Vous êtes sur la page 1sur 94

Lectures on Complex Analysis

M. Pollicott

Contents
1. Introduction
1.1. A little history
1.2. Notation
1.3. Some useful reference books
2. A few basic ideas
2.1. The Riemann sphere
2.2. M
obius maps
2.3. Two applications
2.4. Classification of M
obius maps and their behaviour
2.5. cross-ratios of quadruples of complex numbers
2.6. Problems
3. Analyticity
3.1. Ingredients for the first definition (using power series)
3.2. Ingredients for the second definition (using complex differentiability)
3.3. Ingredients for the third definition (using the Cauchy-Riemann
Equations)
4. Definitions of analyticity
4.1. The main result
5. Integrals and Cauchys Theorem
5.1. Integration on piecewise continuous curves
5.2. parameterization of the curve of integration
5.3. Theorems of Gousart and Cauchy
5.4. Immediate applications of Cauchys Theorem
5.5. Converse to Cauchys Theorem
5.6. Problems
6. Properties of analytic functions
6.1. Logarithms and roots
6.2. Location of zeros: Argument Principle and Rouches Theorem
6.3. Liouvilles Theorem and the Fundamental Theorem of Algebra
6.4. Identity Theorem
6.5. Families of functions and Montels Theorem
6.6. Problems
6.7. More problems
7. Riemann Mapping Theorem
7.1. Riemann Mapping Theorem
7.2. Christoffel-Schwarz Theorem
7.3. examples
7.4. Generalizations of the Schwarz-Christofel theorem
7.5. Application: Fluid dynamics
7.6. Problems
7.7. More problems
8. Behaviour of analytic functions
8.1. The Maximum Modulus Principle
8.2. The Schwarz Lemma
3

5
5
6
7
7
7
9
12
15
18
20
21
22
22
23
24
24
28
28
28
32
36
38
39
40
40
41
45
46
47
49
50
51
51
53
57
60
60
60
62
62
63
63

CONTENTS

8.3. Picks lemma


9. Harmonic functions
9.1. Harmonic conjugates
9.2. Poisson Integral formula
9.3. Mean Value Property and Maximum Principle for harmonic functions
9.4. Schwarz reflection principle for harmonic functions
9.5. Application: Continuous extensions to the boundary of analytic
bijections
9.6. Problems
10. Singularities
10.1. Removable singularities
10.2. Analogues of Cauchys theorem
10.3. Laurents Theorem
10.4. Classification of singularities
10.5. Behaviour at singularities
11. Entire functions, their order and their zeros
11.1. Growth of entire functions
11.2. Factorization of entire functions
12. Prime number theorem
12.1. Riemann zeta function
12.2. A related complex function
12.3. A related counting function
12.4. Finiteness of integrals
12.5. Proof of the Prime Number Theorem
13. Further Topics
13.1. Weierstrasss elliptic function
13.2. Picards Little theorem
13.3. Runges theorem
13.4. Conformality: Property of analytic maps
13.5. Functions of several complex variables

64
65
66
67
71
72
73
74
75
75
76
77
80
80
82
82
84
86
86
87
87
88
90
91
92
92
92
92
93

1. INTRODUCTION

1. Introduction
Complex analysis is one of the classical branches in mathematics with roots
in the 19th century and just prior. Complex analysis, in particular the theory of
conformal mappings, has many physical applications and is also used throughout
analytic number theory. In modern times, it has become very popular through a
new boost from complex dynamics and the pictures of fractals produced by iterating
holomorphic functions. Another important application of complex analysis is in
string theory which studies conformal invariants in quantum field theory.
1.1. A little history. The study complex numbers arose from try to find
solutions to polynomial equations. Al-Khwarizmi (780-850) in his book Algebra
had solutions to quadratic equations of various types. Under the caliph Al-Mamun
(reigned 813-833) Al-Khwarizmi became a member of the House of Wisdom in
Bagdad,
The first to solve the polynomial equation x3 + px = q was Scipione del
Ferro (1465-1526). On his deathbed, del Ferro confided the formula to his pupil
Antonio Maria Fiore, who subsequently challenged another mathematician Nicola
Tartaglia Fontana (1500-1557) to a mathematical contest on solving cubics. (The
name Tartaglia means stammerer a symptom of injuries acquired aged 12 during the french attack on his home town of Bresca). The night before the contest,
Tartaglia rediscovered the formula and won the contest. Tartaglia in turn told the
formula (but not the proof) to an influential mathematician Gerolamo Cardano
(1501-1576), provided he signed an oath to secrecy. However, from a knowledge of
the formula, Cardano was able to reconstruct the proof. Later, Cardano learned
that del Ferro, not Tartaglia, had originally solved the problem and then, feeling
under no further obligation towards Tartaglia, proceeded to publish the result in
his Ars
Magna (1545). Cardano was also the first to introduce complex numbers
a + b into algebra, but had misgivings about it. In the Ars magna he observed,
for example, that the problem
of finding two
numbers that add to 10 and multiply
to 40 was satisfied by 5 + 15 and 5 15 but regarded the solution as both
absurd and useless. Cardan was also said to have correctly predicted the exact date
of his own death (but it has also been claimed that he achieved this by committing
suicide!).
Rene Descartes (1596-1650), the mathematician and philospher coined the term
imaginary: For any equation one can imagine as many roots [as its degree would
suggest], but in many cases no quantity exists which corresponds to what one imagines.
Abraham de Moivre (1667-1754), a protestant, left France to seek religious
refuge in London at eighteen years of age. There he befriended Isaac Newton. In
1698 he mentions that Newton knew, as early as 1676 of an equivalent expression
to what is today known as de Moivres theorem (and is probably one of the best
known formulae) which states that:
(cos() + i sin())n = cos(n) + i sin(n)
where n is an integer. (De Moivre, like Cardan, is famed for predicting the day of
his own death. He found that he was sleeping 15 minutes longer each night and
summing the arithmetic progression, calculated that he would die on the day that
he slept for 24 hours.)

Leonhard Euler (1707-1783) introduced the notation i = 1 in his book


Introductio in analysin innitorum in 1748, and visualized complex numbers as points
with rectangular coordinates, but did not give a satisfactory foundation for complex
numbers. In contrast, there are indications that Carl Friedrich Gauss (1777-1855).
had been in possession of the geometric representation of complex numbers since

CONTENTS

1796, but it went unpublished until 1831, when he submitted his ideas to the Royal
Society of Gottingen. It was Gauss who introduced the term complex number.
Joseph-Louis Lagrange (1736-1813) showed that a function is analytic if it has
a power-series expansion However, it is Augustin-Louis Cauchy (1789-1857) who really initiated the modern theory of complex functions in an 1814 memoir submitted
to the French Academie des Sciences.

Although the term analytic function was not mentioned in his memoir, the concept is present there. The memoir was eventually published in 1825. In particular,
contour integrals appear in this memoir (although Poisson had written a 1820 paper with a path not on the real line). Cauchy also gave proofs of the Fundamental
Theorem of Algebra (1799, 1815) which, as we will see, has analytic proof . In
summary, Cauchy, gave the foundation for most of the modern ideas in the field,
including:
(1)
(2)
(3)
(4)
(5)

integration along paths and contours (1814);


calculus of residues (1826);
integration formulae (1831);
Power series expansions (1831); and
applications to evaluation of denite integrals of real functions

The Cauchy- Riemann equations (actually dating back to dAlembert 1752, then
Euler 1757, dAlembert 1761, Euler 1775, Lagrange 1781) are also usually attributed
to Cauchy 1814-1831 and Riemann 1851.
Cauchy resigned from his academic positions in France in 1830 rather than to
swear an oath of allegence to the new government. However, he felt able to resume
his career in France in 1848, when the oath was finally abolshed.
Regarding subsequent work, Karl Weierstrauss (1815-1897) formulated analyticity in terms of existence of a complex derivative, which is the perspective taken
in most textbooks, and Georg Riemann (1826-1866) made fundamental use of the
notion of conformality (previously studied by Euler and Gauss). Later contributions were made by Poincare to conformal maps and Teichm
uller to quaai-conformal
maps.
1.2. Notation. We denote by C the complex numbers. If z C then we can
write z = x + iy where x, y R. We denote the real and imaginary
p parts by
x = Re(z) and y = Im(z). We then write the absolute value as |z| = x2 + y 2 .
Occasionally it is useful to write complex numbers in radial coordinates, i.e.,
z = rei where r > 0 and 0 < 2.
We denote by z = x iy = rei the complex conjugate. If z, w C then we
write
[z, w] = {z + (1 )w : 0 1}
for the line segment joining them.

2. A FEW BASIC IDEAS

Exercise 1.1. Show that the map : C M2 (R) (= 2 2 real matrices)




x y
: x + iy 7
.
y x
is a monomorphism.
1.3. Some useful reference books.
(1) R. Churchill and J. Brown, Complex Variables and Applications (ISBN
0-07-010905-2). This is a fairly readable account including much of the
material in the course.
(2) I. Stewart and D. Tall, Complex Analysis (ISBN 0-52-128763-4). This is
a popular and accessible book.
(3) L. Alhfors, Complex Analysis: an Introduction to the Theory of Analytic
Functions of One Complex Variable (ISBN 0-07-000657-1). This is a classic
textbook, which contains much more material than included in the course
and the treatment is fairly advanced.
(4) S. Krantz and R. Greene, Function Theory of One Complex Variable
(ISBN 0-82-183962-4). This is a nice textbook, which contains much more
material than included in the course.
(5) S. Krantz, Complex Analysis: The Geometric Viewpoint (0-88-385035-4).
The first chapter gives a nice summary of some of the ideas in the course.
The rest of the book is very interesting, but too geometric for this course.
2. A few basic ideas
2.1. The Riemann sphere. It is convenient to add an extra point to C. In
order to accommodate the extra point , we need to extend the complex plane by
adding this point in to get the Riemann sphere.

b = C {} the Riemann sphere. There is a natural stereoWe denote by C


graphic projection between the sphere (minus the north pole (1, 0, 0)) and the
complex plane
: {(x1 , x2 , x3 ) R3 : x2 + x22 + x23 = 1} {(1, 0, 0)} C
defined by

(x1 , x2 , x3 ) = z :=

x1
1 x3

 

x2
+i
.
1 x3

In particular, 0 it the image of the south pole (0, 0, 1), and the unit circle |z| = 1
is the image of the equator x3 = 0.
x22

Definition 2.1. A circle on the sphere corresponds to the intersection of x21 +


+ x23 = 1 with a plane ax + by + cz = d.

CONTENTS

The following is easy to prove.


Lemma 2.2. The stereographic projection of a circle on the sphere is either a
circle or a line in C.
Proof. The image of the intersection under the projection can be written as
a(z + z) ib(z z) + c(|z|2 1) = d(|z|2 + 1)
If we write z = x + iy then
(d c)(x2 + y 2 ) 2ax 2by + (d + c) = 0.
Case I: If c = d then this is the equation of a straight line.
Case II: If c 6= d then
x2 + y 2

2by
(d + c)
2ax

+
=0
dc dc
dc

which we can rearrange as



2 
2
a
b
a2 + b2 + (c2 d2 )
.
x
+ y
=
dc
dc
(d c)2
It only remains to show that a2 + b2 + c2 d2 > 0 to see this is the equation of a
circle. However,
p
p
|0 | = |ax + by + cz| x2 + y 2 + z 2 a2 + b2 + c2
|
{z
}
=1

by the usual Cauchy-Schrwartz inequality and we are done.

Note that in the proof there is an equality in the last line only when (a, b, c) =
(x1 , x2 , x3 ) for some , i.e., the plane is tangent to the sphere.
Remark 2.3. The inverse images of a points z C is a triple (x1 , x2 , x3 ) lying
on the sphere and satisfying |z|2 = (x21 + x22 )/(1 x3 )2 = (1 + x3 )/(1 x3 ) (since
x21 + x22 + x23 = 1). We can then write
z+z
2
x1 =
since 1 x3 = 2
.
1 + |z 2 |
|z| + 1|
Similarly,
zz
|z|2 1
x2 =
and x3 =
.
2
1 + |z |
1 + |z 2 |
We deduce that


z+z
z z |z|2 1
(x1 , x2 , x3 ) =
,
,
.
1 + |z 2 | 1 + |z 2 | 1 + |z 2 |
This completes the proof.
Exercise 2.4. Prove the converse, i.e., the preimage of a circle or a straight
line in C is a circle on the Riemann sphere.

2. A FEW BASIC IDEAS

b we take the usual


Remark 2.5. To associate the appropriate topology to C
open sets in C plus the complements of compact sets union with the point . This
means that we can interpret zn z in the usual sense if z 6= . However, we say
that zn + if for every K > 0 we have there exists N > 0 such that |z| > K.
Then C is homeomorphic to the ball minus the north pole. The approrpiate
b comes form the standard metric on the ball, i.e.,
metric for C
q
d((x1 , x2 , x3 ), ((x01 , x02 , x03 )) = (x1 x01 )2 + (x2 x02 )2 + (x3 x03 )2
q
= 2(1 x1 x01 + x2 x02 + x3 x03 ).
If z := ((x1 , x2 , x3 )) and w := ((x01 , x02 , x03 )) then we can define a natural metric
b {} by
on C
d(z, w) = d((x1 , x2 , x3 ), ((x01 , x02 , x03 )) = p

|z w|
(1 + |z|2 )(1 + |w|2 )

where (x1 , x2 , x3 ), (x01 , x02 , x03 ) 6= (0, 0, 1) (i.e., z, w 6= ) and


d(z, ) := d((x1 , x2 , x3 ), (0, 0, 1)) = p

2
1 + |z|2

Exercise 2.6. Check all the above formulae.


2.2. M
obius maps. We can consider a, b, c, d C with ad 6= bc,.
b { d } C by f (z) =
Definition 2.7. We define a map f : C
c
b = {}, a single point).
ad = bc then f (C)

az+b
cz+d .

(If

Remark 2.8. We can also assume without loss of generality that ad bc = 1,


since we see that replacing a, b, c, d to a, b, c, d gives the same map. We will
adopt this convention when convenient.
Observe that if c = 0 then f () = , i.e., is a fixed point. On the other
hand, if c 6= 0 then letting z 7 d/c we see that f (d/c) = . Letting z +
we can write that f () = a/c.
Lemma 2.9. If c 6= 0 then the M
obius map f is invertible, and its inverse is
b C
b is a homeomorphism).
also a M
obius map (and thus f : C
dzb
1
Proof. If f (z) = az+b
(z) = cz+a
.
cz+d with ad bc = 1 then we can define f
1
1
We can then explicitly check f f (z) = z and f f (z) = z. For example,


dzb
a cz+a
+b

f (f 1 (z)) = 
dzb
c cz+a
+d

a (dz b) + b (cz + a)
c (dz b) + d (cz + a)
(ad bc) z
=
=z
(ad bc)


Lemma 2.10. If f1 and f2 are M
obius maps then so is the composition f1 f2
Proof. This follows immediately by substitution. (Exercise)
In particular, the set of M
obius maps forms a group.

10

CONTENTS

Example 2.11. There are four fundamental examples of Mobius transformab C.


b
tions f : C
1.z+b
a) Translations: z 7 z + b = 0.z+1
where b C. These are often called
parabolic transformations.

a.z+0
1 where |a| = 1. These are often called
b) Rotations: z 7 az = 0.z+
a
elliptic transformations.

.z+0
1 with > 1 (or
c) Expansions (and Contractions): z 7 z =
0.z+
0 < < 1) and real. These are often called hyperbolic transformations.
d) Inversions: z 7 z1 .
We can use this to deduce the following.
Lemma 2.12. Every M
obius map can be written as a composition of M
obius
maps of the above type.
Proof. Every Mobius map is a combination of three types of maps f (z) = Az,
f (z) = z + B and f (z) = 1/z, where A, B C, since we can write


ad
a
b ad
az + b
a
c (cz + d) + b c
c
=
= +
cz + d
cz + d
c
cz + d
which is a composition of there
z 7 cz 7 cz + d 7

1
7
cz + d


b

ad
c



1
cz + d


7




ad
1
a
b
+
c
cz + d
c


The follow result is very useful. Consider circles in the complex plane of the
form C := {z C : |z z0 | = r} where z0 C and r > 0.
Theorem 2.13. The image of a circle or straight line under a M
obius transformation is a circle or straight line.
Proof. It suffices to consider three different types of Mobius transfomations:
(1) If f (z) = Az where A C with A 6= 0 then the image f (C) is a circle.
(2) If f (z) = z + B where B C then the image f (C) is a circle.
(3) If f (z) = 1/z then the image f (C) is a circle. We first observe if z = x+iy
lies on a circle centred at z0 = x0 + iy0 of radius r > 0 then
(x x0 )2 + (y y0 )2 = r2

(1)

Consider the set of z such that


|z p|
=k
|z q|

(2)

where p, q C and k > 0. If p = u + iv and q = s + it then this becomes:



(x u)2 + (y v)2 = k 2 (x s)2 + (y t)2
(3)
In particular, we can rewrite (1) in terms of (3) by choosing p, q, k such
that
(u k 2 s)
1 k2
v k2 t
y0 =
1 k2

 

u2 + v 2 k 2 (s2 + t2 )
u sk 2
v tk 2
2
r =
+
+
> 0.
1 k2
1 k2
1 k2

x0 =

2. A FEW BASIC IDEAS

Finally we can see that if z satisfies (2) then




1 1
z p
q
|z p|
= k
= k

|z q|
p
z1 1q

11

(4)

i.e., a version of (3) (with p replaced by p1 , q replaced by 1q and k replaced


by pq k). In particlar, the image of the circle given by (2) is a circle given
by (3).

b and distinct w1 , w2 , w3 C
b there
Theorem 2.14. For distinct z1 , z2 , z3 C
b
b
exists a unique M
obius map f : C C such that f (zi ) = wi , for i = 1, 2, 3.
Proof. We first prove existence and then uniqueness.
Existence. Consider the case that z1 , z2 , z3 6= . Let
S(z) =

(z z2 )(z1 z3 )
(z z3 )(z1 z2 )

then we see that S(z1 ) = 1, S(z2 ) = 0, S(z3 ) = . Consider


T (z) =

(z w2 )(z1 w3 )
(z w3 )(z1 w2 )

then we see that T (w1 ) = 1, T (w2 ) = 0, T (w3 ) = . If we define f (z) := S 1 T (z)


then we can then observe that f (zi ) := T 1 S(zi ) = wi (i = 1, 2, 3).
zz2
In the case that z1 = then we let S(z) = zz
and similarly for T .
3
z1 z3
In the case that z2 = then we let S(z) = zz3 and similarly for T .
2
In the case that z3 = then we let S(z) = zzz
and similarly for T .
1 z2
Uniqueness. We can assume without loss of generality that w1 = 1, w2 = 0, w3 =
b C
b taking
. (Otherwise we can additionally compose with a Mobius map g; C
w1 , w2 , w3 to 1, 0, , respectively, and then we can apply the following argument
to show that f1 g 1 = f2 g 1 , which therefore gives us f1 = f2 ). Assume that
b C
b are two M
fj : C
obius maps j = 1, 2 such that fj (1) = z1 , fj (0) = z2 ,
fj () = z3 . Since S(z) = f11 f2 (z) is a Mobius transformation we can write
S(z) =

az + b
cz + d

bC
b fixes the three points 1, 0, . In particular, we see that
Moreover, f11 f2 : C
S(0) = b/d = 0 = b = 0
S() = a/c = = c = 0, and
S(1) = (a + b)/(c + d) = a/d = 1
from which we deduce that S(z) = z for all z, i.e., f1 = f2 .

Let us consider a could of examples of this result.


Example 2.15. Find the M
obius transformation f which maps 1, 0, 1 to the
points i, 1, i.
b
az+b
Assume that f (z) = az+b
cz+d . Since f (0) = d = 1 we have b = d and f (z) = cz+b .
ia+b
Similarly, since f (1) = ia+b
ic+b = 1 = ic ib = a + b and f (1) = ic+b = i =
ic + ib = a + b. Adding the last two equations gives c = ib and subtracting gives
a = ib. Thus
ibz + b
iz + 1
iz
f (z) =
=
=
.
ibz + b
iz + 1
i+z
(Formally, we should also multiply the coefficients by constant to get ad bc = 1)

12

CONTENTS

Example 2.16. Let D = {z C : |z| < 1} be the open unit disk unit and let
H = {z C : Im(z) > 0} be the upper half-plane. Find a surjective M
obius map
f : H D.
Actually, the map in the previous example works (as one might guess from the
three points in the boundary of H being mapped to three points in the boundary
of D). To see this, let z = x + iy with y > 0 then




2
2
i (x + iy) 2 x + i(1 y)) 2


= x + (1 y) < 1,

=
|f (z)| =
x + i(y + 1)
i + (x + iy)
x2 + (1 + y)2
i.e., f (z) D.
Example 2.17. Fix a C and then we can define a Mobius map f :b
C C by
f (z) =

az
1 az

Notice that
|f (z)|2 =

|a z|2
|a|2 2Re(az) + |z|2
=
2
|a az|
1 2Re(az) + |az|2

It is easy to check that if |z| = 1 if and only if |f (z)| = 1. Moreover, if |f (z)| < 1 is
equivalent to
|a|2 + |z|2 < 1 + |az|2 < 1 (1 |a|2 )(1 |z|2 ) > 0 |z| < 1
since |a| < 1. We can conclude that f : D D.
2.3. Two applications.
Application 2.18 (Apollonion Circle Packings). This is related to taking three
circles in the plane.
Theorem 2.19 (Apollonius). Given three mutually tangent circles C1 , C2 , C3
there are precisely two circles C, C 0 tangent to all three.
Proof. We can write this in a number of steps.

Step 1: Let be the point of intersection C1 C2 of two of the circles. We can


1
apply the M
obius transformation f (z) = z
then f () = and C1 {} and
C2 {} are mapped to disjoint (and thus parallel) straight lines L1 and L2 .
Since C1 and C3 are tangent at one point , say, and C2 and C3 are tangent at
one point , say, then the image C4 = f (C3 ) is another circle tangent to L1 and L2 .

2. A FEW BASIC IDEAS

13

Step 2: We can trivially translate C4 between the two parallel lines L1 , L2 to


precisely two more circles C5 and C6 of the same radius such that: C5 and C6 are
both tangent to both L1 and L2 and C3 .

Step 3: We then define C = f 1 (C5 ) and C 0 = f 1 (C6 ). Since again Mobius


transformations take circles to circles (where straight lines are understood as circles
passing through infinity) we deduce that C, C 0 are the only two such circles tangent
to C1 , C2 , C3 .

Theorem 2.20. Descartes Theorem: Given four mutually tangent circles C1 ,
C2 , C3 and C4 whose radii are r1 , r2 , r3 , r4 then F (r1 , r2 , r3 , r4 ) = 0 where
F (r1 , r2 , r3 , r4 ) = 2(r12 + r22 + r32 + r42 ) (r11 + r21 + r31 + r41 )2

(1)

Proof. We begin with two simple estimates. Consider a circle E of radius k.


(1) We can reflect in it a second circle C of radius r whose centres are separated
by d > r + k. In particular, the circles are disjoint and have disjoint
interiors. The image circle f (E 0 ) with have radius k 2 r/(d2 r2 ).
(2) Consider next a straight line L at a distance b > k from the centre of E.
We can reflect it in the circle E and the image is a circle of radius k 2 /2b.

14

CONTENTS

Step 1. Let C1 , C2 , C3 and C4 be four mutually tangent circles. Choose E to be


a (large) circle centred at the tangency point = (x0 , y0 ) between C1 and C2 .
We can invert the four circles in E by f , say, to arrive at a configuration (after
some rotation and translation) with: L1 = f (C1 ) a line given by y = 1; L2 = f (C2 )
a line given by y = 1; and C30 = f (C3 ) and C40 = f (C4 ) being circles of radius
1 with centres (1, 0) and (1, 0), respectively, i.e., the straight lines are circles
with radius r1 = r2 = and the other two circles have radius r3 = r4 = 1. In
particular, we observe that it satisfies (1).

Step 2. We can apply the two estimates to the lines L1 , L2 and the circles C30 , C40
to obtain
k2
k2
k2
k2
,
r(C
)
=
,
r(C
)
=
r(C3 ) = 2
and
r(C
)
=
4
2
1
x0 2x0 + y02
x20 + 2x0 + y02
2(y0 1)
2(y0 + 1)
The result follows by substituting.

Proceeding inductively we can construct circle packings.


Exercise 2.21. If a M
obius map f (z) = az+b
cz+d maps real numbers to real numbers then show that a, b, c, d are all real numbers.
Show that f : H H is a bijection.
Application 2.22 (Hyperbolic Half-plane). The upper half plane H has the
Poincare metric written as (dx2 + dy 2 )/y 2 , i.e., it is similar to the usual Euclidean

2. A FEW BASIC IDEAS

15

metric, except at z = x + iy the distance is scaled by 1/y. We claim that this is


preserved under M
obius maps.
We want to show that any path has the same length as its image f If we let
f (z) = az+b
cz+d then we can write
f 0 (z) =

ad bc
1
c(az + b)
a
=
=

cz + d (cz + d)2
(cz + d)2
(cz + d)2

and

Im(g(z)) = Im

az + b
cz + d


= Im

(az + b)(cz + d)
|cz + d|2


=

1
Im(z)
|cz + d|2

By the chaine rule


Z
Z
|(g 0 (t)|
|(g 0 ( 0 (t))|| 0 (t)|
length(g) =
dt =
dt
Img (t)
Img (t)
Z
1
1
| 0 (t)||c(t) + d|2
dt
=
|c(t) + d|2
Im(t)
Z
| 0 (t)|
=
dt = length()
Im(t)
Alternatively, if z1 , z2 H then can show that




1 z1 z2
d(z1 , z2 ) = 2 tanh
z1 z2
and show that this is preserved by Mobius maps, i.e.,


az1 + b az2 + b


|g(z1 ) g(z2 )| =

cz1 + d cz2 + d


(az1 + b)(cz2 + d) (az2 + b)(cz1 + d)


=

(cz1 + d)(cz2 + d)


(ad bc)(z1 z2 )

=
(cz1 + d)(cz2 + d)
Thus



|z1 z2 | (cz1 + d)(cz2 + d)
=
|z1 z2 | (cz1 + d)(cz2 + d)
|g(z1 ) g(z2 )|
|
{z
}

|g(z1 ) g(z2 )|

=1

2.4. Classification of M
obius maps and their behaviour. We would like
to understand better the behavious different types of Mobius maps We begin with
a simple result.
Lemma 2.23. Mobius maps other than the identity must fix at least one point
and at most two.
2
Proof. Assume that f (z) = az+b
cz+d = z then this is equivalent to cz + (d
a)z b = 0. The solution to this quadratic equation is
p
z = ((a d) (a d)2 + 4bc)/2c C.

Unless f is the identity (a = c and b = d = 0) this has two roots (possibly a single
root repeated if (a d)2 + 4bc = 0).

We would like to classify M
obius maps up to conjugacy.
Definition 2.24. We say that two Mobius maps f1 , f2 are conjugate if f2 =
g 1 f1 g for some M
obius map g.
Definition 2.25. We can associate to the Mobius map f (z) =
ad bc = 1) the value tr(f ) := a + d.

az+b
cz+d

(satisfying

16

CONTENTS

Lemma 2.26. The value tr(f ) is preserved by conjugacy, i.e., if f1 , f2 are conjugate then tr(f1 ) = tr(f2 )
Proof. Since every M
obius map g can be written as a composition of basic
M
obius transformations (translation, inversion and rotation) it suffices to show the
result where g is of each type.
Translations: if g(z) = z + then we can explicitly write
a(z ) + b
+
c(z ) + d
a(z ) + b + (c(z ) + d))
=
c(z ) + d
z(a + c) + (b + d a 2 c)
=
cz + (d c)

f2 (z) = g 1 f1 g(z) =

and we see that tr(f2 ) = (a + c) + (d c) = a + d = tr(f1 ).


Inversions: If g(z) = 1/z then
a/z + b
c/z + d
a + bz
=
c + dz

f2 (z) = g 1 f1 g(z) =

and we see that tr(f2 ) = d + a = tr(f1 ).


Rotations: If g(z) = z then
f2 (z) = g



1 az + b
f1 g(z) =
cz + d
az + b/
=
cz + d

and we see that tr(f2 ) = a + d = tr(f1 ).

Definition 2.27. We say that a rational map is parabolic it it has precisely


b
one fixed point in C
In particular, being parabolic is easily seen to be a conjugacy invariant (i.e.,
if f1 has a single fixed point z0 , say, then f2 has a single fixed point g 1 z0 . In
particular, if f1 has a fixed point f1 (z0 ) = z0 we can choose g so that g() = z0
and then the conjugate map f := f2 fixes .
Lemma 2.28. Let f be a M
obius map with f () = .
(1) Then f is of the form f (z) = z + ;
(2) f has a second distinct fixed point (i.e., it is not parabolic) if and only if
6= 1;
(3) if the second fixed point is 0 (i.e., f (0) = 0) then f (z) = z.
Proof. (1) Since f () = this immediately implies that f2 (z) = z + , for
some , C.
(2) If f has a second point z0 , say, (other than ) then f (z0 ) = z0 + = z0 , i.e,

with 6= 1. In particular, if 6= 1 then f has a second fixed point (in


z0 = 1
addition to ). Conversely, if = 1 then f has no other fixed points except
since f (z) = z + is a straightforward translation. (In the degenerate case = 0
this is just the identity.)
(3) Since f (0) = 0 we see that = 0 and the result follows.


2. A FEW BASIC IDEAS

17

Remark 2.29. In fact M


obius transformations f1 , f2 are conjugate if and only
if tr(f1 )2 = tr(f2 )2
The non-identity M
obius transformations are commonly classified into three
types:
(1) parabolic (conjugate to z 7 z + with a single fixed point );
(2) elliptic (conjugate to z 7 z with || = 1 with points 0, ); and
(3) loxodromic (everythingelse).
The hyperbolic transformations are those conjugate to the expanding and contracting maps z 7 z (with 0 < < 1 or 1 < < +) being a subclass of the
loxodromic ones
bC
b be a non-trivial M
Lemma 2.30. Let f : C
obius transformation. It is
(1) a parabolic M
obius transformation if and only if tr(f ) = 2;
(2) an elliptic M
obius transformation if and only if 2 < tr(f ) < 2;
(3) a hyperbolic M
obius transformation if and only if tr(f ) > 2.
Proof. Assume that f () = (else replace by a conjugate map with this
property).
Case I: f is parabolic . By definition, in this case f has no other fixed points
and we see from the lemma see that f (z) = z + . This always corresponds to
tr(g) = 1 + 1 = 2 (or 2 since we recall that multiplying a, b, c, d by 1 gives the
same map).
Case II: f is not parabolic. In this case, f has a second fixed point. Assume
that f (0) = 0 (else we can replace f by a conjugate map with this property)
and
then by the lemma we can write f (z) = z. Then we require a = d1 = (to
satisfy ad bc = ad = 1).
We can now consider the different values of the trace tr(f ) = a + a1 .
(1) If tr(f ) = 2 then a + 1/a = 2 and so a(a2 2a + 1) = (a 1)2 = 0. But
a = 1 which means that f (z) = z, the identity map, which is excluded
by the non-triviality assumption.
The proofs of parts (2) and (3) are slightly jumbled up.
(a) More generally, if a = rei then
1
tr(f ) = a +
a
ei
(1)
= rei +
r
= (r + 1/r) cos + i(r 1/r) sin
If we assume that the trace in (1) is real (i.e., Im(tr(f )) = (r
1/r) sin = 0) then this is equivalent to having either r = 1/r(= 1)
or = 0, or both. In the first case, if |a| = r = 1 then f is elliptic,
and tr(f ) = 2 cos (2, 2). In the second case, if = 0 or then
f is hyperbolic and tr(f ) = r + 1/r 2.
(b) If we assume that the trace in (1) has a non-zero imaginary part
(i.e., tr(f ) 6 R) then this is equivalent to requiring both r 6= 1/r and
6= 0, . But then since |a| = r 6= 1 this means that the map is
not elliptic (as well as not being parabolic). Thus by definition it is
loxodromic.

Remark 2.31. A loxodromic Mobius map is a composition of a hyperbolic and
elliptic map. In addition, a loxodromic Mobius map has that tr(f ) C [2, 2].
Exercise 2.32. Show that if f is hyperbolic then f n (z) converges to one of the
fixed points.

18

CONTENTS

2.5. cross-ratios of quadruples of complex numbers. We begin with the


definition of a very elegant expression.
b we
Definition 2.33. Given four distinct complex numbers z0 , z1 , z2 , z3 C
define the cross ratio by
(z0 z2 )(z1 z3 )
(z0 , z1 , z2 , z3 ) =
.
(z0 z3 )(z1 z2 )
Recall that we have already proved a theorem to the effect that can find a
b
(unique) M
obius map f (z) = az+b
cz+d that maps any three distinct points z1 , z2 , z3 C,
say, to the reference points 1, 0, in the same order.
b be four distinct points. We have that
Lemma 2.34. Let z0 , z1 , z2 , z3 C
(z0 , z1 , z2 , z3 ) = f (z0 )
i.e., the image of z0 under the unique M
obius map which takes z1 , z2 , z3 to 1, 0, .
The following is trivial.
Example 2.35. (z, 1, 0, ) = z since the associated Mobius map is simply
f (z) = z.
The next lemma shows that the cross ratio is preserved by Mobius maps.
bC
b is (another) M
Lemma 2.36. If g : C
obius map then
(z0 , z1 , z2 , z3 ) = (gz0 , gz1 , gz2 , gz3 ).
Proof. Exercise.

The M
obius transformation has an interesting geometric interpretation:
b then let C denote the
Theorem 2.37. If z1 , z2 , z3 are three distinct points in C
unique circle(or line, if one of the point is equal to ) passing through them. Then
(z0 , z1 , z2 , z3 ) R if and only if z0 C.
Proof. Let f (z) = (z, z1 , z2 , z3 ) be the associated Mobius map, then we want
to show that f 1 (R {}) = C. Suppose w f 1 (R {}) or, equivalently, that
f (w) R {}.
Case 1 (f (w) = ): In this case since
z z2 z1 z3
f (z) =
z z3 z1 z2
we deduce that this is equivalent w = z3 C.
Case 2 (f (w) 6= ): In this case w 6= z3 then since by assumption f (w) R we
see that f (w) = f (w) and thus
aw + b
aw + b
=
(ac ac)|w|2 + (ad bc)w + (bc ad)w + (dd db) = 0.
cw + d
cw + d
This brings us to two subcases.
Case 2a (ac = ac): In this case the equation is of the form
Aw Aw + ik = 0
where k is real and A = ad bc. Rewriting this as
iAw iAw + k = 0
this is the equation of straight line (cf. the usual form Bw + Bw + C = 0).
Case 2b (ac 6= ac): If ac = ac then we can write
|w|2 +

(ad bc)
(bc ad)
dd db
w+
w=
= 0.
ac ac
ac ac
ac ac

2. A FEW BASIC IDEAS

19

Completing the square we can write



2

2




w (ad bc) = |w|2 + (ad bc) w + (bc ad) w + (ad bc)



ac ac
ac ac
ac ac
ac ac

2
dd db (ad bc)
=
(= R2 ).
+
ac ac ac ac
In particular, this is seen to be the equation of a circle of radius R provided the
right hand side is positive. We can rewrite this as
dd db (ad bc) (da ca)
|a|2 |d|2 + |b|2 |c|2 abcd abcd
+
=
ac ac
ac ac ca ca
|ac ac|
Thus it suffices to observe that we can write the numerator as
(ad bc)(ad bc) = |ad bc|2 > 0
using that ad bc 6= 0.
In either case we see that f 1 (R {}) is a circle or straight line

As an immediate corollary we have another presentation of the proof that


M
obius transformations preserve circles (although the proof is essentially the same
as before).
Corollary 2.38. If f is a M
obius transformation then it maps circles (or
lines) to circles (or lines).
Proof. Let C be a circle (or line) determined by three distinct points z1 , z2 , z3
C. Then
z0 C (z0 , z1 , z2 , z3 ) R
(f (z0 ), f (z1 ), f (z2 ), f (z3 )) R
(f (z0 ), f (z1 ), f (z2 ), f (z3 )) f (C) =: C 0
where C 0 is a circle (or line). In particular, the circle (or line) for (z1 , z2 , z3 ) is
mapped to the circle (f (z1 ), f (z2 ), f (z3 )).

Finally, the following is an interesting geometric interpretation of cross ratios.
Remark 2.39 (Cross ratios and hyperbolic geometry). We let
H3 = {(z, t) : z = x + iy C, t > 0}
denote the three dimensional upper half-space with metric
dx2 + dy 2 + dt2
.
t2
We can consider a geodesic : R H3 with end points
ds2 =

z1 = lim (t) := () and z2 = lim (t) := (+)


t

More precisely, for any two distinct t1 < t2 the curve [t1 , t2 ] 3 t 7 (t) is the
shortest path between (t1 ) and (t2 ) in H2 . It appears a semi-circular arc in H3 .
Given z3 , z4 C we can consider the geodesic 0 with end points z3 = 0 ()
and z4 = 0 (+). We denote the Hausdorff distance between the curves and
by
d(, 0 ) = inf d((t1 ), 0 (t2 )).
t1 ,t2 R

The connection between the cross ratio of the four complex numbers z1 , z2 , z3 , z4
(i.e., the end points) and the geometry is the following.
Lemma 2.40 (Fenchel and Ahlfors). |(z1 , z2 , z3 , z4 )| = tanh(d(, 0 ))

20

CONTENTS

The argument of (z1 , z2 , z3 , z4 ) also has a geometric interpretation. It corresponds to the change in the angle by parallel transporting a frame along the
geodesic realizing this distance.
3. Analyticity
We want to recall the definition(s) of analyticity of functions f : U C.

Definition 3.1. Assume that U C is an open set, i.e., for every z0 U there
exists  > 0 such that
B(z0 , ) = {z C : |z z0 | < } U.
It is usually also convenient to additionally assume that:
(1) U is path connected (i.e., for any two points z, w U we can find a
continuous path : [0, 1] U such that (0) = z and (1) = w).

(2) U is simply connected (i.e., any closed path : [0, 1] U with (0) = (1)
can be contracted to a single point or, equivalently, U is homeomorphic
to the unit disk).

We will assume that U always has these properties, unless we explicitly state
otherwise, and frequently call it a domain.
We will present three equivalent definitions once we have introduced three new
lots of notation.
3.1. Ingredients for the first definition (using power series). We want
to start by recalling the more intuitive definition of analytic functions.

3. ANALYTICITY

21

Definition 3.2. Let z0 CPand let (an )


n=0 be a sequence of complex numbers.

We say that an infinite series n=0 an (z z0 )n has a radius of convergence 0


R + (at z0 ) if
1
1/n
= lim sup |an |
.
(1)
R
n+
In particular, we have the folowing:
Lemma 3.3. Assume that R > 0. For any 0 < r < R the series

X
fz0 (z) :=
an (z z0 )n
n=0

converges (uniformly) on B(z0 , r) and defines a function.


3.2. Ingredients for the second definition (using complex differentiability). We begin by what it means for a function to be differentiable as a complex
function.
Definition 3.4. Let U be a domain. We say that f : U C is complex
differentiable at z0 if the limit
f (z) f (z0 )
f 0 (z0 ) := lim
zz0
z z0
exists, i.e., there exists f 0 (z0 ) C such that for all  > 0 there exists > 0 such
that if |z z0 | < then
|f (z) f (z0 ) (z z0 )f 0 (z0 )| |z z0 |.
Remark 3.5. The key point is that z can approach z0 in all directions as a
complex number.
It is easy to see the following result.
Lemma 3.6. If f is complex differentiable at z0 then it is continuous at z0 .
Proof. Given  > 0 there exists > 0 such that whenever |z z0 | < then
we have that


0

f (z0 ) f (z) f (z0 ) < .


z z0
In particular, for |z z0 | < we can bound
|f (z) f (z0 )| (|f 0 (z0 )| + )|z z0 | (|f 0 (z0 )| + ),
which can be made arbitrarily small by choosing > 0 appropriately small.

3.3. Ingredients for the third definition (using the Cauchy-Riemann


Equations). We can write a complex function f : U C in terms of its real and
imaginary parts
f (x + iy) = u(x, y) + iv(x, y)
where z = x + iy U and u(x, y), v(x, y) R.

22

CONTENTS

Definition 3.7. We say that f : U C satisfies the Cauchy-Riemann equations if for each z0 = x0 + iy0 U the partial derivatives
u
u
v
v
(x0 , y0 ),
(x0 , y0 ),
(x0 , y0 ) and
(x0 , y0 )
x
y
x
y
all exist and satisfy
(

u
x (x0 , y0 )
u
y (x0 , y0 )

v
= y
(x0 , y0 )
v
= x (x0 , y0 )

and

for each z0 = x0 + iy0 U . Equivalently, we can write f (z) = u(x, y) + iv(x, y) and




f
f
u
v
u
v
and
i
=i
+i
=
+i
.
x
x
x
y
y
y
{z
}
|
v
i y
+ u
y

using the Cauchy-Riemann equations. Thus we see that


i

f
f
=
x
y

Remark 3.8. We can write this as


f
1
=
z
2

f
f
+i
x
y

f
z


and

= 0 using the Wirtinger derivatives


f
1
=
z
2

f
f
i
x
y


.

4. Definitions of analyticity
We are now in a position to give the equivalent definitions of analyticity.
4.1. The main result. We say that a function f : U C is analytic if is
satisfies any (and thus all) of the following three equivalent conditions.
Theorem 4.1. The following are equivalent:
(1) For each z0 U we can choose  > 0 and (an )
n=0 such that the function
f (z) can be represented as a convergent power series about the point z0 ,
i.e.,
f (z) =

an (z z0 )n

n=0

for z B(z0 , ) U and with radius of convergence R >  > 0;


(2) The function f : U C is complex differentiable at each z0 U ;
u v v
(3) The partial derivatives u
x , y , x , y all exist, are continuous and satisfy
the Cauchy-Riemann equations.
We are not yet in a position to prove the theorem completely (since we will still
need to recall the Cauchy theorem for integrals). However, we will proceed with
the proofs of the parts that we can prove, and then we will return to the remaining
parts once we have established Cauchys Theorem.

4. DEFINITIONS OF ANALYTICITY

23

Proof. (1) = (2). The proof of part (2) assuming part (1) is easy. More
precisely, given z0 U we can write down its expansion
f (z) =

an (z z0 )n

n=0

for z B(z0 , r) in a sufficiently small neighbourhood of z0 . (i.e.,Pr < R)

Let R > 0 be the radius of convergence of a series f (z) = n=0 an (z z0 )n .


We can choose 0 < r < R then for h C sufficiently small
f (z0 + h) f (z0 )
f (z0 + h) a0
a1 =
a1
h
h

X
X
=
an hn1 =:
gn (h)
| {z }
n=2

where we denote

gn (h)

(
an hn1
gn (h) =
0

(1)

n=2

if h 6= 0
if h = 0

for n 2. In the disc {h C : |h| < r} we want to show that


lim

h0

gn (h) = 0,

n=2

P
(i.e., n=2 gn (h) is continuous at 0) from which we deduce, together with (1), that
f (z) is differentiable at z0 . For n 2, we can bound |gn (h)| |an |rn1 . Now since
r < R we have that

X
|an |rn1 < +
n=2

By uniform convergence we can interchange the summation and the limit to get
lim

h0

X
n=2

gn (h) = lim

h0

an hn1 = 0.

n=2

(2) = (1). To prove (1) assuming (2) requires the Cauchy theorem, so we will
return to this later (after we recall that theorem).

24

CONTENTS

(2) = (3). If f (z) = u(x, y) + iv(x, y) is a complex differentiable function of a


complex number z = x + iy C then the derivative at z0 = x0 + iy0 is given by
f (z0 + h) f (z0 )
= f 0 (z0 )
h0
h
lim

If we choose h = t R to be real then this becomes


f (z0 + t) f (z0 )
f
=
(z0 )(= f 0 (z0 ))
t0
t
x
u
v
=
(x0 , y0 ) + i (x0 , y0 )
x
x
lim

If we choose h = it where t R to be purely imaginary then this becomes


lim

t0

1 f
f (z0 + it) f (z0 )
=
(z0 )(= f 0 (z0 ))
it
i x


u
v
= i
(x0 , y0 ) + i (x0 , y0 )
y
y

Complex differentiability means that there is only one derivative and we can
equate the real parts and imaginary parts of these identities and write
u
v
(x0 , y0 ) =
(x0 , y0 )
x
y
(from the real parts) and
v
u
(x0 , y0 ) = (x0 , y0 )
x
y
(from the imaginary parts). This corresponds to the Cauchy Schwartz indentities.
We postpone the proof of continuity of the derivatives (until after we recall
Cauchys theorem).

4. DEFINITIONS OF ANALYTICITY

25

(3) = (2). We can write


f (z) f (z0 )
(u(x, y) u(x0 , y0 )) + i(v(x, y) v(x0 , y0 ))
=
.
z z0
z z0
For each |x x0 |, |y y0 | sufficiently small we can write
u
u
(x, y0 ) (y y0 ) (x0 , y0 )|
x
y
|x x0 |1 (x0 , y0 ) + |y y0 |1 (x, y0 ), and

|u(x, y) u(x0 , y0 ) (x x0 )

v
v
(x0 , y0 ) + (y y0 ) (x, y0 )|
x
y
|x x0 |2 (x0 , y0 ) + |y y0 |2 (x, y0 )

|v(x, y) v(x0 , y0 ) = (x x0 )

by using the triangle inequality


|u(x, y) u(x0 , y0 )| |u(x, y) u(x, y0 )| + |u(x, y0 ) u(x0 , y0 )| and
|v(x, y) v(x0 , y0 )| |v(x, y) v(x, y0 )| + |v(x, y0 ) v(x0 , y0 )|
and the mean value theorem, where 1 (), 2 () are bounds on the partial derivatives.
Moreover, continuity of the partial derivatives means that 1 (), 2 () 0 as z z0 .
Since |x x0 |, |y y0 | |z z0 | then for |z z0 | sufficiently small we can bound
the differences by
u
u
(x0 , y0 ) (y y0 ) (x, y0 )|
x
y
|z z0 | (1 (x0 , y0 ) + 1 (x, y0 )) , and

|u(x, y) u(x0 , y0 ) (x x0 )

v
v
(x0 , y0 ) + (y y0 ) (x, y0 )|
x
y
|z z0 | (2 (x0 , y0 ) + 2 (x, y0 ))

(1)

|v(x, y) v(x0 , y0 ) = (x x0 )

where continuity of the derivatives means that 1 (), 2 () 0 as (x, y) (x0 , y0 ).


By the Cauchy-Riemann equations we have that


u
u
v
v
(x x0 ) (x0 , y0 ) + (y y0 ) (x, y0 ) + i (x x0 ) (x0 , y0 ) + (y y0 ) (x, y0 )
x
y
x
y




u
v
u
v
= (x x0 )
(x0 , y0 ) + i (x0 , y0 ) + (y y0 )
(x0 , y0 ) + i (x0 , y0 )
x
x
y
y
|
{z
}
v
=: x
(x0 ,y0 )+i u
x (x0 ,y0 )




u
v
v
u
= (x x0 )
(x0 , y0 ) + i (x0 , y0 ) + (y y0 ) (x0 , y0 ) + i (x0 , y0 )
x
x
x
x

 

v
v
u
u
(x0 , y0 ) + i (x0 , y0 ) + i (x0 , y0 ) +
(x0 , y0 )
= ((x x0 ) + i(y y0 ))
x
x
x
x


u
v
=
(z z0 )
(x0 , y0 ) + i (x0 , y0 )
| {z }
x
x
(xx0 )+i(yy0 )

(2)
using the Cauchy-Riemann equations. Comparing (1) and (2) we have that


f (z) f (z0 )
u
v
|z z0 |
=
(x0 , y0 ) + i (x0 , y0 ) +
(1 () + i2 ())
z z0
x
x
z z0
u
v

(x0 , y0 ) + i (x0 , y0 )(= f 0 (z0 ))


x
x
as z z0 .


26

CONTENTS

Example 4.2 (Three ways to see f (z) = ez is analytic at z = 0). We let


f : C C be the exponential map f (z) = ez where z = x + iy C.
P
n
(1) We can write f (z) = n=0 (1)n zn! ;
0
(2) We have a complex derivative f (z) = ez (which makes sense as a power
series); and
(3) We can f (z) = u + iv = ex cos y + iex sin y. In particular,
v
u
v
u
= ex cos y =
and
= ex sin y = .
x
y
y
x
The continuity of the partial derivatives in part (3) of the theorem is crucial.
Without the assumption that partial derivatives are continuous, then satisfying the
Cauchy-Riemann equations is not enough to deduce that f is complex differentiable.
Example 4.3. If f : C C is defined by
(
0
f (z) =
exp(1/z 4 )

if z = 0
otherwise

Thus all the expressions in the Cauchy-Riemann equations are zero, but one can
show that f is not differentiable.
5. Integrals and Cauchys Theorem
The most useful theorem in Complex analysis is probably Cauchys theorem.
5.1. Integration on piecewise continuous curves. Given a continuous
curve : [a, b] C we can take the real and imaginary parts (t) = u(t) + iv(t),
where u, v : [a, b] R are real valued.

We say that f (t) is differentiable if u, v : [a, b] R are differentiable and then


we define
f 0 (t) = u0 (t) + iv 0 (t).
Definition 5.1. If f : [a, b] C is continuous then we define
Z b
Z b
Z b
f (t)dt =
u(t)dt + i
v(t)dt.
a

Moreover, for a piecewise continuous function f (t) on a = a0 < a1 < < an = b


we can still define the integral by
n1
X Z ai+1
f dt.
i=0

ai

Lemma 5.2. For f, g piecewise continuous on [a, b] and , C we have that:


Rb
Rb
Rb
(1) a (f (t) + f (t))dt = a f (t)dt + a f (t))dt
Rb
Rb
(2) | a f (t)dt| a |f (t)|dt

5. INTEGRALS AND CAUCHYS THEOREM

27

Proof. The first part follows easily from the definitions.


Rb
For the second part, assume that a f (t)dt 6= 0 (otherwise the result is trivial)
Rb
Rb
Rb
and we can write rei = a f (t)dt. Thus r = | a f (t)dt| = a ei f (t)dt R.
Therefore,
! Z
Z b
Z b
Z b
b

i

i
i


|f (t)| dt
e f (t) dt =
Re e f (t) dt
e f (t)dt =
r = Re
a

and thus
Z

|f (t)| dt

f (t)dt|


5.2. parameterization of the curve of integration. If : [c, d] [a, b]
is differentiable map with continuous positive derivative so that (t) is strictly
increasing then
Z b
Z d
f (s)ds =
f ((t))0 (t)dt
a

by the substitution law and the change of variables law with s = (t).
Definition 5.3. A contour (or curve or arc or path) is a continuous image by
z : [a, b] C
of a closed interval [a, b] in C with the orientation provided by the ordering on [a, b].
Thus z(a) is the first point on the contour and z(b) is the last point on the contour.
Any continuous map giving the curve and orientation is called a parameterization. The contour is called smooth if there is a parameterization z such that z is
differentiable in its interval of definition [a, b] and if z 0 (t) is continuous in [a, b].
The contour is called piecewise smooth if there exists a parameterization z :
[a, b] C for the contour and a partition a = a0 < a1 < < an = b such that
z|[ai , ai+1 ] is smooth.
A contour is simple if can be given by an injective parameterization (i.e., z(a) =
z(b) = a = b)
A contour is closed and simple if it can be given by an injective parameterization
z : [a, b) C (i.e., z(a) = z(b)a = b) and or z(a) = z(b).

Remark 5.4. A piecewise smooth curve may be given by many parameterizations. If : [c, d] [a, b] is strictly increasing then w; [c, d] C defined by
w(t) = z((t)) is another parameterization.
In complex analysis we are usually concerned with piecewise smooth curves
and piecewise smooth parameterizations. A piecewise smooth curve may be given
by many parameterisations, i.e., if : [c, d] [a, b] is strictly increasing then

28

CONTENTS

w : [c, d] C defined by w(t) = z((t)) is also piecewise smooth if z is piecewise


smooth and is differentiable with continuous and positive derivative.
Definition 5.5. We say that two parameterizations z and w are equivalent if
there exists such a .
Exercise 5.6. Show that this is an equivalence relation on parameterizations.
Hint. Clearly we can write w 1 = z and 1 is continuous with positive
derivative. Transitivity follows simply.

If z : [a, b] C is a piecewise smooth parameterization of then we
obtain a piecewise parameterization of (in reverse) by z : [b, a] C with
z (t) = z(t). Let us denote by the reverse of .

Definition 5.7. If 1 is piecewise smooth and 2 is piecewise smooth and if the


last point of 1 is the same as the first point of 2 then we define the composition
1 2 as follows:
If z1 : [a, b] 1 C is a parameterization of 1 and z2 : [c, d] 2 C is a
parameterization of 2 we define z3 : [a, b + d c] 3 C by
(
z1 (t)
if t [a, b]
z3 (t) =
z2 ((t)) if t [b, b + d c]
where is the translation (t) = t + (b c).
We are interested in integration around a piecewise smooth simple contour.
Definition 5.8. Given smooth curves zi : [ai , ai+1 ] i C for
a = a0 < a1 < < an = b
we define
Z
f (z)dz =

n Z
X
i=1

f (z)dz.

(Once we define the integral for smooth contours we can then extend it to piecewise
smooth contours.) If is a smooth contour parameterized by z : [a, b] C
then we define
Z
Z
b

f (z(t))z 0 (t)dt

f (z)dz :=

R
Exercise 5.9. Show that f (w)dw = f (z)dz
R
R a
Proof. We can write f (z)dz = b f (z(t))z 0 (t)dt. Substituting s = t
Rb
R
this becomes: a f (z(s))z 0 (s)ds = f (z)dz.


5. INTEGRALS AND CAUCHYS THEOREM

29

Remark 5.10. More generally, integration with respect to arc length along a
simple contour is rectifiable if it is given by a continuous map z : [a, b] C which
is of bounded variation, i.e., the summations
n1
X

|z(ai+1 ) z(ai )|,

i=0

ranging over all partitions a = a0 < a1 < < an = b, is bounded. The supremum
of these numbers is by defintion the length of . In particular, if is smooth
Rb
represented by z : [a, b] C then is rectifiable and l() = a |z 0 (t)|dt.
The following upper bound is useful.
Lemma 5.11 (Integration along a straight line). If is piecewise smooth and f
is continuous on and |f (z)| M on then
Z



f (z)dz M l().

Proof. Let z : [a, b] C be a parameterization. Then



Z b
Z




0
f (z)dz =
f
(z(t))z
(t)dt




a

Z b

|f (z(t))|.|z 0 (t)|dt
a

Z
M

|z 0 (t)|dt = M l()

as required.

It is useful to illustrate this by a couple of examples .


Example 5.12 (Integration of z n along a straight line). Let = [w1 , w2 ] and
consider the affine parameterization z : [0, 1] C given by
z(t) = w1 (1 t) + w2 t, for 0 t 1,
then z 0 (t) = (w2 w1 ) and by defintion
Z
Z 1
f (z)dz =
f (w1 + t(w2 w1 )t) (w2 w1 )dt.
[w1 ,w2 ]

Consider the specific example f (z) = z n , then if n 6= 1 then


Z 1
wn+1
wn+1
(w1 + t(w2 w1 ))n (w2 w1 )dt = 2
1
n+1 n+1
0
Example 5.13 (Integration around a simple closed curve). Let be the -arc
in the unit circle with z : [0, ] C given by
z(t) = eit for 0 t

30

CONTENTS

then z 0 (t) = ieit and by definition


Z
Z
f (z)dz =

enit ieit dt.

Consider the specific example f (z) = z n then


Z
Z
h
i
i
ei(n+1) 1
f (z)dz =
enit ieit dt =
ei(n+1)t =
(n + 1)i
n+1
0

0
if n 6= 1. On the other hand, if n = 1 then one can easily see that
Z
Z
Z
1
it it
dz =
e ie dt =
idt = i
0
0
z
In particular, if = 2 then
Z
Z
znd =

2
nit

ie

int

dt =

0
2i

if n 6= 1
if n = 1.

5.3. Theorems of Gousart and Cauchy. The last example is a special case
of the following important theorem.
Theorem 5.14 (Gousarts Theorem).
R Let f : U C be an analytic function.
Let U be a closed simple curve then f (z)dz = 0.
There are two basic approaches to proving this. One approach uses Stokes
theorem, but for variety we will describe the other approach (at least in the context
of being a rectangle).
Proof. We begin with a proof of this result in the case that is the boundary
of a rectangle R0 = [a, b] [c, d]. In particular
R0 = [a, b] {c} {b} [c, d] [b, a] {d} {a} [d, c].
We proceed as follows.
Step 1. We can subdivide R0 into four similar sub-rectangles:
(1)

R0 = [a, (a + b)/2] [c, (c + d)/2]


(2)

R0 = [(a + b)/2, b] [c, (c + d)/2)]


(3)

R0 = [a, (a + b)/2] [(c + d)/2, d]


(4)

R0 = [(a + b)/2, b] [(c + d)/2, d]


and then we can write
Z
f (z)dz =
R0

4 Z
X

f (z)dz.
(i)

i=1

R0

(where the integrals along the same curves in different directions cancel).

5. INTEGRALS AND CAUCHYS THEOREM

31

In particular. we can bound


Z
Z






4
max
f
(z)dz



1i4




f (z)dz .
(i)

R0

R0

Let us denote by R1 the subrectangle which maximizes the last term.


Step 2. We can repeat this iteratively to construct a sequence of subrectangles
R0 R1 R2 R3 Rn
such that Rn , n 1, is a sub-rectangle of size
Z

Z



n


f (z)dz 4

R0

(ba)
2n

Rn

(dc)
2n

and we can bound



f (z)dz .

(a)

We can choose z0
n=0 Rn (which implicitly uses compactness of R0 ).
We now want to use that f (z) is complex differentiable (at z0 ). In particular,
for any  > 0 we an choose > 0 so that providing |z z0 | < then
|f (z) f (z0 ) + f 0 (z0 )(z z0 )| |z z0 |.

(b)

Provided n is sufficiently large that the diameter


1p
diam(Rn ) = n (b a)2 + (d c)2
(c)
2
of Rn is smaller than then we have from (b) and (c) that for all z Rn we can
write that
|f (z) f (z0 ) f 0 (z0 )(z z0 )|  diam(Rn )
(d)
 p
n (b a) + (d c)
2

32

CONTENTS

Moreover, since
length(Rn ) =

1
(2(b a) + (d c))
2n

(e)

we can use (a), (d) and (e) to bound


Z
(f (z) f (z0 ) f 0 (z0 )(z z0 ))dz
Rn

sup {|f (z) f (z0 ) f 0 (z0 )(z z0 )|} length(Rn )


|
{z
}
zRn
{z
} 1n (2(ba)(dc))
|
2


2
2
2n

(f )

(ba) +(dc)


 p
2 + (d c)2 2((b a) + (d c)) .
(b

a)
4n |
{z
}
=:C

Step 3. Moreover, we have the following estimates:


For each rectangle Rn :
R
(1) We can evaluate Rn 1dz = 0;
R
(2) We can evaluate Rn (z z0 )dz = 0.
Proof. The first part is easy, since the contribution to the integrals from
opposite sides cancel. For the second part, we can simply explicitly do the integral.

Comparing (a) and (f) and the above claim we can write:


Z



f (z)dz
f (z)dz 4n
Rn
R0
Z
Z
n
=4 |
(f (z) f (z0 ) f 0 (z0 )(z z0 )) dz
(f (z0 ) f 0 (z0 )(z z0 )) dz |
Rn
R
| n
{z
}
= 0 (by claim)


C
4n

= C
4n
R
Thus, since  can be chosen arbitrarily small, we finally deduce that R0 f (z)dz = 0,
as required.

Z


Remark 5.15. For more general simple curves we can approximate by a


piecewise linear curve 0 consisting of horizontal and vertical lines. We can then fill
in the region inside the curve by rectangles and apply the result above.

Corollary 5.16. Let f : U C be analytic (i.e., complex


differentiable). Let
R  1 
dz = 2i.
zz0

U be a simple closed curve containing z0 U . Then

5. INTEGRALS AND CAUCHYS THEOREM

33

Proof. We can assume for simplicity that z0 = 0 (otherwise we simply change


to the translated function for z 7 f (z z0 )). We can also assume that contains
the unit disk D, since otherwise we can scale by z 7 f (zR), for large enough R.
In the previous section we saw that the result holds with replaced by the unit
circle 0 . However, this suffices, because we can change the integral from one along
to one along
0 by introducing an extra arc c joining the two. Then the integrals
R
1
= 0, since the function is analytic inside the closed curve
differ by c0 c1 zz
0
1
c0 c .


We now deduce Cauchys Theorem from this.


Theorem 5.17 (Cauchys Theorem). Let f : U C be an analytic function
(i.e., complex differentiable). Let U be a simple closed curve. Assume that z0
lies inside the curve , then
Z
f (z)
1
f (z0 ) =
dz
(1)
2i z z0

Proof. Step 1. Consider the function


(

f (z)f (z0 )
zz0
0

g(z) =

f (z)

if z 6= z0
if z = z0

Moreover, by continuity of f at z0 ,
g(z)(z z0 ) = lim (f (z) f (z0 ))(z z0 )) = 0
zz0

as z z0 .

Given  > 0 we can choose > 0 such that for |z z0 | < 2 we have
|g(x) g(z0 )| |z z0 |.
Step 2. Let z0 = x0 + iy0 Given a rectangle R we want to divide it into nine
subrectangles, including
R0 = [x0 , x0 + ] [y0 , y0 + ]

34

CONTENTS

then cancellations between the integrals on the boundaries of the eight rectangles
and applying Gousarts theorem to each of these we get
Z
Z
g(z)dz =
g(z)dz
R0

R
0

Moreover, since for z R we can write


|g(z) g(z0 )|




|z z0 |

we can write

Z






8
0


g(z)dz length(R ) sup |g(z)|
sup |f (z)|

zR0
zR0
R0
which can be made arbitrarily small be choice of .
Step 3.
We now apply Gousarts Theorems to deduce that



Z
Z 
Z 
Z 
f (z) f (z0 )
f (z)
1
0=
g(z)dz =
dz =
dz f (z0 )
dz
z z0
z z0
z z0

|
{z
}
2i


5.4. Immediate applications of Cauchys Theorem. Cauchys theorem
has a number of immediate corollaries.
Corollary 5.18. Let f : U C be analytic and let U be a simply connected
domain. LetRz1 , z2 U . Then for any piecewise smooth curve connecting z1 to z2
the integral f (z)dz is independent of .

Proof. Let 1 , 2 be two piecewise smooth curves leading from z1 to z2 contained in U then 1 2 is a closed curve and thus
Z
Z
Z
f (z)dz = 0 =
f (z)dz
f (z)dz
1 2

and the result follows.


Corollary 5.19. There is a formula for the derivatives of the form
Z
f (z)
1
0
dz
f (z0 ) =
2i (z z0 )2

()

where z0 is inside the simple closed curve . More generally, the kth derivative
takes the form
Z
k!
f (z)
(k)
f (z0 ) =
dz
()
2i (z z0 )k+1
Proof. We need to justify differentiating under the integral sign. Using Cauchys
theorem we can write, (for sufficiently small h C):


Z
1
1
1
1
f (z0 + h) f (z0 )
=

f (z)dz
h
2i h (z z0 h) (z z0 )

Z 
1
1
=
f (z)dz
2i (z z0 h)(z z0 )
We claim that the limit exists as h 0 and that it equals
Z
1
f (z)
f 0 (z0 ) :=
.dz
2i (z z0 )2

5. INTEGRALS AND CAUCHYS THEOREM

35

To see this we can bound






Z 
f (z0 + h) f (z0 )

h
1
0



f (z)dz
f (z0 )

h
2i (z z0 h)(z z0 )2



|h|
length()
sup |f (z)|
2 (d(z0 , ) |h|)d(z0 , )2
z
where d(z0 , ) = inf z |z0 z| and observe that the Right Hand Side tends to zero
as |h| 0.
The higher derivative formula comes similarly (cf. next application).


Application 5.20 (Completing the equivalence of the definitions of analyticity


I: Writing analytic functions as power series). We claim that a complex differentaible
function always has a power series expansion.
Let f : U C be a complex differentiable function function. Let z0 U and
choose  > 0 such that
B(z0 , ) = {z C : |z z0 | < R} U

Choose z B(z0 , ) and then := |z0 z| <  and let r be such that < r < .
Then by Cauchys theorem
Z
1
f ()
f (z) =
d
2i C(z0 ,r) z
where C(z0 , r) = {z C : |z z0 | = r} is the circle of radius r about z0 .

For < | z0 | = r <  we have that


1
1
=
z
( z0 ) (z z0 )
1


=
0
( z0 ) 1 zz
z0
!

 
2
1
z z0
z z0
=
1+
+
+
( z0 )
z0
z0


0
which is uniformly convergent on the circle C(z0 , r) since zz
z0 = r < 1. So by
Cauchys theorem integrating around C(z0 , r) gives
Z
1
f (z)
f (z0 ) =
d
2i C(z0 ,r) ( z0 )
Z
Z
1
f (z)
1
f (z)
=
d + (z0 z0 )
d
2i C(z0 ,r) ( z0 )
2i C(a,r) ( z0 )2
Z
1
f (z)
+ (z0 z0 )2
d +
2i C(a,r) ( z0 )3
In particular, we can write this as
f (z0 ) = a0 + a1 (z0 a) + a2 (z a)2 +

36

CONTENTS

where
1
an =
2

Z
C(a,r)

f (z)
d
( z0 )n+1

and we know the series has Radius of convergence R = lim supn+ |an |1/n
r. Thus f (z) is a power series about a.

1

>

Application 5.21 (Completing the equivalence of the definitions of analyticity


II: Continuity of the partial derivatives). Let f : U C be complex differentiable.
For z0 U we can U be a simple closed curve containing z0 . We see from
(*) that f 0 (z0 ) depends continuously on the point z0 . Furthermore, since complex
differentability implies that for f (x + iy) = u(x, y) + iv(x, y) we can write
f 0 (z0 ) =

u
v
v
u
+i
=
i
x
x
y
y

from which we can deduce that the partial derivatives are continuous.
Corollary 5.22. Let f : U C be analytic and let z0 be a unique zero (of
multiplicity one). Let U be a simple closed curve. Then
Z
f 0 (z)
1
z
dz
z0 =
2i f (z)
Proof. We can write f (z) = (z z0 )g(z), where g : U C where g : U C
0
(z)
is non-zero and analytic (and, in particular, g(z0 ) 6= 0). The function ff (z)
is of
1
0
the form zz0 + (z) where : U C is analytic and given by (z) = g (z)/g(z).
Thus, by Cauchys theorem we can write
Z
Z
Z
1
f 0 (z)
1
1
1
z
dz =
z
z(z)dz = z0 .
dz +
2i f (z)
2i z z0
2i
{z
} |
{z
}
|
=z0

=0


Application 5.23 (Perturbation Theory for eigenvalues of matrices). Assume
that A is a matrix and A is a simple eigenvalue. Then letting fA (z) = det(zI A)
then we see that z0 is a simple zero for f (z). By the previous corollary, we see that
Z
1
f 0 (z)
A =
z A dz
2i fA (z)

However, it is easy to see that the coefficients of the polynomials fA (z) and
fA0 (z) depend smoothly on the entries in the matrix A. This is true on and thus
we can deduce the same for A .
5.5. Converse to Cauchys Theorem. The following is a converse to Cauchys
Theorem which is often useful in proofs of later results.
Theorem 5.24 (Moreiras
Theorem). Let U be a domain and let f : U C be
R
continuous and satsify f (z)dz = 0 for all simple closed curves U . Then f
is analytic.
Proof. Fix a point z0 U . We can define F : U C by associating to any
z U a piecewise smooth curve parameterized by z : [0, 1] C where
z(0) = z0 and z(0) = z and then defining
Z
F (z) =
f (z)dz.

6. PROPERTIES OF ANALYTIC FUNCTIONS

37

We first observe that since U is a (simply connected) domain this is well defined,
since if 1 is a second path from z0 to z then we have that
Z
Z
Z
Z
Z
f (z)dz
f (z)dz =
f (z)dz +
f (z)dz =
f (z)dz = 0

(1 )

since (1 ) is a picewise closed curve, and F : U C is analytic, and thus we can


apply Cauchys theorem.

We can now deduce that F 0 (z) = f (z), i.e., f is the complex derivative of F .
But if F it is complex differentiable then it is analytic, and thus all of the derivatives
exist, and us f is also complex differentiable.

6. Properties of analytic functions
We want to find some way to understand better properties of analytic functions.
We will begin with properties of individual functions and move onto families of
functions. However, first we have to recall some basic facts.
6.1. Logarithms and roots. Consider the exponential function z 7 ez for
z C. This is analytic for every z C. We would like to define the inverse map,
the logarithm, in the complex plane but this has a few complications that need to
be addressed.
Lemma 6.1. Let z = x + iy and w = u + iv. Then ez = ew when there exists
n Z such that y = v + 2n.
Proof. If ez ex+iy = ew = eu+iv then taking the modulus gives |ez | = eu =
e = |ew | gives that u = x. This leaves that eiy = eiv and thus y v is an integral
multiple of 2.

x

In particular, exp : z 7 ez maps any strip R [, + 2) bijectively onto


C {0}. However, the inverse to exp() is not, strictly speaking, a well defined
function since there are many points mapped to a single point. However, by the
lemma these points all differ by values 2in, n Z.

38

CONTENTS

We would like to write w = log z when z = ew . If w = u + iv and exp w = z


then
z
eu eiv = |z|
= |z|ei
|z|

z
for some since |z|
= 1. Hence eu = |z| and v = + 2n for any n Z.
Therefore u = log |z| and w = log |z| + i( + 2n) for some n Z.
Definition 6.2. For z 6= 0 define the function arg(z) = arg(z/|z|) = if
z/|z| = ei and < .
For log() there are different branches of arg with different restricted domains.
The principle branch of arg is denoted Arg(z) and is such that Arg(ei ) is the unique
number of the form + 2n which lies between and . This is not defined on
the negative real line (, 0].
Returning to the definition of log:
Definition 6.3. The principle branch of log is denoted Log(z) and is of the
form log |z| + iArg(z). This is not defined on the negative real line (, 0].
We can then use exp and Log to define complex powers of complex numbers.
In particular, we can write that z w = exp(w.Log(z)).
Example 6.4. Find the value of
(1 i)1+i := exp ((1 + i)Log(1 i))

We observe that arg(i i) = 34 and Log(i i) = log 2 34 i. Thus





3
(1 i)1+i = exp ((1 + i)Log(1 i)) = exp((1 + i) log 2 i) .
4
Example 6.5. If N then we can write
z 1/ = exp (log(z)/) = exp ((log |z| + iArg(z))/)).
If z = rei then
z 1/ = exp((1/)(log r + i( + 2n))) = r1/ ei(+2n)/ = r1/ ei( e2in)/ .
These take the values r1/ , r1/ ei2 , , r1/ ei(+2(p1)/p .
In particular




1
1
1
log(z) = exp
log |z| + arg(z) = |z|1/2 eiArg(z)/2 .
z 1/2 = exp
2
2
2
More generally, for z q/p we simply replace z by z q .
Remark 6.6. If is irrational then we can write
z = exp ( log z) = exp( log r + i( + 2n)) .
and the set of values with n Z is infinite. To see this, let z = rei then
exp(i( + 2n)) = exp(i( + 2m)) exp(i2(n m)) = 1
2(n m) = 1 n = m.
(since is irrational).

6. PROPERTIES OF ANALYTIC FUNCTIONS

39

6.2. Location of zeros: Argument Principle and Rouch


es Theorem.
We can apply Gousarts and Cauchys theorem to prove two useful results
Lemma 6.7. Suppose that f : U C is analytic and non-zero then there exists
an analytic function h : U C such that eh(z) = f (z) for all z U .
Proof. Since f is analytic we say that f 0 is analytic and since f does not
vanish then f 0 /f is analytic and so f 0 /f has a primitive h by integrating f , i.e.,
h0 = f 0 /f for h analytic. Thus

d 
f (z)eh(z) = f 0 (z)eh(z) f (z)h0 (z)eh(z) = 0
dz
so f (z)eh(z) = Constant = ea 6= 0. Thus f (z) = ea+h(z)

We can use this lemma to show the following theorem.


Theorem 6.8 (The Argument Principle). Assume f : U C is analytic and
U is a simple closed curve. The general case is similar. Assume that f has no
zeros on . Then the number of zeros N inside is given by
Z 0
1
f (z)
N=
dz.
2i f (z)
Proof. Assume that z0 is a zero of order n 1 for f (z). Assume for simplicity
that we have no other zeros.

Using the lemma we can then write write f (z) = (z z0 )N eh(z) and then
f 0 (z) = (z z0 )N eh(z) h0 (z) + N (z z0 )n1 eh(z)
and

f 0 (z)
N
= h0 (z) +
f (z)
z z0
Since the first term is analytic simple pole at z0 we can apply Gousarts theorem
to the first term and Cauchys theorem to the second term to write
Z 0
1
f (z)
N=
dz.
2i f (z)
With more zeroes the argument naturally generalizes.

Example 6.9. Assume f : D C has a simple zero at 0 (i.e., N = 1) and that


= {z : |z| = r} with 0 < r < 1. Writing
f (z) =

an z n and f 0 (z) =

n=1

X
n=1

(with a0 = 0 and a1 6= 0) we see that


1
f 0 (z)
= + g(z)
f (z)
z

nan+1 z n

40

CONTENTS

where g(z) is analytic on a neighbourhood of 0. By Gousarts theorem


and a direct calculation gives
Z 0
Z
Z
1
f (z)
1
1
1
dz =
dz +
g(z)dz
2i f (z)
2i z
2i
|
{z
} |
{z
}
=1

g(z)dz = 0

=0

Application 6.10 (localizing zeros of Riemann Zeta function). One of the


major open problems in mathematics is the Riemann Hypothesis. For s C with
Re(s) > 1 we define the Riemann zeta function
(s) =

X
1
ns
n=1

P
(It converges since |(s)| n=1 nRe(s) < +). Moreover, it has an analytic
extension to C {1}, i.e., there is an analytic function f : C {1} C such that
f (s) = (s) for Re(s) > 1.

Riemann Hypothesis (Conjecture, 1859) The only zeros for f (s) occur at
s = 2, 4, 6, and on the line
L := {s =

1
+ it : t R}.
2

This is Hilberts 8th problem from his famous list of 23 open problems from 1900,
and one of the 7 Millenium problems from 2000 for which the Clay Institute offered
a million dollars.
Computer searches for counter-examples (so far unsuccessful!) use the Argument Theorem. One considers a small closed curve away from the line L and
R f 0 (z)
1
estimates the integral 2i
dz. One tries to find a such that the integral is
f (z)
non-zero.

Theorem 6.11 (Rouches Theorem: Nearby functions and nearby zeros). Assume that f, g : U C are analytic. Let U be a simple closed curve and
assume that neither f (z) or g(z) have zeros on and that for each z we have
that |g(z)| < |f (z)|. Then f and f + g have the same number of zeros in U .
Proof. This uses the walking the dog method.

6. PROPERTIES OF ANALYTIC FUNCTIONS

41

Let F (z) = fg(z)


(z) then if Nf and Nf +g are the number of zeros inside for each
of the two functions then by the Argument Principle we can write
Z
Z 0
1
(f 0 + g 0 )(z)
1
f (z)
Nf +g =
dz and Nf =
dz.
2i (f + g)(z)
2i f (z)
Thus
1
Nf +g Nf =
2i

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

f 0 (z)
dz
f (z)

f 0 (z)(1 + F (z)) + f (z)F 0 (z)


1
dz
f (z)(1 + F (z))
2i


Z  0
Z 0
0
f (z)
f (z)
1
F (z)
1
=
+

2i f (z)
1 + F (z)
2i f (z)

Z 
F 0 (z)
1
=
dz
2i 1 + F (z)

1
=
2i

f 0 (z)
f (z)

But F (z) 6= 1 for z , since |F (z)| < 1, and so the final expression counts the
number of z inside with F (z) = 1.
Finally, observe that if 0 t 1 we can replace g(z) by tg(z) (and F (z) by
tF (z)) and providing t is small enough we have that the final expression is zero.
But if we assume for a contradiction that Nf +g 6= Nf then it will contradict the
obvious continuity of

Z 
1
tF 0 (z)
t 7
dz
2i 1 + tF (z)

s
Example 6.12. Show that all four of the zeros of z 4 7z 1 lie in B(0, 2) =
{z C : |z| < 2}.
Let f (z) = z 4 then all four zeros lie inside B(0, 2). In fact, z = 0 is a zero of
multiplicity four.
Let g(z) = 7z 1. Let = C(0, 2) = {z C : |z| = 2}. For |z| = 2 we have
that
|g(z)| = | 7z 1| 7|z| + 1 = 15 < 16 = |z 4 | = |f (z)|.
Thus by Rouches theorem we have that f (z) + g(z) = z 4 7z 1 has the same
number of zeros in B(0, 2) as f (z), i.e., four.
Example 6.13. Show that all five of the zeros of z 5 + 3z 3 + 7 lie in B(0, 2) =
{z C : |z| < 2}.
Let f (z) = z 5 then all five zeros lie inside B(0, 2). In fact, z = 0 is a zero of
multiplicity five.
Let g(z) = 3z 3 + 7. Let = C(0, 2) = {z C : |z| = 2}. For |z| = 2 we have
that
|g(z)| = |3z 3 + 7| 3|z 3 | + 7 = 31 < 32 = |z 5 | = |f (z)|.

42

CONTENTS

Thus by Rouches theorem we have that f (z) + g(z) = z 5 + 3z 3 + 7 has the same
number of zeros in B(0, 2) as f (z), i.e., five.
Example 6.14. Show that z 4 7z 1 has precisely one zero in the unit disk
D = {z C : |z| < 1}.
Let f (z) = 7z 1 then f (z) has precisely one zero in D, at z = 71 .
Let g(z) = z 4 . Let = C(0, 1) = {z C : |z| = 1}. For |z| = 1 we have that
|g(z)| = |z 4 | = 1 < 6 = 7 1 = |7z| 1 | 7z 1| = |f (z)|.
Thus by Rouches theorem we have that f (z) + g(z) = z 4 7z 1 has the same
number of zeros in D as f (z), i.e., one.
Remark 6.15. There is a more symmetric version of this theorem. If f1 : U
C and f2 : U C are analytic then |f1 (z) f2 (z)| < |f1 (z)| + |f( z)| for z then
f1 (z) and f2 (z) have the same number of zeros inside the curve.
Corollary 6.16 (Weierstrass-Hurwitz Theorem). Assume fn : U C are a
sequence of non-zero analytic functions. If fn converges uniformly on compact sets
to a continuous function f (i.e., if K U is compact then supzK |f (z)fn (z)| 0
as n +) then:
(1) f : U C is analytic
(2) f either has no zeros or is identically zero.
Proof. For the first part (due to Weierstrass) one only needs to show that for
any simple closed curve U we have that
Z
Z
Z





f (z)dz fn (z)dz + (f (z) fn (z))dz



|
{z
}
=0


0 + length() sup |f (z) fn (z) |
zK
|
{z
}
0

whereRthe first term is zero by Gousarts theorem (since fn : U C is analytic).


Thus f (z)dz = 0 for all simple closed curves, and so f is analytic by Moreiras
Theorem.
For the second part (due to Hurwitz), if f (z) isnt identically zero then the set
of zeros {zi }N
i=1 for f (z) is finite. We need to show that this set is empty. Assume
for a contradiction it isnt empty and let be a closed simple curve which contains
one of these zeros z1 , say. By Cauchys theorem
Z 0
Z 0
1
fn (z)
1
f (z)
0=
dz
dz = 1
2i fn (z)
2i f (z)
as n + which gives a contradiction.

Application 6.17. Recall that the Riemann zeta function

X
1
(s) =
ns
n=1
was claimed to analytic for Res(s) > 1. To check this we can observe that
!
N
N
X
X
1
fN (s) =
=
exp (s log n)
ns
n=1
n=1
is analytic on C. For any compact set K U := {s C : Re(s) > 1}, say, we have
that supsK |(s) fn (s)| 0. Thus by the theorem (s) is analytic on U .

6. PROPERTIES OF ANALYTIC FUNCTIONS

43

6.3. Liouvilles Theorem and the Fundamental Theorem of Algebra.


We begin with another application of the Taylor series version of analyticity.
Theorem 6.18 (Liouvilles Theorem). If a function f : C C is analytic and
bounded then f is constant.
Proof. We can write f (z) as a power series representation centred on 0:

X
f (z) =
an z n .
n=0

Now we use an estimate for an for any R > 0:




1 Z
M 2R
f (z)
M

dz
= n
|an | =
2i C(0,R) z n+1 Rn+1
R
where M is an upper bound for |f (z)| and taking limits as R to get an = 0
except for n = 0, i.e., f (z) = a0 .

Functions which are analytic on all of U = C are called entire functions.
Corollary 6.19 (Fundamental Theorem of algebra). Every non-constant polynomial has a root (i.e., a zero, that is there exists z0 such that p(z0 ) = 0).
Proof. Suppose that the polynomial p(z) does not vanish. Then 1/p(z) is
entire and since |p(z)| + as z + we have that 1/p(z) 0 as z +
and hence |1/p(z)| <  for |z| > R and 1/p(z) is bounded for |z| R so 1/p(z)
is bounded throughout C so 1/p(z) is constant, contradicting the fact that p(z) is
non-constant.

In particular, this implies that the polynomial can be written as:
p(z) = an z n + + a0 = an (z z1 )(z z2 ) (z zn )
Let z1 be a zero and then write
p(z) = an (z z1 +z1 )n + +a0 = an (z z1 )n +a0n1 (z z1 )n1 + +a01 (z z1 )+a00
Clearly a00 = 0, by evaluation at z1 . So p(z) = (zz1 )q(z) where q(z) is a polynomial
of degree n 1 (and the leading coefficient of q(z) is 1). Repeating this on q(z) we
obtain by induction
p(z) = an z n + + a0 = an (z z1 )(z z2 ) (z zn ).
6.4. Identity Theorem. The following is an application of the Taylor series
theorem
Lemma 6.20 (Isolation of zeros). Let r > 0 and let B(z0 , r) = {z C : |zz0 | <
r}. The zeros of a non-trivial f : D(z0 , r) C cannot accumulate within a disk
D(z0 , r ), where  > 0.
Proof. Assume for a contradiction that zk w B(z0 , r ), for some
 > 0, are a convergent sequence of zeros for f (z). We can consider the restriction
f : B(w, ) C and write

X
f (z) =
an (z w)n
n=0

satisfies an = 0, for all n 0, i.e., f (z) is trivial contradicting the hypothesis. More
precisely, assume for a contradiction to the claim that for some N 0 we have that
a0 = a1 = = aN 1 = 0 and |an | > 0. Since f (zk ) = 0 each k 1 we can write

X
f (zk ) = 0 =
an (zk w)n
n=N

44

CONTENTS

If |an | C(2)n , say, then we can bound




X
X
2|zk w|)N

n
(r|zk w|)n = |an | C
0=
an (zk w) |an | C


1 2|zk w|
n=N

n=N

Providing k is sufficiently large the right hand side is positive, giving a contradiction.

Theorem 6.21 (Identity Theorem). Suppose that f, g : U C are analytic
in the domain U and suppose that {zn }, z0 U and zn z0 with zn 6= z0 with
f (zn ) = g(zn ).
Proof. We have already proved this for a disk with centre z0 , say, by the
previous lemma, we see that f = g on a disk with centre z0 contained in U .
Let z1 U then we can see that f (z1 ) = g(z1 ) by parameterizing a polynomial
line from z0 to z1 . Otherwise let w be a point on this line such that f (z) = g(z)
for all points from z0 up to w and for some points arbitrarily close to w we have
f (z) 6= g(z). But w is the limit of a sequence {wn } of points with f (wn ) = g(wn )
and so f = g in a disk cntred at w
This contradicts the definition of w if w is before z1 . Hence z1 = w and
f (z1 ) = g(z1 ).

Corollary 6.22 (Isolation of values). Let f : U R be analytic and nonconstant. Let C then for each z0 U there exists a punctured disc P (z0 , r) =
{z C : |z0 z| < r} {z0 } such that f (z) 6= .
Proof. Assume for a contradiction that for every small punctured disk of
radius r = 1/n, say, there exists zn P (z0 , 1/n) such that f (zn ) = . But this
implies f (z) = on U , i.e., f (z) is constant and contradicts the hypothesis.

Corollary 6.23 (Open Mapping Theorem). Let f : U R be a non-constant
analytic function. Then the image f (V ) of any open set V U is again an open
set.
Remark 6.24. This is very different to case for R. Consider f : R R defined
by f (x) = x2 is not an open map since, for example, f ((1, 1)) = [0, 1), which is
not open.
Proof. Let z0 V then it is enough to show that there exists  > 0 such that
B(f (z0 ), ) f (V ).
Let g(z) = f (z) f (z0 ) then g(z0 ) = 0. Moreover, z0 must be a simple zero
(i.e., multiplicity one) for g otherwise the function g(z) (and hence f (z)) would be
a constant function. In particular, there is a simple closed curve about z0 and
within V , containing which z0 as the unique zero. Moreover, by taking the curve
smaller (if necessary) we can assume that g 0 (z) = f 0 (z) has no zeros within . By
the Argument theorem we see that
Z 0
Z
g ()
1
f 0 ()
1
d =
d.
1=
2i g()
2i f () f (z0 )
Moreover, for fixed, this remains true if we replace f (z0 ) by f (z) with |f (z)
f (z0 )| < , for sufficiently small > 0. i.e, for such f (z) B(f (z0 ), ) there exists
a unique w inside , and thus in V .

6.5. Families of functions and Montels Theorem. Recall that a set K
C is compact if and only if it is closed and bounded.
Definition 6.25. Let {fn }
n=1 be a family of continuous functions fn : U C,
for n 1. We say that this family is normal if there is a subsequence {fnk }
k=1 which
converges uniformly on every compact subset of U to a continuous function f : U

6. PROPERTIES OF ANALYTIC FUNCTIONS

45

C, i.e., for every compact subset K U we have that supzK |fni (z) f (z)| 0
as ni +.
We can also adopt the convention that we allow uniform convergence to
(consistent with the functions being viewed as being valued in the Riemann Sphere).
In the case of analytic functions, we can get that the family is normal by just
assuming the sequence is uniformly bounded.
Theorem 6.26 (Montels Theorem). Let fn : U C, n 1, be a family of
analytic functions. Assume that there is a positive constant M > 0 such that
|fn (z)| M for all z U, n 1
{fn }
n=1

then
analytic.

is a normal family. Moreover, the limiting function f : U C is

The proof is based on a classical result in real analysis. We recall the following
definition.
Definition 6.27. Let K be a compact set. Let {fn }
n=1 be a family of continuous functions fn : K C, for n 1. We say that this family is equicontinuous if
for every  > 0 there is a > 0 such that whenever z, w K with |z w| < then
|fn (z) fn (w)| <  for all n 1.

The corresponding result from real analysis we need is the following.


Theorem 6.28 (Arzela-Ascoli). Any equicontinuous family fn : K C has a
subsequence (fnk ) which is uniformly convergent to a continuous function f : K C
(i.e., supzK |fnk (z) f (z)| 0 as nk +)

Proof. Let z0 U and let r > 0 be so that B(z0 , r) U . Then C U and


B(z0 , r) are disjoint open sets. We can choose > 0 such that if w C U and
z B(z0 , r) then |z w| > .

46

CONTENTS

If z B(z0 , ) then we can consider restriction fn : B(z0 , r) C and by


Cauchys Theorem


M
1 Z
f (z)


0
dz
|f (z)| =

2i |z0 z|= (z z0 )2

Thus for any z1 , z2 B(z0 , ) we have that


!
|fn (z1 ) fn (z2 )|

sup

|f (z)| |z1 z2 |.

|z|B(z0 ,)

In particular, we see that the restrictions fn : B(z0 , ) C, n 1, form an


equicontinuous family. Given  > 0 we simply choose = C > 0.

If K U is compact we can cover it be finitely many such balls and we again


deduce that fn : K C, n 1, forms an equicontinuous family. We can then
invoke the Arzela-Ascoli theorem to deduce that there is a convergent subsequence
fnk : K C, for k 1.
The analyticity of the function f : U C follows from the Hurwitz-Weiestrauss
theorem.

Remark 6.29. There is a stronger version which says that fn : U C, n 1,
is a normal value if they all omit the same three values.
b C
b be a polynomial (or, more
Application 6.30 (Julia sets). Let f : C
generally, a rational map).
Definition. We say that z C is in the Fatou set F if there exists  > 0 such that
the compositions {f n : B(z0 , ) C : n 1} is a normal family.
We define the Juia set J to be the complement of the Fatou set J = C F.
n

Example. If f (z) = z 2 then f n (z) = z 2 . Thus on U = {z C : |z| > 1|} we


have that f n (z) +, i.e., the family of iterates converge to . On the other
hand, on U0 = {z C : |z| < 1|} we have that f n (z) 0, i.e., the family of iterates
converge to . Thus the Fatou set contains U U0 . But if |z| = 1 then in any
disk B(z, ) those points also inside the unit disk will converge to 0 while those
points also outside the unit disk will converge to . i.e., f n (z) doesnt converge to
an analytic function on B(z, ).
Therefore F = U U0 and J = C(0, 1). The Julia set is closed, bounded and
f (J) = J.
Theorem 6.31. The Julia set consists precisely of those points z C such that
for every  > 0 there exists w B(z, ) such that f n (w) +.
Proof. Assume that for every  > 0 there exists w B(z, ) such that
f n (w) +. Since f n (z) J is bounded, we cannot have f n : B(z, ) C
isnt normal and thus z J .

7. RIEMANN MAPPING THEOREM

47

Conversely, given z if there exists  > 0 such that for all w B(z, ) we
have that f n (w) is bounded. Thus we can apply Montels theorem to deduce that
f n : B(z, ) C is normal, i.e., z 6 J .

Exercise: If f is a polynomial and z C then either f n (z) or {f n (z) : n
0} is a bounded set.
Theorem 6.32. The Julia set is equal to the closure of the repelling periodic
points.
The proof uses the other version of Montels theorem.
7. Riemann Mapping Theorem
One of the most interesting results in complex analysis is the Riemann Mapping
Theorem.
7.1. Riemann Mapping Theorem. The most elegant theorem in complex
analysis is the following.
Theorem 7.1 (Riemann Mapping Thoerem). Let U C be an open subset
homeomorphic to D. Then there exists an analytic bijection f : U D.
Proof. The proof is quite elaborate, so we break it down into steps. We
assume for simplicity that U is bounded, i.e., U B(z0 , R) for some R > 0.
Step 1:
that
(1)
(2)
To show

Fix z0 U and let F be the family of analytic functions f : U D such


f is a bijection; and
f (z0 ) = 0.
that F 6= we can construct f : U D by
f (z) =

(z z0 )
2R

then


(z z0 ) |z| + |z0 |
R+R

|f (z)| =

1
2R
2R
2R
and by construction we see that f (z0 ) = 0 and thus f F.
Step 2: Since the functions in F are analytic and bounded (i.e., for f F and
|f (z)| 1) we can apply Montels theorem to deduce that the family is normal.
Step 3: We define
M = sup{|f 0 (z0 )| : f F}.
Claim: We want to show that this is finite (i.e., M < +).

48

CONTENTS

Proof of Claim: Choose r > 0 sufficiently small that B(z0 , r) U . By Cauchys


theorem we have that for any f F,



1
1 Z
1
f (z)


0
|f (z0 )| =
dz
sup |f (z)| .
2
r zD
2i |zz0 |=r (z z0 )
r
{z
}
|
1

Thus the result follows with M

1
r.

Step 4: We now want to show the supremum is realized.


Claim: There exists f0 F such that f00 (z0 ) = M .
Proof of Claim: We can chose fn F, n 1, such that |fn0 (z0 )| M as n +.
By Montels Theorem applied to {fn } we can choose a subsequence {fnk }
k=1 and
an analytic function f : U C such that fnk f uniformly on any compact set
K C. Therefore, we see that |f 0 (z0 )| = M .
If we replace f (z) by ei f (z) then we can choose 0 < 2 so that f 0 (z0 ) = M .
Step 5: We want to show that f : U D is injective.
Claim: f : U D is injective.
Proof of Claim: Let z1 6= z2 be distinct points in U . Let 0 < = |z1 z2 | and
define gk : B(z2 , ) C by
gk (z) = fnk (z) fnk (z1 )
Since the functions fnk ( F) are all injective, the functions gk are all non-zero on
B(z2 , ) (i.e., 6 z B(z2 , ) with f (z) = 0).
By the Hurwitz-Weierstrauss theorem the limit function g(z) = f (z) f (z1 ) is
either identically zero or has no zeros. But, it cannot be identically zero (since we
assume that |f 0 (z0 )| = M > 0). Thus f (z) 6= f (z1 ) for all z B(z2 , ) and thus
f (z2 ) 6= f (z1 ).
Step 6: We want to show that f : U D is surjective.
Claim: f : U D is surjective.
Proof of Claim: Assume for a contradiction that f is not surjective, and thus there
exists w D which is not in the image f (U ).
We can choose a M
obius map w : D D by
zw
w (z) =
1 zw
which maps w to 0. We then consider w f : U D. This map now doesnt have
0 in its image. As we saw in a previous lemma, we can write w f (z) = ek(z) where
k : U C is analytic, and thus we can define g : U C by
g(z) = (w f (z))1/2 (=: ek(z)/2 ).
In particular, g : U D is one-to-one. Finally, we introduce
g(z0 ) (z) =

z g(z0 )
1 zg(z0 )

and define : U D by (z) = g(z0 ) g(z0 ).


We can then compute
0 (z0 ) = 0g(z0 ) (g(z0 )) .g 0 (z0 )
|
{z
}
1
= 1|g(z

0 )|

(1)

7. RIEMANN MAPPING THEOREM

and
g(z0 )2

0

49

= 2g 0 (z0 ).g(z0 )
= 0w (f (z0 )).f 0 (z0 )

(2)

= (1 |w|2 )f 0 (z0 ).
(after a little calculation).)
From (1) and (2) we find that:


1 |w|2
1
f 0 (z0 )
1 |g(z0 )|2
2g(z0 )


1
1 |w|2
=
f 0 (z0 )
1 |w|
2g(z0 )


1 + |w|
f 0 (z0 )
=
2g(z0 )
p
However, since w =
6 0 then 1 + |w| > 1 and g(z0 ) = |w| and thus
!
1 + |w|
0
p
(z0 ) =
f 0 (z0 )
2 |w|
|
{z
}
0 (z0 ) =

>1

giving | (z0 )| > M , a contradiction to the definition of M .


However, this contradicts h(z) having maximum derivative M at z0 .

Remark 7.2. If U is unbounded then we can apply a more elaborate argument


to prove Step 1. Let z0 C U . The function : U C is non-vanishing and we
can write (z) = ek(z) = k(z)2 , where k(z) = ek(z)/2 . Moreover, k(z) is one-to-one
(since (z) is) and there are not distinct points z1 , z2 such that h(z1 ) = h(z2 ) (since
then (z1 ) = (z2 )). This is an open mapping and so we can choose B(b, r) h(U ).
But then B(b, r) h(U ) = .
We may therefore define the holomorphic function
r
f (z) =
.
2(h(z) + b)
Since |h(z) h(z)| r for z U , it follows that f : U D. Since h is one-one
then so is f . Compositing f with a Mobius map that preserves D we can get a
function which is analytic and one-on-one and bounded by 1.
Proposition 7.3 (Caratheodorys Theorem). Let U C be a simply connected
open set whose boundary is a simple closed curve. Let f : U D be the analytic
bijection in the Riemann mapping theorem. Then extends to a homeomorphism
from U to D.
7.2. Christoffel-Schwarz Theorem. The Christoffel-Schwarz Theorem is a
more explicit version of the Riemann mapping theorem where the open region to
be mapped to the unit disk is taken to be the inside of a polygon P.

50

CONTENTS

We recall that the map w = f (z) = i 1z


1+z maps the unit disk to the upper half
2
plane H .
Thus it suffices to find a map g : H P which is analytic and then gf : D P.
The general result is the following:
Theorem 7.4 (Schwarz-Christoffel transform). Let P be a polygon with n sides
such that the internal angles are 1 , , n . There exist w1 , , wn R such
that if we consider the function
C
(1)
F (w) =
(w w1 )11 (w w2 )12 (w wn )1n
Rz
then for z0 H we can choose A, C such that f (z) := A+ z0 F (w)dw is an analytic
bijection from H to the interior of the polygon P (and which extends continuously
to the boundary, taking points w1 , , wk to the vertices of P).
Here C changes the size and orientation of the image and A translates it.
Remark 7.5. Since the external angles of the polygon sum to (n 2) we see
that
n
X
(1 i ) = 2.
i=1

Remark 7.6. It is often convenient to consider the case in which the point at
infinity of the w plane maps to one of the vertices of the polygon (e.g., x1 = and
the vertex with angle 1 ). If this is done, the first factor in the formula is effectively
a constant and may be regarded as being absorbed into the constant C
The following general result is often useful. Let H = {z C : Im(z) > 0} and
H = {z C : Im(z) < 0}.
Lemma 7.7 (Schwarz-reflection principle). If g : H C is analytic and extends
continuously to [a, b] R such that g([a, b]) R then g has an extension as an
analytic function to H [a, b] H .

Proof. For z H we have that z H and we define g(z) = g(z). To deduce


analyticity we can use Moreiras theorem. For example, If is a closed curve that
crosses [a, b] then we can consider two loops 1 , 2 consisting of the pieces of
in each half-plane plus a subinterval of [a, b] between their intersection
R points. By
approximating them by curves wholly in the half planes we see that 1 f (z)dz = 0
R
R
R
R
and 2 f (z)dz = 0, and thus f dz = 1 f (z)dz + 2 f (z)dz = 0. Thus Moreiras
Theorem applies and we are done.


7. RIEMANN MAPPING THEOREM

51

Lemma 7.8 (Osgood-Caratheodory Theorem). If g : P D is an analytic


bijection then it extends to a continuous function g : P D on the boundary.
Proof. Assume that i 1.
(1) Choose one side I1 of P and use a Mobius map f1 : P H such that
S := f1 (P) and f1 (I) = [1, 1]
(2) We can reflect S in the real axis to get S and consider U := S (1, 1)S.
By the Riemann mapping theorem we can find a unique analytic bijection
f2 : U D with f2 (0) = 0 and f20 (0) > 0. However, we see that g2 (z) :=
f2 (z) satisfies the same conclusion and so g2 = f2 and we conclude that
f2 ([1, 1]) R and then f2 (S) = D+ = {z D : Im(z) > 0}
(3) We can choose a Mobius map f3 : D+ D. Then we set f (z) = f3 f3
f1 (z). We thus see that f extends continuously to I1 .
If we fix z1 P then we can choose f3 so that f (z1 ) = 0 and f 0 (z1 ) > 0.
Thus f is unique. In particular, the conclusion that f extends applies to
each side.


This brings us to the main part of the Schwarz-Christoffel theorem.
Theorem 7.9. Every analytic bijection f : H P is necessarily of the form
(1).
The existence of such maps follows from the Riemann Mapping Theorem.
The mapping f (z) extends continuously to the boundary. We can then extend
it to H [] via one of the intervals (xi , xi+1) by the Schwarz reflection principle.

In particular, f (H) corresponds to a polygon P0 obtained from P by reflection in the corresponding side. Using the reflection principle (for another interval

52

CONTENTS

(xj , xj+1 )) gives another brach of the function, which we denote by f1 (z). However,
00
f 00 (z)
(z)
the function ff 0 (z)
= f10 (z) is uniquely defined.
1

We can observe that g(z) = (f (z) f (xi )) i maps the interval (xi , xi + )
onto another line segment. Thus again by the reflection pruinciple we see that g(z)
is analytic. We can then write that f (z) = f (zi ) + g(z)i , for z near to xi and thus
f 00 (z)
(i 1)
f 00 (z)
=
+ hi (z) = 10
0
f (z)
z xi
f1 (z)
where hi : H C is analytic.

Moreover, we can write globally that


k

f 00 (z) X (i 1)
=
+ h(z)
f 0 (z)
z xi
i=1
where h : H C is analytic - and h extends to an entire function. However, w can
deuce that it is the constant function.
Schwartz - Christoffel for D. We can reformulate this in terms of the unit
disk (to look more like the Riemann Mapping Theorem, except that it goes in the
wrong direction).
Let D = {z C : |z| < 1} denote the unit disk. Let P be an open bounded region in the complex plane with polygonal boundary and vertices V = {v1 , , vn },
say. The Schwarz-Christoffel theorem gives a (fairly) explicit formula for a holomorphic bijection : D P.
Corollary 7.10. More precisely, the map : D P defined by:
1 1

k 1
Z z

(z) = A + C
1
1
d.
w1
wk
z0

(1)

where
(1) The exponents 1 , , k correspond to the internal angles of P;
(2) the points W = {w1 , , wk } D on the unit circle correspond to preimages of the vertices V ;

7. RIEMANN MAPPING THEOREM

53

7.3. examples.
Example 7.11 (Maps to a semi-infinite strip). Consider a semi-infinite strip P
in the z-plane. (This may be regarded as a limiting form of a triangle with vertices
0, i, and R (with R real), as R tends to infinity). Now 1 = 0 and 2 = 3 = /2
in the limit.

Suppose we guess that x1 = 1 and x2 = 1 and then we are looking for the
mapping f : H P with f (1) = i, f (1) = 0 (and f () = ). Then f (z) is
given by
Z z
Z z
C
C
p
p
f (z) = A +
dw = A +
dw
(w 1)(w + 1)
(w2 1)
z0
z0
In this special case the integral can be explicitly evaluated to get
f (z) = A + C cosh1 (w)
where A, C C. Requiring that f (1) = i and f (1) = 0 gives A = 0 and C = 1.
Hence the Schwarz-Christoffel mapping is given by f (z) = cosh1 (z).
Example 7.12 (Maps to another semi-infinite strip). Consider the semi-infinite
strip
S = {x + iy : a < x < a and y > 0}
for a > 0. Find an explicit form for the holomorphic bijection f : H S from
the upper half plane to S for which extends to f (1) = a and f () = .

54

CONTENTS

The strip has internal angles /2. We take the two points to be mapped to the
a to be 1. By the Christoffel-Schwarz Theorem this takes the form
Z z
Z z
1
1
dw

f (z) = A
(w 1) 2 (w + 1) 2 dw + B = A
+B
1 w2
0
0
We can evaluate the indefinite integral:
f (z) = B + A sin1 (z)
The image a = f (1) of the first vertex 1 gives that
a = B + A sin1 (1) = B

A
.
2

The image a = f (1) of the second vertex 1 gives that


a = B + A sin1 (1) = B +

A
.
2

from which we deduce that B = 0 and thus A = 2a/ and thus


f (z) =

2a sin1 (z)

Example 7.13 (An equilateral triangle). Let f : H C be defined by


Z z
dw
f (z) =
.
(w(1

w))2/3
0

We claim that the image is an equilateral triangle with vertices at 0, 1, 12 + 23 .


The internal angles are all equal to 3 . It we consider the real axis as the boundary
of H then we see that the interval [0, 1] is mapped to [0, K],say, since the integrand
R1
dt
is real. Here K = 0 (t(1t))
2/3 = 5.299 .

3
The interval [1, ) is mapped to the side [K, K
2 + i 2 K) and the interval

3
(, 0] is mapped to the side ( K
2 + i 2 K, 0]
More generally, if the triangle with angles 1 , 2 and (1 1 2 ) then let
g : H C be defined by
Z z
dw
f (z) = A + C
.
1
1 (w + 1)12
0 (w 1)

(i.e., x1 = 1, x2 = 1 (and x3 = ). The image is a triangle with angles 1 , 2


(and (1 1 2 )).

55

Example 7.14 (A right angle triangle). Consider the triangle in C with vertices at 1, 0 and i and find the analytic bijection f : H which extends to

f (1) = i, f (1) = 1 and f () = 0.

Show that
f (z) =

1 i
2

(z 2 1)3/4 dz +

where

i1
2

(z 2 1)3/4 dz

=
0

The vertices i and 1 have internal angles of 4 and are the images of 1 and 1,
respectively, We can use the Christoffel Schwarz Theorem to write
F (z) = (z + 1)3/4 (z 1)3/4 = (z 2 1)3/4
We have that

Z
f (z) = A

(z 2 1)3/4 dz + B

and then
Z

1
2

(z 1)

i=A

3/4

Z
dz + B and 1 = A

(z 2 1)3/4 dz + B.

If we denote
Z
=
0


Z
(z 2 1)3/4 dz =

(z 2 1)3/4 dz

then A + B = i and A + B = 1 and we can solve for A =


Hence
Z
1 i z 2
i1
.
f (z) =
(z 1)3/4 dz +
2
2
0

1i
2

and B =

Example 7.15 (Squares and Rectangles). Let g : H C be defined by


Z z
dw
f (z) =
.
1/2
0 (w(1 w)( w))
where > 1. The image is a rectangle vertices at 0, K, K iK 0 , iK 0 .

i1
2 .

56

More precisely, the internal angles are all equal to 2 . It we consider the real
axis as the boundary of H then we see that the interval [0, 1] is mapped to the side
[K, 0] where
Z 1
dw
K=
.
(w(1

w)(
w))1/2
0
For 1 < w < there is a single imaginary root and thus the integral from
[1, ] is purely imaginary (with negative imaginary part) and the image of [1, ] is
[K, K iK 0 ] where
Z
1
dw.
K0 =
1/2
1 (w(w 1)( w))
Similarly, we see that the image of [, ) is [K iK 0 , iK 0 ) and the image of
(, 0] is (iK 0 , 0].
The upper half plane H is mapped to the square by
Z
dw
p
z = f (w) =
w(w 1)(w + 1)
7.4. Generalizations of the Schwarz-Christofel theorem. More generally one can consider an analogous problem where we introducce holes into the
picture. Let z1 , . . . , zd D be the centres of subdisks Di = B(zi , ri ) = {z
D : |z zi | < ri } where |zi | + ri < 1. We can then associate a multiply connected
region defined by M = D di=1 Di . When d = 1 this simply corresponds to an annulus. Similarly, we can consider sub-polygons P1 , , Pd P with sets of vertices
V1 , , Vd , respectively. We then let Q = P di=1 Pi
Let Ci = Di and let C i denote the images on reflection in the unit circle
r2 z

j
{z C : |z| = 1} = D. In particular, gi (z) = zi + 1z
satisfies gi (C i ) = Ci . Let
jz
= hg1 , , gd i be the associated Schottky group. By a result of Crowdy et al,
there is an explicit formula for a holomorphic bijection : D P given by

(z) =

C()
z0

k
Y
i=1

w(z, wi )i

|Vi |
d Y
Y

(i)

(j)

w(z, wj )i dz.

(2)

i=1 j=1

where w(, ) is the Schottky-Klein prime function and


(1) the exponents 1 , , k correspond to the internal angles of P and ex(j)
(j)
ponents 1 , , n|Vi | correspond to the internal angles of P(j) ;
(2) the points W = {w1 , , wk } D on the unit circle correspond to
(i)
(i)
preimages of the vertices V and the points W (i) = {w1 , , w|Vi | } Di
on the circle correspond to preimages of the vertices Vi ;
(3) C() is an explicit correction function.
7.5. Application: Fluid dynamics. The streamlines around a circular region of diameter 2a can be described by transforming the horizontal lines past a
slit of width 4a by the associated conformal map f (z) = z + a2 /z. More generally, given a finite number of disjoint circular regions the stream lines described by
transforming the horizontal lines past a slits. ??
8. Behaviour of analytic functions
There are very useful tools in complex analysis: The Maximum Principle, and
the Schwarz and Pick Lemmas

8. BEHAVIOUR OF ANALYTIC FUNCTIONS

57

8.1. The Maximum Modulus Principle. The next theorem shows that the
supremum of the modulus of an analytic function isnt achieved inside U .
Theorem 8.1 (Maximum Modulus Principle). Let f : U C be analytic. Let
U be connected. If z0 U satisfies |f (z0 )| |f (z)| for all z U then f (z) is
necessarily a constant function.
Proof. By multiplying by a constant, if necessary, we can assume without loss
of generality that f (z0 ) R+ . Let
S = {z U : |f (z)| = f (z0 )}.
Since f is continuous we see that S U is a closed set.
We claim that S U is also open. Given z S U , since U is open we can
choose r > 0 sufficiently small that B(z, r) U . Then for any 0 < r0 < r we have
by Cauchys Theorem that


1 Z
f ()

d
f (z0 ) = |f (z)| =
2i C(0,r0 ) z


Z 2
1
f (z + r0 ei ) i

=
(ie )d
2i 0
r0 ei


Z 2
1

=
f (z + r0 ei )d
2i 0
Z 2


1
f (z + r0 ei ) d f (z0 ) = |f (z)|,

2 0
where |f (z + r0 eit )| |f (z)| = f (z0 ), since z S. Thus all the inequalities must be
equalities and so
|f (z + r0 ei )| = f (z0 )
for all 0 < 2 and all 0 < r0 < r. In particular, S contains the open ball
B(z, r) S. But since S is both open and closed and U is connected we deduce
that S = U and thus f takes the constant value f (z0 ).

8.2. The Schwarz Lemma. The next result describes analytic maps of the
disk D = {z C : |z| < 1} to itself.
Theorem 8.2 (Schwarz Lemma). Let f : D D be an analytic map.
(1) Let f (0) = 0 then:
(a) |f (z0 )| |z0 |, z0 D; and
(b) |f 0 (0)| 1.
(2) Moreover, if either |f (z)| = |z| for some z D {0} or |f 0 (0)| = 1 then
f is a rotation of the form f (z) = ei z, for some 0 < 2.
Proof. (1) The function g : D C defined by
(
f (z)
if z 6= 0
z
g(z) =
f 0 (0) if z = 0
is analytic on D. Fix  > 0. We can apply the Maximum Modulus Principle to the
restriction g : B(0, 1 ) C to deduce that for |z| 1  we have that
1
|g(z)|
.
1
Letting  0 we deduce that |g(z)| 1 whenever |z| 1. From the definition of g
this completes the proof of the first part.
(2) For the second part, if |f (z0 )| = |z0 | then |g(z0 )| = 1. The maximum
principle principle applied to g now tells us that g has a constant value, ei , say.
This implies that f (z) is a rotation.

58

CONTENTS

Similarly, if |f 0 (0)| = |g(0)| then the maximum principle principle applied to g


again tells us that g has a constant value, ei , say.

Application 8.3. We denote a Mobius transformations a : D D by
a (z) =

za
1 az

where a D. We can also denote rotations : D D by (z) = ei z, for


0 < 2.
Proposition 8.4. Any analytic function f : D D which is a bijection is of
the form f = a , for some a D and 0 < 2.
Proof. We first observe that a (D) = D. If |z| = 1 then



z a z a



= 1.
|a (z)| =
=
1 az z a
By the Maximum Modulus Principle, since a is not a constant function if |z| < 1
then |a (z)| < |z| < 1. In particular, a (D) D. Observe that 1
a = a , showing
that a : D D is a bijection.
Let us denote f (0) = b, say. Consider g = b f , which is now an analytic
function g : D D for which g(0) = 0. By the Schwarz Lemma, |g 0 (0)| 1.
Similarly, we can apply the Schwarz lemma to g 1 : D D to deduce that
|(g 1 )0 (0)| = |g01(0)| 1. We therefore have |g 0 (0)| = 1. By the uniqueness part of
the Schwarz Lemma we have that g(z) = for some 0 < 2. But then this
means that f (z) = b .

8.3. Picks lemma. There is a generalization of Schwarzs lemma in which
we dont require f (0) = 0
Theorem 8.5 (Picks Lemma). Let f : D D be analytic.
(1) For z1 , z2 D,



f (z1 ) f (z2 ) z1 z2



1 f (z1 )f (z2 ) 1 z1 z2

()

and
|f 0 (z1 )|

1 |f (z1 )|2 )
1 |z1 |2

()

(2) If there is either an equality in (*) for z1 6= z2 , or an equality in (**) for


some z1 , then f must be an analytic bijection.
Proof. Let
(z) =

z1 z
f (z1 ) z
and (z) =
.
1 z1 z
1 f (z1 )z

Then f satisfies the hypothesis of the Schwarz Lemma. Thus for any z D,
| f (z)| |z|.
Setting z = 1 gives the inequality (1).
The Schwarz Lemma also gives that
|( f )0 (z)| 1.
Using the chain rule gives the inequality (2).

9. HARMONIC FUNCTIONS

59

9. Harmonic functions
We want to consider the real valued functions.
Definition 9.1. We say that any function u : U R is harmonic if u = 0
where
2
2
=
+
x2
y 2
Remark 9.2. Any scaler multiple of linear combinations of harmonic functions
are harmonic.
In particular, if f : U C is analytic and f (x + iy) = u(x, y) + iv(x, y) satisfy
the Cauchy-Riemann equations then differentiating
v
u
=
x
y
with respect to x implies that
2u
2v
=
x2
yx
and differentiating
u
v
=
y
x
with respect to y implies that
2u
2v
=

.
y 2
yx
Thus we have shown:
Theorem 9.3. We have that u = Re(f ) is harmonic. Similarly, we see that
v = Im(f ) is harmonic.
Example 9.4. The function f : C C given by f (z) = z 2 is analytic and thus
f (z) = (u(x, y) + iv(x, y))
= (x + iy)2
= x2 y 2 + 2ixy.
where u(x, y) = x2 y 2 and v(x, y) = xy are harmonic.
Example 9.5. Let f : U C be analytic and for z = x + iy and write
f (z) = u(z) + iv(z). Let w(z) := log |f (z)| = 21 log(u(z)2 + v(z)2 ) then


w
1
u
v
=
u(z)
+
v(z)
x
u(z)2 + v(z)2
x
x


w
1
u
v
=
u(z)
+
v(z)
y
u(z)2 + v(z)2
y
y
and
2w
2
=
2
2
x
(u(z) + v(z)2 )2

2w
2
=
y 2
(u(z)2 + v(z)2 )2

u
v
u(z)
+ v(z)
x
x

2

u
v
u(z)
+ v(z)
y
y

2

1
+
2
u(z) + v(z)2

1
+
u(z)2 + v(z)2

u
x

2

u
y

2

v
x

2

2u
2v
+ u(z) 2 + v(z) 2
x
x

v
y

2

2u
2v
+ u(z) 2 + v(z) 2
y
y

Adding these, expressions, the second derivative terms vanish since u and v are
harmonic. The first order terms vanish by the Cauchy-Riemann equations.
Remark 9.6. The equation u = 0 is called Laplaces equation.

60

CONTENTS

9.1. Harmonic conjugates.


Definition 9.7. If u, v are harmonic and satisfy the Cauchy-Riemann equations in u then v is called the harmonic conjugate of u.
Conversely, if u is harmonic then it is the real part of an analytic function.
Theorem 9.8. If u : U R is a C 2 function there there exists a harmonic
conjugate, i.e., there exists a C 2 function v : U R such that the function f : U
C defined by f = u + iv is analytic.
Proof. Consider the case
U = {z = x + iy : x (a , a + ) and y (b , b + )},
and let f, g : U R be C 1 functions defined by
u(z)
u(z)
f (z) =
and g(z) =
.
y
x
Then since u = 0 we have that


g
2u
2v
f
=
= 2 =
.
y
x
y
x2
If we define
Z x
Z y
v(x, y) =

f (t, b)dt +
a

g(x, s)ds

(1)

(2)

then v : U R is a C 2 function such that h


y = g. The derivative with respect to
x is a little more complicated. We can write
Z x

f (t, b)dt = f (x, b).


(3)
x a
For the second term in (2) we have that
Z y
Z y
Z y

f
g
g(x, s)ds =
(x, s) ds =
(x, s)ds = f (x, y) f (x, b).
x b
b y
b
|x {z }
= f
y

by

(4)

(1)

Comparing (2), (3) and (4) gives that


v
= f = f (x, b) + (f (x, y) f (x, b)) = f (x, y).
x
v
u
v
Finally, u
x = g = y and y = f = x .

Exercise 9.9. Given u(x, y) = xy find the harmonic conjugate of u. In particular,
u
v
1
=y=
= v(x, y) = y 2 + f (x)
x
y
2
v
u
1
= f 0 (x) =
= x = f (x) = x2 + C
x
y
2
Therefore v(x, y) = 21 (x2 + y 2 ) + C and
f (z) = u + iv
= xy + i

(x2 + y 2 )
+C
2

i
= (x + iy)2 + C
2
i
= z 2 + C.
2

9. HARMONIC FUNCTIONS

61

which is analytic.
One can always find a harmonic conjugate on a disk, but not on any general
region.
Example 9.10 (Non-example). Consider the function u(x, y) = log(x2 + y 2 )1/2
on the domain U = C {0}. Then
u
x
= 2
x
x + y2
u2
(x2 + y 2 ) 2x2
y 2 x2
=
= 2
2
2
2
2
x
(x + y )
(x + y 2 )2
and by symmetry
u2
x2 y 2
= 2
2
y
(x + y 2 )2
Thus u = 0, i.e., u is harmonic.
Assume for a contradiction we can find v such that f (z) = u + iv is analytic.
Then
u
u
v
u
x
iy
1
1
f 0 (z) =
i
=
+i
= 2
2
=
=
x
y
y
x
x + y2
x + y2
x + iy
z
Thus f (z) = log z + C. But being analytic the function must be continuous contradicting the jump between branches.
9.2. Poisson Integral formula. A harmonic function on the unit disk can
be written be written in terms of its values on the unit circle.
Theorem 9.11 (Poisson integral formula). Let z D and assume that u : U
C be harmonic and U D. We can write
Z 2
1
u(rei ) =
Pr ( t)u(eit )dt
2 0
where
Pr () =

1 r2
.
1 2r cos + r2

Proof. Let f (z) = u(z)+iv(z) be the associated analytic function, where v(z)
is the harmonic conjugate of u(z).
Given z D, let z1 = |z|z 2 C D be the reflection in the unit circle. By
Cauchys theorem (applied twice)
Z
Z
f ()
1
f ()
1
d
d
f (z) =
2i C(0,1) z
2i C(0,1) z1
|
{z
} |
{z
}
=f (z)

=0

62

CONTENTS

Let us denote = eit and z = rei and z1 = 1r ei and then with this parameterization we can write


Z 2
1
1
1
it
f (z) =
f (e )
1 i
ieit dt
it
2i 0
rei eit
e

e
r


Z 2
1
1
1
it
f (e )
dt
=

1 i(t)
2 0
rei(t) 1
1
re
!
Z 2
1 1r ei()
1
1 rei(t)
it
=
f (e )

dt
2 0
1 + r2 2r cos( t) 1 + r12 2r cos( t)


Z 2
1
1 rei(t)
r2 rei(t)
=
f (eit )

dt
2 0
1 + r2 2r cos( t) r2 + 1 2r cos( t)


Z 2
1 r2
1
it
f (e )
dt
=
2 0
1 + r2 2r cos( t)
{z
}
|
R

where in the middle we multiply both expressions by the complex conjugates of


their respective denimonators. Taking real and imaginary components, the result
follows.

Remark 9.12. The Poisson kernel can be written in different ways. For example, we can write for z = rei
Pr ( t) =

1 |z|2
.
|1 zeit |2

Remark 9.13. When z = 0 we have the mean value property:


Z 2
1
u(0) =
u(eit )dt.
2 0
Application 9.14. The Laplace equation appears in the heat equation, one
physical interpretation of this problem is as follows: fix the temperature on the
boundary of the domain according to the given specification of the boundary condition. Allow heat to flow until a stationary state is reached in which the temperature
at each point on the domain doesnt change anymore. The temperature distribution in the interior will then be given by the solution to the corresponding Laplace
problem.
Harmonic functions have various restrictions on their values.
Corollary 9.15 (Harnack). Assume that u : U R is harmonic and U D.
Then for any z D we have that
1 |z|
1 + |z|
u(0) u(z)
u(0).
1 + |z|
1 |z|

9. HARMONIC FUNCTIONS

63

Proof. We observe that for any 0 t < 2 and z D:


1 |z|2
1 + |z|
1 |z|2

=
.
it
2
|e z|
(1 |z|)2
1 |z|
Thus by the Poisson Integral Theorem
Z 2
1 + |z| 1
1 + |z|
u(eit )dt =
u(z)
u(0).
1 |z| 2 0
1 |z|
The other inequality is proved similarly.

The following is the analogue of Montels theorem for analytic functions.


Corollary 9.16 (Harnacks Principle). Assume that uj : U R, j 1, are
harmonic functions with
u1 u2 u3
Then either uj uniformly on compacts sets or there is a harmonic function
u : U R such that uj u uniformly on compact sets.
This is an exercise.
Corollary 9.17 (Jensen). Let f : U R be analytic and let U D. Let
a1 , , aN D be the zeros of f (z) with |aj | < 1, for j = 1, , N , and assume
there are no zeros on C(0, 1). Then
Z 2
N
X
1
log |f (0)| =
log |aj | +
log |f (eit )|dt
2
0
j=1
Proof. If f (z) has no zeros then u(z) := log |f (z)| is harmonic and by the
Poisson Integral Formula (at z = 0) we have that
Z 2
1
log |f (eit )|dt.
(1)
log |f (0)| =
2 0
More generally, if f (z) has zeros a1 , , aN D then we can define

N
Y
1

za
j
f (z)
F (z) :=
z

a
j
j=1
which is now a zero free analytic function. We can apply (1) to F (z) to deduce that

N
X
1
log |F (0)| = log |f (0)| +
log
ai
i=1
Z 2
1
log |F (eit )|dt
=
2 0



Z 2
N
X
1 ai eit
1


=
log it
log |f (eit )| +
dt
2 0
e aj
i=1
| {z }
=1

1
=
2
since the maps z 7

1zai
zai

log |f (eit )|dt

preserve the unit circles.




Application 9.18 (The Dirichlet problem). We can construct harmonic functions on D from a continuous function on the unit circle.

64

CONTENTS

Theorem 9.19. Let U : C(0, 1) R be continuous on the boundary of the unit


disk D. Then the function u : D R defined by
(
U (z)
if z C(0, 1)
u(z) = 1 R 2
it
if z = rei D
2 0 Pr ( t)f (e )dt
is harmonic (i.e., u = 0)
Proof. We can rewrite
Pr ( t) =

1 |z|2
eit
eit
=
+
1
|z eit |2
eit z
eit z

where z = rei . Then


1
u(z) =
2

Pr ( t)U (eit )dt


0

Z 2  it
1
e
eit
=
+ it
1 U (eit )dt.
2 0
eit z
e z
For the first integral we have observe that


Z 2 
Z 2 
eit
2eit
2
it
U (e )dt =
U (eit )dt
x2 0
eit (x + iy)
(eit (x + iy))3
0
and


Z 2 
Z 2 
2
eit
2eit
it
U
(e
)dt
=
U (eit )dt
it (x + iy))3
y 2 0
eit (x + iy)
(e
0
and thus adding these two expressions gives

Z 
eit
U (eit )dt = 0.

eit (x + iy)
The second integral is similarly. The final term is a constant. Thus three terms are
harmonic, and thus so is the sum.
To see that u(z) is continuous (on the boundary) we can write

Z 2
Z 2 
1
1
1 |z|2
it
U (eit )dt
Pr ( t)U (e )dt =
2 0
2 0
|z eit |

Z 2  it
1
e +z
=
U
dt
2 0
1 + zeit
w+z
using the change of variables z (w) = 1+zw
(with |0z (w)| =
z D. Moreover, since for each C(0, 1) we have
 it

e +z
lim
=
z 1 + zeit

1|z|2
|1+zw|2 )

for each

if 6= eit . The result follows from bounded convergence of integrals.

9. HARMONIC FUNCTIONS

65

More generally, we see that given a simple closed curve and a continuous
function U : C we can find a harmonic function u inside with agrees with U
on . It suffices to solve the problem for the unit circle, as above, and then use the
Riemann mapping theorem.
However, for the circle we can use the Poisson integral formula.
9.3. Mean Value Property and Maximum Principle for harmonic
functions. The following is a characteristic property of harmonic functions.
Theorem 9.20 (Mean Value Property). Let u : U R be harmonic. For any
B(z, r) U we have that
Z 2
1
u(z + rei )d.
u(z) =
2 0
for any sufficiently small r > 0.
Proof. Let r1 > r > 0 such that B(z, r1 ) U . Let v : B(z, r1 ) R be
the harmonic conjugate for u : B(z, r1 ) R and denote the analytic function
f : B(z, r1 ) C defined by f (z) = u(z) + iv(z).

By the Cauchy theorem


f (z) = u(z) + iv(z) =

1
2i

Z
C(z,r)

f ()
d
z

Z 2
1
f (z + rei ) i
=
ire d
2i 0
rei
Z 2
Z 2
1
1
=
u(z + rei )d + i
v(z + rei )d
2 0
2 0
Taking real parts gives the result.

This is also the analogue of the maximum principle for analytic functions.
Theorem 9.21. Let u : U R be harmonic. If there exists z0 U such that
u(z0 ) = supzU |u(z)| then the function u is constant.
Proof. We can give two proofs:
First proof (using harmonic conjugates): Let
M = {w U : u(w) = sup |u(z)|}.
zU

Clearly, z0 M =
6 . It suffices to show that M U is open, since U being
connected would imply M = U .
Let z M. For small r > 0 we can consider B(z, r) U . Let v : B(z, r) R
be the harmonic conjugate for u : B(z, r) R and denote the analytic function
f : B(z, r) C defined by f (z) = u(z) + iv(z). We define g(z) = ef (z) then

66

CONTENTS

|g(z)| = supB(z,r) |g()|. We can apply the maximum modulus principle for analytic functions to deduce that g(z) is constant on B(z, r), and thus u(z) = log |g(z)|
is constant on B(z, r). Thus M is open.
Second proof (using mean value property): Let L := supzU |u(z)| and let
M = {w U : u(w) = L}.
By hypothesis z0 M 6= . Since g is continuous we see that M is closed. It
suffices to show it is also open.
Let z M and choose r > 0 such that B(z, r) U and then
Z 2
1
g(z + rei )d L
L = g(z) =
2 0
and thus g(z + rei ) = L for 0 < 2, for all sufficiently small r > 0. We deduce
that M is open.

We can use this to get the converse to the first result.
Theorem 9.22. If h : U R satisfies the mean value theorem then it is
harmonic.
Proof. Assume that B(z, r) U . By using the Poisson Integral Theorem we
have a harmonic function u : B(z, r) R which coincides with h on the boundary
C(z0 , r). It suffices to show that u and h therefore coincide on all of B(z, r).
Let us write g(z) = u(z) h(z), which vanishes on C(z0 , r). Moreover, it
satisfies the mean value theorem since this holds for both u (since it is harmonic)
and h (by hypothesis). By the previous lemma we see that g 0 (since otherwise
the supremum would occur in B(z, r) and this would make the function constant,
giving a contradiction). Replacing w by w we deduce that w = 0, i.e., u(z) = h(z)
on B(z, r). Since the disk is arbitrary we deduce that h(z) is analytic on U .

9.4. Schwarz reflection principle for harmonic functions. We can now
formulate an analogue of the Schwarz reflection principle for functions.
Theorem 9.23. Given a domain U and suppose that U R = (a, b). Let
U + = {z U : Im(z) > 0}. Assume v : U + R is harmonic and for each
z0 (a, b) we have that limzz0 v(z) = 0. Let U = {z : z U } and define

if z U +
v(z)
w(z) = 0
if z (a, b)

v(z) if z U
Then w : U + (a, b) U R is harmonic.
Proof. We claim that the Mean Value Theorem holds on a sufficiently small
disk about any point. If z U + then it follows form u being harmonic. Similarly,
it holds for z U .
If z (a, b) then for sufficiently small r > 0 we have that
Z 2
1
w(z + rei )d
2 0
Z
Z
1
1
=
w(z + rei )d +
w(z + rei+i )d
2 0
2 0
Z
Z
1
1
i
=
w(z + re )d +
w(z rei )d
2 0
2 0
Z
Z
1
1
i
=
w(z + re )d
w(z + rei )d = 0.
2 0
2 0

9. HARMONIC FUNCTIONS

67

since w(x + iy) = w(x iy). Since w(z0 ) = 0 the Mean Value Property holds.


9.5. Application: Continuous extensions to the boundary of analytic
bijections. We can use the Schwarz reflection principle for analytic functions to
give another (and more generalizable) proof that if f : U D is an analytic bijection
where U C is a domain whose boundary consists of straight line segments, or,
more generally, circular arcs, then it must extend continuously to the boundary
(i.e., the Osgood Caratheodory theorem).
We can consider the case that U = P is a polygon, the other cases being similar.
Theorem 9.24. Let f : P D be an analytic bijection then it extends continuously to the boundary.
Proof. It is easy to see that if zn z0 P then |f (zn )| 1 as n +.
(Otherwise we could find a subsequence with f (znk ) accumulating in the interior of
D and then this would pull back to a closed set containing znk , necessarily bounded
1
away from P and
 giving
 a contradiction). Let B(0, 2 ) be the closed ball about 0.

Thus K = f 1 B(0, 12 ) P and we can assume that zn 6 K for sufficiently large


n.
Let us assume that z0 is not a vertex of P then we can consider a small neighbourhood D(z0 , ) := B(z0 , ) P which is a half-disk, provided  > 0 is sufficiently
small. Moreover, provided  > 0 is small enough it is disjoint K and we have that
f : D(z0 , ) C
is non zero. In particular, then write
u(z) + iv(z) = i log |f (z)|.
However, since |f (zn )| 1 we have that v(z) = Im(log |f (zn )|) 0. We can now
apply the Schwarz reflection principle for harmonic functions to v(z)

Application 9.25 (Annuli and eccentricity). Two annuli are conformally equivalent iff the ratio of the outer radius and the inner radius is the same for the two.
Thus each is conformally equivalent to a unique standard annulus r < |z| < 1
with 0 < r < 1.
Theorem 9.26 (Riemann Mapping Theorem for Annuli). There is an analytic
bijection between the annuli A1 = {z C : 1 < |z| < r1 } and A2 = {z C : 1 <
|z| < r2 } if and only if r1 = r2 .
Proof. Assume we have an analytic map f : A1 A2 and thus that f extends
continuously to the closed annuli by the previous result.
By using the Schwarz reflection principle we can show that f extends to an
(1)
analytic map between A1 = {z C : r11 < |z| < r1 } and A12 = {z C : r12 <
|z| < r2 }, i.e., the annuli obtained by reflecting A1 and A2 in the unit circle.

68

CONTENTS
(2)

We can reflect again to show that f extends to an analytic map between A1 =


(2)
{z C : r11 < |z| < r12 } and A2 = {z C : r12 < |z| < r22 }. Continuing iteratively,
(n)

we have an analytic map f : C {0} C {0} (since C {0} =


n=1 A1
(n)

n=1 A2 ).

Moreover, we can assume that limz0 f (z) = 0 and limz f (z) = (the
other cases being similar). In particular, we can deduce that f (z) is a Mobius map
and thus a rotation.

10. Singularities
We have previously considered analytic functions f : U C and their zeros.
What about the possibilities of singularities (where the function isnt analytic)?
10.1. Removable singularities. The first thing to consider is whether we
really have a singularity at all, or if we could just extend the analytic function to
the missing point.
Definition 10.1. We say that f : U C has an isolated singularity at z0 if
f is defined and analytic in a punctured neighbourhood P (z0 , r) = B(z0 , r) {z0 },
but not defined at z0 .
Example 10.2. Let f (z) = 1/z then z0 = 0 is a singularity.
The first result show that at singularities the function blows up, in some sense.
Theorem 10.3 (Removable Singularities Theorem). Assume f : U C is
analytic expect possibly at z0 U . However, if U is bounded on some punctured ball
P (z0 , ) = {z C : 0 < |z z0 | < } then f (z) is analytic at z = z0 as well.

Proof. Since f is analytic and bounded on P (z0 , ) we see




!
(z z )2 Z



f
()


0
(z z0 )2 f 0 (z) =
d 4
sup |f ()|
2
|z |
2i

P (z0 ,)
|z0 |= 20 (z )

(1)

10. SINGULARITIES

69

Thus (z z0 )2 f 0 (z) is bounded on P (z0 , )

We can then define


(
g(z) =

(z z0 )2 f (z)
0

if z 6= z0
if z = z0

then by hypothesis we have limzz0 g(z) = 0 and by (1) we have that



lim g 0 (z) = lim (z z0 )2 f 0 (z) + 2(z z0 )f (z) = 0
zz0

zz0

In particular, g is C 1 , and by Moreiras theorem we can deduce


P that g(z) is analytic.
We can consider the power series expansion g(z) = n=0 an (z z0 )n . Since
g(0) = 0 and g 0 (0) = 0 we deduce that

g(z) =

X
n
a
(z

z
)
an (z z0 )n = (z z0 )2
n
0

n=2
n=2
{z
}
|

f (z)

and thus f (z) = g(z)/(z z0 )2 is analytic, as required.

Example 10.4 (Non-example). Let f (z) = sin(z)/z then this has a singularity
at z0 = 0, but this is a removable singularity.
10.2. Analogues of Cauchys theorem. We next look for analogues of
Cauchys theorem in a neighbourhood of the singularity (and in a neighbourhood
of ). Assume that f is defined on the annulus
A(a, r, R) = {z C : r < |z a| < R}

where r may be zero and R may be +.

70

CONTENTS

Lemma 10.5 (Analogue of Cauchys integral formula). Let f : A(a, r, R) C


be analytic. Let r < r2 < |z0 a| < r1 < R then
Z
Z
1
f (z)
1
f (z)
f (z0 ) =
dz
dz
2i C(a,r1 ) z z0
2i C(a,r2 ) z z0

Proof. Define
(
F (z) =

f (z)f (z0 )
zz0
f 0 (z0 )

if z A {z0 }
if z = z0

then since f is analytic (i.e., complex differentiable) at z0 we deduce that F is continuous on A, and analytic on A {0}. In particular, by the removable singularities
theorem we see that F is analytic. Hence
Z
Z
F (z)dz =
F (z)dz
(1)
C(a,r1 )

C(a,r2 )

(by joining C(a, r1 ) and C(a, r2 ) by a line, along which the integrals in different
directions cancel, and applying Cauchys Theorem).
Observe that that
Z
Z
1
1
dz = 2i and
dz = 0.
(2)
z

z
z

z0
0
C(a,r1 )
C(a,r2 )

Then comparing (1) and (2) gives:


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

z
z

z
z

z0
0
0
C(a,r1 )
C(a,r1 )
C(a,r1 )
Z
f (z)
=
dz 2if (z0 )
z
z0
C(a,r1 )
Z
Z
f (z) f (z0 )
f (z) f (z0 )
=
dz =
dz
z z0
z z0
C(a,r2 )
C(a,r2 )
Z
f (z)
=
dz.
C(a,r2 ) z z0

10. SINGULARITIES

71

Rearranging these equalities gives that


Z
Z
1
f (z)
1
f (z)
f (z0 ) =
dz
dz
2i C(a,r1 ) z z0
2i C(a,r2 ) z z0
as required.

10.3. Laurents Theorem. As a consequence of the last theorem we next


have the analogue of power series expansions for analytic functions.
Theorem 10.6 (Laurents Theorem). Suppose that f : A(a, r, R) C is analytic. Then there exists (an )nZ such that

f (z0 ) =

an (z0 a)n

n=

for all z0 A(a, r, R). Moreover, for any r < r1 < R we can write
Z
f (z)
dz, for n Z.
an =
n+1
C(a,r1 ) (z a)
Moreover, this expansion is unique.
Proof. We present the proof in two parts: First the existence of the expansion
and then the uniqueness.
Existence of expansion. Let r < r2 < |z0 a| < r1 < R then (by joining C(a, r1 )
and C(a, r2 ) by a line, along which the integrals in different directions cancel, and
applying Cauchys Theorem) we have
Z
Z
1
f (z)
f (z)
1
f (z0 ) =
dz
dz
2i C(a,r1 ) (z z0 )
2i C(a,r2 ) (z z0 )
|
{z
} |
{z
}
=:I1

=:I2

where
I1 =

an (z a)n where an =

n=0

1
2i

Z
C(a,r1 )

f (z)
dz for n 0
(z a)n+1

by analogy with the proof of the power series expansion for analytic functions.

The evaluation of I2 is similar, in as much as we can write


1
1

=
z z0
(z0 a) 1
1
=
(z0 a)

za
z0 a


1+

za
z0 a


+

za
z0 a

2
+

72

CONTENTS

for |z a| = r2 < |z0 a|. Thus


Z

X
1
f (z)
I2 =
dz =
an (z0 a)n
2i C(a,r2 ) (z z0 )
n=1
where
an

1
=
2i

n1

f (z)(z a)
C(a,r2 )

1
dz =
2i

f (z)(z a)n1 dz

C(a,r1 )

since (z a)n1 f (z) is analytic, which is justified by by the the uniform convergence
of the geometric series and integration. Hence

X
f (z0 ) = I1 I2 =
an (z0 a)n
n=

where an is as given.

Uniqueness
P of expansion. To see the uniqueness of the coefficients, suppose
f (z) = n= bn (z a)n then dividing by (z a)k+1 and integrating around a
circle C with radius between r and R, i.e.,


Z
Z

X
f (z)
1
1
nk1
dz =
bn
(z a)
dz = bk
2i C (z a)k+1
2i C
n=
showing uniqueness.

Example 10.7. We can consider the function


1
1
1
f (z) =
=

(z 1)(z 2)
(z 2) (z 1)
on various domains. Observe that we can expand
( P

z n if |z| < 1
1
= P n=0n
z1
if |z| > 1
n=1 z
and
1
=
z2

( P

n=0 2n1 z n
P n1 n
z
n=1 2

if |z| < 2
if |z| > 2.

Case I: For z D(0, 1) we have an expansion


f (z) =

(1 2n1 )z n .

n=0

This is a power series which is analytic (and with no negative powers involved),
Case II: For z A(0, 1, 2) we have a Laurent series expansion
f (z) =

z n

n=0

2n1 z n .

n=0

Case III: For z A(0, 2, ) (i.e., |z| > 2) we have a Laurent series expansion
f (z) =

X
n=0

z n

(2n1 1)z n

n=0

This brings us to the following estimate.


1The integrals around C(a, r ) and C(a, r ) can now be replaced by integrals around any
1
2
circle of radius between r and R (by connecting the circles C(a, r1 ) and C(a, r2 ) by line segments
and integrating in both directions (so that these integrals will cancel).

10. SINGULARITIES

73

Corollary 10.8. Let f : A(0, r, R) C be analytic and have a Laurent series

f (z) =

an (z a)n

n=

then for any r < r1 < R we can bound |an |

M (r1 )
r1n

where M (r) = sup|za|=r1 |f (z)|.

Proof. Since
1
an =
2i

Z
C(a,r1 )

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

we have that
|an |

2r M (r)
M (r1 )
.
=
n+1
2 r1
r1n


10.4. Classification of singularities. We can now classify the different types


of singularities. Consider f : P (a, R) C be analytic in a punctured disk P (a, R)
centred at a singularity a and let

f (z) =

an (z a)n

n=

be its Laurent series on a punctured disk P (a, R) = A(a, 0, R) (where R > 0).
Definition 10.9. If an 6= 0 for infinitely many positive n then a is called
an essential singularity. If an = 0 for only many positive n then a and the least
integer n such that an 6= 0 is negative then a is called a pole of order n. In the
particular case that a1 6= 0 and ak = 0 for k 2 we call a a simple pole.
Remark 10.10. If the least integer n such that an 6= 0 is non-negative is nonnegative then a is called a removable singularity for in this case
f (z) = an (z a)n + an+1 (z a)n+1 +
for 0 < |z a| < R and n 0. There fore we can define (or extend) f to a by
(
0
if n > 0
f (a) = lim =
za
a0 if n = 0
And with this value we can define f : D(a, R) C and it is analytic. We shall
always adopt the convention of removing singularities, in the future.
Example 10.11. Let f (z) = sin(z)/z = 1 z 2 /3! + z 4 /5! then we can
remove the singularity at z = 0 by setting f (0) = 1.
Definition 10.12. If the least n such that that an is strictly positive then f (z)
is said to have a zero at a of order n.
The order of a function at

n > 0
Ord(f, a) =

a is defined by
if
if
if
if

a
a
a
a

is
is
is
is

an essential singularity
a pole of order n
a removable singularity and f (a) 6= 0
a zero of order n

Definition 10.13. A function is said to be meromorphic if it is analytic in a


domain U except at a set of points which are poles (with no essential singularity).

74

CONTENTS

10.5. Behaviour at singularities. We can consider the behaviour of a function near a zero or pole
Theorem 10.14. Ord(f, a) = n where n is finite if and only if , , > 0 such
that
|z a|n |f (z)| |z a|n
whenever 0 < |z a| <
Proof. Assume that Ord(f, a) = n, then we can write
f (z) = an (z a)n + an+1 (z a)n+1 + = (an + g(z))(z a)n
where an 6= 0. Moreover, g(z) 0 as z 7 a and thus given 0 <  < 1 we can write
|g(z)| < |an | provided |z a| < , provided is small enough. In particular, we
can write
f (z)(z a)n = an + g(z).
and thus
|an |(1 ) |f (z)(z a)n | |an |(1 + )
if is small enough, giving the required bound.
Conversely, suppose
|z a|n |f (z)| |z a|n
when |z a| is small enough then f (z)(z a)n is bounded in a punctured neighbourhood of a and hence a is a removable singularity. In particular,
f (z)(z a)n = an + an+1 (z a) + an+2 (z a)2 +
in a punctured neighbourhood P (a, ) of a (since f (z)(z a)n is analytic at a by
the removable singularity). We want to show that an 6= 0. But from the left hand
inequality |f (z)(z a)n | is bounded below since > 0, which would not be the
case where an is equal to zero. Hence an 6= 0 and Ord(f, a) = n.

The previous result gives a rough idea of the behaviour of f near a pole (when
n is negative). The next result describes the behaviour near an essential singularity
(i.e., f (P (a, )) C is dense).
Theorem 10.15 (Casorati-Weierstrass). Let f have an essential singularity at
a. Then for any b C and every , > 0 there exists z such that 0 < |z a| <
such that |f (z) b| < .
Proof. Assume for a contradiction that there exist b, , such that for all z
such that 0 < |z a| < and |f (z) b| . Consider the function g : P (a, ) C
defined by
1
,
(1)
g(z) =
f (z) b
which, by assumption, is bounded by 1/ in P (a, ). Then g has a removable
singularity at a (and thus is bounded). In particular, we can write
g(z) = an (z a)n + an+1 (z a)n+1 +
where an 6= 0 and thus
g(z) = (z a)n h(z)
(2)
for an analytic function h : B(a, ) C with h(a) =
6 0. Thus comparing (1) and
(2) gives
1
f (z) b =
.
(3)
(z a)n h(z)
Moreover, 1/h(z) is analytic in a neighbourhood of a (since h does not vanish in a
neighbourhood of a by continuity and the fact h(a) 6= 0). But (3) shows that f has
a pole at a of order n, contradicting the hypothesis of an essential singularity. 

10. SINGULARITIES

75

Remark 10.16 (Picards Big Theorem). The following is am improvement on


the WeierstrassCasorati theorem.
Theorem 10.17 (Picards Big Theorem). If an analytic function f (z) has an
essential singularity at a, then on any punctured disk B(a, the function f (z) takes
on all possible complex values, with at most a single exception, infinitely often.
.

1. Show that the following series define a meromorphic function on C and determine
the set of poles and their orders
P (1)n
(1)
n!(n+z)
Pn=0

1
(2)
z 2 +n2
Pn=0

1
(3)
n=0 (n+z)2


P
1
+ n1
(4) z1 + nZ{0} zn
P (1)n
(5)
n=0 n!(n+z)
2. Show that
f (z) =

X
n=1

z2
+8

n2 z 2

is defined and continuous for the real values of z. Determine the region of he
complex plane in which this function is analytic. Determine its poles.
3. Let f : C C be a non-constant entire function. Show that the image of f is
dense in C.
4. Show that the order of a pole is invariant under analytic bijections (i.e., if
f : U C has a pole at z0 of order n and : V U is an analytic bijection with
(w) = z then w is a pole for f : V C of order n.
5. Let f : C C be a meromorphic function with lim|z|+ |f (z)| = . Prove
that f (z) is a rational function (i.e., the ratio of two polynomials).
6. Show that the Laurent series of a function on an annulus can be differentiated
term by term to give the derivative of the function on the annulus.
7. Find the Laurent series for the following functions:
(1) z/(z + 2) for |z| > 2,
(2) sin(1/z) for z 6= 0,
(3) cos(1/z) for z 6= 0,
(4) 1/(z 3) for |z| > 3.
8. Prove
(1)
(2)
(3)

the following expansions:


P 1
ez = e P
+ e n=1 n!
(z 1)n ,

n
1/z = n=0P
(1) (z 1)n for |z 1| < 1

2
1/z = 1 + n=1 (n + 1)(z + 1)n for |z + 1| < 1

9. Let f (z) be analytic on the annulus 0 < r |z| R. Prove that there is a
function f1 (z) analytic on |z| R and f2 (z) analytic on |z| r and
f = f1 + f2
on the annulus.

76

CONTENTS

11. Entire functions, their order and their zeros


11.1. Growth of entire functions. Let f : C C be an entire function.
Definition 11.1. We say that f (z) is of finite order if there exists a > 0 and
r > 0 such that
a
|f (z)| e|z| for |z| > r
The least such a (i.e the largest lower bound) is called the order of the entire
function.
Example 11.2.
(1) Any polynomial is of order 0.
(2) The function f (z) = sin(z) is of order 1.
(3) The function f (z) = sin(z 2 ) is of order 2.
We can denote
Mf (r) = max |f (z)|.
|z|=r

This function conveys a lot of information about the zeros of the entire function.
Lemma 11.3. Let f : C C be an entire function. If N Z+ and
Mf (r)
< +
lim inf
r+
rN
then f (z) is a polynomial of degree at most N + 1.
Proof. Assume that f (z) has a power series expansion (about z = 0)

X
f (z) =
an z n
n=0

and consider the associated polynomial P (z) =


defined by

PN

n=0

an z n . The function g : C C

g(z) := z N 1 (f (z) P (z)) = aN +1 + aN +2 z + aN +3 z 2 + .


is analytic and not identically zero. Assume that rnN Mf (rn ) is bounded for some
sequence rn + then
Mf (rn )
0 and thus Mg (rn ) 0
rnN +1
as n tends to infinity. In particular, we can choose R > 0 such that
sup{Mg (r) : 0 r R} > R.
By the maximum modulus principle applied to g(z) the disk B(0, R) we deduce
that g(z) is constant (i.e., g(z) = aN +1 ) and that f (z) = P (z) + aN +1 z N +1 is a
polynomial.

There is also a close connection between Mf (r) and the of zeros of f (z).
Definition 11.4. We denote by n(r)the number of zeros of f (z) in the ball
B(0, r) (i.e., n(z) = Card{z B(0, r) : f (z) = 0}).
Lemma 11.5. Let f : C C be an entire function with f (0) = 1. Then
n(r)

Mf (2r)
log 2

for all r > 0.


Proof. We can apply Jensens formula to the function F (z) := f (z/2r), say,
on the unit disk D to write


Z 2
n(2r)
X

2r
1
0 = log |f (0)| =
log
+
log |f 2rei |d
(1)
|ai |
2 0
k=1

11. ENTIRE FUNCTIONS, THEIR ORDER AND THEIR ZEROS

77

where a1 , , an(2r) are the zeros of f (z) with |ai | < 2r (and a1 , , an(r) are the
zeros of f (z) with |ai | < r). In particular,

n(r)

n(r) log 2

log 2

|ai |
k=1
|{z}
X

n(2r)

log 2

|ai |
| {z }

X
k=1

1
2
|

Z
0


log |f 2rei |d
{z
}
by (1)

log Mf (2r)

As a corollary, we have the following relationship between all the zeros and the
order.
Lemma 11.6. Let f : C C be an entire function of order a with f (0) = 1. If
(ai )
i=1 are the zeros for f (z) ordered by modulus, i.e.,
|a1 | |a2 | |a3 | |an |
then

1
< +.
a
|a
n|
n=1
Proof. Let  > 0, then for r > 0 sufficiently large we have that
M (r) exp(ra+ ).
Thus by the previous theorem
n(r)r(r2)

r(a)
r(a) a+
M (2r)
2r
0 as r +.
log 2
log 2

In particular, there exists r0 such that providing r > r0 have n(r) Cra+2 . Thus
since n n(|a|n ) |an |a+2 we have that
a+1
  a+2
1
1

.
|an |a+1
n
Providing 0 <  <

1
2

we have that

a+1

  a+2
X
1
1

< +.
a+1
|a
|
n
n
n=1
n=1


11.2. Factorization of entire functions. One of the most important applications of entire functions of finite order is their representation as infinite products.
This can be thought of as a generalization of the simple representation of a polynomial P (z) = an z n + + a1 z + a0 in terms of a finite product

n 
Y
z
P (z) = (an a0 )
1
i
i=1

78

CONTENTS

where 1 , , n are zeros of P (z).


Definition 11.7. We can define the elementary entire functions


z2
zp
Ep (z) = (1 z) exp z +
+ +
2
p
We have the following bound.
Lemma 11.8. If z D then |Ep (z) 1| |z|p+1 .
Proof. We leave the proof as an exercise. It requires showing that Ep (z) has
a power series expansion

X
Ep (z) = 1 +
bn z n
n=p+1

where bn 0. Then since 0 = Ep (1) = 1+ n=p+1 bn we have that 1 =


and then

X
X
|Ep (z) 1| = |
bn z n | |z|p+1
|bn | .
n=p+1

n=p+1

|bn |

n=p+1

{z

=1

}


This leads to the following.


Lemma 11.9 (Weierstrass). If an occur with multiplicity pn then
f0 (z) :=

Epn (

n=1

z
)
|an |

is an entire function with zeros an having multiplicity pn .


Proof. Let r > 0 then there exists N such that for n N we have that
|an | > r. We can then use the previous lemma to bound
|1 Epn (

z
1
)| | |pn +1
|an |
an

and thus


X


Epn ( z 1) < +.


|an |

n=N

In particular,

Y
n=1

E pn (




Y
z
z
)=
1 + E pn (
)1
|an |
|an |
n=1

which is enough to guarantee convergence.

This brings us to the statement of the main result.


Theorem 11.10 (Hadamard Factorization). If f (z) is an entire function of
order a and with zeros (an ) then we can write
f (z) = eQ(z)

Epn (

n=1

where Q(z) is a polynomial of degree at most a.

z
)
|an |

12. PRIME NUMBER THEOREM

79

Proof. The detailed proof is omitted. However, we will sketch the outline.
The idea is that given f (z) we can consider its zeros and thus construct the function
f0 (z). Moreover eg(z) := f (z)/f0 (z) will be an entire function.
Let r > 0 and let a1 , , an be the zeros for f (z) with |ai | < r. We can apply
the Poisson formula to log |f (z)| to write

2
 i

Z 2
n
X
r ai z
1
re + z


+
Re
log |f (z)| =
log |f (rei )d
(1)
log

i z
r(z

a
)
2
re
i
0
i=1
We can apply

i y
to (1) to get that

X 1
X
ai
1
f 0 (z)
=
+
+
r a z
f (z)
z

a
r
2
i
i
i
i

2rei
(rei z)2

log |f (rei )d

We can then differentiate this a times with respect to z to get




Z
X
X a!ai a+1
a!
(h + 1)! 2
2rei

+
+
log |f (rei )d
a+1
2a z
i z)a+2
(z

a
)
r
2
(re
i
i
0
i
i
We want to let r + and it is possible to show that this becomes

X
i

a!
.
(z ai )a+1

(2)

In particular, we have that


g a+1 (z) = Da

f0
f

(z) Da

f00
f0


(z)

The second term on the Right Hand Side of (2) is precisely (2). We deduce that
g a+1 = 0 and thus g(z) is a polynomial of degree a.

Example 11.11. We can consider sin(z) which is entire, of order 1, and zeros
at n, where n Z. Thus
Y
z 
.
sin(z) = C
1
n
nZ

(The function g(z) = az + b must be constant by the periodicity under z 7 z + 2.


12. Prime number theorem
Perhaps one of the most interesting applications of complex analysis to another
discipline is the proof of the prime number theorem in number theory.
Theorem 12.1. Let (x) denote the number of prime numbers less than x then
(x) logx x as x +, i.e., limx+ x/(x)
log x = 1.
Example 12.2. We can compute: (10) = 4, (100) = 25, (1, 000) = 168,
(1, 000, 000) = 78, 498 and (1, 000, 000, 000) = 50, 847, 534.
x
In 1808 Legrendre conjecture that (x) log(x)1.08366
. The correct form was
proved independently by Hadamard and de la Valle-Poussin in 1896, using complex
analysis and the zeta function introduced by Riemann in 1859.
There are proofs which do not use complex analysis, including the elementary
proof of Selberg and Erd?s from 1949, but these are even more difficult.

80

CONTENTS

12.1. Riemann
Pzeta function. Recall that the Riemann zeta function is
defined by (s) = n=1 ns , provided Re(s) > 1.
Lemma 12.3. (s)

1
s1

is analytic for Re(s) > 0.

Proof. For Re(s) > 1 we can write



Z
Z n+1 

X
X
1
1
1
1
1

dx =
s dx
(s)
=
s 1 n=1 ns
xs
ns
x
1
n=1 n
and bound

Z x

1


1
|s|
1 = s
du
.
ns
xs n us+1 nRe(s)+1
This leads to the result.

1
Remark 12.4. The zeta function (s) s1
actually extends to an entire
function. The extension to Re(s) > 0 is given at the end.

Lemma 12.5. We can write


Y
1
(s) =
1 ps
for Re(s) > 1
p

where the product is over all primes p.


Proof. This follows from the prime factorization of natural numbers.

12.2. A related complex function. We define


(s) =

X log p
ps

for Re(s) > 1

where the sum is over all primes p.


1
Lemma 12.6. (s) s1
is meromorphic for Re(s) > 1/2. Moreover, the poles
occur at the zeros of (s) in this region.

Proof. For Re(s) > 1 we have that


0 (s)

=
log (s)
(s)
s
| {z }
meromorphic for Re(s) > 0

=
s
=

X
p

!
X

log(1 p

log p
(ps 1)

X log p
|

ps
{z

=:(s)

log p
1)
p
|
{z
}
converges for Re(s) > 1/2
X

ps (ps


Lemma 12.7. (s) has no zeros on Re(s) = 1. (In particular, (s) is analytic
on Re(s) = 1).

12. PRIME NUMBER THEOREM

81

Proof. We can write for s = + it


()3 |( + it)|4 |( + 2it)|

X X 1
3 + 4 cos(mt log p) + cos(2mt log p) 1(1)
= exp
m |
{z
}
mp
p m=1
0

(since 3 + 4 cos + cos(2) = 2(1 + cos ) 0). Assume for a contradiction that
1 + it0 is a zero for (s). Since s = 1 is a simple pole for (s) we see that
lim ()3 |( + it0 )|4 |( + 2it0 )| = 0

giving the contradiction to (1).

12.3. A related counting function. We denote


X
(x) =
log p for x > 0
px

Lemma 12.8. We can bound (x) (8 log 2)x


Proof. Using the binomal expansion we can bound

 
Y
X
2n
22n = (1 + 1)2n

p = exp
log p = e(2n)(n) .
n
n<p2n

n<p2n

i.e., (2n) (n) n(2 log 2). If 2k < x 2k+1 , for some k > 0, then we can now
bound
k
X

(2m+1 ) (2m )
(x) (2k+1 )
|
{z
}
m=0
2m (2 log 2)

k
X

!
2m

m=0
k+1

= 2(2

2 log 2

1)2 log 2

(8 log 2)x

12.4. Finiteness of integrals. The core of the proof is a finite integral, using
Cauchys theorem.
Lemma 12.9. The following integral os finite:
Z y
Z
(x) x
(x) x
lim
dx
=
dx < +
y+ 1
x2
x2
1
Proof. We break the poof up into simpler steps.
Step 1 (Relating the integral to (s)): For Re(s) > 1 we have
Z
X log p
(x)
(s) =
=
s
dx.
s
p
xs+1
1
p
Moreover, by a change of variables (x = et ) we can write
Z
Z
Z
Z
(x)
1
1
st
t
e
(e
)dt
and
est dt.
dx
=
=
dx
=
s+1
s+1
x
s
x
0
1
0
1
Let us denote f (t) := (et )et 1 then by combining (1) and (2) we have that
Z
(s + 1) 1
g(s) :=
f (t)est dt =

(s + 1)
s
0

(1)

(2)

82

CONTENTS

for Re(s) > 0. Moreover, this is analytic on a open set U {z : Re(z) 0}.
Step 2 (g(z) as a limit): For T > 0 we can consider the entire functions
!
Z log T
Z T
(x) x
zt
dx
gT (z) =
f (t)e dt =
x2
1
0
We need to show that
Z
lim
T

f (t)dt = lim gT (0) = g(0) =


T

f (t)dt.
0

Let C = C(R, ) be the boundary of the D-shaped region defined my


D = {z C : |z| R, Re(z) }.

Given R, provided > 0 is sufficiently small, we have that g(z) gT (z) is


analytic on a neighbourhood of D (by Step 1) and thus by Cauchys theorem we
can write


Z
1
ezT
z2
g(0) gT (0) =
(g(z) gT (z))
1 + 2 dz.
(3)
2i C
z
R
We can write C = C + C where C + = {z C : Re(z) 0} and C =
{z : Re(z) 0}
Step 3 (Integral over C+ ): For z C+ we can bound the integrand of (3) using
Z



zt

|g(z) gT (z)| =
f (t)e dt
T
Z
f (t)eRe(z)t dt
max |f (t)|
t0
(4)
T
{z
}
|
=:B


B

eRe(z)T
Re(z)

and



zT
z 2 1
2Re(z)
e
1
+
= eRe(z)T
.


2
R
z
R2

(5)

(since if z = Rei then |1 + z 2 /R2 | = |1 + ei2 | = 2 cos() = 2Re(z)). Together (4)


and (5) give that the integral over C+ has a bound



1 Z
ezT
z2

B
(g(z) gT (z))
1 + 2 dz .
(6)

2i C+
R
z
R

12. PRIME NUMBER THEOREM

83

Step 4 (Integral of gT (z) over C ): By analyticity of gT (z) we can write






Z
Z
ezT
z2
ezT
z2
gT (z)
1 + 2 dz =
gT (z)
1 + 2 dz
0
z
R
z
R
C
C

(7)

0
where we replace the integral over C by that over C
= {z C : |z| = R, Re(z) <

0}.

Furthermore, we can again bound


Z T
Z
zt
|gT (z)| = |
f (t)e dt| B
0

|ezt |dt =

BeRe(z)T
|Re(z)|

for Re(z) < 0 (compare with (4)). The analogue of (5) is





zT
z 2 1
2|Re(z)|
e
1+ 2
= eRe(z)T
.

R
z
R2
Together (4) and (5) give us a bound on the contribution of (7) to (3) of



1 Z
ezT
z2

B
gT (z)
1 + 2 dz

2i C0
R
z
R

(40 )

(50 )

(8)

(compare with (6)).


Step 5 (Integral of g(z) over C ): Finally, we can bound the remaining contribution to (3) coming from the integralover C by


Z
g(z)
z2
1 + 2 ezT dz 0 as T 0
(9)
z
R
C
|
{z
}
Independent of T
since Re(z) < 0 on C .
Step 6 (Putting the bounds together): Finally, from the identity (3) and the
estimates (6), (8) and (9) we have that
lim sup |g(0) gT (0)|
T +

2B
.
R

However, since the R can be arbitrarily large we deduce that limT gT (0) =
g(0).

84

CONTENTS

12.5. Proof of the Prime Number Theorem. It is actually easier to first


show the following.
Lemma 12.10. (x) x as x +
Proof. Assume that > 1 we can find xn + such that (xn ) xn .
Since (x) is monotone increasing:
!
Z xk
Z xk
Z

X
(x) x
1
1
dx
xk
dx
dx
x2
x2
x
1
xk
xk
k=1
!
Z
Z

X
1
1
dx
=
dx
2
1 x
1 x
k=1
|
{z
}
(By change of variables)



X
1
=
1
log = +.

k=1 |
{z
}
>0

This contradicts the integral converging.


On the other hand, assume that < 1 we can find xn + such that
(xn ) xn we again show that the integral doesnt converge.
Either way, we see that (x) x as x +.

This leads to the final result.
Theorem 12.11 (Prime Number Theorem).
x
(x)
as x +.
log x
Proof. We can bound
(x) =

log p

px

log x = (x) log x

px

For any  > 0,


(x)

log p

x1 px

(1 ) log x

x1 px

(1 ) log x (x) (x1 )


| {z }
x1

Since  > 0 is arbitrary, together these show that (x)

x
log x .

Application 12.12 (More on the Riemann zeta function). Recall that we define
the Riemann zeta function by

X
1
(s) =
s
n
n=1
It is possible to show that this has an extension to C. We can write
Z
(s) =
xs1 ex dx
0

13. FURTHER TOPICS

85

for = Re(s) and then


ns (s) =

xs1 enx dx.

Thus summing over n we have


(s)(s) =

X
1
s
n
n=1

(s) =
0

xs1
dx.
ex 1

Lemma 12.13. For = Re(s) > 1,


(s) =

(1 s)
2i

Z
C

(z)s1
dz
ez 1

where C runs from to 0, an infinitesimal circle about 0, followed by 0 to .


Proof. We can write the integral as
Z
Z
(1 s) xs1 e(s1)i
(1 s) xs1 e(s1)i
dx
+
dx

2i
2i
ee(s1)i 1
ee(s1)i 1
0
0
which is equal to 2i sin((s 1)(s)(s)) (using that (s)(1 s) = / sin(s)). 
In particular, the Right Hand Side of the formula is meromorphic for all s. The
integral is entire in s. Moreover, (1 s) is meromorphic with poles at s = 1, 2, .
(The poles at n 2 must actually cancel with the zeros of the integral since we
already know that (s) is analytic there. However, the integral at s = 1 has value
1.
Corollary 12.14. (s)(s 1) is an entire function (i.e., (s) has a simple
zero at s = 1).
13. Further Topics
There are a number of topics which would have fitted nicely into the course
with more time.
13.1. Weierstrasss elliptic function. If 1 , 2 C such that Im(1 /2 )
then we can consider the set of points
L = {2n1 1 + 2n2 2 : n1 , n2 Z} C
Definition 13.1. We can define the Weierstrass elliptic functions are defined
by
P (z) =

X
1
1
1
+
2
2
z
z w2
w
L

In particular, we have that P (z + 2i ) = P (z) for i = 1, 2..


Theorem 13.2. This has a Laurent series expansion
1
1
1
P (z) = 2 + g2 z 2 + g3 z 4 + O(z 6 ).
z
20
28
where
X
X
1
g2 (1 , 2 ) = 60
and g3 (1 , 2 ) = 60
4
w
L{0}

The coefficients satisfy g2 (1 , 2 ) =


for > 0.

L{0}

1
w6

g2 (1 , 2 ) and g3 (1 , 2 ) = 6 g3 (1 , 2 ),

We can let = 1 /2 with Im( ) > 0 be the period ratio and consider g2 ( ) =
g2 (1, 2 /1 ) and g3 ( ) = g3 (1, 2 /1 ).

86

CONTENTS

Definition 13.3. The modular discriminant is defined as


( ) = g23 27g32
This is studied in its own right, as a cusp form, in modular form theory (that
is, as a function of the period lattice).
Theorem 13.4. The discriminant is a modular form of weight 12. That is,
under the action of the modular group, it transforms as


z + b

= (c + d)12 ( )
c + d
with a, b, c, d being integers, with ad bc = 1.
13.2. Picards Little theorem. These describe the values of entire functions.
Theorem 13.5 (Picards Little Theorem). A non-constant entire function takes
every value , except at most one.
13.3. Runges theorem. This deals with approximations on compact sets.
Theorem 13.6 (Runge). Let U be a (simply-connected) domain. Then any
function analytic in U can be approximated uniformly on compact sets K U by
polynomials, i.e., for each  > 0 there is a polynomial p(z) with complex coefficients
such that |f (z) p(z)| <  for all z K.
13.4. Conformality: Property of analytic maps. Any complex number
z = rei where r = |z| and = arg(z) is defined up to a multiple of 2.
Let f : U C be analytic and let be a smooth contour in U represented bt
a parameterization z : [a, b] C and let a < t0 < b be such that z 0 (t0 ) 6= 0.
Definition 13.7. Let U be an open neighbourhood. We say that f : U C
is conformal on U if for each z0 U the function f is complex differentiable at z0
and f 0 (z0 ) 6= 0 exists.
This has a more natural interpretation.
Lemma 13.8. f is conformal if and only if it preserves angles, i.e., given two
smooth paths 1 , 2 : [0, 1] C with 1 (0) = 2 (0) we have that
arg (10 (0) 20 (0)) = arg ((f 1 )0 (0) (f 2 )0 (0)) .
Proof. We can write j (t) = xj (t) + iyj (t) and denote z0 = 1 (0) = 2 (0).
Let f (z) = f (x + iy) = u(x, y) + iv(x, y). Suppose that 10 (0) = 20 (0). We can write
(f j )0 (0) = f 0 (j (0).j (0) for j = 1, 2. Thus
arg((f j )0 (0)) = arg(f 0 (j (0)) + arg(j (0))
up to a multiple of 2. We can then write
arg((f 1 )0 (0)) arg((f 2 )0 (0)) = arg(10 (0)) arg(20 (0))
up to a multiple of 2.

Theorem 13.9. Analytic maps are conformal at all points at all points z0 such
that f 0 (z0 ) 6= 0
Method I: This can be illustrated by considering two two orthogonal families
families of contours such as the vertical lines and horizontal lines (or concentric
circles and radial lines). Let f (z) = u(x, y) + iv(x, y) with vertical lines x = and
horizontal lines y = .
Consider the the image of 2 perpendicular curves passing through z0 , i.e., z0 =
+ i. The image of the line x = is now z = u(, y) + iv(, y) for y R. The
image of the line y = is now z = u(x, ) + iv(x, ). for x R.

14. PROBLEMS

87

Method II: Consider the so-called level curves of f (z) = u(z) + iv(z), i.e.,
u(x, y) = and v(x, y) = . In particular, these curves are the ones mapped
to the straight lines u = and = v.
Provided f 0 (z0 ) 6= 0 we have that the angle between these preimages is = /2.
Example 13.10 (Method I). Let f (z) = ez then f 0 (z) = ez 6= 0 everywhere.
Writing z = x + iy gives f (z) = ex (cos + i sin ).
If x = then u = e cos y and v = e cos y with y R a parameter, i.e.,
2
u + v 2 = e2 which describes a circle.
If y = then u + iv = ex (cos + i sin ). Thus uv = tan(), i.e., a straight line
through the origin at angle .
Example 13.11 (Method II). Let f (z) = z 2 then f 0 (z) = 2z =6= 0 except at
z = 0. Writing f (z) = u + iv, the preimage of u = are those x + iy such that
u(x, y) = and the preimage of v = are those x + iy such that v(x, y) = .
If x = then u = e cos y and v = e cos y with y R a parameter, i.e.,
2
u + v 2 = e2 which describes a circle.
If y = then u + iv = ex (cos + i sin ). Thus uv = tan(), i.e., a straight line
through the origin at angle .
Thus z 2 = (x+iy)2 = (x2 y 2 )+i2xy. The level sets correspond to x2 y 2 =
or 2xy = .
Remark 13.12. In practise, it can be difficult to find the appropriate constants
for these mappings. This involves numerical evaluation of elliptic integrals.
13.5. Functions of several complex variables. We can consider analytic
functions f : U C, where U C2 is an open set. We can define analyticity in terms of the convergence of power series in two variables (on poly disks
B(z0 , w0 , 1 , 2 ) = {(z, w) C2 : |z z0 | < 1 , |w w0 | < 2 }) upon which we can
apply the Cauchy theorem twice to write
Z
Z
1
f (z, w)
f (z, w) = 2
dzdw
4 C(z0 ,1 C(z0 ,1 (z z0 )(w w0 )
There are two particularly interesting theorems: The Bochner Tube Theorem and
Hartogs Theorem
Theorem 13.13 (Bochner Tube theorem). Let R2 be a connected open set
and let
co() = {x + (1 )y : x, y , 0 1}
be the convex hull. If f : + iR2 C is analytic then f : co() + iR2 C
Theorem 13.14 (Hartog Theorem). Assume that f : B(z0 , w0 , 1 /2, 2 /2) C
is analytic in two variables and f (, w) : B(z0 , r1 ) C is analytic in one variable
for each w B(w0 , r2 ). Then f : B(z0 , w0 , 1 /2, 2 /2) C is also analytic.
14. Problems
14.1. Problems 1. .
1. Show that the preimage of a circle or a straight line in C is a circle on the
Riemann sphere under stereographic projection.
b coming from the stereographic projection of the usual
2. Show that the metric on C
Euclidean metric is given as follows: If z := ((x1 , x2 , x3 )) and w := ((x01 , x02 , x03 ))
then
|z w|
d(z, w) = d ((x1 , x2 , x3 ), (x01 , x02 , x03 )) = p
(1 + |z|2 )(1 + |w|2 )

88

CONTENTS

whenever (x1 , x2 , x3 ), (x01 , x02 , x03 ) 6= (0, 0, 1) (i.e., z, w 6= ) and


d(z, ) := d((x1 , x2 , x3 ), (0, 0, 1)) = p

2
1 + |z|2

whenever (x1 , x2 , x3 ) 6= (0, 0, 1) (i.e., z 6= )




a b
3. Given a complex matrix M =
we can associate a Mobius map TM :
c d
C C of the form TM (z) = az+b
obius maps, relate MT1 T2 to MT1
cz+d . Given two M
and MT2 .
4. Find the M
obius transformation that takes:
(1) The points i, 1, i to the points i, 0, i, respectively;
(2) The points 1 + i, 0, 1 i to the points 1, i, i, respectively;
(3) The points 0, 1 + i, 1 i to the points 1, i, i, respectively.
5. What is the image of the unit disk D = {z C : |z| < 1} under the following
M
obius maps:
z+i
;
(1) T (z) = zi
iz1
(2) T (z) = iz ;
(3) T (z) =
(4) T (z) =
(5) T (z) =
(6) T (z) =

(1+i)z+(i1)
;
iz+1
2z1
;
z2

2z1+i ;
2z+1i
(1+i)z1i
.
z+1

b is given by
6. The cross ratio of four points z0 , z1 , z2 , z3 C
(z0 , z1 , z2 , z3 ) =

(z0 z2 )(z1 z3 )
.
(z0 z3 )(z1 z2 )

If = (z0 , z1 , z2 , z3 ) show that different permutations of the four numbers gives


cross-ratios with the values
1
1

, 1 , 1 ,
,
.
1 1
Show that these are all possible values.
Apollation circle packing
1. Show that if E is a circle of radius k then:
(1) if we reflect in E a circle C of radius r whose centres are separated by
d > r + k then the image f (C) is a circle with radius k 2 r/(d2 r2 ).
(2) if we reflect in E a straight line at a distance b > k from the centre of E
then the image f (C) is a circle of radius k 2 /2b.
2. If a M
obius map f (z) = az+b
cz+d takes real numbers to real numbers (i.e., t R =
f (t) R) then show that a, b, c, d are all real numbers.

Classification of M
obius maps
3. Show that any M
obius map other than the identity must fix at least one point
and at most two.

14. PROBLEMS

89

4. Since we are assuming ad bc = 1 show that we can write (a d)2 + 4bc =


(a d)2 + 4ad 4 = 0
5. Let f : C C be a M
obius map. Show that
(1) If f () = then f (z) = z + , for some , C.
(2) If f () = and f (0) = 0 then f (z) = z, for some C.
6. We say that two linear fractional transformations f1 , f2 are conjugate if f2 =
g 1 f1 g for some linear fractional map g. We can associate to the Mobius map
f (z) = az+b
cz+d the value tr(f ) := a + d. Show that the value tr(f ) is preserved by
conjugacy, i.e., if f1 , f2 are conjugate then tr(f1 ) = tr(f2 ).
bC
b be a M
7. Let f : C
obius transformation. The non-identity Mobius transformations are commonly classified into three types:
(1) parabolic (conjugate to z 7 z + with a single fixed point );
(2) elliptic (conjugate to z 7 z with || = 1 with two points 0, ); and
(3) loxodromic (everything else).
Show that:
(1) Elliptic M
obius transformations correspond to tr(f )2 [0, 4);
(2) Parabolic M
obius transformations correspond to tr(f )2 = 4;
(3) Hyperbolic M
obius transformations (i.e., conjugate to z 7 z with 0 <
< 1 or (1, )) correspond to tr(f )2 (4, ).
14.2. Problems 2. .
Here <z and =z stand for real and imaginary parts of the complex number z,
respectively.
0. Let f (z) : C C be a continuous function. Is it true that
(1) <f (z) and =f (z)
(2) |f (z)| and (f (z))
are continuous functions?
1. Show that if f : U C is analytic and there exist a C and a convergent
sequence zn z with zn , z U such that f (zn ) = a then f |U = a. What happens
if z 6 U ?
2. Show that the equivalence of parameterizations of curves is an equivalence relation.
3. Show that

f (w)dw =

f (z)dz for any closed curve .

4. Show that if R = [a, b] R [c, d] is a rectangle then


(1) We can evaluate RR dz = 0;
(2) We can evaluate R (z z0 )dz = 0.
5. Write the Cauchy-Riemann equations in polar coordinares.
6. Let f : D C be a function well-defined in the unit disc, such that partial
derivatives exist everywhere and Cauchy-Riemann equations hold true. Is it true
that f is complex differentiable everywhere in the unit disc? Hint: consider f (z) =
exp(z 4 ).

90

CONTENTS

8. Find all holomorphic functions that map the unit disc to the real line.
9. Find domains of complex differentiablity of the following functions
(1) f1 (z) = |z|;
(2) f2 (z) = z z;
(3) f3 (z) = z.
14.3. Problems 3. .
P
= n=0 Bn!n z n . Prove that
(
1 if n = 1
B1
Bn1
B0
+
+ +
=
n!0! (n 1)!1!
1!(n 1)!
0 if n > 1.

1. Define the Bernoulli numbers by the series

2. Compute

z
ez 1

zez dz where:

(1) = [i, i + 2]; and


(2) = {x + ix2 : 0 x 1}.
3. Compute

sin(z)dz where = {x + ix : 0 x 1}.

4. Let f : D D be an analytic
P map which is injective (i.e., f (z1 ) = f (z2 ) implies
z1 = z2 ) of the form f (z) = n=1 an z n . Show that the image has area:
Area(f (D)) =

n2 |an |2 .

n=1

5. Let f : D D C be a continuous function such that


(1) for w D the map D 3 z 7 f (z, w) C is analytic; and
(2) for z D the map D 3 zw 7 f (z, w) C is analytic.
P P
Show that we can write f (z, w) = n=0 m=0 anm z n wm , which converges absolutely for z, w D.
6. Use Cauchys theorem to evaluate

Z
C(0,1)

za
zb

2
dz

where b < 1 and C(0, 1) = {z C : |z| = 1}.


7. Let f : U C be an analytic function. Let z0 U with f 0 (z0 ) 6= 0. Let C be a
sufficiently small circle centred at z0 then show
Z
2i
dz
=
0
f (z0 )
f (z) f (z0 )
8. If f is analytic in U and z0 U define
(
g(z) =

f (z)f (z0 )
zz0
f 0 (z0 )

if z 6= z0
if z = z0 .

Show that this is analytic.


9. Use Rouches theorem to show that :
(1) z 7 4z 3 + z 1 = 0 has three zeros inside the unit circle;
(2) z 7 + 5z 3 z 2 = 0 has three zeros inside the unit circle.

14. PROBLEMS

91

10. Use Rouches theorem to show that z 87 + 36z 57 + 71z 4 + z 3 z + 1


(1) has 4 zeros inside the unit circle.
(2) has 87 zeros inside the circle {z C : |z| = 2}.
14.4. Problems 4. .
1. If |a| > e then show that ez = az n has n roots in the unit disk D = {z C : |z| <
1}.
2. Consider the following polynomials:
(1) How many zeros of z 7 2z 5 + 6z 3 z + 1 are in the unit disk D;
(2) How many zeros of z 4 6z + 3 are in the annulus {z C : 1 < |z| < 2};
(3) How many zeros of z 4 4z 3 +6z 2 4z+3 are in the set {z C : |z1| < 1}.
3. Let h be analytic in the unit disk D and assume that
|h(z) h(w)| < |z w|
for all z, w D. Show that the function f : D C defined by f (z) = h(z) + z is
one-to-one.
4. Let a > 0. Show that if f : C C is analytic and satisfies |f (z)| < |z|a then f
is a polynomial of degree at most [a].
5. Show that if f : C C is analytic and satisfies |f (z)| > 1 whenever |z| > 1 then
f is a polynomial.
6. Show that if f : C C is analytic, non-constant and has no zeros then there
must exist a sequence of points |zn | such that f (zn ) 0.
7. Let f : D C be analytic and continuous on the closed unit disk D = {z
C : |z| 1}. Moreover, assume that:
(1) |f (z)| M whenever |z| = 1 and Im(z) 0; and
(2) |f (z)| N whenever |z| = 1 and Im(z) < 0.

Then show that |f (0)| N M .


8. Show that if f : C C is analytic and Im(f (z)) 6= 0 whenever |z| 6= 1 then f is
constant.
9. Show that if f is a polynomial and z C then either f n (z) as n + or
{f k (z) : k 0} is a bounded set.

14.5. Problems 5. 1. Consider the family of functions fn : D C defined


by fn (z) = z n , n 1. Show that for all z D = {z C : |z| < 1} we have that
fn (z) 0 as n and that the convergence is uniform on compact sets, but
that it is not uniform on D.
2. Give an example of a family of analytic functions fn : D C, n 1, for which
fn (z) has precisely one zero, but fn converges uniformly on D to the function f = 0
(i.e., the function which is identically zero).
3. Give an example of a family of analytic functions fn : C C, n 1 such
that fn uniformly on C (i.e., for all compact sets K C and M > 0 we
have that inf zK |fn (z)| M for sufficiently large n) but for which the derivatives
fn0 : C C, n 1, neither tend to nor form a normal family.

92

CONTENTS

4. Given two domains U1 , U2 and two points z1 U1 and z2 U2 show that there
exists an analytic bijection f : U1 U2 such that f (z1 ) = z2 . [Hint: Use the
Riemann Mapping Theorem].
5. Find the image of the unit disk D with respect to the map z 7 z + z1 . Find the

image of the upper half plane H with respect to the map z 7 z 2 1


6. Find an analytic bijection between the following domains:
(1) R (0, ) and H = {z C : Im(z) > 0};
(2) H [0, ir] and H;
(3) D and C (, 41 ];
(4) (R (0, 2)) ((i , i ] [i + , i + )) and H;
(5) H and the equilateral triangle.
7. Let U C be defined by
{z C : 0 < Re(z) < 1, 0 < Im(z) < 1}



[
1
1 1
,
+
i
.
2n 2n
2
n=1

(1) Show that U is (path) connected;


(2) Show that U is simply connected;
(3) Show that 0 U in the boundary cannot be reached by a continuous
path from any point in U ;
(4) Show that any analytic bijection f : U D cannot extend to a continuous
map from the boundary U (of U ) to the unit circle D.
8. Let U C be defined by
{z C : Im(z) > 0 and |z| < 1}


[  1
0, eip/q .
q

p/qQ

Show that every point in the boundary U can be reached by a continuous path
from any point in U . However, show that any analytic bijection f : U D cannot
extend to a continuous map from the boundary U of U to the unit circle D.
9. Consider an analytic bijection f : U D that satisfies that for z0 U R we
have that f (z0 ) = 0 and f 0 (z0 ) > 0. Show that if U is symmetric about the real
axis then f (z) = f (z).
10. Show that there is no analytic bijection between D and C.
14.6. Problems 6. .

1. Let f : U D be an analytic bijection. Fix z0 U . Show that if f (z0 ) = 0 and


f 0 (z0 ) > 0 then f is uniquely determined (i.e., uniqueness in the Riemann Mapping
Theorem).
2. Let x1 < x2 < < xn be on the real line. Let 0 < 1 , , n < 2. Let
C
.
F (z) =
1
n
(z x1 )
(z xn )1n
Show that if xi1 < t < xi then
Arg(F (t)) = Arg(C) (i + i+1 + + n ).
Deduce that the image of R with respect to the map f (z) =
piecewise linear curve. Is this a closed curve?

Rz
z0

F (w)dw is a

14. PROBLEMS

93

3. Let f : P D be continuous. Assume that f extends to a bijection f : P D


on the boundary. Show that f : P D is surjective.
4. Let g : U C be analytic. Let D U be a closed region whose boundary curve
= D is simple. Assume that g : g() is one-to-one. Show that g : D g(D)
is one-to-one. (This is Darbouxs theorem). [Hint: Use the Argument Principle]
5. Do questions 2 4 give another proof that the Schwartz-Christoffel formula gives
an analytic bijection f : H P for a polygon P?
6. Consider the semi-infinite strip
S = {x + iy : a < x < a and y > 0}
for a > 0. Use the Schwarz-Christoffel theorem to find the analytic bijection f :
H S which extends to f (1) = a and f () = .
7. Consider the triangle in C with vertices at 1, 0 and i. Use the SchwarzChristoffel theorem to find the analytic bijection f : H which extends to
f (1) = i, f (1) = 1 and f () = 0.
8. Derive the version of the Schwarz-Christoffel theorem for an analytic bijection
f : D P from the unit disk to a polygon.
Rz
9. Describe the image of the unit disk D under the map f (z) = 0 (1 n )2/n d,
for n 3.
Rz
10. Describe the image of the unit disk D under the map f (z) = 0 (1 4 )1 (1 +
4 )1/2 d.
14.7. Problems 8. .

1. Let f : U C be analytic. Assume that f never vanishes. If there is a point


z0 U such that |f (z0 )| |f (z)|, for all z U then f is a constant.
2. Suppose that both f (z) and g(z) are analytic on D with |f (z)| = |g(z)| for
|z| = 1. Show that if neither f (z) nor g(z) vanishes in D then there exists C
with || = 1 such that f = g.
3. Suppose that f (z) is analytic on D and for 0 r < 1 define A(r) = max|z|=r Ref (z).
Show that unless f is a constant, A(r) is a strictly increasing function of r.
4. Does there exist an analytic function f : D D with f ( 21 ) =

3
4

and f 0 ( 21 ) = 34 ?

5. Let f (z) = (z + 1)2 on U = {z = x + iy : 0 x 2 and 0 y 1}. Where


does |f (z)| have its maximum and minimum values?
6. Show that the function w(x, y) = ex sin(y) is harmonic.
7. Let f : U C be analytic and non-zero. Show that log |f (z)| is harmonic.
8. Show Harnacks principle: Let u1 u2 be harmonic functions on an open
set U . Then either uj uniformly on compacts sets or there is a harmonic
function u on U such that uj u uniformly on compact sets. [Hint: Use Harnacks
Inequality].
9. Find the harmonic conjugate of u(x, y) = y 3 3x2 y.

94

CONTENTS

10. Describe suitable domains for the following harmonic functions, and find their
harmonic conjugates of the following functions:
(i) 2x(1 y);
(ii) 2x x3 + 3xy 2 ;
(iii) sinh(x) sin(y);
(iv) y/(x2 + y 2 ).
11. Show that if v1 , v2 are both harmonic conjugates of u then v1 v2 is a constant.

Vous aimerez peut-être aussi