Complex Variables-LectureNotes
Complex Variables-LectureNotes
Lecture Notes
Victor Ivrii
Department of Mathematics,
University of Toronto
Contents i
i
Preface
These are Lecture Notes for MAT 334 “Complex Variables” at Faculty of
Arts and Science, University of Toronto. This is a junior class for all but
Math Specialist students.
I was teaching it for several years and the last time it at Fall of 2020.
This time the class was taught online due to COVID-19 pandemic and I
made beamer slides for lectures. These slides were reformatted to a book
format.
These Lecture Notes are addition rather than substitution for our stan-
dard textbook Complex Variables, 2nd Edition, by Stephen D. Fisher (re-
ferred as Textbook).
We cover Chapters 1–3 from Textbook, however some material removed
and other material added, exposition is different.
1
Chapter 1
Introduction to MAT334
We start a class called “Complex Variables” but more precisely it should be
called Functions of a Complex Variable and even more precisely Functions
of One Complex Variable.
In late 17-th century I. Newton diskovered that the decomposition of
functions of one real variable into power series (which we call today Taylor
series) is a very powerful tool:
∞
X
f (x) = an (x − x0 )n (1.1.1)
n=0
with
1 (n)
an = f (x0 ). (1.1.2)
n!
Starting from I. Newton mathematicians began to use this tool extensively
and for more than 100 years a function was something that is given by a
power series.
They did not care where this series converges (if it converges at all,
except for x = x0 ), and if it converges to f (x). Today we call these functions
analytic or, more precisely real analytic.
There were some problems with this approach: many functions are not
differentiable, even for infinitely differentiable functions series (1.1.1) may
not converge (except at x = x0 ), or converge not to f (x). Also comparison
of decompositions at different points was not very obvious. So, if a function
f (x) depends only on x, why we need to deal with x0 ?
2
Chapter 1. The Complex Plane 3
then one can plug instead of x a complex number z and get a function of
the complex variable
∞
X
f (z) = an (z − x0 )n (1.1.3)
n=0
Think differently!!
Think like a complex analyst!!
y
z
z1 + z2 = z 1 + z 2 ,
αz = αz for real α,
z = z.
z1 z2 = z2 z1 , (1.1.6)
z(z1 + z2 ) = zz1 + zz2 , (1.1.7)
z1 (z2 z3 ) = (z1 z2 )z3 , (1.1.8)
z1 z2 = z 1 z 2 . (1.1.9)
1 · 1 = 1, i · 1 = 1 · i = i, i · i = −1.
Chapter 1. The Complex Plane 6
z1 |z1 |
= , (1.1.15)
z2 |z2 |
z1
arg = arg(z1 ) − arg(z2 ). (1.1.16)
z2
Corollary 1.1.2. As z ̸= 0, n ∈ Z
|z n | = |z|n , (1.1.17)
arg(z n ) = n arg(z). (1.1.18)
Further,
′
(eiθ )′ = cos(θ) + i sin(θ) = − sin(θ) + i cos(θ) = i cos(θ) + i sin(θ)
= ieiθ
Thus
(eiθ )′ = ieiθ . (1.1.22)
Chapter 1. The Complex Plane 8
Properties (1.1.20) and (1.1.22) show that this function behaves like an
exponent, which justifies this definition.
Furthermore, if we define for λ = α + iβ, α, β ∈ R,
z n = w, w = ρeiφ . (1.1.28)
Re(nz) = C n = A + iB ̸= 0. (1.2.4)
n·z =C (1.2.5)
y ℓ
n
d x
Figure 1.1: d > 0 above dashed line, d < 0 below dashed line.
1.2.3 Circles
A circle with a center at p ∈ C and radius r ≥ 0 is defined by equation
|z − p| = r. (1.2.7)
Consider now two distinct points, p, q ∈ C and a curve which is the locus
of the points, satisfying
c 2 ρ2 |c|2
(1 − ρ2 ) w − =
1 − ρ2 1 − ρ2
c ρ|c|
⇐⇒ w − 2
= ;
1−ρ 1 − ρ2
ρ|c| c
Thus, w is on the circle of radius R = 1−ρ 2 centered at the point w0 = 1−ρ2
and so z lies on the circle of the same radius R centered at the point
p ρ2 q
z0 = − . (1.2.9)
1 − ρ2 1 − ρ2
Point z0 belongs to a straight line passing through p and q, on the
ρ2 |c|
distance 1−ρ 2 from p and further from q than p. Recall that |z − p| = ρ|z − q|
p 2
ρ q
and z0 = 1−ρ 2 − 1−ρ2 . When ρ ↘ 0, circles became smaller and smaller
and shrink to p, and when ρ ↗ 1 then circles become larger and larger,
approximating L and their centers go to the right infinity (in our picture),
when ρ jumps over 1 we get circles on the other side, and their centers far
away to the left, and when ρ ↗ ∞ these circles shrink to q and their centers
move to from the left to q.
Chapter 1. The Complex Plane 12
q p L′
q p L′
circle is orthogonal to each blue circle at the point of intersection (two lines
are orthogonal at the point of intersection if their tangent lines at this point
are orthogonal).
See the proof on pp. 18–20 of the Textbook.
Remark 1.2.1. (a) In this class we will encounter many mutually orthogo-
nal families of lines.
(b) We see that straight lines could be also included in the family of circles.
We will see that a straight line is a circle, passing from the infinity
(the exact meaning will be explained later).
Translations.
C ∋ z 7→ w = z + a ∈ C, a ∈ C, (1.2.10)
Scalings.
C ∋ z 7→ w = λz ∈ C, λ > 0, (1.2.11)
Rotations.
C ∋ z 7→ w = zeiφ ∈ C, φ ∈ R, (1.2.12)
Inversion.
it is called an inversion.
Later we will prove that it transforms circles and straight lines into
circles and straight lines (but some straight lines can become circles and
some circles can become straight lines–think, which are suspect!)
We can solve easier problems today:
z2
w3 z1
Show that a straight line passing through
0 (with an angle φ) becomes the straight z3
line passing through 0 (with an angle −φ). w2
w1
Therefore
1 (1 + e−it )
z = r(1 + eit ) =⇒ w = =
r(1 + eit ) r(1 + eit )(1 + e−it )
(1 + e−it ) (1 + cos(t) − i sin(t)) 1 i sin(t)
= = = −
2r(1 + cos(t)) 2r(1 + cos(t)) 2r 2r(1 + cos(t))
1
and Re(w) = 2r
and
sin(t) 1
Im(w) = − = − tan(t/2)
2r(1 + cos(t)) 2r
(check it!) and when t runs from 0 to ±π, Im(w) runs from 0 to ∓∞.
So, we got a vertical straight line {z : Re(z) = 1/(2r)}.
Since inversion is self-inverse we conclude also that a straight line, not
passing through 0 becomes a circle passing through 0.
r(1 + eit )
0 x
r(1 + e−it )
1
2r (1 − i tan( 2t )
1
Figure 1.2: r = 2
Solution. Note that (1.2.14) does not change under translations, rotations
and even scalings. So without any lost of generality one can assume that
p = r and q = −r (shifting their midpoint to 0 and then rotating). Assume
that 0 < φ < π (otherwise we permute p and q).
Also note that arg(z − p) is an angle between vector y
⃗ and the positive direction of axis x, and arg(z − q)
pz
is an angle between vector qz
⃗ nd the positive direction z
of axis x. Therefore (1.2.14) means that
φ
z−p
φ = arg =
z−q
arg(z − p) − arg(z − q) = ∠(qzp)
z z
Dε (z) Dε (z)
Remark 1.3.1. The notion of an open set depends where we consider it (and
the same will be true for the notions of the interior of a set, a closed set, a
closure of the set). For example Open interval (a, b) ⊂ R is an open subset
of R but not of C and its interior in C is empty:
(c) This set is not connected (d) This set is not connected
Domains are the natural setting for the study of analytic and harmonic
functions.
Obviously, any convex open set is also starlike, and any starlike set is
also connected.
(a) Convex set (b) Starlike but not con- (c) Connected but not star-
vex set like set
K′
K′
N
Chapter 1. The Complex Plane 21
1 − |z|2
1+z (1 + z)(1 − z)
Re = Re =
1−z (1 − z)(1 − z) |1 − z|2
1.4.2 Graphs
We would like to plot functions of complex variables, but this is not really
possible. Indeed, the plot of a real-valued function of a real variable is a
two-dimensional picture.
The plot of a real-valued function of the complex variable (that is of two
real variables) is a three-dimensional picture which could be plotted.
But the plot of a complex-valued function of a complex variable is four
dimensional picture, so we draw domain and range. F.e. in Example 1.4.2
we show a picture like this:
Chapter 1. The Complex Plane 23
1+z
w=
1−z
1.4.3 Limits
The notions of limits and continuity is due to the same notions from Calculus
I and Calculus II.
Definition 1.4.2. (a) A sequence of complex numbers {zn }∞
n=1 converges
to A (equivalently, has a limit A),
zn → A as n → ∞ or lim zn = A
n→∞
if for any ε > 0 there exists N = N (ε) such that n ≥ N =⇒ |zn −A| <
ε.
(b) Otherwise we say that {zn }∞
n=1 diverges.
Definition 1.4.3. Consider f (z) in the domain D and let z0 belong to the
closure of M , that is is either in M or on its boundary. We say that f (z)
tends to L as z tends to z0 , or f (z) → L as z → z0 , or f (z) has a limit L
at z0 , or
lim f (z) = L
z→z0
lim f (z) = L
z→∞
(ii) f (z)n → A, |f (z)| → |A|, Re(f (z)) → Re(A), Im(f (z)) → Im(A) as
z → z0 ;
f (z) A
(iii) If B ̸= 0 then → .
g(z) B
1.4.4 Continuity
Definition 1.4.5. Function f (z) is continuous at point z0 if f is defined in
z0 and
Theorem 1.4.3 (Theorem 3 from page 37 of the Textbook). Let f (z) and
g(z) be continuous at z0 , α ∈ C. Then
(a) It is convergent
PN and its sum its sum is S if the sequence of partials
sums SN = n−1 an converges to S.
(b) It is divergent if the sequence of partial sums diverges.
See Examples on pages 38–40 of the Textbook.
The following theorem is proven the same way as in Calculus I:
Theorem 1.4.5. Let series ∞
P
n=1 an converges absolutely, that is
∞
X
|an | < ∞.
n=1
Then it converges.
Power series will play a crucial role in this class.
These functions will coincide with the standard functions of real variable ex ,
cosh(x), sinh(x), cos(x) and sin(x) when z = x ∈ R, and satisfy the same
identities. It will turn out later that these functions are analytic.
We also introduce the inverse functions log(z), arccos(z) and arcsin(z),
which will be multi-valued functions of complex variable z, and their single-
valued variants Log(z), Arccos(z) and Arcsin(z). These functions also
coincide with the corresponding functions of the real variable when z = x ∈ R
and it will turn out later that these functions are analytic.
ez+w = ez ew (1.5.1)
ez+2πi = ez (1.5.3)
0 ≤ Im(z) < 2π
(a) z (b) w = ez
Remark 1.5.1. (a) When x in the horizontal lines runs from (−∞, ∞),
|w| runs (0, ∞) (so rays are outward),
(b) When y in the vertical segments runs from 0 to 2π, arg(w) runs
(0, 2π) (so circles go counter-clockwise).
(c) One can see that half-strip {z : 0 ≤ Im(z) < 2π, Re(z) < 0} is
mapped to the unit disk with a punched center {0 < w : |w| < 1}, while
half-strip {z : 0 ≤ Im(z) < 2π, Re(z) > 0} is mapped to the exterior of the
unit disk {w : |w| > 1}.
https://mathworld.wolfram.com/WeierstrassEllipticFunction.html
(it is way too advanced for our class, we are not covering it, but look at this
beauty).
and take Arg(z) ∈ (−π, π) (we are opportunistic here and change the
definition as we see fit!); then Log(z) maps C \ (−∞, 0] onto strip {w : − π <
Im(w) < π}.
We selected a branch Arg(z) of arg(z). Two different sides of the cut
go to two different horizontal straight lines. Cuts always have two sides and
in mappings they usually separate.
Chapter 1. The Complex Plane 30
(a) z (b) w = z α
Again different sides of the cut are mapped onto different pieces of the
boundary. Blue rays and red arcs on both pictures have same orientations.
Chapter 1. The Complex Plane 31
Remark 1.5.2. (a) Case α > 1 is reduced to this one since z α is inverse to
z 1/α . Then picture (a) and (b) interchange.
(b) One can consider −1 < α < 0, pictures remain the same, except the
sector now is {w : | Arg(w)| < −απ} and blue rays and red arcs on
both pictures have opposite orientations.
(a) The disk with a cut {z : |z| < R, | Arg(z)| < π},
(b) The exterior of the disk with a cut {z : |z| > r, | Arg(z)| < π},
(c) The ring with the cut {z : r < |z| < R, | Arg(z)| < π}.
eiz + e−iz
cos(z) = ,
2
and
eiz − e−iz
sin(z) =
2i
because we have these formulas for real z.
Let also
sin(z) cos(z)
tan(z) = , cot(z) = ,
cos(z) sin(z)
and
1 1
sec(z) = csc(z) =
cos(z) sin(z)
(b) Horizontal segments {z : 0 < x = Re(z) < 2π, y = Im(z) = const} are
mapped to w = u + iv with v = − sin(x) sinh(y), u = cos(x) cosh(y)
which are the same ellipses as before.
The same is true for sin(z), except strips are shifted and flipped due to
sin(z) = cos( π2 − z).
sin(z)
Finally, tan(z) = cos(z) gives us a picture of Circles of Appolonius with
= i and q = −i, but explanation why it happens and the discussion of
Chapter 1. The Complex Plane 34
-i
Therefore
√
arccos(w) = − i log w ± 1 − w2 .
Similarly
√
arcsin(w) = − i log iw ± 1 − w2
Chapter 1. The Complex Plane 35
and
i 1 − iw
arctan(w) = log .
2 1 + iw
In particular, range of cos(z) and sin(z) in C (all values are achieved), and
range of tan(z) is C \ {i, −i} (all values except ±i are achieved).
For single-valued functions we have formulas
√
Arccos(w) = − i log w ± 1 − w2 ,
√
Arcsin(w) = − i log iw + 1 − w2
and
i 1 − iw
Arctan(w) = Log
2 1 + iw
√
where on the domain (with cuts removed) Re( 1 − w2 ) > 0,
which defines branch uniquely.
their properties and properties and formulas for inverse hyperbolic functions
follow from the properties of trigonometric functions.
Example 1.5.1. Range of tanh(z) is C \ {1, −1}.
(c) The curve γ is closed if γ(a) = γ(b), that is the initial point γ(a)
coincides with the end-point γ(b).
Remark 1.6.1. The famous Jordan Curve Theorem asserts that the comple-
ment of the range of a curve, which is simple and closed, consists of two
disjoint open connected sets, one bounded and the other unbounded.
The bounded piece is the inside of the curve and the unbounded piece
the outside.
Despite the almost painful obviousness of this statement, the theorem is
hard to prove. We shall accept it as true.
outside
inside
Remark 1.6.2. (a) It is very common and convenient to refer to the range
of γ(t) as the curve γ and to γ(t) itself as the parametrization of the
curve.
(b) With this use of the word curve, a curve becomes a concrete geometric
object such as a circle or a straight line segment and hence is easily
Chapter 1. The Complex Plane 38
visualized. The difficulty with this view is that a particular curve has
many different parameterizations. However, our results under very
broad assumptions would not depend on parametrizations.
(c) What is more, for closed curves the results would not depend on the
choice of start-point (which is also an end-point).
(d) We extend the notion of the curve, requiring γ(t) to be only piecewise
continuous, that is, consisting of several curves in the old understand-
ing.
θ1 θ0
c
z1
z0
(c) Also
Z Z Z
f (z) dz = f (z) dz + f (z) dz
γ1 +γ2 γ1 γ2
(e) Integral over closed contour does not depend which point is start- and
end-point. Indeed, if such points are a and b then the path with start-
and end-point a is γ = γ1 + γ2 , and the path with start- and end-point
b is γ ′ = γ2 + γ1 :
γ1 b
γ2
I
(f) For integral over closed contour a special notation often is used:
γ
But recall: orientation matters!
We will use the following important inequality, which follows from the
standard properties of integral :
Theorem 1.6.1.
Z Z
| f (z) dz| ≤ |f (z)| |dz|.
γ γ
Chapter 1. The Complex Plane 41
D γ1
Why? Let us make an infinitely thin cut (if N > 1 we need N cuts)
so that domain D∗ after cuts will be connected and simply-connected. Its
boundary γ ∗ = γ0 + ℓ + ℓ′ + γ1 is connected and the counter-clockwise
orientation of γ0 implies clockwise orientation of γ1 (and γ2 , . . . , γN ) and
the proper orientations of ℓ′ and ℓ.
Chapter 1. The Complex Plane 42
γ0 γ0
ℓ′ ℓ
γ1 γ1
D∗ D
∗ ∗
RR RR
Applying to D , γ Green’s formula we see that D ∗ = D
, and
′
H R R R R R R H
γ∗
= γ0
+ ℓ
+ ℓ′
+ =
γ1 R γ0R
+ γ1
= γ
because ℓ = −ℓ (have opposite
directions) and therefore ℓ + ℓ′ = 0, so corresponding integrals cancel one
another.
(b) Let
∂N ∂M
− = 0. (1.6.2)
∂x ∂y
Then if a closed curve γ ′ is properly oriented and bounds a domain
D′ ⊂ D then
I ZZ
∂N ∂M
M dx + N dy = − dxdy = 0. (1.6.3)
γ′ D′ ∂x ∂y
(c) In simple connected domains every simple closed curve bounds some
subdomain; it is not so in domains which are not simply-connected:
(d) Therefore in simply-connected domains (1.6.2) implies that
I
M dx + N dy = 0 (1.6.4)
γ
(we slightly cut corners here, since there could be an infinite number
of self-intersections).
(f) Then integral over curve which is not closed depends only on it’s start
and end-points:
γ2
γ1
Therefore
x dy − y dx
= d θ, θ = arg(x + yi);
x2 + y 2
x dy − y dx
I
= 2π
γ x2 + y 2
Remark 1.6.6. Why these notations (1.6.8) and (1.6.9)? Let us write
∂f ∂f
df = dx + dy. (1.6.10)
∂x ∂y
On the other hand dz = dx + idy, dz = dx − idy imply that dx = 12 (dz + dz)
and dy = − 2i (dz − dz) and therefore
1 ∂f ∂f 1 ∂f ∂f
df = −i dz + +i dz
2 ∂x ∂y 2 ∂x ∂y
and in these notations we extend the usual formula to complex variables:
∂f ∂f
df = dz + dz (1.6.11)
∂z ∂z
∂f ∂f
even if ∂z
and ∂z
are not partial derivatives in the sense of Calculus II.
Chapter 1. The Complex Plane 46
Corollary 1.6.5. I ZZ
∂f
f dz = 2i dxdy, (1.6.12)
γ D ∂z
where D is a bounded domain in C and γ = ∂D is it’s border, properly
oriented.
∂f ∂f
df = dz + dz. (2.1.1)
∂z ∂z
with
∂f 1 ∂f ∂f
:= +i , (2.1.2)
∂z 2 ∂x ∂y
∂f 1 ∂f ∂f
:= −i . (2.1.3)
∂z 2 ∂x ∂y
Looks like the usual formula from Calculus II?
∂f ∂f
df = dx + dy. (2.1.4)
∂x ∂y
47
Chapter 2. Basic Properties of Analytic Functions 48
Lemma 2.1.1. dz and dz are linearly independent: for any complex num-
bers A and B
A dz + B z = 0 ∀dz =⇒ A = B = 0.
(A + B) dx + i(A − B) dy = 0 ∀dx, dy
Corollary 2.1.2.
df = A dz + Bdz ∀dz
∂f ∂f
implies that A = ∂z
, B= ∂z
defined by (2.1.2)–((2.1.3).
(b) Then we call A(z) derivative of analytic f (z) function f (z) and denote
df
it by f ′ (z) and dz .
Then we have:
Chapter 2. Basic Properties of Analytic Functions 49
∂f 1 ∂f ∂f
:= +i = 0, (2.1.6)
∂z 2 ∂x ∂y
df ∂f
f ′ (z) = := . (2.1.7)
dz ∂z
Remark 2.1.1. It is equivalent to definition on page 77 of the Textbook:
Example 2.1.4. (a) cos(z), sin(z), cosh(z) and sinh(z) are entire analytic
functions and usual expression for derivatives hold.
(b) tan(z), sec(z) are analytic functions except when cos(z) = 0 (so except
{z = π2 + πn, n ∈ Z}) and usual expression for derivatives hold.
(c) cot(z), csc(z) are analytic functions except when sin(z) = 0 (so except
{z = πn, n ∈ Z} and usual expression for derivatives hold.
Chapter 2. Basic Properties of Analytic Functions 51
Remark 2.1.3. In these examples we can make a cut along any ray from 0
to ∞ and select a corresponding branch.
Later we will prove
∂v ∂u ∂v ∂u
= M := − , = N := .
∂x ∂y ∂y ∂x
It immediately implies
Case 1. Series
X
S(z) := an (z − z0 )n (2.2.1)
n=1
Case 3. Series (2.2.1) converges for some z ̸= z0 but not for all of them.
Explore Case 3. So, there are z1 ̸= z0 for each series (2.2.1) converges, and
z2 ̸= z0 for which it diverges.
By Theorem 2.2.1 it is impossible that |z1 − z0 | > |z2 − z0 | and therefore
|z1 − z0 | ≤ |z2 − z0 |. Therefore, there exists R : 0 < R < ∞ such that
(b) In Case 2 (series converges for all z ∈ C only) the radius of convergence
is R = ∞ and the disk of convergence is C.
(c) In Case 3 the radius of convergence is R (0 < R < ∞) and the disk of
convergence is {z : |z − z0 | < R}.
Theorem 2.2.2. Consider a power series (2.2.1) with the radius of conver-
gence R.
Chapter 2. Basic Properties of Analytic Functions 57
1 |an+1 |
= lim . (2.2.2)
R n→∞ |an |
Proof. Proof follows from the ratio test and root test for numerical series.
∞
X (z − 1)n
Example 2.2.2. (a) The radius of convergence of is R =
n=0
n!
∞ and it converges in C.
∞
X n(z + 1)n
(b) The radius of convergence of is R = 5 and the disk
n=0
5n
of convergence is {z : |z + 1| < 5}.
∞
X
(c) The radius of convergence of n!!(z − i)n is R = 0 and it
n=0
converges only in {i}.
∞
X zn P∞ 1
(b) converges for all z : |z| = 1 ( n=1 n2 < ∞).
n=1
n2
The last part of this answer is important, because it shows, that you
understand, that the boundary of the disk is a circle, not just two points.
Then
∞
f (z + h) − f (z) −2
X
| − g(z)| ≤ |h|δ |a|n (R − δ)n
h n=2
(ii) In particular,
∞
1 X
(b) Integrating 2
= (−1)n z 2n , converging as |z| < 1, we get
1+z n=0
∞
X (−1)n z 2n+1
Arctan(z) =
n=0
2n + 1
Chapter 2. Basic Properties of Analytic Functions 61
α(α + 1) 2
(1 − z)−α = 1 + αz + z + ...
2
α(α + 1) · · · (α + n − 1) n
+ z + . . . (2.2.9)
n!
Unless α is a non-positive integer we get an infinite series. In particular,
1
plugging α = 12 and z := z 2 we get a decomposition for √ .
1 − z2
Integrating, we get a decomposition for Arcsin(z).
Remark 2.2.1. One can prove that
(a) In Series (2.2.9) coefficient an has a magnitude nα−1 and therefore it
converges absolutely on C = {z : |z| = 1} for α < 0; diverges on C for
α ≥ 1 and converges on C except z = 1 for 0 < α < 1;
(b) Series for Log(1 − z) converges on C except z = 1;
(c) Series for Arctan(z) converges on C except z = ±i;
(d) Series for Arcsin(z) converges on C but not absolutely.
Results “converges but not absolutely” are more difficult and beyond our
reach.
also converges in D.
Chapter 2. Basic Properties of Analytic Functions 62
∂f
and since = 0 is our definition of the analyticity, this is 0.
∂z
Remark 2.3.1. We also proved in Subsection 1.6.4 “Green’s Formula: Dis-
cussion” that
- formula (1.6.7) holds in the case when γ consists of several pieces,
provided they are properly oriented
- and if the integrand in the double integral is 0 then the orientation
does not matter,
Chapter 2. Basic Properties of Analytic Functions 63
(we did it for Green’s formula in real variables, but Green’s formula in
Complex variables simply inherits these properties).
Still, read Theorem 2 of Section 2.3 in the Textbook (pages 108–109).
Figure 2.1: Green lines are deformations of blue lines, but red lines are not.
where the right-hand expression means integral over any curve γ, with a
start-point z0 (some fixed point in D) and an end-point z.
This is justified since the right-hand expression does not depend on the
choice of γ with a start-point z0 (a fixed point in D) and an end-point z.
Indeed, if γ1 and γ2 are tw such points, then
Z Z Z
f (z) dz − f (z) dz = f (z) dz = 0
γ1 γ2 γ1 −γ2
1
Example 2.3.2. (a) If f (z) = and γ is a circular curve, centered at 0
z
with radius r, with counter-clockwise orientation, then
Z Z Z 2π
dz
f (z) dz = = idθ = 2πi ̸= 0
γ γ z 0
Chapter 2. Basic Properties of Analytic Functions 66
with
1 1 1 1
= =
ζ −z (ζ − z0 ) − (z − z0 ) (ζ − z0 ) 1 − z − z0
h i
ζ − z0
and decompose
∞
X z − z0 n
1
z − z0 =
1− n=0
ζ − z0
ζ − z0
z − z0
which converges as |z − z0 | < r and |ζ − z0 | ≥ r and therefore < 1.
ζ − z0
∞ Z
X 1 f (ζ) dζ
f (z) = n+1
×(z − z0 )n .
n=0
2πi γ (ζ − z0 )
=an
Example 2.4.1.
∞ ∞ ∞
z
X zn X z 2n X z 2n+1
e = , cosh(z) = , sinh(z) = ,
n=0
n! n=0
(2n)! n=0
(2n + 1)!
∞ n 2n ∞
X (−1) z X (−1)n z 2n+1
cos(z) = , sin(z) =
n=0
(2n)! n=0
(2n + 1)!
have R = ∞.
Chapter 2. Basic Properties of Analytic Functions 68
(b)
∞ ∞
1 X X (−1)n z 2n+1
= (−1)n z 2n , Arctan(z) =
1 + z2 n=0 n=0
2n + 1
for any r < dist(z, ∂D), which is the distance from z to ∂D–the boundary
of D.
Proof. Immediately from Theorem 2.3.1 and equality f (n) (z0 ) = n!an (where
in the last moment we plug z instead of z0 ). Indeed, since the radius of
convergence is at least R = dist(z, ∂D), then |an | ≤ C(r)r−n for r < R.
Remark 2.4.2. (a) If we consider functions of one real variable, then exis-
tence of n derivatives does not imply existence of (n + 1) derivatives.
(b) f (n) (x) does not satisfy (2.4.3) or any other inequality. In fact, for any
sequence bn there exists an infinitely smooth function f (x) such that
f (n) (x0 ) = bn .
lim an Rn = 0. (2.4.4)
n→∞
(b) for Re(α) ≥ 1 it does not converge at any point z : |z| = 1 (common
term does not tend to 0)
Then f (z) = 0 in D.
Chapter 2. Basic Properties of Analytic Functions 71
z0
z1
√
Example 2.4.7. Consider f (z) = z as Re(z) > 0.
Using the same method we can expand it to the red disk. Then to blue,
and to brown, and to green
√ but where brown and green intersect, we got
two different branches of z!
We know that not all the coefficients here are 0 (otherwise f (z) would
be identically 0), so there is an integer m, such that
a0 = a1 = . . . = am−1 = 0, am ̸= 0
m m+1
⇐⇒ f (z) = am (z − z0 ) + am+1 (z − z0 ) + ..., am ̸= 0
(k) (m)
⇐⇒ f (z0 ) = 0 k = 0, . . . , m − 1, but f (z0 ) ̸= 0.
Definition 2.4.1. (a) In this case we say that z0 is a zero of f (z) and m
is an order of zero.
(b) Simple zeroes are zeroes of order 1, double zeroes are zeroes of order 2,
triple zeroes are zeroes of order 3, and so on.
for every triangle γ that lies, together with its interior, in D, then f is
analytic on in D.
We skip the proof, see proof of Theorem 2 on page 129 of the Textbook.
Chapter 2. Basic Properties of Analytic Functions 74
|F (z)| + |F (0)| 2M
|g(z)| ≤ ≤ .
R R
For any ζ ∈ C let us take R > 2|ζ|. By Cauchy’s formula
Z
1 g(z) dz
g(ζ) = .
2πi |z|=R z − ζ
Then
1 2M 2πR
|g(ζ)| ≤ →0 as R → ∞.
2π R R − |ζ|
with integral taken along any piecewise smooth curve from z0 to z; since
f ′ (w)
f (w)
is an analytic in D (think, why) this integral does not depend on the
curve.
′
Then h′ = ff and
Then e−h(z) f (z) = c with non-zero constant c and finally f (z) = ceh(z) .
Since h(z0 ) = 0 we have c = Log(f (z0 )) and g(z) = h(z) − Log(f (z0 ))
we proved
eg(z) = f (z), z ∈ D.
Then the product h(z) = f (z)g(z) is analytic in the disk {z : |z| < R}.
Hence, by Theorem 2.4.1 , h has a power series expansion in this same disk
∞
X h(n) (0)
h(z) = cn z n with cn = .
n=0
n!
But
f (k) (0) g (n−k) (0)
= ak , = bn−k
k! k!
and therefore
n
X
ck = ak bn−k .
k=0
As these examples show, there are three possible modes of behaviour for
f (z) when 0 < |z − z0 | < r:
(a) |f (z)| remains bounded as z → z0 ; then we call z0 a removable singu-
larity.
(b) lim |f (z)| = ∞; then we call z0 a pole.
z→z0
√ √
(b) z0 = 0 and f (z) = z, or 3 z, or z α with α ∈ / Z because it cannot
be defined as a single-valued analytic function near 0. It is called a
branching point.
Chapter 2. Basic Properties of Analytic Functions 78
g(z) = b0 + b1 (z − z0 ) + b2 (z − z0 )2 + b3 (z − z0 )3 + . . .
g(z) = b2 (z − z0 )2 + b3 (z − z0 )3 + . . .
=⇒ f (z) = b2 + b3 (z − z0 ) + . . . (2.5.1)
Remark 2.5.1. Compare with Calculus I: if f (x) is bounded it does not mean
that it has a limit, and even if it has a limit, it’s derivative could be really
bad:
1
Example 2.5.3. 1. f (x) = sin is bounded but has no limit as x → 0.
x
1
However in complex variables f (z) = sin has an essential singularity
x
at 0.
2.5.3 Poles
Pole are genuine singularities, but we can “tame” them. Assume that
1
Indeed, if needed we can reduce r. Then g(z) = is analytic and
f (z)
bounded in the punctured disk {z : 0 < |z − z0 | < r} and therefore for g(z)
point z0 is a removable singularity.
So we can define g(z0 ) = limz→z0 g(z) = 0 (because of (2.5.2)) and g(z)
becomes analytic in the whole disk {z : |z − z0 | < r}.
However, since g(z0 ) = 0 point z0 is also a zero of g(z). Let m = 1, 2, . . .
be an order of zero. Then
g(z) = (z − z0 )m h(z)
1
Therefore H(z) = is analytic in {z : |z − z0 | < r} and H(z0 ) ̸= 0.
h(z)
Then
1 1 H(z)
f (z) = = m
= .
g(z) (z − z0 ) h(z) (z − z0 )m
Since
H(z) = b0 + b1 (z − z0 ) + b2 (z − z0 )2 + . . . , b0 ̸= 0
we conclude that
f (z) = a−m (z − z0 )−m + a−m+1 (z − z0 )−m+1 + a−m+2 (z − z0 )−m+2 + . . .
(2.5.4)
with an = bn+m and a−m ̸= 0.
Remark 2.5.2. Decomposition (2.5.4) is an example of what we call later
Laurent’s series.
1
We see that f (z) has a pole at z0 iff and only if g(z) = (with
f (z)
g(z0 ) = 0) has a zero at z0 .
Definition 2.5.2. (a) We say, that f (z) has a pole of order (multiplicity)
1
m at z0 if g(z) = has a zero of order (multiplicity) m at z0 .
f (z)
(b) Poles of multiplicity 1 are called simple poles, poles of multiplicity 2
are called double poles, poles of multiplicity 3 are called triple poles,
and so on.
We the got the following
Theorem 2.5.1. f (z) has a pole of order m at z0 if and only if decomposition
(2.5.4) holds with a−m ̸= 0 holds in {z : 0 < |z − z0 < r}.
Example 2.5.4. (a) cot(z) has simple poles at zn = πn.
(b) cot2 (z) has double poles at zn = πn.
(c) z cot2 (z) has double poles at zn = πn, n =
̸ 0 and a simple pole at
z0 = 0.
(d) z −1 cot(z) has simple poles at zn = πn, n ̸= 0 and a double pole at
z0 = 0.
Chapter 2. Basic Properties of Analytic Functions 81
z0
z
γ1 γ2
z0
Observe that
∞ Z
X 1
- f (ζ)(ζ − z0 )n dζ × (z − z0 )−n−1
n=0
2πi γ1
−1 Z
X 1 f (ζ)
= m+1
dζ × (z − z0 )m
m=−∞
2πi -γ1 (ζ − z0 )
We can rewrite it as
∞
X
f (z) = an (z − z0 )n (2.5.5)
n=−∞
with Z
1 f (ζ) dζ
an = (2.5.6)
2πi γ (ζ − z0 )n+1
because due to Cauchy’s theorem integrals over γ, γ2 and −γ1 are equal.
−γ1 γ2
Chapter 2. Basic Properties of Analytic Functions 85
∞
1 X (−1)n
(b) cos = 2n
;
z n=0
(2n)!z
∞
1 X (−1)n
(c) sin = .
z n=0
(2n + 1)!z 2n+1
1
= exp z + z1 .
(b) z = 0 is an essential singularity for f (z) = exp(z) exp
z
It’s decomposition is
∞ ∞ ∞
X zm X zn X
. = bp z p
m=0
m! n=0
n! p=−∞
with
X 1
bp = .
n≥0,m : n−m=p
n!m!
2.5.7 Singularities at ∞
Assume that ∞ ∈ D ⊂ C;b that means that ∞ has a neighbourhood contained
in D; in other words D ⊃ {z : |z| > r} for sufficiently large r. Like in the
case z0 ∈ C we have Definitions:
Definition 2.5.4. Let f (z) be an analytic function in {z : |z| > r}. Then
∞ is an isolated singularity of f (z).
Definition 2.5.5. (a) ∞ is a removable singularity if |f (z)| ≤ M for
z : |z| > r with sufficiently large M , r;
Chapter 2. Basic Properties of Analytic Functions 87
√
(b) z0 = ∞ and f (z) = z because it cannot be defined as a single-valued
analytic function near ∞. It is called a branching point.
analytic in {w : 0 < |w| < r−1 } and type of singularity of g(w) at 0 coincides
with a type of singularity of f (z) at ∞.
Since
∞
X
g(w) = a−n wn
n=−∞
(b) ez , sin(z), cos(z) have infinite number of positive powers in their decom-
positions; and we again conclude that they have essential singularity
at infinity.
(c) Assume that all singularities of f (z) on C b are isolated and they
are poles. Then it can have only a finite number of singularities
z1 , . . . zN , which are poles of multiplicities m1 , . . . , mN . Then g(z) =
(z − z1 )m1 · · · (z − zn )mn f (z) is an entire analytic function and since
we assumed that ∞ is also a pole, we have |g(z) ≤ C(|z| + 1)M . Then
it is a polynomial P (z) and
P (z)
f (z) = Q(z) = (z − z1 )m1 · · · (z − zn )mn
Q(z)
is a rational function.
Conversely, any rational function has all singularities on C
b isolated,
and they are poles.
Then coefficient a−1 with the opposite sign is called the residue of f (z) at
∞ and denoted as Res(f ; ∞):
Res(f ; ∞) = −a−1 . (2.5.9)
Remark 2.5.4. (a) Residue at non-isolated singularity is not defined.
(b) If z0 ̸= ∞ is a removable singularity of f (z), then Res(f, z0 ) = 0.
(c) On the contrary, even if ∞ is a removable singularity of f (z), it may
happen that Res(f, ∞) ̸= 0.
One can ask, why these definitions? Why we select coefficient a−1 among
others? And why we take an opposite sign at a−1 at ∞?
The answer to the first two questions follows from the equality:
Z
f (z) dz = 2πia−1 = 2πi Res(f ; z0 ) (2.5.10)
γ
γ
z0
(a) z0 ∈ C (b) z0 = ∞
g(z)
f (z) = , g(z0 ) ̸= 0, h(z0 ) = 0, h′ (z0 ) ̸= 0, (2.5.15)
h(z)
and g(z), h(z) analytic at z0 . Then
g(z0 )
Res(f ; z0 ) = ′ . (2.5.16)
h (z0 )
g(z0 ) 1 + b1 (z − z0 ) + . . .
f (z) = ′
h (z0 )(z −
z0 ) 1 + c1 (z − z0 ) + . . .
g(z0 )
= ′ 1 + d1 (z − z0 ) + . . .
h (z0 )(z − z0 )
g(z0 )
= ′ (z − z0 )−1 + a0 + a1 (z − z0 ) + . . .
h (z0 )
1 + b1 (z − z0 ) + . . .
F (z) :=
1 + c1 (z − z0 ) + . . .
H(z)
f (z) = (2.5.17)
(z − z0 )m
Chapter 2. Basic Properties of Analytic Functions 92
Since cot(z) is odd, cot2 (z) is even and the required residue is 0 due to Hint
(B).
Example 2.5.14. Calculate Res(tan3 (z); z = πn + π2 ), n ∈ Z. Using Hint (A),
π π
Res(tan3 (z); z = πn + ) = Res(tan3 (z + + πn); z = 0)
2 2
= Res(cot3 (z); z = 0).
Then
z2
3 z2
3
3 cos3 (z) 1− 2
+ ... 1 1− 2
+ ...
cot (z) = 3 = 3 = 3 3
sin (z) z3 z 1− z2
z− 6
+ ... 6
+ ...
where . . . denote terms which do not contribute to term with z −1 in the
final decomposition (we need to get no more than z 2 in the numerator and
z 3 in denominator); therefore
3z 2 3z 2
cot3 (z) =z −3 1 − + . . . = z −3 − z −1 + . . .
+ ... 1 +
2 6
and the required residue is −1.
z1
z3
z2
and we arrive to
Z N
X
f (z) dz = 2πi Res(f ; zk ). (2.6.1)
γ k=1
z1
z3
z2
assuming that
(a) m ≤ n − 2,
Proof. We want to apply Theorem 2.6.1 and for this we need to take a
+
bounded domain. We take a half-disk DR = {z : |z| < R, Im(z) > 0} which
contains all roots of Q(z) in the upper half-plane.
−R R
+
Then, according to Theorem 2.6.1, with ΓR the boundary of DR
Z m
P (z) X P (z)
dz = 2πi Res ; zk (2.6.4)
ΓR Q(z) k=1
Q(z)
for sufficiently large |z|. If you cannot prove it by yourself, look at pages
154–155 of the Textbook.
P (z)
Therefore, Q(z) ≤ M1 Rm−n on γR and
Z
P (z)
| dz| ≤ πR × M1 Rn−m → 0 as R → ∞
γR Q(z)
P (x)
≤ M2 (|x| + 1)m−n ∀x ∈ R.
Q(x)
Indeed, it is so for |x| ≥ c and for |x| ≤ c it follows from the assumption
that Q(x)Rdoes not have real roots.
∞
Since −∞ M2 (|x| + 1)m−n dx < ∞ again because m − n + 1 < 0 we
conclude that
Z R Z ∞
P (x) P (x)
dx → dx as R → ∞
−R Q(x) −∞ Q(x)
Remark 2.6.2. Using lower half-disk D′ = {z : |z| < R, Im(z) < 0} one
could get
Z ∞ r
P (x) X P (z)
dx = −2πi Res ; zk (2.6.3)′
−∞ Q(x) k=m+1
Q(z)
Z ∞
dx π
Example 2.6.1. = α > 0.
−∞ x2 +α 2 α
Indeed, there is just one root z = iα in the upper half-plane, and
1 1 1
Res( ; αi) = z=αi
=
z2 +α 2 2z 2αi
Z ∞
dx π
Example 2.6.2. = α > 0, β > 0.
−∞ (x2 + α2 )(x2 + β 2) αβ(α + β)
(a) Consider first α ≠ β. Then there are two simple roots z1 = αi and
z2 = βi in the upper half-plane and
1 1 1
Res 2 2 2 2
; αi = = .
(z + α )(z + β ) 2z(z + β ) z=αi 2α(β − α2 )i
2 2 2
(b) As α = β one can consider a double root, and thus a double pole, but
also just plug α = β into the result above (not into each residue!).
Proof. Since cos(z) and sin(z) are unbounded in both the upper (and lower)
complex half-plane observe first that
Z ∞ P (x)eix
I = Re dx (2.6.7)
−∞ Q(x)
and
Z ∞ xP (x)eix
J = Im dx (2.6.8)
−∞ Q(x)
−R R
xP (x)
where f (x) = Q(x)
→ 0 as r → +∞ and since
−R + iR R + iR
[h!] −R R
Solution. Since the integrand is an even function, we see that we can take
integral from −∞ to ∞ and then to halve it. Since Q(x) = x2 + a2 has just
one root αi in the upper complex half-plane, and this root is simple,
1 h zeiz i 1 zeiz πe−α
I = Im 2πi Res 2 , αi = × Im 2πi = .
2 z + α2 2 2z z=αi 2
−R + iR R + iR
−R −ε ε R
iz
Due to Cauchy’s Theorem Γε,R e zdz = 0 (there are no singularities
R
inside). Exactly like in the proof of Theorem 2.6.4 it is easy to show that
integrals over both vertical lines and upper horizontal line tend to 0 as
R → ∞ and that
Z ∞ ix Z −1 ix
e dx e dx
and
1 x −∞ x
Chapter 2. Basic Properties of Analytic Functions 102
Taking the same curve Γε,R as in the previous example, we see that due to
iz
Cauchy’s Theorem Γε,R e zdz = 0 (there are no singularities inside).
R
Again, integrals over both vertical lines and upper horizontal line tend
to 0 as R → ∞. Further,
Z ∞ Z −1
(1 − e2ix ) dx (1 − e2x ) dx
and
1 2x2 −∞ 2x2
converge absolutely. Therefore since 1 − e2iz = −2iz + (1 + 2iz − e2iz ),
Z (1 − e2iz ) dz
I = − lim+ Re
ε→0 γε 2z 2
−i dz (1 − e2iz + 2iz) dz
Z Z
= − lim+ Re +
ε→0 γε z γε 2z 2
Chapter 2. Basic Properties of Analytic Functions 103
where γε is an arc of radius ε. Here the first integral equals −π and the
second tends to 0 as ε → 0 (as in the previous example). Finally, I = π.
Remark 2.6.4. To calculate
Z ∞
P (x) sin(x) dx
(2.6.11)
−∞ xQ(x)
and
∞
P (x) sin2 (x) dx
Z
(2.6.12)
−∞ x2 Q(x)
where P (x) and Q(x) are even real-valued polynomials of degrees m and
n correspondingly, m ≤ n, and Q(x) does not have real roots, we use the
same curve Γε,R and the same approach, albeit instead of Cauchy’s theorem
we use the Residue theorem.
Example 2.6.7.
Z
sin(x) dx
I := , α > 0, (2.6.13)
x(x2 + α2 )
and
sin2 (x) dx
Z
J := , α > 0. (2.6.14)
x2 (x2 + α2 )
Solution. Analysis of the previous two examples shows that the contribution
of 0 (or, more precisely, of −γε with ε → 0+ ) is απ2 (because there is an extra
1 1
factor z2 +α 2 z=0 = α2 ).
π eiz π eiz
I= + Im 2πi Res ; αi = + Im 2πi
α2 z(z 2 + α2 ) α2 2z 2 z=αi
π
= 2 (1 − e−α )
α
and
π 1 − e2iz π (1 − e2iz )
J= + Re 2πi Res ; αi = + Re 2πi
α2 2z(z 2 + α2 ) α2 4z 3 z=αi
−2α
π 1−e
= 2 1− .
α 2α
Chapter 2. Basic Properties of Analytic Functions 104
where P (x) and Q(x) are even real-valued polynomials of degrees m and n
correspondingly, and −1 < p < 1, p ̸= 0, m + p + 1 < n and Q(x) does not
vanish on the real line.
This integral is understood as (a garden variety) improper integral.
Example 2.6.8.
∞
xp dx
Z
I= , −1 < p < 1, α > 0. (2.6.16)
0 x2 + α 2
zp
Solution. Consider f (z) = 2 and a curve Γε,R which is already famil-
z + α2
iar:
−R −ε ε R
z p dz zp (αi)p
Z
2 2
= 2πi Res 2 ; αi = 2πi
Γε,R z + α z + α2 2z z=αi
iπp
= παp−1 e 2 (2.6.17)
iπp
because ip = eZ2 on the selected branch.
R
xp dx
Let Iε,R = 2 2
denote a required integral, but from ε to R.
ε x +α
Chapter 2. Basic Properties of Analytic Functions 105
and, finally
iπ
παp−1 e 2 παp−1 παp−1 pπ
I= iπp
= iπp iπp = sec .
1+e e− 2 + e 2 2 2
Observe that the final answer is real and positive (as it should be).
Example 2.6.9.
Z ∞
ln(x) dx
I := , α > 0. (2.6.19)
0 x2 + α 2
Solution. We repeat arguments of Example 2.6.8, selecting the same domain
Log(z)
and curve, and f (z) = 2 where we select a branch Log(z) which is
z + α2
ln(x) as z = x > 0 and analytic inside Γε,R . What will change?
First, we have a different right-hand expression in (2.6.17):
Log(z) Log(z) π π iπ
2πi Res ; αi = 2πi = Log(αi) = ln(α) +
z2 + α 2 2z z=αi α α 2
because of the branch selection.
RR
Second, while Iε,R = ε ln(x) dx
x2 +α2
,
−ε Z −ε
ln(|x|) + πi dx
Z
Log(x) dx
=
−R x2 + α 2 x2 + α 2
R
−R Z R
ln(|x|) + πi dx
Z
dx
= = Iε,R + πi ,
ε x2 + α 2 2
ε x +α
2
Chapter 2. Basic Properties of Analytic Functions 106
where we changed x := −x passing from the first line to the second one.
R
Again, Iε,R → I as ε → 0+ and R → ∞, γR . . . → 0 as R → ∞ and
R
γε
. . . → 0 as ε → 0+ and
Z ∞
dx π iπ
2I + πi = ln(α) +
0 x2 + α 2 α 2
π ln(α)
and taking real part we get 2I = π ln(α) and I = while imaginary
R ∞ dx α π
2α
part will just bring us trivial 0 x2 +α2 = 2α .
The same trick would work in many cases, like with an extra factor xp
with −1 < p < 1.
Example 2.6.10.
Z ∞
x ln(x) dx
I := , α > 0, β > 0, α ̸= β. (2.6.20)
0 (x + α2 )(x2 + β 2 )
2
Solution. The same trick out of the box would not work here because without
ln(x) we have an odd function and
Z −ε Z R
x(ln(|x|) + πi) dx x(ln(|x|) + πi) dx
2 2 2 2
=− 2 2 2 2
−R (x + α )(x + β ) ε (x + α )(x + β )
RR
and adding ε (x2x+α
ln(|x|) dx
2 )(x2 +β 2 ) we get just integral without logarithmic factor.
On the other hand, the integral expression after we take the limit as R → ∞
and ε → 0+ is
Z ∞ 2
x ln2 (x) dx
Z 0
x ln(|x|) + iπ dx
+
0 (x2 + α2 )(x2 + β 2 ) 2 2 2
−∞ (x + α )(x + β )
2
Z ∞ Z ∞ 2
x ln2 (x) dx x ln(x) + iπ dx
= −
0 (x2 + α2 )(x2 + β 2 ) (x2 + α2 )(x2 + β 2 )
Z ∞ 0
x 2 ln(x) + iπ dx
= − iπ .
0 (x2 + α2 )(x2 + β 2 )
This expression must be equal to the last expression on the previous slide.
We take only imaginary parts (taking real parts will bring a much easier
result)
ln(α) − ln(β) ln(α) + ln(β)
−2πI = π .
α2 − β 2
Taking α = β and calculating the limit we would automatically bring result
in this case as well:
ln(α)
−2πI = π .
α2
Example 2.6.11.
Z ∞
xp dx
I := , −1 < p < 1, p ̸= 0, α > 0. (2.6.21)
0 (x − a)2 + α2
Here the radius of the large arc is R, the radius of small arc is ε and the
width of the cut is 0!
zp
We take f (z) = where z p is the branch which on the upper
(z − a)2 + α2
side of the cut, when z = x + i0, x > 0, is equal to xp . Then on the lower
side of the cut, z = x − i0, x > 0, is equal to e2iπp xp .
One can see easily, that the integral over γR (the large arc) tends to 0 as
R → ∞, integral over γε (the small arc) tends to 0 as ε → 0+ , and integrals
over upper and lower sides of the cut tend to I and −e2iπp I correspondingly,
because while we integrate from ε to R on the upper side of the cut, we
integrate from R to ε on the lower side!
So the integral would be
(1 − e2iπp )I
in the end.
Now we have not one, but two poles z± = a ± αi. This is why we leave
keyhole domain as a last reserve.
Calculating residues
zp zp 1
Res 2 2
; a ± αi = =± (a ± αi)p .
(z − a) + α 2(z − a) z=a±αi 2αi
Therefore
π
1 − e2iπp I = (a + αi)p − (a − αi)p
α
π
= (a + αi)p − e2iπp (a − αi)p
α
where in the first line ζ p is a branch defined on 0 < arg(z) < 2π, while in
the second line ζ p is a branch defined on −π < arg(z) < π. You just need
to look at arguments of ζ and ζ p .
Then
π
e−ipπ − eiπp I = e−ipπ (a + αi)p − eiπp (a − αi)p .
α
p
Taking imaginary part and observing that in these new settings ζ p = ζ we
get
2π −iπp
−2 sin(πp)I = Im e (a + αi)p .
α
Chapter 2. Basic Properties of Analytic Functions 109
Finally,
π
I = − csc(πp) Im e−iπp (a + αi)p .
α
Remark 2.6.5. (a) For a = 0 this answer coincide with the answer obtained
in Example 1.
(b) One can consider examples with the factor xp lnq (x).
(c) For
Z ∞
ln(x) dx
I := , α > 0,
0 (x − a)2 + α2
Log2 (z)
one should take a keyhole domain (if a ̸= 0) and f (z) = 2 with
z + α2
Log(z) = ln(x) on the upper side of the cut and Log(z) = ln(x) + 2πi
on it’s lower side (think–why square).
(d) With keyhole domain one cannot consider examples with factors cos(x)
or sin(x) because eiz is bounded only in the upper half-plane.
Example 2.6.12.
Z 2π
dθ
I := , a > |b| > 0. (2.6.24)
0 a − b cos(θ)
Solution. Then
Z Z
dz 2i dz
I= z 1
=−
|z|=1 iz a − b( 2 + 2z ) |z|=1 bz 2 + b − 2az
Z
2 dz
=− .
bi |z|=1 (z − z1 )(z − z2 )
√ √
with z1 = 1b a − a2 − b2 , z2 = 1b a + a2 − b2 = z1−1 , |z1 | < 1 < |z2 |.
Example 2.6.13.
Z 2π
dθ
I := , a > |b| + |c|, |c| > 0. (2.6.25)
0 a2 − (b + c cos(θ))2
Solution. Then
Z
dz
I=
− [b + c( z2 + 2z
1 2
|z|=1 iz a2 )]
Z
4iz dz
= .
|z|=1 [2bz + c(z 2 + 1)]2 − 4a2 z 2
1
p
The denominator has two roots inside z1,2 = c
a±b− (a ± b)2 − c2 and
1
two roots outside z3,4 = z1,2 . Then
Z
4i z dz
I= 2 ,
c |z|=1 (z − z1 )(z − z2 )(z − z3 )(z − z4 )
112
Chapter 3. Analytic Functions as Mappings 113
(c) Equivalently,
Remark 3.1.1. This theorem states that zeroes of analytic function are
isolated and cannot accumulate to a point in D.
However they can accumulate to the point on the boundary, including
∞. For example, zeroes of sin(z) are zn = πn and they accumulate to ∞,
and zeroes of sin z1 which is analytic in C \ {0} are accummulating to 0.
2
when this
√ parameter changes, like z + ε has two different simple zeroes
z1,2 = ± −ε as ε ̸= 0, and one double zero as ε = 0. More generally, it is
the same for roots of polynomials P (z) when coefficients change.
Proof of Lemma 3.1.2. (i) If f (z) has a zero of order m at z0 , then f (z) =
(z − z0 )m g(z),
where
p q
f′ f′ X f′
Z
1 X
dz = Res ; zj + Res ; wk
2πi γ f j=1
f k=1
f
p q
X X
= mj + (−nk ) = N (f ; γ) − P (f ; γ)
j=1 k=1
where equality between the first and the second lines follows from Lemma 3.1.2.
3.1.5 Infinity
What about infinity? Recall that
(a) f (z) has a zero of order m at ∞ if f (z) = z −m g(z) with g(z) having
removable singularity at ∞ and g(∞) ̸= 0. Then
f′ g′ f′
= −mz −1 + =⇒ Res ;∞ = m
f g f
since at ∞ residue is calculated with the opposite sign!
(b) f (z) has a pole of order n at ∞ if f (z) = z n g(z) with g(z) having
removable singularity at ∞ and g(∞) ̸= 0. Then
f′ −1 g′ f′
= nz + =⇒ Res ; ∞ = −n
f g f
since at ∞ residue is calculated with the opposite sign!
Therefore,
Remark 3.1.3. (a) Statement of Lemma 3.1.2 holds for ∞ as well!
Chapter 3. Analytic Functions as Mappings 117
(b) Further, Statement of Theorem 3.1.3 holds not only for inside γ but
also for outside γ (in which case one must assume that outside γ f
is analytic except for a finite number of poles, and ∞ is also either a
pole or zero, and it is included then in calculating N (f ; γ) and P (f ; γ).
Also γ must be properly oriented, so going along it leaves outside on
the left!
Example 3.1.1. (a) f (z) = z (simple zero at 0) then, when z goes once
around 0 and its argument changes from 0 to 2π, arg(f (z)) changes
with the same speed from 0 to 2π;
(b) f (z) = z 2 (double zero at 0) then, when z goes once around 0 and its
argument changes from 0 to 2π, arg(f (z)) changes with the double
speed from 0 to 4π;
(c) f (z) = z 3 (triple zero at 0) then, when z goes once around 0 and
its argument changes from 0 to 2π, arg(f (z)) changes with the triple
speed from 0 to 6π;
(d) f (z) = z −1 (simple pole at 0) then, when z goes once around 0 and
its argument changes from 0 to 2π, arg(f (z)) changes with the same
speed but in the opposite direction from 0 to −2π;
(e) f (z) = z −2 double pole at 0) then, when z goes once around 0 and its
argument changes from 0 to 2π, arg(f (z)) changes with the double
speed but in the opposite direction from 0 to −4π;
(f) f (z) = z −3 triple pole at 0) then, when z goes once around 0 and
its argument changes from 0 to 2π, arg(f (z)) changes with the triple
speed but in the opposite direction from 0 to −6π.
iR
z + e−z = λ
has exactly one solution in the right half-plane {z : Re(z) > 0}.
Chapter 3. Analytic Functions as Mappings 120
iR
−iR
Proof. Note that the assumption (3.1.7) (strict inequality! ) ensures that
neither f nor g is zero on γ. We may assume further that all the common
zeros of f and g inside γ have been canceled; this affects neither the
assumptions nor the conclusion.
g
Let h = ; then
f
|h(z) + 1| < 1 ∀z ∈ γ
so the range of h on γ lies in the open disk of radius 1 centered at the point
−1.
In particular, arg(h(z)) has no net change as z traverses γ because
w = f (z) cannot go around 0. By the Argument Principle, the number of
zeros of h inside γ equals the number of poles of h inside γ.
But the number of zeros of h is just the number of zeros of g, and the
number of poles of h just the number of zeros of f .
Remark 3.1.5. (a) Suppose that f and g satisfy the hypotheses of Theo-
rem 3.1.5 except that
It is then still true that f and g have equally many zeros inside γ. The
reasoning is elementary: according to Theorem 3.1.5, the assumption
(3.1.8) implies that f and −g have equally many zeros inside γ.
However, a point is a zero of −g exactly when it is a zero of g.
(b) So, we need to construct a a reference function g(z), satisfying (3.1.7)
or (3.1.8) and such that we can easily calculate the number of its zeros.
Example 3.1.4. Show that all the zeros of
p(z) = 3z 3 − 2z 2 + 2iz − 8
|f (z) + 8| ≤ 3 + 2 + 2 = 7 < 8,
so p(z) and f (z) = 8 have the same number of zeros within {z : |z| = 1}, by
Rouché’s Theorem; thus, p(z) does not vanish in the disk {z : |z| = 1}.
Chapter 3. Analytic Functions as Mappings 122
0
2
1
ez = e2 z
N (f ) = P (f ) (3.1.9)
Proof. Indeed, G has just one pole, namely, ∞ and the order of it is equal
to the order of polynomial G.
Remark 3.1.6. The last statement is known as a Fundamental Theorem of
Algebra (of polynomials).
We can draw an immediate conclusion from the work that preceded the
statement of Theorem 3.2.1.
Chapter 3. Analytic Functions as Mappings 125
Remark 3.2.2. (a) This is a very special case of the uniqueness theorem
in PDE: let u be a harmonic function in the bounded domain D ⊂ Rn ,
and u = 0 on its boundary. Then u = 0 in D. Moreover, if u and v
are two harmonic functions in D; then u − v is also harmonic in D
and if u = v everywhere on B, then u = v everywhere in D.
| sin(θ)|
g(θ) =
2 + cos(θ)
sin(θ) 2 cos(θ) + 1
g(θ) = =⇒ g ′ (θ) = =0
2 + cos(θ) (2 + cos(θ))2
1 2π 1
=⇒ cos(θ) = − =⇒ θ = =⇒ g(θ) = √ .
2 3 3
However we need to explore the ends θ = 0 and θ = π, where g(θ) = 0.
So
1
(a) maxD (u) = √ achieved as θ = 2π 3
and
3
θ = 4π
3
(don’t forget
√
about lower half circle),
3 3 3
or points − 2 ± i 2 .
Example 3.2.2.
Let
f (z) = z 2
and
D = {z : |z − 1| ≤ 1}.
Find
max Im(f (z)) and min Im(f (z))
z∈D z∈D
and
ϕ′ (t) = a cos(t) + cos2 (t) − sin2 (t) = 2 cos2 (t) + a cos(t) − 1 = 0
1 √
=⇒ cos(t) = −a ± a2 + 8 .
4
We get two values
√ √
1 3 3 3
cos(t1 ) = =⇒ x1,2 = , y1,2 = ± =⇒ Im(z1,2 ) = ± ,
2 2 2 2
and
cos(t3 ) = −1 =⇒ x3 = 0, y3 = 0 =⇒ Im(z3 ) = 0.
√
3 1 √
Therefore, maxD f (z) = , achieved at 3 + 3i and
√ 2 2
3 1 √
minD f (z) = − , achieved at 3 − 3i .
2 2
Example 3.2.3.
Let
f (z) = ze−z
and
D = {z : | Im(z)| < Re(z), |z| > 1}.
Find
max |f (z)|
z∈D
Therefore
√
max |f (z)| = 2e−1
z∈D
achieved as z = (1 ± i).
Example 3.2.4.
Let
1
f (z) =
1 + z2
and
D = {z : | Im(z)| ≤ Re(z)}.
Find
max Im(f (z)) and min Im(f (z))
z∈D z∈D
1 1
max Im(f (z)) = √ and min Im(f (z)) = − √
z∈D 2 z∈D 2
(ii) Equality can hold for some z ̸= 0 only if f (z) = λz with a constant
λ : |λ| = 1.
Proof. (i) Since f (0) = 0, we know that g(z) = f (z)/z is also analytic
in D. For |z| = r, we have |g(z)| = |f r(z) ≤ 1r .
By the maximum-modulus principle, the inequality |g(z)| ≤ 1/r is true
for |z| < r as well. Since r can be arbitrarily close to 1, we must conclude
that |g(z)| ≤ 1 if |z| < 1; thus, |f (z)| ≤ |z| in D.
Remark 3.2.3. Schwarz’s Lemma does not hold for functions of real variable.
For example, f (x) = x22x+1 satisfies |u(x)| ≤ 1 for all x ∈ R and u is smooth,
u(0) = 0 but u(x) > |x| for x ∈ (−1, 1).
Chapter 3. Analytic Functions as Mappings 131
where γ = {z : |z − z0 | = r}.
where γ = {z : |z − z0 | = r}.
Example 3.2.5.
| sin(θ)|
g(θ) =
2 + cos(θ)
Solution.
2π π
| sin(θ)| dθ | sin(θ)| dθ
Z Z
1 1
u(0) = =
2π 0 2 + cos(θ) π 0 2 + cos(θ)
1 θ=π 1
= − ln(2 + cos(θ)) θ=0
= ln(3).
π π
a1 z + b 1 a2 w + b 2
Proof. (i) Let w = T1 (z) = and T2 (w) = . Then
c1 z + d 1 c2 w + d 2
a1 z + b 1
a2 + b2
c1 z + d1 a2 (a1 z + b1 ) + b2 (c1 z + d1 )
(T2 ◦ T1 )(z) = =
a1 z + b 1 c2 (a1 z + b1 ) + d2 (c1 z + d1 )
c2 + d2
c1 z + d 1
(a2 a1 + b2 c1 )z + (a2 b1 + b2 d1 ) a3 z + b 3
= = (3.3.2)
(c2 a1 + d2 c1 ) + (c2 b1 + d2 d1 ) c3 z + d3
!
1 0
(c) Finally, to T (z) = z is assigned matrix I = .
0 1
(d) However, the same transform T corresponds to the whole bunch of
quadruplets {a, b, c, d}, different by a common factor λ ̸= 0. Requiring that
∆ := ad − bc = 1 (3.3.4)
would ensure that to each transform T correspond exactly two such quadru-
plets, different by factor −1, and thus two matrices M [T ] with determinant
1, different by a factor −1.
(e) Under this condition to T −1 is assigned matrix M [T ]−1 .
(f) On the other hand, inverse map M 7→ T (where M is any 2×2-matrix
with determinant 1) is single valued and according to Part 1
(g) T [M2 M1 ] = T [M2 ] ◦ T [M1 ],
(h) T [M −1 ] = (T [M ])−1 ,
(i) and T [±I] = id is an identical map z → z.
We can say few mathematical jargon words showing how learned we are,
but we are not in Math Specialist class.
Theorem 3.3.2. A linear fractional transformation
a−z
T (z) = λ with a, λ ∈ C, |a| < 1 and |λ| = 1 (3.3.5)
1 − az
maps a unit disk D := {z : |z| < 1} onto itself.
Proof. Observe that if T is defined by (3.3.5) then
|a − z|2 |a|2 − 2 Re(az) + |z|2
|T (z)|2 = =
1 − az|2 1 − 2 Re(az) + |a|2 |z|2
and |T (z)| < 1 if and only if
which is true iff |z| < 1 that means z ∈ D. Therefore, it maps D onto
itself.
Chapter 3. Analytic Functions as Mappings 135
because |a1 |2 + |a2 |2 < 1 + |a1 |2 |a2 |2 (since |a1 | < 1, |a2 | < 1).
z−a w − a2
(ii) If w = λ then z = λ2 with λ2 = λ and a3 = −λa,
1 − az 1 − a2 w
|λ2 | = 1 and |a2 | < 1 (Check it! ).
(iii) Finally w = z is given by (3.3.5) with λ = 1, a = 0.
Theorem 3.3.4. Let |z1 | < 1, |z2 | = 1 and |w1 | < 1, |w2 | = 1. Then
exists exactly one transformation L of type (3.3.5) such that L(z1 ) = a1 and
L(z2 ) = w2 .
Chapter 3. Analytic Functions as Mappings 136
T (z) = z. (3.3.8)
az + b
Proof. z is a fixed point of fractional linear transformation T (z) =
cz + d
with c ̸= 0 if and only if z is a root of the quadratic equation
cz 2 + (d − a)z − b = 0
̸ 0 consider by
which has at most two distinct roots. Case c = 0, d =
yourself.
Corollary 3.3.7. Let T and S be two fractional linear transformations
that are equal at three distinct points: T (zj ) = S(zj ) for j = 1, 2, 3. Then
T (z) = S(z) for all z.
Proof. If T (zj ) = S(zj ) then (S −1 ◦ T )(zj ) = zj for j = 1, 2, 3 and since
S −1 ◦ T is also a fractional linear transformation we conclude that (S −1 ◦
T )(z) = z for all z, and T (z) = S(z) for all z.
Theorem 3.3.8. Let z1 , z2 and z3 be three distinct complex numbers, and
let w1 , w2 and w3 be also three distinct complex numbers. Then there exists
a unique linear fractional transformation L with L(zj ) = wj for j = 1, 2, 3.
Proof. Let
z − z1 z2 − z3
T (z) = . (3.3.9)
z − z3 z2 − z1
Then T (z1 ) = 0, T (z2 ) = 1 and T (z3 ) = ∞. Let us define S(w) by the
same formula (3.3.9) with w, w1 , w2 and w3 instead of z, z1 , z2 and z3 . Then
S(w1 ) = 0, S(w2 ) = 1 and S(w3 ) = ∞. Then L = S −1 ◦ T is a required
transformation. Due to Corollary 3.3.7 it is unique.
Finally,
z − z1 z2 − z3 w − w1 w2 − w3
= . (3.3.10)
z − z3 z2 − z1 w − w3 w2 − w1
Remark 3.3.4. We can extend these arguments to the case when one of
2 −z3
z1 , z2 , z3 or/and w1 , w2 , w3 is ∞. Indeed, if z1 = ∞, we can take T (z) = zz−z 3
;
z−z1 z−z1
if z2 = ∞, then T (z) = z−z3 ; and if z3 = ∞, then T (z) = z2 −z1 .
Example 3.3.1. Find a linear fractional transformation, that sends 0, 1, 2 to
−1, 0, 4 respectively.
Chapter 3. Analytic Functions as Mappings 138
1 − 12 z − 12
2z − 1
1 z =
1− 2 1− 2 z−2
1 1
w−2
1+ 2 2w − 1 5z − 4
1 w =− =⇒ w = .
−1 − 2 1− 2 2−w 4z − 5
Since U and W are linear transformations, which map circles onto circles
and straight lines onto straight lines, it is sufficient to prove the declared
property for inversion map V . The equation
α(x2 + y 2 ) + βx + γy = δ
α + βu − γv = δ(u2 + v 2 )
Remark 3.3.5. One can see easily that the inversion map transforms
(a) a straight line or a circle passing through the origin into a straight
line,
(a) z (b) w
(b) a straight line or a circle not passing through the origin into a circle.
Chapter 3. Analytic Functions as Mappings 140
(a) z (b) w
The curve γ has a tangent vector z ′ (t0 ) at z0 , and we suppose that z ′ (t0 ) ̸= 0.
Remark 3.4.1. If we consider another parametrization of the same curve γ,
z = z(t(s)), a′ < s < b′ , then we will get another tangent vector, but it will
be proportional with a positive coefficient if we assume that the direction in
which curve is passed is the same for both parameterizations, that is, t(s) is
increasing function of s.
What happens to the curve γ when we apply an analytic function f (z)
to it? It is transformed into a new curve Γ in the w-plane; Γ is given by
In particular,
f ′ (z0 ) ̸= 0 (3.4.4)
we conclude that
(a) the tangent vector is scaled in length by a factor |f ′ (z0 )|, and
Definition 3.4.1. Let γ1 and γ2 be two smooth oriented curves that intersect
at the point z0 = z1 (t0 ) = z2 (s0 ). Then the angle between γ1 and γ2 at z0
is the angle θ measured counter-clockwise from the tangent vector z1′ (t0 ) to
the tangent vector z2′ (s0 ), provided neither of these tangent vectors is zero:
γ2
z2′ (s0 )
z0 θ
γ1 z1′ (t0 )
Remark 3.4.2. This definition of the angle between two curves breaks down
if one of the vectors z1′ (t0 ) or z2′ (s0 ) is 0.
Chapter 3. Analytic Functions as Mappings 142
Definition 3.4.2. Suppose now that ϕ(z) is a function, perhaps not analytic,
defined in the disk {z : |z − z0 | < r and satisfying ϕ(z) ̸= ϕ(z0 ) in the
punctured disk {z : < |z − z0 | < r}.
Then map ϕ is conformal at z0 if, whenever two oriented curves γ1 and
γ2 meet at z0 , the angle from Γ1 = ϕ(γ1 ) to Γ2 = ϕ(γ2 ) is equal to the angle
from γ1 to γ2 .
Remark 3.4.3. Mercator projection in geography also preserves angles (but
stretches near polar domains more than near equatorial ones).
Theorem 3.4.1. If f is analytic near z0 and f ′ (z0 ) ̸= 0 then f is conformal
at z0 .
Proof. Indeed, if γ1 and γ2 intersect at z0 then Γ1 and Γ2 intersect at
w0 = f (z0 ) and equation (3.4.3)
(e) One can prove the converse to Theorem 3.4.1: if f is conformal, then
it is analytic and f ′ ̸= 0.
(b) |L(z)| < |z| for for all z : 0 < |z| < 1.
But case (b) is impossible. Indeed, we can apply the same arguments to
inverse map L−1 and conculde that |L−1 (w)| ≤ |w| for all w : |w| < 1 and
plugging w = L(z) we conclude that |z| ≤ |L(z)| < |z| for all z : 0 < |z| < 1.
Contradiction.
Therefore
λz − a z−b
(T −1 ◦ S)(z) = λz =⇒ S(z) = T (λz) = =λ
1 − aλz 1 − bz
Proof. Indeed, any conformal map of D onto D is a map of type (3.4.6) and
for them this statement has been proven.
3.4.3 Examples
Example 3.4.1. w = z 2 , it maps the right half-plane {z : − π2 < arg(z) < π2 }
conformally onto plane with a cut {w : − π < arg(w) < π}. What happens
with level lines of z? {z = x + bi} become parabolas {w = (x2 − b2 ) + 2bxi}
and {z = a + yi} become also parabolas {w = (−y 2 + a2 ) + 2ayi}:
(a) z (b) w
Chapter 3. Analytic Functions as Mappings 145
(a) z (b) w
−1
−2
−2 −1 0 1 2
Chapter 3. Analytic Functions as Mappings 146
−1
−2
−2 −1 0 1 2
√
Example 3.4.4. w = z, it maps the plane with a cut {z : −π < arg(z) < π}
3
(a) z (b) w
Chapter 3. Analytic Functions as Mappings 147
q p L′
(a) z (b) w
w−p
Example 3.4.6. z = log w−q
̸ q, it is defined on the twice punctured
with p =
plane {w : |w − q| > 0, |w − q| > 0}. Consider inverse map; then we get
Circles of Appolonius. Indeed Re(z) = a means that |w−p| |w−q|
= ea and we get a
family of blue circles, while Im(z) = b means that arg(z − p) − arg(z − q) = b
and we get a family of arcs of red circles between p and q (see W2L1).
Example 3.4.7. w = cos(z). It maps conformally stripe {z : 0 < Re(z) < π}
onto plane with cuts C \ ((−∞, −1] ∪ [1, ∞)) (see W3L2):
u2 v 2
+ 2 =1
a2 b
with a = cosh(y), b = | sinh(y)|.
C \ [−1, 1]. Indeed, circles z = reit with fixed r and t ∈ [0, 2pi)are mapped
onto confocal ellipses u = A cos(t), v = −B sin(t) with A = 12 (r + r−1 ),
Chapter 3. Analytic Functions as Mappings 149
Remark 3.5.1. (a) The proof of this theorem is completely out of our
reach. Actually there are many proofs and none is constructive.
While The Riemann Mapping Theorem says how domains are mapped,
it does not discuss the mapping of their boundaries. This is covered by
Carathéodory’s theorem below. But first we need the following
Jordan curve has an important property, which looks obvious but in fact
is very non-trivial and deep:
(a) f : γ1 → γ2 ;
γ
γ z
z
Mγ (z)
Mγ (z)
O
(a) (b)
Figure 3.1: On (a) z and Mγ (z) are on the equal distances from γ; if γ ⊂ R
then Mγ (z) = z. On (b) O is a center of the circle, R its radius and
R2
|OMγ (z)| = R2 |Oz|−1 ; if γ ⊂ {z : |z| = R} then Mγ (z) = .
z
Corollary 3.5.6. Let D1 = {z : r1 < |z| < R1 } and D2 = {w : r2 < |w| <
R2 } and f : D1 → D2 conformal one-to-one map of D1 onto D2 . Then
r1 r2
(i) = ;
R1 R2
Chapter 3. Analytic Functions as Mappings 152
µ r2 R2
(ii) Either f (z) = λ|z| or f (z) = with |λ| = = and |µ| = r1 R2 =
z r1 R1
r2 R1 .
Proof. Assume that f maps inner circle onto inner circle and outer circle
onto outer circle. Otherwise we replace f (z) by f (z)−1 , r2 by R2−1 and R2
by r2−1 .
Using mirror reflection we conclude that f maps conformally punctured
disk {z : 0 < |z| < R1 } onto punctured disk {w : 0 < |w| < R2 } and therefore
disk onto disk, and f (0) = 0. Then f (z) = λz and |λ| = R1−1 R2 :
(a) z (b) w
Since f maps inner circle onto inner circle, it should be also |λ| = r1−1 r2 ,
and it implies both (i) and (ii).
3.5.3 Examples
Example 3.5.1 (Upper half-plane to unit circle). (a) Consider the following
conformal map from H+ := {z : Im(z) > 0} onto D = {w : |w| < 1}.
According to Carathéodory’s theorem it’s boundary, {z : Im(z) = 0} should
be mapped onto {w : |w| = 1} and one of the way to achieve it is to take
z−i
w = f (z) =
z+i
we took “−” in the numerator and “+” in the denominator because otherwise
it would be singular at i ∈ H+ . And we can see that if z = x + yi with
y > 0, then |w| < 1 (check it!).
Further, since it is a linear fractional transformation, it must transform
lines {z : Im(z) = b} into either circles or straight lines, but those must be
circles, passing through 1 (when |z| → ∞, w → 1).
Chapter 3. Analytic Functions as Mappings 153
And indeed:
x + (y − 1)i [x + (y − 1)i][x − (y + 1)i]
w = f (z) = =
x + (y + 1)i x2 + (y + 1)2
(a) z (b) w
f (z) = A(z − x0 )β + B
where x0 and β are real numbers, 0 < β < 2, and 0 ̸= A and B are complex
numbers.
Basically it is a composition of the linear function w = Aω + B which
shifts and rotates, ω = ζ β which maps upper half-plane onto sector {ω : 0 <
arg(ω) < απ} and a shift along real axis ζ = z − x0 .
Observe that f ′ (z) = βA(z − x0 )α with α = β − 1 =⇒ −1 < α < 1.
Consider
with x1 < x2 < . . . < xN and |αj | < 1 (so f (z) will be primitive of it). Then
θ0 + θ1 + . . . + θN = 2π (3.5.7)
w0 θ4
w4
θ0
w2
θ2
θ1 θ3
w1 w3
Let αj = −θj /π. Then there exist real numbers x1 < x2 < . . . < xN and
a constant A such that the function f (z) whose derivative is
(b) Polygon in this content is something more general, than the bounded
polygon, leave alone bounded convex polygon, considered in geometry.
It can be unbounded and contain cuts.
(c) While there are plenty of examples in the Textbook, we consider few,
those where we can achieve a simple answer, may be even we already
know.
Example 3.5.4. Let P = {w : − a < Re(w) < a, Im(w) > 0}.
Chapter 3. Analytic Functions as Mappings 156
w2
w2
w3
w1 w3 w1