Ma 231
Ma 231
Stefan Adams
2 Visualisation of functions f : Rm → Rn 4
2.1 Scalar fields, n = 1 . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Vector fields, n > 1 . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Curves and Surfaces . . . . . . . . . . . . . . . . . . . . . . . 10
3 Line integrals 16
3.1 Integrating scalar fields . . . . . . . . . . . . . . . . . . . . . . 16
3.2 Integrating vector fields . . . . . . . . . . . . . . . . . . . . . 17
5 Surface Integrals 33
5.1 Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5.2 Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5.3 Kissing problem . . . . . . . . . . . . . . . . . . . . . . . . . . 40
8 Integration by Parts 59
10 Stokes’s theorem 72
10.1 Stokes’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . 72
10.2 Polar coordinates . . . . . . . . . . . . . . . . . . . . . . . . . 78
ii
Preface
Health Warning: These notes give the skeleton of the course and are not a
substitute for attending lectures. They are meant to make note-taking easier
so that you can concentrate on the lectures. An important part in vector
analysis are figures and pictures. These will not be contained in these notes.
For any figure which appears on the blackboard in my lectures I leave some
empty space with a reference number which coincides with the number I am
using in the lectures. You can fill the diagrams and figures by your own.
These notes grew out of hand written notes from Jochen Voß who gave
this course 2005 and 2006. I thank him very much for letting me using his
notes.
Any remarks and suggestions for improvements would help to create better
notes for the next year.
Stefan Adams
Motivation
What is Vector Analysis?
In analysis differentiation and integration were mostly considered in one di-
mension. Vector analysis generalises this to curves, surfaces and volumes in
Rn , n ∈ N. As an example consider the “normal” way to calculate a one
dimensional integral: You may find a primitive of a function f and use the
fundamental theorem of calculus, i.e. for f = F 0 we get
Z b
f (x) dx = F (b) − F (a).
a
Notation
One of the main problems in vector analysis is that there are many books
with all possible different notations. During the whole course I outline alter-
native notations in use. It is one of the objectives to acquaint you with the
different notations and symbols. Note that most of the material originated
from physics and hence many books are using notations and symbols known
by people in physics.
iii
p
Vectors: x ∈ Rn with x = (x1 , . . . , xn ) and with norm kxk = x21 + · · · + x2n .
Alternative notations are: ~x, x, x and in dimension n = 3 ~r = (x, y, z) for
the vector and |x| or r for the norm of ~x, x, x or ~r.
Properties of the norm are (i)kxk ≥ 0; (ii)kλxk = |λ|kxk for λ ∈ R; (iii)kx +
yk ≤ kxk + kyk for x, y ∈ Rn .
Functions: f : Rm → Rn with component functions f1 , . . . , fn : Rm → R.
Alternative notations are: f~ or f or f .
Partial derivatives: Let ei , 1 ≤ i ≤ n, be the canonical basis vectors in Rn
(i.e. hei , ej i for i, j+1, . . . , n). The partial derivative of a function f : Rn → R
with respect to the i-th direction at point x ∈ Rn is given by
iv
Part I: Real Vectoranalysis
1 Gradients and Directional Derivatives
In this section we ask how does a function f : Rn → R change when we move
from x ∈ Rn in direction of a vector y ∈ Rn .
Figure 1
We can reduce the problem to one dimension. We define for the given x ∈ Rn
and y ∈ Rn the function
ϕ(x) : R → R
t 7→ ϕ(x) (t) = f (x + ty).
The change of f in direction y equals the change of ϕ(x) and is thus given by
ϕ0(x) .
Definition 1.1 The directional derivative Dy f (x) of a function f : Rn →
R at the point x ∈ Rn in direction of y ∈ Rn is given by
f (x + ty) − f (x)
Dy f (x) = lim
t→0 t
if the indicted limit exists. Dy f (x) is also called the derivative along the
vector y at the point x.
1
Remark 1.2 (a) With ϕ(x) (t) = f (x + ty) the directional derivative of f
at the point x ∈ Rn in direction y ∈ Rn is given as
(b) We have defined the directional derivative via the limit as t → 0 for
the differential quotient. If one takes the limit t ↓ 0, that is t > 0 and
t → 0, one gets the directional derivative from the right hand side. The
same applies to the limit t ↑ 0 for the directional derivative from the
left.
Figure 2
2
Recall ϕ(x) (t) = f (x + ty), t ∈ R, x, y ∈ Rn . The function ϕ(x) : R → R is
the composition of the function g : R → Rn , t 7→ x + ty, and the function
f : g(R) → R (i.e. the restrition of f on g(R) ⊂ Rn ), that is
Examples for vector fields are the magnetic, the electric or the velocity (vec-
tor) field, whereas temperature and pressure are scalar fields.
Our calculation in (1.1) shows that the directional derivative Dy f at any
point x ∈ Rn is linear in y and we only need to know the gradient ∇f (x) in
order to calculate Dy f (x)
3
Remark 1.6 Recall the Cauchy-Schwarz inequality
We get
|Dy f (x)| = |h∇f (x), yi| ≤ k∇f (x)kkyk,
and for y ∈ Rn with kyk = 1 we have |Dy f (x)| ≤ k∇f (x)k. Assume that
∇f (x)
∇f (x) 6= 0 and pick y = k∇f (x)k
with kyk = 1. This gives
∇f (x)
Dy f (x) = ∇f (x), = k∇f (x)k.
k∇f (x)k
2 Visualisation of functions f : Rm → Rn
Visualisation of functions is hard and needs practice. One of the objectives
of the course is to learn basic techniques to create sketches/figures showing
basic features of a given function.
4
Figure 3:
f −1 (c) = {x ∈ D : f (x) = c}
5
Example 2.2 (a) f : R2 → R, (x, y) 7→ f (x, y) = x2 + y 2 .
Figure 4:
f (x, y, z) = 0 ⇔ r2 := x2 + y 2 = z 2 ⇔ r = |z|,
p
where r = x2 + y 2 .
Let c > 0:
√ p
f (x, y, z) = c ⇔ r = z 2 − c with z 2 ≥ c ⇔ z = ± x2 + y 2 + c.
6
Figure 5: cone
Let c < 0:
√ p
f (x, y, z) = c ⇔ r = z2 − c ⇔ z = ± x2 + y 2 − c.
7
The level set (surface) is the single-sheeted hyperboloid of revolution
√ around
the z axis, which intersects the x − y plane in the circle of radius −c.
8
Figure 8:
2 2 x
(b) f : R → R , (x, y) 7→ f (x, y) = , see figure 9 below.
y
Figure 9:
9
2 2 y
(c) f : R → R , (x, y) 7→ f (x, y) = , see figure 10 below.
x
Figure 10:
10
Example 2.4 (a)
cos t
f : [0, 2π] → R3 , t 7→ f (t) = sin t .
t
(b)
s cos t
f : [0, 1] × [0, 2π] → R3 , (s, t) 7→ f (s, t) = s sin t .
t
This mapping defines the helicoid seen in figure 12.
Note that for fixed parameter t we have lines within the surface of the helicoid
and through any point of the helicoid there is a helix going through that
point.
C ⊂ Rn , can be given as the level set of some real valued function, e.g.
consider the function f : R2 → R, (x, y) 7→ f (x, y) = x2 + y 2 . The curve C is
then the level set
C = {(x, y) ∈ R2 : x2 + y 2 = 1},
which is again the circle line in x − y plane.
For a parametric curve ϕ : R → Rn any point ϕ(t) gives the ’position’ at
’time’ t. The derivative with respect to the parameter t gives the velocity
vector ϕ0 (t) at time “time” t. Both are vectors in Rn with the following
components
ϕ(t) = (ϕ1 (t), . . . , ϕn (t)) ∈ Rn and ϕ0 (t) = (ϕ01 (t), . . . , ϕ0n (t)) ∈ Rn .
If ϕ0 (t) 6= 0, then ϕ0 (t) is a tangent vector of the curve. The tangent line
Tϕ(t) at a point ϕ(t) is given by
11
Figure 11: helix
12
Figure 12: helicoid
13
Definition 2.5 A vector x ∈ Rn is orthogonal to a parametric curve
ϕ : R → Rn at the point ϕ(t) if hx, ϕ0 (t)i = 0, i.e. if it is orthogonal to
the tangent line.
The chain rule gives for the derivative of the composition f ◦ ϕ with respect
to t at t = 0 as
where the ◦-operation on the right hand side is the matrix product which in
this case is the corresponding scalar product in Rn . With that we get
d
0= f (ϕ(t)) t=0 = h∇f (ϕ(0)), ϕ0 (0)i
dt
= h∇f (a), ϕ0 (0)i.
This implies ∇f (a) ⊥ ϕ0 (0) and ϕ0 (0) is tangent vector, i.e. it is in L(f (a)).
2
Recall the notion of (total) differentiability in the following definition. It
generalises the notion of differentiability for real-valued functions defined on
the real line. Roughly speaking, the existence of the differential quotient is
equivalent to a linear approximation (tangent line) to that function.
14
Definition 2.7 (Differentiability) Let D ⊂ Rn be a domain and f : D →
Rm , m ≥ 1, f = (f1 , . . . , fm ). We say that f is differentiable at x0 ∈ D if
the partial derivatives of f exist at x0 and if there exists a linear mapping
L : Rn → Rm with
kf (x) − f (x0 ) − L(x − x0 )k
lim = 0,
x→x0 kx − x0 k
where the linear mapping L is given by the so-called (m×n)− Jacobi matrix
at the point x0 , i.e.
∂f ∂f1 ∂f1
1
∂x1
(x0 ) ∂x 2
(x 0 ) . . . ∂x n
(x 0 )
∂f2
(x0 ) ∂f2 (x0 ) . . .
∂x ∂x
. . .
L = Df (x0 ) =
...
1 2 .
... ... ...
∂fm
... ... . . . ∂xn (x0 )
Df : D → Lin(Rn , Rm ),
where the ◦-operation on the right hand side means composition of the linear
mappings which corresponds to the matrix product of the matrices describing
the linear mappings.
15
3 Line integrals
In this section we want to take integrals along a curve C ⊂ Rn . A curve is a
one-dimensional object in Rn .
Figure 13:
R R R
Alternative notations are: γ u, C u ds and γ u ds and the integral is
sometimes also called the path integral along the path C.
16
Note that the scalar field u needs only to be defined along the curve, i.e.
u : C → R.
As
0 − sin t p
γ (t) = and kγ 0 (t)k = (− sin t)2 + cos2 t = 1,
cos t
R 2π
kγ 0 (t)k dt = 2π.
R
we have γ
1= 0
17
Figure 14:
Hence, heuristically the total amount of work shall be an integral (the sum
of all these small contributions)
Z
− F~ ·d~x.
C
R R R
Alternative notations are C
f · d~s, γ
f ·d~s or γ
f ·T̂ ds.
18
0
where f (γ(t), kγγ 0 (t)
(t)k
is the projection of f onto the tangent line, see figure
R
15 below. Hence, the tangent line intergal γ f of a vector field f is the scalar
line integral of the component of f along the tangent direction.
Figure 15:
Example 3.5 (Tangent line integral) We calculate the work done when
moving a mass m> 0 along a line/curve with parametrisation γ : [0, π] →
t
R2 , t 7→ in the gravity field f : R2 → R2 , (x, y) 7→ f (x, y) =
− cos t
0
. Here g is a constant, called the earth acceleration. Then the
−mg
0 π
work to be done moving the mass from to is given by
−1 1
Z Z π D E
0 1
− f =− , dt
γ 0 −mg sin t
Z π
= mg sin t dt = 2mg.
0
19
Figure 16:
Example 3.6 (Scalar line integral) Consider the curve C with parametri-
sation γ : [−π, π] → R3 , t 7→ γ(t) = (cos t, sin t, t). The scalar line integral of
u : R3 → R, (x, y, z) 7→ x2 + y 2 along the curve C is
Z Z Z π
2
u = (x + y ) = 2
(cos2 (t) + sin2 (t))kγ 0 (t)k dt,
C C −π
√ √
and as γ 0 (t) = (− sin t, cos t, 1) with kγ 0 (t)k = sin2 t + cos2 t + 1 = 2 we
get Z √
(x2 + y 2 ) = 2π 2.
C
20
4.1 FTC for gradient vector fields
Definition 4.1 A vector field f : D → Rn , D ⊂ Rn , is called a gradient
vector field , if there exists a differentiable scalar field Φ : D → R with
f (x) = ∇Φ(x) for all x ∈ D. (4.5)
Φ is called the potential of the vector field f if (4.5) is satisfied. Sometimes
a gradient vector field is also called a vector field of gradient type.
The most important result about gradient vector fields is the following
theorem.
21
Thus we get
Z Z b Z b
0
f= hf (γ(t)), γ (t)idt = (Φ(γ(t))0 dt
C a a
= Φ(γ(b)) − Φ(γ(a)),
In Example
3.5 we have a curve C with parametrisation γ : [0, π] → R2 , t 7→
t
γ(t) = , hence the work (tangent line intergal) can be computed
− cos t
directly with the potential
Z
− f = mg(cos π − (−1)) = 2mg.
C
22
Figure 17: Simple and not simple curve
23
Figure 18:
Figure 19:
24
and a parametrised path γ2 : [a2 , b2 ] → Rn from w to v. Taking the direction
into account one gets
Z
f = Φ(γ1 (b1 )) − Φ(γ1 (a1 )) = Φ(w) − Φ(v) = −(Φ(γ(a2 )) − Φ(γ2 (b2 ))
γ1
Z
=− f
γ2
25
particular path taken by the curve. That is
Z Z
f= f
C1 C2
for any two curves connecting two points a ∈ Rn and b ∈ Rn in the same
direction, i.e. both curves travel form a to b.
Then one can show that a vector field f is of gradient type if it is conservative.
This provides us with another method to check if a given vector field f
is of gradient type or not. We are left to check if an integral (tangent line)
along a closed curve is zero. However, here one has to ensure that the closed
curve lies in the domain of definition of the vector field.
where Φ(0) is a constant and where γx is the straight line from the origin 0
to x, that is
γx : [0, 1] → Rn , t 7→ γx (t) = tx.
26
Figure 21:
We can choose any other reference point x0 ∈ Rn instead the origin. In that
case one needs a path connecting x0 and x. It is important that the point and
the path connecting any point with the reference point are lying in the domain
of definition of Φ and f . If Φ1 and Φ2 are two potentials for the vector field
f , then
If one considers the flow of a liquid in a pipe, then the vector field f
yields the velocity vector field of the parametric curve (path) γ. The velocity
vector of the fluid is tangent to a flow line, see figure 22 below.
27
Figure 22:
If one is given a vector field it is easy to draw the flow line. It is the line
threading its way through the vector field in the plane as shown in figure 23.
Figure 23:
28
Example 4.10 (Finding a potential) Let f : R2 → R2 , (x, y) 7→ f (x, y) =
1
2
(−x, y). We sketch the vector field in figure 24 below.
Figure 24:
1 x2
∂x Φ(x, y) = − x ⇔ Φ(x, y) = − + c1 (y)
2 4
1 y2
∂y Φ(x, y) = y ⇔ Φ(x, y) = + c2 (x),
2 4
y2 x2
i.e. Φ(x, y) = 4
− 4
is a solution (C1 = C2 = 0).
29
we can conclude
∂x Φ(x, y) = 2y ⇔ Φ(x, y) = 2xy + C1 (y)
1
∂y Φ(x, y) = x + y ⇔ Φ(x, y) = xy + y 2 + C2 (x).
2
But it is impossible to find any such functions C1 and C2 because one can
never match 2xy versus xy. Thus f is not a gradient vector field.
Example 4.12 (Radial vector fields) For the scalar valued function g : (0, ∞) →
R define the vector field f : Rn → Rn by
x
, for x ∈ Rn \ {0}
(
g(kxk) kxk
f (x) = .
0 , for x = 0
Assume that we can always find a primitive G : (0, ∞) → R with G0 (r) = g(r)
for all r > 0.
We say that f : Rn → Rn is a radial vector field if kf (x)k is constant
within concentric spheres, i.e.
Define
q
n
Φ : R \ {0} → R, x 7→ G(kxk) = G x21 + · · · + x2n .
30
(a) Gravitational force field: Put the origin at the centre of the earth or
some other planet having acceleration gplanet and mass M > 0. The gravita-
tional force on some testing body with mass m > 0 is the vector field
gplanet mM
F : R3 \ {0} → R3 , x 7→ F (x) = − x,
kxk3
The minus sign in the force field ensures that the force is directed to the
centre of the planet.
Since the potential Φ is constant on the level surface of Φ they are called
equipotential surfaces. Note that the force field is orthogonal to the
equipotential surfaces, compare the example of radial vector fields above.
There, the force field is radial and the equipotential surfaces are concentric
spheres.
We finish this section coming back to the notion of a flow line of a vector
field. We connect it to first order differential equations. Geometrically, a flow
line for a given vector field f is a parametric curve (path) that threads its way
through the domain of the vector field in such a way that the tangent vector
of the parametric curve (path) coincides with the vector field (see figure 23
above). A flow line may be viewed as a solution of a system of differential
equations. Let Γ(x, t) for t ≥ 0 and x ∈ Rn be the position at time t of the
point on the flow line through x (at time t = 0) after time t has elapsed.
31
Figure 25:
32
Similarly, Dx Γ(x, t) is a n × n-matrix.
5 Surface Integrals
We want to extend the notion of line integrals to surface integrals in R3 . One
can generalises this to arbitrary k-dimensional surfaces in Rn , 1 < k ≤ n, but
we deal first with the case k = 2 and n = 3.
5.1 Surfaces
We need methods for describing surfaces in R3 . A surface S ⊂ R3 as a subset
of points in R3 can be described by two methods:
1.) Level sets of real-valued functions f : R3 → R. That is,
S = f −1 (c) = {x ∈ R3 : f (x) = c}
for some c ∈ R.
2.) Parametrisation
α1 (s, t)
α : Q → R3 , (s, t) 7→ α(s, t) = α2 (s, t)
α3 (s, t)
Example 5.2 (a) Cylinder (volume). Let R > 0, H > 0. The parametri-
sation
r cos ϕ
α : [0, R] × [0, 2π] × [0, H] → R3 , (r, ϕ, h) 7→ α(r, ϕ, h) = r sin ϕ
h
describes a cylinder (see figure 26 below). Note that when the angle variable
varies only in a subset of [0, 2π] one gets not the whole cylinder but one
where a ’piece of cake’ is removed. If one fixes the radius r ∈ (0, R] one gets
the parametrisation of the cylindrical surface of radius r and height H.
33
Figure 26:
34
Figure 27:
The set S1 = α1 (Q1 ) is the circle line in the plane and the set Sk = αk (Qk )
is a k-dimensional surface. One can prove that
Sk = {x ∈ Rk+1 : kxk = 1}. (5.8)
Sk is called the k-dimensional sphere in Rk+1 . For k = 2 and n = 3 we get
the polar coordinates (earth coordinates)
cos ϕ1 cos ϕ2
α2 (ϕ1 , ϕ2 ) = sin ϕ1 cos ϕ2 ,
sin ϕ2
where the angle ϕ1 ∈ [0, 2π] is the geographical longitude and respectively the
angle ϕ2 ∈ [−π/2, π/2] is the geographical latitude. Note that sometimes one
35
takes the latitude angle Θ = π2 − ϕ2 ∈ [0, 2π] where one takes the angle away
from the north pole. In the latter case one take the angle from the equator
counting negative towards the north pole and positive towards south pole.
Recall that sin ϕ2 = cos Θ respectively cos ϕ2 = sin Θ for ϕ2 ∈ [−π/2, π/2]
and Θ ∈ [0, π].
5.2 Integral
For an integration over a surface in R3 we need surface unit normal vectors N b.
The unit normal vectors have norm 1 and they are perpendicular to the plane
at each single point of the surface (more precisely they are perpendicular to
the tangent plane at that point).
If we describe our surface as a level set of a differentiable real-valued function
f : R3 → R we know from Lemma 2.6 that the gradient is orthogonal to the
level surfaces, i.e.
∇f ⊥ S ⇒ N b := ∇f ∈ R3 .
k∇f k
What is the tangent plane at a point to a given surface S ⊂ R3 ? When our
surface S ⊂ R3 is given with a parametrisation α : Q → R3 we see from figure
29 that at each point α(s, t) of the surface S the vectors ∂α
∂s
(s, t) and ∂α
∂t
(s, t)
36
are both tangent to the surface (that is, they are lying in the tangent plane).
Hence, the cross-product
∂α ∂α
(s, t) × (s, t)
∂s ∂t
is a normal to the surface at the point α(s, t).
We normalise this vector to get the unit normal for the surface S
Remark 5.5 (a) If we integrate 1 over a surface we get the surface area of
S, Z ZZ
∂α ∂α
1= (s, t) × (s, t) dsdt = area (S).
S Q ∂s ∂t
(b) If we take a small rectangle q ⊂ Q in the parameter domain Q whose left
bottom corner is (s, t) ∈ Q and whose side length are ∆s and ∆t respectively,
we can map this rectangle with the parametrisation to R3 . The area of the
image rectangle α(q) can be computed as
∂α ∂α
area (α(q)) = (s, t) × (s, t) ∆s∆t.
∂s ∂t
38
Figure 30: Image of the small rectangle q
39
5.3 Kissing problem
Example 5.7 (a) Spherical cap S: Fix θ ∈ (0, π/2) and define
cos ϕ1 cos ϕ2
α : [−π, π] × [θ, π/2] → R3 , (ϕ1 , ϕ2 ) 7→ α(ϕ1 , ϕ2 ) = sin ϕ1 cos ϕ2
sin ϕ2
We want to compute the surface area of the spherical cap defined by the
parametrisation α. We compute the surface normal vectors at (ϕ1 , ϕ2 ):
− sin ϕ1 cos ϕ2 − cos ϕ1 sin ϕ2
∂α ∂α
= cos ϕ1 cos ϕ2 = − sin ϕ1 sin ϕ2
∂ϕ1 ∂ϕ2
0 cos ϕ2
with norm
∂α ∂α p
× = cos2 ϕ2 = cos ϕ2 ,
∂ϕ1 ∂ϕ2
40
where the last equality follows due to θ ∈ (−π/2, π/2). The surface area is
then given as the surface integral of 1,
Z Z π/2 Z π ϕ2 =π/2
area (S) = 1= 1 cos ϕ2 dϕ1 dϕ2 = 2π sin ϕ2
S θ −π ϕ2 =θ
(c) Newton’s kissing problem How many unit balls can simultaneously
touch a given ball in R3 ? We calculate the shadowed surface area seen in
figure 32.
balls. Hence, the number in question is less than 14. Is the maximum number
now 12, 13 or 14 ?
What about the other dimensions?
41
Kissing in 1d
It is easy to see (and to prove) that in two dimensions the kissing number
is 6.
42
Kissing in 2d
43
Kissing in 3d
44
dimension kissing ] year
1 2
2 6
3 12 (1874)
4 24 (2003)
8 240 (1979)
24 196560 (1979)
45
Figure 34: Velocity not parallel
f~·d~s.
R R
Alternative notations: S
f ·N
b ds or
S
46
Figure 35: box Ω
The unit normal vector N btop can already be seen in figure 35 but we show
the exact calculation.
1 0 0
∂αtop ∂αtop
× = 0 × 1
= 0 = 1.
∂x ∂y
0 0 1
0
Thus, Ntop (x, y, bz ) = 0 with (x, y) ∈ [ax , bx ] × [ay , by ]. Hence, the flux
b
1
through the top is
Z Z bx Z by
hf, Nbi = f3 (x, y, bz ) dydx.
Stop ax ay
2.) Flux through the bottom. This is the same calculation but with the
opposite direction, hence a minus sign appears for the unit surface normal
47
0
Nbbottom (x, y, az ) = 0 with (x, y) ∈ [ax , bx ] × [ay , by ] and the component
−1
function has to be evaluated at az for the third entry.
Z Z bx Z by
hf, N
bi = − f3 (x, y, az ) dydx.
Sbottom ax ay
3.) and 4.) Similarly we derive the contribution from the left and the right
surface, Sl and Sr , as
Z Z bx Z by Z bz
∂f1
hf, N i =
b (x, y, z) dzdydx.
Sl ∪Sr ax ay az ∂x
5.) and 6.) Finally we get the contribution from the back and the front
Z Z bx Z by Z bz
∂f2
hf, N i =
b (x, y, z) dzdydx.
Sback ∪Sfront ax ay az ∂y
Taking the sum over all contributions of the six surfaces gives the total flux
across the surface
as
Z Z bx Z by Z bz ∂f ∂f2 ∂f3
1
hf, N
bi = + + dzdydx. (6.12)
∂Ω ax ay az ∂x ∂y ∂z
6.2 Divergence
In the following we will often write ∂Ω for the surface of some domain Ω ⊂ R3 .
The integrand in (6.12) motivates the following definition.
48
Definition 6.3 Let f : Rn → Rn be a differentiable vector field. The diver-
gence of the vector field f = (f1 , . . . , fn ) is the scalar field
div f : Rn → R
n
X ∂fi
x 7→ div (f )(x) = (x).
i=1
∂xi
~ · f~ or ∇ · f .
Alternative notations are: ∇ · f or ∇
Note the following relations of scalar and vector fields with the corresponding
operations.
grad
Φ : Rn → R −→ grad Φ : Rn → Rn
div
div f : Rn → R ←− f : Rn → Rn
∆Φ : Rn → R
n
X ∂Φ2
x 7→ ∆Φ(x) = div grad Φ(x) = (x).
i=1
∂x2i
Note: n n
X ∂ ∂Φ X ∂ 2 Φ
∆Φ = div (grad Φ) = = 2
.
i=1
∂x i ∂x i i=1
∂x i
49
where vol(Ω) is the volume of the small box Ω, implies that
R
∂Ω
hf, N
bi
div f (a) ≈ . (6.13)
vol(Ω)
The divergence of a vector field gives the outgoing flux per volume.
As seen div f (a) corresponds to the amount of flux of the vector field out
of the small volume divided by the volume. This is a rate of ’expansion’
or ’stretching’ of the vector field (see figure 36 above). The divergence is
div f (x) = 1, x ∈ R3 , and the figure 36 may represent a gas which is expand-
ing.
50
Figure 37: vector field contracting
This vector field is contracting (see figure 37), and its divergence is div f (x) =
−1, x ∈ R3 .
51
This vector field is neither expanding nor contracting (see figure 38), and its
divergence is div f (x) = 0, x ∈ R3 .
Remark 7.2 (a) The theorem also works for dimension n 6= 3 but one has
to define the surface integral.
R
n = 2 : ∂Ω is a line and ∂Ω hf, N b i is a (tangent) line integral.
R
(b) The left hand side of (7.14),
RRR Ω div (f )(x) dx, is just a volume Integral,
that is an iterated integral Ω div f (x, y, z) dxdydz (Fubini’s theorem,
compare Analysis I).
R R
Alternative notations are: Ω div f dV for n = 3 or Ω div f dA for
n = 2.
Example 7.3 (a) Let Ω = B(0, R) ⊂ Rn a ball of radius R > 0, B(0, R) =
{x ∈ Rn : kxk ≤ R}, and denote by ∂Ω the surface (sphere) of Ω, i.e.
∂Ω = ∂B(0, R) = {x ∈ Rn : kxk = R}. The unit outward normal is
then
N b (x) = x , for all x ∈ ∂Ω.
R
n n
Define f : R → R , x 7→ f (x) = x and check that div f (x) = n for all
x ∈ Rn . Gauss’s Divergence Theorem gives
Z Z Z
x
n dx = hx, i = R.
B(0,R) ∂B(0,R) R ∂B(0,R)
52
(b) Differentiability in Gauss’s Divergence Theorem is needed as the fol-
lowing example shows.
2 2 1 x
f : R → R , (x, y) 7→ f (x, y) = 2 .
x +y y
2
∂ x ∂ y
div f (x, y) = ∂x 2
x +y 2 + 2
∂y x +y 2 = 0. Set Ω = B(0, 1) ⊂ R2
2 x
having unit normal N : ∂Ω → R , (x, y) 7→ N (x, y) =
b b . We get
R y
Ω
div f (x, y) dxdy = 0 but
Z Z
hf, N i =
b 1 = length(∂Ω) = 2π.
∂Ω ∂Ω
The theorem does not work due to the singularity at the origin of the
vector field f .
Sketch for the proof of Theorem 7.1. Let Ω ⊂ R3 be a region in R3
and divide it in finitely many small regions Ωi centred at ai ∈ Ω with outward
surfaces ∂Ωi (see figure 39) having unit normal N bi .
53
The approximation becomes exact in the limit vol(Ωi ) → 0. Hence, multiply
(7.15) by vol(Ωi ) and sum up to get
X XZ
div f (ai )vol(Ωi ) ≈ hf, N
b i. (7.16)
i i ∂Ωi
Here, the left hand side will become the volume integral in the limit vol(Ωi ) →
0 (Riemann sum). But what about the right hand side in (7.16)? Consider
in figure 40 two small adjacent regions Ω1 and Ω2 .
Along the common surface of ∂Ω1 and ∂Ω2 we get hf, N b1 i + hf, N
b2 i = 0.
All contributions from the interior of Ω to the sum of the right hand side in
(7.16) cancel out, leaving only the surface integral over the exterior surface
b i. 2
R
∂Ω. Hence, in the limit the right hand side of (7.16) becomes ∂Ω hf, N
54
Figure 41: Cylinder
Surface integral: The surface S of the cylinder Ω has the following three
single parts S = Sbottom ∪ Stop ∪ Scyl.
Sbottom = {(x, y, z)R3 : x2 + y 2 ≤ 4, z = 0}:
0 4x
Nb (x, y, z) = − 0 and f (x, y, z) = −2y 2 for (x, y, z) ∈ Sbottom .
1 0
55
R
As hf (x, y, z), N
b (x, y, z)i = 0 for (x, y, z) ∈ Sbottom we get
Sbottom
hf, N
b i = 0.
Stop = {(x, y, z)R3 : x2 + y 2 ≤ 4, z = 3}:
0 4x
b (x, y, z) = 0 and f (x, y, z) = −2y 2 for (x, y, z) ∈ Stop .
N
1 9
A parametrisation is given by
s cos(t)
αtop : [0, 2] × [0, 2π] → R3 , (s, t) 7→ αtop (s, t) = s sin(t)
3
with
cos(t) −s sin(t)
∂αtop ∂αtop
(s, t) = sin(t) and (s, t) = s cos(t)
∂s ∂t
0 0
and cross-product
0
∂αtop ∂αtop
× = 0 .
∂s ∂t
s
Hence, the top contributes to the flux
Z Z 2π Z 2
hf, N i =
b 9s dsdt = 36π.
Stop 0 0
gives
−2 sin(t) 0
∂αcyl. ∂αcyl.
(t, z) = 2 cos(t)
and (t, z) = 0
∂t ∂z
0 1
and
2 cos(t)
∂αcyl. ∂αcyl. ∂αcyl. ∂αcyl.
× = 2 sin(t) ⇒ × = 2.
∂t ∂z ∂t ∂z
0
56
2 cos t
bcyl. (t, z) = 1 2 cos t, and the cylinder barrel contributes to the
Hence, N 2
0
flux
Z Z 2π Z 3
16 cos2 (t) − 8 sin3 (t) dzdt
hf, N i =
b
Scyl. 0 0
Z 2π Z 3
16 cos2 (t) − 8 sin3 (t) dzdt
=
Z0 2π Z0 3
16 cos2 (t)) + 8 sin(t) cos2 (t) − 8 sin(t) dzdt
=
0 0
Z 2π
16 cos2 (t)) + 8 sin(t) cos2 (t) − 8 sin(t) dt
=3
Z0 2π h1
2 1 i2π
= 48 cos (t) dt = 48 cos(t) sin(t) + t = 48π,
0 2 2 0
where we used that
Z 2π 2π
sin(t) dt = − cos(t) 0 = 0
0
Z 2π 2π
sin(t) cos2 (t) dt = − cos3 (t) 0 = 0.
0
We have thus proved the Divergence Theorem follows for this example, i.e.
Z ZZZ
hf, N i =
b div f (x, y, z) dzdydx.
S Ω
57
RRR
Ω
%(x, t) dx1 dx2 dx3 . The rate of mass flow into the domain Ω is given by
the flux integral Z
− h%~v , N
b i,
∂Ω
where the minus sign signals that N b points outward and where ~v : R3 ×R+ →
R gives the velocity vector ~v (x, t) of the fluid at point x ∈ R3 and time t.
3
Physical law: mass is conserved, that is, the rate of change of mass
in Ω equals the rate at which mass enters Ω.
The right hand side can be written as the volume integral for the divergence of
%~v and the time derivative can be interchanged with the volume integration.
All this gives ZZZ
∂%
(x, t) + div (%~v )(x) dx = 0.
Ω ∂t
As the domain is arbitrary the integrand equals identically zero. Hence, we
have derived the conservation law for the mass of a fluid.
58
Mass conservation law
∂%
+ div (%~v ) = 0.
∂t
8 Integration by Parts
This section introduces integration by parts techniques. We let Ω ⊂ R3 be
some region, i.e., a bounded open subset of R3 , and by ∂Ω we denote the
surface of Ω such that Ω = Ω ∪ ∂Ω. The surface normal is the vector field
Nb : ∂Ω → R3 , N
b (x) = (N
b1 (x), N
b2 (x), N
b3 (x)).
Proof. Without loss of generality choose i = 1 and define the vector field
∂f
v : R3 → R3 , x 7→ v(x) = (f (x), 0, 0). Then div v(x) = ∂x 1
and v · N
b =
hv, N
bi = fNb1 . The claim follows then by the Divergence Theorem, where we
put v = (0, f, 0) and v = (0, 0, f ) for the other cases. 2
59
(c) If f : Ω → R is two times continuously differentiable,
Integration by part (IBP) - vector case
Z Z Z
∆f (x)h(x) dx = bi −
hh∇f, N h∇f (x), ∇h(x)i dx (8.20)
Ω ∂Ω Ω
Proof. (a) Apply Lemma 8.1 to the function f (x) = g(x)h(x), x ∈ Ω, and
use the chain rule
∂f ∂g ∂h
(x) = (x)h(x) + g(x) (x)
∂xi ∂xi ∂xi
for i = 1, 2, 3.
∂g
(b) Use again Lemma 8.1 for f (x) = ∂x i
(x) for i = 1, 2, 3 to obtain
Z 2 Z
∂ g ∂g b
2
(x) dx = Ni .
Ω ∂xi ∂Ω ∂xi
Let Ω ⊂ R3 be some piece of material and fix the temperature f (x) ∈ R+ for
every point x ∈ ∂Ω on the surface (boundary) of Ω. In a steady state the
temperature is a scalar field T : Ω ∪ ∂Ω → R+ solving the following boundary
value problem
∆T (x) = 0 for all x ∈ Ω,
(8.21)
T (x) = f (x) for all x ∈ ∂Ω.
The existence of solutions for (8.21) is difficult, but we can use Proposition 8.2
to show the uniqueness of a solution for (8.21). Assume that T : Ω∪∂Ω → R+
and Te : Ω ∪ ∂Ω → R+ are both a solution to (8.21). Define the scalar field
D : Ω ∪ ∂Ω → R+ , x 7→ D(x) = T (x) − Te(x). This solves the following
60
Proposition 8.2 implies that
Z Z Z
0 = (∆D(x))D(x) dx = − h∇D(x), ∇D(x)i dx + b i.
Dh∇D, N
Ω Ω ∂Ω
Henceforth (because the second integral of the right hand side vanishes)
Z
k∇D(x)k2 dx = 0.
Ω
61
Figure 43: rectangle Ω
We get similar results for the top and the bottom edge. Summing up these
contributions we get the tangent line integral along ∂Ω
Z Z d Z b
hf, Tbi = f2 (b, y) − f2 (a, y) dy + f1 (x, c) − f1 (x, d) dx
∂Ω c a
Z dZ b Z bZ d
∂f2 ∂f1
= (x, y) dxdy − (x, y) dxdy
∂x c ∂y
Zc a a
∂f2 ∂f1
= − (x, y) dxdy,
Ω ∂x ∂y
where we used the Fundamental Theorem of calculus in the second line. The
integrand in the last line motivates the following definition of a scalar field.
62
Definition 9.1 Let f : R2 → R2 be a differentiable vector field. The curl of
the vector field f is the scalar field curl f : R2 → R defined by
curl f : R2 → R
∂f2 ∂f1
(x, y) 7→ curl (f )(x, y) = (x, y) − (x, y).
∂x ∂y
R
Remark 9.2 ∂Ω
hf, Tbi is the circulation around ∂Ω.
−y
Example 9.3 (a) The vector field v : R2 → R2 , (x, y) 7→ v(x, y) =
x
has positive curl, curl (v)(x, y) = 1 − (−1) = 2 for all (x, y) ∈ R2 .
See the sketch of the vector field in figure 44, where the arrows of the
vector field are circulating in positive direction around the origin (that is,
anti-clockwise direction).
2 2 −x
(b) The vector field v : R → R , (x, y) 7→ v(x, y) = has vanishing
−y
curl, curl (v)(x, y) = 0 for all (x, y) ∈ R2 . See the sketch of the vector field
in figure 45, where the arrows are directed towards the origin.
63
Figure 45: zero curl
0
(c) The vector field v : R2 → R2 , (x, y) 7→ v(x, y) = has positive curl,
x
curl (v)(x, y) = 1 for all (x, y) ∈ R2 . In figure 46 the arrows are pointing
upwards in the positive half plane and downwards in the negative half plane.
64
Figure 46: positive curl, a small wheel centred at y-axis would
turn
The calculation above for the line integral along the boundary of a rectangle
showing that the integral is the surface integral of the curl of the vector field
can be proved for any region Ω ⊂ R2 . This is the content of the following
theorem.
Theorem 9.4 (Green’s theorem - Stokes’s theorem in R2 ) Let Ω ⊂ R2
be a bounded region and Tb : ∂Ω → R2 be positively oriented unit tangent
vectors for the boundary line ∂Ω of the region Ω. Let Ω = Ω ∪ ∂Ω. If
f : Ω → R2 , f = (f1 , f2 ), is continuously differentiable, then
Z Z
hf, Tbi = curl (f )(x, y) dxdy. (9.23)
∂Ω Ω
Proof. We give a sketch of the proof only, because one can use the proof of
the Divergence Theorem for it. The contributions of the interior cancel out
here as they do for the Divergence Theorem. To apply that theorem directly
define the vector field
2 2 f2 (x, y)
v : R → R , (x, y) 7→ v(x, y) = .
−f1 (x, y)
65
Then,
∂f2 ∂f1
div v(x, y) = (x, y) − (x, y) = curl f (x, y),
∂x ∂y
and !
−N D
b2 E
hv, N
b i = f2 N
b1 − f1 N
b2 = f, = hf, Tbi.
Nb1
2
In the next proposition we show how one can apply Green’s theorem.
Proof.
Z ZZ ZZ
hf, Tbi = (1 + 1) dxdy = 2 dxdy = 2area (Ω).
∂Ω Ω Ω
9.2 Application
Proposition 9.5 has a direct application for the construction and the use of
a planimeter. A planimeter is a measuring instrument used to measure the
surface area of an arbitrary two-dimensional shape. The most common use
is to measure the area of a plane shape. There are many different kinds of
planimeters but all operate in a similar way. A pointer on the planimeter is
used to trace around the boundary of the shape. This induces a movement
in another part of the instrument and a reading of this is used to establish
the area of the shape. The precise way in which they are constructed varies,
the main types of mechanical planimeter being polar; linear; and Prytz or
66
”hatchet” planimeters. In the linear and polar planimeter, as one point on
a linkage is traced along the shape’s perimeter, that linkage rolls a wheel
along the drawing. The area of the shape is proportional to the number
of turns through which the measuring wheel rotates when the planimeter is
traced along the complete perimeter of the shape. The concept having been
pioneered by Hermann in 1814, Swiss mathematician Jakob Amsler-Laffon
built the first modern planimeter in 1854, the operation of which can be
justified by appealing to Green’s theorem and in particular to Proposition 9.5.
Let us study briefly the so-called linear planimeter, see figure 47 below. The
measuring arm (elbow) can move up and down along the y-axis. Note that
b is the y-coordinate of that measuring arm. The scalar product of that arm
with a vector field N~ = (Nx , Ny ) is
~ ,N
hEM ~ i = xNx + yNy .
Example 9.6 Verify Green’s theorem for the following vector field
2 2 xy + y 2
f : R → R , (x, y) 7→ f (x, y) = ,
x2
67
Figure 47:
68
Prytz
69
Planimeter
Planimeter
70
Figure 48: region Ω
We split the boundary in two pieces ∂Ω = ∂Ω1 ∪ ∂2 (see figure 48) with
parametrisation
t 1
∂Ω1 : γ1 : [0, 1] → R2 , t 7→ γ1 (t) = 2 , γ10 (t) =
t 2t
2 1−t 0 −1
∂Ω2 : γ2 : [0, 1] → R , t 7→ γ2 (t) = , γ2 (t) =
1−t −1
Note that
γ10 (t)
1 1
T1 (γ1 (t)) = 0
b =√ .
kγ1 (t)k 1 + 4t 2t
2
Then,
71
Z Z 1 D E
hf, Tb1 i = f (γ1 (t)), Tb1 (γ1 (t)) kγ10 (t)k dt
∂Ω1 0
Z 1D 3 E
t + t4 1
= 2 , dt
0 t 2t
Z 1
19
t4 + 3t3 dt = .
=
0 20
and
1 D (1 − t)2 + (1 − t)2 −1 E√
Z Z
1
hf, Tb2 i = ,√ 2 dt
∂Ω2 0 (1 − t)2 2 −1
Z
= −3 (1 − t)2 dt = −1.
1
R
This gives ∂Ω
hf, Tbi = − 20 and thus the proof of Green’s theorem for this
example.
10 Stokes’s theorem
Stokes’s theorem gives an alternative expression for the surface integral of
the curl of a vector field. This is given by the flow along the boundary.
Remark 10.2 (a) Positively oriented means that the right hand rule for the
pair (N
b , Tb) is satisfied. That is, if you walk around the boundary, the surface
should be on your left.
(b) curl
72
f : R3 → R3 , f = (f1 , f2 , f3 ) differentiable, then the curl in R3 is the vector
field
∂2 f3 (x, y, z) − ∂3 f2 (x, y, z)
curl f : R3 → R3 , (x, y, z) 7→ curl (f )(x, y, z) = ∂3 f1 (x, y, z) − ∂1 f3 (x, y, z) .
∂1 f2 (x, y, z) − ∂2 f1 (x, y, z)
Proof of Theorem 10.1. Following the last remark we prove the theorem
by ’lifting’ Green’s theorem from R2 to R3 . On the left hand side in figure
49 we have the 2d-world and on the right hand side of figure 49 we have the
3d-world.
73
Figure 49: lifting from R2 to R3
Note that the norm drops out when we compute the tangent line integral.
Henceforth,
Z Z bD
∂α γ1 ∂α ∂γ2 E
hf, TS i =
b f (α(γ(u))), (γ(u)) (u) + (γ(u)) (u) du
∂S a ∂s ∂u ∂t ∂u
! !
hf, ∂α i hf, ∂α
i
Z bD E Z D E
∂s 0 ∂s
= , γ (u) du = , T
bΩ du.
a hf, ∂α
∂t
i ∂Ω hf, ∂α
∂t
i
74
Similarly we get (after some computation and application of the chain rule)
Z ZZ D
∂α ∂α E
hcurl f, N
bi = curl f, × dsdt
S Ω ∂s ∂t
!
hf, ∂α i
ZZ
∂s
= curl ∂α
dsdt.
Ω hf, ∂t i
Example 10.3 We prove Stokes’s theorem for the following example. Let
the surface S = {(x, y, z) ∈ R3 : z = x2 + y 2 , z ≤ 4} with parametrisation
x
α : B(0, 2) → R3 , (x, y) 7→ α(x, y) = y ,
x2 + y 2
−2x
b (x, y, z) = ∇s(x, y, z) = 1 −2y ,
N (x, y, z) ∈ S = α(B(0, 2)).
k∇s(x, y, z)k 3
1
75
and therefore
Z Z 0 −2x E
0 , 1 −2y 3 dxdy
D
hcurl f, N
bi =
S B(0,2) 3
2x − 1 1
Z Z
= 2x dxdy − dxdy = 0 − vol(B(0, 2)) = −4π.
B(0,2) B(0,2)
Example 10.5 (Faraday’s law) Let the time-dependent electric (E) and
magnetic (H) field be given,
E : R × R3 → R3 ,
H : R × R3 → R3 .
76
which is time-dependent, and the time-dependent magnetic flux is given by
the flux integral Z
hH, N
b i.
S
(ii) For any two oriented simple closed curves C1 ⊂ R3 and C2 ⊂ R3 having
R R
the same starting and terminal point: C1 hf, Tb1 i = C2 hf, Tb2 i.
77
10.2 Polar coordinates
As a reminder we summarise briefly important facts about spherical coordi-
nates.
Polar coordinates in R3
r cos ϕ cos θ
x : [0, ∞) × [−π, π] × [0, 2π] → R3 , (r, θ, ϕ) 7→ x(r, θ, ϕ) = r sin ϕ cos θ .
r sin θ
Here the single coordinates have the following names.
r ∈ [0, ∞) is called the radius,
θ ∈ [−π, π] is called the (geographical) latitude,
ϕ ∈ [0, 2π] is called the (geographical) longitude.
Compare with Example 5.2 part (c) where we defined iteratively polar co-
ordinates for any dimension. A function f : R3 → R can be given in polar
coordinates by writing
f : [0, ∞) × [−π, π] × [0, 2π] → R, (r, θ, ϕ) 7→ f (r, θ, ϕ) = f (x(r, θ, ϕ)).
Basis vectors
cos(ϕ) cos(θ)
∂x er
er = = sin(ϕ) cos(θ) , ker k = 1, êr = = er ,
∂r ker k
sin(θ)
78
Any vector field can be expressed in this basis,
V : [0, ∞)×[−π, π]×[0, 2π] → R3 , (r, θ, ϕ) 7→ Vr (r, θ, ϕ)êr +Vθ (r, θ, ϕ)êθ +Vϕ (r, θ, ϕ)êϕ .
Warning:
∂Vr ∂Vθ ∂Vϕ
div V 6= + + .
∂r ∂θ ∂ϕ
Applying the chain rule one obtains after some calculation the differential
operators.
Differential operators:
∂f 1 ∂f 1 ∂f
grad (f ) = êr + êφ + êθ
∂r r cos θ ∂φ r ∂θ
1 ∂ 2 1 ∂vφ 1 ∂
div (v) = r vr + + cos(θ)vθ
r2 ∂r r cos θ ∂φ r cos θ ∂θ
1 ∂vθ ∂ 1 ∂vr ∂
curl (v) = − cos(θ)vφ êr + − rvθ ) êφ
r cos θ ∂φ ∂θ r ∂θ ∂r
1 ∂ 1 ∂vr
+ rvφ − êθ
r ∂r cos θ ∂φ
79
Part II
where
Re(f ) = u : R2 → R2 , (x, y) 7→ u(x, y) is the real part,
Im(f ) = v : R2 → R2 , (x, y) 7→ v(x, y) is the imaginary part.
80
Figure 50: Graphical representation of a complex number
z = x + iy = Re(z) + iIm(z)
p
= r cos(ϕ) + i sin(ϕ) with r = x2 + y 2 = |z|
and the argument ϕ = arctan( xy ) = arg(z). We call x = Re(z) the real part
p
and y = Im(z) the imaginary part of z = x + iy ∈ C and |z| = x2 + y 2 the
modulus of z = x + iy ∈ C.
(i) z = z.
(ii) z + w = z + w.
(iii) zw = zw.
81
(iv) |z| = |z|.
Moivre-Laplace:
z n = 1 ⇔ |z|n einϕ = 1
⇔ |z| = 1 and cos(nϕ) + i sin(nϕ) = 1
2kπ
⇔ |z| = 1 and ϕ = k = 0, 1, . . . , n − 1.
n
Hence, the distant roots of z n = 1 are given by
2kπi
zk = e n , k = 0, 1, . . . , n − 1,
zn → z as n → ∞ ⇔ |zn − z| → 0 as n → ∞
⇔ |xn − x| → 0 as n → ∞ and |yn − y| → 0 as n → ∞.
Note that | · | is used in the first line as the modulus for complex numbers and
in the second line as the modulus for real numbers.
Lemma 11.2 (Rules for converging sequences) Let (zn )n∈N , (wn )n∈N be
complex sequences with zn → z ∈ C and wn → w ∈ C as n → ∞. then
82
(a) zn + wn → z + w as n → ∞.
(b) zn wn → zw as n → ∞.
zn z
(c) wn
→ w
as n → ∞ if w 6= 0.
Proof. Exercise. 2
Notation 11.4 Let ε > 0. The set
B(z, ε) = {z 0 ∈ C : |z 0 − z| < ε}
is the open ball (circle) in C around z ∈ C with radius ε, see figure 51. The
circle line is ∂B(z, ε) = {z 0 ∈ C : |z 0 − z| = ε} and the closed ball is
B(z, ε) = B(z, ε) ∪ ∂B(z, ε) = {z 0 ∈ C : |z 0 − z| ≤ ε}.
83
The circle centre a ∈ C and radius r > 0 is the locus of points at distance
r from a so has equation |z − a| = r. There is another equation: let α, β ∈ C
with α 6= β and let λ ∈ R, with λ > 0 and λ 6= 1. The equation
z−α
= λ, λ > 0, λ 6= 0, (11.26)
z−β
84
These points lie at opposite ends of a diameter and writing the equation
of the circle in the form |z − a| = r we get
1 1
a = (z1 + z2 ) and r = |z1 − z2 |.
2 2
Hence
1 1
α − a = λ(z2 − z1 ) and λ(β − a) = (z2 − z1 ).
2 2
Thus
1
(α − a)(β − a) = (z2 − z1 )(z2 − z1 ) = r2 .
4
The points α and β in C are inverse points with respect to the circle
|z − a| = r if and only if they satisfy (α − a)(β − a) = r2 . Note that we must
always have one of α and β inside the circle and the other outside see figure
52.
85
Figure 53: Riemann sphere - stereographic projection
Correspondence C
e and Σ:
We shall add some arithmetic rules for the extended complex plane C:
e
86
projects onto to a circle in C, and that every circle in C arises in this manner.
It is thus natural to regard lines as ’circles through infinity’, and to adopt
the collective name circline for a circle or straight line in C. e Henceforth,
(11.26) with λ > 0 represents a circline which is a line if λ = 1 and a circle
(in C) otherwise.
Möbius transformations
It is easy to see that mappings of the first three types take straight lines to
straight lines and circles to circles. Consider the inversion z 7→ w = 1/z
as a map from C e to C.
e The image of the line Re z = 1 is described in
inverse-point form as |z| = |z − 2| (points equidistant from 0 and 2). We get
that |1/w| = |1/w − 2|, that is, |2w − 1| = 1. As inversion is self-inverse,
we see also that the circle |2z − 1| = 1 maps to the line Re w = 1 under
z 7→ w = 1/z. Henceforth under inversion a line may map to a circle and a
circle may map to a line.
f −1 : C e f −1 (w) = dw − b .
e → C,
a − cw
Proposition 11.6 (Circlines under Möbius transformations) Let C be
a circline with inverse points α and β in C and let f be a Möbius transfor-
mation. Then f maps C to a circline, with inverse points f (α) and f (β).
87
Remark 11.7 In real analysis convergence of a sequence is defined with the
so-called ε − n0 -criterion. For complex valued sequences (zn )n∈N it reads as
follows:
f : D → C is continuous at z ∈ D
⇔ ∀ε > 0 ∃ δ > 0 : w ∈ D and |z − w| < δ ⇒ |f (z) − f (w)| < ε
⇔ ∀ε > 0 ∃ δ > 0 : w ∈ B(z, δ) ⇒ f (w) ∈ B(f (z), ε).
f (z + h) − f (z)
f 0 (z) := lim
h→0 h
exists, that is, if the limit of the sequence f (z+hhnn)−f (z) exists for every sequence
(hn )n∈N in C with hn → 0 as n → ∞, and this limit is (when it exists) denoted
by f 0 (z).
So far there seems to be no much difference with real analysis. But the
following example sheds some light onto the new features of complex analysis.
88
(b) f : C → C, z 7→ f (z) = z = x − iy for z = x + iy. As
zn = xn + iyn → z as n → ∞ ⇒ xn → x and yn → y as n → ∞
⇒ z n → z as n → ∞,
f (z + h) − f (z) z+h−z h
= = .
h h h
If we choose hn = n1 , n ∈ N, we get hn
hn
= 1 for any n ∈ N. But if we choose
hn = i n1 , n
∈ N, we get hn
= −1. Hence, the limit of the differential quotient
hn
does not exist and therefore f is not differentiable at z ∈ C.
89
The Cauchy criterion is used in the proof of the following lemma.
∞
X
f (z) := fn (z)
n=0
The converges of ∞
P
n=0 Mn implies that Mn → 0 as n → ∞. Hence, (|sM (z)−
sN (z)| → 0 as N, M → ∞, and thus f (z) exists (i.e., the sum converges).
For the proof of the continuity part see real analysis. 2
We write R = ∞ if ∞ n
P
n=0 cn (z − a) converges for arbitrarily large |z|,
i.e., for all z ∈ C.
90
P∞ n
(b) n=0 cn z does not converge for z ∈ C with |z| > R.
(d) If ∞
P n
P∞ n
n=0 cn (z − a) diverges for z1 ∈ C then n=0 cn (z − a) diverges
for all z ∈ C with |z − a| > |z1 − a|.
Remark 12.5 (a) The series may or may not converge for |z| = R. This
has to be analysed in detail for any example. Recall that ∂B(0, R) is the circle
line of radius R around the origin and B(0, R) is the closed ball. The set of
complex numbers for which the power series converges includes the open ball
B(0, R) and possibly some points from the boundary line ∂B(0, R).
f is C ∞ on B(0, R) with
dk f X
(z) = cn n(n − 1) · · · (n − k + 1)z n−k .
dz k n=k
91
(d) We have put a = 0 in the theorem because, by shifting,
∞
X
g(z) := cn (z − a)n
n=0
converges for all z ∈ B(a, R) and f (z) = g(z + a), z ∈ B(0, R).
If
|cn+1 |
→ α ∈ [0, ∞] as n → ∞,
|cn |
P∞ P∞
then
P∞ n=0 |c n | converges if α < 1 (i.e. n=0 cn converges absolutely), and
n=0 |cn | diverges if α > 1. If α = 1 then the test gives no information.
If p
n
lim |cn | = α ∈ [0, ∞),
n→∞
P∞ P∞
then
P∞ n=0 |cn | converges if α < 1 (i.e. n=0 cn converges absolutely), and
n=0 |cn | diverges if α > 1. If α = 1 then the test gives no information.
Proof.
A: Let α < 1. Then there a q ∈ (α, 1). Hence, there is n0 ∈ N such that
cn+1
cn
< q for all n ≥ n0 . W.l.o.g. we can assume that the last statement
holds for all n ∈ N as any addition of finitely many terms does not change
the convergence properties of the series. We get
cn+1 cn c2 cn+1
··· = , ∀n ∈ N.
cn cn−1 c1 c1
92
many terms). This proves the first (convergence)
p part. If α > 1 we have
a subsequence (cnk )k∈N with limk→∞ nk |cnk | = α such that there is k0 ∈ N
with |cnk | > 1 for all k ≥ k0 . But this contradicts the necessary condition
cn → 0 as n → ∞ for the convergence of the series. Therefore the series
diverges for α > 1. P∞ 2
n
The convergence set of a power series n=0 cn (z −a) is the set of complex
numbers z for which the series converges. This set includes the open disc
|z − a| < R if R is the radius of convergence. It may also contain points of
the boundary, however, this has to check for each example separately.
Example 12.6 (a) For which complex numbers does the series
∞
X ∞
X
n
cn (z − a) = (3 + i)(2i)n (z + i)n
n=0 n=0
93
What happens on the boundary of that open ball? For complex numbers
z ∈ C with |z + i| = 12 we get |c|cn+1
n|
|
= 1, and hence (|cn |)n∈N and there-
fore (cn )n∈N does not converge to zero. Thus the series is divergent for any
point on the boundary ∂B(−i, 21 ). The series converges only in the open ball
B(−i, 12 ).
|cn+1 | n
= |z| → |z| as n → ∞,
|cn | n+1
gives R = 1. We know that on the boundary line ∂B(0, 1) the series converges
for z = −1 but it diverges for z = 1 (harmonic series). The convergence set
is shown in figure 55.
Figure 55: convergence set (b)
94
Definition 12.7 For z ∈ C the following power series are defined
∞
X zn
ez := ,
n=0
n!
∞
X (−1)n
cos(z) = z 2n ,
n=0
(2n)!
∞
X 1 2n
cosh(z) := z , (12.28)
n=0
(2n)!
∞
X (−1)n 2n+1
sin(z) := z ,
n=0
(2n + 1)!
∞
X 1
sinh(z) := z (2n+1) .
n=0
(2n + 1)!
Proposition 12.8 The power series on the right hand side of (12.28) have
radius of convergence R = ∞.
Proof. Exercise. The ratio test shows, easily, that all series have infinite
radius of convergence. 2
How do the functions defined in (12.28) behave for z ∈ R? From the
Power Series Theorem 12.4 we know for example that
∞ ∞
d z X z n−1 X z n
e = n = = ez .
dz n=1
n! n=0
n!
The following lemma gives a representation of some functions with the ex-
ponential function.
eiz − e−iz
sin(z) = ,
2i
eiz + e−iz
cos(z) = ,
2
ez − e−z
sinh(z) = ,
2
ez + e−z
cosh(z) = .
2
95
Proof. First note that
n n 2in ; if n = 2k is even
i + (−i) =
0 ; if n is odd
Using the power series for the exponential function we get
∞
1 iz −iz
1 X (iz)n (−iz)n
e +e = +
2 2 n=0 n! n!
∞
1 X zn n
i + (−i)n
=
2 n=0 n!
∞
X z 2k (−1)k
= = cos(z).
k=0
(2k)!
(a) e0 = 1,
(b) ez+w = ez ew for all z, w ∈ C,
(c) ez 6= 0 for all z ∈ C.
Proof. (a) Follows from the power series. (b) We first sketch the long
method (Cauchy product rule for series). Put
X z n wm
pl = for all l ∈ N0 .
n,m∈N0 : n+m=l
n! m!
96
Then
∞ n X∞ ∞ ∞ l ∞
X z wm X X 1 X l n l−n X (z + w)l
= pl = z w = .
n=0
n! m=0
m! l=0 l=0
l! n=0
n l=0
l!
Periodicity
Proof. We prove only (a) as (b) and (c) follow immediately analogously.
Put z = x + iy. Then
arg(z) := {θ ∈ R : z = |z|eiθ }.
Note that arg(z) is a countably infinite set consisting of all numbers of the
form θ + 2kπ, k ∈ Z.
97
(a) arg(i) = {(4k + 1) π2 : k ∈ Z}.
98
Figure 56: the roots of z 6 = −1
13 Holomorphic functions
In this section holomorphic functions will be defined. The starting point is
Example 11.12(b). The complex conjugation is not differentiable at any point
z ∈ C. We want to understand this. Therefore we assume that the function
f : C → C, z 7→ f (z) = u(x, y) + iv(x, y) is differentiable at z = x + iy ∈ C.
We study the following two cases for the limit of the differential quotient.
99
The function f is differentiable at z = x + iy and therefore all the limits
of the differential quotients exist and are equal. Thus we get the following
equations
∂u ∂v ∂v ∂u
= and =− . (13.29)
∂x ∂y ∂x ∂y
Proof. (a) This follows as indicated above. (b) Apply the mean value
theorem of real analysis (exercise). 2
100
The Cauchy-Riemann equations are satisfied for all (x, y) ∈ R2 and the
partial derivatives are continuous on R2 , hence the function f is differentiable
on C.
The condition in Theorem 13.1 that there is a small ball in the domain of
definition is crucial. In the following we will assume that the domain for any
function has the property that one can find around any point of that domain
a small ball contained in the domain. Such sets have a special name.
101
f
(c) Suppose g(z) 6= 0 for all z ∈ D. Then g
is holomorphic in D with
f 0 f 0 (z)g(z) − f (z)g 0 (z)
(z) = , z ∈ D.
g g 2 (z)
102
(a) f 0 (z) = 0 for all z ∈ D.
(b) |f | is constant in D.
Our aim is to apply the theorems of vector analysis for the study of
complex valued functions on C. Let f : C → C be a holomorphic function on
C with real and imaginary part u respectively v. Define the two-dimensional
vector field
2 2 F1 (x, y) u(x, y)
F : R → R , (x, y) 7→ F (x, y) = = .
F2 (x, y) −v(x, y)
Then
∂u ∂v
div F = − =0
∂x ∂y
F2 ∂F1
curl F = −
∂x ∂y
∂v ∂u
=− − = 0.
∂x ∂y
Hence, the theorems of Gauss and Green are useful here. The missing piece is
the notion of integration for complex valued functions defined on the complex
plane. This is the content of the next section.
14 Complex integration
The complex line integral will be defined, the Fundamental Theorem of Cal-
culus will be proved and examples of non-vanishing integrals along closed
paths are studied. A curve (path) C ⊂ C is a set points in the complex plane
for which one can find a one-dimensional parametrisation (i.e. a parametri-
sation that depends only on one real variable). Let a parametrisation for a
curve C in the complex plane C be given, i.e.
where a ≤ b and where x(t) is the real part Re(γ(t)) and y(t) is the imaginary
part Im(γ(t)) of the curve. Recall that γ is a parametrisation of the curve
C if γ is piecewise C 1 and γ([a, b]) = C. The derivative with respect to the
real parameter t is γ 0 (t) = x0 (t) + iy 0 (t), t ∈ [a, b].
103
Definition 14.1 The line (path) integral of a function f : D → C, D ⊂ C,
along the curve (path) C ⊂ D ⊂ C with parametrisation γ : [a, b] → C is
defined as
Z Z b
f= f (γ(t))γ 0 (t) dt
γ a
Z b Z b
(14.31)
0 0
= Re(f (γ(t))γ (t)) dt + i Im(f (γ(t))γ (t)) dt
a a
R R
Alternative notation: C
f and γ
f (z) dz.
The following properties are easily proved (compare part I of the lecture).
104
Lemma 14.3 Let C ⊂ C be a curve in the complex plane with parametri-
sation γ : [a, b] → C (i.e., γ([a, b]) = C) and let f : C → C be a continuous
function.
(a) If γ − : [a, b] → C, t 7→ γ − (t) = γ(a + b − t) is the reversed parametrisa-
tion, then Z Z
f =− f.
γ− γ
The line integral for the complex conjugate, i.e., for the function f : C →
C, z 7→ f (z) = z, along C is computed as follows. A parametrisation for
105
the circle line is given by γ(t) = i + 2eit for t ∈ [0, 2π] with derivative
γ 0 (t) = 2ieit , t ∈ [0, 2π]. Hence,
Z Z 2π Z 2π
− i + 2e−it 2ieit dt
it
f= it
i + 2e 2ie dt =
γ 0 0
Z 2π it 2π
e
2eit + 4i dt = 2
= + 8πi = 8πi.
0 i t=0
Why does the integral in Example 14.4 does not vanish, although we
have a closed path? We will come back to this later. We now study the most
important integral in complex analysis, the so-called Fundamental Integral.
106
points on the curve and/or at points surrounded by the curve. However,
example 14.5 shows that there are cases when the line integral along the
closed circline vanishes although the function is not holomorphic at points
inside the circle line. We therefore expect to have vanishing line integrals
along closed paths as long the integrand functions are holomorphic. The
next theorem is the first step towards the so-called Cauchy-Theorem which
proves this conjecture.
Proof.
Z Z b Z b
df df d
= (γ(t))γ 0 (t) dt = (f (γ(t))) dt
γ dz a dz a dt
Z b Z b
d d
= Re(f (γ(t))) dt + i Im(f (γ(t))) dt
a dt a dt
= f (γ(b)) − f (γ(a)),
where we used the chain rule and the FTC for the real functions Re(f (γ(t)))
and Im(f (γ(t))). 2
15 Cauchy’s theorem
In this section the Cauchy Theorem is proved and analysed. The proof
involves to specify an admissible class of paths/curves. Before that we shall
discuss the connection to Green’s and Gauss’s theorem. Let γ : [a, b] →
C, t 7→ γ(t) = x(t) + iy(t) be a parametrisation of a curve (path) C ⊂ C and
107
let a continuous function f : C → C, f = u + iv, be given. We compute
Z Z b
f= (u(γ(t)) + iv(γ(t))(x0 (t) + iy 0 (t)) dt
γ a
Z b b Z
0 0
u(γ(t))y 0 (t) + v(γ(t))x0 (t) dt
= u(γ(t))x (t) − v(γ(t))y (t) dt + i
a a
Z bD 0 E Z bD 0
u(γ(t)) x (t) u(γ(t)) y (t) E
= , 0 dt + , dt
a −v(γ(t)) y (t) a −v(γ(t)) −x0 (t)
Z Z
= hF, T i + i hF, N
b b i,
γ γ
0 0
0 x (t) y (t)
where T = γ (t) =
b 0 and N =
b 0 and F : R2 → R2 , (x, y) 7→
y
(t) −x (t)
u(x, y)
F (x, y) = . Note that N
b is orthogonal to Tb. At the end of
−v(x, y)
Section 13 we showed that div F = curl F = 0. Hence, we can easily derive
Cauchy’s theorem from Green’s theorem and Gauss’ s theorem. However, it
will turn out that the topological properties of the curves (paths) are linked
to the calculus of complex valued function defined on domains in C. We only
indicate this issue briefly and defer a detailed study of this relationship to
later courses. Hence, we will not define properly what deformation in the
following definition means. A deformation of a curve (path) is a continuous
transformation in the complex plane.
108
Proof. We sketch the proof ignoring the dependence on the topological
details of the curve C ⊂ D. By Ω we denote the region surrounded by the
curve (path), that is the boundary ∂Ω = C. See figure 59.
Figure 59: region D with Ω ∪ ∂Ω
where the last equality follows from Green’s and Gauss’s theorem respectively
using the fact that γ is a closed path. 2
Remark 15.3 The Theorem also holds for more general curves, see figure
60,
109
Figure 60: γ union of γ1 ∪ γ2
R R R
where the integral γ f = γ1 f + γ2 f = 0 vanishes because Green’s and
Gauss’s theorem can be applied to Ω1 and Ω2 respectively. But the region
the curve (path) is encircling has to be simply connected, i.e., it must be
possible to contract it to a single point. This contraction is not possible in
the example of an annulus in figure 61,
110
Figure 61: annulus with ∂Ω ⊂ D and Ω 6⊂ D
where γ = ∂Ω but Ω 6⊂ D.
(c) A circline path is a path which is the join of finitely many paths of
type (a) and (b), and a contour is a simple closed circline path.
111
(e) The curve (path) parametrised by γ : [0, 2π] → C, t 7→ γ(t) = a + reit
is called the positively oriented circle with centre a and radius r,
and the curve (path) with parametrisation γ : [0, 2π] → C, t 7→ γ(t) =
a + re−it is called the negatively oriented circle. Positively oriented
means anti-clockwise in the following.
112
Figure 63: ∂B(0, R)
where γ is the parametrisation for ∂B(i, 2), a circle line of radius 2 around
i. As 4π is the area of that circle one can guess that the following holds.
Let γ be the parametrisation of any closed curve (path) C. Then
Z
z = 2i(area(I(γ))), (15.35)
γ
where I(γ) is the region enclosed by the curve (path). This can be seen as
follows. Recall that for any a, b ∈ C one can prove that Im(ab) = 2(area of
the triangle spanned by a and b).
113
Figure 64: z, z + ∆
114
Figure 65: γ = γ1 ∪ γ2
The proof follows with the following theorem which we cite here without
proof.
115
Z Z
f= f.
γ ∂B(a,r)
Z Z
f= f.
γ γ
e
R
Example 15.8 Integrate ∂B(0,2) f , where f : C → C, z 7→ f (z) = 2(4z 2 −
1)−1 . The function has two singularities inside the circline ∂B(0, 2). We
write
1 1
f (z) = − , z ∈ C.
2z − 1 2z + 1
The Deformation Theorem 15.7 implies that
Z Z Z
1 1
f= −
∂B(0,2) 2z − 1 2z + 1
Z∂B(0,2) Z∂B(0,2)
1 1
= −
∂B( 12 , 14 ) 2z − 1 ∂B(− 12 , 14 ) 2z + 1
1 1
= 2πi − 2πi = 0,
2 2
where we used the circlines ∂B( 12 , 14 ) and ∂B(− 12 , 14 ) such that we avoid one
of the singularities respectively, see figure 66 (other circline (paths) would do
the same).
116
Figure 66: B( 12 , 41 ) and B(− 21 , 14 )
16 Cauchy’s formulae
Armed with Cauchy’s theorem we can prove a host of striking results about
holomorphic functions. In Subsection 16.2 we prove Liouville’s theorem,
differentiability of any order, and Taylor’s theorem.
• Z
1
dw = 2πi
∂B(a,r) w−a
• f (w) − f (a) → 0 as w → a
117
Theorem 16.1 (Cauchy’s integral formula) Let the function f be holo-
morphic inside and on a positively oriented contour with parametrisation
γ : [a, b] → C. Then
Z
1 f (w)
f (z) = dw for z ∈ I(γ). (16.37)
2πi γ w−z
Proof. Pick z ∈ I(γ). Then there is an ε > 0 with B(z, ε) ⊂ I(γ). Use the
Deformation Theorem 15.7, add and subtract an integrand with numerator
f (z) to obtain
Z Z
f (w) f (w)
dw = dw
γ w−z ∂B(z,ε) w − z
f (w) − f (z)
Z Z
f (z)
= dw + dw
∂B(z,ε) w − z ∂B(z,ε) w−z
=: I + II.
Remark 16.2 Holomorphic function are special: the values of the function
at the boundary of a closed curve γ determine the value at z ∈ I(γ).
Can we also get integral formulae for the derivatives? Let the function f
be holomorphic inside and on a positively oriented contour parametrised by
γ. Cauchy’s formula (16.37) implies
Z
1 f (w)
f (z) = dw for z ∈ I(γ),
2πi γ w − z
118
so we may differentiate the expression on the right hand side with respect to
z neglecting the integral. We obtain
Z
0 1 f (w)
f (z) = dw for z ∈ I(γ),
2πi γ (w − z)2
Proof. Pick ∈ I(γ). Then there is an ε > 0 such that B(z, 2ε) ⊂ I(γ).
Using formula (16.37) for the differential quotient (for h small enough)
f (z + h) − f (z)
Z
1 1 1
= f (w) − dw
h 2hπi ∂B(z,2ε) w−z−h w−z
Z
1 f (w)
= dw
2πi ∂B(z,2ε) (w − z − h)(w − z)
Z
1 f (w)
= dw
2πi ∂B(z,2ε) (w − z)2
Z
1 1 1
+ f (w) − dw
2πi ∂B(z,2ε) (w − z − h)(w − z) (w − z)2
Z Z
1 f (w) 1 hf (w)
= dw + dw.
2πi ∂B(z,2ε) (w − z)2 2πi ∂B(z,2ε) (w − z − h)(w − z)2
119
We have to show that the second term on the right hand side vanishes as
h → 0. For that choose h so small that |h| < ε and conclude for all w ∈
∂B(z, 2ε) that |w − z − h| ≥ |w − z| − |h| > ε. As f is continuous there is
a M > 0 with |f (w)| ≤ M for w ∈ ∂B(z, 2ε). Hence the second term on
the right hand side is bounded by |h|M 2ε
ε(2ε)2
which converges to zero as h → 0.
The higher derivatives are proved via induction using formula (16.37) for the
difference f (k+1) (z + h) − f (k) (z). 2
where γ is the parametrisation of a circle line ∂B(a, r), 0 < r < R, or of any
positively oriented contour in B(a, R) enclosing the point a.
Proof. Pick z ∈ B(a, R) and choose r > 0 such that |z − a| < r < R and
let γ be the parametrisation of the circline ∂B(a, r). The formula (16.37)
gives Z
1 f (w)
f (z) = dw. (16.41)
2πi ∂B(a,r) (w − z)
Because of |z−a| < |w−a| = r for all w ∈ ∂B(a, r) we have |z−a|/|w−a| < 1
and can expand (in a geometric series)
1 1 1
= .
w−z w − a 1 − z−a
w−a
120
Inserting this into the integrand in (16.41) and interchanging summation
and integration (to be justified later in Theorem 18.7) we conclude with
the existence of the coefficients and their representation by
P∞(16.40). The
uniqueness follows from the following. Assume that f (z) = n=0 cn (z − a)n
for all z ∈ B(a, r) for some 0 < r < R. Provided summation and integration
can be interchanged we have for any n ∈ N0
Z Z ∞
f (z) X
dz = ck (z − a)k (z − a)−n−1 dz
∂B(a,r) (z − a)n+1 ∂B(a,r) k=0
∞
X Z
= ck (z − a)k−n−1 dz = 2πicn
k=0 ∂B(a,r)
Hence
1 |a − b|2πRM
|f (a) − f (b)| ≤ → 0 as R → ∞,
2π ( 12 R)2
implying f (a) = f (b). 2
The following example shows how Cauchy’s formulae can be used to evaluate
integrals.
121
(b)
2πi d3 1
Z
1 2πi
dw = =− .
∂B(−1,3) (w − 4)(w + 1)4 3! dz 3 z − 4 z=−1 54
The problem when singularities are on the curve is studied in the next section.
17 Real Integrals
This section will demonstrate the usefulness of Cauchy’s Theorem for evalu-
ating real integrals. We focus on the following example. What is the integral
Z ∞
sin(x)
dx ?
−∞ x
eiz
f : C → C, z 7→ f (z) = ,
z
which is holomorphic in the punctuated complex plane C \ {0}.
Pick some R > 0, and choose ε > 0 with ε < R. Consider the closed curve
(path) γ in figure 67. It is a union of the curves (paths) γ1 , γ2 , γ3 and γ4 .
122
Figure 67: path γ
γ1 is the half-circle of radius R > 0 from (R, 0) to (−R, 0), γ2 is the straight
line from (−R, 0) to (−ε, 0) and γ4 is the straight line from (ε, 0) to (R, 0).
To avoid the singularity at the origin the path γ3 is the small half-circle of
radius ε from (−ε, 0) to (ε, 0). This path is negatively oriented (clockwise).
iz
The function f (z) = ez is inside γ and on γ ∗ holomorphic. Hence, Cauchy’s
theorem gives Z Z
f= f = 0. (17.43)
γ ∪4i=1 γi
The contributions from the segments on the real axis are evaluated in the
limit ε → 0.
Z Z Z −ε ix Z R ix
e e
f+ f= dx + dx
γ2 γ4 −R x ε x
Z −ε Z R
cos(x) + i sin(x) cos(x) + i sin(x)
= dx + dx
−R x ε x
Z −ε Z R
i sin(x) i sin(x)
= dx + dx
−R x ε x
Z R
sin(x)
−→ i dx as ε → 0.
−R x
123
From (17.43) we will conclude
Z Z Z R Z Z
sin(x)
lim f+ f =i dx = − lim f+ f . (17.44)
ε→0 γ2 γ4 −R x ε→0 γ3 γ1
To prove (17.42) we divide (17.44) by i and take the limit R → ∞. For that
we compute the integrals on the right hand side of (17.44).
(1) Adding and subtracting an integrand we get for the first integral on the
right hand side of (17.44)
Z iw Z iw
e −1 e −1
Z Z Z
1 1 1
f= dw + dw = − dw + dw
γ3 γ3 w γ3 w 2 ∂B(0,ε) w γ3 w
Z iw
e −1
= −iπ + dw =: −iπ + II,
γ3 w
where the minus sign is due to the negative orientation of γ3 and the 12 is
there because we only have the half-circle. It remains to show that the second
term II on the right hand side of the last equation vanishes as ε → 0.
eiw − 1
|II| ≤ hlength(γ3 )i max ≤ M πε → 0 as ε → 0,
w∈∂B(0,ε) w
iw
because e w−1 is continuous on ∂B(0, ε) (attains its maximum value because
the circline is closed and bounded). Hence,
Z
lim f = −iπ. (17.45)
ε→0 γ3
Thus we need to show in (2) that the second integral on the right hand side
of (17.44) vanishes in the limit R → ∞. The other term does not depend on
R.
124
t 2t
The inequality follows from the estimation sin(t) ≥ π/2 = π
for all t ∈
[0, π/2], see figure 68. A proof can be found in Lemma 17.1.
Figure 68: estimation sin(t) on [0, π/2]
2t
This implies e−R sin(t) ≤ e−R π for all t ∈ [0, π/2]. We conclude from (17.44)
and (17.45)
Z R Z Z
sin(t)
lim i dx = − lim lim f+ f = iπ
R→∞ −R x R→∞ ε→0 γ3 γ1
to get (17.42). 2
Lemma 17.1 (Jordan’s inequality)
2 sin(x) π
≤ ≤1 for 0 < x ≤ .
π x 2
Proof. It suffices to show that the function sin(x)
x
decreases for x ∈ (0, π2 ],
i.e., to show that
d sin(x) x cos(x) − sin(x) π
= 2
≤ 0, x ∈ (0, ].
dx x x 2
This follows easily from [x cos(x) − sin(x)]|x=0 = 0 and from
d
x cos(x) − sin(x) = −x sin(x) ≤ 0.
dx
2
125
18 Power series for Holomorphic functions
This section completes our study of power series and its proofs are based on
Cauchy’s theorem and Cauchy’s formulae.
Proof. We only sketch the P ideas, parts of the proof follow analogously to
the real case. Assume that ∞ c
n=0 n z n
converges for |z| < R. Pick 0 < ρ < R
|z| n
and let |z| < ρ and assume z 6= 0. Now ρ < 1 implies that ∞
|z| P
n=0 n ρ
converges (Ratio test for example). Hence there is a M > 0 such that
|z| n
n ≤M for all n ∈ N0 .
ρ
Thus
M
|ncn z n−1 | ≤ |cn ρn | for all n ∈ N0 ,
|z|
gives (a) because the series ∞ n
P
n=0 |cn ρ | converges. To conclude with (b) we
have to show that
∞
f (z + h) − f (z) X (z + h)n − z n
(∗) := − g(z) = − nz n−1 → 0 as h → 0,
h n=1
h
126
For z, z + h ∈ B(0, R) we must prove that
n
(z + h)n − z n n−1
X n k−2 n−k
− nz =h h z
h k=2
k
n−2
X n!
=h hr z n−2−r
r=0
(n − (r + 2))!(r + 2)!
Fix z and pick ρ with |z| < P ρ < R, so that |z| + |h| < ρ whenever |h| <
ρ − |z|. Using part (a) twice, ∞ n=2 n(n − 1)|cn |ρ
n−2
converges to a constant
0
independent of h. We conclude that f (z) does exist and equals g(z). The
proof of (c) follows wit (16.38). 2
P∞
Example 18.2 We know that the geometric series n=0 z n has radius of
convergence R = 1. Theorem 18.1 implies that
∞
−2 d X
(1 − z) = (1 − z)−1 = nz n−1 , |z| < 1.
dz n=1
1 X n
= z n−k , |z| < 1, k ∈ N.
(1 − z)k+1 n≥k k
127
Theorem 18.1 gives
f holomorphic ⇔ f analytic.
128
Example 18.6 (a) Let f (z) = sin2 (z) + cos2 (z) and g(z) = 1 defined for
all z ∈ C, both functions are holomorphic on B(0, R) for all R > 0. As
f (x) = g(x) for all x ∈ R we conclude with Theorem 18.5 that f (z) = g(z)
for all z ∈ C.
129
Proof. Let p(z) = cn z n + cn−1 z n−1 + · · · + c1 z + c0 with c1 , . . . , cn ∈ C and
1
cn 6= 0. Assume that p has no root. Then f : C → C, z 7→ f (z) = p(z) is
holomorphic on C. As |p(z)| → ∞ as |z| → ∞, there exists R > 0 such that
1
|f (z)| = | p(z) | < 1 for |z| > R. Hence, f is bounded and therefore constant
due to Liouville’s theorem 16.5. This gives the required contradiction. 2
19.1 Index
Definition 19.1 (Index) Let γ be a parametrisation of a closed path, and
consider the complement Ω := C \ γ ∗ (recall that γ ∗ = γ([a, b])). The index
of the curve (path) is the function Indγ : Ω → C defined by
Z
1 dw
Indγ (z) = , z ∈ Ω. (19.46)
2πi γ w − z
Without proof we cite the following theorem about the values of the index
function.
130
Figure 69: components of Ω of a path γ
131
Figure 70: three index functions
Is there a connection with the index? This is indeed the case, and we shall
show it later.
132
Figure 71: example with Indγ ((1 + i)) = 2
We shall explore the meaning of the condition that the complement of the
domain of definition lies outside of the curve (path). We do this with a couple
of examples. In figure 72 we have two circle lines around the origin, ∂B(0, R)
and ∂B(0, r), r < R , one circle line is lying inside the other one. Consider
the path γ as the union of γ1 , the outer circle line in positive direction, the
straight connection of the outer circle with the inner circle γ2 , the inner circle
line γ3 with negative orientation and the straight line back (on the same line
as before but in different direction) γ4 .
133
Figure 72: paths γ1 , γ2 , γ3 , γ4
but inside of the inner circle ∂B(0, r) we have points which do not belong to
the domain of definition and do not belong to the outside of the whole path
γ. A way out of this is the following. Pick ε > 0 and open both circle lines
by a small piece of length ε to get the key-hole path in figure 73.
134
Figure 73: key-hole path
The curves (paths) in figure 74 and 75 are encircling both two singularity
points. We color the two components around the two singularities by red (the
upper one) and by blue (the lower one). Let γ be the closed paths enclosing
(surrounding) the two points respectively- one (figure 74) puts the points on
the outside and one (figure 75) encircles the upper point for example in one
movement, see the figure 74 and 75 respectively. Is there a difference in the
methods of encircling? In which components is the index Indγ 6= 0 ? If we
consider the curve (path) as a fence keeping some animals we would see no
difference at all. But there is a difference.
135
Figure 74: enclosing two singularities - method (a)
In figure 74 both the red and blue component have zero index but in figure
75 the blue component has a non-zero index. This shows that for the curve
(path) γ in figure 74 the two singularities are lying outside γ.
136
Figure 75: enclosing two singularities - method (b)
This discussion shall motivate the following final version of Cauchy’s theorem.
137
19.2 Laurent series
We study now power expansions with negative powers. We start with the
following Binomial expansions derived from the geometric series.
Binomial expansions
For |z| < 1 the geometric series gives the expansion of (1 − z)−1 with positive
powers
∞
1 X
= zn.
1−z n=0
19.7 A series ∞
P
Definition
P∞ P∞an , an ∈ C, converges (to s = s1 + s2 ) if
n=−∞
n=0 an converges (to s1 ) and n=1 a−n converges (to s2 ). A power series
with negative and positive powers is called a Laurent series.
∞
X
f (z) = cn (z − a)n for z ∈ A, (19.47)
n=−∞
138
where the coefficients are given by
Z
1 f (w)
cn = dw, n ∈ Z, (19.48)
2πi γ (w − a)n+1
with γ ∗ = ∂B(a, r) for any r with R < r < S.
Pick z ∈ A and choose R < P < |z| < Q < S and consider the two closed
paths γ
e and γ e is the closed path enclosing z ∈ A and
b in figure 76, where γ
z ∈ O(bγ ). From (16.37) we get
Z
1 f (w)
f (z) = dw,
2πi γe w − z
and from Cauchy’s theorem 19.5 we have
Z
1 f (w)
0= dw.
2πi γb w − z
The integrals along the line segments connecting the two circlines ∂B(0, Q)
and ∂B(0, P ) in figure 75 cancel out each other and the remaining gives
Z Z
1 f (w) 1 f (w)
f (z) = dw − dw. (19.49)
2πi ∂B(0,Q) w − z 2πi ∂B(0,P ) w − z
139
We can expand the integrands (w − z)−1 in (19.49) in positive powers for
the first integral and in negative powers for the second integral in (19.49)
respectively (see Binomial expansions).
∞
1 1 1 1 X z n
= = for all w ∈ ∂B(0, Q)
w−z w 1− z w n=0 w
w
140
For the annulus A2 we write
1 1
f (z) = ,
(z − 1)3 1 + 1
z−1
Theorem 19.8 indicates that the isolated singularities are the candidates
for Laurent expansions. We need the following classification of isolated sin-
gularities.
141
The first term on the right hand side of (19.50) is called the principal part
of the Laurent expansion. The isolated singularity a is said to be
The example 1/(z − 1)2 has at z = 1 a pole of order 2, also called double
pole.
We are now able to answer our questions about when integrals of functions
along closed paths vanish or not. If we surround poles of the integrand the
integral does not vanish, because there is a reminder which is given by the
residue of that pole. This is the content of the last theorem.
Z N
X
f (z) dz = 2πi res{f ; ak }. (19.51)
γ k=1
n < −mk . Integration of each principal part gives (see fundamental integral)
mk
c(k)
Z X −n
fk (z) dz = = 2πic(k)
−1 = 2πi res{f ; ak }.
γ n=1
(z − ak )n
142
PN
The function g : I(γ) → C, given by g(z) = f (z)− k=1 fk (z), is holomorphic
and Theorem 19.5, i.e.,
Z Z N Z
X
0= g(z) dz = f (z) dz − fk (z) dz,
γ γ k=1 γ
Lemma 19.13 Let B(a, R) be the open ball around a ∈ C with radius R > 0,
and assume that the function f : B(a, R) → C is holomorphic on B(a, R) \
{a}.
res{f ; a} = g(a).
h(a)
res{f ; a} = .
k 0 (a)
g(z)
(c) Let f have a pole at a of order m > 1 and let f (z) = (z−a) m,z ∈
Proof. (a) This follows immediately from the the Laurent series expansion
of the function f .
(b)
h(z) z−a h(a)
lim (z − a) = h(a) lim = 0 .
z→a k(z) z→a k(z) − k(a) k (a)
143
(c) Cauchy’s formulae give
(m − 1)! (m − 1)!
Z Z
(m−1) g(z)
g (a) = m
dz = f (z) dz
2πi ∂B(a,R/2) (z − a) 2πi ∂B(a,R/2)
144