COMPLEX ANALYSIS - Semester 2016-1: Robert Oeckl - CA NOTES - 01/12/2015
COMPLEX ANALYSIS - Semester 2016-1: Robert Oeckl - CA NOTES - 01/12/2015
Contents
1 Holomorphic functions 3
1.1 The complex derivative . . . . . . . . . . . . . . . . . . . . . 3
1.2 Elementary properties of holomorphic functions . . . . . . . . 5
1.3 The exponential function . . . . . . . . . . . . . . . . . . . . 6
1.4 Power series and analytic functions . . . . . . . . . . . . . . . 9
2 Complex integration 15
2.1 Integration along paths . . . . . . . . . . . . . . . . . . . . . 15
2.2 Closed paths and winding . . . . . . . . . . . . . . . . . . . . 18
2.3 Integrable functions . . . . . . . . . . . . . . . . . . . . . . . 19
2.4 The Cauchy Integral Formula . . . . . . . . . . . . . . . . . . 23
2.5 General Cauchy Theory . . . . . . . . . . . . . . . . . . . . . 27
3 Basic properties 31
3.1 From local to global structure . . . . . . . . . . . . . . . . . . 31
3.2 Zeros . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.3 Holomorphic logarithms and roots . . . . . . . . . . . . . . . 36
4 Singularities 39
4.1 Types of singularities . . . . . . . . . . . . . . . . . . . . . . . 39
4.2 Meromorphic functions . . . . . . . . . . . . . . . . . . . . . . 41
4.3 Laurent Series . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.4 Residues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5 Conformal mappings 49
5.1 Conformal mappings as holomorphic functions . . . . . . . . 49
5.2 The Open Mapping Theorem . . . . . . . . . . . . . . . . . . 50
5.3 Biholomorphic mappings . . . . . . . . . . . . . . . . . . . . . 51
5.4 Conformal automorphisms of C and C× . . . . . . . . . . . . 52
5.5 Conformal automorphisms of D . . . . . . . . . . . . . . . . . 55
5.6 Möbius Transformations . . . . . . . . . . . . . . . . . . . . . 56
5.7 Montel’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . 59
5.8 The Riemann Mapping Theorem . . . . . . . . . . . . . . . . 61
2 Robert Oeckl – CA NOTES – 01/12/2015
6 Harmonic functions 65
6.1 Mean value and maximum . . . . . . . . . . . . . . . . . . . . 65
6.2 The Dirichlet Problem . . . . . . . . . . . . . . . . . . . . . . 68
1 Holomorphic functions
1.1 The complex derivative
The basic objects of complex analysis are the holomorphic functions. These
are functions that posses a complex derivative. As we will see this is quite a
strong requirement and will allow us to make far reaching statements about
this type of functions. To properly understand the concept of a complex
derivative, let us recall first the concept of derivative in Rn .
Definition 1.1. Let U be an open set in Rn and A : U → Rm a function.
Given x ∈ U we say that A is (totally) differentiable at x iff there exists an
m × n-matrix A′ such that,
A(x + ξ) = A(x) + A′ ξ + o(∥ξ∥)
for ξ ∈ Rn sufficiently small. Then, A′ is called the derivative of A at x.
Recall that the matrix elements of A′ are the partial derivatives
∂Ai
A′ij = .
∂xj
Going from the real to the complex numbers, we can simply use the de-
composition z = x + iy of a complex number z into a pair of real numbers
(x, y) to define a concept of derivative. Thus, let U be an open set in C and
consider a function f : U → C. We view U as an open set in R2 with coordi-
nates (x, y) and f = u + iv as a function with values in R2 with coordinates
(u, v). The total derivative of f , if it exists, is then a 2 × 2-matrix f ′ given
by ( )
∂u ∂u
f′ = ∂x
∂v
∂y
∂v .
∂x ∂y
So far we have only recited concepts from real analysis and not made use
of the fact that the complex numbers do not merely form a 2-dimensional
real vector space, but a field. Indeed, this implies that there are special
2 × 2-matrices, namely those that correspond to multiplication by a complex
number. As is easy to see, multiplication by a+ib corresponds to the matrix,
( )
a −b
.
b a
The crucial step that leads us from real to complex analysis is now the
additional requirement that the derivative f ′ take this form. It is then more
useful to think of f ′ as the complex number a + ib, rather than this 2 × 2-
matrix.
4 Robert Oeckl – CA NOTES – 01/12/2015
Proof. The proofs are completely analogous to those for real functions on
open subsets of the real line with the ordinary real differential. Alternatively,
1.-3. follow from statements in real analysis by viewing C as R2 .
Note that items 1.-3. imply that O(D) is an algebra over the complex
numbers.
where exp, cos and sin are the functions known from real analysis.
Proposition 1.11. The complex exponential function has the following
properties:
Robert Oeckl – CA NOTES – 01/12/2015 7
1. exp is entire.
5. exp(C) = C \ {0}.
6. For each z ∈ C \ {0} there is a unique angle θ ∈ [0, 2π), called the
argument or phase, so that z = |z| exp(iθ).
Remark 1.13. This Proposition shows in particular that the complex ex-
ponential function is uniquely determined by the properties exp′ = exp and
exp(0) = 1.
√
Exercise 6. Define a function z 7→ z on C or on a subset of C. Is this
function holomorphic and if yes, where? Comment on possible choices in
the construction.
That is, the power series converges absolutely in the open disk Br (z0 ) to a
complex function f : Br (z0 ) → C. For any 0 < ρ < r the convergence is
uniform in the open disk Bρ (z0 ). It diverges for z outside of the closed disk
Br (z0 ).
Proof. Exercise.
Proof. (Adapted from Rudin.) Fix z0 ∈ D and r > 0 such that Br (z0 ) ⊆ D.
Suppose f is given by the power series (1) and converges in Br (z0 ). Consider
the power series
∞
∑
g(z) := (n + 1)cn+1 (z − z0 )n .
n=0
It is then enough to show that g(z) converges in Br (z0 ) and that g(z) is the
complex derivative of f for all z ∈ Br (z0 ). The statement (2) about the k-th
derivative follows then by iteration.
Firstly, it is clear by Lemma 1.18 that g(z) has the same radius of con-
vergence as f (z). In particular, g(z) converges in Br (z0 ). Fix z ∈ Br (z0 )
and define ξ := z − z0 . Then, set ρ arbitrarily such that |ξ| < ρ < r. Let
ζ ∈ Bs (0) \ {0} where s := ρ − |ξ| and set
f (z + ζ) − f (z)
h(ζ) := − g(z).
ζ
where
(ξ + ζ)n − ξ n
an (ζ) := − nξ n−1 .
ζ
Note that a0 (ζ) = 0 and a1 (ζ) = 0. By explicit computation we find for
n ≥ 2,
∑
n−1
an (ζ) = ζ kξ k−1 (ξ + ζ)n−k−1 .
k=1
1
|an (ζ)| < |ζ| n(n − 1)ρn−2 .
2
This implies
∞
1∑
|h(ζ)| < |ζ| |cn |n(n − 1)ρn−2 .
2 n=2
However, since ρ < r, the sum converges by Lemma 1.18 showing that there
is a constant M such that
|h(ζ)| < |ζ|M.
Robert Oeckl – CA NOTES – 01/12/2015 11
This completes the proof of (2). Finally , the uniqueness of the coefficients
cn follows from the special case of (2) given by
f (k) (z0 ) = k!ck .
= 2πcn rn .
12 Robert Oeckl – CA NOTES – 01/12/2015
Lemma 1.24 (Gutzmer Formula). Let z0 ∈ C and r > 0, and suppose the
power series
∞
∑
f (z) = cn (z − z0 )n
n=0
Proof. Since
∞
∑
f (z0 + reiθ ) = cn rn e−inθ
n=0
we have,
2 ∞
∑
f (z0 + reiθ ) = cn rn f (z0 + reiθ )e−inθ ,
n=0
This shows the claimed equality. The stated inequality is obtained by esti-
mating the integral through the maximum of its integrand.
n!M
|f (n) (z0 )| ≤ .
rn
Proof. Exercise.
|f (z)| ≤ a + b|z|c ∀z ∈ C,
2 Complex integration
2.1 Integration along paths
Definition 2.1. Let I = [a, b] ⊂ R be a compact interval. A continuous
map γ : I → C is called a curve in C. We denote the image of the curve
γ by |γ|. If γ(a) = γ(b), the curve is called closed. A curve γ : I → C
is called a path iff it is piecewise continuously differentiable. That is, there
exist points a = x0 < x1 < · · · < xn = b such that γ restricted to [xk−1 , xk ]
is continuously differentiable for all k ∈ {1, . . . , n}.
Definition 2.2. Let γ : [a, b] → C and γ̃ : [ã, b̃] → C be paths. We say that
γ̃ is a reparametrization of γ iff there exists a monotonous, continuous and
piecewise continuously differentiable map ϕ : [ã, b̃] → [a, b] with ϕ(ã) = a
and ϕ(b̃) = b and such that γ̃ = γ ◦ ϕ.
(b − t)x1 + (t − a)x2
γ(t) :=
b−a
with end points x1 and x2 , then its length should be |x2 − x1 | where we use
the standard Euclidean inner product on C. In general, we can approximate
a path by subdividing the interval on which it is defined and replacing the
pieces of paths in subdivisions by straight lines. The length of the path
should then be the limit of the sum of the lengths of these straight lines
when we make the subdivisions arbitrarily fine. That this limit exists is
due to the piecewise continuous differentiability property we have imposed.
(The limit does not necessarily exist for arbitrary curves, even if their image
is bounded.) The result is the following, which we state as a definition.
16 Robert Oeckl – CA NOTES – 01/12/2015
Exercise 10. (a) Show that the definition indeed agrees with the result of
the procedure described above. (b) Give an example of a curve that has
bounded image, but no well defined length.
Exercise 11. Show that the length of a path is invariant under reparametriza-
tion. That is, show that if γ is a path and γ̃ is a reparametrization of γ,
then l(γ) = l(γ̃).
Definition 2.4. Let U ⊆ C be open and f : U → C be a continuous map.
Let γ : I → C be a path such that |γ| ⊂ U . We define the complex integral
of f along the path γ as follows,
∫ ∫ b
f (z) dz := f (γ(t))γ ′ (t) dt. (3)
γ a
Proof. Exercise.
Proof. Exercise.
18 Robert Oeckl – CA NOTES – 01/12/2015
1 l(γ)
|Indγ (z)| ≤ < 1.
2π |z| − M
On the other hand Indγ (z) ∈ Z, so we must have in this case Indγ (z) = 0.
This completes the proof.
Proof. Exercise.
Conversely, suppose that F satisfies the stated formula for every path γ
in D. Let z ∈ D and choose r > 0 such that Br (z) ⊆ D. For ξ ∈ Br (0) let
γξ : [0, 1] → C be the path γξ (t) := z + tξ. By assumption,
∫ ∫ 1
F (z + ξ) − F (z) = f (ζ) dζ = f (z + tξ)ξ dt
γξ 0
For ξ ̸= 0 we get,
∫
F (z + ξ) − F (z) 1
= f (z + tξ) dt.
ξ 0
The right hand side of this expression converges to f (z) when |ξ| → 0 since,
(∫ ) ∫
1 1
f (z + tξ) dt − f (z) ≤ |f (z + tξ) − f (z)| dt
0 0
≤ sup |f (z + ζ) − f (z)|,
ζ∈B|ξ| (0)
where the right hand side expression converges to zero for |ξ| → 0 by conti-
nuity of f . Thus, F is complex differentiable at z with the differential being
F ′ (z) = f (z). This completes the proof.
Proof. If f is integrable, then by Theorem 2.12 the integral along any close
path must be zero. Conversely, suppose the integral of f along any closed
path is zero. Choose z0 ∈ D arbitrarily. Define
∫
F (z) := f (ζ) dζ,
γz
∆n into four triangles ∆n,1 , . . . , ∆n,4 by subdividing each of its sides into
two pieces of equal length. Now choose k ∈ {1, 2, 3, 4} such that the absolute
value ∫
f (z) dz
∆n,k
is maximized and set ∆n+1 := ∆n,k . This defines a sequence of triangles.
∩
Note that the intersection n∈N ∆n is a single point z0 ∈ D.
By the addition property of the integral along paths we have for every
n ∈ N the identity
∫ ∫ ∫ ∫ ∫
f= f+ f+ f+ f.
∂∆n ∂∆n,1 ∂∆n,2 ∂∆n,3 ∂∆n,4
and thus, ∫ ∫
f ≤ 4 n−1
f .
∂∆ ∂∆n
For the circumference of the triangles we obtain the relation,
1
l(∂∆n ) = n−1 l(∂∆). (5)
2
Now set ϵ > 0 arbitrarily and choose r > 0 such that Br (z0 ) ⊆ D and
|g(z)| ≤ ϵ|z − z0 |, where g(z) := f (z) − f (z0 ) − f ′ (z0 )(z − z0 )
for all z ∈ Br (z0 ). (This is possible since f is complex differentiable at z0 .)
Now fix n ∈ N such that ∆n ⊂ Br (z0 ). Note that the constant function and
the identity function are integrable so that with Proposition 2.13 we have,
∫ ∫ ∫
( )
f (z) dz = f (z0 ) + f ′ (z0 )(z − z0 ) + g(z) dz = g(z) dz.
∂∆n ∂∆n ∂∆n
Proof. Fix ϵ > 0. Denote the corner points of ∆ by (p, x, y). Define the
triangle ∆t for t ∈ [0, 1] as the triangle with corner points (p, xt , yt ), where
xt := p + t(x − p) and yt := p + t(y − p). Then, l(∂∆t ) → 0 as t → 0. By
continuity of f on the compact set ∆, Proposition 2.6 implies that there
exists t > 0 such that ∫
f < ϵ.
∂∆t
Now, subdivide the triangle ∆ into the triangle ∆t and the triangles with
corners given by (xt , x, y) and (xt , y, yt ). The integral over boundary paths
24 Robert Oeckl – CA NOTES – 01/12/2015
of the latter two triangles vanishes by Proposition 2.16. On the other hand,
the sum of the integrals over the boundary paths of the three triangles equals
the integral over the boundary path of ∆. Thus,
∫ ∫
f= f < ϵ.
∂∆ ∂∆t
The Cauchy Integral Formula is often used in the special case where the
path is the boundary of a disk: Let D ⊆ C be a region, f ∈ O(D), z ∈ D
and r > 0 such that Br (z) ⊂ D. Then,
∫
1 f (ζ)
f (z) = dζ.
2πi ∂Br (z) ζ −z
Lemma 2.21. Let U ⊆ C be open, f : U → C continuous and γ a closed
path in U . Define the function F : C \ |γ| → C via
∫
f (ζ)
F (z) := dζ.
γ ζ −z
Then, F is analytic in C \ |γ|. Moreover, for all n ∈ N0 ,
∫
(n) f (ζ)
F (z) = n! dζ.
γ (ζ − z)n+1
Proof. Fix z0 ∈ C \ |γ| and define for all n ∈ N0 ,
∫
f (ζ)
cn := dζ.
γ (ζ − z0 )n+1
Set r := inf t∈[a,b] |γ(t) − z0 |. We proceed to show that the power series
∞
∑
G(z) := cn (z − z0 )n
n=0
converges in Br (z0 ) and agrees there with F (z). Fix z ∈ Br (z0 ). Define the
partial sums gn : |γ| → C for n ∈ N0 via,
∑
n
f (ζ)(z − z0 )k
gn (ζ) := .
k=0
(ζ − z0 )k+1
Since |ζ −z0 | ≥ r > |z−z0 | and f is bounded on |γ|, the sequence of functions
{gn }n∈N0 converges uniformly on |γ|. Thus, by Proposition 2.7,
∫ ∫
G(z) = lim gn (ζ) dζ = lim gn (ζ) dζ.
n→∞ γ γ n→∞
This Theorem finally yields the remarkable result that holomorphic func-
tions are analytic. Together with Theorem 1.20 this means that the proper-
ties of holomorphicity and analiticity are really equivalent. Furthermore, it
implies that the derivative of a holomorphic function is again a holomorphic
function.
Robert Oeckl – CA NOTES – 01/12/2015 27
2. f is analytic in D.
3. f is locally analytic in D.
4. f is locally integrable in D.
Proof. Exercise.
Exercise 19. Calculate the following integrals. [Hint: Use the Cauchy
Integral formula]
1. ∫
ez
dz
∂B2 (0) (z + 1)(z − 3)2
2. ∫
1
dz
∂B2 (−2i) z2 +1
Exercise 20. Determine all entire functions f ∈ O(C) which satisfy the
differential equation f ′′ + f = 0.
is holomorphic.
3. Intγ ⊂ D.
However, the series on the right hand side converges for all ρ < r to a
continuous function which goes to 0 when ρ → 0. Thus, g is continuous
at (z, z). Since z was arbitrary in D this completes the proof that g is
continuous in D × D. Exercise.Show that g is holomorphic in the second
argument. Now we apply Lemma 2.27 to conclude that the function h :
D → C defined by ∫
h(z) := g(ζ, z) dζ
γ
is holomorphic in D.
Now observe that for z ∈ D \ |γ| we have
∫ ∫
f (ζ)
h(z) = g(ζ, z) dζ = −2πi f (z)Indγ (z) + dζ.
γ γ ζ −z
30 Robert Oeckl – CA NOTES – 01/12/2015
3 Basic properties
3.1 From local to global structure
Definition 3.1. Let T be a topological space and A a subset. We say that
p ∈ A is an isolated point of A in T iff there exists a neighborhood U ⊆ T
of p such that U ∩ A = {p}. We say that A is discrete in T iff all its points
are isolated.
Theorem 3.2 (Riemann Continuation Theorem). Let D ⊆ C be a region
and A ⊂ D a discrete and relatively closed subset. Suppose that f ∈ O(D \
A). Then, the following assertions are equivalent.
1. f extends to a holomorphic function on D.
2. The coincidence set {z ∈ D|f (z) = g(z)} is not empty and not discrete.
3. There exists a point z0 ∈ D such that f (n) (z0 ) = g (n) (z0 ) for all n ∈ N.
Proof. The implication 1.⇒2. is trivial. We show 2.⇒3. Let h := f − g.
Suppose z0 ∈ {z ∈ D|h(z) = 0} is not an isolated point. Suppose there
exists m ∈ N0 such that h(m) (z0 ) ̸= 0 and choose the smallest such m. Since
h is holomorphic in D it is also analytic by Theorem 2.22 and has a power
series expansion around z0 for some radius r > 0, given by
∞
∑ h(n) (z0 )
h(z) = (z − z0 )n = (z − z0 )m k(z),
n=m n!
such that Bρ (z) ⊆ U . Then, the power series converges with radius at least
ρ and for 0 < r < ρ we have, by Lemma 1.24,
∞
∑
|cn |2 r2n ≤ ∥f ∥2∂Br (z) ≤ ∥f ∥2U = |f (z)|2 = |c0 |2 .
n=0
This implies ck = 0 for all k ∈ N, i.e., f is constant in Bρ (z). But then the
Identity Theorem (Theorem 3.3) ensures that f is constant in all of D.
Proof. Exercise.
Exercise 22. Let f, g be entire functions satisfying |f (z)| ≤ |g(z)| for all
z ∈ C. Show that there is a ∈ C such that f = ag.
1. f (D ∩ R) ⊆ R.
Exercise 26. For each of the following properties give an example for a
holomorphic function defined in some disk around 0 with that property or
show that there can be no such function.
3.2 Zeros
Definition 3.9. Let D ⊆ C be a region, z0 ∈ D and f ∈ O(D) such that
f (z0 ) = 0. We say that f has a zero of order n at z0 iff there exists g ∈ O(D)
such that g(z0 ) ̸= 0 and f (z) = (z − z0 )n g(z) for all z ∈ D.
Proof. Exercise.
p(z) = cn (z − a1 ) · · · (z − an ).
Proof. Exercise.[Hint: First show the existence of one zero and factorize
it, then proceed recursively.]
∑ ∫
m
1 f ′ (z)
nk Indγ (ak ) = dz.
k=1
2πi γ f (z)
where g ∈ O(D) has no zeros in D. Using the product rule for the derivative
we find for z ∈ D \ {a1 , . . . , an },
f ′ (z) g ′ (z) ∑m
nk
= + .
f (z) g(z) k=1 z − ak
1. na (f g) = na (f ) + na (g).
4 Singularities
4.1 Types of singularities
Definition 4.1. Let D ⊆ C be a region, a ∈ D and f ∈ O(D \ {a}).
Then, we say that f has an isolated singularity at a. Moreover, a is called
a removable singularity iff f can be extended to a holomorphic function on
all of D.
g(z)
f (z) = ∀z ∈ D \ {a}.
(z − a)m
Proof. Since f has a pole at a there exist r > 0 such that Br (a) ⊆ D
and f (z) ̸= 0 for all z ∈ Br (a) \ {a}. Thus we can define h ∈ O(Br (a) \
{a}) by h(z) := 1/f (z). But limz→a h(z) = 0, so by Theorem 3.2, a is a
removable singularity of h and h can be extended to a holomorphic function
on all of Br (a). By Proposition 3.10 there exists a unique m ∈ N such
that h(z) = (z − a)m k(z), where k ∈ O(Br (a)) and k(a) ̸= 0. Moreover,
k(z) ̸= 0 for all z ∈ Br (a) so we can invert it, defining g ∈ O(Br (a)) by
g(z) = 1/k(z). But notice that g(z) = (z − a)m f (z) for all z ∈ Br (a) \ {a},
which obviously extends to a holomorphic function on D \ {a}. So g really
extends to a holomorphic function on all of D. Observe also that g(a) ̸= 0.
This completes the proof.
∑
m
bn
f (z) = g(z) + ∀z ∈ D \ {a}.
n=1
(z − a)n
Proof. Exercise.
The second term on the right hand side of the equation above is also
called the singular part of f at a.
We now turn to essential singularities. In some sense they are more
“wild” than poles, as shows the following Theorem.
Proof. We start with the implication 1.⇒2. Assume the contrary of 2. Let
U ⊆ D be a neighborhood of a such that f (U \ {a}) is not dense in C.
Thus, there exists p ∈ C and r > 0 such that f (U \ {a}) ∩ Br (p) = ∅.
This implies |f (z) − p| ≥ r for all z ∈ U \ {a}. Define g ∈ O(U \ {a}) by
g(z) := 1/(f (z)−p). Then, |g(z)| ≤ 1/r for all z ∈ U \{a} so by Theorem 3.2,
g has a removable singularity at a. Thus, c := limz→a g(z) exists. If c ̸= 0,
f (z) = p + 1/g(z) is bounded near a and thus has a removable singularity
at a. If c = 0, then limz→a |f (z)| = ∞ and f has a pole at a. In both cases,
a is not an essential singularity, contradicting 1. Exercise.Complete the
proof.
Exercise 30. Find and classify the isolated singularities of the following
functions and specify the order in case of a pole:
z4 1 − cos(z)
1. 2.
(z 4 + 16)2 sin z
1
3. exp(1/z) 4.
cos(1/z)
Robert Oeckl – CA NOTES – 01/12/2015 41
Exercise 32. Investigate how the different types of singularities behave with
respect to addition, multiplication, quotienting and composition (whenever
the corresponding operations make sense)!
Proof. Exercise.
Exercise 35. Show that the set of rational functions forms a proper subfield
of M(C).
∑ ∑ ∫
1 f ′ (z)
N (z)Indγ (z) − N (z)Indγ (z) = dz,
z∈Z z∈P
2πi γ f (z)
Then,
∑ ∑ ∑ ∑
N (z)Indγ (z)− N (z)Indγ (z) = N (z)Indγ (z)− N (z)Indγ (z),
z∈Zf z∈Pf z∈Zg z∈Pg
Proof. First, note that the inequality also implies |γ|∩Zf = ∅ and |γ|∩Zg =
∅. Set U := D \ (Zf ∪ Zg ∪ Pf ∪ Pg ) and h(z) := f (z)/g(z) for all z ∈ U .
Then, h ∈ O(U ). Note that the hypothesis is equivalent to the inequality
Exercise 39. Determine the number of zeros (counted with order) of the
following functions in the specified domain:
1. z 5 + 13 z 3 + 14 z 2 + 1
3 in B1 (0) and in B1/2 (0).
This type of region is called an (open) annulus. Note the special case of the
punctured disk A0,r (z) = Br (z) \ {z}.
Proposition 4.12. Let {cn }n∈Z be an indexed set of complex numbers. De-
fine r1 , r2 ∈ [0, ∞] via
r1 := lim sup |c−n |1/n and 1/r2 := lim sup |cn |1/n .
n→∞ n→∞
converges absolutely for all z ∈ Ar1 ,r2 (0) and uniformly on Aρ1 ,ρ2 (0) where
r1 < ρ1 < ρ2 < r2 . Moreover, it diverges for z ∈ C \ Ar1 ,r2 (0).
Proof. Exercise.[Hint: Split the series into the parts with positive and
negative indices and apply Lemma 1.18.]
Proof. Exercise.
where the series converges absolutely for all z ∈ Ar1 ,r2 (z0 ) and uniformly on
As1 ,s2 (z0 ), when r1 < s1 < s2 < r2 . Also, the coefficients are given by
∫
1 f (ζ)
cn = dζ,
2πi ∂Br (z0 ) (ζ − z0 )n+1
which converges pointwise in B1/r1 (0) and uniformly in B1/s1 (0) for any
s1 > r1 . Thus
( ) ∞
∑
− 1
f (z) = g = bn (z − z0 )−n
z − z0 n=1
converges pointwise in C \ Br1 (z0 ) and uniformly on C \ Bs1 (z0 ) for any
s1 > r1 . On the other hand, the power series expansion
∞
∑
f + (z) = cn (z − z0 )n
n=0
converges pointwise in Br2 (z0 ) and uniformly on Bs2 (z0 ) for any 0 < s2 < r2 .
Summing both expansions and setting c−n := bn for all n ∈ N yields the
Laurent series with the desired properties.
Set r1 < r < r2 . Using Lemma 2.10 together with convergence of the
Laurent series and interchangeability of limit and integral (Proposition 2.7)
yields the desired formula for the coefficients cn .
Proof. Exercise.
Give the Laurent series expansion of f in the following regions: A0,1 (0),
A1,2 (0), A2,∞ (0).
4.4 Residues
Definition 4.18. Let a ∈ C and 0 < r, f ∈ O(Br (a) \ {a}) and
∑
f (z) = cn (z − a)n
n∈Z
Proof. Define à := Intγ ∩ A. This is finite since Intγ ∪ |γ| is compact. Thus,
suppose à = {a1 , . . . , an }. Observe that the sum in the statement really
only runs over Ã, since the index of the other elements of A vanishes. Now,
decompose f into a sum
and notice that g has no singularities in Intγ left. Note that the integral
over g along γ vanishes by Theorem 2.28. Thus, the Theorem reduces to
proving the identity,
∫
1
Res(h, a)Indγ (a) = h(z) dz
2πi γ
for functions h ∈ O(C \ {a}) such that lim|z|→∞ h(z) = 0. Consider the
Laurent series of h around a,
−1
∑
h(z) = cn (z − a)n .
n=−∞
Since this converges uniformly on the compact set |γ|, we can interchange
integration and summation,
∫ −1
∑ ∫
h(z) dz = cn (z − a)n dz.
γ n=−∞ γ
Now note that (z − a)n has a primitive if n ≤ −2, i.e., is then integrable in
C \ {a}. Thus, by Proposition 2.13 its integral vanishes. Hence,
∫ ∫
h(z) dz = c−1 (z − a)−1 dz = Res(h, a)2πiIndγ (a).
γ γ
5 Conformal mappings
5.1 Conformal mappings as holomorphic functions
Recall that we have the standard Euclidean scalar product on the complex
plane, by viewing C as a two-dimensional real vector space. That is, we
have
⟨z, z ′ ⟩ := aa′ + bb′ = ℜ(zz ′ ),
√
where z = a+ib and z ′ = a′ +ib′ . Recall also that |z| = ⟨z, z⟩. In geometric
terms we have,
⟨z, z ′ ⟩ = |z||z ′ | cos θ,
where θ is the angle between z and z ′ , viewed as vectors in the complex
plane.
We shall now be interested in mappings A : C → C that preserve angles
between intersecting curves. First, we consider R-linear mappings. Then,
for A to be angle-preserving clearly is equivalent to the identity,
Proof. Exercise.
Proof. Exercise.
In particular, inf ζ∈∂Br (z) |f (ζ) − a| > |f (z) − a|. Thus, by Proposition 3.8
f − a must have zeros in the region Br (z). That is, there exists ξ ∈ Br (z)
such that f (ξ) = a.
Proof. Let a ∈ D and p := f (a). Suppose first that f ′ (a) = 0. Then, either
f is constant or f − p has a zero of order m ≥ 2 at a. In the first case
the lack of local injectivity is trivial. In the second case consider an open
neighborhood U ⊆ D of a. Applying Theorem 5.8, there exists ϵ > 0 such
that for q ∈ Bϵ (p) \ {p} the equation f (z) = q has at least two distinct
solutions for z ∈ U . In particular, f is not injective in U . Since U was
arbitrary, f is not locally injective at a.
Now suppose f ′ (a) ̸= 0. Then, f − p has a simple zero at a. Applying
Theorem 5.8, there exist ϵ > 0 and δ > 0 with Bδ (a) ⊂ D such that for
all q ∈ Bϵ (p) the equation f (z) = q has exactly one solution in Bδ (a). By
continuity of f , U := f −1 (Bϵ (p)) ∩ Bδ (a) is an open neighborhood of a.
Clearly, f is injective in U , showing that f is locally injective at a.
52 Robert Oeckl – CA NOTES – 01/12/2015
Exercise 47. Show that the group generated by translations, rotations and
scalings of C consists precisely of the biholomorphic transformations C → C
of the form
z 7→ az + b with a ∈ C \ {0}, b ∈ C.
z 7→ az + b for some a ∈ C× , b ∈ C.
Proof. Let
∞
∑
f (z) = cn z n
n=0
Replacing z by f (z) in the second inequality yields, |z| ≤ |f (z)| for all z ∈ D.
Thus, we actually find |f (z)| = |z| for all z ∈ D. By Lemma 5.17 this implies
that there exists a ∈ C with |a| = 1 and f (z) = az, i.e., f is a rotation.
Proof. Exercise.
Exercise 52. Verify that the upper triangular matrices (with non-vanishing
determinant) form a subgroup of GL2 (C). Show that the image of this
subgroup under the map GL2 (C) → Möb is the group Aut(C). Identify
the upper triangular matrices corresponding to translations, rotations and
dilations.
Exercise 53. Verify that the other Möbius transformations also define bi-
holomorphic mappings. Between which regions?
Recall that GL+
2 (R) is the group of orientation-preserving general linear
transformations of R2 , i.e., these are 2 × 2-matrices with real entries and
positive determinant.
Proposition 5.21. The restriction of the map GL2 (C) → Möb to the sub-
group GL+ 2 (R) yields Möbius transformations that are conformal automor-
phisms of H. That is, we obtain a group homomorphism GL+ 2 (R) → Aut(H).
Proof. Exercise.
Exercise 57. Show that PGL+ 2 (R) is isomorphic to SL2 (R)/Z2 , where Z2
is the subgroup of SL2 (R) consisting of {1, −1}.
{rϵ
}
Choosing δ := min r, 4M yields the estimate
Proof. Let z0 ∈ D and set r > 0 such that Br (z) ⊂ D. By Proposition 2.17
fn is integrable in Br (z0 ). For any closed path γ in Br (z0 ) we thus have
∫ ∫ ∫
f= lim fn = lim fn = 0,
γ γ n→∞ n→∞ γ
where we have used Proposition 2.7 to interchange the integral with the
limit. Thus, f is integrable in Br (z0 ) and hence holomorphic there by Theo-
rem 2.25. Since the choice of z0 was arbitrary we find that f is holomorphic
in all of D.
Fix k ∈ N and consider z0 ∈ D. Choose r > 0 such that B2r (z0 ) ⊆ D.
Now for each z ∈ Br (z0 ) we have the Cauchy estimate (Proposition 1.25),
k! k!
|fn(k) (z) − f (k) (z)| ≤ k
∥fn − f ∥∂Br (z) ≤ k ∥fn − f ∥B2r (z0 ) .
r r
For ϵ > 0 there is by uniform convergence of {fn }n∈N an n0 ∈ N such
that |fn (z) − f (z)| < ϵ rk /k! for all n ≥ n0 and all z ∈ B2r (z0 ). Hence,
(k) (k)
|fn (z) − f (k) (z)| < ϵ for all n ≥ n0 and all z ∈ Br (z0 ). That is, {fn }n∈N
converges to f (k) uniformly on some neighborhood of every point of D.
To obtain uniform convergence on a compact subset K ⊂ D it is merely
necessary to cover K with finitely many such neighborhoods.
62 Robert Oeckl – CA NOTES – 01/12/2015
g ∈ O(D) with g 2 = f . If g(z1 ) = g(z2 ) then (g(z1 ))2 = (g(z2 ))2 and
so z1 = z2 since f is injective. Therefore also g is injective. Moreover, if
g(z1 ) = −g(z2 ) we can draw the same conclusion z1 = z2 , but this time we
get a contradiction, since g is zero-free. Thus, if z ∈ C is in the image of
g, then −z cannot be in the image of g. Now since g is not constant the
Open Mapping Theorem 5.6 ensures that g(D) is open. In particular there
exists w ∈ C and r > 0 such that Br (w) ⊂ g(D). But applying the previous
statement to all elements of Br (w) we obtain Br (−w) ∩ g(D) = ∅. It is now
easy to see that the function h ∈ O(D) defined by h(z) := r/(g(z) + w) is
also injective and satisfies h(D) ⊆ D. Setting v := h(z0 ), we have Dv ◦ h ∈ F
since Dv ∈ Aut(D) and Dv (v) = 0.
Since D is open, there exists r > 0 such that Br (z0 ) ⊂ D. Using the
Cauchy estimate (Proposition 1.25) we find the bound |f ′ (z0 )| < 1/r for all
f ∈ F . This implies that
M := sup{|f ′ (z0 )| : f ∈ F }
is well defined. On the other hand we will show that if f (D) ̸= D for some
f ∈ F , then there exists g ∈ F such that |g ′ (z0 )| > |f ′ (z0 )|. This implies
that h ∈ F is a biholomorphism D → D if |h′ (z0 )| = M . We will then show
that such an h exists.
Consider some f ∈ F such that f (D) ̸= D. Choose p ∈ D \ f (D).
Since Dp ∈ Aut(D), the composition Dp ◦ f is injective and Dp ◦ f (D) ⊂ D.
Furthermore, Dp ◦f is zero-free since Dp−1 (0) = {p}. Since D is homologically
simply connected we can find a holomorphic square root g ∈ O(D) with
g 2 = Dp ◦ f according to Theorem 3.18. In fact, it is clear that g is injective
and g(D) ⊆ D. Set w := g(z0 ). Then h := Dw ◦ g ∈ F . Consider now the
holomorphic map k : D → D given by k(z) = Dp ((Dw (z))2 ). Then, f = k ◦ h
and applying the chain rule for derivatives we obtain
Noting that k(0) = 0 we can apply the Schwarz Lemma 5.17. Since k is not
a rotation, this implies |k ′ (0)| < 1. Hence, |f ′ (z0 )| < |h′ (z0 )| since h′ (z0 ) ̸= 0
by injectivity of h.
The image of all functions in F is contained in the bounded set D, so in
particular F is locally bounded. According to Montel’s Theorem 5.31 this
implies that F is normal. Consider now a sequence {fn }n∈N of elements
of F such that |fn′ (z0 )| → M as n → ∞. Since F is normal, there is a
subsequence {fnk }k∈N which converges uniformly on any compact subset of
D to a function f ∈ O(D) by Proposition 5.32. By the same Proposition
64 Robert Oeckl – CA NOTES – 01/12/2015
Proof. Exercise.
6 Harmonic functions
6.1 Mean value and maximum
We coordinatize the complex plane by coordinates (x, y) ∈ R2 with z =
x + iy ∈ C. The Laplace operator on the complex plane is then given by
∂2 ∂2
∆ := + .
∂x2 ∂y 2
Proposition 6.2. The real and the imaginary part of a holomorphic func-
tion are harmonic.
Proof. Exercise.
Proof. Exercise.
Note that the interchange of differentiation and integration in the first step
is permitted since the integrand is continuously differentiable and the in-
tegration range compact. On the other hand, differentiating by ∂/∂y we
obtain
vy (x, y) = ux (x, y).
Proof. If D = C then Lemma 6.4 directly applies and we are done. Suppose
therefore that D ̸= C. By the Riemann Mapping Theorem 5.36 there exists
a biholomorphic map f : D → D. By Proposition 6.3, u ◦ f −1 : D → R is
harmonic. Applying Lemma 6.4, there exists a harmonic function w : D → R
such that u ◦ f −1 + iw : D → C is holomorphic. Define v : D → R by
v := w ◦ f . Then, v is harmonic by Proposition 6.3 and u + iv : D → C is
holomorphic.
Proof. Exercise.
Proof. Choose s > r such that Bs (a) ⊆ D. By Theorem 6.5 there exist a
harmonic function v : Bs (a) → R such that f := u + iv : Bs (a) → C is
holomorphic. Applying the Cauchy Integral Formula (Theorem 2.20) to f
at the point a with path ∂Br (a) we obtain,
∫ ∫ 2π ( )
1 f (ζ) 1
f (a) = dζ = f a + reiθ dθ.
2πi ∂Br (a) ζ −a 2π 0
Taking the real part on both sides yields the desired result.
Robert Oeckl – CA NOTES – 01/12/2015 67
Proof. Define
A := {z ∈ D : u(z) = u(a)}.
Since u is continuous, A must be closed in D. We proceed to show that A is
also open. Let z0 ∈ A and r > 0 such that Br (z0 ) ⊆ D. Choose b ∈ Br (z0 )
and set s := |b − z0 |. By the mean value property
∫ 2π ( )
1
u(z0 ) = u z0 + seiθ dθ.
2π 0
Proof. Exercise.
Exercise 61. Show the following version of the maximum principle, which
is more similar to Theorem 3.5: Let D ⊆ C be a region and f : D → C
a continuous function satisfying the mean value property. Suppose that |f |
has a maximum at some point a ∈ D, i.e., that |f (z)| ≤ |f (a)| for all z ∈ D.
Then f is constant. [Hint: Consider the function g(z) := ℜ(f (z)/f (a)).]
68 Robert Oeckl – CA NOTES – 01/12/2015
1. P is harmonic.
1 − |z|2 1 − r2
Pr (θ) = P (z) = = .
|1 − z|2 1 − 2r cos θ + r2
7. For all 0 < r < 1 and 0 < |δ| < |θ| ≤ π we have Pr (θ) < Pr (δ).
8. For each 0 < δ < π and ϵ > 0 there exists 0 < ρ < 1 such that for all
ρ < r < 1 and δ < |θ| ≤ π we have |Pr (θ)| < ϵ.
7. This follows easily from 2. 8. Fix 0 < δ < π and ϵ > 0. Then, Pr (δ) → 0
for r → 1− using 2. Thus, there is 0 < ρ < 1 so that |Pr (δ)| < ϵ if ρ < r < 1.
Using 7. completes the proof of 8.
Proof. Define u(z) for z ∈ D by the stated formula and u(z) := b(z) for
z ∈ ∂D. We first show that u is harmonic in D. We note that for z ∈ D,
∫ ( )
1 π 1 + ze−iϕ ( )
u(z) = ℜ b eiϕ dϕ
2π −π 1 − ze−iϕ
( ∫ )
1 π eiϕ + z ( iϕ )
=ℜ b e dϕ .
2π −π eiϕ − z
Here, we have used the properties of the Poisson kernel given in Proposi-
tion 6.12 parts 3. and 6.
It remains to show uniqueness of the function u. Suppose there was
another function v : D → R with the required properties. Then, the dif-
ference u − v would be continuous on D and harmonic in D. Furthermore,
(u − v)|∂D = 0, so by Proposition 6.10, u − v = 0, i.e., u = v.
Proof. Exercise.
Remark 7.4. The topological space Ĉ together with the structures intro-
duced in the preceding Proposition is called the Riemann sphere. It is an
example of a complex manifold. The maps ϕ0 , ϕ∞ are called charts.
Exercise 62. Let {zn }n∈N be a sequence of complex numbers. Show that
limn→∞ zn = ∞ in Ĉ if and only if for each M > 0 there exists n0 ∈ N such
that |zn | > M for all n ≥ n0 .
Exercise 63. Consider the symmetric function d : Ĉ × Ĉ → R+
0 given by
2|z − z ′ |
d(z, z ′ ) := √ ∀z, z ′ ∈ C
(1 + |z|2 )(1 + |z ′ |2 )
2
d(∞, z) := √ ∀z ∈ C
1 + |z|2
d(∞, ∞) := 0.
Show that d defines a metric on the Riemann sphere that is compatible with
its topology.
Remark 7.5. The metric introduced above can be obtained from the stere-
ographic projection of Ĉ identified with the unit disk to the complex plane.
74 Robert Oeckl – CA NOTES – 01/12/2015
7.2 Functions on Ĉ
Exercise 64. Let D ⊆ Ĉ be a region and f : D → C be continuous. Let
a ∈ D \ {0, ∞}. Show that f ◦ ϕ−1
0 is holomorphic/conformal at ϕ0 (a) iff
−1
f ◦ ϕ∞ is holomorphic/conformal at ϕ∞ (a).
Definition 7.6. Let D ⊆ Ĉ be a region and f : D → C be continuous.
Let a ∈ D. If a ̸= ∞, we say that f is holomorphic/conformal at a iff
f ◦ ϕ−1
0 is holomorphic/conformal at ϕ0 (a). If a ̸= 0, we say that f is
−1 is holomorphic/conformal at ϕ (a).
holomorphic/conformal at a iff f ◦ ϕ∞ ∞
We say that f is holomorphic/conformal in D iff f is holomorphic/conformal
at each point a ∈ D.
Exercise 65. Let D ⊆ Ĉ be a region and a ∈ D\{0, ∞}. Let f ∈ O(D\{a}).
Show that the type and order of the singularity of f ◦ ϕ−10 at ϕ0 (a) is the
same as the type and order of the singularity of f ◦ ϕ−1
∞ at ϕ∞ (a).
Definition 7.7. Let D ⊆ Ĉ be a region, a ∈ D and f ∈ O(D \ {a}).
If a ̸= ∞, we say that f has a removable singularity/a pole of order n/an
essential singularity at a iff f ◦ϕ−1
0 has a removable singularity/a pole of order
n/an essential singularity at ϕ0 (a). If a ̸= 0, we say that f has a removable
singularity/a pole of order n/an essential singularity at a iff f ◦ ϕ−1 ∞ has a
removable singularity/a pole of order n/an essential singularity at ϕ∞ (a).
Proposition 7.8. Let f ∈ O(Ĉ). Then, f is constant.
Proof. Exercise.
Proof. Exercise.
Theorem 7.16. Let (a, b, c) and (a′ , b′ , c′ ) be triples of distinct points in Ĉ.
Then, there exists exactly one Möbius transformation f such that f (a) = a′ ,
f (b) = b′ , f (c) = c′ .
Proof. Exercise.