Download as pdf or txt
Download as pdf or txt
You are on page 1of 53

Conformal Mapping and its

Applications

Mohamad A. Mehdi

Thesis Advisor: Dr. Roger Nakad

THESIS

Submitted in partial fulfillment of the requirements

for the degree of Master of Sciences in Mathematics

in the Department of Mathematics and Statistics in

the Faculty of Natural and Applied Sciences

of Notre Dame University-Louaize


Lebanon

May 29th, 2017

1
.

2
Conformal Mapping and its
Applications

Mohamad A. Mehdi

Approved by:

Dr. Roger Nakad

-
Dr. Bassem Ghalayini

Dr. Holem Saliba

3
.

4
This thesis is dedicated to my two lovely daughters Tala and Sara. I love
you both a lot.

5
.

6
Acknowledgments

I would like to express my deepest appreciation and gratitude to my supervisor,


Dr. Roger Nakad for the patient guidance, encouragement and insight he has provided
throughout my time as his student. I have been extremely lucky to have a supervisor
who cared so much about my work, and who responded to my questions and queries so
promptly.

I would also like to thank all the members of the math department at Notre Dame
University-Louaize. Dr. Melhab Keirouz, I would like to thank you for the continuous
inspiration and eagerness you provided me with. I am also thankful to Dr. Georges Eid,
Dr. Bassem Ghalayini, Dr. Holem Saliba, and Dr. Joesph Malkoun.

I must express my gratitude to Zeinab Dia, my wife, for her continuous support
and encouragement. Also I am forever in debt to my mother, father, brothers, and
sister who were there for me with all the ups and downs of my research. Special thanks
are extended to my student Zahraa for her willingness to proof read countless pages of
meaningless mathematics. I would like to thank Dr. Yasser Fadlallah for standing by
my side throughout the research time.

Completing this work would not have happened without the support and help of my
close friends Sara, Hussein, Rida and Mohanad. I shall always be in their dept.

7
.

8
Introduction

Complex analysis, commonly known as the theory of functions of a complex vari-


able, is one of the important branches of classical mathematics. Not only it is useful in
several mathematical branches such as algebraic geometry, number theory, analytic com-
binatorics and applied mathematics, but also proved to be highly valuable in physics
including hydrodynamics and thermodynamics, and in engineering fields such as nu-
clear, aerospace, mechanical and electrical engineering. One important application of
complex analysis is in the string theory, which studies conformal invariants in quantum
field theory [Dit-Geo]. Applying analytic methods to partial differential equations allows
researchers to study the evolution of a system that is changing in manner governed by
precise constraints which may be used to represent many important problems in physical
world, relating for example diffusion of heat, fluid flows or quantum mechanics. Euler,
Gauss, Riemann, Cauchy and Weierstrass are all well acknowledged mathematicians
who worked extensively on complex analysis.

Conformal mapping is an important technique used in complex analysis and has


many applications in different physical situations. It is a mapping that preserves ori-
ented angles locally, allowing it to be efficiently implemented in creating solutions to the
Laplace equation on complicated planar domains. This can be used to simplify regions
with complicated shapes by preserving certain physical aspects and maintaining small-
scaled shapes. In other words, the technique of a conformal mapping is to transpose
a problem from a complicated region to a simpler domain, whereby it would be easier
to solve; then to be able to translate the solution to the original problem by using the
inverse of a conformal mapping to map back to the complicated region. Moreover,
the crucial point is that an analytic-mapping requires only a non-vanishing derivative
to be conformal and then to be able to simplify a governing and rather complicated
principle.

Conformal mappings have long played an important role in hydrodynamics and aero-
dynamics; this can be projected through Halseys work [Hal]. Siegel and Snyder suc-
ceeded in applying conformal mappings to the analysis of heat transfers in porous regions
with curved boundaries [Sie-Sny] and Theodorsen stressed on the importance of numeri-
cal conformal mapping when working on wing sections [Thd]. Contributions in applying
conformal mapping techniques to determine the exact frequencies of vibrating plates of
complicated boundary shapes carrying concentrated masses were made by Schinzinger
and Laura [Sch-Lau]. The application of conformal mapping can also conquer the pat-

9
tern recognition field. In fact, Wechsler and Zimmerman mapped the arrangement of
pixels on a machine-vision image from a log-spiral configuration into a rectangular plane
[Wec-Zim]. Conformal mapping methods has been also successfully applied in a wide
variety of fields varying from acoustic wave guide studies to plate buckling theory, from
the analysis of vibrating printed circuits to the determination of the temperature field
in the nuclear field arrangement of fast breeder reactors, from studies on airflow past
multielement airfoils to the solution of geophysical and oceanographic problems, and
from microstrip design to analysis of the earth-ionosphere wave guide.

The goal of this thesis is to introduce conformal mappings, prove their prop-
erties and discuss some of their applications.

We start in Chapter 1 by recalling basic facts in complex analysis including ana-


lycity of functions with complex variables, harmonic functions, complex integration,
residues and poles of complex functions.

In Chapter 2, we study some important complex mappings: the mapping w = z1 ,


w = sin z and linear fractional transformations. We show how certain regions are trans-
formed and prove some local properties of those mappings. Then, we introduce the
concept of a conformal map, give examples and prove crucial properties. More precisely,
we emphasize on the notable connection between conformal mappings and harmonic
functions of two variables: If a function is harmonic (i.e. it satisfies the Laplace equa-
tion) then the transformation of such a function via conformal mapping is also harmonic.
So, equations pertaining to any field that can be represented by a potential function
can be solved via conformal mapping. If the physical problem can be represented by
complex functions but the geometric structure becomes inconvenient then by an appro-
priate mapping it can be transferred to a problem with much more convenient geometry.

Three applications of conformal mappings are discussed and examined in Chapter


3. The first application is in fluid flow (Section 3.1). We start by recalling some
preliminaries in boundary value problems and fluid flow. Then, we derive velocity
potential and complex velocity of fluid flow in horizontal plate, circular disk, tilted
plates and airfoils. As a second application (Section 3.2), we determine the electrostatic
potential inside a non-coaxial cylindrical cable with prescribed constant potential
values on the two bounding cylinders. Such potential will be deduced from the well
known potential along a coaxial cable. In the last application (Section 3.3), we examine
problems of steady temperatures and discuss temperature functions in a half plane,
quadrant and semi-infinite slab.

10
Contents

1 An Overview of Complex Analysis 13


1.1 Functions of a complex variable . . . . . . . . . . . . . . . . . . . . . . . 13
1.2 Harmonic functions and harmonic conjugates . . . . . . . . . . . . . . . . 16
1.3 Complex integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.4 Complex series, residues and poles . . . . . . . . . . . . . . . . . . . . . . 20

2 Conformal Mapping 25
2.1 Mappings by elementary functions . . . . . . . . . . . . . . . . . . . . . . 25
2.1.1 Linear transformations . . . . . . . . . . . . . . . . . . . . . . . . 25
2.1.2 The transformation w = z1 . . . . . . . . . . . . . . . . . . . . . . 26
2.1.3 The mapping w = sin z . . . . . . . . . . . . . . . . . . . . . . . . 27
2.1.4 Linear fractional transformations . . . . . . . . . . . . . . . . . . 28
2.1.5 Linear fractional transformations of the upper half plane . . . . . 29
2.2 Conformal mappings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.2.1 Definition of conformal mappings . . . . . . . . . . . . . . . . . . 31
2.2.2 Examples of conformal mappings . . . . . . . . . . . . . . . . . . 33
2.3 Properties of conformal mappings . . . . . . . . . . . . . . . . . . . . . . 36

3 Applications of Conformal Mapping 40


3.1 Application 1: Fluid flow . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.1.1 Boundary value problems and fluid flow . . . . . . . . . . . . . . 40
3.1.2 Horizontal plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.1.3 Circular disk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.1.4 Tilted plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.1.5 Airfoils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2 Application 2: A non-coaxial cable . . . . . . . . . . . . . . . . . . . 45
3.3 Application 3: Steady temperatures . . . . . . . . . . . . . . . . . . 47
3.3.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.3.2 Steady temperature in a half plane . . . . . . . . . . . . . . . . . 48
3.3.3 Steady temperature in a semi-infinite slab . . . . . . . . . . . . . 49
3.3.4 Steady temperature in a quadrant . . . . . . . . . . . . . . . . . . 50

11
12
Chapter 1

An Overview of Complex Analysis

In this chapter, we survey the continuity, differentiability and analyticity results of


functions of a complex variable. The main goal is to introduce complex integrals which
are extremely important in the study of functions of complex variables and much richer
than real integrals (in one variable). For more details, we refer to [Bro-Chu, Gam,
Pon-Sil]

1.1 Functions of a complex variable


Let S be a subset of the complex plane C. A complex function f : S C is an
application that assigns to each complex number z = x + iy S a complex number
w = f (z) called the image of z by f . The set S is called the domain of definition of f .
We can write
f (z) = U (x, y) + iV (x, y),
where U (x, y) and V (x, y) are real functions of x and y.

It is known that a complex number z = x + iy can be written in polar form (or expo-
nential form) as
z = rei = r cos + ir sin ,
p
where r is the length of z, r = |z| = x2 + y 2 and is the angle that z makes with the
positive real axis when z is interpreted as a radius vector. has an infinite number of
possible values that differ by integral multiples of 2. Each value is called an argument
of z and the set of all such values is denoted by arg z. The principal value of arg z
denoted by Arg z is that unique value such that < . Evidently, then,
arg z = Arg z + 2n (n = 0, 1, 2, ).
Using the polar form of a complex number, f (z) can be also written as
f (z) = f (rei ) = U (r, ) + iV (r, ),
where U (r, ) and V (r, ) are real functions of r and .

13
A neighborhood of a complex number z0 is the set of complex numbers z such that
|z z0 | <  for some  > 0.

Definition 1.1.1. Let f be a complex function defined for all points z in a certain neigh-
borhood of z0 . We denote the limit of f as z approaches z0 is w0 in C by limz z0 f (z) =
w0 . It means that

 > 0 , > 0 such that |f (z) w| <  whenever |z z0 | < .

If  > 0, > 0 such that |f (z) L| <  whenever |z| > , we say that the limit of f
when z tends to infinity is L. It is denoted by
1
lim f (z) = lim f = L. (1.1)
z + z 0 z
A function f is continuous at z0 if limzz0 f (z) exists, f (z0 ) exists and limzz0 f (z) =
f (z0 ).

Definition 1.1.2. Let f be a complex function whose domain of definition contains a


neighborhood of a point z0 . We say that f is differentiable at z0 if and only if

f (z) f (z0 )
lim (1.2)
z z0 z z0
exists. This limit (when it exists) is denoted by f 0 (z0 ).

Now we provide a fast method to check whether a given function is differentiable or


not at a given point z0 .

Proposition 1.1.1. Let the function f (z) = U (x, y) + iV (x, y) be defined through-
out some neighborhood of a point z0 = x0 + iy0 . Suppose that the first-order partial
derivatives of the functions U and V with respect to x and y exist everywhere in that
neighborhood and those partial derivatives are continuous at (x0 , y0 ) and satisfy the
Cauchy-Riemann equations, i.e.,


Ux = Vy ,
(1.3)
Uy = Vx ,

at (x0 , y0 ). Then f 0 (z0 ) exists and we have f 0 (z0 ) = Ux (x0 , y0 ) + iVx (x0 , y0 ). Here Ux
(res. Uy ) denotes the partial derivative of U w.r.t the variable x (resp. w.r.t the variable
y) and Vx (res. Vy ) denotes the partial derivative of V w.r.t the variable x (resp. w.r.t
the variable y).

Using the polar expression, we can restate Proposition 1.1.1 as follows:

Proposition 1.1.2. Let the function f (z) = U (r, ) + iV (r, ) be defined throughout
some neighborhood of a point z0 = r0 ei0 . Suppose that the first-order partial derivatives
of the functions U and V with respect to r and exist everywhere in that neighborhood

14
and those partial derivatives are continuous at r0 ei0 and satisfy the Cauchy-Riemann
equations, i.e.,


rUr = V ,
(1.4)
U = rVr ,

at (r0 , 0 ). Then f 0 (z0 ) exists and we have f 0 (z0 ) = ei0 (Ur (r0 , 0 ) + iVr (r0 , 0 )). Here
Ur (resp. U ) denotes the partial derivative of U w.r.t the variable r (resp. w.r.t the
variable ) and Vr (res. V ) denotes the partial derivative of V w.r.t the variable r (resp.
w.r.t the variable ).
A subset D of the complex plane is a domain if D is open and connected. It means
that D is open and any two points in D can be joined by a broken line segment in D.
A domain D is said to be star-shaped with respect to z0 if for all points z1 in D, the
straight line segment joining z0 and z1 is contained in D. A domain D is said to be
star-shaped if it is star-shaped with respect to at least one of its points.

We are now ready to introduce the concept of an analytic function. A function f is


analytic at a point z0 if it has derivative at each point in some neighborhood of z0 . A
function f is analytic in an open set if it has derivative everywhere in that set. If we
should speak of a function f that is analytic in a set S which is not open, it is to be
understood that f is analytic in an open set containing S.

For instance, the function f (z) = z1 is analytic at each nonzero point in the finite
plane but the function f (z) = |z|2 is not analytic at any point since its derivative exists
only at z = 0 and not throughout any neighborhood. An entire function is a function
that is analytic at each point in C. Since the derivative of a polynomial exists every-
where, it follows that every polynomial is an entire function.

If a function f fails to be analytic at a point z0 but is analytic at some point in every


neighborhood of z0 , then z0 is called a singular point, or singularity, of f . The point
z = 0 is evidently a singular point of the function f (z) = z1 . The function f (z) = |z|2 ,
on the other hand, has no singular points since it is nowhere analytic. A necessary,
but by no means sufficient, condition for a function f to be analytic in a domain D
(open and connected) is clearly the continuity of f throughout D. Satisfaction of the
Cauchy- Riemann equations is also necessary, but not sufficient. Sufficient conditions
for analyticity in D are provided by Theorem 1.1.1 and Theorem 1.1.2.
Example 1.1.1. 1. For all z C, we define the complex exponential function ez by
ez = ex+iy = ex (cos y + i sin y). It is clear that ez is entire and (ez )0 = ez .
2. As in real analysis, solving the equation ez = w, where w is a non-zero com-
plex number, introduces the logarithmic complex function. Writing w = rei , we
get eu eiv = rei and so eu = r which is equivalent to u = lnr, and eiv = ei , so
v = + 2k where k Z. This leads to
log z = ln r + i( + 2k), (1.5)

15

where k Z and is the principal argument of z. For example, for z = 1 i 3,
we have log z = ln 2 + i(
3
+ 2k), where k Z.

The principal value of log z is the value obtained from Equation (2) when k = 0
and it is denoted by Log z, i.e., Log z = ln r + i. Note that

log z = Log z + 2ki.

The function f (z) = Logz is analytic on D with f 0 (z) = z1 , where D denotes the
set of the complex plane deprived of the points which are real and non-positive.
It is called the principal branch of Logz. More general, the function f (z) = log z
is analytic on D with f 0 (z) = z1 and where D denotes the entire complex plane
deprived of the points on the semi line starting from the origin and making an
angle with the positive x-axis, i.e., D = C {z, |z| =
6 0, < arg z < + 2}.
d n
3. It is clear that f (z) = z n is entire and that dz
z = nz n1 .

4. We may use the exponential function to define the complex sine and cosine of
complex number. In fact, we have

eiz eiz eiz + eiz


sin z = and cos z = .
2i 2
d d
Moreover, sin z and cos z are entire and we have dz
cos z = sin z, dz
sin z = cos z.

Theorem 1.1.1. Consider a function f analytic in a domain D. If f 0 (z) = 0 in D then


f is constant over D.

1.2 Harmonic functions and harmonic conjugates


A real-valued function H(x, y) of 2 variables x and y is said to be harmonic in a given
domain D of C if throughout that domain, it has continuous partial derivatives of the
first and second order and satisfies the partial differential equation

Hxx + Hyy = 0 (1.6)

known as Laplaces equation. Here Hxx (resp. Hyy ) denotes the second partial derivative
of H with respect to x (resp. y).

Harmonic functions play an important role in applied mathematics. For example,


the temperatures T (x, y) in thin plates lying in the xy-plane are often harmonic. A
function V (x, y) is harmonic when it denotes an electrostatic potential that varies only
with x and y in the interior of a region of three-dimensional space that is free of charges.

Example 1.2.1. [Bro-Chu] The function T (x, y) = ey sinx is harmonic in any domain
of the xy-plane and, in particular, in the semi-infinite vertical strip 0 < x < , y > 0 .

16
It also assumes the values on the edges of the strip. More precisely, it satisfies all of the
conditions
Txx + Tyy = 0
Thus, T is harmonic in any domain of the xy-plane particularly in the semi - infinite
vertical strip (0 < x < , y > 0). More precisely, it satisfies all the conditions:


Txx (x, y) + Tyy (x, y) = 0,
T (0, y) = 0,


T (, y) = 0, (1.7)
T (x, 0) = sin x,




limy T (x, y) = 0,

which describe steady temperatures T (x, y) in a thin homogeneous plate in the xy-plane
that has no heat sources or sinks and is insulated except for the stated conditions along
the edges.
The next theorem provides a source of harmonic functions. In fact, using the Cauchy-
Riemann equations, we can state the following:
Theorem 1.2.1. If a function f = U + iV is analytic in a domain D, then U and V
are harmonic.
If two given functions U and V are harmonic in a domain D and their first order
partial derivatives satisfy the Cauchy-Reimann equations throughout D, then V is said
to be a harmonic conjugate of U .
Theorem 1.2.2. A function f (z) = U + iV is analytic in a domain D if and only if V
is the harmonic conjugate of U .
Remark 1.2.1. Let U and V be 2 functions in a domain D.
1. If V is a harmonic conjugate of U , then it is not always true that U is a har-
monic conjugate of V . Consider U (x, y) = x2 y 2 and V (x, y) = 2xy. A simple
calculation shows that Uxx + Vyy = 0, Ux = Vy and Uy = Vx , and as U and V are
continuous it follows that V is a harmonic conjugate of U and f (z) = U + iV = z 2
is analytic on any domain D. On the other hand g(z) = 2xy + i(x2 + y 2 ) is not
analytic on any domain D, which means that in this case U is not a harmonic
conjugate of V .
2. The harmonic conjugate of a function is unique up to adding a constant. Consider
the continuous function U (x, y) = y 3 3x2 y. We want to find a function V which
is a harmonic conjugate of U. Knowing that U and V satisfy the Cauchy-Reimann
equations, we get Ux = Vy which means that Vy = 6xy and thus V = 3xy 2 + (x).
Similarly, Uy = Vx so 3y 2 3x2 = 3y 2 0 (x) and 0 (x) = 3x2 and hence (x) =
x3 . Now we can write the harmonic conjugate of U which is V = 3xy 2 + x3 + C
where C is any real constant. It means that all the functions:
fC (z) = y 3 3x2 y + i(3xy 2 + x3 + k) = i(z 3 + C)
are analytic for all values of C R.

17
1.3 Complex integration
We are now able to define the integral of a complex function f . Complex integration
affords a lot of freedom due to how different the complex plane is from R.

Let f (t) be a complex-valued function of a real variable t, i.e., f (t)= U (t) + iV (t),
then the definite integral of f (t) over the interval a t b is given by
Z b Z b Z b
f (t)dt = U (t)dt + i V (t)dt.
a a a

Integrals of complex-valued functions of a complex variable are defined on curves in


the complex plane, rather than on just intervals of the real line. Classes of curves that
are adequate for the study of such integrals will be introduced now.
Definition 1.3.1. A set of points z = (x, y) in the complex plane is said to be an arc if
x = x(t), y = y(t) (a t b),
where x(t) and y(t) are continuous functions of the real parameter t. This definition
establishes a continuous mapping of the interval a < t < b into the xy-plane and the
image points are ordered according to increasing values of t. It is convenient to describe
the points of (C) by means of the equation
z = z(t) (a < t b),
where z(t) = x(t) + iy(t).
The arc (C) is a simple arc, or a Jordan arc, if it does not cross itself; that is, (C) is
simple if z(t1 ) 6= z(t2 ) when t1 6= t2 . When the arc C is simple except for the fact that
z(b) = z(a), we say that (C) is a simple closed curve, or a Jordan curve. An arc (C) is
positively oriented when it is in the counterclockwise direction.

An arc (C) is smooth if it has non vanishing, continuous derivative. Finally a con-
tour, or piecewise smooth arc (C) is an arc consisting of a finite number of smooth arcs
joined end to end. When only the initial and final values of z(t) are the same, a contour
C is called a simple closed contour. The length of a contour or a simple closed contour
is the sum of the lengths of the smooth arcs that make up the contour.

We turn now to integrals of complex-valued functions f of the complex variable


z. Such an integral is defined in terms of the values f (z) along a given contour (C),
extending from a point z = z1 to a point z = z2 in the complex plane. It is, therefore, a
line integral; Rand its value depends, in general, on the contour as well as on the function.
It is written (C) f (z)dz and defined by
Z Z b
f (z)dz = f (z(t))z 0 (t)dt,
(C) a

where z = z(t)(a t b) represents a contour (C) extending from a point z1 = z(a) to


a point z2 = z(b).

18
Theorem 1.3.1. [Greens Theorem] Let (C) be a positively oriented simple closed
curve in a plane, and let D be the region bounded by (C). If P and Q are two functions of
x and y defined on an open region containing D and have continuous partial derivatives
then, Z ZZ 
P Q 
P dx + Qdy = dxdy. (1.8)
(C) D x y
Definition 1.3.2. P (x, y) and Q(x, y) are continuously differentiable functions.
P Q
1. The differential P dx + Qdy is said to be closed on a domain D if y
= x
on D.
2. The differential of P dx + Qdy is said to be exact if there exists a differentiable
complex function h where dh = P dx + Qdy.
From the previous definition, we can say that exact differentials are closed (which
follows immediately from Cauchy-Riemann equations applied to the function h). More-
R (x ,y )
over, the integral (x01,y01) P dx + Qdy is independent of the path from (x0 , y0 ) to (x1 , y1 )
if the differential P dx + Qdy is exact, and in this case
Z x1 ,y1 )
P dx + Qdy = h(x1 , y1 ) h(x0 , y0 )
(x0 ,y0 )

Theorem 1.3.2. [Gam] Let P and Q be continuously differentiable complex functions


defined over a domain D. Suppose that:
1. D is a star-shaped domain (as a disk or a rectangle).
2. The differential P dx + Qdy is closed on D.
Then the differential P dx + Qdy is exact.
A complex function F is an antiderivative of f if F is analytic on D and F 0 (z) = f (z)
for all z D.
Proposition 1.3.1. [Bro-Chu] Suppose that a function f (z) is continuous in a domain
D, then the following statements are equivalent.
1. f (z) has an antiderivative F (z) in D.
2. The integrals of f (z) along contours lying entirely in D and extending from any
fixed point z1 to any fixed point z2 have the same value, and so they are independent
of the path chosen.
3. The integrals of f (z) along closed contours lying entirely in D have the value 0.
Next, we list some important integration theorems:

1. Cauchy-Goursat theorem[Bro-Chu]. If a function f is analytic at all points


interior to and on a simple closed contour (C) then:
Z
f (z)dz = 0
(C)

19
2. Cauchy integral formula[Bro-Chu]. Let f be an analytic function everywhere
inside and on a simple closed contour (C) taken in the positive sense. If z0 is any
point interior to (C) then:
Z
1 f (z)dz
f (z0 ) =
2i (C) z z0
and more general, we have
Z
(n) n! f (z)dz
f (z0 ) =
2i (C) (z z0 )n+1

3. Liouvilles theorem[Bro-Chu]. If f is an entire and bounded function in the


complex plane, then f (z) is constant throughout the plane.
Example 1.3.1. 1. As f (z) = ez is analytic at all points interior to and on the unit
circle then: Z
ez dz = 0
|z|=1

2. Let (C) is a positively oriented circle with equation |z| = 2, then by Cauchy integral
z
formula applied to the function f (z) = 9z 2 we get

Z Z z  i 
zdz 9z 2
2
= dz = 2if (i) = 2i =
(C) (9 z )(z + i) (C) z (i) 10 5

1.4 Complex series, residues and poles


Now
P we represent analytic functions using infinite series representations. Recall that
k=1 ak , where ak C, is said to converge to S if the sequence of partial sums {sk }
converges to S, where sk = a0 + a1 + . . . + ak .

A power series centered at z0 is a series having the form k


P
k=0 ak (z z0 )P . Making
a change of variable w = a z0 makes the powerP series centered at 0. If k=0 ak z
k
k
is a power series then there exists R > 0 such that k=0 ak z does not converge P whenk
|z| > R , converges absolutely for |z| < R , and for each r < R, the series k=0 ak z
converges uniformly when |z| r. The number R is called the radius of convergence of
the power series.

The power series representation of an analytic function f throughout the disk


|z z0 | < R0 is given by:

X f (n) (z0 )
f (z0 ) = (z z0 )k
n=0
n!

It is called the Taylor series expansion of f at z0 . When z0 = 0, it is called the


Maclaurin series expansion of f at z0 .

20
Suppose that f is analytic throughout an annular domain D : R1 < |z z0 | < R2 ,
and let (C) denote any positively oriented simple closed contour around z0 lying in D,
then the Laurent series representation of f on D is

X
n
X bn
f (z) = an (z z0 ) +
n=0 n=1
(z z0 )n

1
R f (z)dz 1
R f (z)dz
where an = 2i (C) (zz0 )n+1
and bn = 2i (C) (zz0 )n+1
.

Example 1.4.1. 1. The function f (z) = ez is an P


entire function with f (n) (z) = ez
and thus the Maclaurin expansion of ez is ez = zn
n=0 n! with |z| < .

1 1 1
2. [Bro-Chu] Consider the function f (z) = (z1)(z2) = z1 z2 which clearly has
two singular points z = 1 and z = 2. The function f is analytic in D1 P : |z| < 1,
D2 : 1 < |z| < 2 and D3 : 2 < |z| < . In the domain D1 , we have 1z = 1 n
n=0 z ;

now as | z2 | < 1, then 1(1 z ) = n=0 ( z2 )n . Noting that f (z) = 1z
1
+ 12 11 z , we get
P
2 2
that the Maclaurin expansion of f is:

X 1 X z n X n1
n
f (z) = z + ( ) = (2 1)z n
n=0
2 n=0 2 n=0

1 1 1 1
Similarly, in D2 we can write f (z) = z 1 z1
+ 2 1 z2
to get for 1 < |z| < 2 the series


X zn X 1
f (z) = +
n+0
2n+1 n=1 z n

1 1 1 1
In D3 , f (z) = z 1 z1
z 1 z2
and so


X 1 2n1
f (z) =
n+0
zn

3. Let be a closed positively oriented contour inside D1 : |z|R< 1 around 0 and let
1 1
f be the function defined by f (z) = (z1)(z2) . The integral z2 (z1)(z2) dz can be
computed directly using the Maclaurin series of f in D1 as

1
Z Z
f (z)dz
2
dz = = 2ia1
z (z 1)(z 2) z2

We already proved that an = 2n1 1 so a1 = 43 and

1
Z  3 1
2
dz = 2i = i
z (z 1)(z 2) 4 2

21
1
R
We can also compute z2 (z1)(z2) dz using the Cauchy integral formula. Take
1
f (z) = (z1)(z2) which is analytic everywhere inside and on (positively oriented).
Now as 0 is an interior point of , then by Cauchy integral formula:

1
Z Z
f (z)dz 2i 00
2
dz = 2
= f (0)
z (z 1)(z 2) z 2!
00 2 2 00
Knowing that f (z) = (z1)3
(z2)3
, we get f (0) = 12 , so:

1
Z
2i 00 1
dz = f (0) = i
z 2 (z 1)(z 2) 2! 2

To integrate and differentiate power series we use the following proposition:

Proposition 1.4.1. [Bro-Chu] Let (C)Pdenotes any contour interior to the circle of
convergence of the power series S(z) = n
n=0 an (z z0 ) .

1. The series formed by multiplying each term of the power series by g(z) can be
integrated term by term over (C) where g(z) is any continuous complex function
on (C). That is:
Z
X Z
g(z)S(z)dz = an g(z)(z z0 )n dz
(C) n=0 (C)

2. S(z) is differentiable and we have



X
0
S (z) = nan (z z0 )n1 .
n=1

Definition 1.4.1. 1. A point z0 is called a singular point if f is not analytic at z0 but


is analytic in every neighborhood of z0 . A singular point is isolated if in addition
to being singular, we can find a deleted neighborhood 0 < |z z0 | < ,  > 0 , of
z+1
z0 where f is analytic. The function f (z) = z2 (z 2 +1) has singular points z = 0 and

z = i.

2. Let z0 be an isolated singular point of a function f, then a positive number R2


domain D: 0 < |z P
such that f is analytic in the P z0 | < R2 , and so f has a Laurent
series representation f (z) = n=0 na (z z0 ) n
+ bn
n=1 (zz0 )n . The residue of f at
1
R
the singular point z0 is Resz=z0 f (z) = b1 = 2i (C)
f (z)dz, where (C) is any closed
positively oriented contour in D around z0 .

3. An isolated singular point z0 of a function f is said to be a pole of order m if f(z)


S(z)
can be written in the form f (z) = (zz 0)
m where S(z) is analytic and non zero at

z0 and m N. Moreover,

(a) Resz=z0 f (z) = S(z0 ) if m=1.

22
S (m1) (z0 )
(b) Resz=z0 f (z) = (m1)!
when m 2.
Theorem 1.4.1 (Cauchys residue theorem). [Bro-Chu]. Let (C) be a simple closed
positively oriented contour. If a function f is analytic inside (C) and on (C) except on
a finite number of singular points zk (k = 1, 2, . . . , n) inside (C) then:
Z X n
f (z)dz = 2i Res f (z). (1.9)
(C) z=zk
k=1

5z2
R
Example 1.4.2. Consider (C) z(z1) dz, where (C) is the circle |z| = 2. The function
5z2
f (z) = z(z1) has two isolated singularities inside (C) : z = 0 and z = 1, now we need to
find r1 and r2 which are the respective residues of f at z = 0 and z = 1 . As 0 < |z| < 1,
1
it is clear that f (z) = 5z2
z 1z
= (2 z2 )(1 z z 2 . . .), and so r1 = 2 as it is b1 in
the Laurent series. Similarly in 0 < |z 1| < 1, we can write the Laurent series to get
r2 = b2 = 3 and thus
Z X2
f (z)dz = 2i Res f (z) = 2i(r1 + r2 ) = 2i(5) = 10i
(C) z=zk
k=1

Another approach for finding r1 and r2 is by noting that z = 0 and z = 1 are both poles
5z2
z1 5z2
of order 1 with respect to f , as f (z) = z
, so r1 = g(0) = 2, where g(z) = z1
. To
5z2
5z2
find r2 , take f (z) = z
z1
and so r2 = h(1) = 3, where h(z) = z
.
Definition 1.4.2. A number z0 is a zero of f with order m if the function f is analytic
at z0 , all derivatives f (n) (z) at z0 exist, f (z0 ) = 0, f (k) (z0 ) = 0 (k = 1, 2, . . . m 1) and
f (m) (z0 ) 6= 0 .
Next, we provide some theorems relating poles and zeroes of analytic functions.
Theorem 1.4.2. [Bro-Chu] A function f analytic at z0 ) has z0 a zero of order m if and
only if there exists a function g, which is analytic and nonzero at z0 such that
f (z) = (z z0 )m g(z)
From the previous theorem, if two functions P (z) and Q(z) are both analytic at z0
and P (z0 ) 6= 0 and z0 is a zero with order m of P (z), then z0 is a pole of order m for
P (z)
f (z) = Q(z) . Now, if P and Q are two analytic functions such that P (z0 ) 6= 0 and z0 is
P (z)
a zero with order 1 for Q(z), then z0 is a pole with order 1 for f (z) = Q(z)
and

P (z0 )
Res f (z) =
z=z0 Q0 (z0 )
Example 1.4.3. To find the residue of the function f (z) = z4z+4 at the isolated singu-
larity z = 1+i, we take P (z) = z and Q(z) = z 4 +4 as P (1+i) = 1+i 6= 0, Q(1+i) = 0
and Q0 (1 + i) = 8 + 8i 6= 0, then 1 + i is a zero of order 1 for Q and so a pole of order
1 for f . Hence,
P (1 + i) 1+i i
Res f (z) = 0 = =
z=1+i Q (1 + i) 8 + 8i 8

23
Definition 1.4.3. A function f is meromorphic on D if f (z) has only isolated singu-
larities in D, which are all poles.

Theorem 1.4.3. [The argument principle][Bro-Chu]. Suppose that f(z) is analytic


and non-zero on a closed contour (C), and meromorphic inside (C) then:

f 0 (z)
Z
1
dz = N0 (f ) Np (f ) (1.10)
2i (C) f (z)

Where N0 (f ) and Np (f ) denote the number of zeroes and poles of f in D respectively,


1
R f 0 (z)
counted with multiplicities. In particular, if the function f is analytic then 2i (C) f (z)
dz
represents the number of zeroes of f

Proof. Since the zeroes and poles of f are isolated, we can deform (C) into a set
of small disjoint external circles (C ) centered at each zero or pole. It follows that it
is enough to prove the theorem for each circle separately and then add the results. In
such a circle (C ) centered at z0 we can write f (z) = (z z0 )m h(z), where m > 0 if z0
is a zero with order m and m < 0 if z0 is a pole with order m and where h(z0 ) 6= 0.

f 0 (z) m(z z0 )m1 h(z) + h0 (z)(z z0 )m m h0 (z)


= = + .
f (z) (z z0 )m h(z) z z0 h(z)

Thus, we have
f 0 (z)
Z Z
1 1 m
dz = +0=m
2i (C ) f (z) 2i (C ) z z0

24
Chapter 2

Conformal Mapping

In this chapter, we introduce the concept of a conformal mapping with emphasis on


properties, fundamental examples and connections between such mapping and harmonic
functions ([Gam, Bro-Chu, Pon-Sil, Olv, Kar, Bak-New, Blair, Gre-Kra, Neh]).

2.1 Mappings by elementary functions


We start by giving examples on mappings by elementary complex functions.

2.1.1 Linear transformations


To geometrically illustrate a mapping of a complex variable it is convenient to consider
two different planes with rectangular coordinates. These planes are called the z-plane
and the w-plane. Mappings of a complex variable can be illustrated graphically by
indicating correspondences between sets of points in these two planes. Let z = x + iy
represents a complex number in the z-plane and let w = u + iv represent the image of
z in the complex w-plane. For S a set of complex numbers in the z-plane we define
a single valued mapping f on S as a rule which assigns to each value z = x + iy in S
one and only one complex number w = u + iv = u(x, y) + iv(x, y) of the w-plane. The
functional relation is then denoted by:

w = u + iv = f (z) = f (x + iy) = u(x, y) + iv(x, y)

And so if f (z) = z 2 the image of A(1,1) is the point A(0,2) and so A has coordinates
(1,1) in the z-plane and A has coordinates (0,2) in the w-plane.
A general linear transformation transforming a point with affix z into an image point
with affix w is given by w = Az + B, where A and B are nonzero complex constants.
We begin with the study of the transformation with complex form w = z + B.

1. The transformation with complex form w = z + B, where B is a nonzero complex


number, is the translation of vector V~ , where V~ has an affix B. Analytically
speaking, it means that this transformation maps the points with coordinates
(x, y) onto points with coordinates (x + xV , y + yV ), where V~ has coordinates

25
(xV , yV ). The translation maps all figures in the z-plane into congruent figures in
the w-plane.

2. Consider the transformation defined by its complex form w = Az, with the con-
dition that A = rei 6= 0, so we can write w = aei z and it follows directly that
this transformation expands or contracts radius vectors with a scale |A| = r and
rotates them with angle = arg(A) about the origin. So this transformation maps
all figures in the z-plane onto similar figures extended/contracted and rotated in
the w-plane.

3. The general linear transformation with complex form w = Az + B is often called


a direct plane similitude, where A = rei 6= 0. This transformation is clearly the
composite of w = Az and w = z + B. So this transformation is basically an ex-
pansion/contraction and a rotation followed by a translation and so it maps figures
in the z-plane onto similar figures in the w-plane.

Proposition 2.1.1. Linear transformations preserve oriented angles.

Proof. Let (~a, ~b) be an oriented angle in the z-plane and (a~0 , b~0 ) be its image in the
w-plane by a translation thus ~a = a~0 and ~b = b~0 and so (~a, ~b)= (a~0 , b~0 ) up to adding a
multiple of 2 and thus a translation preserves oriented angles, a similar argument but
now the mapping w = Az is considered where A = rei then:
z   Az  z 
b~0 ~b ~b
(a~0 , b~0 ) = arg = arg = arg = (~a, ~b) + 2ki where k Z.
za~0 Az~a z~a

Thus, linear transformations preserve oriented angles.

Example 2.1.1. Consider the linear transformation with complex representation w=


(1 + i)z + 2. This transformation expands figures with a scale factor |1 + i| = 2
and rotates them about the origin by an angle = arg (1 + i) = 4 [2] followed by a
translation of the vector V~ (2, 0). The image of the circle (C)with center O(0, 0) and
radius r = 2 is the circle (C 0 ) with center A(2, 0) and radius 2 2.

1
2.1.2 The transformation w = z
Consider for this section that all complex numbers involved are different than zero.
The mapping with complex form w = z1 will be described as the composite of two
transformations with respective complex forms w1 = z1 and w2 = z. Clearly, the first
transformation is an inversion with pole O and power 1, or simply an inversion with
respect to the unit circle, and the second transformation is the symmetry with respect
to the axis of abscissas. Noting that an inversion transforms lines into lines and cir-
cles and circles into lines and circles depending whether or not O belongs to the curve,
while the symmetry conserves geometric figures but not oriented angles, it will lead
to the result that the mapping w = z1 will transform circles into lines and circles and
circles into lines and circles. The mapping w = z1 is an involution, which means that
it is its own inverse mapping. Now, we study how this mapping maps circles and

26
straight lines. As an initiative, we need to find the analytic form of this mapping. Let
z = x + iy and w = u + iv. By comparing real parts and imaginary part we get that
x y
u = x2 +y 2 and v = x2 +y 2 . As this mapping is an involution we can directly conclude
u v
that x = u2 +v 2 and y = u2 +v 2 . The general equation of a line or a circle is given by the

following equation: A(x2 + y 2 ) + Bx + Cy + D = 0, where B 2 + C 2 4AD > 0. This


equation represents a circle when A 6= 0, while when A = 0, it represents a straight line.
This can be seen clearly when rewriting the equation after completing its square to get
B 2 C 2 2 2 4AD 2
(x + 2A ) + (y + 2A ) = ( B +C 2A
) . If M (x, y) is a point on the general curve it
u
means that its coordinates verify A(x + y 2 ) + Bx + Cy + D = 0, and as x = u2 +v
2
2 and
v 2 2
y = u2 +v2 then the coordinates of the image of M verify D(u + v ) + Bu Cv + A = 0
which is also a general equation of a line or circle. This lead to the following conclusions:

1. The image of a circle (A 6= 0) which passes through the origin (D = 0) will be


a straight line with equation Bu Cv + A = 0 which doesnt pass through the
origin.

2. The image of a circle (A 6= 0) which does not pass through the origin (D 6= 0) is a
circle with equation D(u2 + v 2 ) + Bu Cv + A = 0 which does not pass through
the origin.

3. The image of a straight line (A = 0) which passes through the origin (D = 0) is a


line with equation Bu Cv = 0 which passes through the origin.

4. The image of straight line (A = 0) which does not pass through the origin (D 6= 0)
is a circle with equation D(u2 + v 2 ) + Bu Cv = 0 which passes through the ori-
gin. We could have foreseen this result with the mapping w = z1 being its own
inverse, and from part (1) we know that circles passing through the origin will
be mapped onto straight lines not passing through origin and so straight lines
not passing though the origin should be mapped onto circles passing through the
origin. All the above curves are such that they are deprived of the origin in case
they pass through the origin.

2.1.3 The mapping w = sin z


We know that sin z = sin(x + iy) = sin x cosh y + i cos x sinh y and thus the mapping
w = sin z can be regarded as w = u(x, y) + iv(x, y) where u(x, y) = sin x cosh y and
v(x, y) = cos x sinh y. Our aim here is to show that w is a one-to-one mapping of the
semi-infinite strip 2 x 2 , y 0 in the z-plane onto the upper half plane v 0 in
the w-plane. We will start by proving that the boundary of this strip is mapped in a
one-to-one manner onto the real axis in the w-plane. Taking A and E be any points on
x = 2 with y 0 gives that A and E will have coordinates: u = cosh y and v = 0
and so A and E are on the real u-axis in the w-plane, moreover as cosh y is increas-
ing and greater or equal to 1 when y 0 thus the two semi lines x = 2 with y 0
will be mapped onto the intervals ] 1] [1, +[. Now we consider any point in
the strip 2 < x < 2 and y = 0 which will have coordinates (sin x, 0) in the w-plane
and thus is mapped onto the interval ] 1, 1[ in the w-plane; now we can consider the

27
interior of this strip; each point in the region 2 < x < 2 and y 0 lies on one of
the vertical semi line x = c1 with y 0 where c1 ] 2 , 2 [ and those semi lines will
be mapped onto u = sin c1 cosh y and v = cos c1 sinh y where y [0, +[, noting that
cosh2 y sinh2 y = 1 we get the the image of the lines are branches of the hyperbolas
on the upper side of the v-axis in the w-plane (as v = cos c1 sinh y 0) with equation:

u2 v2
=1
sin2 c1 cos2 c1
p
whose foci have at the points sin2 c1 + cos2 c1 = 1 in the w-plane; see Figure 2.1.
So we notice that the images of those semi lines are distinct and constitutes the entire
half plane v 0. More precisely if we are moving c1 from 0 to 2 say a point L then
L will be on a right branch of a hyperbolas which will be opening wider and its vertex
(sin c1 , 0) is moving towards the origin w = 0, a similar argument is made for a point M
c1 is moving from 2 to 0, and this achieves our aim.

Figure 2.1: w = sin z

2.1.4 Linear fractional transformations


The transformation with complex form w = az+b cz+d
where (ad bc 6= 0) is called linear
fractional transformation, Mobius transformation or bilinear mapping. To study this
transformation we rewrite its complex form as w = ac + bcad 1
c cz+d
. This equation reveals
that this mapping can be studied as the composite of three mappings with respective
complex forms w1 = cz + d, w2 = z1 and w3 = ac + bcad c
z, and this means that Mobius
transformations map circles and lines into circles and lines. Now we extend T (z) = az+b
cz+d
to have as a domain of definition the extended z-plane denoted by C . Knowing
that T (z) = az+b
cz+d
when ad bc 6= 0, we now write T () = if c = 0, T () = ac when
c 6= 0 and T ( dc ) = when c 6= 0 . This new function T is continuous and clearly one
to one and thus the inverse transformation T 1 exists with the following properties:
1. T 1 (w) = z if and only if T (z) = w.
dw+b
2. T 1 (w) = cwa
, so T 1 is itself a fractional transformation.

3. T 1 () = when c = 0.

4. T 1 ( ac ) = if c 6= 0.

5. T 1 () = dc if c 6= 0.

28
Example 2.1.2. The linear fractional transformation with complex form:
z
(z) = with || < 1
z 1
maps the unit disk into itself, moving the origin z=0 into = . To illustrate this we
note the following relations:

|z |2 = (z )(z ) = |z|2 z z + ||2 (2.1)

|z 1|2 = (z 1)(z 1) = ||2 |z|2 z z + 1 (2.2)


Subtracting Equation 2.2 from Equation 2.1, we get:

|z |2 |z 1|2 = (1 ||2 )(|z|2 1),

which is negative whenever || < 1 and |z| < 1 which means that || < 1.

2.1.5 Linear fractional transformations of the upper half plane


Let us determine all linear fractional transformations that map the upper half plane
Im z > 0 onto the open disk |w| < 1, and the boundary Im z = 0 onto the boundary
|w| = 1.

We select three points on Im z = 0: z = 0, z = 1 and z = , which are to be trans-


formed onto |w| = 1, and lets determine a linear fractional transformation T with com-
plex representation w = az+b
cz+d
satisfying those conditions. As the image of z = 0 is on
b
|w| = 1, it follows that | d | = 1, and so |b| = |d| =
6 0. Now the image of should also
be on |w| = 1; we know that T () = c when c 6= 0 which means that | ac | = 1 and so
a
b
a z+ a
|a| = |c| =
6 0. Hereby we can rewrite the complex form of T as T (z) = c z+ d
, and as
c
| ac | = 1, then we can write the complex number a
c
in its exponential a
form c = ei , and
b
i z+ a
so T (z) = e z+ dc
= ei zz
zz1
0
, where |z0 | = |z1 |. In addition, we know that the image of
z = 1 is also on |w| = 1, which gives that |1 z0 | = |1 z1 | and so

(1 z0 )(1 z0 ) = (1 z1 )(1 z1 ).

Knowing that z0 z0 = z1 z1 = 1, it follows immediately that z1 = z0 or z0 = z1 , so z0 = z1


since if z1 = z0 then T = ei which is constant. This mapping transforms z0 into the
origin w = 0 which is inside |w| = 1 which means that Im z0 > 0. So the linear fractional
transformations mapping the upper half plane Im z > 0 onto the open disk |w| < 1 and
zz0
the boundary Im z = 0 into the boundary |w| = 1 have the form w = ei z z0
, where
Im z0 > 0. Now it remains to prove conversely that any map having the form w = ei zz 0
zz0
with Im z0 6= 0 has the desired property which is very easy to see as
z z0 z z0
|w| = |ei | = |ei || | = 1.
z z0 z z0

29
It is very clear to see the result geometrically. If z is above the x-axis, then the distance
joining z to z0 is smaller than the distance from z to z0 as z0 is situated above the
x-axis, and so for any point z above the x-axis, its image w will have |w| < 1 and thus
the region y > 0 will be mapped onto |w| < 1; while if a point z is below the x-axis, its
image w will have |w| > 1, and if a point z = k where k R is on the x-axis, we have
|k zo | = |k z0 | = |k z0 | and thus its image w will be on |w| = 1.
Example 2.1.3. [Bro-Chu, page 327-328]
1. The transformation w = iz
i+z
zi
can be written as w = ei zi
and thus maps the upper
half plane Im z > 0 onto the open disk |w| < 1 and the boundary Im z = 0 into the
boundary |w| = 1.
z1
2. Consider the transformation w = z+1 . Note that this transformation doesnt satisfy
the conditions for a linear fractional transformation that maps the upper half plane
Im z > 0 onto the open disk |w| < 1 and the boundary Im z = 0 into the boundary
|w| = 1 as, here, z0 = 1 with Im z0 = 0. Writing z = x + iy and w = u + iv gives
that this mapping maps the upper half plane y > 0 onto the upper half plane v > 0
2y
and the x-axis onto the u-axis. To make this clear, v = Im w= |z+1| 2 which means

that y and v have the same sign and v = 0 when y = 0.


3. Let T be the transformation with complex form
z1
w = Log , (2.3)
z+1
where the principle branch of the logarithmic function is used. T is recognized as
the composite of two transformations with respective complex forms
z1
Z= (2.4)
z+1

w = Log Z, (2.5)
where z = x + iy, Z = X + iY and w = u + iv. We have seen (part 2 above) that
z1
the mapping Z = z+1 maps the upper region y > 0 into the upper Y > 0; now, we
check the image of the upper region by Log Z. We can write Z = Rei and
Log Z = ln R + i,
where pi < < and R > 0.
We see that as a point Z = Rei0 (0 < 0 < ) moves outward from the origin
along the ray = 0 , its image is the point whose rectangular coordinates in the
w-plane are (ln R, 0 ). That image evidently moves to the right along the entire
length of the horizontal line v = 0 . Since these lines fill the strip 0 < v <
as the choice of 0 varies between 0 and , the mapping of the half plane Y > 0
onto the strip is, in fact, one to one. This shows that the composition (2.3) of
the mappings (2.4-2.5) transforms the plane y > 0 onto the strip 0 < v < (see
Figure 2.2).

30
Figure 2.2: w = Log Z

2.2 Conformal mappings


We are now ready to define a conformal mapping and study its properties.

2.2.1 Definition of conformal mappings


Definition 2.2.1. Given two curves (C1 ) and (C2 ) meeting at a point A then the angle
between the two curves at A is the angle between the two tangents at A to (C1 ) and (C2 ).

Example 2.2.1. 1. The angle between the line (m) with equation x = 0 and the line
(n) with equation y = x at the origin is clearly 4 , now we check the angle of their
images under f (z) = z 2 in the w-plane. The image (m0 ) of (m) has an equation
v = 0 with u < 0 while the image (n0 ) of (n) has an equation u = 0 with v > 0,
meeting at the origin and thus making a right angle at the origin.

2. Consider the two curves (C1 ) and (C2 ) with respective equations y = x and y = x2
meeting at the point A(1, 1). An equation of the tangent to (C1 ) at A is y = x and
s1 s2
to (C2 ) is y = 2x , now the angle between the curves is such that = | 1+s 1 s2
|,
where s1 and s2 are the respective slopes of the tangents to (C1 ) and (C2 ). We
have tan = 13 and = tan1 ( 13 ). In general we do not need to find the equation
dy
of the tangent as all we need is to find dx at the point A which represents the slope
of the tangent.

3. Consider the mapping f (z) = z 2 from the z-plane onto the w-plane and lets see
the angle between the image of the vertical line (d) with equation x = 1 and the
image of the horizontal line (L), y = 2 in the w-plane at the image B of the point
A(1, 2). First, writing z = x + iy, we get that w = u + iv = x2 y 2 + 2xyi and
so u = x2 y 2 and v = 2xy. The image of (d) is found by letting x = 1 and so
u = 1 y 2 and v = 2y and so the image of (d) is the parabola (P ) in the w-plane
with equation v 2 = 4u + 4. Similarly for (L), we get its image in the w-plane
which is the parabola (Q) with equation v 2 = 16u + 64 meeting at B(3, 4). Clearly
dv
(d) and (L) meet. The slope of tangent to (P ) at B in the w-plane is du and as we
have the relation between u and v given implicitly, we use implicit differentiation
dv
to get 2v du dv
= 4, meaning that du = 2
v
= 21 at the point B. Similarly for (Q),
dv 8
we get du = v = 2 at the point B, and it is clear that they meet at a right angle
as 12 2 = 1.

31
If two curves (C1 ) and (C2 ) meet at a point A making an angle then it is not
necessary that the images of those curve (meeting at the image of A ) under a mapping
f meet making the same angle as we saw in the previous examples. This motivates us
to define the mappings which preserve angles.
Definition 2.2.2. A mapping f : U V is conformal (or angle preserving) at a point
z0 U if it preserves the oriented angles between all curves through z0 . A mapping
f : U V is said to be conformal over U if it is conformal at every point z in U.
A more useful definition for conformal mapping is as follows:
Definition 2.2.3. [Olv, page 2.] Let f : U R2 defined by f (x, y) = u(x, y)+iv(x, y),
be a a mapping with continuous partial derivatives and let (x0 , y0 ) be any point in U ,
then f is said to be conformal at (x0 , y0 ) if r R+ and [0, 2[ such that for any
differentiable curve : [a, b] U with (0) = (x0 , y0 ) we have:

|(f )0 (0)| = r| 0 (0)|,


arg(f )0 (0)) = + arg( 0 (0)).

This means that when any differentiable curve through (x0 , y0 ) is mapped by f, its
tangent vector at (x0 , y0 ) is stretched by a factor r and rotated by an angle where r
and are independent of the curve. This means that oriented angles will be preserved.
In complex terms the word conformality is nothing but differentiability as we will
see in the next proposition.
Proposition 2.2.1. [Bro-Chu] Let f : U V where U, V C be a complex function
with continuous partial derivatives and let z0 U , then f is conformal at z0 if and only
if f is analytic at z0 and f 0 (z0 ) 6= 0.
Proof. Suppose that a given function f is analytic at z0 with f 0 (z0 ) 6= 0. Let
r = |f 0 (z0 )| , and let = arg(f 0 (z0 )) where [0, 2[. If is any differentiable curve
with (0) = z0 and using the chain rule we get

(f )0 (z0 ) = f 0 ((z0 )) 0 (0) = f 0 (z0 ) 0 (0)

|(f )0 (z0 )| = |f 0 (z0 )|| 0 (0)| = r| 0 (0)|


Moreover:
arg(f )0 (z0 ) = arg(f 0 (z0 )) + arg( 0 (0)) = + arg( 0 (0)),
and so f is conformal at z0 . Now, conversely, suppose that f = u + iv is conformal
at z0 = x0 + iy0 as f has continuous partial derivatives, then we need to show that f
satisfies the Cauchy-Riemann equations at z0 in order to be analytic at z0 . Consider
the curve (t) = z0 + t starting at z0 where t is close to 0. Then:

(f )(t) = u(x0 + t, y0 ) + iv(x0 + t, y0 )

Derive with respect to x to get (f )0 (t) = ux (x0 + t, y0 ) + ivx (x0 + t, y0 ) and so

(f )0 (0) = ux (x0 , y0 ) + ivx (x0 , y0 ) = fx (x0 , y0 ).

32
Similarly, taking the curve (t) = z0 + it then deriving with respect to y we get: (f
)0 (0) = uy (x0 , y0 )+ivy (x0 , y0 ) = fy (x0 , y0 ). Noting that 0 (0) = i 0 (0) we get | 0 (0)| =
| 0 (0)| and arg( 0 (0)) = 2 + arg( 0 (0)). As f is conformal at z0 , then r > 0 and
[0, 2[ such that |(f )0 (0)| = r|0 (0)| and arg(f )0 (0) = + arg(0 (0) for any
differentiable curve starting at z0 , and so we have:

|fx (x0 , y0 )| = |(f )0 (0)| = r| 0 (0)| = r| 0 (0)| = |(f )0 (0)| = |fy (x0 , y0 )|

   
arg fx (x0 , y0 ) = arg (f )0 (0)
 
0
= + arg (0)
 
= + arg 0 (0)
  2
0
= arg (f ) (0)
  2
= arg fy (x0 , y0 )
2
This means that fx (x0 , y0 ) = ify (x0 , y0 ), and so:

ux (x0 , y0 ) + iv(x0 , y0 ) = vy (x0 , y0 ) iux (x0 , y0 )

Taking the real and imaginary parts we get that the Cauchy-Riemann equations are
verified and thus f is analytic at z0 with f 0 (z0 ) 6= 0 as a direct conclusion from the
existence of the scale factor r > 0 as f being conformal.

2.2.2 Examples of conformal mappings


1. The exponential function is conformal at all points z. Every nth power func-
tion is conformal except at z = 0 . Every branch of the complex logarithm
or any nth root function is conformal over its domain (which cannot include
branch points).

2. Linear fractional transformations are conformal indeed. To prove this we saw


earlier that translation, inversion then reflection about the real axis (w = z1 ), ho-
mothecy and rotation preserve oriented angles and are thus conformal. We will see
later that the composite of two conformal mappings is conformal and we can prove
that any linear fractional transformation with complex form f (z) = az+b
cz+d
such that
ad bc 6= 0 is indeed conformal as f = f4 f3 f2 f1 where f1 (z) = z + dc which
is a translation, f2 (z) = z1 which is an inversion followed by reflection about the
real axis, f3 (z) = bcad
c2
z which is a homothecy and rotation and f4 (z) = z + ac
which is again a translation. Remark that later we might write that fractional
linear transformations map circles into circles, this is true as a straight line will
be considered as a circle passing through .

33
3. The Joukowski map [Olv] was first deployed to study flows around airplane
wings by the pioneering Russian aero and hydro-dynamics researcher Nikolai
Zhukovski (Joukowski). Joukowswis map is defined as follows:
1 1
(z) = (z + ).
2 z
d
Since dz = 12 (1 z12 ) = 0 if and only if z = 1, then the Joukowski map is con-
formal at all points except z = 1, as well as the singularity z = 0 where it is
not defined. Consider z = rei . It follows that (z) = 12 r(ei + ei ) and so when
z = ei is on the unit circle centered at the origin then (z) = cos and thus this
mapping maps the unit circle onto the interval [1; 1] on the real axis, and so
the points inside and outside the unit circle will be mapped to fill the rest of the
plane. Finding the inverse map is determinedpby finding z(), and a simple
calculation leads to the following result: z = 2 1, this means that every
6= 1 comes from two different points z, if / [1, 1] then came from two
points p p
z1 = + 2 1 and z2 = 2 1,
where z1 is clearly outside the unit circle and z2 is inside the unit circle whereas
if ] 1, 1[ we saw that both points z1 and z2 lie on the unit circle with same
abscissa in the z-plane.

Defining the Joukowski map over the region |z| > 1 defines a one to one conformal
mapping onto the exterior of segment [-1,1] in the -plane C\[1, 1]. Under the
Joukowski map, the concentric circles |z| = r 6= 1 are mapped into ellipses with
foci at 1 in the -plane, the effect of this mapping on circles not centered at the
origin can take wide variety of shapes depending whether the circle passes through
the singular z = 1, and the image of such circles are not smooth anymore but has
a cusp at = 1.

Figure 2.3: Image of circles |z| = r (see [Olv])

34
Figure 2.4: Airfoils Obtain Under the Joukowski map [Olv])

4. The Stereographic projection [Olv] is also conformal. Let S2 = {(x, y, x)


R2 : x2 + y 2 + z 2 = 1} be the unit sphere and let n denote the north pole (0,0,1),
the stereographic projection map : S2 n C is defined by
x + iy
(x, y, z) = .
1z
The Stereographic projection is described geometrically as a mapping that maps
any point p S2 other than n into a point (p) in the complex plane such that
the image is where the straight line joining n and p cuts the xy-plane, that is:
(p) = l(n, p) R2 where l(n, p) = {(1 t)n + tp : t R} is the line through n
and p. So to find the intersection point (p), any point on the line has coordinates
(1 t)(0, 0, 1) + t(x, y, x), in particular the last coordinate is 1 t tz, and it is
1 x y
null when t = 1z and thus (p)( 1z , 1z , 0). Conformal mapping means that it
preserves the angles between curves. Take a point p S2 n, let Tp denotes the
tangent plane to S2 at p, and let Tn denotes the tangent plane to S2 at n. We
have equal angles and right angles between the radii and the tangent plane, see
Figure 2.5 , hence equal angles , hence equal angles 0 and so equal lengths b.

Figure 2.5: Side view stereographic projection map

Now, let be a smooth curve on S2 through p, and let t be its tangent at p and
t be the intersection of the plane containing n and t with R2 . Clearly t is the

35
tangent to at (p), since and t are curves in R3 with the same tangent t
at p, it follows that and t = t are curves in R2 with the same tangent
at (p). Since t is its own tangent at (p), it is also tangent to there. The
lengths b are equal and so finally the angles are also equal. Similarly, taking a
second curve through p completes the proof (see Figure 2.6).

Figure 2.6: Top view stereographic projection map [Olv]

2.3 Properties of conformal mappings


We finish this chapter by giving crucial properties of conformal mappings.

Proposition 2.3.1. The composition g f of two conformal maps f and g is conformal


on a domain D when f 0 (z) 6= 0 and g 0 (f (z)) 6= 0 for all z D.
 
Proof. This results directly from using the chain rule as dz d
g f (z) = g 0 f (z) f 0 (z).

Theorem 2.3.1 (Rouches theorem). For any two complex-valued functions f and g
analytic inside some region K with closed contour K, if |g(z)| < |f (z)| on K, then f
and f + g have the same number of zeros inside K, where each zero is counted as many
times as its multiplicity.

Proof. For t [0, 1], let:

f 0 (z) + tg 0 (z)
Z
1
(t) = dz.
2i (K) f (z) + tg(z)

As,
| |f (t)| t|g(t)| | | |f (t)| |g(t)| | > 0

36
thus the denominator in the integrand is never zero. Noting that is continuous in the
interval [0,1] and is integer valued. Clearly from the argument principle (see Theorem
1.4.3), (t) is the number of zeroes of the function f + tg inside (K). Being continuous
and integer valued on the connected set [0, 1] then it must be constant. In particular
1
R f 0 (z)
(0) = (1). But (0) = 2i (K) f (z)
is the number of zeroes of f as f is analytic and
1 f 0 (z)+g 0 (z)
(1) = 2i f (z)+g(z)
is the number of zeroes of f + g, and thus the result follows.
Lemma 2.3.1. Suppose that a non-constant function f from D to D0 is analytic, and
let z0 D such that f 0 (z0 ) = 0, then f cannot be a one-to-one mapping in any disk
containing z0 .
Proof. Let w0 = f (z0 ). Since f (z) is not constant then the function g(z) = f (z) w0
is not an identically zero function. Since f 0 (z0 ) = 0, then g(z) has a zero of finite order
at least 2 at z0 . Since zeroes of analytic functions are isolated, then we can find r > 0
such that g(z) and f 0 (z) have no zeroes in the punctured disk {0 < |z z0 | r}.
Then m = min |g(z)| for z A, where A = {z : |z z0 | = r}. Let w C such that
0 < |w w0 | < m, then it follows from Rouches theorem that the functions g(z)
and g(z) + (w0 w) have the same number of zeroes inside A and so the function
g(z) + (w0 w) = f (z) w has at least two zeroes inside A none of them is z0 as
w 6= w0 . Since f 0 (z) 6= 0 inside the punctured disk, it follows that the zeroes must be
simple and distinct, and so f cannot be one-to-one inside any disk containing z0 .
Corollary 2.3.1. Suppose that a function f is analytic and one-to-one in a domain D,
then f is conformal at every point in D.
Proof. Let z0 D, since f is one-to-one and analytic in D, we have by Lemma 2.3.1
that f 0 (z0 ) 6= 0 and thus f is conformal at z0 and thus in D.
Lemma 2.3.2. Suppose that D and D0 are domains, and let f be an analytic one-to-one
and onto mapping, then the inverse mapping f 1 from D0 to D is analytic in D.
Proof. The proof just follows from the fact that
z z0  w w0 1
lim = lim 6= 0
w w0 w w0 z z0 z z0

This limits being not zero is a consequence from f being one-to-one and onto which
guarantees f 0 (z) 6= 0 by Corollary 2.3.1.
Remark 2.3.1. If f is analytic and one-to-one on a domain D then by applying the
Cauchy-Riemann equations on the inverse map f 1 we have:
x y
= , (2.6)
u v

x y
= . (2.7)
v u
Moreover, we have
2x 2x 2y 2y
+ =0 and + =0 (2.8)
u2 v 2 u2 v 2
37
Theorem 2.3.2. Suppose that D and D0 C are domains, and let f be an analytic
one-to-one and onto mapping from D to D0 . Suppose further that for every z in D, we
write w = f (z), where z = x + iy and w = u + iv, with x, y, u, v R. Then a function
(x, y) is harmonic in D, if and only if the function (u, v)is harmonic in D0 , where
(u, v) = x(u, v)), y(u, v) .
 
Proof. Note that (u, v) = x(u, v)), y(u, v) can be written as (w) = f 1 (w)


and that the function f 1 is analytic on D0 . Using the chain rule we get:
x y
= +
u x u y u
Now,
2  x y  2 x x   2 y y  
= + = + + 2 + .
u2 u x u y u u2 x u u x u y u u y
On the other hand, we have:
  x   y   x 2 y 2
= + = + .
u x u x x u y x u x u xy
Similarly we get:
  x 2 y 2
= + .
u y u xy u y 2
Remark here that order partial differentiating x or y does not matter due to the conti-
nuity of the functions. Combining all terms to get:
2 2 x 2 y  x 2 2 x y 2  y 2 2
= + + +2 + (2.9)
u2 u2 x u2 y u x2 u u xy u y 2
By symmetry of differentiation and in a similar manner we get:
2 2 x 2 y  x 2 2 x y 2  y 2 2
= 2 + + +2 + (2.10)
v 2 v x v 2 y v x2 v v xy v y 2
Now adding Equations 2.9 and 2.10 to get
 2 x 2 x   2 y 2 y 
4 = + + +
x u2 v 2 y u2 v 2
 x y x y  2
+2 +
u u v v xy
 x 
2 x 2  2  y 2 y 2  2
+ + ) + + ( (2.11)
u v x2 u v y 2
Using (2.8), (2.6) and (2.7), we get:

4 = |f 0 (z)|2 4 (2.12)

This clearly shows that is harmonic in D if and only if is harmonic in D0

38
Remark 2.3.2. This clearly shows that harmonic functions are conformally mapped
into harmonic functions, the importance of the previous theorem is that when solving
a problem, among other things when we change variables we sometimes dont want to
change the physical significance of this problem, we may be solving examples in the xy-
plane and for convenience we map the problem into the uv-plane. It may happen that
certain physical properties are present in the xy-plane (potential or force) or maybe one
family of lines intersect another family of lines at right angles and we want a mapping
into the uv plane where images of these families will meet at right angles.

39
Chapter 3

Applications of Conformal Mapping

Conformal mappings can be effectively used for constructing solutions to the Laplace
equation on complicated planar domains that appear in a wide range of physical prob-
lems, including fluid mechanics, aerodynamics, thermomechanics, electrostatics, and
elasticity. In this chapter, we will develop the basic techniques and theorems of complex
analysis that impinge on the solution to boundary value problems associated with the
planar Laplace equations and will expose three applications of conformal mappings
[Saf-Sni, Rud, Hen, Ahl].

3.1 Application 1: Fluid flow


Conformal mapping techniques can be used in a clever way to analyze problems in fluid
flow mechanics in two-dimentional domains, when the flow is incompressible, irrotational
and steady. Two-dimentional fluid mechanics problems are relevant when the fluid is
thin in the third dimension, in which case the fluid velocity is often negligible in that
direction, or otherwise uniform along that direction.

3.1.1 Boundary value problems and fluid flow


A boundary value problem is a differential equation with a set of additional constraints,
called boundary conditions. A solution of such boundary value problem satisfies the
differential equation and the boundary conditions. Boundary value problems arise from
several fields in physics as in wave equations, potential forces and steady temperatures.
Among the earliest boundary valued problems studied is the Dirichlet problem of
finding harmonic functions. There are several types of boundary conditions. A boundary
condition which specifies the value of the function itself at the boundary is a Dirichlet
boundary condition, while a boundary condition which specifies the value of the normal
derivative of a function at the boundary and not the value of the function itself is a
Neumann boundary condition. When the boundary has the form of a curve or surface
that gives a value to the normal derivative and the variable itself then this condition is
a Cauchy boundary condition.

40
Consider a planar steady state fluid flow, with velocity vector field:

V(x, y) = (u(x, y), v(x, y))T

where (x, y) C where represents the domain occupied by the liquid, while the
vector V(x, y) represents the instantaneous velocity of the fluid at (x, y). Recall that
the flow is incompressible if and only if it has vanishing divergence, that is:
u v
OV = + = 0.
x y
Incompressibility means that the fluid volume does not change as it flows. Most liquids,
including water, are, for all practical purposes, incompressible. On the other hand, the
flow is irrotational if and only if it has vanishing curl, that is:
v u
OV = =0
x y
Irrotational flows have no vorticity, and hence no circulation. A flow that is both
incompressible and irrotational is known as an ideal fluid flow. In many physical regimes,
liquids (and, although less often, gases) behave as ideal fluids. The velocity vector field
V(x, y) induces an ideal fluid flow if and only if f (z) = u(x, y) iv(x, y) is analytic. The
corresponding complex function f is called the complex velocity of the fluid flow. Under
the flow induced by the velocity vector field V = (u(x, y), v(x, y))T , the fluid particles
follow the trajectories z(t) = x(t) + iy(t) obtained by integrating the system of ordinary
differential equations:
dx dy
= u(x, y) and = v(x, y).
dt dt
Now, suppose that the complex velocity f (z) admits a complex primitive

(z) = (x, y) + i(x, y)

Then:
d
= i = u iv
dz x y
Thus,O = V, and hence the real part (x, y) of the complex function (z) defines a
velocity potential for the fluid flow. For this reason, the anti-derivative (z) is known as a
complex potential function for the given fluid velocity field. Since the complex potential
is analytic, its real part, the potential function, is harmonic, and therefore satisfies the
Laplace equation 4 = 0. Conversely, any harmonic function can be viewed as the
potential function for some fluid flow. The real fluid velocity is its gradient: V = O,
and is incompressible and irrotational. The harmonic conjugate (x, y) to the velocity
potential also plays an important role, and, in fluid mechanics, is known as the stream
function. It also satisfies the Laplace equation 4 = 0, and the potential and stream
function are related by the Cauchy- Riemann equations:

=u= and =v=
x y y x

41
Example 3.1.1. Consider a complex potential function (z) = z = x + iy, the velocity
potential is (x, y) = x and the steam function is (x, y) = y. The complex derivative
of the potential is the complex velocity
d
f (z) = =1
dz
Let () = (, ) + i(, ) be an analytic function representing the complex po-
tential function for a steady state fluid flow in a planar domain D. Composing the
complex potential () with a one to one conformal map = g(z) leads to a trans-
formed complex potential (z) = (g(z)) = (x, y) + i(x, y) on the corresponding do-
main = g 1 (D), thus we can use conformal maps to construct fluid flows in compli-
cated domains from known flows in much simpler domains. We will restrict our study
to planar fluid flows around a closed and bounded subset D R2 = C, representing a
cross-section of a cylindrical object see, Figure 3.1.

Figure 3.1: Cross section of cylindrical object [Olv]

The (complex) velocity and potential are defined on the complementary domain
= C\D occupied by the fluid. The velocity potential (x, y) will satisfy the Laplace
equation 4 = 0 in the exterior domain . For a solid object, we should impose the
homogeneous Neumann boundary conditions

=0 (3.1)
n
on the boundary = D indicating that there is no fluid flux into the object. We
note that, since it preserves angles and hence the normal direction to the boundary,
a conformal map will automatically preserve the Neumann boundary conditions. We
shall assume our object is placed in a uniform horizontal flow, e.g., a wind tunnel, as
sketched in Figure 3.2.
Thus, far away, the object will not affect the flow, and so the velocity should ap-
proximate the uniform velocity field V = (1, 0)T , where, for simplicity, we choose our
physical units so that the fluid moves from left to right with an asymptotic speed equal
to 1. Equivalently, the velocity potential should satisfy:

(x, y) x so O (1, 0) when x2 + y 2 0.

An alternative physical interpretation is that we are located on an object that is


moving horizontally to the left at unit speed through a fluid that is initially at rest.

42
Figure 3.2: Flow pass a solid object [Olv]

Think of an airplane flying through the air at constant speed. If we adopt a moving
coordinate system by sitting inside the airplane, then the effect is as if the plane is
sitting still while the air is moving towards us at unit speed.

3.1.2 Horizontal plate


A simple example is the flat plate moving horizontally through the fluid. The plates
cross-section is a horizontal segment, and for convenience we may consider this segment
to be D = [1, 1] located on the real axis. If the plate is thin and smooth, it will affect
the horizontal flow and, in this case, the velocity potential is given by:

(x, y) = x for x + iy = C\[1, 1]

Note that O = (1, 0)T , and hence this flow satisfies the Neumann boundary conditions
3.1 on the horizontal segment D. The corresponding complex potential is (z) = z
with complex velocity 0 (z) = f (z) = 1

3.1.3 Circular disk


We have seen previously that the Joukowski conformal map defined by = 21 (z + z1 )
squashes the unit circle |z| = 1 into the real line segment [-1,1] in the -plane, and so
= g(z) will map the fluid flow outside the unit disk to the fluid flow past the line
segment. The complex potential for horizontal plate flow was () = , and thus the
complex potential for the fluid flow outside a unit disk will be:
1 1
(z) = g(z) = (z + )
2 z

3.1.4 Tilted plate


Let us next consider the case of a tilted plate in a uniformly horizontal fluid flow. The
cross-section will be the line segment z(t) = tei where 1 t 1 obtained by rotating
the horizontal line segment [-1,1] through an angle - . The goal is to construct a fluid
flow past the tilted segment that is asymptotically horizontal at large distance. The air
flow will be going from left to right, is called the attack angle of the plate or airfoil

43
relative to the flow. The key observation is that, while the effect of rotating a plate in
a fluid flow is not so evident, rotating a circularly symmetric disk has no effect on the
flow around it. Thus the rotation w = eiz maps the disk potential:
1 x y
(z) = z + = (x + 2 2
) + i(y 2 )
z x +y x + y2
to the complex potential:

ei
(w) = (ri w) = ei w + .
w
The streamlines of the induced flow are no longer asymptotically horizontal, but rather
at an angle , If we now apply the original Joukowski map to the rotated flow, the circle
is again squashed down to the horizontal line segment, but the stream lines continue to
be at angle at large distances. Thus, if we then rotate the resulting flow through an
angle -, the net effect will be to tilt the segment to the desired angle while rotating
the streamlines to be asymptotically horizontal giving the result:
p
(z) = ei (z cos i sin z 2 e2i )

3.1.5 Airfoils
We have seen earlier that Joukowski map to off-center disks will, in favorable con-
figurations, produce airfoil-shaped objects. The fluid motion around such airfoils can
thus be obtained from the flow past such an off-center circle. Note that the affine
transformation w = z + transforms the disk |z| 1 onto the disk |w | ||. In
particular, the boundary circle will continue to pass through the point w = 1 provided
that || = |1 |, now apply the Joukowski transformation to get:
1 1 
= z + +
2 z +
The resulting complex potential for the flow past the airfoil is obtained by substituting
the inverse map: p
w + 2 1
z= =

into the disk potential whereby:
p p
+ 2 1 ( 2 1)
() = +
2 + 1 2
Finally, if the airfoils position is with an attack angle with the horizontal line, we
replace by ei to get:
i
p i
p
e + (ei )2 1 (e (ei )2 1)
(ei ) = +
2 + 1 2ei

44
3.2 Application 2: A non-coaxial cable
The goal of this section is to determine the electrostatic potential u inside a non-coaxial
cylindrical cable with prescribed constant potential values on the two bounding cylin-
ders, see Figure 3.3, where u(x, y) is the potential function (decribed in the previous
section).

Figure 3.3: A non-coaxial cable

Suppose that the larger cylinder has a radius of 1 and is centered at the origin,
while the smaller cylinder has a radius of 52 and centered at z = 25 . The resulting
electrostatic potential u(x, y) will be independent of the longitudinal coordinate, and
so can be viewed as a planar potential in the annular domain contained between the
two circles representing the cross section of the cylinders = {|z| < 1 and |z 25 | > 25 }.
The desired potential must satisfy the Dirichlet boundary value problem:

4u = 0 on
u = b on |z| = 1 (3.2)
2 2
u = a on |z 5 | = 5

Figure 3.4: The domain

To solve this problem we start by studying the mapping = 2z1


z2
which is a frac-
tional linear transformation. In the domain (see Figure 3.4) contained between two

45
concentric circles, the larger circle with radius 1 and the smaller circle with radius 25 ,
with center z = 25 , the mapping takes those non-concentric circles into concentric an-
nulus of the form A = {r < || < 1}, to see this note that we proved earlier that linear
z
transformations with form = z1 with || < 1 map the unit disk into itself, moreover,
linear fractional transformations map circles into circles, so if we seek a value of to
make the inner circles |z 25 | = 25 a circle of the form || = r which is centered at the
origin, choosing to be real and trying to map the points 0 and 54 on the inner circle
to the points r and -r on the circle || = r, we get:
1. (0) = = r
4

2. ( 45 ) = 4
5
1
= r
5

Equating r in both equations gives:


2 2 2
+ =0
5 5
and so = 21 or = 2 and so as || < 1 then = 12 and r = 12 and hence = 2z1
z2
is our
2
conformal map transforming these non-concentric circles = {|z| < 1 and |z 5 | > 25 }
into concentric annulus A = { 21 < || < 1}. The corresponding transformed potential is
U (, ) having the following Dirichlet boundary conditions:

U = a on || = 21

(3.3)
U = b on || = 1
It is clear that coaxial potential energy must be a symmetric solution of Laplace equation
and thus having the form:
U (, ) = log || +
Replacing the initial conditions, we find that:
ba

= log 2 (3.4)
=b
And so:
ba ba
U (, ) =log || + b = log( 2 + 2 ) + b (3.5)
log 2 2 log 2
Equation 3.5 represents the conformal map of the potential of the concentric (see Figure
3.5), then by Theorem 2.3.2 we can find the expression of the potential of the non-coaxial
cable u = U to get:
2 2
ba 2z 1
+ b = b a log (2x 1) + y + b
u(x, y) = log
log 2 z2 2 log 2 (x 2)2 + y 2

Remark 3.2.1. The same harmonic function determines the equilibrium temperature
of an annular plate whose inner boundary is kept at temperature u = a while the outer
boundary is kept ar temperature u = b

46
Figure 3.5: Electrostatic potential between coaxial and non-coaxial cylinders [Olv]

3.3 Application 3: Steady temperatures


In the theory of heat conduction, an assumption is made that heat flows in the direction
of decreasing temperature. Another assumption is that the time rate at which heat flows
across the surface area is proportional to the component of the temperature gradient in
the direction perpendicular to the surface area.

3.3.1 Background
In the theory of heat conduction, the flux across a surface within a solid body at a point
on that surface is the quantity of heat flowing in a specified direction normal to the
surface per unit time per unit area at the point. Flux is, therefore, measured in such
units as calories per second per square centimeter. It is denoted here by , and it varies
with the normal derivative of the temperature T at the point on the surface:
dT
= K (K > 0). (3.6)
dN
Relation 3.6 is known as F ourier0 s law and the constant K is called the thermal
conductivity of the material of the solid, which is assumed to be homogeneous. The
points in the solid are assigned rectangular coordinates in three-dimensional space, and
we restrict our attention to those cases in which the temperature T varies with only
the x and y coordinates. Since T does not vary with the coordinate along the axis
perpendicular to the xy-plane, the flow of heat is, then, two-dimensional and parallel to
that plane. We agree, moreover, that the flow is in a steady state; that is, T does not
vary with time. It is assumed that no thermal energy is created or destroyed within the
solid. That is, no heat sources or sinks are present there. Also, the temperature function
T (x, y) and its partial derivatives of the first and second order are continuous at each
point interior to the solid. This statement and 3.6 for the flux of heat are postulates
in the mathematical theory of heat conduction, postulates that also apply at points
within a solid containing a continuous distribution of sources or sinks. The temperature

47
function satisfies Laplaces equation at each interior point of the solid so
Txx + Tyy = 0. (3.7)
And noting the continuity of T leads it to be a harmonic function of x and y in the
domain representing the interior of the solid body. The surfaces T (x, y) = c1 , where c1
is any real constant, are the isotherms within the solid. They can also be considered as
curves in the xy-plane; then T (x, y) can be interpreted as the temperature at a point
(x, y) in a thin sheet of material in that plane, with the faces of the sheet thermally
insulated. The isotherms are the level curves of the function T . The gradient of T is
perpendicular to the isotherm at each point, and the maximum flux at a point is in the
direction of the gradient there. If T (x, y) denotes temperatures in a thin sheet and if S
is a harmonic conjugate of the function T , then a curve S(x, y) = c2 has the gradient
of T as a tangent vector at each point where the analytic function T (x, y) + iS(x, y) is
dT
conformal. The curves S(x, y) = c2 are called lines of flow. If the normal derivative dN
is zero along any part of the boundary of the sheet, then the flux of heat across that
part is zero. That is, the part is thermally insulated and is, therefore, a line of flow.

3.3.2 Steady temperature in a half plane


The boundary value problem of the expression for the steady temperatures T (x, y) in a
thin semi-infinite plate y > 0 whose faces are insulated and whose edge y = 0 is kept at
temperature zero except for the segment 1 < x < 1, where it is kept at temperature
unity is:
Txx (x, y) + Tyy (x, y) = 0 ( < x < +) (3.8)
and T (x, 0) = 1 when |x| < 1 and T (x, 0) = 0 when |x| > 1 besides |T (x, y)| < M where
M is a positive constant. Clearly this is a Dirichlet problem for the upper half of the
xy-plane. Our method of solution will be to obtain a new Dirichlet problem for a region
in the uv-plane, thus we need a conformal mapping w = f (z) that is analytic in the
domain y > 0 except at the point z = 1, where it is undefined. This conformal map is
z1
w = f (z) = Log z+1 as we have seen in Example 2.1.3. A bounded harmonic function of
u and v that is zero on the edge v = 0 of the strip and unity on the edge v = is nothing
but T = 1 v. This function is harmonic as it is the imaginary part of the function ( 1 w).
As we know the solution in the uv-plane we can get back the solutions in the xy-plane
by changing to x and y coordinates (see Figure 3.6) by means of the equation:
|z 1| z1
w = ln + iArg( )
|z + 1| z+1
which clearly gives that:
h x2 + y 2 1 + i2y i 2y  
v = Arg = arctan .
(x + 1)2 + y 2 x2 + y 2 1
The range of the arctangent here is from 0 to as Arg z1

z+1
= 1 2 [0, ]. and
so:
1  2y 
T = arctan 2
x + y2 1

48
 
r1
Figure 3.6: w = Log z1
z+1 r2
> 0, 2 < 1 2 < 3
2
[Olv]

The isotherms T (x, y) = c1 (0 < c1 < 1) are arcs of the circles with equation:

x2 + (y cot(c1 ))2 = csc2 (c1 )

note that this result follows directly from re-writing T (x, y) = c1 in its canonical form.
These circles pass through (1, 0) and with centers on the y-axis.

Remark 3.3.1. If the assumption of unit temperature condition is dropped to become


another constant T0 along 1 < x < 1, then as the product of a harmonic function by a
scalar is also harmonic we can conclude that the steady temperatures in the given half
region is given by:
T0  2y 
T = arctan 2
x + y2 1

3.3.3 Steady temperature in a semi-infinite slab


Consider a semi-infinite slab in the three-dimensional space bounded by the planes
x = 2
and y = 0 when the first two surfaces are kept at temperature zero and the
third at temperature unity (see Figure 3.7). Lets find a formula for the temperature
T (x, y) at any interior point of the slab. The boundary value problem here is:

1.  
Txx + Tyy = 0, < x < ,y > 0
2 2
2.
 
T , y = T , y = 0, y > 0
2 2
3.  
T (x, 0) = 1, <x<
2 2
where T (x, y) is bounded. The map needed in this case is w = u + iv = sin z which
transforms this boundary value problem into the same one as the previous application
and thus:
1  2v 
T = arctan 2 (3.9)
u + v2 1

49
The change of variable in z = u + iv = sin z = sin(x + iy) leads to u = sin x cosh y and
v = cos x sinh y and thus the steady temperatures are given by the equation:
1  2 cos x sinh y 
T = arctan .
sin2 x cosh2 y + cos2 x sinh2 y 1
Simplifying the desired function to get T (x, y) in its simplest form as:
2  cos x 
T = arctan
sinh y
Remark 3.3.2. If the temperature on the third surface is T0 instead of 1 then the steady
temperature is given by the function:
2T0  cos x 
T = arctan
sinh y

Figure 3.7: Semi-infinite slab [Bro-Chu, page 366]

3.3.4 Steady temperature in a quadrant


Let us find the steady temperatures in a thin plate having the form of a quadrant if a
segment at the end of one edge is insulated, if the rest of that edge is kept at a fixed
temperature, and if the second edge is kept at another fixed temperature. The surfaces
are insulated, and so the problem is two-dimensional (see Figure 3.8). The temperature
scale and the unit of length can be chosen so that the boundary value problem for the
temperature function T becomes:


Txx (x, y) + Tyy (x, y) = 0
Ty (x, 0) = 0 when 0 < x < 1

(3.10)

T (x, 0) = 1 when x > 1
T (0, y) = 0 when y > 0,

where T (x, y) is bounded in the quadrant. The transformation w = u + iv = sin z is


a one-to-one mapping of the semi-infinite strip 0 u 2 and v 0. Observe now that

50
the existence of an inverse is ensured by the fact that the given transformation is both
one to one and onto. Since transformation w = sin z is conformal throughout the strip
except at the point w = 2 , the inverse transformation must be conformal throughout
the quadrant except at the point z = 1. This inverse transformation maps the segment
0 < x < 1 of the x-axis onto the base of the strip and the rest of the boundary onto
the sides of the strip. The required temperature function T for the new boundary value
problem is clearly: T = 2 u. Now to write T in terms of x and y, we use u = sin x cosh y
and v = cos x sinh y to get when 0 < u < 2 :

x2 y2
=1 (3.11)
sin2 u cos2 u
So for a fixed u the hyperbola 3.11 has foci the points z = 1, and the length of the
major axis, which is the line segment joining the two vertices, is 2 sin u. The required
temperature function is:

2 h p(x + 1)2 + y 2 p(x 1)2 + y 2 i


T = arcsin
2

Figure 3.8: Temperature in a quadrant[Bro-Chu, page 368]

51
Bibliography

[Ahl] L. Ahlfors, Complex analysis, McGraw-Hill, New York, 1966.

[Bak-New] J. Bak and D. J. Newman, Complex analysis, third edition, Springer, 2010.

[Blair] D. Blair, Inversion theory and conformal mapping, American Mathematical So-
ciety, 2000.

[Bro-Chu] J. W. Brown and R. V. Churchill, Complex variables and applications, Edition


8, McGraw-Hill International Edition, 2001.

[Dit-Geo] P. Dita and V. Georgescu, Conformal invariance and string theory, Academic
Press, Inc., Boston, MA (USA), 1989, 570, ISBN 0-12-218100-X.

[Gam] Th. W. Gamelin, Complex analysis, UTM, Springer-Verlag (2001).

[Gre-Kra] R. Greene and S. Krantz, Function theory of one complex variable, third edi-
tion, Graduate studies in Mathematics, volume 40, American Mathematical Society,
2006.

[Hal] N. D. Halsey, Potential flow analysis of multielement airfoils using conformal


mapping, AIAA Journal 17.12 ,1979.

[Hen] P. Henrici, Applied and computational complex analysis, vol. 1, J. Wiley and Sons,
New York, 1974.

[Kar] V. Karunakaran, Complex analysis, second edition, Alpha Science International


Ltd. Harrow, UK, 2005.

[Neh] Z. Nehari, Conformal mapping, Dover publications, INC. New York, 1952.

[Olv] P. Olver, Complex analysis and conformal mapping, University of Minnesota, 2013.

[Pon-Sil] S. Ponnusamy and H. Silverman, Complex variables with applications,


Birkhauser Basel, 2006.

[Rud] W. Rudin, Real and complex analysis, third edition, McGraw-Hill, New York,
1987.

[Saf-Sni] E. B. Saff and A. D. Snider, Fundamentals of complex analysis, third edition,


Prentice-Hall, Inc., Upper Saddle River, N.J., 2003.

52
[Sch-Lau] R. Schinzinger and P. AA Laura, Conformal mapping: methods and applica-
tions, Courier Corporation, 2003.

[Sie-Sny] R. Siegel and A. Snyder, Heat transfer in cooled porous region with curved
boundary, Journal of Heat Transfer 103 (4), 765-771, 1981.

[Thd] Th. Theodorsen, Theory of wing sections of arbitrary shape, National Advisory
Committee for Aeronautics. Langley Aeronautical Lab. Langley Field, VA, United
States, 1993.

[Wec-Zim] H. Wechsler and G. Lee Zimmerman, 2-D invariant object recognition using
distributed associative memory, IEEE Transactions on Pattern Analysis and Machine
Intelligence, Vol.10, Issue 6, 1988.

53

You might also like