then is bounded, then constant by Liouvilles theorem.
Theorem. Let R ( C be open, R 6= , and R 6= C. Choose a z0 R. Then there exists a
unique, one-to-one, holomorphic map : R U such that (z0 ) = 0 with 0 (z0 ) R0 .
Proof. Uniqueness: suppose 1 , 2 : R U both satisfy the theorem. Then : 2 1
1 :
U U is one-to-one, onto, and holomorphic, and (0) = 2 (1 (0)) = 2 (z0 ) = 0.
By the Schwarz lemma, we know that |(z)| |z| for all z U , and if is onto,
(z) = ei z. Then 1 (0) = ei R0 = ei = 1 = (z) = z. This implies
0 (0)
2 = 1
= Id = 2 = 1 , as 0 (0) = 02 (0) > 0. So uniqueness is easy. Lets work
1
1
out an example for the other parts:
Example. Let R = U , : R = U U . If (0) = 0, then by Schwarzs lemma, we

65

know that is onto |0 (0)| = 1 = |0 (0)| is the largest possible among all maps
U U if 0 7 0.
z
is such that B : U U , B (0) = ,
If : U U, (0) = 6= 0, then B (z) = 1z
B () = 0; e.g. B swaps 0 and . Then |( B )0 (0)| = |0 ()||B0 (0)| |0 ()| is
constant. It turns out that is one-to-one and onto = |0 ()| is maximized.
In general, we want to find : R U thats one-to-one and holomorphic, with (z0 ) =
0, |0 (z0 )| maximized. Even more precisely, we will take
F = {f : R U : f is one-to-one, holomorphic, and f 0 (z0 ) = 0}.
Well show
(a) F 6=
(b) supf F f 0 (z0 ) = M <
(b) There exists f F : f 0 (z0 ) = M .
(c) If = f from above, then : U U is one-to-one and onto, (z0 ) = 0, and
0 (z0 ) > 0.
In this case, the function we will cook up looks like B .
First we want f : R U that is one-to-one and holomorphic. Suppose there exists

. Then this
D(p0 , ) := {z : |z p0 | < } such that D(p0 , ) R . Let f (z) = zp
0

is one-to-one since its a Mobius transform, if z R = |f (z)| = |zp0 | < = 1.


q
zp0
What is no such disc exists? Choose any p0
/ R, and define g(z) =
z0 p0 with
g(Z0 ) = 1, one-to-one since g(z)2 is. Do this by choosing the appropriate square root. How

do we choose a branch of
? Well,
r

and


log

z p0
z0 p0

z p0
= exp
z0 p0

1
log
2

z p0
z0 p0



= log(z p0 ) log(z0 p0 ) =
z0

This makes sense by the closed curve theorem, as

1
p0

d
.
p0

is holomorphic on R.

Claim. There exists > 0 such that |g(z) + 1| > for all z R.
Proof. If not, then z0 , z1 , ... R such that g(zi ) 1. Then

zi p0
z0 p0

1, and

zi p0
z0 p0

squaring
= 1 = lim(zi p0 ) = lim(z0 p0 ) = lim zi = z0 . But by continuity of g,
lim g(zi ) = g(z0 ) = 1, a contradiction. This proves (a).
Proof of (b). Since R is open, D(z0 , 2) R for some > 0. If f F, then


Z
Z
1 Z

0
f
(z)
|f 0 (z)|
1
1
1
1


0
|f (z0 ) =
dz
dz
dz =
2
2
2i C(z0 ,) (z z0 )2 2
|z z0 |
2

= sup |f 0 (z)|
f F

66

1
<

For (b), choose f1 , f2 , ... F such that limn |fi (z0 )| = M . Suppose that we knew
that fi uniformly, where is holomorphic on R.
Claim. F.
Proof. fi0 (z) 0 (z0 ) = M . Also, for all z R, |(z)| = lim |fn (z)| 1 so : R U =
closed ball.
By the open mapping theorem, (R) is open, so (R) U .
Why is one-to-one? We need the fact that a limit of one-to-one functions is again oneto-one. Suppose not. Then z1 , z2 R such that (z1 ) = (z2 ) = a U , and let D1 , D2
be discs around z1 , z2 such that D1 D2 = . Now a is not constant, since a has
a zero on Di . Thus fi a has a zero on D1 for infinitely many i, which implies that fi a
has no zeroes on D2 for all such i. Thus fi (z0 ) a. Apply Hurwitz to the subsequence of
such fi on D2 , so a has no zeroes on D2 , which implies (z1 ) = a, a contradiction. So
weve checked that is one-to-one, holomorphic, and onto.
We need to show that the fi converge uniformly on compact subsets K R. Well only use
that |fi | 1 for all i. This is Montels theorem.
Theorem. If f1 , f2 , ... : R U is any sequence of functions, then there is some subsequence fi1 , fi2 , ... that converges uniformly on compact subsets K R.
Proof. Let {1 , 2 , ...} R be a countable dense subset, e.g. {x + iy R : x, y Q}.
Since {fn (i )}
n=i U is bounded, there is a subsequence {f1,n } {fn } such that f1,n (i )
some number (i ). Likewise, there is {f2,n } {fn } such that f2,n (i ) (i )...
For every m we get {fm,n } such that fm,n (i ) (i ) for all i m. Let n = fn,n , the
diagonal elements, so that {n }n6=m {fm,n }nm . Thus n (m ) (n ) for all m 1.
Claim. {n }
n=1 converges on all of R.
Proof. Let D = D(x0 , 3d) R be some disk, and let K = D(z 0 , d) D(z 0 , 3d) R be a
compact subset of R. Moreover, for all z K, C(z, d) D. Since |n | < 1, z K we have


1 Z
1

()


n
|0n (z)| =
d
.
2i C(z,d) ( z)2 d
= z1 , z2 K,
Z

|n (z1 ) n (z2 )| =

z2

z1

Z

n (z)dz

z2

z1

Z

|n (z)|dz

z2

z1


dz 1
= |z1 z2 |.
d d

Hence {n } is equicontinuous on K in that n, z1 , z2 K, if |z1 z2 | < d, then |n (z1 )


n (z2 )| < .
For z K, n, m 0, k 0,
|n (z) m (z)| |n (z) n (k )| + |n (k ) m (k )| + |m (k ) m (z)|.
If we choose k K such that |z k | < 13 d, and choose N s.t. n, m N , |n (k )
m (k )| < /3, then
|n (z) m (z)| < /3 + /3 + /3 = .
67

Thus {n (z)} is Cauchy, and n (z) (z). Weve proven n pointwise.


Claim. is uniformly continuous on K if |z1 z2 | < d.
|(z1 ) (z2 )| = lim |n (z1 ) n (z2 )| .
Uniform convergence on K? Suppose n . Let  > 0, and Sj := {z K : |n (z)
(z)| <  for all n > j}. Sj is open: indeed, if z1 Sj and |z1 z2 | < d for > 0, then
|n (z1 ) (z2 )| |n (z1 ) n (z1 )| + ... |n (z1 ) (z)| + 2 < /2
for all n > j. Taking small enough, z2 Sj . This implies D(z, ) Sj .

Sj = K

j=1

Compactness implies K + Sj for some j implies uniform convergence. All you do now is
fiddle around with B s.

68

CHAPTER

19
RIEMANN MAPPING THEOREM AND INFINITE
PRODUCTS

We will continue with the Riemann mapping theorem, which we got through most of last
time:

Theorem. (Riemann mapping) Let R C be open and simply connected (for us, use CR
is connected, i.e. can apply closed curve theorem). Also suppose R 6= C. Then ! map f that
is holomorphic, with holomorphic inverse, s.t. f : R D, with f () = 0, R0 3 f 0 () > 0.
Improperly Stated Generalization. (Uniformization theorem) Every simply connected Riemann surface is conformal to, and has a holomorphic map with holomorphic inverse to,
precisely one of
= S2
(1) C
(2) C
(3) D
The nice thing is that these correspond to different geometries. D is the hyperbolic geometry, S 2 youve done on homework, and C is Euclidean. We dont have the tools to do this
properly yet.
Proof of the Riemann mapping theorem so far. Let
F = {f : R D : f is holomorphic, 1-1, and f 0 () > 0}.
We show that
(A) F 6= .
(B) supf F |f 0 ()| < and g F achieving this sup, e.g. |g 0 ()| = supf F |f 0 ()|.
(C) Given this g achieving the sup, g is surjective and g(w) 6= 0 for w R (so that we
get a holomorphic inverse).
We can also draw pictures for A and B, which we showed last time with the help of the
following:
Lemma. (Montels theorem) A uniformly bounded sequence of holomorphic functions
has a subsequence which converges on compact subsets of the domain, and the limit is
69

holomorphic (using Moreras theorem).


For more info here, see also the Arzela-Ascoli theorem.
Lets finish up with C.
Claim. This limit g has the property that g() = 0 and is surjective (any 1-1 function
f can have f 0 (w) = 0 for any w. ).
Proof. g sends 0 7 0. If not, g() = w 6= 0, and Bh g : 7 0. Then |(Bw g)0 ()| =
1
0
0
|Bw (g())g 0 ()| = 1|w|
2 |g ()| > 1, and there is a larger such g.
g is surjective. Suppose not, that that it misses w. Wlog w = 0, and consider (Bw g),
0
which sends 7 w and misses 0. Then |(Bw g)0 ()| = |Bw
(g())||g 0 ()| = (1|w|2 )|g 0 ()|.
Now consider the following claim:
Claim. a holomorphic branch of
Proof. We can write

Bw g (which maps R D)
Z

log(Bw g)(z) log(Bw g)() =

Then take e

1
2 ()

(Bw g)0 (z)


dz = .
(Bw g)(z)

We have ( Bw g) : w 7 w by construction. Then ( Bw g)0 ()| = | 0 (w)|(1

1
2
0
2
0
|w| )|g ()| = 2|w|1/2 (1 |w| )|g ()|. Also, (Bw Bw g) : 7 0, and (Bw

1
1
0
Bw g)0 () = |B
( w)|... = 1|w|
(1 |w|2 )|g 0 ()|.
w
2|w|1/2
Set |w| = r2 , for r R0 (with w not 0 b/c g 0 () > 0). We claim that

 
1
1
(1 r4 ) > 1
2
1r
2r
for 0 < r < 1. To see this, note that
1 + r2 2r = (r 1)2 > 0
for 0 < r < 1, so

1+r 2
2r

> 1.

So g is surjective. Hence g is injective. We claim that g 0 (w) 6= 0 for any w. To see


this, suppose g 0 (w) = 0. Wlog suppose w = 0. Then g 0 (0) = 0 and g(z) = z k h(z) for k 2,
with h holomorphic and h0 (0) 6= 0. Hence
Z
g 0 (z)
dz = Z(g)
C (0) g(z)
inside D (0) = k. Take with || small; then
Z
g 0 (z)
k=
dz = Z(g )
C (0) g(z)
in D (0). If so, g(z) is 0 at inverse images of , but all of the same small ball around 0
is an inverse image of something. So g 0 (0) = 0 and g is constant.

70

Here are some consequences of the Riemann mapping theorem:


(1) Every connected and simply connected region in R2 is topologically equivalent (in
the sense of homeomorphic) to a disk. For example, consider taking out a Cantor set in R2 .
(2) See moduli spaces of curves and their relations to string theory.
Back to infinite products and the gamma function. Lets finish the proof of the
following theorem:
Theorem.



Y
z2
sin(z) = z
1 2 .
k
k=1

Corollary. Set z = 1/2, and


1=





Y
1
Y (2k 1)(2k + 1)
10
=
.
2
(2k)2
2
(2k)(2k)
k=1

Then

k=1

=
2

22
13



44
35



66
57


....

This has a taste of number theory.


Proof of the theorem. The gist of the proof is to write
Q(z) =

sin(z)
k=1 (1

z2
k2 )

,
3/2

which has growth order of at most something like Ae|z| . We conclude that Q(z) is a
constant, which taking limits we see to be = 1. See the textbook for a more thorough proof.

This is presumably the flavor of analytic number theory, where we cant show things
definitely but can use complex analysis to give a (poor) bound that translates into something
nice. Well continue with the gamma function (z) next time.

71

CHAPTER

20

ANALYTIC CONTINUATION OF GAMMA AND ZETA

Analytic continuation of gamma and zeta. So we have these two functions:


Z

X
1
(z) =
et tz1 dt,
(z) =
z
n
0
n=1
The first is holomorphic for Re(z) > 0 and the second for Re(z) > 1; this is just an
application of Moreras theorem. Well talk about the gamma function for a while; what
wed like is for these functions to be holomorphic in all of the complex plane. It turns
out that the properties of the gamma function will be important for understanding prime
numbers. Lets start with gammarecall what happened last time. We have
Z
Z 1
Z
(z) =
et tz1 dt =
et tz1 dt +
et tz1 dt.
0

The right integral uniformly converges on BR (0), i.e. it converges on compact subsets of C.
Now look at the left integral. We saw that it converges uniformly for compact subsets of
Re(z 1) > 1, i.e. Re(z) > 0.
How do we extend this to all of C? The extension will have various poles, and is meromorphic in C.
Claim. (z + 1) = z(z).
Proof. We integrate by parts. Consider
Z
(z) =

et tz1 dt.

Take u = tz1 , dv = et dt = du = (z 1)tz2 dt, v = et . Then


Z
Z

et tz1 dt = et tz1 0 +
(z 1)et tz2 dt = (z 1)(z 1),
0

as desired.

Corollary. (n) = (n 1)! for n Z1 .


72

Proof. Check (1) =

R
0

et dt = 1 from the claim.

From the above, we can see that


(z + 1)
z
is meromorphic (holomorphic) for Re(z + 1) > 0, i.e. Re(z) > 1 with a pole only at z = 0.
Repeating this k times, we see that
(z) =

(z) =

(z + k + 1)
z(z + 1)...(z + k)

is meromorphic for Re(z + k + 1) > 0, i.e. Re(z) > k + 1, with poles only at nonnegative
integers. This gives a continuation of into the left side of C.
The residues at the poles is given by
Res((z); 0) =
Res((z); k) =

(1)
= 1,
1

(1)
(1)k
=
.
(k)(k + 1)...(1)
k!

These poles are indeed simple by our calculations above. We will show another nice claim:
Claim. (z)(1 z) =

sin(z) .

Corollary. has no zeroes.


Proof. This property follows directly from the claim. sin(z) is nonzero for z not an
integer; for z an integer, we already know that has a pole if z Z0 or (z) = (z 1)! 6= 0
if z Z1 .

To prove the initial claim, well need to show another fact, that for
n

t
t
e = lim 1
n
n
we have
Z
n


1

(z) = lim

t
n

n

tz1 dt,

R
e.g. we can switch the lim and . Then
n
Z n
Z n
t
1
lim
1
tz1 dt = lim n
(n t)n tz1 dt,
n 0
n n
n
0
and from integration by parts we have
n Z n z
Z n

tz
t
n z1
n
(n t) t dt = (n t) +
n(n t)n1 dt.
z
z
0
0
0
So

1
n nn

lim

1 n
n nn z

(n t)n tz1 dt = lim

73

Z
0

tz (n t)n1

Z n
1
n(n 1)...1
tz+n1 dt
n n2 z(z + 1)...(z + n 1) 0
n
 z
 

n
nn nz
1
n
tz+n
= lim
=
...
since
.
n
z z+1
z+n
z + n 0
z+n
= lim

This gives
z
1
= lim
(z) n nz

z+1
1


...

z+n
n

n
z
z Y
1
+
,
n nz
k

= lim

k=1

which looks something like the expression for sin(z) we saw before. Lets rewrite the last
expression as
n


z Y
z
z
z  Y h
z  z/k i
z
z
1
+
=
lim
zn
+
+
...
+
1+
e
,
exp
z
+
z
n n
n
k
2 3
n
k

lim

k=1

k=1

and the product by itself converges as shown before. Rewriting once again with nz =
e(log n)z ,


z
z
z  Y h
z  z/k i
lim znz exp z + + + ... +
1+
e
n
2 3
n
k
k=1
"
!#
n
X
Y h
1
z  z/k i
= lim exp z
log n
z
1+
e
.
n
k
k
k=1

k=1

Also,
"
lim

n  
X
1

k=1

#
log n = 0.577,

the Euler-Mascheroni constant. Then


h
Y
z  z/k i
1
= ez z
1+
e
.
(z)
k
k=1

From this we see that






Y
Y
1
1 z z
z2
z2
sin(z)
1
=
=
e
e z(z)
1 2 =z
1 2 =
,
(z)(1 z)
z(z)(z)
z
k
k

k=1

which is a cute equation. It also gives the values


 
 

1
3

= ,
=

(1/2)! =
.
2
2
2

The zeta function. We turn to (z), defined as


(z) =

X
1
,
nz
n=1

74

Re(z) > 1.

k=1

We claim the following:


Claim. (z) has a meromorphic extension to all of C with a simple pole only at 1.
Well show next time that (z) has no zeroes on Re(z) = 1, and from this deduce the prime
number theorem, which relates to the growth rate of the number of prime numbers. You
may have also heard of the Riemann hypothesis, which says that the nontrivial zeros of are
all on the line Re(z) = 1/2; this gives us a stronger estimate of the growth rate of the primes.
Claim. For Re(z) > 1,
(z) =

Y 
p prime

1
1 pz


.

Proof. We first write


1
1
1
1
= 1 + z + 2z + 3z + ...
1 1/pz
p
p
p
k
1
This gives us something of the form 1 + 1/2z + 1/3z + .... If we write n = p
1 ...pk , then

1
1
1
= 1 ... k
nz
p1 z pk z
in the product. Then




N
Y 
X

1
1
1


1 + z + ... + M z


z

n
p
p

n=1
p<N
for some M . Moreover,






 Y 
 Y 

Y



1
1
1
1


.
1 + z + ... + M z




z
z
p
p
p<N
p<N 1 p
p prime 1 p

In the limit, we thus have






X

1
1 Y





z
z


n
1

p


n=1
p prime

For the reverse direction, we have


Y 

1+

p<N

etc. Then

and

Y 
p<N

1
pz

1
1 z
2


(z) =

1
1
+ ... + M z
z
p
p

(z) = 1 +

N
X
1
,
nz
n=1

1
1
+ z + ...
3z
5

X
n has no prime factors<N

75

1
1 as N ,
nz

giving the desired equality.

Now lets extend (z) to all of C (with a pole). With the change of variables s = nt,
we have
Z
Z
sz1 ds
es z1
ent tz1 dt =
= nz (z).
n
n
0
0
Then

1
(z) =
(z)
n=1

nt z1

1
dt =
(z)

!
nt

z1

n=1

1
dt =
(z)

tz1
dt
et 1

for Re(z) > 1. This is defined on the entire complex plane:


Z 1 z1
Z z1 
1
t
t
(z) =
dt +
dt .
t1
(z) 0 et 1
e
1
The right integral is holomorphic by Moreras theorem and the fact that it converges uniformly on compact subsets of C. For the first integral, we break the integrand up. Note
that the function f (z) = ez11 has a simple pole at z = 0 near the real line, and is otherwise
holomorphic. Its Laurent expansion around z = 0 converges on A(0, 2), and uniformly and
absolutely on A(, 2 ):
1
1
= + c0 + c1 t + c2 t2 + ...
et 1
t
Note that c1 = 1. Rewriting the integral with the Laurent expansion, we have
1

Z
0

and we can commute the


simplify:

Z 1
X
tz1
dt
=
tz1+n cn dt,
et 1
n=1 0

and

by absolute convergence of the integrand. Then we can



X
c0
c1
cn
1
+
+
+ ....
=
z
+
n
z

1
z
z
+1
n=1

[Alternatively, the

R1
0

is holomorphic and converges since


Z
0

X
Bm
xs1
dx
=
,
x
e 1
m!(s + m 1)
m=0

where Bm denotes the m-th Bernoulli number defined by

X
x
Bm m
=
x .
ex 1 m=0 m!

Then B0 = 1, and since ezz1 is holomorphic for |z| < 2, we must have lim supm |Bm /m!|1/m =
1/2.] Anyway, we can claim
Claim. This gives a holomorphic function on all of C.
76

P cn
bRc
Claim.
n=1 z+n converges uniformly for z in BR (0) n=1 B (n). The same ar1
gument also shows its meromorphic by excluding the term z+k
and B (k).
The punchline here is that there exists a meromorphic function on C which equals
for Re(z) > 1 and has a pole only at z = 1.

77

1
n=1 nz

CHAPTER

21

ZETA FUNCTION AND PRIME NUMBER THEOREM

Functional equations towards the prime number theorem. Last time we had
Z z1
1
t
(z) =
dt,
Re(z) > 1,
(z) 0 et 1
where

1
1
= + c0 + c1 t + ...
et 1
t

We used this to see that (z) extends to a meromorphic function in all of C with a simple
pole at z = 1 (and this is the only pole). We can see this by breaking the integral down into
1
(z)

Z
0

1
tz1
dt =
et 1
(z)

Z

tz1
dt +
et 1

Z
1


tz1
dt ,
et 1

and noting that the right hand integral is holomorphic by Moreras theorem and uniform
convergence on compact subsets of C, while the left hand integral is also holomorphic and
converges.
Functional equation. Define a theta function (t) as
(t) :=

en t .

n=

This has the property that


(t) = t1/2 (1/t)
for t R>0 , which is on the problem set and uses the Fourier transform.
Theorem.
z/2 (z/2)(z) =

1
2

uz/21 ((u) 1)du =

Z
0

78

uz/21 (u)du,

where

(u) 1 X n2 u
(u) :=
=
e
.
2
n=1

Proof. First we rewrite


1
2

z/21

((u) 1)du =

Note that

!
u

z/21 n2 u

du.

n=1

uz/21 ((u) 1)du

converges nicely because (u) 1 is bounded by the decreasing exponential:





X
X
eu


mu
n2 u
e
e
.
=


1 eu


m=1

n=1

So for u = 1, this is bounded and the integral converges. Near 0, we consider


Z
1 1 z/21
u
((u) 1)du.
2 0
and as u 0, (u) is well-behaved and the integral converges. Take the original integral and
swap the sum and integral (justifying the interchange of summation and integral yourself)
to get
!
Z X

Z
X
2
z/21 n2 u
u
e
du =
uz/21 en u du
0

n=1

n=1

Let t = n2 u, dt = n2 du. Then


Z
Z
X
X
z/21 n2 u
u
e
du =
n=1

n=1

z/21 t

2 z/2

e dt (n )

(z/2) z/2 nz

n=1

= (z/2) z/2 (z),


giving us the desired result.

What do we do with this theorem? Lets define the xi function


(z) = (z/2) z/2 (z).
Theorem. (z) is holomorphic for Re(z) > 1, extends to a meromorphic function on all of
C with poles at 0 and 1, and
(z) = (1 z).

Proof. First we have (u) = u1/2 (1/u). Then


(u) =

u1/2 (1/u) 1
(1/u) 1
1
(u) 1
=
= u1/2
+ 1/2 1/2
2
2
2
2u
79

= u1/2 (1/u) +

1
1/2.
2u1/2

We break the integral up into two pieces:


Z 1
Z
z/21
(z) =
u
(u)du +
0

uz/21 (u)du.

Let t = 1/u, du = 1/u2 du, du = 1/t2 dt. Then



Z 1
Z z/2+1 
t
t1/2
1
z/21
1/2
t (t) +
u
(u)du =

dt
t2
2
2
0
1

Z 
Z
1
tz/21/2 tz/21 dt
tz/21/2 (t)dt +
=
2
1
1


Z

1 tz/2
1 t(z1)/2
+
=
tz/21/2 (t)dt +

2 z1
2 z/2 1
1
2
1
Z
1
1
=
+
tz/21/2 (t)dt
z1 z
1
Z


1
1
+
(u) uz/21 + uz/21/2 du,
= (z) =
z1 z
1
and the whole thing is symmetric through z 7 1 z.

The conclusion is that is symmetric about x = 1/2. There are no zeroes with x > 1,
and by symmetry there are no zeroes with x < 1. Note that (z) has simple poles at
nonnegative integers, while (z/2) has simple poles at 0, 2, 4, .... Then
(z) =

z/2 (z)
(z/2)

has trivial zeroes at 2, 4, ..., and well show that there are no zeroes right of x = 1 (also,
there is neither a pole nor zero at 0).
Claim. (z) has no
Q zeroes in {Re(z) > 1}.
Proof. (z) = p prime (1 1/pz ) 1 for Re(z) > 1.

Theorem. (z) 6= 0 for Re(z) = 1.


Proof. Deferred to below.
Theorem. (Prime Number) Let (N ) = # of primes N . Then
(N )
i.e.

N
,
log N

(N )
1
N/ log N

as

N .

We wont show this theorem today, but we will show that (z) has no zeroes for Re(z) = 1.
80

To do this we need a few other results:


Lemma.
X

log (z) =

mZ,p prime

1
.
mpmz

Proof.
Y 

(z) =

p prime

log (z) =

X
p prime

1
log 1 z
p

1
1 1/pz

X X
(pz )m
=
=
m
m=1
p prime

X
mZ,p prime

1
.
mpmz

Note that this is much like the zeta function:

X X
(pz )m
=
m
m=1

p prime

where

X
mZ,p prime

X
cn
1
=
,
z
mpmz
n
n=1

(
1/m n = pm
cn =
0
otherwise

and cn 0. This is an instance of a Dirichlet series, which we wont talk about in detail.
(Take Math 229x for this.)

Lemma. 3 + 4 cos + cos 2 0.
Proof. Trivial.

Lemma. For z = x + iy, x > 1,


log |(x)3 (x + iy)4 (x + 2iy)| 0.
Proof. We have
log |(x)3 (x + iy)4 (x + 2iy)| = 3 log |(x)| + 4 log |(x + iy)| + log |(x + 2iy)|
= 3Re(log (x)) + 4Re(log (x + iy)) + Re((x + 2iy))
X
=
nx (3 + 4 cos(y log n) + cos(2y log n)) 0,
n

as desired.

Proof of the theorem. We wish to show that (z) 6= 0 for Re(z) = 1. Suppose (x + 2iy) = 0.
Then consider (x)3 (x + iy)4 (x + 2iy) = A. If (1 + iy) = 0, then as x 1+ , A 0, a
contradiction. This is because (z) has a simple pole at z = 1, (z) has zero at z = 1 + iy,
and (z) is holomorphic (no pole) at z = 1 + 2iy.

Chebyshev/Tchebychevs (x) function. (unrelated to (u) earlier) Define
X
X
X  log x 
(x) :=
log p =
(n) =
log p,
log p
pm <x
n<x
p<x
81

for the von Mangoldt function defined as


(
log p
(n) =
0

n = pm
n 6= pm ,

and p a prime.
Proposition. (x) x (x) logx x .
Proof. We will only show the forward direction. Note that
X  log x 
X
(x) =
log p
log x = log x (x).
log p
p<x
p<x
So

(x)
(x) log x

,
x
x

and if (x)/x 1, then


1 = lim inf

(x) log x
.
x

For the opposite inequality, note that


X
X
(x)
log p
log p ((x) (x )) log(x )
x <p<x

p<x

for (0, 1). Hence


(x)/x +
i.e.

So if

(x ) log x
(x) log x

,
x
x

(x) log x
(x) log x
+
.
x
x
x
(x)
x

1,
(x) log(x)
x

1 lim sup
for all (0, 1), so
1 lim sup
and (x)/x 1 =

(x) log x
x

(x) log x
x

1.

82

CHAPTER

22
PRIME NUMBER THEOREM

Prime number theorem. Recall from last time that if (x) is the number of primes x,
then the prime number theorem states that (x) logx x . Well go about proving this today.
A bit of history: while the prime number theorem was proved in the 1800s, an elementary
proof of it was still in the air in the 1940s. Selberg and Erdos actually had a bitter fight over
it: Selberg came up with a formula that implied the Prime Number Theorem but didnt
work out the details until Erd
os had already did, and they disagreed on who got the credit
for it. Eventually Selberg won out, much to Erdoss chagrin. There is an even simpler proof
using basic complex analysis, given by Newman, one of the authors of our textbook.
Anyways, recall that we had
log (z) =

X
p prime,n1

1
npnz

for Re(z) > 1. Since


X
p prime,n2


X  1
1
1
=
+
+
...
npnz
2p2z
3p3z
p prime

is holomorphic where Re(z) > 1/2,


X
p prime

1
= log (z)
pz

X
p prime,n2

1
npnz

is holomorphic where Re(z) P


> 1/2. Note that log(z 1)(z) is entire, holomorphic
on a
P
p
subset of Re(z) 1, so that p prime p1z + log(z 1) is as well, along with p prime plog
+
z
1
.
z1
Let be defined as
X log p
(z) =
.
pz
p prime

Lemma. (z)

1
z1

is holomorphic where Re(z) 1.


83

Proof. See above.


We also consider the function (x) =

px

log p for real x.

Theorem. (x) x = (x)


P x.
P
Proof. We have (x) = px log p px log x = (x) log x.

Chebyshev showed that (x) C x in limit, i.e. lim sup (x)


< c, which we will show
x
later. For the moment, we turn to proving another useful theorem first.
Theorem. Suppose
Z

ezt f (t)dt

g(z) =
0

for f bounded and integrable, and is holomorphic for Re(z) 0. Then


Z
g(0) =
f (t)dt
0

is well-defined.
RT
Proof. Suppose T > 0 and let gT (z) := 0 ezt f (t)dt. Let C be the boundary of the
region {z : |z| R, Re(z) for small = (R) > 0 and large R so that g(z) is holomorphic
on the region and its boundary. Applying Cauchys integral formula, we get


Z
1
z2
Tz
g(0) gT (0) =
(g(z) gT (z))e
1 + 2 (1/z)dz.
2i C
R
Now bound this: on the part of C on the upper half-plane (denote this as C1 ), we can write

Z
Z


BeT Re(z)
tz

|g(z) gT (z)| =
e f (t)dt B
|etz |dt =
,
Re(z)
T
T
2

2Re(z)
z
where B = maxt0 |f (t)| and since |eT z | = eT Re(z) and |1 + R
2| =
R , we see that
R
.
For
the
other
piece
of
the
contour
C
,
we
can
replace
the
integral
for gT (since
| C1 | B
2
R
gT is entire) by Rthe contour of a semicircle in the lower half plane C20 (which, similar to
the above, has | C 0 | B
R ), while the other integral 0 as T by a simple bounding
2
argument. Taking R concludes the proof.


Now we can get a bound for (x) in the flavor of Chebyshev:


Lemma. (x) (4 log 2)x.
Proof. 2n
by definition contains every prime number from n + 1, ..., 2n in its factoriza
Qn
tion; thus n<p2n p 2n
n . But we can also expand
(1 + 1)2n =

  
 
 
2n
2n
2n
2n
+
+ ... +
=
4n .
0
1
2n
n

Thus
Y
n<p2n

p 4n =

X
n<p2n

84

log p n log 4.

If n takes on values from 1, 2, 4, ..., 2M , where M is the highest such that 2M +1 x, we get
X
(x)
< 2M +1 log 4 < 2x log 4 = (4 log 2)x,
1<p2M +1

as desired.

Were ready to finish off the proof of the prime number theorem. If Re(z) > 1, then
Z
X log p
(z) =
ezt (et )dt,
=
z
z
p
0
p
and from the above, we have limx
(z + 1)
=
z+1

(x)
x2

e(z+1)t (et )dt =

= 0. This implies

(z + 1) 1
=
z+1
z

e(z+1)t ((et )et 1)dt.

We can then take g(z) = 0 f (t)ezt dt and f (t) = (et )et 1, so that the lemma we
proved above tells us
Z
Z
(x) x
dx < .
f (t)dt =
x2
0
1
This furthermore implies that (x) x:
Lemma. If g(x) is nondecreasing and
Z
1

g(x) x
dx < ,
x2

then g(x) x.
Proof. Almost trivial.

So weve proved the prime number theorem.

85

CHAPTER

23
ELLIPTIC FUNCTIONS

Elliptic Functions. These are doubly periodic, meromorphic functions on C. We wont


talk too much about the history of the word elliptic. If w1 , w2 are two nonparallel vectors
w1

/ R), then an elliptic function f are periodic in both directions:


in C ( w
2
f (z) = f (z + w1 )

and

f (z) = f (z + w2 ).

These had origins in determining elliptic arclengths. By precomposing with multiplication


2
by a nonzero complex number , we may assume w1 = 1, w2 = = w
w1 : that is, take
g(z) = f m (z), g(z) = f (z). Taking = w1 gives g(z + 1) = f (w1 z + w1 ) = g(z) and
g(z + ) = f (w1 z + w2 ) = f (w1 z) = g(z).
Wlog, we have H, i.e. Im( ) > 0. Then Im( ), Im(1/ ) have opposite signs because
Im( ) 6= 0 or else its collinear with 1.
The objective now is to study these meromorphic functions f on C such that f (z + 1) =
f (z) and f (z + ) = f (z). The fundamental domain for translation by the lattice
= {a + b : a, b Z}
is then given by the parallelogram
P = {a + b : (a, b) [0, 1)2 }.
Now we have some facts about f :
1. Fact. If f is holomorphic, then f is constant.
Proof. f is bounded by its maximum on P .
P
2. Fact.
P Res(f ; ) = 0.
Proof. First, if there are no poles on P , then
Z
X
f (z)dz 2i
Res(f ; ).
P

By taking the integral along P , we put


Z
Z 1
Z 1
Z
f (z)dz =
f (t)dt +
f (1 + t) dt
P

86

Z
f (1 + t)dt

f ( t)dt.
0

The integrals cancel in pairs by f being elliptic, and we are left with 0. If there is a
pole along P , then the claim is that ,  : P +  has no pole where = (1 + ) C
is not collinear with 1 and . To see this, suppose not; then we get an accumulation
of poles at distinct points. Now note that P +  is also a fundamental domain, and
we can reason similarly to the above.
Remark on terminology: The order of an elliptic function is the total order of its poles in
the fundamental domain P .
Corollary. There are no elliptic functions with orders 0 or 1.
Proof. If f has order 0, then it is holomorphic, giving a contradiction. If f has order 1,
then there is one simple pole, and the residue is 6= 0, giving another contradiction.

Proposition. If f is an elliptic function of order k, then TFAE:
(a) f has k total order of zeroes in P .
(b) f (z) c has total order of zeroes equal to k.
Proof. (a) = (b). Use the argument principle. (b) = (a) follows similarly.

Weierstrass functions. These are the order 2 elliptic functions with double poles
at the lattice points (e.g. 0). Heres a first attempt to define such a function:
X
X
1
1
=
.
2
(z + w)
(z + m + n )2
w

m,nZ

The problem here is that this is not absolutely convergent.


Lemma.
Proof.
X
mZ

1
(m,n)6=(0,0) (|m|+|n|)r

< for r > 2.

X
X
1
1
1
1
1
1
1
=
+
2
=
+
2

+
2
dl
r
r
r
r
r
(|m| + |n|)r
|n|r
(|m|
+
|n|)
|n|
l
|n|
l
|n|
m=1
l=|n|+1

=
Sum this over n:

1
|n|r+1
+
2
.
|n|r
r + 1


X 1
1
+
C
|n|r
|n|r1
n

summable for r 1 > 1, i.e. r > 2.


P
1
Corollary.
(m,n)6=(0,0) (m+n )r converges absolutely for any r > 2, H.
Proof. Left as an exercise.

The right way to define such a function is


1
(z) = 2 +
z

X
w{0}

1
1
2
2
(z + w)
w


.

The claim is that this converges uniformly on compact subsets of C to a meromorphic


function with double poles at .
87

Claim. is the function were looking for.


Proof. Let |z| < R. Then
(z) =


X 
1
1
1
+

,
z2
(z + w)2
w2
|w|<2R

which is a finite sum. We add on another term:




X 
X 
1
1
1
1
1
+
2 +
2 .
z2
(z + w)2
w
(z + w)2
w
|w|<2R

The claim is that

|w|>2R

|w|>2R

1
(z+w)2

1
w2

with w converges uniformly to a holomorphic


2

(z+w) w
1
z +2zw
1
function on |z| < R. to see this, note that (z+w)
2 w2 =
(z+w)2 w2 = w2 (z+w)2 , where
the denominator is like w4 and the numerator is like w where |z| is bounded. So the sum
converges uniformly on |z| < R. Thus the whole sum converges uniformly.


Proposition. is elliptic: (z + 1) = (z) and (z + ) = (z).


Proof. Consider 0 (z) on the swiss cheese disk. We have the limit fo holomorphic
functions, so the derivative converges uniformly. Note that
0 (z) = 2

X
w

1
(z + w)2

is doubly periodic. Then we have that (z + 1) = (z) + a where a is a fixed constant


because the derivative is 1-periodic, while (z + ) = (z) + b where b is fixed because the
derivative is -periodic. The fact that a = b = 0 follows from the fact that is even: in
particular, (1/2) = (1/2) and ( /2) = ( /2).

Remark. We note that is an order 2 elliptic function, as it has a double pole at only
the origin in P . Likewise, 0 is an order 3 elliptic function, since it has a triple pole at only
the origin in P .
Theorem. Every elliptic function is a rational function in and 0 .
Proof. First lets understand the relationship between and 0 . We note that 0 is odd,
and 0 (1/2) = 0 (1/2) = 0 (1/2), so that 0 (!/2) = 0 and similarly for /2, (1 + )/2.
These are all the zeroes of 0 since 0 is of order 3, while 0 has a pole of order 3 at the
origin.
For , this means that (z) () = 0 for = 1/2, /2, (1 + )/2, and has a double
root (and nothing else since is of order 2). (z) () = 0 has a simple root at j
distinct, for 6= 0, 1/2, /2, (1 + )/2 P .
Theorem. We have
0 (z)2 = 4((z) e1 )((z) e2 )((z) e3 )
for e1 = (1/2), e2 = ( /2), e3 = ((1 + )/2).
Proof. Look at the zeroes and the poles of each side. 0 (z)2 has a pole of order 6 at 0,z
eroes of order 2 at 1/2, /2, (1 + )/2, and no others. The right hand side has poles of order
88

6 at 0, zeroes of order 2 at the same points, and no others. The function


0 (z)2
4((z) e1 )((z) e2 )((z) e3 )
is holomorphic, doubly periodic, and hence constant. Whats the constant? If we Laurent
expand at 0, we get
2
1
(z) = 2 + ... and 0 (z) = 2 + ...,
z
z
so


1
4((z) e1 )((z) e2 )((z) e3 ) = 4
+
...
z6
and
0 (z)2 =

4
+ ...
z6

so the constant is 1.

with order 2. Note that


Geometric interpretation. is surjective from P to C
2

P/ = T .
A (sketch of a) geometric picture is as follows: consider skewering your donut on a stick
through four points on the lattice; then taking the quotient by flipping over, we get halftorus that is one-to-one except for two points that are two-to-one.
Theorem. Every elliptic function that is doubly periodic under 1 and is a rational
function in , 0 .
Proof. First, we do this for even elliptic functions. Let F denote one such, which has
even order (possibly 0) of zeroes and poles at 0. Hence F (z)(z)m , m Z has no zero or
pole at 0. Let 1 , ..., k denote the zeroes of F with multiplicity; these come in pairs i
if i 6= 1/2, /2, (1 + )/2. Likewise, list the poles 1 , ..., k of F ; the same comment holds.
The claim is that
Q
((z) (i ))
Q
G(z) = i
j ((z) (j ))
has the same zeroes and poles with multiplicity as F , so F = cG.
(z)
, Fodd =
Now for general elliptic F , let F = Feven + Fodd . Then Feven = F (z)+F
z
F (z)F (z)
, and they are both elliptic. Feven is a rational function in , while Fodd /0 is
2
even and elliptic, and thus a rational function in . The result follows.

Note that, for different , P/ are not generally the same torus.
Anyway, this story of elliptic curves has applications to number theory and cryptography,
and algebraic geometry and other areas of math. Next time, well start with the formula
(0 )2 = 40 g2 g3 for some g2 , g3 , and get into other uses and generalizations for elliptic
functions.

89

CHAPTER

24

WEIERSTRASSS ELLIPTIC FUNCTION AND AN


OVERVIEW OF ELLIPTIC INVARIANTS AND MODULI
SPACES

Elliptic functions. There is a relationship between elliptic functions and elliptic integrals.
Consider integrals of the form
Z
Z
dx
dx

or
.
quartic
cubic
These integrals show up when we want to compute the arc length of ellipses. The second is a
Schwarz-Christoffel transformation from H to a quadrilateral. As before, starting from their
double-periodicity and some transformation properties, we want to study the fundamental
function (z).
Today, well start with (z), and consider the equation
(0 (z))2 = 4(z)3 g2 (z) g3 ,
which arises when we consider the integral
Z
z

dz
p

4z 3 g2 z g3

0 (z)

=
(z)

0 (w)dw
p
0 (w)2

and substitute (w) = z; this gives as the inverse of the integral. Note that, under the
elliptic integral given above, the three roots of the cubic in the denominator and go
to the vertices of the quadrilateral in the image, since the elliptic integral is basically a
Schwarz-Christoffel map. One can show here that an elliptic function is the inverse to the
Schwarz-Christoffel map.
Recap on what we know about . First we recall that Weierstrasss elliptic function
is defined for any fixed as

X 
1
1
1
(z) = 2 +

,
z
(z + w)2
w2

90

where = {0} and the lattice for a fixed is = {n + m : m, n Z} C. is


meromorphic with double poles at lattice points. We also had
0 (z) = 2

X
w

X
1
1
=
2
(z + w)3
(z + n + m )3
n,mZ

We showed that if (1/2) = e1 , ( /2) = e2 , and ( 1+


2 ) = e3 , then
0 (z)2 = 4((z) e1 )((z) e2 )((z) e3 ).
Today, well compute 0 (z) = 4(z)3 g2 (z) g3 . We do this by introducing Eisenstein
series:
Eisenstein series. These are functions given by
Ek ( ) =

X 1
=
w2

X
(m,n)6=(0,0)

1
.
(m + n )k

Here are a few facts, which well leave to the reader to prove:
Fact 1: If k 3, then Ek ( ) is a holomorphic function of in H: the sum converges
uniformly in for Im( ) > > 0.
Fact 2: If k is odd, then Ek ( ) = 0.
Fact 3: We have the following transformation relations:
Ek ( + 1) = Ek ( )

and

Ek (1/ ) = k Ek ( )

E.g. for the last relation we have (m + n(1/ ))k = k (m n)k .


These properties tell us that E2k is a weakly modular function of weight 2k. Note that, since
the Eisenstein series converge depending on , we can just write them as constants.
Proposition. If z is near 0, then

(z) =

X
1
1
+
(2k + 1)E2k+2 z 2k = 2 + 3E4 z 2 + 5E6 z 4 + ...
2
z
z
k=1

Proof. We can write



X 
1
1
1
(z) = 2 +
2 ,
z
(z w)2
w

where we reindex the lattice under the transformation w 7 w. Then


1
1
1
= 2
(z w)2
w (1 z/w)2
1
1

1
2
Noting that 1x
= 1 + x + x2 + ... for |x| < 1 and (1x)
2 = x ( 1x ) = 1 + 2x + 3x + ...,
we have for |z/w| < 1 that


 z 2
1
1
z
=
1
+
2
+
3
+
...
.
(z w)2
w2
w
w

91

So

X 1 X
X
1
1
zn
(z) = 2 +
=
+
(n
+
1)
z
w2 n=1
wn
z 2 n=1
w
P
1
Since En+2 = w wn+2
, we have

X
w

1
wn+2

!
(n + 1)z n

(z) =

X
1
+
(2k + 1)E2k+2 z 2k ,
2
z
k=1

as En+2 = 0 when n is odd.

Theorem. If 0 (z)2 = 4(z)3 g2 (z) g3 , then g2 = 60E4 and g3 = 140Eg .


Proof. We can write out some terms of the identity we derived above:
(z) =

1
+ 3E4 z 2 + 5E6 z 4 + ...
z2

2
+ 6E4 z + 20E6 z 3 + ...
z3
4
1
0 (z)2 = 4(z)3 + ... = 6 24E4 2 80E6 + ...
z
z
1
1
(z)3 = 6 + 9E4 2 + 15E6 + ....
z
z
Now take the difference
0 (z) =

0 (z)2 4(z)3 + 60E4 (z) + 140E6 .


This is holomorphic near 0; each term is holomorphic away from and is doubly-periodic,
so the difference is holomorphic, doubly-periodic, and takes 0 to 0. Then, from earlier
problems, we know that its 0, and the equality
0 (z)2 = 4(z)3 g2 (z) g3
gives g2 = 60E4 and g3 = 140Eg as desired.

What is this explicit formula telling us? On the one hand, we have C/, the domain
of and 0 which we identify with a torus. Consider a map to C2 given by
z 7 ((z), 0 (z));
we can draw C2 (which we dont really discuss in class) by drawing R R on the axes
in 2-space, and the curve it traces out is the set of solutions to the polynomial function
y 2 = 4x3 g2 x g3 , a cubic curve. What youre seeing is that this complex torus is a cubic
curve in C2 . We can check injectivity and surjectivity and other nice things, but lets get
out of this story portion and move into....
The space of elliptic curves. Now we want to consider all possible s, and we note
that the lattice doesnt change under the action of P SL2 Z Aut(H) = P SL2 R. This is
just the space of transformations of the form
7

a + b
c + d
92

for a, b, c, d Z, generated by 7 + 1 and 7 1/ ; its what you can do to to not


change your lattice.
(The elliptic functions compose with dilations; e.g. theyre maps from coplex tori from
one direction to the other direction, just by dilation. You cant just stretch it in one
direction, since then it wont be holomorphic.)
So let me show you a fundamental domain D in H. This is the part of H above the circle
|z| = 1 and bounded by the lines Re(z) = {1/2, 1/2}; e.g. D = {z H : |z| > 1, 21 <
Re(z) < 12 }. This is the moduli space of complex 1-tori (elliptic curves). In other words,
D is the fundamental domain of the action of P SL2 Z. H/P SL2 Z is the moduli space of
elliptic curves.
Lets do some justice to why these things are nice. E2k ( ) is sort of a function on
H/P SL2 Z, defined on D, and transforms under the involution 7 1/ :
E2k (1/ ) = 2k E2k ( ).
When we say these things are called modular forms, we mean that they dont change the
correct way, and the reason here is the power coefficient. We can, however, write
lim
Im( )

E2k ( ) =

lim
Im( )

X
(m,n)6=(0,0)

X
1
1
=
2
= 2(2k),
2k
(m + n )2k
m
m=1

since for n 6= 0 the center experssion has the n term 0 for Im( ) . We also have
E4 () = 2(2k).
Then

4
2
) = 4 .
90
3
8 6
g3 () = 140E6 () = 140(2)((6)) =
.
27
Setting the modular discriminant = g23 27g32 , we note that () = 0; the right hand
side expression, g23 27g32 , is the discriminant of 4x3 g2 x g3 . This is a weight 12 modular
form, by which we mean that E2k is a transformation of weight 2k, and its modular if
limIm( ) E2k ( ) exists and holomorphic at , E2k ( ) is holomorphic on H, and E2k ( +
1) = E2k ( ). We have another modular form of weight 12, which is simply g23 . Define
g2 () = 60E4 () = 60(2)(

j=

1728g23
.

This has weight 0, so its really defined as a function on H/P SL2 Z (e.g. a meromorphic
function from H C invariant under the action of SL2 Z), which is the moduli space of all
elliptic curves. Its called a j-invariant of elliptic curves. Theres a theorem that states:
Theorem. This is holomorphic on H, invertible under P SL2 Z, has j() = , with
residue 1 at , and gives a surjection H C that injects from the quotient H/P SL2 Z.
Proof. See Serres Course in Arithmetic.

I wont say too much more about this subject; weve got j as an invariant in the moduli space, but we wont talk about why number theorists and cryptographers are interested
in it. For this, consult any reference on elliptic curves. Well use the remainder of the time
for questions.
93

APPENDIX

SOME THINGS TO REMEMBER FOR THE MIDTERM

Complex expressions:
1. Remember polar form: z = rei .
u
v
2. Stereographic projection is given by : S 2 {p} R2 , : (u, v, w) 7 ( 1w
, 1w
), 1 :
2

2y
x +y 1
2x
(x, y) 7 ( x2 +y
2 +1 , x2 +y 2 +1 , x2 +y 2 +1 ).
iz

iz

iz

3. cos z = e +e
, sin z = e e
2
2i
4. log z = log r + i + i(2n).

iz

Holomorphic functions:
1. To show that a function f is holomorphic, use the Cauchy-Riemann equations and
show that the partials are continuous. You can also any rule from real analysis (the
chain rule, the product rule, the quotient rule, etc.), show that f has a power series
development, or use Moreras theorem.
2. Recall that the Cauchy-Riemann equations are
(
v
u
= y
x
v
u
x = y .
R
3. Moreras theorem: If f (z) is continuous on a region D and f (z)dz = 0 for a
triangle (or closed curve), then f (z) is holomorphic. This is useful for:
1. Showing limn fn = f is holomorphic.
2. Showing some functions defined by sums or integrals are holomorphic.
3. Showing functions that are continuous and holomorphic on all but some set (e.g.
some point) are holomorphic.
4. f is holomorphic f complex analytic f smooth; this means you can substitute
any holomorphic function with its power series, and can differentiate it endlessly.
5. The closed curve theorem tells us that the integral of a function that is holomorphic
in the open disk D over a contour C D is 0.
94

6. Per the antiderivative theorem, if f (z) is holomorphic on an open disk D, then it


has an antiderivative F (z) such that F 0 (z) = f (z).
7. Cauchys integral formula allows us to determine the value of a holomorphic function f at a point. In particular,
Z
1
f (z)
f (w) =
dz,
2i C z w
where f is holomorphic in an open disk D and C contains w.
8. The uniqueness theorem tells us that holomorphic functions are determined uniquely
up to their values in a certain region.
9. The mean value theorem allows us to evaluate a holomorphic function at point. If
f (z) is holomorphic on D (R), then r : 0 < r < R, we have
Z 2
1
f ( + rei )d.
f () =
2 0
10. The maximum modulus principle: if f is holomorphic in any region, then its
maximum is achieved at the boundary.
11. Harmonic functions are characterized by the fact that their Laplacian is zero, and
possess nice properties as you discovered in the problem sets.
Entire and meromorphic functions:
1. Entire functions are functions that are holomorphic in all of C.
2. Liouvilles theorem tells us that any bounded entire function is constant. A generalization is that if |f (z)| A + B|z|N and f (z) is entire, then f (z) is a polynomial of
degree N .
3. Likewise, if f (z) is entire and f (z) as z , then f (z) is a polynomial.
4. A nonconstant polynomial always has a root, per the fundamental theorem of algebra.
5. Remember the different types of singularities and why theyre useful, namely poles of
order k and essential singularities.
In other words, they have a
6. Meromorphic functions are holomorphic functions on C.
finite set of removable singularities, and can always be expressed as as a ratio of two
holomorphic functions.
7. A weak formulation of Picards little theorem tells us that if f is entire, then its
image is dense.
Evaluating integrals:
1. There are generally two ways were learned to do nontrivial integrals:
1. Use the closed curve theorem to give a contour integral as zero, then decompose
the contour and bound the individual parts to find a specific integral.
2. Use the residue theorem by first calculating residues:
95

(0) Compute the integral

R
C ()

f (w)dw

(1) Compute the Laurent series for f centered at , and take c1


(2) If f has a simple pole, then c1 = limz (z )f (z)
(3) If f (z) = A(z)/B(z) for B having a simple pole and A 6= 0, then c1 = Res(f ; ) =
A()/B 0 ().
The value of the integral over a curve containing these poles is 2i (sum of residues).
Series:
1. Know the definitions of radius of convergence, lim|an |1/n = L, and convergence
tests/arguments for power series. These may be things that range from proving that
a power series of a holomorphic function is uniformly convergent to finding the radii
of convergence of some functions.
2. Laurent series are power series that extend in both and directions, and are
typically defined on annuli (why?). To compute Laurent series, you can generally
1. Start with known power series and make the appropriate substitutions/adjustments.
2. Use known geometric series. Being able to decompose a function into its partial
fraction expansion is a good trick to know here.
Others:
1. Know the topological aspects: polygonally connected (path connected), polygonally
simply connected (simply connected), what is meant by a region, etc. Note any open
path connected set is simply connected.
R
R
PP PR
2. Know when to interchange
,
, , lim , etc. To deal with interchange of
sums, show absolute convergence. To deal with interchange of limits, show uniform
convergence. (This is covered in real analysis, and also in baby Rudin.)
3. A helpful tool is that you can bound the absolute value of any integral over CR by
2R times the integrand; this is known as the ML theorem or the estimation lemma.

96

APPENDIX

B
ON THE FOURIER TRANSFORM

This is an addendum to the class notes on Fourier transforms, and is based on Chapter 4 of
Stein and Shakarchis Complex Analysis. It is important to note that a function f : R C
must satisfy appropriate regularity and decay conditions to possess a Fourier transform. If
it does (well make this notion precise below), then the Fourier transform is defined by

f() =

f (x)e2ix dx,

R.

(B.1)

f()e2ix d,

x R.

(B.2)

The Fourier inversion formula is then


Z

f (x) =

There is a deep and natural connection between complex analysis and the Fourier transform.
For example, given a function f initially defined on the real line, the possibility of extending
it to a holomorphic function is closely related to the rapid decay at infinity of its Fourier
transform f.
2

Proposition. The functions f (x) = ex and f (x) =


forms.
Proof. Show this yourself.

1
cosh x

are their own Fourier trans

Definition. For > 0, denote by F the class of all functions that satisfy the following two conditions:
(i) f is holomorphic in the horizontal strip S = {z C : |Im(z) < a}.
(ii) There is a constant A > 0 :
|f (x + iy)|

A
1 + |x|1+

for any  > 0 and all x R and |y| < a. This condition is known as moderate decay. As
we will soon see, F can be regarded as the class of functions with well-defined Fourier
97

transforms.
Intuitively, one can think of F as those holomorphic functions of S that are of moderate decay on each horizontal line Im(z) = y, uniformly in a < y < a. Some functions
that belong to F include
2
f (z) = ez
for all , and
f (z) =

1
c
2
c + z2

for 0 < a < c.


Proposition. If f F , then f (n) F for 0 < < .
Proof. Use the Cauchy integral formula.

Proposition. If f F for some > 0, then


|f()| Be2||
for some B and any 0 < .
Proof. Use contour integration.

Lets drop the in F to make notation simpler. The previous proposition tells us that if
f F, then f has rapid decay at infinity. The Fourier inversion formula then works in the
following sense:
Proposition. If f F, then the Fourier inversion formula holds, and we have
Z
f (x) =
f()e2ix d

for all x R.

Properties of the Fourier transform.


Now that we have some basis for knowing when a Fourier transform or inversion is welldefined, lets talk about some of the things we do with them.
Theorem. (Basic properties) Let f, g F. Then the following are true:
= af + b
(1) If h(x) = af (x) + bg(x) for a, b C, then h
g.
0
0
0

(2) If h(x) = f (x x ) for x R, then f () = e2ix f().


0

(3) If h(x) = e2ix f (x) for 0 R, then h()


= f(  0
).
1

(4) If h(x) = f (ax) for a R {0}, then h() = f


.
|a|

(5) If h(x) = f (x), then h()


= f().
(6) Let F denote the Fourier transform. F 2 (f )(x) = f (x) and F 4 (f ) = f .
Proof. These are all easy to verify.

98

Theorem. (Relation to differentiation)



F

d
dx

m


 n
d
m
((ix) f ) = (i)
f,
d
n

where F denotes the Fourier transform. In particular, the Fourier transform of (ix)f (x)
d
is d
f (), and the Fourier transform of f 0 is (i)f().
Proof. Use differentiation under the integral sign and integration by parts.
Theorem. (Relation to convolution) If f g denotes the convolution of f and g defined by
Z
(f g)() :=
f ()g( t)dt,

then f[
g() = f g.
This latter theorem is important for functional analysis, but we wont discuss it to make this
note self-contained. We can also prove an important tool known as the Poisson summation
formula:
Theorem. (Poisson summation formula) If f F, then
X
X
f (n) =
f(n).
nZ

(B.3)

nZ

Proof. Let f F and choose a : 0 < < . The function

1
e2iz 1

has simple poles

f (z)
e2iz 1

1
2i

with residue
at the integers. So
has simple poles at the integers n, with residues
f (n)/2i. Applying the residue formula to the contour N below,

we get
X
|n|N

Z
f (n) =
N

f (z)
dz.
1

e2iz

Letting
N , and recalling the moderate decay of f , we see that the sum converges to
P
f
(n),
and that the integral over the vertical segments goes to 0. So in the limit we
nZ
get
Z
Z
X
f (z)
f (z)
f (n) =
dz

dz,
(B.4)
2iz 1
2iz 1
e
e
L1
L2
nZ

99

where L1 and L2 are the real line shifted down and up by b, respectively. Now we use the
fact that if |w| > 1, then

X
1
= w1
wn
w1
n=0
to see that on L1 , where |e2iz | > 1, we have
1
e2iz 1
If |w| < 1, then

1
w1

= e2iz

e2inz .

n=0

n=0

w , so on L2 we have
1
e2iz 1

e2inz .

n=0

Substituting these equalities into equation (4) above, we see that


!
!
Z
Z

X
X
X
2iz
2inz
2inz
f (n) =
f (z) e
e
dz +
f (z)
e
dz
nZ

Z
X
n=0

L1

f (z)e2i(n+1)z dz +

L1

L2

n=0
Z
X
n=0

f (z)e2inz dz =

L2

n=0

f(n + 1) +

n=0

X
n=0

where we shift L1 and L2 back to the real line by replacing x 7 x ib.

f(n) =

f(n),

nZ

The Poisson summation formula gives many useful identities in complex analysis and beyond (namely, number theory). For example, the Fourier transform of the function f (x) =
2
2
et(x+a) is f() = t1/2 e /t e2ia . Applying the Poisson summation formula to this
pair gives the following identity:

X
n=

et(n+a) =

t1/2 en

/t 2ina

n=

This identity
P can be 2used to prove the following transformation law for the theta function
(t) := n= en t :
(t) = t1/2 (1/t),
for t > 0. It can also be used to prove the key functional equation of the Riemann zeta
function, which gives its analytic continuation:
 s 
(1 s)(1 s)
(s) = 2s s1 sin
2
Finally, Fourier transforms can tell us much about a function. For example, the following
theorem describes the nature of those functions whose Fourier transforms are supported;
in particular, it relates the decay properties of a function with the holomorphicity of its
Fourier transform:
Theorem. (Paley-Wiener) Suppose f is continuous and of moderate decrease on R. Then
f has an extension to the complex plane that is entire with |f (z)| Ae2M |z| for some
A > 0 if and only if f is supported in the interval [M, M ].
100

Applications of the Fourier transform.


One can think of the Fourier transform as a mapping that converts a function f (t) in the
time domain (t has units of seconds) to a function f() in the frequency domain ( has
units of hertz or something).
Fourier transforms can be used to solve differential equations (c.f. below), and is clearly
widely used in physics, probability, engineering, and many other areas. For example, one
usually proves the central limit theorem in probability theory using Fourier transforms, and
can prove a mathematical formulation of the Heisenberg uncertainty principle. For all the
Fourier analysis you can get, Stein and Shakarchis Fourier Analysis is a good standard
textbook (and one you can find online...).

Examples involving the Fourier transform.


Here is a flavor of some of the problems one can ask about Fourier transforms.
2

Example 1. This example generalizes some of the properties of ex related to


that it is its own Fourier transform. Suppose f (z) is an entire function that
R
2
2
|f (x + iy)| ceax +by for some a, b, c > 0. Let f() = f (x)e2ix dx.
0 2
0 2
is an entire function of that satisfies |f( + i)| c0 ea +b for some a0 , b0 , c0

the fact
satisfies
Then f
> 0.

d
with the
Proof. Note that f is clearly an entire function since we can just commute dz
2ix

integral, and f (x)e


is entire. Thus f does not have any poles, and in particular we
can change the contour from the real axis to the line x iy for some y > 0 fixed and
< x < without affecting the value of the integral. Assume > 0, and observe that
Z
Z
Z
f() =
f (x)e2ix dx =
f (x iy)e2i(xiy) dx =
f (x iy)e2(xi+y) dx.

Then note
Z
Z


2(xi+y)

f (x iy)e
dx

ax2 +by 2 2(xi+y)

ce

ce

ax2 2xi+by 2 2y

Z

dx =



dx .

Since |eax 2xi | 0 as |x| , we see that f() = O(eby 2y ). Setting y = d where
2
2
2
d is a small constant, we have that f() = O(eb(d) 2(d) ) = O(e(bd 2d) ), as desired.
Also observe that
Z
Z




2ix(+i)
2(xi+y)(+i)




|f ( + i)| =
f (x)e
dx =
f (x iy)e
dx

Z

=

Z

f (x iy)e2(xix+y+yi) dx

ceax

+by 2 2(xix+y+yi)



dx

2
2
Repeating the above argument, we see that f() = O(e(by 2y)+(ax +2x) ). Now let y =
2
2
2
2
d and x = d0 for small d, d0 in the bound to see that |f(+i)| = O(e(bd 2d) +(ad0 +2d0 ) ).
0 2
0 2
So |f( + i)| c0 ea +b for a0 = 2d bd2 , b0 = 2d0 ad20 (which are > 0 for small
d, d0 ), and some suitable scale factor c0 > 0.


101

Example 2. The problem is to solve the differential equation


an

dn
dn1
u(t) + an1 n1 u(t) + ... + a0 u(t) = f (t),
n
dt
dt

where a0 , a1 , ..., an are complex constants, and f is a given function. Here we suppose that
f has bounded support and is smooth.
Proof. We do this inRsteps:

(a) Let f(z) = f (t)e2izt dt. Observe that f is an entire function, and using
A
integration by parts show that |f(x + iy)| 1+x
2 if |y| a for any fixed a 0.
n
(b) Write P (z) = an (2iz) + an1 (2iz)n1 + ... + a0 . Find a real number c so that
P (z) does not vanish on the line L = {z : z = x + ic, x R}.
(c) Set
Z 2izt
e
u(t) =
f(z)dz.
L P (z)
Check that

n
X

d
aj ( )j u(t) =
dt
j=0

and

e2izt f(z)dz =

e2izt f(z)dz

e2ixt f(x)dx.

We then conclude by the Fourier inversion theorem that


that the solution u depends on the choice of c.

Pn

j=0

d j
) u(t) = f (t). Note
aj ( dt

d
(a) f is an entire function because we can commute dz
with the integral and the integrand is also an entire function. We know that f C 2 , so now we use integration by parts:
1
let u = f (t), du = f 0 (t)dt, dv = e2izt dt, and v = 2iz
e2izt so that


Z
Z 0
f (t) 2izt

f (t) 2izt
f (t)e2izt dt =
e
]t= +
e
dt
|f(z)| =
2iz

2iz

Now let u = f 0 (t), du = f 00 (t)dt, dv =

e2izt
2iz dt,

1
2izt
and v = (2iz)
so that
2e




Z
f (t) 2izt

f 0 (t) 2izt
f 00 (t) 2izt
.
|f(z)| =
e
]t= +
e
]
+
e
dt
t=

2
2
2iz
(2iz)
(2iz)
Since f is entire, we can (as in the previous exercise) change the contour of integration
since f does not have any poles. So we change the contour from t to t is, as in the
argument on page 115 of the text. Rewriting, we have



f 0 (t is)

f (t is) 2i(x+iy)(tis)
2i(x+iy)(tis)




e
]t= +
e
]t=
|f (x + iy)|
2i(x + iy)
(2i(x + iy))2
Z



f 00 (t + is)
2i(x+iy)(tis)
e
dt
+
.
2
(2i(x + iy))
102

If f 0 and f 00 also decay, we can bound this again:








M1
M0
2s(x+iy)
2s(x+iy)



+
e
|f (x + iy)|
e


2
2i(x + iy)
(2i(x + iy))
Z



M2
2s(x+iy)

+
e
dt
.
2
(2i(x + iy))
Note that the first two terms are bounded, and examining the integral we see that we
can extract the z12 term. Since y is also bounded, we have that |f(x + iy)| = O( x12 ), and
A
can therefore conclude that |f(x + iy)| 1+x
2 for a suitable scale factor A.
(b) P is a polynomial of degree n, so it only has finitely many roots and we can choose
such a c so that the line L does not pass through a root. So we can choose c1 = 1, and let
cn = cn1 + 1 if P vanishes with the choice of cn1 . Then c will be the cn for which P does
not vanish, and itll surely be found after n iterations.
We can find some constraints on c if we notice that we want |P (x + ic)| > 0 and can
bound |P (x + ic)| as follows:

n1
n1
X
X
|P (x + ic)| |an ||x + ic|n
|aj ||x + ic|j |an ||x + ic|n
|aj | |x + ic|n1
j=1

j=1

for |x + ic| > 1 (we can repeat the same procedure with a different bound of |x + ic| instead
of |x + ic|n1 in the rightmost term if |x + ic| < 1, and again solve accordingly). Since we
Pn1
Pn1 |a |
want |an ||x + ic|n ( j=1 |aj |)|x + ic|n1 > 0, we see that |x + ic| > j=1 |anj | . We also
Pn1 |a |
know that |x + ic| |c|, so if |x + ic| > 1 we can choose c > max(1, j=1 |anj | ) to satisfy the
inequality and were done. In particular, one c for which the polynomial does not vanish is
Pn1 |a |
c = max(1, j=1 |anj | ) + 1.
(c) We differentiate under the integral sign to see that
n
X
j=0

Z
= an
L

aj (

X
d j
d
) u(t) =
aj ( )j
dt
dt
j=0

(2iz)n

Z
L

X
e2izt
f (z)dz =
aj
P (z)
j=0

Z
(
L

d j e2izt
)
f (z)dz
dt P (z)

Z
Z
e2izt
e2izt
e2izt
f (z)dz+an1
f (z)dz+...+a0 (2iz)0
f (z)dz
(2iz)n1
P (z)
P (z)
P (z)
L
L
Z
Z
P (z)e2izt
=
f (z)dz =
e2izt f(z)dz,
P (z)
L
L

as desired.
Furthermore, we claim that the equality
Z
Z
e2izt f(z)dz =
L

e2ixt f(x)dx

follows by observing that, as in the previous problem, e2izt f(z) is clearly an entire function
d
and thus its integral is also entire by commuting dz
. So the integral does not have any
poles, and in particular we can change the contour of integration from the line L = {z :
z = x + ic, x R} (where c is chosen so that P (z) does not vanish) to the real axis without
affecting the value of the integral.
103

Finally, because were applying the Fourier inversion to the Fourier transform f of f
and both functions satisfy
conditions
R in the hypothesis, we get from the Fourier
Pn the decay
d j
) u(t) = e2ixt f(x)dx = f (t), and we are done. Thus
inversion theorem that j=0 aj ( dt
the solution to the differential equation
an

dn1
dn
u(t)
+
a
u(t) + ... + a0 u(t) = f (t),
n1
dtn
dtn1

is
Z
u(t) =
L

e2izt
f (z)dz,
P (z)

for P (z) = an (2iz)n + an1 (2iz)n1 + ... + a0 and c R so that P (z) does not vanish on
the line L = {z : z = x + ic, x R}.


104

APPENDIX

C
REFERENCES

We cite many sources in these notes. The main course textbook is:
J. Bak and D. J. Newman, Undergraduate Texts in Mathematics: Complex Analysis. Springer,
2010.
One that is oftentimes referenced (and is a gem of a textbook), is
E. M. Stein and R. Shakarchi, Princeton Lectures in Analysis 2: Complex Analysis. Princeton UP, 2003.
And of course, the classic textbook is none other than Ahlfors:
L. V. Ahlfors, Complex Analysis: An Introduction to the Theory of Analytic Functions
of One Complex Variable. McGraw-Hill, 1979.

105

AFTERWORD

You survived! If you enjoyed what you saw in Math 113, you should consider Math 213a
(graduate complex analysis) or Math 229x (analytic number theory). Math 113 provides
adequate (dare I say ample) background for both in most years. 213a deals with complex
analysis, but on a graduate level: youll revisit some of the topological basics, and learn
some nice and beautiful mathematics like infinite product expansions and more on elliptic
and meromorphic functions. Math 229x is number theory using complex analysis: youll
begin with things like the prime number theorem and move on quickly. These courses can
vary a bit, so do take our recommendation with caution.
If you arent pursuing further work in mathematics, we hope that you enjoyed the course
and remember how nice complex analysis is.
Thanks for being a great class, in terms of problem sets and everything else. Keep in
touch with Andy or any of us CAs if youd like, and have a great summer!

Felix and Anirudha

106

Anda mungkin juga menyukai