Complex Analysis - Script[1]
Complex Analysis - Script[1]
Andreas Compagnoni
Contributions by:
Alberto De Stefani
Simeon Barbey
Viktor Kusar
Noah Talarico
This work follows quite accurately the course taught by Professor O.Imamoglu at
ETH Zürich in the years 2023 and 2024, based once again on the book by Stein and
Shakarchi [SS10], properly cited during the proceedings. The material has then under-
gone some changes in the notation and also some slight changes in the reformulation of
a few concept. It does not contain nevertheless any change in the content itself, which
is a good thing, given the title on the title-page.
In case the reader shall notice any imprecision, mistake, typo or similar, we kindly
encourage the reader to report them by sending an e-mail to:
acompagnoni@ethz.ch
specifying “Complex Analysis - ” in the object, immediately followed by the topic under
discussion. Please note that “relatively long” times of response ought to be expected.
(i) It is possible that the reader might still see some parts of this work being written
in red. These are my comments or remarks about the content: they might of
interest (expecially to understand the notation), but are not necessarily content
covered by the Professor in class. This also applies to the Appendix B.
(ii) To avoid misinterpretation of notation, when possible, hence when a line ends with
a mathematical symbol with nothing following (except a displayed mathematical
formula), the last punctuation symbol of the above mentioned line is omitted.
We shall conclude by showing the very last blackboard of the course, to keep in mind
while reading.
iii
iv
Sincerely,
Andreas Compagnoni
Contents
0 Introduction 1
v
vi CONTENTS
References 193
Chapter 0
Introduction
f :C→C
We will see that the study of Complex Function Theory is not simply the study of
functions on R2 : in fact, the theory of functions of one real variable is in many ways
more complicated than the theory of functions of a complex variable.
1. It is not too difficult to find a function of a real variable that is in Dn (R) but not
in D∞ (R). Consider
2. There are functions f : R → R that are infinitely many times differentiable, whose
Taylor series does not represent f , i.e. f is not analytic. E.g.
exp −1
2 , x ∈ R \ {0}
f : R → R, x 7→ f (x) = x
0 , x ∈ {0}
1
2 CHAPTER 0. INTRODUCTION
3. There are plenty of C ∞ (R) functions of a real variable that are bounded, e.g.
sin(x), cos(x)
In contrast: we will see that if f : C → C is differentiable and bounded, then it
is constant (Liouville’s Theorem 2.8)
4. For two functions of a real variable f, g, they both can ”agree” (be equal) on an
open set without being equal.
In contrast: if f, g : C → C are two differentiable functions which coincide on
an arbitrarily small disc (or even a convergent sequence (zn )n∈N ), then f = g
(Analytic continuation principle 2.10)
Remark 0.1. The power of Complex Function Theory comes from this ”robustness”
or rigidity. It is a field in which, in some sense, Analysis, Geometry and Algebra come
together.
This, we will see, allows one to prove Theorems that a priori seem to have nothing to
do with complex numbers.
Example 0.1. 1. The integral
Z ∞ Z ∞
√
2 2 2π
cos(t )dt = sin(t )dt =
0 0 4
2. Let π(x) := #{p ∈ P : p ≤ x} with P denoting the set of prime numbers. Then
x
π(x) ≈x→∞
log(x)
Result know as the Prime Number Theorem A.5, as
π(x)
lim x =1
x→∞
log(x)
4. Let r4 (n) := #{(m1 , ..., m4 ) ∈ Z4 : 4l=1 m2l = n}, then r4 (n) = 8 d|n d
P P
4∤d
C := {x + iy : x, y ∈ R and i2 = −1}
∃ιR : R → C, r 7→ r + i · 0
Re(z) := x
3
4 CHAPTER 1. PRELIMINARIES TO COMPLEX ANALYSIS
z̄ := x − iy
(iii) z ∈ R ⇐⇒ z = z̄
(iv) z ∈ iR ⇐⇒ z = −z̄
2x x+iy+x−iy z+z̄
Proof. (i) Re(z) = x = 2
= 2
= 2
Algebraic Structure of C
Complex numbers can also be represented as ordered pairs of real numbers in R2 , so
for z ∈ C we have that
z∼= (x, y)
1.1. THE COMPLEX NUMBERS AND THE COMPLEX PLANE 5
z + w := (x + u) + i(y + v)
or as pairs in R2
z+w ∼
= (x + u, y + v)
or as pairs in R2
z·w ∼ = (xu − yv, xv + yu)
Note that i ∼
= (0, 1) and (0, 1) · (0, 1) = (−1, 0) ∼
= −1, as i2 = −1
R2 with these two operations +, · becomes a field, i.e. (R, +, ·) satisfies the following:
Additive inverse of z ∈ C:
−z ∼
= (−x, −y)
6 CHAPTER 1. PRELIMINARIES TO COMPLEX ANALYSIS
Multiplicative inverse of z ∈ C:
z̄ x −y
z −1
= 2 ∼= , 2
|z| x + y x + y2
2 2
arg(z) := θ ∈ R : z ∈ |z|eiθ
Definition 1.6. The argument of z ∈ C chosen in the interval (−π, π] is called the
principal argument of z ∈ C and denoted by Arg(z) ∈ arg(z)
π
It holds that: Arg(i) = 2
and ∀c ∈ (0, +∞) : Arg(−c) = π
Proof. Here is enough to observe that for t ∈ [−1, 1], arcsin is the unique number
u ∈ −π , π such that sin(u) = t. By using the various mirroring in the unit circle
2 2
(indeed we divided by |z|), we obtain the result.
arg(z −1 ) = − arg(z)
arg(zw) = arg(z) + arg(w)
= − θ ∈ R : z = |z|eiθ = − arg(z)
= θz + θw ∈ R : zw = |zw|ei(θz +θw ) =
= θz + θw ∈ R : zw = |z|eiθz |w|eiθw =
= arg(z) + arg(w)
Remark 1.2. Despite these results, it is not always the case that for z ∈ C we have
Arg(z −1 ) = − Arg(z)
Arg(zw) = Arg(z) + Arg(w)
as shown in the following example.
8 CHAPTER 1. PRELIMINARIES TO COMPLEX ANALYSIS
On the other hand, we have zw = (ac − bd) + i(bc + ad). The multiplication in C hence
corresponds to the multiplication of the respective matrices in R2×2
Topological results
Proposition 1.4. The following properties hold for the complex norm | · |:
(i) ∀z ∈ C : |z| = 0 ⇐⇒ z = 0
and since this holds for all z1 , z2 ∈ C, then by swapping z1 and z2 we obtain the
same inequality in absolute value, namely
√ √ √ √
(iii) Let z1 , z2 ∈ C, then |z1 z2 | =
z1 z2 z¯1 z¯2 = z1 z¯1 z2 z¯2 = z1 z¯1 z2 z¯2 = |z1 ||z2 |
p
(iv) For z ∈ C we have |z̄| = |x − iy| = x2 + y 2 = |x + iy| = |z|
|Re(z)|2 = x2 ≤ x2 + y 2 = |z|2
|Im(z)|2 = y 2 ≤ x2 + y 2 = |z|2
√
Applying the · function, which is a monotonically increasing function, we obtain
the wished result.
zw = rsei(θ+ν) ∈ C∗
10 CHAPTER 1. PRELIMINARIES TO COMPLEX ANALYSIS
Since both r, s are non-zero then also their product is not, hence zw ∈ C∗
Next, we recall some definitions that we need from Topology and Analysis.
Definition 1.7. We denote the open disc of radius r > 0 centred at z with
Dr (z) or D(z, r) and the closed disc of radius r ≥ 0 centred at z with Dr (z) or
D(z, r). They are both defined as follows
Dr (z) := {w ∈ C : |w − z| < r}
Dr (z) := {w ∈ C : |w − z| ≤ r}
Remark 1.3. If r > 0, then ∀z ∈ C : Dr (z) = Dr (z), hence in this case the closed disc
is equal to the closure of the open disc.
∀z ∈ U ∃r > 0 : Dr (z) ⊆ U
The set of all such open subsets of C is called the standard topology of open
sets of C and denoted by OC
∗
Proposition 1.7. K ⊆ C is compact, if and only if every sequence (zn )n∈N∗ ∈ U N
has a subsequence that converges to a point in U
We mention here that in any euclidean space: a connected open set is automatically
open and path-connected and vice versa. Any two distinct points z0 , z1 in an open
connected set A ⊆ C can be connected by a polygonal path lying in A
E.g. ∅, C, Dr (z), Dr (z), Cr (z) and R are connected, whereas Z, Q and R ∪ D1 (2i) are
disconnected.
(iii) ∀ε > 0∃N ∈ N∗ ∀m, n ≥ N : |zm − zn | < ε, i.e. (zn )n∈N∗ is a complex Cauchy-
sequence.
lim f (z) = w0
z→z0
z∈U
• f is called holomorphic at z0 ∈ Ω, if
Remark 1.4. Regular or complex differentiable are other words used for holomorphic.
Example 1.2. Let f ∈ CC , f (z) = z. Then f is entire, since
f (z0 + h) − f (z0 ) z0 + h − z0 h
lim = lim = lim = 1
h→0 h h→0 h h→0 h
H (Ω) := {f ∈ CΩ : f is holomorphic on Ω}
(i) The set H (Ω) is a C-vector space. More precisely, if f, g ∈ H (Ω), then
∀α, β ∈ C : αf + βg ∈ H (Ω)
and
∀α, β ∈ C : (αf + βg)′ = αf ′ + βg ′
(The zero-function 0 ∈ CΩ is the zero element of the vector space).
(f g)′ = f ′ g + f g ′
f
(iii) If g(z0 ) ̸= 0, then g
is holomorphic at z0 and
′
f f ′ (z0 )g(z0 ) − f (z0 )g ′ (z0 )
(z0 ) =
g g 2 (z0 )
Proof. The claims of this Propositions are very similar to the case of Real Analysis;
therefore here we will only show the fourth.
Since
f (z) − f (z0 )
lim F (z) = lim = f ′ (z0 ) = F (z0 )
z→z0 z→z0 z − z0
we have that F is continuous at z0
Similarly, G is continuous at w0 . Hence, since f is differentiable at z0 and hence
continuous at z0 , G ◦ f is continuous at z0
with E ∈ CΩ satisfying
E(z, z0 )
lim =0
z→z0 z − z0
Here we have that c = f ′ (z0 )
Example 1.3. 1. Example 1.2 and Proposition 1.8 applied repeatedly show that any
polynomial p ∈ C[X] is complex differentiable at every point z ∈ C
F : R2 → R2
(x, y) 7→ (x, −y)
Recall: A function
F : R2 → R2
(x, y) 7→ u(x, y), v(x, y)
or equivalently
Recall: We can view C as a 1-dimensional vector space over C with basis BC = {1} or
as a 2-dimensional real vector space with basis BR = {1, i}
A map T : C → C is C-linear if
T (z) = λz
for some λ ∈ C. On the other hand, a map T : C → C is R-linear if
with
1
λ = (T (1) − iT (i))
2
1
µ = (T (1) + iT (i))
2
z+z̄ z−z̄
using that x = 2
and y = 2i
Remark 1.6. Note that in Example 1.3, the function f (z) = z̄ as map R2 → R2 is
differentiable with Jacobian equal to
1 0
0 −1
but which is not of the above form.
Our next goal is to see how this Linear Algebra fact about R-linear versus C-linear maps
is reflected in the case of a linear function f : C → C and of its complex differentiability.
f˜ : R2 → R2
(x, y) 7→ u(x, y), v(x, y)
As specified in 1.1, we can identify the two spaces R2 and C as equal, hence f˜ ∼ = f =⇒
˜
f =f
This said, the derivative limz→z0 f (z)−f
z−z0
(z0 )
exists independently of how z → z0
We can conclude from this that the usual partial derivatives ux (z0 ), vx (z0 ) exist and
hence also the partial derivative fx (z0 ) = ux (z0 ) + ivx (z0 ) exists and that
∂f
(z0 ) = 0
∂ z̄
and
∂f ∂u
f ′ (z0 ) = (z0 ) = 2 (z0 )
∂z ∂z
If we write f (z) = f˜(x, y), then f˜ : R2 → R2 , (x, y) 7→ u(x, y), v(x, y) is complex
1.2. HOLOMORPHIC FUNCTIONS 19
differentiable with
ux uy ux uy ux −vx
Df˜(x0 , y0 ) = = =
vx vy −uy ux vx ux
and
˜
u u 2
det Df (x0 , y0 ) = x y = f ′ (z0 ) = u2x (z0 ) + u2y (z0 ) = u2x (z0 ) + vx2 (z0 )
vx vy
Remark 1.7. Recall that we can represent any complex number z = a + bi with a 2 × 2
real matrix:
a −b
a + ib ←→
b a
If f (z) = u(z) + iv(z) is complex differentiable at z0 , then f ′ (z0 ) = ux (z0 ) + ivx (z0 ) has
matrix representation
′ ux (z0 ) −vx (z0 )
f (z0 ) ←→
vx (z0 ) ux (z0 )
On the other hand, the corresponding function
ux = vy
vx = −uy
If one remembers the general form of the Jacobian of a function g : R2 → R2 and the
matrix representation of a complex number, then can remember the Cauchy-Riemann
equations.
h→0
with ε2 (h) −−→ 0. By then summing these two together we obtain
h→0
where ε(h) = (ε1 + ε2 )(h) −−→ 0. Using now the Cauchy-Riemann Equations we can
reform the previous expansion in what follows
h→0
Hence f (z0 + h) − f (z0 ) = (∂x u − i∂y u)h + |h|ε(h) where ε(h) −−→ 0
∂x u(z) = 2x ∂x v(z) = 2y
∂y u(z) = 2y ∂y v(z) = 2x
Remark 1.8. Many books distinguish between complex differentiability at a point and
holomorphicity at a point as follows:
• Holomorphicity at a point is instead given instead when the limit in 1.15 exists in
a neighbourhood of that point.
[SS10] does not make such distinction (and in this course we will not need it).
22 CHAPTER 1. PRELIMINARIES TO COMPLEX ANALYSIS
A quick summary
Let Ω be an open subset of C, then
1. For f ∈ CΩ with f (z) = u(z) + iv(z) we have that if f is holomorphic on Ω, then
u, v satisfy the Cauchy-Riemann Equations ux = vy and uy = −vx . Moreover it
holds that
f ′ (z) = ux (z) − iuy (z)
L : R2 → R2
x a −b x
7→
y b a y
1 1
Moreover, with the convention that 0
:= ∞ and ∞
:= 0, R is given by
1 1
= lim sup |an | n
R n→∞
Hence , ∞ n−1
P
n=1 nan z has the same radius of convergence. Repeated applications of
this same argument show that the sum
∞
X n!
an z n−k
n=k+1
(n − k − 1)!
Let now z ∈ DR (0) ⊆ C and choose δ > 0 such that |z| + δ < R, e.g. one can take
δ = R−|z|
2
, with h ∈ C such that |h| < δ, then
∞ ∞
an (z + h)n − an z n
f (z + h) − f (z) X n−1
X
n−1
− nan z = − nan z ≤
h n=1 n=1
h
∞ n
! !
X 1 X n
≤ |an | hk z n−k − z n − nz n−1 =
h k Bin
n=1 k=0
∞ n
X X n
= |an | hk−1 z n−k ≤
k Bin
n=2 k=2
∞ n
X X n
≤ |an | |h|k−1 |z|n−k ≤
k Bin
n=2 k=2
1 ∞ n
(∗ ) X X n−2
≤ |an |n(n − 1) |h|k−2 |z|n−k |h| ≤
k − 2 Bin
n=2 k=2
∞
X n−2
≤ |an |n(n − 1) |z| + |h| |h| ≤
n=2
∞ n−2
(∗2 )
X R + |z|
≤ |h| |an |n(n − 1)
n=2
2
| {z }
independent of h
1 n n n−1 n(n−1) n − 2
using in the (∗ )-step that ∀k ≥ 2 : = k = k(k−1) ≤
k Bin k − 1 Bin k − 2 Bin
n−2
n(n − 1) and using lastly in the (∗2 )-step that |h| < R−|z|
k − 2 Bin 2
We therefore obtain
∞ n−2
X R + |z| h→0
|h| |an |n(n − 1) −−→ 0
n=2
2
1.2. HOLOMORPHIC FUNCTIONS 25
2. Trigonometric functions
∞
eiz − e−iz X (−1)n z 2n+1
sin(z) := =
2i n=0
(2n + 1)!
∞
eiz + e−iz X (−1)n z 2n
cos(z) := =
2 n=0
(2n)!
with 2 2
ei − e−i e2 − 1
sin(i) = =i
2i 2e
and 2 2
ei + e−i e−1 + e e2 + 1
cos(i) = = =
2 2 2e
For the records, these functions are not bounded.
3. The series ∞
X zn
n=1
n2
P∞ 1
has convergence radius 1, i.e. it converges for all z ∈ D1 (0), since n=1 n2 <∞
4. The geometric series
∞
X
zn
n=0
5. The series ∞
X zn
(−1)n−1
n=1
n
converges for all z ∈ D1 (0). Moreover, for z = 1 the series converges (Leibniz’
criteria), while for z = −1 the series diverges (Harmonic series).
γ : [a, b] → C
t 7→ γ(t) = x(t) + iy(t)
Here, we consider
γ(a + h) − γ(a)
γ ′ (a) := lim
h↘0 h
and
γ(b + h) − γ(b)
γ ′ (b) := lim
h↗0 h
as the right and left derivatives respectively.
• A curve is simple if it’s not self intersecting, i.e. γ(t) ̸= γ(s) unless s = t or
s = a and t = b
1.3. COMPLEX LINE INTEGRALS (INTEGRALS ALONG CURVES) 27
Remark 1.10. For us in this course the curves will always be piecewise smooth. From
now on when we say “a curve” we mean “a piecewise smooth one”, even if we forget to
write it.
Remark 1.11. We will often work with a particular parametrisation, since most im-
portant notions will be independent of parametrisation (for example path integrals).
Because of this independence, we often describe curves by drawing them as geometric
objects in the plane.
There are two elementary methods to modify or combine paths in order to obtain new
paths.
γ − : [a, b] → C
t 7→ γ − (t) := γ(b + a − t)
• If γ1 : [a1 , b1 ] → C and γ2 : [a2 , b2 ] → C are two paths such that γ1 (b1 ) = γ2 (a2 ),
then the concatenation or sum of the paths γ1 , γ2 is a path
γ1 ⊎ γ2 : [a1 , b1 + b2 − a2 ] → C
γ1 (t) , t ∈ [a1 , b1 ]
t 7→ (γ1 ⊎ γ2 )(t) :=
γ2 (t − b1 + a2 ) , t ∈ [b1 , b1 + b2 − a2 ]
28 CHAPTER 1. PRELIMINARIES TO COMPLEX ANALYSIS
γ : [0, 1] → C
t 7→ (1 − t)z1 + tz2
γ : [0, 4] → C
t , t ∈ [0, 1]
1 + i(t − 1) , t ∈ [1, 2]
t 7→ γ(t) :=
(3 − t) + i , t ∈ [2, 3]
i(4 − t) , t ∈ [3, 4]
To define the complex line integrals we recall that a continuous function g : [a, b] → R
Rb
is Riemann integrable, i.e. a g(t)dt exists.
30 CHAPTER 1. PRELIMINARIES TO COMPLEX ANALYSIS
Definition 1.19. For a complex valued function g ∈ C[a,b] we can define the integral
Z b Z b Z b
g(t)dt := u(t)dt + i v(t)dt
a a a
Since g(t) = f γ(t) γ ′ (t) ∈ C[a,b] is continuous on [a, b], the integral on the right is
Let γ̃ : [c, d] → C be another parametrisation of im(γ), such that γ̃(s) = (γ ◦ σ)(s) for
some σ : [c, d] → [a, b] with σ ∈ C 1 [c, d], [a, b] and σ ′ (s) > 0. Then we have
Z Z d Z d
′
f γ σ(s) γ ′ σ(s) σ ′ (s)ds
f (z)dz = f γ̃(s) γ̃ (s)ds =
γ̃ c c
′
By letting t = σ(s) and consequently dt = σ (s)ds, we obtain
Z d Z b Z
′ ′ ′
f γ σ(s) γ σ(s) σ (s)ds = f γ(t) γ (t)dt = f (z)dz
c a γ
The following properties of path integrals follow from the properties of the Riemann
integral.
(iii) Exercise.
An immediate Corollary is
Corollary 1.1. [SS10, Corollary I.3.3] If γ is a closed curve (i.e. γ(a) = γ(b)) in an
open set Ω, f ∈ C 0 (Ω; C) and has a primitive in Ω, then
Z
f (z)dz = 0
γ
Proof of Theorem 1.4. Let F = U (x, y) + iV (x, y) and γ : [a, b] → C. We first assume
that γ is smooth. We define a function
G : [a, b] → C
t 7→ F γ(t) = F x(t), y(t)
We need to check the compatibility of the real derivative of G and the complex deriva-
tive of F .
We have that G ∈ C 0 [a, b]; C is a continuous function, hence
′
′
G (t) = U x(t), y(t) + iV x(t), y(t) =
= Ux x(t), y(t) x′ (t) + Uy x(t), y(t) y ′ (t) +
+ i Vx x(t), y(t) x′ (t) + Vy x(t), y(t) y ′ (t)
1.3. COMPLEX LINE INTEGRALS (INTEGRALS ALONG CURVES) 33
by using the Chain Rule from Vector Analysis [EW22]. Now, applying the Cauchy-
Riemann Equations for F we get
′ ′
= Ux x(t), y(t) x (t) − Vx x(t), y(t) y (t) +
+ i Vx x(t), y(t) x′ (t) + Ux x(t), y(t) y ′ (t) =
= Ux x(t), y(t) + iVx x(t), y(t) x′ (t)+
′
+ − Vx x(t), y(t) + iUx x(t), y(t) y (t) =
= Ux x(t), y(t) + iVx x(t), y(t) x(t) + iy(t) =
= F ′ γ(t) γ ′ (t) =
where in the step denoted by (∗) we used the Fundamental Theorem of Analysis on G
[EW22].
If γ is piecewise smooth, then there is a dissection of [a, b] of the form
[
[a, b] = [aℓ−1 , aℓ ]
ℓ∈{1,...,n}
with a =: a0 and b =: an in accordance with the curve; this means such that for
ℓ ∈ {1, ..., n} we can define γℓ := γ|[aℓ−1 ,aℓ ] and dissect the curve as follows
n
]
γ= γℓ
ℓ=1
Then
Z n Z
X n
X
f (z)dz = f (z)dz = F (aℓ ) − F (aℓ−1 ) = F (an ) − F (a0 ) = F (b) − F (a)
γ ℓ=1 γℓ ℓ=1
using the newly obtained result for the smooth case in the third step and acknowledging
the telescopic character of the sum in the second last one.
Another Corollary of Theorem 1.4 is the following:
34 CHAPTER 1. PRELIMINARIES TO COMPLEX ANALYSIS
Corollary 1.2. [SS10, Corollary I.3.4] Let f ∈ H (Ω) on an open and connected
subset Ω ⊆ C. If f ′ = 0, then f is constant.
Proof. We want to show that for any two points z, w ∈ Ω, it holds f (z) = f (w)
But given that f ′ = 0, the integral on the left is equal to zero and therefore f (z) = f (w)
The arbitrariness of the choice of z, w ∈ Ω concludes the proof.
f : C∗ → C
1
z 7→
z
f has no primitive on C∗ . To see this let γ parametrise the circle centred at 0 and of
radius 1, namely C1 (0), i.e. γ : [0, 2π] → C, t 7→ eit , then
Z Z 2π Z 2π Z 2π
it
it 1 it
f (z)dz = f e ie dt = ie dt = i dt = 2πi ̸= 0
γ 0 0 eit 0
R
Example 1.9. What is γ z 2 dz, if γ : [0, 1] → C, t 7→ t + πit2 ?
z3
Since F (z) = 3
is a primitive of z 2 , using Theorem 1.4 we have that
(1 + πi)3
Z
z 2 dz = F γ(1) − F γ(0) =
γ 3
or Z Z 1
2
z dz = (t + πit2 )2 (1 + 2πit)dt = ...
γ 0
Cauchy’s Theorem, as we will see, has many applications, e.g. Liouville’s Theorem,
which in return gives a proof of Fundamental Theorem of Algebra.
The interior of a path is not easy to define for a general curve. We will work around
this difficulty by first proving Cauchy’s Theorem for curves, whose interior is easy to
define, namely for triangles and rectangles (Goursat’s Theorem).
We then use Goursat’s Theorem to show that a holomorphic function on an open disc
has a primitive in that disc. This then will give us a Corollary: Cauchy’s Theorem on
a disc.
We first need the following Proposition about nested compact sets, but before that we
define a useful notational tool to approach this type of scenarios, namely the diameter
of a set.
diam(S) := sup |a − b|
a,b∈S
35
36 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
Then
∃!z0 ∈ C∀n ∈ N∗ : z0 ∈ ∆(n)
Proof. Choose zn any point in ∆(n) . For any n, m ∈ N∗ such that n ≥ m ≥ 1 we have
|zn − zm | ≤ diam ∆(m) < +∞
n→∞
Since diam ∆(n) −−−→ 0, this says that (zn )n∈N∗ ∈ n∈N∗ ∆(n) is a Cauchy sequence,
Q
Proof. Note that a triangle is a closed curve, which is the union of three line segments.
If T has three corners at z1 , z2 and z3 , then we will write
3
]
T = αl =: ⟨z1 , z2 , z3 ⟩
l=1
U
(Note that the sign does NOT represent a “flipped gravestone” ∼ Greta from Lo-
carno1 )
1
Locarno is a southern Swiss town and municipality in the district Locarno (of which it is the
2.1. GOURSAT’S THEOREM 37
with
α1 : [0, 1] → C, t 7→ α1 (t) = z1 + (t − 0)(z2 − z1 )
α2 : [0, 1] → C, t 7→ α2 (t) = z2 + (t − 1)(z3 − z2 )
α3 : [0, 1] → C, t 7→ α3 (t) = z3 + (t − 2)(z1 − z3 )
We now define the simplex
∆ := {z ∈ C : z = t1 z1 + t2 z2 + t3 z3 and 0 ≤ t1 , t2 , t3 and t1 + t2 + t3 = 1}
such that it is the smallest convex set containing z1 , z2 , z3 . We hence have that im(T ) ⊆
∆, in fact im(T ) = ∂∆
capital), located on the northern shore of Lake Maggiore at its northeastern tip in the canton of Ticino
at the southern foot of the Swiss Alps. It has a population of about 16,000 (proper), and about 56,000
for the agglomeration of the same name including Ascona besides other municipalities.
38 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
i.e. we are at each step partitioning ∆ using lines parallel to the sides and passing
through their midpoints.
(n)
The triangular paths Tk with k ∈ {1, ..., 4} are all entirely contained in ∆ and we
have that
Z 4 Z
X
f (z)dz = f (z)dz
(n+1)
T (n) k=1 Tk
therefore we get
Z Z
f (z)dz ≤ 4 max f (z)dz
k∈{1,...,4} (n+1)
T (n) Tk
(n+1)
We choose T (n+1) as one of Tk with k ∈ {1, ..., 4} so that
Z Z
f (z)dz ≤ 4 f (z)dz
T (n) T (n+1)
We also have that the closed filled triangles ∆(n) , defined with respect to their T (n) in
a similar fashion as ∆ with T , are nested compact sets
Moreover, if dn and Pn are the diameter and the perimeter of ∆(n) respectively, then
for all n ∈ N we have
d0 d P0 P
dn = n
=: n and Pn = n
=: n
2 2 2 2
Hence
n→∞
diam ∆(n) −−−→ 0
We can now apply the result anticipated by Proposition 2.1. In order to do so, though,
we remember that f is holomorphic at z0 ∈ ∆(n) ⊆ Ω, hence it holds that
|E(z)|
with limz→z0 |z−z0 |
= 0. It is clear that E(z) is continuous at z0 and that limz→z0 E(z) =
0
R R
Let ε > 0, we will show that I := T
f (z)dz ≤ εP d, which will show that T
f (z)dz = 0
Using Corollary 1.1, the first integral on the right is zero, since g(z) = f (z0 )+f ′ (z0 )(z −
2
z0 ) has a primitive (take G(z) = f (z0 )z + f ′ (z0 ) (z−z2 0 ) as such) and since the curve is
closed.
By letting Z
In := E(z)dz
T (n)
Proof. This follows immediately from the Goursat’s Theorem 2.1 and by dividing the
rectangle R into two triangles T1 , T2 as shown in the picture.
Therefore, we obtain
Z Z Z
f (z)dz = f (z)dz + f (z)dz = 0
R T1 T2
For future results, for example for the deduction of Cauchy’s Integral Formula, a minor
extension of this result is useful.
in which L is the length of the perimeter of R, PRlj = Ln is the length of the perimeter
of Rlj and M = maxz∈R |f (z)| = ∥f ∥∞,R is the maximum of the continuous function
|f | on the compact set R
By construction, the point z0 cannot belong to more than four subrectangles: the point
in which touches most subrectangles is a common vertices of all the four. Hence we can
finally consider the following estimate
Z X Z X Z L n→∞
f (z)dz = f (z)dz ≤ f (z)dz ≤ 4M −−−→ 0
R z0 ∈Rlj ∂Rlj z0 ∈Rlj Rlj n
We are going to prove the following version which assumes that f is continuous in D
and that its integral along rectangles whose sides parallel to the coordinate axes vanish.
Which then we are going to use to give a slightly stronger form of Cauchy’s Theorem.
Theorem 2.4. Let D be an open disc in C and f ∈ C 0 (D) with the property that
Z
f (z)dz = 0
R
for every closed rectangle R ⊆ D with ∂R = R in D and whose sides are parallel
to the coordinate axes. Then f has a primitive in D
Theorem 2.5 (Cauchy’s Theorem for a disc). [SS10, Theorem II.2.2] Suppose D ⊆
C is an open disc in C and f ∈ H (D), or more generally f ∈ C 0 (D) with f |D\{z0 } ∈
H (D \ {z0 }) for some z0 ∈ D. Then
Z
f (z)dz = 0
γ
for every closed rectangle R ⊆ D with ∂R = R (including the ones, whose sides are
parallel to axes). By Theorem
R 2.4 f has a primitive in D and by Theorem 1.4 and
Corollary 1.1 we have that γ f (z)dz = 0 for every piecewise smooth path γ in D
Proof of Theorem 2.4. Let f ∈ C 0 (D) be continuous on the disc D and let z0 = x0 +iy0
be the center of D
For an arbitrary point z = x+iy ∈ D such that z ̸= z0 , let z1 := x+iy0 and z2 := x0 +iy
2.2. LOCAL EXISTENCE OF PRIMITIVES AND CAUCHY’S THEOREM IN A DISC43
By assumption
Z Z Z Z
f (w)dw + f (w)dw + f (w)dw + f (w)dw = 0
[z0 ,z1 ] [z1 ,z] [z,z2 ] [z2 ,z0 ]
where we denote the path along the line segment linking two points p1 , p2 ∈ D, from
p1 to p2 , with [p1 , p2 ]
R R
This sum represent either R
f (w)dw or − R
f (w)dw, depending on the location of z
and Z x Z y
F (z) = f (t + iy0 )dt + i f (x + it)dt
x0 y0
Using both the above results and the Fundamental Theorem of Analysis [EW22]:
Z x
d
g(t)dt = g(x)
dx a
if g ∈ C 0 Dr (a)∩R; C with g(t) = f (t+iy), then we have that Fx (z) = f (x+iy) = f (z).
Similarly, since Z y
d
h(t)dt = h(y)
dy a
44 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
if h ∈ C 0 Dr (a) ∩ R; C with h(t) = f (x + it), we have Fy (z) = if (x + iy) = if (z)
(for both parts we used that the first integral in the equations are independent of x, y
respectively). Hence, it follows that Fx , Fy exist and are continuous, so F ∈ C 1 (D)
At this point, since Fx (z) = f (z) and Fy (z) = if (z), if we write F (z) = u + iv, then
this gives
f (z) = Fx (z) = ux + ivx = −iFy (z) = −i(uy + ivy ) = vy − iuy
Hence ux = vy and vx = −uy and so the the Cauchy-Riemann equations hold.
Finally, we know that F ∈ C 1 (D) and that F satisfies that Cauchy Riemann Equations,
therefore by Theorem 1.1 F ∈ H (Ω) and
∂F
F ′ (z) = (z) = f (z)
∂x
hence F is a primitive of f
Corollary 2.2. [SS10, Corollary II.2.3] Let f ∈ H (Ω) for Ω ⊆ C an open set
containing a circle C and its interior, then
Z
f (z)dz = 0
C
R
Sketch of a proof. Simply consider F (z) := γz
f (w)dw as in the following illustration.
Remark 2.1. Corollary 2.2 is in fact valid whenever we can define the interior of a
contour unambiguously and construct polygonal paths in an open neighbourhood of both
the contour and its interior. In [SS10] these contours are called toy contours.
Remark 2.2. Note that Cauchy’s Theorem 2.5 does not say anything about integrals
of functions over arbitrary open sets and arbitrary closed curves. Indeed recall Example
1.8.
2.2. LOCAL EXISTENCE OF PRIMITIVES AND CAUCHY’S THEOREM IN A DISC45
Example 2.1. We can show by parametrizing the circle that for all r > 0 we have:
Z
1
dz = 2πi
Cr (z0 ) z − z0
Indeed, the circle of center z0 and radius r has parametrization σ(t) = z0 + reit for
t ∈ [0, 2π] and hence
Z Z 2π
1 1
dz = it
ireit dt = 2πi
|w−z0 |=r z − z0 0 re
For each k ∈ {1, ..., 4} we choose an open disc Dk , so that the trajectory of the closed
path γk ⊎ βk− is in Dk and so that f (z) = z−z
1
0
is holomorphic in Dk
46 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
where
γ1 : [0, R] → C, t 7→ γ1 (t) = t
h πi
γ2 : 0, → C, t 7→ γ2 (t) = Reit
4
π
γ3 : [0, R] → C, t 7→ γ3 (t) = tei 4
2.2. LOCAL EXISTENCE OF PRIMITIVES AND CAUCHY’S THEOREM IN A DISC47
Claim. 2
π 1 − e−R
Z
iz 2 R→+∞
e dz ≤ −−−−→ 0
γ2 4R
Proof of the claim. Expanding the integral we have
Z Z Z π Z π
4 4
iz 2 iz 2 i(Reit )2 it 2 e2it )
e dz ≤ e |dz| = e iRe dt = R ei(R eit dt =
γ2 γ2 0 0
Z π
4 2 (cos(2t)+i sin(2t))
=R ei(R eit dt =
0
Z π
4 2 2
=R eiR cos(2t) e−R sin(2t)
eit dt =
0 | {z } |{z}
=1 =1
Z π Z π
4 R R→∞ 4
= dt −−−→ 0dt = 0
0 eR2 sin(2t) 0
Using that the last integrand in bounded for R ∈ [0, +∞) and therefore applying the
Dominant Convergence Theorem [Da 24] to swap the limit and the integral.
Therefore we have
Z R
√ Z ∞ √
it2 2 −t2 2π
lim e dt = (1 + i) e dt = (1 + i)
R→∞ 0 2 0 2 2
R∞ √
2 π
since from [EW22] we have that 0
e−t dt = 2
Remark 2.3. Note that the use of compact paths to solve this kind of integrals in a
very useful tool. One first solves the compact case and then by letting some parameter
approach its limit case of interest, if the integrand is nice enough, the complexity of the
problem can be considerably reduced.
48 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
Remark 2.4. Note that Cauchy’s Integral Formula says that the values of f on D are
determined by their boundary values on the circle C
Proof. Let z0 ∈ Ω, since Ω is open, there exist r > 0 such that Dr (z0 ) ⊂ Ω, still since
Ω is open, ∀z ∈ ∂Dr (z0 ) : z ∈ Ω and so ∃ε > 0 : Dr+ε (z0 ) ⊂ Ω. We obtain from this
that Cr (z0 ) ⊂ Dr+ε (z0 ). We let z ∈ Dr (z0 ) and define
f (w)−f (z)
, w ̸= z
g : Dr+ε (z0 ) → C, w 7→ g(w) = w−z
′
f (z) ,w=z
Then it holds that g ∈ C 0 Dr+ε (z0 ) and away from z we also have that g is holomor-
phic, i.e. g|Dr+ε (z0 )\{z} ∈ H Dr+ε (z0 ) \ {z}
i.e.
f (w) − f (z)
Z
dw = 0
Cr (z0 ) w−z
Note that on Cr (z0 ) it holds that w ̸= z, since z ∈ Dr (z0 ). Hence
Z Z
f (w) 1
dw = f (z) dw
Cr (z0 ) w − z Cr (z0 ) w − z
γ : [0, 2π] → C
t 7→ γ(t) = z0 + reit
γ̃ : [0, 2π] → C
s 7→ γ̃(s) = z + ρ(s)eit
where ρ : [0, 2π] → R, s 7→ ρ(s) := γ t(s) − z = |γ̃(s) − z| with σ : [0, 2π] →
[0, 2π], s 7→ t(s) as reparametrization of the control variable. Clearly ρ is smooth,
moreover
γ̃ ′ (s) = ρ′ (s)eis + iρ(s)eis
We proceed now to develop the integral form above:
Z 2π ′
ρ (s)eis + iρ(s)eis
Z
1
dw = ds =
Cr (z0 ) w − z 0 ρ(s)eis
Z 2π ′ Z 2π
ρ (s)
= ds +i ds =
0 ρ(s) 0
| {z }
real integral
s=2π
= ln ρ(s) s=0 +2πi =
| {z }
=0
= 2πi
Here t changes with s and ρ = ρ(s) = γ t(s) − z . As mentioned before
γ̃(s) = z + ρ(s)eit
50 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
grable on every [a, b] and which has converging −∞ f (t) dt, its Fourier Transform
Z ∞
fˆ(ξ) := f (x)e−2πiξx dx
−∞
2
Example 2.3. We’ll show that e−πx is its own Fourier Transform. We want to show
that if f (x) = e−πx , then fˆ(ξ) = e−πξ
2 2
Where γ1 is the path on the real axis, γ2 is the right vertical path, γ3 is the upper
horizontal one and γ4 is the left vertical one. Hence, using Cauchy’s Theorem 2.5
Z
f (z)dz = 0
γR
Note that on γ1 :
Z Z R
2
f (z)dz = e−πx dx
γ1 −R
while on γ3 :
Z Z −R Z R Z R
−π(x+iξ)2 −π(x2 +2πixξ) πξ 2 πξ 2 2
f (z)dz = e dx = − e e dx = −e e−πx e−2πixξ dx
γ3 R −R −R
As R → ∞, the first integral over γ1 equals 1, while the integral over γ3 gives
Z R
πξ 2 2
−e e−πx e−2πixξ dx
−R
For a fixed ξ, the integral can be bounded using Proposition 1.10 with
Z
2 2 2
f (z)dz ≤ ξ sup e−πR e−πiRy eπy ≤ Ce−πR
γ2 y∈[0,ξ]
Next we are going to see that Cauchy’s theorem 2.5 and the Cauchy’s Integral Formula
2.6 will imply fundamental properties of holomorphic functions.
3. The Fundamental Theorem of Algebra holds, i.e. any polynomial p(z) ∈ C[z] of
degree n ∈ N, has n roots in C (counted with multiplicity).
4. If f, g ∈ H (Ω) and f (z) = g(z) for all z in some sequence of distinct points with
a limit point in Ω, then ∀z ∈ Ω : f (z) = g(z). In particular, if f, g agree on an
open set U of Ω, then they agree on all of Ω
Theorem 2.7. [SS10, Theorem II.4.4] Suppose that Ω ⊆ C is an open set and
f ∈ H (Ω). Let z0 ∈ Ω and r > 0 be such that Dr (z0 ) ⊂ Ω. Then f has a power
series extension at z0
∞
X
∀z ∈ Dr (z0 ) : f (z) = an (z − z0 )n
n=0
f [n] (z0 )
Z
1 f (w)
∀n ∈ N : an = = dw
n! 2πi Cr (z0 ) (w − z0 )n+1
Proof. Let z0 and r be defined as above, so that Dr (z0 ) ⊂ Ω. Fix s ∈ (0, r) and let
Cs (z0 ) be the circle of radius s with center z0 . By setting γ a path in Dr (z0 ) such that
im(γ) = Cs (z0 ) ⊂ Ω and using Theorem 2.6, we obtain
Z
1 f (w)
f (z) = dw
2πi γ w − z
Since we are integrating on γ, for w ∈ im(γ) (and for z ∈ Ds (z0 )) we have for the term
in the last fraction that
z − z0 |z − z0 |
= <1
w − z0 s
This means that we can rewrite the whole last fraction as a geometric series (see [EW22])
∞ n
1 X z − z0
z−z0 =
1 − w−z 0 n=0
w − z0
for w ∈ im(γ) and z ∈ Ds (z0 ). The convergence of the series is uniform, since the
bound |z−z
s
0| z−z0
for w−z 0
is independent of w ∈ im(γ). Hence we can interchange the series
and the integral (again, see [EW22]), obtaining
∞
!
(z − z0 )n
Z
1 X
f (z) = f (w) n+1
dw =
2πi γ n=0
(w − z0 )
∞ Z
n 1 f (w)
X
= (z − z0 ) dw =
n=0
2πi γ (w − z0 )n+1
∞
X
= an (z − z0 )n
n=0
Hence, we have that in Ds (z0 ) the function f is the sum of the power series
∞
X
an (z − z0 )n
n=0
54 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
We have seen that series in their disc of convergence are differentiable with derivatives
given by termwise differentiation, as stated in Theorem 1.3. Hence, for all z ∈ Ds (z0 )
we have
X∞
′
f (z) = nan (z − z0 )n−1 =
n=1
∞
X
= (k + 1)ak+1 (z − z0 )k
k=0
Being a power series, f ′ (z) is also holomorphic in Ds (z0 ), i.e. f ′ |Ds (z0 ) ∈ H Ds (z0 ) .
Inductively, we get that f is differentiable infinitely many times for z ∈ Ds (z0 ) and
evaluating it and its derivatives at z = z0 gives
a0 = f (z0 )
a1 = f ′ (z0 )
..
.
n!an = f [n] (z0 )
[n]
Hence an = f n!(z0 ) is independent of s and we have ∞ n
P
n=0 an (z − z0 ) converging for all
z ∈ Ds (z0 ), this for any arbitrary s ∈ (0, r). By taking the limit for s → r we notice
that γ expands to Cr (z0 ), which is still in Ω: the above calculations and step are still
valid, especially the uniform convergence of the series, also in this case is granted, as
|z − z0 | < r. Viewing this as (constant for each z ∈ Ds (z0 )) continuous function of s,
we can extend by continuity to r. This concludes the proof.
Remark 2.5. In all these Theorems with Ω ⊆ C open by assumption, it always follows
by definition of open set in C that
∃r > 0 : Dr (z0 ) ⊂ Ω
Moreover, the above proof of Theorem 2.7 gives a method for determine the radius of
convergence of the series expansion of any f ∈ H (Ω), as done above by expanding s
to r: we start by a fully contained Ds (z0 ) ⊂ Ω for some s ∈ (0, +∞) and progressively
increase it until we meet ∂Ω, thus external boundary of the set or some singularity
points. As long as Ds (z0 ) ⊂ Ω, the success of the operation is granted.
Remark 2.6. Note that the proof also gives that if f is complex differentiable at z0 ,
then in facts it is complex differentiable infinitely many times at z0 . It follows that for
every n ∈ N
f [n] (z0 )
Z
1 f (w)
an = dw =
2πi Cs (z0 ) (w − z0 )n+1 n!
In fact we have
2.3. CAUCHY’S INTEGRAL FORMULAS 55
Corollary 2.3 (Cauchy’s Integral Formula for derivatives). [SS10, Corollary II.4.2]
If f ∈ H (Ω), then f is infinitely often complex differentiable in Ω and in particular
∀n ∈ N : f [n] ∈ H (Ω)
Proof. The fact that f ′ is complex differentiable (or holomorphic) follows from the fact
that, since Ω is open, ∀z0 ∈ Ω∃r > 0 : Dr (z0 ) ⊂ Ω. By Theorem 2.7, f has a power
series expansion there, which is therefore holomorphic in Dr (z0 ). Since power series are
infinitely often complex differentiable in their disc of convergence, we have that f is
complex differentiable infinitely often at z0 . Since z0 was arbitrary, f in infinitely often
complex differentiable in Ω
To prove this, by induction on n ∈ N, note that the base case with n = 0 is simply the
Cauchy’s Integral Formula in Theorem 2.6. We are now going to show the induction
step, therefore suppose that
(n − 1)!
Z
[n−1] f (w)
f (z) = n
dw
2πi Cr (z0 ) (w − z0 )
for any z ∈ Dr (z0 ). For h ∈ C small enough, so that z + h and z are both away from
Cr (z0 ), we have
!
[n−1] [n−1]
(z + h) − f (n − 1)!
Z
f (z) f (w) 1 1
= n − dw
h 2πi Cr (z0 ) h w − (z + h) (w − z)n
We then use the equality (in C)
n−1
!
X
an − bn = (a − b) an−1−k bk
k=0
1 1 h→0
with a = w−(z+h) and b = w−z and take the limit as h → 0: note that a−b h
1
−−→ (w−z)2
Pn−1 n−1−k k h→0 n
and k=0 a b −−→ (w−z)n−1 , to get that
Remark 2.7. Theorem 2.7 says that a holomorphic function f can be locally developed
as a power series around each point of the P definition domain Ω. Explicitly, for each
z0 ∈ Ω, ∃Dr (z0 ) ⊂ Ω and a power series ∞ n
n=0 an (z − z0 ) , which converges for all
z ∈ Dr (z0 ) and represents the function f in Dr (z0 )
Due to this power series expansion we have that holomorphic functions are exactly the
functions which are everywhere representable as a power series (with a positive radius
of convergence). Recall that any power series represents a holomorphic function in their
disc of convergence. This is why we have the words ”holomorphic” and ”analytic” used
interchangeably in various sources.
Corollary 2.4 (Cauchy Inequality). [SS10, Corollary II.4.3] With the assumptions
as in Theorem 2.3, we have for every n ∈ N that
Corollary
Pn 2.5 (Fundamental Theorem of Algebra). Every polynomial p(z) = a0 +
k
k=1 ak z with deg(p) = n ≥ 1 has precisely n roots in C, counted with multiplicity.
If these roots are w1 , ..., wn (with possible repetitions), then
n
Y
p(z) = an (z − wk )
k=1
By contradiction, suppose that there is no such root and that consequently the function
1
Q(z) = p(z) ∈ H (C)
|z|→∞ |z|→∞
Hence p(z) −−−−→ ∞ and consequently Q(z) −−−−→ 0, from this result we deduce
that ∃r > 0 : Q(z) ≤ 1 whenever |z| ≥ r (Q is a continuous function, i.e. Q ∈ C 0 (C)),
but Q is continuous and hence bounded on the compact set Dr (0), say Q < m for
some m ∈ R
using the Binomial Theorem [EW22], new coefficients bn−1 , ..., b0 and bn = an . Since
p(w1 ) = 0, we must then have b0 = 0
Hence !
n
X
p(z) = (z − w1 ) bk (z − w1 )k−1 = (z − w1 )p̃(z)
k=1
Remark 2.9. 1. Holomorphic functions can have infinitely many zeroes, but we we
are going to see that these zeroes are isolated, i.e. for each zero z0 of a holomorphic
function f there exists a neighbourhood of z0 with no other zero.
E.g. cos(z) and sin(z) have respectively zeroes for z = (2k + 1) π2 and z = πk, with
k∈Z
Example 2.4. If Ω = [−1, 1] ∪ {2i}, then the zn ̸= z0 condition avoids the case 2i is a
limit point of Ω, since otherwise we could take ∀n ∈ N∗ : zn = 2i
We next define the order of zero of f at z0
Proof. (i) Let f be holomorphic in Ω with Ω open. By Theorem 2.7 and using that
Ω is open, ∃r > 0 is such that Dr (z0 ) ⊂ Ω and on it we have
∞
X f [n] (z0 )
∀z ∈ Dr (z0 ) : f (z) = (z − z0 )n
n=0
n!
Hence, if we define
∞
X f [m+k] (z0 )
∀z ∈ Dr (z0 ) : h(z) := (z − z0 )m
m=0
(m + k)!
Note that, since h ∈ H Dr (z0 ) , it is also continuous there and since h(z0 ) ̸= 0,
it holds that
∃ε ∈ (0, r)∀z ∈ Dε (z0 ) : h(z) ̸= 0
Moreover, h and n are unique, since if we assumed not, hence
Hence, if f [k] (z0 ) = 0 = g [k] (z0 ), then also (g + f )[k] (z0 ) = 0. This implies that
ordz0 (f + g) ≥ min{ordz0 (f ), ordz0 (g)}
By part (ii), instead, we write
with (h1 h2 )(z0 ) ̸= 0 From this, using the power series expansion of f g or the
uniqueness of n and h in part (ii) we get
As Corollary we get that the zeroes of an holomorphic function are isolated. More
precisely we have
Proof. Using Proposition 2.2 we have that ∃r > 0 : Dr (z0 ) ⊂ Ω in which we write
f |Dr (z0 ) (z) = (z − z0 )n h(z) with n = ordz0 (f ) and h(z0 ) ̸= 0. Let z ∈ Ḋr (z0 ), then
f (z) = 0 ⇐⇒ h(z) = 0
since (z − z0 )n ̸= 0 for z ̸= z0
Being h(z0 ) ̸= 0 and h is continuous on Dr (z0 ), we have that
2
It is useful to this purpose to recall the definition of a region in C, see 1.11.
2.3. CAUCHY’S INTEGRAL FORMULAS 63
/Z
limit point limn→∞ zn =: z0 ∈ Ω such that z0 ∈
Remark 2.10. 1. The reason this result is called Principle of Analytic Continua-
tion 2.10 is the following:
3. The condition that the limit point of zeroes is in Ω is also crucial, as shown in
the following Example 2.5.
Example 2.5. Take Ω = C∗ and
f :Ω→C
iπ −iπ
π ez −e z
z 7→ sin =
z 2i
π −e−π
It holds that f ∈ H (C∗ ) and f ̸= 0, since already f (i) = e 2i
̸= 0
Consider a sequence of zeroes given by { n1 }n∈N∗ with limn→∞ 1
n
= 0, as f ( n1 ) = sin(πn) =
64 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
0 for all n ∈ N∗ . The limit point of zeroes is not in Ω, causing the result. This example
shows that the zeroes can converge to a boundary point. Note that we do have that the
zeroes { n1 }n∈N∗ are isolated.
Another interesting example regarding the above concept of continuation is given here
at Example 2.6.
We are now going to prove the following theorem, which proves Theorem 2.10.
(i) f = 0
(i) f = g
Therefore f (z) ̸= 0 on Ḋε (a). Hence Zf ∩ Ḋε (a) = ∅, but this says that a is not
a limit point of Zf . Hence ∀n ∈ N : f [n] (a) = 0
(ii)=⇒(i): Let A := z ∈ Ω : ∀n ∈ N : f [n] (a) = 0 . By assumption a ∈ A,
hence A ̸= ∅. We will show that A = Ω via the characterisation of connectedness,
hence that f = 0
Recall: for an open subset Ω ⊆ C, connected means that the only both open and
closed subsets of Ω are ∅ and Ω (It is not possible to find two disjoint non-empty
open sets Ω1 , Ω2 such that Ω = Ω1 ∪ Ω2 ).
Since A ̸= ∅, if we can show that A is both open and closed, then A = Ω. Hence:
A is open: To see this, let c ∈ A and let r > 0 such that Dr (c) ⊂ Ω. Then
by Theorem 2.7
∞
X
∀z ∈ Dr (c) : f (z) = an (z − c)n
n=0
with
f [n] (c)
an = =0
n!
since c ∈ A, hence f |Dr (c) = 0. This means that Dr (c) ⊆ A. Hence, for an
arbitrary c ∈ A, we found a neighbourhood Dr (c) ⊆ A, which shows that A
is open.
66 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
Remark 2.11. 1. (a) The Identity Theorem 2.7 makes it clear that the real func-
tions
sin, cos, exp : R → R
can be uniquely extended to complex numbers, via their form on the real line.
E.g. sin(z) and cos(z) are entire functions. For any z = x ∈ R we have that
sin2 (x) + cos2 (x) = 1. Thus we define f ∈ CC , z 7→ f (z) = sin2 (z) + cos2 (z)
and g ∈ CC , z 7→ 1. Since f and g agree on the real line, they have to agree
on all of C, from which
(b) The functional equations can also be transferred from reals to complex num-
bers.
E.g. from
∀x, y ∈ R : exp(x + y) = exp(x) exp(y)
we first conclude
P∞ 1
We have g : Ω1 → C, z 7→ n=0 z n and f : Ω2 ⊇ Ω1 → C, z 7→ 1−z
. Therefore, if
z ∈ D1 (0), we then have
1
f (z) = g(z) =
1−z
Example 2.6. Let Ω = D1 (0), then it holds that
∞
X
∀z ∈ Ω : f (z) = zn
n=0
1
Let Ω̃ = C \ {1}, then Ω ⊆ Ω̃ and F (z) = 1−z is defined on all of Ω̃ and it agrees with
P∞ n 1
n=0 z whenever z ∈ D1 (0), so F (z) = 1−z is the analytic continuation of f to C\{1}
P∞
Warning: This does not say that F represents n=0 z n on the complement C \ D1 (0).
Note that in the Identity Theorem 2.7 we have two holomorphic functions defined on
the same set Ω. Here we have instead:
f : D1 (0) → C F : C \ {1} → C
∞
X 1
z 7→ zn z 7→
n=0
1−z
68 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
n! n
converges by comparison to the geometric series, since |z|P∞< |z| . When we look at
n!
series of holomorphic functions, for any ε > 0 we see that n=0 z converges absolutely
and uniformly on compact subsets of D1−ε (0), but f cannot be extended anywhere beyond
D1 (0)
f g = 0 =⇒ f = 0 or g = 0
The assumption
∀z ∈ Ω : f (z)g(z) = 0
then implies that ∀z ∈ Dε (a) : g(z) = 0. But then
Using the Identity Theorem 2.7 applied to g : Ω → C and the zero function 0 : Ω →
C, w 7→ 0 gives
g|Ω = 0|Ω = 0
– ∀a, b, c ∈ R : a(b + c) = ab + ac
– ∀a, b, c ∈ R : (b + c)a = ba + ca
The last Theorem 2.12 says that if Ω is open and connected, then the ring of analytic
functions on it has no zero divisor, hence is an integral domain3 .
Our next application is the Morera’s Theorem, which is a converse to Goursat’s Theo-
rem 2.1.
Recall: Goursat’s Theorem 2.1 says the following: let f : Ω → C for Ω open be a
holomorphic function and let T in Ω be a triangle, whose interior is also contained in
Ω, then Z
f (z)dz = 0
T
Theorem 2.13 (Morera’s Theorem). [SS10, Theorem II.5.1] Let Ω ⊆ C open and
f ∈ C 0 (Ω; C). Assume that for any open disc D with D ⊂ Ω and any triangular
path T such that im(T ) ⊆ D we have that
Z
f (z)dz = 0
T
Proof. Let z0 ∈ Ω and r > 0 such that Dr (z0 ) ⊂ Ω. For z ∈ Dr (z0 ) we define
Z
F (z) := f (w)dw
γz :=ℓ[z0 ,z]
3
In algebra, an integral domain is a non-zero commutative ring in which the product of any two
non-zero elements is non-zero. Integral domains are generalizations of the ring of integers and provide
a natural setting for studying divisibility. In an integral domain, every non-zero element a has the
cancellation property, that is, if a ̸= 0, an equality ab = ac implies b = c
70 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
for any T in Dr (z0 ), in particular for im(T ) = ⟨z0 , z, z +h⟩, we state the following claim.
It follows that
F (z + h) − F (z)
− f (z) ≤ sup |f (w) − f (z)|
h w∈ℓ[z,z+h]
2.4. SEQUENCES OF HOLOMORPHIC FUNCTIONS 71
N∗
Definition 2.5. A sequence (fn )n∈N∗ ∈ CΩ of functions defined on an open set
Ω ⊆ C is called uniformly convergent in Ω to the limit limn→∞ fn = f ∈ CΩ , if
It turns out that we only need local uniform convergence or equivalently uniform con-
vergence on compact subsets.
N∗
Definition 2.6. Let Ω ⊆ C be open and (fn )n∈N∗ ∈ CΩ a sequence of functions.
Ω
N∗
(fn )n∈N∗ ∈ C is called locally uniformly convergent or compactly con-
vergent or uniformly convergent on compact sets, if the following equivalent
conditions are satisfied:
(i) ∀a ∈ Ω∃ε > 0 : Dε (a) ⊆ Ω and fn |Dε (a) converges uniformly in Dε (a)
n∈N∗
Remark 2.15. Note that since continuity is a local property even in the case of real
valued functions, local uniform convergence of continuous functions will imply continuity
of the limit function.
With this result in mind we continue to the main Theorem that we have in this section.
N∗
Theorem 2.14. [SS10, Theorem II.5.2] Let Ω ⊆ C be open and (fn )n∈N∗ ∈ H (Ω)
a sequence of holomorphic functions. If (fn )n∈N∗ converge uniformly to a function
f in every compact subset of Ω, then f ∈ H (Ω)
Proof. Since each of the fn ’s is holomorphic, they are all also continuous, hence by the
above Proposition 2.3 their limit f is also continuous.
2.4. SEQUENCES OF HOLOMORPHIC FUNCTIONS 73
To show that f is also holomorphic we will use Morera’s Theorem 2.13 and the fact
that the set described by any triangle T is compact.
Let D = Dr (z0 ) ⊆ Ω an open disc in Ω for a z0 and a r > 0. Let T be any triangle with
inside contained in D. By Goursat’s Theorem 2.1, we have
Z
∀n ≥ 1 : fn (w)dw = 0
T
since fn (z) → f (z) uniformly on compact sets. Being the set described by T compact,
we have that fn (z) → f (z) uniformly on it, so
Z Z Z
fn (z)dz − f (z)dz ≤ |fn (z) − f (z)| |dz| ≤
T T T
n→∞
≤ sup |fn (z) − f (z)| LT −−−→ 0 · LT = 0
z∈T
for LT ∈ R the length of the perimeter of T and since (fn )n∈N∗ converges uniformly on
the set delimited by T . We hence have that
Z Z
lim fn (z)dz = f (z)dz
n→∞ T T
| {z }
=0
and so finally f ∈ H (Ω), using Morera’s Theorem 2.13 to conclude the proof.
We can extend the previous result to the following generalisation.
N∗
Theorem 2.15. [SS10, Theorem II.5.3] Let Ω ⊆ C be open and (fn )n∈N∗ ∈ H (Ω)
a sequence of holomorphic functions such that fn ⇒ f , i.e. fn converges uniformly
to f , on every compact subset K ⊆ Ω.
N∗
Then (fn′ )n∈N∗ ∈ H (Ω) converges uniformly to limn→∞ fn′ = f ′ on every com-
pact subset of Ω
74 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
Proof. Let z0 ∈ Ω and r > 0 such that Dr (z0 ) ⊂ Ω, then (fn )n∈N∗ converges uniformly
to f also on Dr (z0 ) as subset of Ω
To bound the denominator for both w ∈ Cσ (z0 ) and z ∈ Dr (z0 ), it holds that
|w − z| = |w − z0 + z0 − z| = w − z0 − (z − z0 ) ≥ |w − z0 | − |z − z0 | = |σ − r|
and since fn ⇒ f on the compact set Cσ (z0 ), we have that fn′ ⇒ f ′ on Dr (z0 )
Since by definition every compact set is contained in a union of finitely many such discs,
the result follows.
These Theorems are often used to prove holomorphicity of functions defined by infinite
series.
N∗
Corollary 2.8. Let Ω ⊆ C be open and (fn )n∈N∗ ∈ H (Ω) , if
∞
X
∀z ∈ Ω : F (z) = fn (z)
n=1
Proof. This Corollary is a direct consequence of the previous Theorem 2.15 applied on
the sequence (SN )N ∈N∗ defined as follows: for each N ∈ N∗
N
X
SN : Ω → C, z 7→ SN (z) = fn (z)
n=1
For series of functions, we have also the following useful Theorem of Weierstrass, called
Weierstrass M-test.
N∗
Theorem 2.16 (Weierstrass M -test). Let (fn )n∈N∗ ∈ CΩ be a sequence of
functions and ∅ =
̸ U ⊆ Ω. Suppose that
N∗ X
∃(Mn )n∈N∗ ∈ R≥0 : ∀n ∈ N∗ ∀z ∈ U : fn (z) ≤ Mn and Mn < ∞
n∈N∗
P
Then n∈N∗ fn converges absolutely and uniformly on U
Then we have
Proposition 2.4. [SS10, Proposition VI.2.1] The series that represents the Rie-
mann Zeta function ∞
X 1
ζ(z) :=
n=1
nz
converges absolutely and uniformly on every half plane
since
∞ ∞ ∞ ∞
X 1 X 1 X 1 X 1
|ζ(z)| = z
≤ z
≤ Re(z)
≤ 1+δ
<∞
n=1
n n=1
|n | n=1
n n=1
n
Applying the previous Weierstrass M -test 2.16 we obtain the result.
Hence ∞ 1
P
n=1 nz converges uniformly on every half plane as ∀δ > 0 : Re(z) ≥ 1 + δ > 1
and therefore defines a holomorphic function if Re(z) > 1 (every compact subset of
{z ∈ C : Re(z) > 1} is contained in such a half plane in which Re(z) ≥ 1 + δ).
2.4. SEQUENCES OF HOLOMORPHIC FUNCTIONS 77
θ:H→C
∞
2πin2 z 2z
X X
z 7→ θ(z) := e =1+2 e2πin
n∈Z n=1
Proposition 2.5. For all z ∈ H, the function θ(z) is well defined, i.e. the series
converges and defines a holomorphic function there.
Proof. We are going to show that θ converges uniformly on any subset of the form
Hδ := {z ∈ C : Im(z) ≥ δ} ⊂ C with δ > 0. Since any compact subset of H is
contained in such a set, by Theorem 2.14 this will imply the result.
P∞ 2πin2 z
Hence, n=0 e converges uniformly on Hδ for any δ > 0 and therefore also on
every of their compact subset. Thus it defines a holomorphic function on H, as argued
above.
Example 2.9 (Fourier analysis and the Theta function). The Fourier Transform serves
as important tool to Number Theory, in this settings we can think of the Theta function
as such a transform of a certain function. The uniqueness of the coefficients of this
transform allows the Theta function to identify the squares, for instance. So,
∞ ∞ 1 , m=0
2πin2 z 2πin2 z
X X
1+2 e = α(m)e =⇒ α(m) = 2 , m is a square
0 , else
n=1 n=0
78 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
Many special functions in mathematics are defined in terms of integrals of the type
Z b
f (z) := F (z, t)dt
a
(i) F ∈ C 0 (Ω × I)
Proof. The idea is to use the Riemann sums to approximate the integral: let
n−1
b−aX b−a
fn (z) := F z, a + j
n j=0 n
We want to show that (fn )n∈N∗ converges to f uniformly on compact subsets. Then
using Theorem 2.15 we can conclude that f is holomorphic.
Hence
∀ε > 0∃δ > 0∀(z1 , t1 ), (z2 , t2 ) ∈ K × I :
ε
|z1 − z2 | < δ and |t1 − t2 | < δ =⇒ |F (z1 , t1 ) − F (z2 , t2 )| <
b−a
Let now n ∈ N∗ be such that b−a
n
< δ, then let z ∈ K and consider
!
n−1 Z a+(j+1) b−a
X n b−a
fn (z) − f (z) = F z, a + j − F (z, t) dt
j=0 a+j b−a
n
n | {z }
| {z } dependent of t
independent of t
using
Z b n−1 Z
X a+(j+1) b−a
n
f (z) = F (z, t)dt = F (z, t)dt
a j=0 a+j b−a
n
and
n−1
b−aX b−a
fn (z) = F z, a + j =
n j=0 n
n−1 Z a+(j+1) b−a
X n b−a
= F z, a + j dt =
b−a
j=0 a+j n
n
n−1
F z, a + j b−a
X
n
= b−a
j=0 n
80 CHAPTER 2. CAUCHY’S THEOREM AND ITS APPLICATIONS
and finally
n−1
ε Xb−a ε b−a
|fn (z) − f (z)| ≤ = n =ε
b − a j=0 n b−a n
which gives the uniform convergence of fn to f on K, since z was arbitrary in that set.
Being K arbitrary we obtain by Theorem 2.15 that f is holomorphic.
Remark 2.17. With some more work one can also show that f ′ is given by
Z b Z b Z b
′ d ′ d
∀z ∈ Ω : f (z) = F (z, t)dt = F (z, t)dt = F ′ (z, t) dt
dz a a dz a | {z }
=ft′ (z)
Remark 2.18. Many special functions that appear as solutions of differential equa-
tion, for example Bessel functions, have integral representations,e.g. Jn (z) is defined as
solution of Bessel’s complex differential equations:
d2 f (z) df (z)
z2 2
+z + (z 2 − n2 )f (z) = 0
dz dz
For n ∈ Z, it holds that
Z π
1
Jn (z) = eiz sin(t) e−int dt
2π −π
Z π
1
and J0 (z) = eiz sin(t) dt
2π −π
The goal is to extend Cauchy’s Theorem 2.5 and the Cauchy Integral Formula 2.6 from
holomorphic functions to functions which might have singularities.
Recall: Cauchy’s Theorem 2.5 states that for any f ∈ H (Ω), any closed curve γ such
that the image of the curve and its interior are both contained in Ω, we have
Z
f (z)dz = 0
γ
The Cauchy Integral Formula 2.6 states instead that for any z ∈ D ⊂ Ω, with D a disc,
such that ∂D =: C and for any f |D ∈ H (D), we have
Z
1 f (w)
f (z) = dw
2πi C w−z
To this end, we are first going to look at isolated singularities of a function f . We will
see that there are three prototypes for these:
• removable singularities
• poles
• essential singularities
1
given respectively by the following functions at z = 0: sin(z)
z
, z1 , e z . For instance, for the
first function it holds that
∞
sin(z) X (−1)n z 2n
=
z n=0
(2n + 1)!
81
82 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
shows that z = 0 is a “removable” singularity. One can view the RHS1 as an analytic
continuation to all C of the LHS2 = sin(z)
z
. For the second function we have
1
lim =∞
z→0 z
1
whereas for the third function |e z | oscillates. In fact,
1 x↘0
• if z → 0 on positive real numbers, then |e x | −−→ ∞
1 x↗0
• if z → 0 on negative real numbers, then |e x | −−→ 0
These three examples of singularities are what we call a removable singularity, a pole
and an essential singularity respectively.
We will prove a generalisation of Cauchy’s Theorem 2.5 to functions that are holomor-
phic except for finitely many isolated points. This will lead us to the
Residue Formula: if f ∈ H (Ω) for an open Ω containing a circle C and its interior,
except finitely many points z1 , ..., zn with n ∈ N inside C, then
Z N
X
f (z)dz = 2πi Reszk (f )
C k=1
| {z }
a−1
where G ∈ H Dr (z0 )
This Theorem, like Cauchy’s Theorem 2.5 can be used to evaluate many real integrals
and complex line integrals. It will also lead to many theoretical results just like Cauchy’s
Theorem 2.5 did.
Argument principle: which allows us to count the number of zeroes (and poles) of
holomorphic (meromorphic) functions inside closed curves.
or if
∃Uz0 ∈ OCΩ : f |Uz ∈ H Uz0 \ {z0 }
0 \{z0 }
Note that in the above Definition 3.1 the term ”possible” means that the singularity in
question at that point can possibly be removed.
Example 3.1. Let tan z1 be defined in C without some countable set has singularities
at
2 2 2
, , ,...,0
π 3π 5π
Note that 0 is not an isolated singularity of f . The points
2 3 2
, ,...,
π 2π (2k + 1)π
f : C∗ = C \ {0} → C
z 7→ z
f : C∗ = C \ {0} → C
1
z 7→
z
has a singularity at z = 0, which cannot be removed.
Proof. Exercise.
If a point satisfies one of these equivalent conditions, we call it a removable singularity.
As a consequence of Riemann’s continuation Theorem we have Riemann’s Theorem
on removable singularities.
3.1. ZEROES AND POLES 85
Note that this represents (iii)=⇒(i) in the previous Theorem which states (iii)⇐⇒(i)
sin(z) cos(z)
lim = lim =1
z→z0 z z→z 0 1
sin(z)
which
imply that f (z) = z is bounded in a neighbourhood of 0, i.e. f ∈
B Ḋr (0) . In fact, it holds that
sin(z)
lim =1
z→0
z̸=0
z
sin(z) 1
−1 <
z 2
and therefore f ∈ B Ḋr (0) . One can also use that on C
∞
X (−1)n z 2n+1
sin(z) =
n=0
(2n + 1)!
sin(z)
is the holomorphic extension of z
to all of C
86 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
near z0 . We can then ask whether its unboundedness is similar to the unboundedness
of (z−z1 0 )n , i.e. we can ask whether (z − z0 )n f (z) is bounded near z0 for n sufficiently
large. If such an n ∈ N exists, then z0 is called a pole of f , let us hence first define this
once for all.
∃n ∈ N∗ ∃r > 0 : Ḋr (z0 ) ⊂ Ω and (z − z0 )n f |Ḋr (z0 ) ∈ B Ḋr (z0 )
is called the order of the pole of f at z0 . Poles of first order are called simple
poles.
Remark 3.1. To compute the order of pole of f at z0 , one can can compute the order
of zero of f1 at z0 , as shown later.
1
Example 3.4. The function f (z) = (z−z0 )m
for z ∈ C \ {z0 } has a unique pole of order
m at z = z0
We will see soon that poles arise from reciprocals of holomorphic functions with zeroes.
Before we make this more precise, let us recall that zeroes of holomorphic functions are
isolated and we have the following Proposition for their behaviour near a zero.
g(z)
∀z ∈ Ḋr (z0 ) : f (z) =
(z − z0 )m
Proof. (i)=⇒(ii): The fact that f has a pole of order m at z0 means that (z − z0 )m f (z)
is bounded near z0 and that m is minimal, thus ∃r > 0 : Dr (z0 ) ⊂ Ω on which
f in bounded. The Riemann’s Theorem on removable singularities 3.2 states that
∃g ∈ H Dr (z0 ) such that g(z) = (z − z0 )m f (z) whenever z0 ̸= z ∈ Dr (z0 )
g(z) = (z − z0 )g̃(z)
is bounded near z0 and this would contradict the minimality of m. Hence g(z0 ) ̸= 0
and together with it we get that f (z) = (z − z0 )−m g(z) for z ∈ Ḋr (z0 )
(ii)=⇒(iii): Suppose that ∃r > 0 as in (ii) and that ∃g ∈ H Dr (z0 ) such that
g(z0 ) ̸= 0 and f (z) = (z − z0 )−m g(z) for z ∈ Ḋr (z0 ). Since g(z0 ) ̸= 0 and g is contin-
uous, then ∃r > 0 : ∀z ∈ Dr (z0 ) : g(z) ̸= 0, meaning that r can be chosen adequately
since the beginning.
then h(z) ̸= 0 for z ∈ Ḋr (z0 ) and therefore h ∈ H Dr (z0 ) and with h(z0 ) = 0. Also,
(iii)=⇒(i): Suppose that ∃r > 0 : Dr (z0 ) ⊂ Ω and ∃h ∈ H Dr (z0 ) such that h(z) ̸= 0
1
∀z ∈ Ḋr (z0 ) : f (z) =
h(z)
Since h has a zero of order m at z0 , then by Proposition 2.2 ∃k ∈ H Dr (z0 ) such that
1 1
∀z ∈ Ḋs (z0 ) : f (z) = = (z − z0 )−m
h(z) k(z)
1
would imply that (z − z0 )m f (z) = k(z) is holomorphic on Ḋs (z0 ) and has the holo-
1
morphic extension k on Ds (z0 ) (given that k1 is holomorphic on Ds (z0 ) since k ̸= 0 on
Ds (z0 )).
(z − z0 )m f (z)
1
1. The function f (z) = ez −1
has a pole of order 1 at z = 0, since
∞
! ∞ ∞
!
1 X zn X zn X z n−1
= ez − 1 = −1= =z 1+
f (z) n=0
n! n=1
n! n=1
n!
1
has a zero of order 1 at z = 0 (note that ez −1
has simple poles at z = 2πik for
k ∈ Z).
z
2. The function f (z) = z 2 −1
has poles of order 1 at z = ±1, since
−1 z
f (z) = (z − 1)
z+1
z
and h(z) = z+1
is holomorphic and non-vanishing on D 1 (1). Similarly,
2
−1 z
f (z) = (z + 1)
z−1
z
and h̃(z) = z−1
is holomorphic and non-vanishing on D 1 (−1)
2
The next Theorem is the analogue of the power series expansion of a holomorphic func-
tion.
at z0 , then
n
X a−k
f (z) = G(z) +
k=1
(z − z0 )k
g [n−k] (z0 )
where ∀k ∈ {1, ..., n} : a−k := (n−k)!
P∞
Remark 3.2. The function f (z) := −n an (z − a)k is a special case of a Laurent
series (see Serie 9 and Serie 10).
Definition 3.4. • The number a−1 , i.e. the coefficient of (z − z0 )−1 in Theorem
3.4, is called the residue of f at the pole z0 , denoted by
Resz0 (f ) := a−1
3.1. ZEROES AND POLES 91
• The function n
X a−j
Pz0 (f, z) := Pzf0 (z) :=
j=1
(z − z0 )j
g [j] (z0 )
∀j ∈ {1, ..., n} : a−j :=
j!
as in latter proof.
Conversely, if limz→z0 (z −z0 )f (z) exists and is non-zero, then (z −z0 )f (z) is bounded in
some neighbourhood of z0 . Hence, z0 is a pole of f of order at most 1 by our definition
of pole.
If the limit exists and is equal to 0, then it means that f has a removable singularity at
z0
at z0 , then
1 dn−1
(z − z0 )n f (z)
Resz0 (f ) = lim n−1
z→z0 (n − 1)! dz
Proof. Let
f (z) = Pzf0 (z) + G(z)
92 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
a−j
and G ∈ H (Dr (z0 )). Then it holds that
Pn
for z ∈ Dr (z0 ) with Pzf0 (z) = j=1 (z−z0 )j
n
X
n n
(z − z0 ) f (z) = G(z)(z − z0 ) + a−j (z − z0 )n−j
j=1
for some G̃ ∈ H (Dr (z0 )) by the General Leibniz Rule [EW22]. Hence, the final result
is obtained by some final algebraic manipulations, so
1 dn−1 n
Resz0 (f ) = a−1 = lim (z − z0 ) f (z)
z→z0 (n − 1)! dz n−1
1
2. The function f (z) = has two poles of order 2 at z = ±i
(z 2 +1)2
1 d 2 1 d −2
Res±i (f ) = lim (z ± i) 2 = lim =
z→±i (2 − 1)! dz (z + 1)2 z→±i dz (z ± i)2
−2 ±1
= lim 3
=
z→±i (z ± i) 4i
Remark 3.4. The following one is a useful tool to calculate residues of simple poles.
Proof. It is clear that if g has a simple zero at z0 , then g(z) = (z − z0 )g̃(z) where
g̃(z0 ) ̸= 0, non-zero and holomorphic in some Dr (z0 ) ⊂ Ω for some r > 0
f (z) f (z)
= (z − z0 )−1
g(z) g̃(z)
where fg̃(z)
(z)
∈ H Dr (z0 ) and non-zero at z0 . So fg(z)
(z)
has a simple pole at z0 . We now
f
apply Theorem 3.5 to g , so
f f (z) (z − z0 )
Resz0 = lim (z − z0 ) = lim f (z) =
g z→z 0 g(z) |{z} z→z 0 g(z) − g(z0 )
g(z0 )=0
(z − z0 ) f (z0 )
= f (z0 ) lim = ′
z→z0 g(z) − g(z0 ) g (z0 )
since g is holomorphic at z0
Example 3.7. 1. Let g(z) = z21+1 with a simple pole at z = i, then
1 1 1 1
Resi = Resi = =
z2 + 1 g(z) g ′ (i) 2i
3
2. We want to determine Resi z2z+1 , hence we can either use partial fraction ex-
pansion as
z3 z 1 1 1 1
2
=z− 2 =z− −
z +1 z +1 2z −i 2z +i
3
and get Resi z2z+1 = − 21 , or use the above Lemma 3.1
3
z f (i) i3 1
Resi 2
= ′
= =−
z +1 g (i) 2i 2
with f (z) = z 3 and g(z) = z 2 + 1 (notably faster in this case).
Remark 3.5. Note that, if f (z) = Pzf0 (z) + G(z) for z ∈ Dr (z0 ), where Pzf0 is the
principal part of f at z0 and G is a holomorphic function, and C(z0 ) is any circle
centred at z0 and contained in Dr (z0 ), then
Z Z n
f
X a−k
Pz0 (z)dz = k
dz = 2πia−1
C(z0 ) C(z0 ) k=1 (z − z0 )
since3 Z
1 0 , n ̸= 1
dz =
C(z0 ) (z − z0 )n 2πi , n = 1
3
Using the Cauchy’s Integral Formula 2.6 with the constant function 1|D(z0 ) : D(z0 ) → C, z 7→ 1
and C = ∂D
94 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
Before we give the proof we will look at simple examples which use this formula to
calculate integrals.
Example 3.8. 1. Let γ be the circle such that |z| = 3 and consider the integral
Z
dz 1 1
= 2πi Resi + 2πi Res−i =
γ (z 2 + 1)2 (z 2 + 1)2 (z 2 + 1)2
1 −1
= 2πi + =0
4i 4i
3.1. ZEROES AND POLES 95
By Theorem 3.4
f (z) = Pzf0 (z) + G(z)
where G is holomorphic in a neighbourhood Dr (z0 ) of z0 and
n
X a−k
Pzf0 (z) =
k=1
(z − z0 )k
is the principal part of f at z0 , note that Pzf0 (z) is holomorphic in all of C \ {z0 }.
Another way to say this is that the function f (z) − Pzf0 (z) extends holomorphically to
Ω, as for the function
f (z) − Pzf0 (z) , z ∈ Ω \ {z0 }
g : Ω → C, z 7→ g(z) =
G(z0 ) , z ∈ Dr (z0 )
Recall: The Cauchy Integral Formula for derivatives 2.3. Let C = ∂D be any circle
whose interior D is contained in Ω. Then for F ∈ H (Ω) and any z ∈ D it holds that
Z
[n] n! F (w)
F (z) = dw
2πi C (w − z)n+1
hence,
2πi dz n−1
Z
dz 0 , n−1≥1
= (1) =
γ (z − z0 )n n! dz n−1 2πi , n − 1 = 0
and so we get Z
f (z)dz = 2πia−1 = 2πiz0 Res(f )
γ
For the general case consider f holomorphic in Ω except for finitely many points z0 , ..., zn
and let Pzfk be the principal part at zk for k ∈ {0, ..., n} =: I, which is holomorphic in
C \ {zk }
3.1. ZEROES AND POLES 97
This gives an extension of g to (Ω \ S) ∪ {z0 } = Ω \ {z1 , ..., zn }, we can do this for each
zk ∈ S with k ∈ I to get a holomorphic extension of g to all Ω. By Cauchy’s Theorem
2.5 we obtain that Z
g(z)dz = 0
γ
which in return gives Z XZ
f (z)dz = Pzfk (z)dz
γ zk ∈S γ
k∈I
Remark 3.6. 1. Another way to prove this is the following: first assume there is
just one pole inside of γ. Consider the following contour Γε,δ
inside of it f is holomorphic and we can show using Cauchy’s Theorem 2.5 that
Z
f (z)dz = 0
Γε,δ
Here we went around the pole z0 with a circle of radius ε > 0. The width of the
corridor is δ > 0
We can then make the width of the corridor narrower by letting δ → 0 and use
continuity of f to show that the two sides of the corridor cancel each other. The
remaining part consists of two curves, the larger circle γ and the small circle
Cε (z0 ) with clockwise orientation; we therefore get
Z Z
f (z)dz + f (z)dz = 0
γ Cε (z0 )
2. The best way to understand and generalise the Residue Formula 3.6 (and Cauchy
Integral Formula 2.6) is via homotopy. It is based on the following principle.
Let f be holomorphic on an open set Ω. For example, the space between two
circles.
3.1. ZEROES AND POLES 99
The principle is that if two closed curves can be deformed to each other while
remaining in Ω, then Z Z
f (z)dz = f (z)dz
γ1 γ2
Moreover,
r(t) = γ(t) − z0
and θ(t) is a continuous choice of argument along the line segment γ̃(t) = γ(t)−z0 ,
lastly we have that
γ(t) − z0
eiθ(t) =
γ(t) − z0
100 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
(This is similar to the parametrisation of a circle using a point inside other than
the centre)
Before we give more theoretical applications of the Residue Theorem 3.6, let us give
some applications to the evaluation of real integrals.
This of course can be evaluated easily using arctangent. We though give another proof
of it, using the Residue Theorem 3.6.
Idea: To choose a function f and a closed contour, so that part of the contour leads to
the real integral after taking limits.
1
In this particular case we take f (z) = 1+z 2
as function and γR as the contour path.
3.1. ZEROES AND POLES 101
for ΓR as the semicircular part of the closed path. As R → ∞, the first integral gives
Z ∞
1
2
dx
−∞ 1 + x
and similarly, as R → ∞, we see that over the semicircle ΓR the integral goes to zero.
This is because on ΓR : |z 2 + 1| > R2 − 1 and hence z21+1 < R21−1 ≈ R12 . Therefore, we
have that Z
1 1 π R→∞
2
dz < 2 πR ≈ −−−→ 0
ΓR 1 + z R −1 R
Hence finally we conclude that
Z ∞
1
dx = π
−∞ 1 + x2
Example 3.10. The same technique works to calculate the integrals of the form
Z ∞
P (x)
dx
−∞ Q(x)
where P, Q ∈ C[z], Q has no zero in the real axis and deg(Q) ≥ deg(P ) + 2
Note that we need this bound for the degrees of P and Q in order to get
Z
P (z) R→∞
dz −−−→ 0
ΓR Q(z)
102 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
namely
deg(Q) ≥ deg(P ) + 2
Once we get this result, we can proceed with the calculation of the initial integral: we
remember that Z Z R Z
P (z) P (x) P (z)
dz = dx + dz
γR Q(z) −R Q(x) ΓR Q(z)
gives as R → ∞ Z
P (z) X P
dz = 2πi Resz̃
γR Q(z) Q
z̃∈F ∩int(γR )
P
for F the set of all poles of Q
(therefore we then consider only the ones inside γR ).
Example 3.11. Z ∞
1 π
dx = 3
−∞ (x2 2
+a )2 2a
1
Consider the function f (z) = (z2 +a 2 )2 with poles at ±ai of order 2 each. Without loss
of generality, we assume that a > 0, since this would only switch the poles between
themselves in a manner that would way in the same result as the following.
1 d 2
d 1
Resai = lim (z − ai) f (z) = lim =
(z 2 + a2 )2 z→ai dz z→ai dz (z + ai)2
−2 −2 −2 −i
= lim = = =
z→ai (z + ai)3 (2ai)3 8a3 i 4a3
3.1. ZEROES AND POLES 103
As above we conclude Z
1
lim dz = 0
R→∞ ΓR (z 2 + a2 )2
and we get Z ∞
1 π
dx = 3
−∞ (x2 2
+a )2 2a
Example 3.12. The same contour can be used to evaluate integrals of rational functions
times sin(ax), cos(ax), for a ∈ R, i.e. of the form
Z ∞
P (x)
cos(ax)dx
−∞ Q(x)
e−a
iaz
e eiaz
Resi = lim =
z2 + 1 z→i z + i 2i
Hence
e−a
Z
f (z)dz = 2πi = πe−a
γR 2i
Since |eiaz | < 1 on the upper half plane, we have that
eiaz 1
2
≤ 2
z +1 R −1
and hence that
eiaz
Z
πR R→∞
2
dz ≤ 2 −−−→ 0
Γr z +1 R −1
and consequently
∞
eiax
Z
∀a > 0 : dx = πe−a
−∞ x2 + 1
104 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
cos(ax) eiax
we now note that x2 +1
= Re x2 +1
, hence by taking the real part of the function we
get Z ∞
cos(ax)
dx = πe−a
−∞ x2 + 1
This also shows that Z ∞
sin(ax)
dx = 0
−∞ x2 + 1
sin(ax)
which can also be directly seen, as x2 +1
is an odd function.
Example 3.13 (Integrals of trigonometric functions). The Residue Theorem 3.6 can
be used to evaluate real integrals of the form
Z 2π
P cos(t), sin(t)
dt
0 Q cos(t), sin(t)
where P, Q are polynomials and where it holds that ∀x, y ∈ R : (x2 + y 2 = 1) =⇒
Q(x, y) ̸= 0 (every pair (x, y) on C1 (0) has Q(x, y) ̸= 0.)
We now turn to more theoretical applications of the Residue Theorem 3.6. We start
by giving one more description of an isolated singularity, which is a pole. Namely, we
have the following Proposition.
Proof. If f has a pole of order k ∈ N∗ at z0 , then by Theorem 3.3 we have that there
exists some r > 0 for which f (z) = (z − z0 )−k h(z) on Ḋr (z0 ) with a bounded function
h ∈ H Dr (z0 ) for which it holds that h(z0 ) ̸= 0. Then
z→z
f (z) = (z − z0 )−k h(z) −−−→
0
∞
z→z
0
since h(z) −−−→ h(z0 ) ̸= 0 and k ≥ 1
z→z
0
Conversely, if f (z) −−−→ ∞, then we can find some r > 0 such that f (z) ≥ 1 (or any
0 ). In particular, f (z) ̸= 0 on Ḋr (z0 )
ε > 0 would also be suited for the case) on Ḋr (z
by Theorem 3.3 and there is a h ∈ H Ω \ {z0 } such that
1
h(z) :=
f (z)
is holomorphic in Ḋr (z0 ) and h(z) ≤ 1 there. Furthermore, it holds that limz→z0 h(z) =
1
limz→z0 f (z) = 0. By the Riemann’s Theorem on removable singularities 3.2, h extends
1
to a holomorphic function h̃ in Dr (z0 ) by defining h̃(z0 ) := limz→z0 f (z)
= 0 and other-
wise h̃ = h|Ḋr (z0 )
1
Therefore, if N ∈ N is the order of zero of h̃ at z0 , then f (z) = h(z)
has a pole of order
N ∈ N at z0
We have seen that an isolated singularity z0 of f is removable if f is bounded near z0 ,
z→z0
at the same time z0 is a pole if f (z) −−−→ ∞
1
As we saw in the very beginning, the function e z has a more exotic behaviour. E.g.
1 1
e x → 0 as x ↗ 0 from the negative reals, whereas e x → ∞ as x ↘ 0 from the positive
reals.
106 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
tial singularity at z0
Then the image of Ḋr (z0 ) under f , namely f Ḋr (z0 ) , is dense in C
Remark 3.7. The Casorati-Weierstrass Theorem 3.7 states that the image of a punc-
tured disc Ḋr (z0 ), no matter how small, effectively fills up the whole complex plane
(where z0 is an essential singularity). In fact, a remarkable Theorem of Picard says
Theorem 3.8 (Picard’s Theorem (1879)). If f ∈ H Ḋr (z0 ) and has an essential
singularity at z0 , then C \ f Ḋr (z0 ) contains at most one point.
1
The function f (z) = e z maps each punctured disc centred at z = 0 to C∗ , i.e. it does
not take the value 0, so the “exceptional value” permitted by Picard’s Theorem 3.8
may in fact exist.
Proof of Theorem 3.7. We want to show that for the given r > 0
and to do so we argue by contradiction and show that this will forces the singularity at z0
to be either removable or a pole, hence contradicting the assumption that z0 is essential.
Assume on the contrary that the image in question is not dense in C, hence for the
given r > 0
∃δ > 0∃w0 ∈ C∀z ∈ Ḋr (z0 ) : f (z) − w0 ≥ δ
1
Let g : Ḋr (z0 ) → C, z 7→ g(z) := f (z)−w0
, then on Ḋr (z0 ) we have that g is bounded
1
by δ , hence has a removable singularity at z0 by Riemann’s Theorem on removable
singularities 3.2. Hence, there is a holomorphic extension of g, i.e. we can define g at
z0 , so that g becomes holomorphic in Dr (z0 )
1
Since f (z) − w0 ≥ δ and g(w) = f (z)−w 0
, clearly g has no zero in Ḋr (z0 ), hence its
1
reciprocal g has an isolated singularity at z0 ∈ Dr (z0 ). This singularity is either a
removable one or a pole, depending on whether limz→z0 g(z) = 0 or not, respectively.
In turn, this gives that the singularity of f = w0 + g1 at z0 can be at most a pole, giving
the anticipated contradiction. Note that the limit limz→z0 g(z) exists, since g has a
removable singularity at z0
3.2. MEROMORPHIC FUNCTIONS 107
f : C∗ → C
1
z 7→
z
can be extended to
fˆ : C → C ∪ {∞}
1
z 7→
z
∀z ∈ C : ∞ ± z = z±∞ = ∞
∀z ∈ Ĉ \ {0} : ∞ · z = z·∞ = ∞
z
∀z ∈ C : ∞
= 0
z
∀z ∈ Ĉ \ {0} : 0
= ∞
lim |zn | = ∞
n→∞
∗
where |zn | n∈N∗ ∈ RN . Similarly, we say that limz→z0 f (zn ) = ∞, if
lim f (zn ) = ∞
z→z0
Talking now about functions, we can extend our definition of holomorphic function to
a more general notion.
108 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
is discrete in Ω
Note that the set of poles for meromorphic functions is discrete, as the set of zeroes for
holomorphic functions.
Then f ∈ M (C), since f is holomorphic outside the finite zeroes’ set of Q(z). If
z0 is a zero of Q, then f has a pole at z0 , as
P (z)
lim f (z) = lim =∞
z→z0 z→z0 Q(z)
If we have two functions f, g ∈ M (Ω) with pole sets Sf and Sg , then f +g is holomorphic
in Ω \ (Sf ∪ Sg ) and we can define f + g at points in this set using the usual definition:
where Pzf0 , Pzg0 are the principal parts of f and g at z0 (one of them can be zero if f or
g does not have a pole at z0 ) and where f˜, g̃ ∈ H Dr (z0 ) . Then
so f + g has a pole of order ≥ 1 at z0 and we assign the value ∞ to that point under
f + g, this unless Pzf0 (z) + Pzg0 (z) = 0 (which can happen). Hence, f + g ∈ M (Ω) with
Sf +g ⊆ Sf ∪ Sg
(iv) If 0 ̸= f ∈ M (Ω) and the zeroes of f do not have a limit point in Ω, then
1
f
∈ M (Ω)
Proof. (i) Obvious, but note that we identified a holomorphic function f ∈ H (Ω)
with the corresponding function f˜ ∈ M (Ω), where f˜ = i ◦ f and i : C ,→ Ĉ
l terms and G ∈ H
1
where Pzf0g is a linear combination of (z−z 0 )
Dr (z0 ) for some
r > 0 by Theorem 3.4 (G is the sum of the respective functions in the case of f
110 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
and g separately), so
∞
! ∞
!
X X
fg = ak (z − z0 )k bl (z − z0 )l =
k=−n l=−m
∞
X X
= ak bN −l (z − z0 )N
N =−(n+m) k,l
k+l=N
1 z→z0
−−−→ 0
f (z)
Recall: If f ∈ CΩ , Ω open and connected and f ∈ H (Ω), then the zeroes of f do not
have a limit point in Ω (see Theorem 2.10).
For an open and connected set Ω, the same is true for f ∈ M (Ω)
3.2. MEROMORPHIC FUNCTIONS 111
∗
Proof. Assume on the contrary that ∃(zn )n∈N∗ ∈ ZfN of distinct points such that
limn→∞ zn = b ∈ Ω
But b ∈
/ Sf either, since if b were a pole of f , then limz→b f (z) = ∞ and it would
mean that for ε > 0 we would have f (z) > 0 for z ∈ Ω with |z − b| < ε. But this is
impossible, since if zn → b, then |zn − b| < ε for n ≥ n0 and f (zn ) = 0
Remark 3.10. Let f ∈ M (Ω) and z0 a pole of f , since Sf has no limit point in Ω,
then it exists a punctured neighbourhood Ḋr (z0 ) of z0 for some r > 0 such that
Ḋr (z0 ) ∩ Sf = ∅
If the order of the pole of f at z0 is k, then f (z) = (z − z0 )−k g(z) with an analytic
function g(z) ∈ H Dr (z0 ) by Theorem 3.3.
Hence, locally every meromorphic function is the quotient of two holomorphic functions.
Here:
g(z)
f (z) =
(z − z0 )k
It is a non-trivial result that if Ω is non-empty, open and connected (i.e. a region), then
we can do this globally.
For any f ∈ M (Ω) with Ω open and connected, there exist g, h ∈ H (Ω) such that f = g
h
Algebraically, we can state this as follows: recall that if Ω is open and connected, then
H (Ω) has no zero divisor, hence it is an integral domain. Consequentially, it has a
quotient field (or field of fractions):
ng o
Q H (Ω) = : g, h ∈ H (Ω) and h ̸= 0 = M (Ω)
h
This is similar to the construction of Q as field of fractions of the integral domain Z
112 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
In particular,
Combining what we know about the behaviour of functions near zeroes and poles (The-
orem 2.2 and Theorem 3.3), we get:
z 2 2
Example 3.15. Let f (z) = (ez −1) 2 , z has zero of order 1 at z = 0, while (e − 1) has
Explicitly, π is given by
x1 x2 x1 x2
π(p) = π(x1 , x2 , x2 ) = , ,0 = + i
1 − x3 1 − x3 1 − x3 1 − x3
Note that the equation of the ray that starts at N and goes through p is:
N + t(p − N ), t≥0
and consequently
π(p) = N + t0 (p − N )
where t0 is unique positive real number, so that (0, 0, 1) + t0 (x1 , x2 , x3 − 1) has third
coordinate equal to 0. Solving this equation for t0 gives the formula for π(p) above.
Defining π(N ) = ∞ gives a bijection
π : S 2 → Ĉ
Conversely, given z ∈ C one checks that
|z|2 − 1
−1 2x 2y
π (z) = 2
, 2 , 2 ∈ S 2 \ {N }
|z| + 1 |z| + 1 |z| + 1
and π −1 (∞) := N gives the inverse map. Here we get S 2 homeomorphic to Ĉ, since
both maps are continuous.
114 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
Before we study the values of holomorphic functions using the Residue Formula 3.6, let
us mention that we can also talk about meromorphic functions on Ĉ (as opposed to
M (Ω) with Ω ⊆ C).
1
If a function f is analytic for large values of z, i.e. |z| > R
for some R > 0, then the
function
1
g(z) := f
z
is holomorphic in a deleted neighbourhood of 0, i.e. in ḊR (0)
Definition 3.10. • For a function f , which is analytic for |z| > R1 for some
R > 0, we say that f has an isolated singularity
1
at ∞, which will be called
removable, a pole or essential, if g(z) = f z has an isolated singularity at
0 (which is removable, a pole or essential respectively).
Example 3.16. 1. An entire function is analytic in ḊR (∞) for every R > 0
E.g. the function f (z) = ez has an isolated singularly at ∞, which is essential,
1
because e z has an essential singularity at 0. Hence, ez is not meromorphic in Ĉ,
but it is meromorphic on C
Pn k
2. The function p(z) ∈ C[z] has
Pn ak a pole at ∞. If p(z) = a 0 + k=1 ak z for some
1
n ∈ N, then p z = a0 + k=1 zk has a pole of order n at 0
3. The function f (z) = tan(z) does not have an isolated singularity at ∞: Each
ḊR (∞) includes poles of f with z = π2 + kπ, k ∈ Z
n −1 o
1 π
Also note that g(z) = tan z
has singularities S := 2
+ kπ : k ∈ Z , which
1
accumulate at z = 0. Hence, the singularity of tan z
at z = 0 is not isolated.
The following Theorem for meromorphic functions on Ĉ
Proof. Exercise.
f′
Lemma 3.2. Let Ω ⊆ C be open and connected, also 0 ̸= f ∈ M (Ω), then f
∈
′
M (Ω), the logarithmic derivative of f , is also meromorphic in Ω. Moreover, ff
has poles of order 1 at all z0 ∈ Ω, for which ordz0 (f ) ̸= 0, i.e. either z0 is a zero or a
pole of f ′
f
∀z0 ∈ Zf ∪ Sf : Resz0 = ordz0 (f )
f
Proof. Since f ̸= 0, Ω open and connected, being f ∈ M (Ω), the zeroes of f do not
have a limit point in Ω, and f1 ∈ M (Ω)
−n h′ (z)
f ′ (z) = h(z) + =
(z − z0 )n+1 (z − z0 )n
1
= h′ (z)(z − z0 ) − nh(z)
| {z } (z − z0 )n+1
=:h̃(z)
where h̃ ∈ H Dr (z0 ) and h̃(z0 ) = −nh(z0 ) ̸= 0 by definition. Hence, for all z ∈ Ḋr (z0 )
f′
Hence, f
has a pole of order 1 at the points where ordz0 (f ) ̸= 0
where g ∈ H Dr (z0 ) , g(z) ̸= 0 for all z ∈ Dr (z0 ) and n = ordz0 (f ). Hence, we have:
and
f ′ (z) n(z − z0 )n−1 g(z) + (z − z0 )n g ′ (z)
∀z ∈ Dr (z0 ) : = =
f (z) (z − z0 )n g(z)
n(z − z0 )n−1 g(z) (z − z0 )n g ′ (z) n g ′ (z)
= n
+ n
= +
(z − z0 ) g(z) (z − z0 ) g(z) (z − z0 ) g(z)
| {z }
=G(z)∈H Dr (z0 )
Hence
f′
Resz0 = n = ordz0 (f )
f
Lemma 3.2 immediately gives, using the Residue Theorem 3.6, the following result:
Theorem 3.10 (Argument Principle). [SS10, Theorem III.4.1] Let Ω ⊆ C open and
connected, f ∈ M (Ω) and let im(γ) ⊆ Ω be a circle (or any other curve such that
the Residue Formula 3.6 holds). If f has no zeros or poles on γ, then
Z ′
1 f (z) X X X
dz = ordz0 (f ) + ordz0 (f ) = ordz0 (f )
2πi γ f (z)
z0 ∈Zf ∩int(γ) z0 ∈Sf ∩int(γ) z0 ∈int(γ)
ordz0 (f )̸=0
′
f
Proof. This follows from the previous Lemma 3.2 and Resz0 f
= ordz0 (f )
f ′ (z) f′
Z
1 X
dz = Reszk = Z −P
2π γ f (z) f
zk ∈int(γ)
where Z is the number of zeroes of f inside γ counted with multiplicity (to express the
order of zero) and P is the number of poles of f inside γ counted with multiplicity (to
express the order of pole).
118 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
Corollary 3.1. Let f ∈ C[z] be a polynomial. Choose R > 0 large enough, so that
all zeroes of f are inside DR (0), then
f ′ (z)
Z
dz = deg(f )
CR (0) f (z)
We have the following Corollary of the Argument Principle Theorem 3.10, which says
that a holomorphic function, when perturbed slightly, does not change its number of
zeroes.
Proof. For t ∈ [0, 1], define ft (z) = f (z) + tg(z) so that f0 (z) = f (z) and f1 (z) =
(f + g)(z). Note that for z ∈ C it results that
given the assumption. Hence, we have that ∀z ∈ C : ft (z) > 0 and therefore ft has
no zero in C
Note that if we can show that
Z ′
1 ft (z)
nt := dz
2πi C ft (z)
is a constant, this will show that f and f +g have the same number of zeroes inside of C
By Argument Principle 3.10 applied to ft (z) (which by the above consideration has no
zero in C), we have that Z ′
1 ft (z)
nt = dz
2πi C ft (z)
counts the number of zeroes of ft inside of C, in particular it is integer valued.
f ′ (z)
Note that nt is a continuous function of t, since ftt (z) is jointly continuous on [0, 1] for
all z ∈ C, as both ft′ (z) and ft (z) are continuous and ft (z) ̸= 0 for all z ∈ C
3.3. APPLICATIONS OF THE RESIDUE THEOREM 119
h : [a, b] × [c, d] → R
Rd
is continuous on [a, b] × [c, d], then F (t) := c h(t, x)dx is continuous on [a, b].
Using this gives that Z ′
1 ft (z)
dz
2πi C ft (z)
is continuous and since nt is also integer valued, it must be a constant (otherwise the
Intermediate Value Theorem [EW22] gives the existence of t0 ∈ [0, 1] such that nt0 is
not integral). Hence, n0 is the number of zeroes inside of f and n1 is the number of
zeroes inside of f + g, so finally
Z ′
(f + g)′ (z)
Z
1 f (z) 1
dz = n0 = n1 = dz
2πi C f (z) 2πi C (f + g)(z)
Example 3.17. We use Rouché’s Theorem 3.11 to show that the polynomial
p(z) = z 6 + 8z 4 + z 3 + 2z + 3
The idea is to write p = Big + Small on C1 (0), such that Big(z) = 8z 4 = f (z) and
Small(z) = z 6 + z 3 + 2z + 3 = g(z), so: g(z) = |z 6 + z 3 + 2z + 3| < |8z 4 | = f (z) on
C1 (0), so with |z| = 1
Hence, by Rouché’s Theorem 3.11, f (z) = 8z 4 and (f + g)(z) = p(z) have the same
number of zeroes inside the unit circle. f has four zeroes (counted with multiplicity)
and so does p
Example 3.18. Rouché’s Theorem 3.11 can also be used to give a “nice4 ” or simple
proof of the Fundamental
P Theorem of Algebra.
d d−1 k
Let p(z) = z + k=1 ak z + a0 for which, if |z| is large enough, the term |z|d
Pd−1
dominates. Choose R large enough, so that for f (z) = z d and g(z) = k=1 ak z k + a0
we have:
f (z) > g(z)
on CR (0) and hence p = f + g and f have the same number of zeroes inside CR (0). We
obtain that f has d zeroes inside CR (0) and so does p
Rouché’s Theorem also leads us to two other important Theorems:
4
(subjectively) “nice” means “objectively nice”, if said from the Professor.
120 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
Let r > 0 such that Dr (z0 ) ⊂ U and such that ∀z ∈ D˙ r (z0 ) : f (z) − w0 ̸= 0; this is
allowed, since the zeroes of f˜(z) := f (z) − w0 are isolated. In particular, we have that
f (z) − w0 ̸= 0 for z on the circle Cr (z0 )
Since Cr (z0 ) is compact and f (z) − w0 ̸= 0 on Cr (z0 ), we can find a δ > 0 such that
f (z) − w0 ≥ δ
We want to show that F has a zero inside the circle Cr (z0 ). This will show that
∃z ∈ Dr (z0 ) : f (z) = w and hence that w ∈ f Dr (z0 )
We now apply Rouché’s Theorem 3.11 to f˜, g̃ on the circle Cr (z0 ), where we have that
and so f˜ and F̃ = f˜+ g̃ have the same number of zeroes inside Dr (z0 ). Since f˜ = f −w0
conclude that ∃z ∈ Dr (z0 ) : F (z) = 0, i.e.
has a zero inside Dr (z0 ), namely z0 , we must
∃z ∈ Dr (z0 ) : f (z) = w, so w ∈ f Dr (z0 ) as wanted.
Remark 3.12. This Theorem 3.12 states, for example, that if f ∈ H (Dr (z0 )) and non-
constant for some z0 ∈ C, then it is not possible to that f (z) ∈ R for all z ∈ Dr (z0 ),
since any subset of R is not open in C
Example 3.19 (Necessity of the boundedness of Ω). The assumption that Ω is bounded,
hence compact, is crucial, as shown in this example.
Let Ω = z ∈ C : Im(z) ∈ − π2 , π2
be open in C and connected, it is not hard to see
that the set Ω is not bounded. Consider the function f (z) = exp (ez ) on Ω, then
π
f |∂Ω (z) = exp ex±i 2 = exp(±iex )
Proof of Theorem 3.13. We first note that ∃ maxz∈Ω f (z) , since Ω is a compact set
and f ∈ C 0 (Ω), as shown in [EW22].
Let f ∈ H (Ω) and non-constant. Suppose also on the contrary that f attains a
maximum at z0 ∈ Ω. By the Open mapping Theorem 3.12, if f is an open map, by
letting D = Dr (z0 ) ⊆ Ω, then f (D) is open in C and contains f (z0 )
Hence f (D) contains a disc D̃ around f (z0 ) and therefore there are points z ∈ D such
that
f (z0 ) < f (z)
which contradicts the assumption that |f | attains its maximum at z0 (in any disc in C
one can find such points).
If Ω is bounded and f is non-constant and continuous, then f (z) attains its maximum
on Ω, since it is a continuous function on a compact set, as shown in [EW22]. By the
first part of the Theorem, this point where f attains its maximum cannot be inside Ω.
Hence, it has to be on the boundary Ω \ Ω = ∂Ω
for the two paths γ0 : [a, b] → Ω and γ1 [a, b] → Ω, either closed or with fixed endpoints.
3.4. HOMOTOPY AND SIMPLY CONNECTED DOMAINS 123
Not closed curves of this type curves are called homotopic with fixed endpoints: this
means that for each s ∈ [0, 1] it exists a curve γs in Ω parameterised by γs (t), hence
γs : [a, b] → Ω such that
and such that at s = 0 we have γs (t)|s=0 = γ0 (t) and at s = 1 we have γs (t)|s=1 = γ1 (t).
All this should be done continuously.
• Let γ0 : [a, b] → Ω and γ1 : [a, b] → Ω be two curves such that γ0 (a) = γ1 (a)
and γ0 (b) = γ1 (b), i.e. they have the same endpoints.
We says that γ0 is homotopic to γ1 in Ω with fixed endpoints, and denote
it by γ0 ∼Ω γ1 rel ∂[a, b], if
such that
such that
Note that if clear enough, the notation for fixed endpoints, namely “rel ∂[a, b]” will here
be omitted.
Example 3.20. 1. If Ω = C, then any two closed curves γ0 , γ1 are homotopic. In
particular, every closed curve is homotopic to the constant curve σc : [a, b] →
C, t 7→ c for every c ∈ C
124 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
H : [a, b] × [0, 1] → C
(t, s) 7→ (1 − s)γ0 (t) + sγ1 (t)
H(t, 0) = γ0 (t)
H(t, 1) = γ1 (t)
H(a, s) = (1 − s)γ0 (a) + sγ1 (a)
H(b, s) = (1 − s)γ0 (b) + sγ1 (b)
Since γ0 (a) = γ0 (b), γ1 (a) = γ1 (b) and ∀s ∈ [0, 1] : H(a, s) = H(b, s). Hence
γs : [a, b] → C are all closed curves.
Note that geometrically H is defined using the line segment between γ0 (t) and γ1 (t)
for each fixed t ∈ [a, b]. Hence, s ∈ [0, 1] varies over the line segment between
γ0 (t) and γ1 (t) for each fixed t ∈ [a, b].
For the constant curve σc , we can take the homotopy between σc and γ as
H : [a, b] × [0, 1] → C
(t, s) 7→ (1 − s)c + sγ(t)
Note that the same definition we used for closed curves γ0 , γ1 also gives a homotopy
with fixed points, if γ0 : [a, b] → Ω and γ1 : [a, b] → Ω are two curves with fixed
endpoints, i.e. with γ0 (a) = γ1 (a) and γ0 (b) = γ1 (b)
2. Note that the same formula for the homotopy works for any convex Ω ⊆ C. I.e.
if we have two curves γ0 , γ1 either closed or with a fixed endpoints in a convex set
3.4. HOMOTOPY AND SIMPLY CONNECTED DOMAINS 125
Ω, then since for a convex set the line segment between any two points is also in
the set, the function defined by
H : [a, b] × [0, 1] → C
(t, s) 7→ (1 − s)γ0 (t) + sγ1 (t)
gives a homotopy in Ω.
In particular, this works for Ω with the form of a disc.
γ0 : [0, π] → Ω
t 7→ eit
and
γ1 : [0, π] → Ω
t 7→ e−it
We will see a simple proof of this when we will see the homotopy version of
Cauchy’s Theorem 2.5. Intuitively, to deform γ0 to γ1 we have to go through 0,
which is not in Ω though.
4. The set Ω = C \ (−∞, 0] is not convex and therefore we cannot use the previ-
ous formula, but still any two closed curves γ0 , γ1 in Ω are homotopic, i.e. we
R deform γR0 to γ1 in the following way. Then for f ∈ H (Ω) we have that
can
γ0
f (z)dz = γ1 f (z)dz
126 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
The idea is to choose any point on the real line, say c ∈ R, and the constant curve
σc : [a, b] → Ω
t 7→ c
(ii) or γ0 , γ1 have the same endpoints and are homotopic with fixed endpoints.
Example 3.21. 1. Let Ω = DR (z0 ) and R > 0 and let im γ0 = Cr (z0 ) with r ∈
(0, R) as in the following picture.
Then γ0 can be deformed into the point z0 by dilation (which can be be thought as
the constant curve for which ∀t ∈ [a, b] : γ1 (t) = z0 ), so for f ∈ H (Ω)
Z Z
f (z)dz = f (z)dz = 0
Cr (z0 ) σz0
128 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
2. Let Ω = C∗ and
γ0 : [0, π] → Ω γ1 : [0, π] → Ω
t 7→ e it
t 7→ e−it
The two paths are not homotopic with fixed endpoints, since if they were, then we
would get that for z1 ∈ H (Ω) it would hold that
Z Z
1 1
dz = dz
γ0 z γ1 z
and therefore Z Z Z
1 1 1
dz − dz = dz = 0
γ0 z γ1 z C1 (0) z
but Z
1
dz = 2πi ̸= 0
C1 (0) z
We now look at the proof of the Homotopy Theorem 3.14. We will look at the case of
closed curves (in [SS10] one finds instead the version where the endpoints of a curve
are fixed).
Proof of the Theorem 3.14. 1. A simpler version of the proof is given with the ex-
tra assumption that the homotopy H(t, s) has continuous second order partial
derivatives and
∂ 2H ∂ 2H
∀(t, s) ∈ [a, b] × [0, 1] : (t, s) = (t, s)
∂s∂t ∂t∂s
For this we first recall from Real Analysis:
G : [0, 1] → R
Z b
s 7→ G(s) := h(t, s)dt
a
=h(t,s)
We want to show that I(0) = I(1) by showing that I(s) is constant. So consider
Z b
′ ∂ ∂H
I (s) = (f ◦ H)(t, s) (t, s) dt =
a ∂s ∂t
Z b
′ ∂H ∂H ∂ ∂H
= (f ◦ H)(t, s) (t, s) (t, s) + (f ◦ H)(t, s) (t, s) dt
a ∂s ∂t ∂s ∂t
and note that what is inside the round parenthesis is also equal by assumption to
∂ ∂H
(f ◦ H)(t, s) (t, s)
∂t ∂s
2. For the general proof the idea is the following: if we make a small deformation
of one of the curves γs (t), say γ0 (t) to γ 1 (t), so that if we look at a small piece
N
around a point of γ0 (t), say t ∈ (tm , tm+1 ) for m ∈ {0, ..., N − 1} and N the
cardinality of the dissection of [a, b] and of [0, 1], then we can show that these are
contained in a small disc in Ω
Let t ∈ (tm , tm+1 ) for m ∈ {0, ..., M − 1} and M the cardinality of the dissection
of [a, b]
H(t, 0) = γ0 (t)
1
H t, = γ 1 (t)
N N
Now, we move over to the whole curve γ0 , γ 1 using small discs contained in Ω to
N
get Z Z
f (z)dz = f (z)dz
γ0 γ 1
N
1
|wnk − znk | <
nk
so (wnk )k∈N∗ also converges to z. (wnk )k∈N∗ is also in C \ Ω, so its limit point is
z ∈ C \ Ω and this is a contradiction.
This Lemma 3.3 together with (b) will allow us to find the small discs that are
contained in Ω. This is because we can divide the rectangle [a, b] × [c, d] into small
rectangles such that the images of these small rectangles are contained in small
132 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
discs of radius ε
More precisely, let (tm )m∈{0,...,N } ∈ [a, b]{0,...,N } and (sn )n∈{0,...,N } ∈ [0, 1]{0,...,N }
represent the dissections of the two intervals and ε > 0. Since H is uniformly
continuous on [a, b] × [c, d], then it exists an N ∈ N∗ such that
whenever
2
(t, s) − (tm , sn ) <
N
b−a n
for zm,n = H(tm , sn ) with tm := a + N
m and sn := N
for m, n ∈ {0, ..., N }
Let Qm,n := [tm , tm+1 ] × [s , s ] for m, n ∈ {0, ..., N − 1}, since the diameter of
√ n n+1
2
Qm,n is diam(Qm,n ) = N < N2 , it follows that
H(Qm,n ) ⊆ Dε (zm,n )
n−1 n
where γ n−1 = H t, N
and γ Nn = H t, N
N
and let σm := ℓ[zm,n−1 ,zm,n ] be line segment between zm,n−1 and zm,n , thus σm+1 =
ℓ[zm+1,n−1 ,zm+1,n ] be the line segment between zm+1,n−1 and zm+1,n ; as special case
we consider σ0 = [z0,n−1 , z0,n ] and σN = [zN,n−1 , zN,n ]
Now we apply Cauchy’s Theorem 2.5 in the disc Dε (zm,n−1 ) and obtain
Z Z Z Z
f (z)dz − f (z)dz − f (z)dz + f (z)dz = 0
(m) (m)
γn σm+1 γ n−1 σm
N N
In spaces like C, C\(−∞, 0] or any convex set, we observed that any two closed curves, or
any two curves with the same endpoints, are homotopic. This leads us to the following
definition.
3.5. THE HOMOTOPY THEOREM 135
for any closed curve γ in Ω and that any two primitives differ by a constant.
Proof. Let Ω ⊆ C be simply connected and fix z0 ∈ Ω, we define F ∈ CΩ such that for
any z ∈ Ω
Z
F (z) := f (w)dw
γz
Let z ∈ Ω, if we choose h small enough, so that the image of the line segment
im ℓ[z,z+h] ⊆ Ω, then
Z
F (z + h) − F (z) = f (w)dw
ℓ[z,z+h]
136 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
Arguing as in the proof of Theorem 2.3 or using continuity of f as below, we get that
F (z + h) − F (z)
lim = f (z)
h→0 h
which shows that F is a primitive of f in Ω, as z was arbitrary. I.e.
Z
F (z + h) − F (z) = f (w) − f (z) + f (z) dw =
ℓ[z,z+h]
Z Z
= f (z) dw + f (w) − f (z) dw
ℓ[z,z+h] ℓ[z,z+h]
| {z }
=h
so Z
f (w) − f (z) dw ≤ h sup f (w) − f (z)
ℓ[z,z+h] w∈im ℓ[z,z+h]
hence
F (z + h) − F (z)
− f (z) ≤ sup f (w) − f (z)
h
w∈im ℓ[z,z+h]
h→0
Being f is of C 0 - class, it implies that sup f (w) − f (z) −−→ 0
w∈im ℓ[z,z+h]
log(z) = log(r) + iθ
where log(r) is the usual logarithm log : R+ → R of the positive real number r. The
problem is that this is not single valued, as θ is only unique up to an integer multiple
of 2π. Indeed, the argument is multivalued.
Example 3.22. Let z = 1 ∈ C, it holds that e0 = 1, but also that for any w ∈ 2πiZ
the equality ew = 1 holds.
exp ◦ℓ = id
If Ω is clear from the text, sometimes this function is also denoted by log
Note that any function f ∈ H (Ω) that meets that condition is a branch of logarithm
on Ω, but when one is fixed, then it it denoted by logΩ
Remark 3.14. 1. Since exp(z) ̸= 0 for all z ∈ C, in order for logΩ to exist we need
that 0 ∈
/Ω
C
2. If Ω = C∗ , even though exp ∈ C∗ is surjective, there is no branch of the
logarithm on Ω. Indeed, if there were a f ∈ H (Ω) such that ∀z ∈ Ω : exp f (z) =
z, then differentiating on both sides would give
Let G : CΩ such that z 7→ G(z) := z exp − f (z) , since f ′ (z) = z1 we have that
G′ (z) = −zf ′ (z) exp − f (z) + exp − f (z) = − exp − f (z) + exp − f (z) = 0
Since Ω is connected, it follows that G(z) = ze−f (z) = a (constant) necessarily, for some
a ∈ C. Moreover, since exp ̸= 0, z ̸= 0, also a ̸= 0, so we obtain that it exists b ∈ C
such that a = exp(b) and therefore, by algebraic manipulations on G, that
z
exp f (z) =
a
Hence, let F ∈ CΩ such that z 7→ F (z) := f (z) + b, then
exp F (z) = exp f (z) + b = exp f (z) exp(b) = z
| {z } | {z }
= az =a
Definition 3.14. Let C− := C\(−∞, 0]. The principal branch of the logarithm
on C− is the unique function logC− ∈ H (C− ) such that logC− (1) = 0. This particular
logC− is also denoted by Log (with capital L) or LogC− .
Proposition 3.5. If z = reiθ ∈ C− with r > 0 and θ ∈ (−π, π), then the principal
branch of the logarithm is given by the formula
LogC− (z) = log(r) + iθ = log |z| + i Arg(z)
−
hence, a primitive of z1 by Theorem
R 13.15, where we take a path γz in C which starts
at 1 and ends at z. Note that γ1 w dw = 0, hence logC− (1) = 0. This implies that
this branch of logarithm is equal to the principal one (we are going to use the usual
notation for this branch from now on).
If z = reiθ with r < 1, take the path γz that goes on the real line from 1 to r, then on
the circular arc to z, so
1 −θ
−ire−it
Z Z
dx
Log(z) = − + dt = log(r) +iθ
r x 0 re−it | {z }
| {z } | {z } the real one
on the x-axis on the arc z = re−it for t ∈ (0, −θ)
does not hold in general for all z, w, zw ∈ C− , but if w = reiα , z = seiβ and
zw = sreiθ with α, β, θ ∈ (−π, π), then exists γ ∈ {−2π, 0, 2π} such that
θ =α+β+γ
Then
Log(zw) = log(rs) + iθ =
= log(r) + log(s) + i(α + β + γ) =
= log(r) + iα + log(s) + iβ + iγ =
= Log(w) + Log(z) + iγ
In particular,
The condition is satisfied whenever Re(w) > 0 and Re(z) > 0, hence in the real
positive half plane H.
2. For the principal branch of the logarithm, Log, we have the following Taylor ex-
pansion
∞
X (−1)n−1
∀z ∈ D1 (1) : Log(z) = (z − 1)n
n=1
n
To see this, we differentiate both sides: on the left we have z1 , while on the right
we have ∞ ∞
X
n−1 n−1
X 1 1
(−1) (z − 1) = (1 − z)n = =
n=1 n=0
1 − (1 − z) z
(−1)n−1
Hence, log and ∞ (z − 1)n only differ by a constant. Looking at z = 1
P
n=1 n
we obtain that the constant is 0 and with it the result.
3. We have
3.6. COMPLEX LOGARITHM 141
and consequently
θ = Arg(z) + α
Note that this definition depends on the choice of the branch of the logarithm logΩ :
if we choose another branch of the logarithm, call it ℓ, as ℓ = logΩ +2πik for some
k ∈ Z, then
[z α ]ℓ = exp α logΩ (z) + 2πik = z α e2πikα
If we set Ω = C− , if we choose the principal branch of the logarithm with Log(1) =
0 and α = m1 for m ∈ N, then
1 1
z m = e m Log(z)
142 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
satisfies
m m
1 1 1
Y
zm = e m Log(z) = em m Log(z) = eLog(z) = z
j=1
Example 3.23. Let Log be the principal branch of the logarithm on C− , then
1 1
z 2 = e 2 Log(z)
1
Note that for z ∈ R+ , the value of z 2 is the usual positive square root.
then
1 1 1 θ+ikπ 1 θ 1 θ
h 1i
z = exp
2 logC− ,k = r 2 ei 2 = r 2 ei 2 eikπ = r 2 ei 2 (−1)k = (−1)k z 2
2
For the choice logC− ,1 (z) = log(r) + i(θ + 2π) we have (always with r > 0 and Arg(z) ∈
(π, 3π))
3.6. COMPLEX LOGARITHM 143
f′
Proof. Exercise. Define g a primitive of f
Proof. Let
1 1
h(z) = exp log(f ) = exp g(z)
2 2
from Theorem 3.17, then
h2 = exp g(z) = f (z)
Before we move to conformal maps in the next section, we mention that there are
various ways to look at simply connected regions. This is taken up in the book in the
Appendix B [SS10].
144 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
We have seen that if Ω is simply connected (i.e. such that any two curves in Ω with
same endpoints are homotopic), then for all closed curve γ in Ω and for all f ∈ H (Ω)
we have Z
f (z)dz = 0
γ
Clearly, we have with Cauchy’s Theorem 2.5 or the Homotopy Theorem 3.14 that
The proof of the direction: Ω bounded and simply connected ⇒ C \ Ω connected, uses
the notion of winding numbers, which we are going to briefly discuss as next, since it
also leads to the natural generalisation of the Residue Theorem 3.6.
Remark 3.17. In the above Theorem 3.19, the assumption that Ω is bounded in C is
important, since the infinite strip is simply connected and unbounded, its complement
has 2 components. However, if the complement is taken in Ĉ = C∪{∞}, the conclusion
holds if Ω is bounded or not.
3.7. WINDING NUMBERS 145
We have seen that for f ∈ H (Ω), for some Ω ⊆ C simply connected, if γ1 , γ2 are two
closed curves such that γ1 ∼Ω γ2 , then by Theorem 3.14
Z Z
f (z)dz = f (z)dz
γ1 γ2
with ∀t ∈ [0, 2π] : γ(t) = z0 + r(t)eiθ(t) in Ω ⊆ C for Ω open and for some functions r, θ
of class C 1 such that ∀t ∈ [0, 2π] : r(t) > 0 and r(0) = r(2π), θ(0) = θ(2π)
The same proof we gave for the Residue Formula 3.6 for a circle works also here for
Z X
f (z)dz = 2πi Resw (f )
γ w∈(Sf ∩int(γ))
The Homotopy Theorem 3.14 gives the following first generalisation of the Residue
Theorem 3.6.
Proof. (i) This is a special case of the Homotopy Theorem 3.14, since f ∈ H (V )
and γ1 ∼V γ2
(ii) Follows from the previous point and the Residue Formula 3.6.
To look at more general curves, we first introduce the winding number of a curve.
2. On the other hand, if γ(t) = z0 + reit for t ∈ [0, 2π], but we are looking at a point
z0 ̸= z1 ∈ C
Z
1 dz X 1 0 , z1 ∈/ int(γ)
wγ (z1 ) = = Resz̃ =
2πi γ z − z1 z − z1 1 , z1 ∈ int(γ)
z̃∈Sf ∩int(γ)
3.7. WINDING NUMBERS 147
1 dz
R
So at least, when γ is a circle, then the integral 2πi γ z−z1
indeed tells us whether
γ wraps around z1 or not.
“To get an intuition”: For a general smooth γ : [a, b] → C with γ(a) = γ(b) and z0
inside the path, the following imprecise and really not completely correct argument
might give an insight as to why it is called winding number.
Z b
γ ′ (t)
Z
dz
wγ (z0 ) = = dt
γ z − z0 a γ(t) − z0
But of course this is not correct, because γ(t) − z0 is complex valued and if γ wraps
around a point z0 , then we cannot define an analytic branch of log γ(t)−z0 on C\{z0 }
If we think of Log(z) = log |z| + i Arg(z) and recall that the difficulty in defining the
logarithm comes from choosing the correct value of the Arg(z), we can look at
Z
1
dz = log γ(b) − z0 − log γ(a) − z0 =
γ z − z0
= log γ(b) − z0 + i Arg γ(b) − z0 − log γ(a) − z0 + i Arg γ(a) − z0 =
= i Arg γ(b) − z0 − Arg γ(a) − z0
The ambiguity in defining Arg γ(t) − z0 for t = a, t = b must be an integral multiple
of 2π and this integer counts the number of times γ wraps around z0
148 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
We have indeed the following Proposition, which shows that ωγ (z) is always an inte-
ger.
wγ : Ω → C
Z
1 du
z 7→
2πi γ u − z
by the latter result. Hence, H is constant by Corollary 1.2 and so H(t) = γ(t) −
−G(t)
z e = c for some c ∈ C. We have that ∀t ∈ [a, b] : γ(t) − z = ceG(t) and
From this it follows that eG(b) = 1 and so G(b) ∈ 2πiZ (c ̸= 0, since γ(t) ̸= z and
e−G(t) ̸= 0).
Since G(b) = 2πiwγ (z), this shows that wγ (z) is integer valued: being
Z b
1 γ ′ (s)
wγ (z) = ds
2πi a γ(s) − z
3.7. WINDING NUMBERS 149
Finally, if M := maxt∈[a,b] γ(t) (it exists, since [a, b] is compact and γ piecewise of class
C 0 ) and |z| > M , then
Z
1 dw 1 L(γ)
wγ (z) = ≤
2π γ w − z 2π |z| − M
we then have
1 L(γ) |z|→∞
wγ (z) ≤ −−−−→ 0
2π |z| − M
Hence |wγ (z)| < 1 once |z| is large enough, but being an integer means that wγ (z) = 0,
if |z| is large enough.
We can now give the general Residue Formula.
for some a−j (z0 ) ∈ C and with N (z0 ) = − ordz0 (f ) (since f has a pole at z0 , the order
of that point is negative).
Case 1: Let Sf be finite, then
X
f˜ := f − Pzf0
z0 ∈Sf
Hence
Z Z X X Z
f dz = Pzf0 (z) dz = Pzf0 (z)dz
γ γ z0 ∈Sf z0 ∈Sf γ
dz 1 −1 1
R
Recall: γ (z−z0 )j
= 0 if j ̸= 1, since (z−z0 )j
has primitive (z−z0 )j−1 j−1
and γ is
closed.
So
Z Z X a−1 (z0 )
f (z)dz = dz =
γ γ z0 ∈Sf
z − z0
X Z a−1 (z0 )
= dz =
z0 ∈Sf γ z − z0
Z
X dz
= 2πia−1 (z0 ) =
z0 ∈Sf γ z − z0
| {z }
=wγ (z0 )
X
= 2πia−1 (z0 ) wγ (z0 ) =
z0 ∈Sf
X
= 2πi Resz0 (f )wγ (z0 )
z0 ∈Sf
Case 2: Sf is infinite. Pick R > 0 such that wγ (z) = 0 if |z|≥ R and γ is homo-
topic (so that for the homotopy in question H it hold that im H [a, b]×[0, 1] ⊆
DR (0)) to the constant cure in Ω ∩ DR (0) (since Ω is simply connected γ ∼Ω σ
(constant curve), which only involves a bounded set). Then Sf ∩ DR (0) is finite,
since Sf is a discrete set.
Let
X
f˜ := f |Ω∩DR (0) − Pzf0 ∈ H Ω ∩ DR (0)
Ω∩DR (0)
z0 ∈Sf
|z0 |<R
Hence
Z X Z
f dz = Pzf0 dz =
γ z0 ∈Sf γ
|z0 |<R
X
= 2πi Resz0 (f )wγ (z0 ) =
z0 ∈Sf
|z0 |<R
X
= 2πi Resz0 (f )wγ (z0 )
z0 ∈Sf
Corollary 3.3. Let Ω be open and simply connected, f ∈ H (Ω) and γ a closed
cure in Ω, then
Z
1 f (w)
∀z ∈ Ω \ im(γ) : dw = f (z)wγ (z)
2πi w−z
Proof. This is the generalised Residue Theorem 3.20 applied to the function fw−z
(w)
=
g(w), which is meromorphic in Ω and has a simple pole at w = z and residue f (z)
Motivating questions:
1. Given two open sets U, V ⊆ C, when does there exists a holomorphic bijection
between them, i.e. when is there a bijective f ∈ V U such that f ∈ H (U, V )?
We are going to see that the inverse map f −1 ∈ U V is automatically also holo-
morphic (compare open sets using holomorphic functions).
2. We will then prove Schwarz’s Lemma, which says any f ∈ RD such that f (0) = 0
must satisfy
Note that there can be no holomorphic bijection f ∈ DC between C and D, since in that
case f would be bounded and entire; hence by Liouville’s Theorem 2.8 f is constant.
3.9. CONFORMAL MAPS AND THE RIEMANN MAPPING THEOREM 153
Note that for Ω to be connected is also a necessary condition, since D is connected. The
same is true for simply connected, since if f ∈ V U is a holomorphic bijection with U
simply connected, then so is V
Remark 3.21. Note that there is a small difference in the definition of conformal
compared to the book [SS10]: in the book f is taken to be bijective.
Also, from now on, unless otherwise specified, we are going to assume that U, V ⊆ C
and that they are open in C
∀z ∈ U : f ′ (z) ̸= 0
Proof. Suppose f is injective and holomorphic, but on the contrary ∃z0 ∈ U : f ′ (z0 ) = 0
We want to show that f cannot be injective. Let h : U → V, z 7→ h(z) := f (z) − f (z0 ),
this implies that h (z0 ) = 0 and h′ (z0 ) = 0
Therefore, we can assume that k < ∞, by Theorem 2.7 we have that ∃r > 0 such that
for all z ∈ Dr (z0 )
154 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
f (k) (z0 )
f (z) − f (z0 ) = (z − z0 )k + G(z) (z − z0 )k+1 =
k!
= a (z − z0 )k + G(z) (z − z0 )k+1
k
where f k!
(z0 )
=: a ̸= 0 and z ∈ Dr (z0 ). Since the zeroes of f ′ are isolated, we can also
choose r > 0 such that f ′ (z) ̸= 0 for z ∈ Ḋr (z0 )
The idea is to use Rouché’s Theorem 3.11 to show that for any w ∈ C
Since f ′ (z) ̸= 0 for z ∈ Ḋr (z0 ), we have that gw′ (z) = f ′ (z) ̸= 0 for z ∈ Ḋr (z0 )
Hence, each zero has order 1 and they are distinct, but that means that there exist k
distinct points z1 , z2 , ..., zk such that f (zi ) = f (z0 ) + w, i.e. f is not injective.
To show that in some neighbourhood of z0 the function gw (z) = f (z) − f (z0 ) − w has
k zeroes, we write the following expansion for z ∈ Dr (z0 )
We apply Rouché’s Theorem 3.11 as follows: let c := sup|z−z0 |= r2 |G(z)|, c exists since
k
G is continuous. Pick s ∈ 0, min{ 2r , 1} and assume that |w| < |a| 2s . On Cs (z0 ),
that g(z) = f (z)−f (z0 )−w has the same number of zeroes in Ds (z0 ) as a (z − z0 )k −w,
3.9. CONFORMAL MAPS AND THE RIEMANN MAPPING THEOREM 155
n o
s k |a| r
for |w| < |a| 2
and s < min , ,1
c2k 2
as wanted.
Note that if w = reiθ , then the zeroes of a (z − z0 )k − w are at {zn }k−1 n=0 , where
1
1 θ+2πn
zn − z0 = wa k ei( k ) for n ∈ {0, ..., k − 1}, but then |zn − z0 | = |w|
k
|a|
< 2s < s.
Hence, all k roots of a (z − z0 )k − w are inside Ds (z0 )
f −1 (w) − f −1 (w0 ) 1 1
lim = lim =
w→w0 w − w0 z→z0 f (z)−f (z0 ) f′ (z0 )
z−z0
T : H (V ) → H (U )
ϕ 7→ ϕ ◦ f
T −1 : H (U ) → H (V )
φ 7→ φ ◦ f −1
for a, b ∈ C and ϕ1 , ϕ2 ∈ H (V )
Example 3.25 (The disc and the upper half plane). Let H := {z ∈ C : Im(z) > 0}
be the upper half plane and D := D1 (0) = {z ∈ C : |z| < 1} be the unit disc. Then the
map
f :H→D
z−i
z 7→
z+i
is a conformal equivalence, so
1+w
f −1 (w) = i
1−w
This example shows that the property that a set is bounded is not preserved under
conform equivalence.
Proof of the Example 3.25. First note that for any z ∈ H it holds:
z−i
|f (z)| = <1
z+i
since the distance from z to i is shorter than the distance from z to −i, which is in the
lower half plane.
3.9. CONFORMAL MAPS AND THE RIEMANN MAPPING THEOREM 157
1+w
− i 1+w
i 1−w 1−w
im g(w) = =
2i
1 1 + w 1 + w̄ 1 (1 − w̄)(1 + w) + (1 − w)(1 + w̄)
= + =
2 1 − w 1 − w̄ 2 |1 − w|2
1 − |w|2
= > 0 , since |w| < 1
|1 + w|2
Hence g indeed goes from D to H. Finally a direct calculation verifies that f (g(w)) = w
and (g ◦ f )(z) = z and so g = f −1
Note that the map f from Example 3.25 takes the real line to the boundary of the disc
with f (0) = −1, f (1) = −i and f (”∞”) = 1
f :U →H
z 7→ z 2
g:H→U
1 1
z 7→ z = exp
2 Log(z)
2
is its inverse.
158 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
• (f is injective) Let z12 = z22 , then z1 = ±z2 and only one of z2 , −z2 can be in U .
Since z1 , z2 ∈ U we have that z1 = z2
In general, let n ∈ N∗ and let the sector Sn = {z ∈ C : Arg(z) ∈ 0, πn }, then the map
f : Sn → H
z 7→ z n
with inverse
f −1 : H → Sn
1 1
w 7→ w = exp
n Log(w)
n
f : H → C−
z 7→ −z 2
f˜ : H → C+
z 7→ z 2
maps H to the slit plane cut of the positive reals C+ := C \ [0, +∞).
3.9. CONFORMAL MAPS AND THE RIEMANN MAPPING THEOREM 159
In particular, for V = C+
Riemann’s Theorem 3.21 says that any simply connected domain U , which is a proper
subset of C, i.e. U ̸= ∅ and U ̸= C, is conformally equivalent to D.
Ω ∼c C
Corollary 3.5. Any two proper simply connected open subsets of C are conformally
equivalent.
The proof is though not constructive, as we will see. In general, it is not easy to find
an explicit map. During the rest of the course we will prove this Theorem. The strategy
of he proof is as follows:
f1 : Ω → D
f2 : Ω → D
then
f2 ◦ f1−1 : D → D
is an automorphism of D
2. If ∅ =
̸ Ω ̸= C, we will show that there is a conformal map f : Ω → D with
f (z0 ) = 0. Hence Ω is conformally equivalent to an open subset of D, hence
Ω ∼c f (Ω) ⊆ D
3. The second step shows that the set F = f ∈ DΩ : f is conformal and f (z0 ) = 0 ̸=
∅. We will see that s := supf ∈F f ′ (z0 ) exists and we will show that ∃f ∈ F
such that f ′ (z0 ) is maximal, i.e. the supremum s is taken. This f has “maximal
expansion speed”.
3.9. CONFORMAL MAPS AND THE RIEMANN MAPPING THEOREM 161
f (0) = eiθ α
f ′ (0) = eiθ |α|2 − 1
Remark 3.25. 1. Note that an immediate Corollary of Theorem 3.22 is that the
only automorphisms of D that fix 0 are rotations, since
f (0) = eiθ α = 0 =⇒ α = 0
Proof of the uniqueness of F in Theorem 3.21. If f1 , f2 are two such maps with f1 (z0 ) =
f2 (z0 ) = 0 and f1′ (z0 ) , f2′ (z0 ) > 0, then g := f2 ◦ f1−1 : D → D is an automorphism of
D. Hence, by Theorem 3.22
α−z
g(z) = eiθ
1 − ᾱz
for some θ ∈ R and α ∈ D. Since f1 (z0 ) = 0, f2 (z0 ) = 0 we have g(0) = f2 ◦ f1−1 (0) =
Hence
f2′ (z0 )
= −eiθ > 0
f1′ (z0 )
since fk′ (z0 ) > 0 for k ∈ {1, 2}. It follows that −eiθ ∈ R>0 and hence that θ = π + 2πk
as eiθ = −1. Also, α = 0, so we can conclude that
g(z) = z ⇒ f1 = f2
Lemma 3.4 (Schwarz). [SS10, Lemma VIII.2.1] Let f ∈ H (D, D) with f (0) = 0.
Then
(iii) f ′ (0) ≤ 1 and equality holds if and only if f is a rotation, i.e. ∃θ ∈ R : f (z) =
eiθ z
This holds for all z ∈ Dr (0). By the Maximum Modulus Principle 3.13, we have
1
∀z ∈ Dr (0) : g(z) ≤
r
(the holomorphic function g cannot attain a maximum in Dr (0))
This is true for all z ∈ D such that |z| < r < 1, then by letting r → 1 it follows
that
|g(z)| ≤ 1
and hence
∀z ∈ D : f (z) ≤ |z|
(ii) We proved that (i) gives supz∈D g(z) ≤ 1, but the assumption f (z0 ) = |z0 |
for some z0 ∈ D \ {0} implies that g has a local maximum at z0 ∈ D. By the
Maximum Modulus Principle 3.13 this can only happen if g is constant, hence
∃c ∈ C∀z ∈ D : f (z) = zg(z) = cz
Since f (z0 ) = |z0 | for that z0 ∈ D \ {0}, it follows that |c| = 1. Hence, c = ei θ
for some θ ∈ R and f (z) = eiθ z
(iii)
f (z) f (z) − f (0)
g(0) = lim = lim = f ′ (0)
z→0 z z→0 z−0
so f ′ (0) = g(0) ≤ 1. If f ′ (0) = 1, then again 0 is a local maximum of g and
we conclude as in (ii) that f (z) = eiθ z for some θ ∈ R, namely a rotation.
1 − |α|2 z = 1 − |α|2 w
z=w
Hence φα is a conformal map φα : D → D
164 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
3. φα (D) ⊆ D: This point might look superfluous, but checking this condition guar-
antees that the map is well-defined on its domain of definition, as any argument
will have an image in the codomain. So if |z| = 1, then z = eiθ and
α − eiθ α − eiθ
−iθ w
iθ
= e−iθ
φα e = iθ −iθ =e −iθ
e (e − ᾱ) e − ᾱ −w̄
w
φα eiθ = e−iθ
=1
−w̄
By the Maximum Modulus Principle 3.13, we have that ∀z ∈ D : φα (z) < 1 (not
being φα (z) a constant map, it cannot have a local maximum inside of D).
4. We have that
α−z
α− 1−ᾱz α − |α|2 z − α + z (1 − |α|2 ) z
(φα ◦ φα ) (z) = α−z
= = =z
1 − ᾱ 1−ᾱz
1 − ᾱz − |α|2 + ᾱz 1 − |α|2
Hence
α−z
(R ◦ φα ) (z) = eiθ
1 − ᾱz
is on automorphism of D
∀z ∈ D : g(z) ≤ |z|
Since g −1 (0) = 0, we can also apply the Schwarz Lemma 3.4 to g −1 and get
∀w ∈ D : g −1 (w) ≤ |w|
Combined with g(z) ≤ |z| we get that |g(z)| = |z|. Once again by Schwarz
Lemma 3.4 (b), g(z) = eiθ z is a rotation with some θ ∈ R. Hence eiθ z =
(f ◦ φα )(z) = g(z)
Remark 3.28. It is of interest that one can see that AutH (H) can be described via the
action of SL2 (R) ↷ H via fractional linear transformations.
Remark 3.29. Being these maps invariant under re-scaling by a real factor, we dis-
cover that these (conformal) automorphism are in fact represented in GL2 (R) quotient
by R∗ , namely the projective general linear group PGL2 (R), but only considering
those elements with positive determinant, hence a subgroup of it. These elements form
PSL2 (R), the projective special linear group, this being the quotient of SL2 (R) and
{±1}, and can be seen as a subgroup of the former (of the orientation-preserving trans-
formations).
Over R these two groups are different, in particular the second one being a strict sub-
group of the first one, but for instance over C surprisingly
PGL2 (C) ∼
= PSL2 (C)
This is going to have a greater relevance later in geometry, in particular in the context
of Möbius Transformations.
166 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
γ : Aut(D) → Aut(H)
φ 7→ γ(φ) = F −1 ◦ φ ◦ F
γ −1 : Aut(H) → Aut(D)
β 7→ γ −1 (β) = F ◦ β ◦ F −1
Proposition 3.9. [SS10, Step 1 in Section VIII.3.3, p.228] Let Ω ⊂ C such that
∅≠ Ω ̸= C, open and simply connected. Then there exists a conformal map f ∈ DΩ
such that 0 ∈ f (Ω), i.e. Ω is conformally equivalent to a subset of D, which contains
the origin.
Proof. Let δ > 0, such that D2δ (z0 ) ⊆ Ω and let f ∈ F. The Cauchy Integral Formula
2.6 gives Z
′ 1 f (z)
f (z0 ) = dz
2πi Cδ (z0 ) (z − z0 )2
Hence, using the standard estimate one finds that
1 f (z) 1
f ′ (z0 ) ≤ 2πδ max 2
≤
2π z∈Cδ δ δ
since f (z) ≤ 1 for all z ∈ Ω. Hence f ′ (z0 ) is bounded by 1δ for an arbitrary f ∈ F,
thus there exists an upper bound and by consequence of it also the supremum.
The next Proposition is key and states that the supremum
s = sup f ′ (z0 )
f ∈F
is taken.
The proof of this Proposition 3.10 uses a compactness argument which we will come back
to, but we first see why this is key, in the sense that it gives the conformal equivalence
that we are looking for, between Ω and D (see Step 3 in Section 3.3 in [SS10], p.231).
Remark 3.30. The step 2 shows that Ω is conformally equivalent to an open subset of
D, which contains 0
Why are we looking for an extremal function that realises the extremal value s =
supf ∈F f ′ (z0 ) ?
3.9. CONFORMAL MAPS AND THE RIEMANN MAPPING THEOREM 169
We can assure without loss of generality that Ω is an open subset of D that contains 0,
so we can assume that z0 = 0
We want to choose a function in F with “maximal expansion”, but what does “expand-
ing” mean? Consider
f (0) = 0 =⇒ f (z) ∼ f ′ (0)z
for z near 0, so if f ′ (0) > 1, we say that f is expanding, since the distances between
nearby points are expanding
To do this we are going to use φα and the square root map. Let
φ = φα : D → D
α−z
z 7→
1 − ᾱz
be the automorphism of D, with φα (0) = α and φα (α) = 0
Since 0 ∈
/ (φ ◦ f )(Ω), and Ω is simply connected, a logarithm and a square root of of
φ ◦ f exist, i.e.
∃f˜ ∈ H (Ω)∀z ∈ Ω : f˜2 (z) = (φ ◦ f )(z)
′
α ◦f )
One can simply take g̃ as primitive of (φ
(φα ◦f )
, so that exp g̃(z) = (φα ◦ f )(z)
Note that f˜ is also injective: if f˜(z) = f˜(w), then (φ ◦ f )(z) = (φ ◦ f )(w). Since φ ◦ f
is conformal and f, φ are injective, we have that z = w
φβ : D → D
β−z
z 7→
1 − β̄z
with φβ (β) = 0. Finally, let g(z) := φβ ◦ f˜ ∈ DΩ . Then g (z0 ) = 0 and g ∈ H (Ω, D),
since f˜ ∈ H (Ω, D) and φβ ∈ AutH (D)
Moreover, g is injective, since φβ and f˜ are injective, also g ∈ F by definition.
Claim. g ′ (z0 ) > f ′ (z0 )
This will give the contradiction that we are looking for.
Proof of the Claim. Recall: we first looked at φα ◦ f as
f φα
φα ◦ f : Ω →
− D −→ D
√
Then we took the · function, call it h
h : (φα ◦ f ) (Ω) → D
1
w 7→ exp w
2
3.9. CONFORMAL MAPS AND THE RIEMANN MAPPING THEOREM 171
f˜ := h ◦ φα ◦ f : Ω → D
so that f˜2 = φα ◦ f
Then we composed it with φβ , to get g ∈ DΩ such that
g := φβ ◦ f˜ = φβ ◦ h ◦ φα ◦ f
| {z }
f˜
2
⇒ φ−1
β ◦g = φα ◦ f
−1
2
⇒ φ−1
α ◦ φβ ◦ g =f
−1
f = φ−1
α ◦ s ◦ φβ ◦g = Φ ◦ g
| {z }
=:Φ
Note that Φ is not injective. Now Φ ∈ H (D, D) by composition and since φ−1
β (0) = β
−1
Φ(0) = φ−1 (0) = φ−1 β2
α ◦ s ◦ φβ α
2
˜ 2 ˜
But recall that f (z0 ) = β, it follows that β = f (z0 ) = (φα ◦ f ) (z0 ). Hence
Φ(0) = φ−1
α ◦ φα ◦ f (z0 ) = f (z0 ) = 0
Hence, we can apply Schwarz Lemma 3.4 (iii) to get Φ′ (0) < 1 (note that |Φ′ (0)| = ̸ 1,
iθ
since if it were so, then Φ(z) = e z for some θ ∈ R and it would mean that Φ is
injective, but Φ cannot be such, since the squaring function is not injective). Using the
chain rule applied to f = Φ ◦ g we have
∗
Recalling the definition of supremum, we take a sequence (fn )n∈N∗ ∈ (F)N with
n→∞
fn′ (z0 ) −−−→ s. We want to show that this sequence has a limit f in F
Note that the proof will not be constructive and will only guarantee the existence
of a limit f ∈ F
So we are looking for an analogue of this for F. This is provided in the following
by Montel’s Theorem.
∗
3. In application to Riemann’s Theorem 3.21 we have a sequence (fn )n∈N∗ ∈ F N ,
so that ∀z ∈ Ω∀n ∈ N∗ : |fn (z)| ≤ 1 (not only compact sets, as D is bounded).
∗
Hence we can apply Montel’s Theorem 3.24 to find a sequence (fnk )k∈N∗ ∈ F N
which converges uniformly on compact sets and this will give the limk→∞ fnk = f
3.9. CONFORMAL MAPS AND THE RIEMANN MAPPING THEOREM 173
(holomorphic) that we are looking for, provided that we can show that f ∈ F
We now need to argue for the injectivity of such f , to this purpose we use the
following result:
∗
Proposition 3.12. Let (fn )n∈N∗ ∈ F N be a sequence in F and suppose that
fn → f uniformly on any compact set K ⊆ Ω. Then, either f is constant or
f ∈ F. Moreover, for any z ∈ Ω
Proof. We will suppose that f is not injective and show that then f is constant.
Suppose that for z1 ̸= z2 ∈ Ω : f (z1 ) = f (z2 ). If f is not constant, since the
zeroes of holomorphic functions are isolated, we can find a disc Dδ (z2 ) ⊆ Ω, so
that f (z) − f (z2 ) ̸= 0 in Ḋδ (z2 ). Hence, in particular
∀z ∈ C δ (z2 ) : f (z) − f (z2 ) ̸= 0
2
We apply the Argument Principle 3.10 to the function f (z) − f (z1 ) which has a
zero, namely z2 , in D δ (z2 ) to get
2
f ′ (z)
Z
1
dz ≥ 1
2πi C δ (z2 ) f (z) − f (z1 )
2
174 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
since the integrals on the right counts the zeroes of the holomorphic function
fn (z) − fn (z1 ) in C δ (z2 ), so none by the injectivity of fn , but the integer on the
2
left is ≥ 1, which is a contradiction. Hence, f must be a constant.
∗
Remark 3.31. Finally note that in the case (fn )n∈N∗ ∈ F N with limn→∞ fn′ (z0 ) = s
and limn→∞ fn′ (z0 ) = f (z0 ) we have that
1. ∀n ∈ N∗ : fn′ (z0 ) ̸= 0, using Proposition 3.8 and that fn ’s are conformal (hence
∀z ∈ Ω : fn′ (z) ̸= 0).
such that limn→∞ fn′ (z0 ) = s > 0. Hence f ′ (z0 ) ̸= 0 and f is not constant.
1. The first part is about the complex behaviour of sequences of holomorphic func-
tions, which says
This is a complex behaviour in the sense that it is not true for sequences of real
functions. For example, fn (x) = sin(nx) on (0, 1) is uniformly bounded on com-
pact sets, but not equicontinuous.
The family (fn )n∈N∗ ∈ R[0,1] on [0, 1], given by fn (x) = xn , is not equicontinuous,
even though each
fn : [0, 1] → R
x 7→ xn
is uniformly continuous on [0, 1]. The family (fn )n∈N∗ is not equicontinuous. For
n→∞
example take any w ∈ (0, 1), then fn (1) − fn (w) −−−→ 1
2. The second part is known as Arzelà-Ascoli Theorem [EW22], which says that any
family F of functions, which is uniformly bounded and equicontinuous on com-
pact subsets of Ω, has a subsequence which converges uniformly on every compact
subset of Ω (the limit need not be in F).
Rm ).
Assuming the Arzelà-Ascoli Theorem 3.25, the proof of Montel’s Theorem 3.24 reduces
to proving that every sequence of holomorphic functions, which is uniformly bounded
176 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
This uses Cauchy’s Integral Formula 2.6 together with the following Lemma:
Lemma 3.7. [SS10, Lemma VIII.3.4] Let Ω ⊆ C be open. Then ∃ compact sets
{Kl }l∈N∗ such that
1. ∀l ∈ N∗ : Kl ⊆ int(Kl+1 )
Proof. Exercise.
Now, assuming the Arzelà-Ascoli Theorem 3.25 and Lemma 3.7 we can give the proof
of Montel’s Theorem 3.24.
Proof of Theorem 3.24. Let (fn )n∈N∗ be a sequence of holomorphic functions, which are
uniformly bounded on compact sets. We want to show that (fn )n∈N∗ is equicontinuous.
1. Let K ⊆ Ω be compact, let r > 0 such that D3r (z) ⊆ Ω for z ∈ K (we can
choose r so that 3r is less than the distance from K to the boundary of Ω, i.e.
3r < d(K, ∂Ω)).
1 1 |z − w| |z − w|
− = ≤
ξ−z ξ−w |ξ − w||ξ − z| 2r · r
1 1 |z − w|
fn (z) − fn (w) ≤ 2π(2r)|z − w| 2 Mk ≤ M
2π 2r r
3.9. CONFORMAL MAPS AND THE RIEMANN MAPPING THEOREM 177
since ∀ξ ∈ C2r (w)∀n ∈ N∗ : fn (ξ) ≤ M , where M in the uniform bound for all
fn ∈ F in D2δ (w)
So, for any ε > 0 we get
as soon as n εr o
|z − w| ≤ min r,
M
2. To extract a subsequence which converges uniformly on all compact sets we use
a standard trick, called the “diagonal argument”.
Let {Kl }l∈N∗ be the sequence of compact sets given by the last Lemma 3.7.
By the first step and the Arzelà-Ascoli Theorem 3.25, there is a subsequence of
(fn )n∈N∗ converging uniformly on K1 , say
(fn )n∈L1
Now let
n1 := min {n ∈ N∗ : n ∈ L1 }
n2 := min {n ∈ N∗ : n ∈ L2 \ {n1 }} ∈ L2 ⊂ L1
n3 := min {n ∈ N∗ : n ∈ L3 \ {n1 , n2 }} ∈ L3 ⊂ L2 ⊂ L1
..
.
Note that L := {n1 , n2 , ...} has the property that L \ Lk is finite for each k
For each k ∈ N∗ , (fn )n∈L = (fnj )j∈N is up to finitely many terms (which has no
effect on convergence) a subsequence of (fn )n∈Lk
Hence (fnj )j∈N converges uniformly on every Kk , since any compact set K is
contained in Kk for some k ∈ N∗
178 CHAPTER 3. MEROMORPHIC FUNCTIONS AND RESIDUE FORMULA
Appendix A
1. We have seen
P∞ 1
Proposition A.1. [SS10, Proposition VI.2.1] The series ζ(s) = n=1 ns con-
verges absolutely and uniformly on every half plane
∀δ > 0 : Uδ := {s ∈ C : Re(s) ≥ 1 + δ}
3. RFor a function f ∈ CR , which is Riemann integrable on every [a, b] and for which
∞
−∞
f (t) dt converges, its Fourier transform is defined as
Z ∞
fˆ(ξ) := f (x)e−2πiξx dx
−∞
We also have shown that f (x) = e−πx has fˆ(ξ) = e−πξ . What we have not seen, but
2 2
can be proved, are the following results about Fourier Transform, which can be found
179
180APPENDIX A. THE ANALYTIC CONTINUATION OF THE RIEMANN ZETA FUNCTION
in Chapter IV in [SS10].
For a > 0, denote by Fa the class of all functions f that satisfy the following two
conditions:
1. f is holomorphic in the horizontal strip Sa := z ∈ C : Im(z) < a
2
Corollary A.1. The Poisson Summation Formula applied to ft (x) = e−πtx , for
t ∈ R>0 gives
∞ ∞
2 −1 −πn2
X X
e−πtn = t2e t
n=−∞ n=−∞
Hence
X
−πtn2 −1
X −πn2 1 1
ϑ(t) := e =t 2 e t =√ ϑ
n∈Z n∈Z
t t
181
So
1 1
ϑ(t) = √ ϑ
t t
and note that ϑ(t) = θ(it)
We can now use this transformation of the θ-function to give analytic continuation and
functional equation of ζ(s), namely
Theorem A.3. [SS10, Theorem VI.2.3] Let for Re(s) > 1 and
s
−s/2
Λ(s) := π Γ g(s)
2
Then Λ(s) has a meromorphic continuation to all of s-plane, with simple poles at
s = 0, 1 and satisfies the functional equation
Λ(1 − s) = Λ(s)
(−1)n
Res−n (Γ) =
n!
Proof of Theorem A.3. Idea: to relate Λ(s) and ϑ(t) = θ(it) via an integral transform
and use the transformation property of ϑ(t) = √1t ϑ 1t inside the integral to analytically
continue Λ(s).
We start by collecting growth and decay property of ϑ(t), for example
−1 t→0
ϑ(t) ≤ Ct 2 −−→ 0
(follows from the functional equation) and v(t) − 1 ≤ Ce−πt for some C > 0 and for
all t ≥ 1. Since for t ≥ 1, we get
2
X X
2 e−πn t ≤ 2 e−πnt ≤ Ce−πt
n∈N∗ n∈N∗
182APPENDIX A. THE ANALYTIC CONTINUATION OF THE RIEMANN ZETA FUNCTION
The relation between Λ and ϑ is now given by the fact that for any s ∈ C with Re(s) > 1
we have Z ∞
−s
s s dt
Λ(s) = π Γ2 ζ(s) = ϑ(t) − 1 t 2
2 0 t
This is based on the simple observations that
R∞ 2 s −s −s
1. 0 e−πn t t 2 dtt = (πn2 ) 2 Γ(s) = π 2 Γ(s)n−s
2. ϑ(t) − 1 = 2 ∞ −πn2 t
P
n=1 e
R
Pthe estimates on ϑ(t) as t → 0 and as t → ∞, one can justify the change of
Using
and to get for Re(s) > 1
Z ∞ ∞ Z ∞ ∞
1 s dt X −πn2 t s dt −s
X 1
ϑ(t) − 1 t 2 = e t
2 = π Γ(s)
2 = Λ(s)
2 0 t n=1 0 t n=1
ns
Z ∞
1 s dt
Λ(s) = ϑ(t) − 1 t 2
2 0 t
Now, we will see that we can make sense of the right hand side for s ∈ C. Now
1
1 1 −1
Z Z
1 s dt 1 −1 s dt
ϑ(t) − 1 t 2 = t2 ϑ − 1 + t 2 − 1 t2 =
2 0 t 2 0 t t
Z 1 Z 1
1 1 s dt
Z
1 1 s−1 dt 1 s−1 dt
= ϑ −1 t 2 + t 2 − t2 =
2 0 t t 2 0 t 2 0 t
Z 1 " s−1 #1 s 1
1 1 s−1 dt 1 t 2 1 t2
= ϑ −1 t 2 + − =
2 0 t t 2 s−1
2
2 2s 0
0
1 1
Z
1 s−1 dt 1 1
= ϑ −1 t 2 + −
2 0 t t s−1 s
1 du
Now we make the change of variables u = t
with u
= − dtt we get
Z 1 Z ∞
1 1 s−1 dt 1−s du
ϑ −1 t 2 = ϑ(u) − 1 u 2
2 0 t t 0 u
183
1 1
R∞ s 1−s
dt
Both s−1 − s and 1 ϑ(t) − 1 t + t 2 2
t
are invariant under s 7→ 1 − s, hence
Λ(s) = Λ(1 − s)
−s
Λ(s) = π 2 Γ 2s ζ(s) has analytic continuation to all s ∈ C except for the simple poles
at s = 0 and s = 1
Since Γ 2s has poles at 2s = 0, −1, −2, ...; ζ(s) does not have a pole at s = 0 and must
vanish at s = −2, −4, −6, ...; since Λ(s) does not have poles at s = −2, −4, ...
Due to the Euler product, we have that ζ(s) ̸= 0 for Re(s) > 1 and by the functional
equation, that ζ(s) ̸= 0 for Re(s) < 0
Definition A.1 (Riemann Hypothesis). If s ̸= −2, −4, ..., s ∈ C with ζ(s) = 0, then
Re(s) = 12
Then
π(x)
lim x =1
x→∞
log(x)
Other results
Part of the reason for this limit to integration is linkable to the Homotopy Theorem
3.14, as a change of variable, once composed with the path, deforms it in possibly
non-viable ways. For instance, exiting the domain of the integrand function, crossing
a singularity, etc. To name one more, also changes in the “simplicity” of the curve can
have effects.
Theorem B.1. Let Ω ⊆ C be an open subset and let γ ∈ C 1 [a, b]; Ω be a path
in Ω. If ϕ : Ω → ϕ(Ω) is biholomorphic, i.e. is a bijection that is holomorphic and
185
186 APPENDIX B. OTHER RESULTS
A useful result to accompany this Theorem is the following Lemma, which states the
form that biholomorphisms need to assume from C to C, namely Affine maps, denoted
by AutH (C)
Proof. We want to show this in two steps, first showing that such a function is a
polynomial and second that its degree is 1
Let f ∈ AutH (C). First, the series expansion of f at 0 is
∞
X
f (z) = aℓ z ℓ
n=0
and converges over all of C, since f is entire. Composing f with the inversion η : C∗ →
C, z 7→ z1 it results that for
∞
X
g := f ◦ η : C∗ → C, z 7→ aℓ z −ℓ
n=0
There are many more (conformal) automorphism between two proper subset of C with
specific characteristics, these can be obtained using the Riemann mapping Theorem
3.21.
f (z) f ′ (z)
lim = lim ′
z→a g(z) z→a g (z)
N∗
(iv) If a sequence (fn )n∈N∗ ∈ H Ḋr (z0 ) is such that fn ⇒ f , then
n→∞
∞
! ∞
X X
Resz0 fn = Resz0 (fn )
n=0 n=0
(ii) There is no “nice” formula for Resz0 (f g). However, when f and g just have poles
at z0 it is possible to compute Resz0 (f g) in a ”finitary” fashion from the Laurent
expansions of f and g as follows: assuming z0 = 0 for simplicity we have
∞
X
f (z) = ak z k
k=−m
X∞
g(z) = bℓ z ℓ
ℓ=−n
for n, m ∈ N; therefore
n−1
! m−1
!
X X
f (z)g(z) = ak z k bℓ z ℓ + h(z)
k=−m ℓ=−n
h(z)
f˜(z) =
z
for some holomorphic h in the same neighbourhood. Then
1. Draw a horizontal line to the right of each point and extend it to infinity.
2. Count the number of times the line intersects with polygon edges.
3. A point is inside the polygon if either count of intersections is odd or point lies on
an edge of the polygon. If none of the conditions are true, then point lies outside.
191
192 APPENDIX B. OTHER RESULTS
References
[SS10] E.M. Stein and R. Shakarchi. Complex Analysis. Princeton lectures in anal-
ysis. Princeton University Press, 2010. isbn: 9781400831159. url: https :
//books.google.ch/books?id=0ECHh9tjPUAC.
[CE16] R. Cavalieri and Miles E. Riemann Surfaces and Algebraic Curves: A First
Course in Hurwitz Theory. Cambridge University Press, 2016. url: https:
//www.cambridge.org/core/books/riemann-surfaces-and-algebraic-
curves/4D91A22369F62D519F43A8A0713E29D9#.
[EW22] M. Einsiedler and A. Wieser. Analysis I und II. ETH Zürich, 2022. url:
https://metaphor.ethz.ch/x/2023/fs/401- 1262- 07L/sc/Analysis-
Skript.pdf.
[Ima23] Ö. Imamoḡlu. Lecture Notes in Complex Analysis. ETH Zürich, 2023. url:
https://metaphor.ethz.ch/x/2023/hs/401-2303-00L/.
[Da 24] F. Da Lio. Analysis III. ETH Zürich, 2024. url: https://people.math.
ethz.ch/~fdalio/lecture-notes.
193