Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 452

PERIYAR INSTITUTE OF DISTANCE EDUCATION

(PRIDE)

PERIYAR UNIVERSITY
SALEM - 636 011

M.Sc. MATHEMATICS
SECOND YEAR
PAPER - VI : COMPLEX ANALYSIS

1
Prepared by :
M. RAJASEKARAN
Department of Mathematics
Government Arts College
Salem - 636 007.

2
UNIT - I
INTRODUCTION TO COMPLEX PLANE

It is fundamental that real and complex numbers obey the same basic
laws of arithmetic. We begin our study of complex function theory by
stressing and implementing this analogy.
Arithmetic Operations.
From elementary algebra the reader is acquainted with the imaginary
unit i with the property i  1 . If the imaginary unit is combined with two
2

real numbers  ,  by the processes of addition and multiplication, we obtain a


complex number   i  .  and  are the real and imaginary part of the
complex number. If   0 , the number is said to be purely imaginary; if   0
, it is course real. Zero is the only number which is at once real and purely
imaginary. Two complex numbers are equal if and only if they have the same
real part of the same imaginary part.
Addition and multiplication do not lead out from the system of complex
numbers. Assuming that the ordinary rules of arithmetic apply to complex
numbers we find indeed

(1)
   i       i         i     
and

(2)
   i      i         i     
In the second identity we have made use of the relation i  1 . It is
2

less obvious that division is also possible. We wish to show that


   i  
   i  is a complex number, provided that    i   0 . If the
quotient is denoted by x  iy , we have
   i      i   x  iy 
By (2) this condition can be written
   i    x  y   i  x  y 
And we obtain the two equations
  x  y
  x  y
This system of simultaneous linear equations has the unique solution.

3
  
x
 2  2
  
y 2 2
 

For we know that    is not zero. We have thus the result.


2 2

  i      
 2 i 2
(3)   i      2
2

Once the existence of the quotient has been proved, its value can be
found in a simplex way. If numerator and denominator are multiplied with
  i , we find at once

  i    i     i        i     
 
  i    i     i   2  2

As a special case the reciprocal of a complex number  0 is given by


1   i
 2
  i    2
n
We note that i has only four possible values: 1,i, -1, -i. they
correspond to values of n which divided by 4 leave the remainders 0, 1, 2 , 3,.
Square Roots.
We shall now show that the square root of a complex number can be
found explicity. If the given number is   i , we are looking for a number x
+ iy such that
 x  iy 
2
   i
This is equivalent to the system of equations.
(4)
x 2  y2  
2xy  
From these equations we obtain

x  y 2    x 2  y 2   4x 2 y 2   2  2
2 2 2

Hence we must have


x 2  y 2   2  2
where the square root is positive or zero. Together with the first equation (4)
we find
(5)

4
x2 
1
2

   2  2 
1

y 2     2   2
2

Observe that these quantities are positive or zero regardless of the sign
of  .
The equation (5) yield, in general, two opposite values for x and two for
y. but these values cannot be combined arbitrarily, for the second equation (4)
is not a consequences of (5). We must therefore be careful to select x and y so
that their product has sign of  . This leads to the general solution.
(6)
   2  2    2   2 

  i    i
 2  2 
 

Provided that   0. For   0 the values are   if   0, i  if


  0 . It is understood that all square roots of positive numbers are taken with
the positive sign.
We have found that the square root of any complex number exists and
has two opposite values. They coincide only if   i  0 . They are real if
  0,   0 and purely imaginary if   0,   0 . In other words, except for
zero, only positive numbers have real square roots and only negative numbers
have purely imaginary square roots.
Since both square roots are in general complex, it is not possible to
distinguish between the positive and negative square root of a complex number.
We could of course distinguish between the upper and lower sign in (6), but
this distinction is artificial and should be avoided. The correct way is to treat
both square roots in a symmetric manner.
Justification.
So far our approach to complex numbers has been completely
uncritical. We have not questioned the existence of a number system in which
the equation x  1  0 has a solution while all the rules of arithmetic remain in
2

force.
We begin by recalling the characteristic properties of the real-number
system which we denoted by R. in the first place, R is a field. This means that
addition and multiplication are defined, satisfying the associative,
commutative, and distributive laws. The number 0 and 1 are neutral elements
under addition and multiplication, respectively:   0  , .1   for all .

5
Moreover, the equation of subtraction   x   has always a solution, and the
equation of division x   has a solution whenever   0.
One shows by elementary reasoning that the neutral elements and the
results of subtraction and division are unique. Also, every field is an integral
domain:   0 if and only if   0 or   0 .
These properties are common to all fields. In addition, the field R has
    or    
an order relation . It is most easily defined in terms of the set
R of positive real numbers:    if and only if     R . The set R  is
 

characterized by the following properties: (1) 0 is not a positive number; (2) if


  0 either  or   is positive; (3) the sum and the derives all the usual
rules for manipulation of inequalities. In particular one finds that every square
 2 is either positive or zero; therefore 1  12 is a positive number.

By virtue of the order relation the sums 1,1  1,1  1  1,.... are all
different. Hence R contains the natural numbers, and since it is a field a it must
contain the subfield formed by all rational numbers.
Finally, R satisfies the following completeness conditions: every
increasing and bounded sequence of real numbers has a limit . let
1   2   3  .....   n  ...., and assume the existence of a real number B such

that  n  B for all n. then the completeness condition requires the existence of
a number A = lim n   n with the following property; given any  0 there
exists a natural number n 0 such that A    n  A for all n  n 0 .
Our discussion of the real-number system is incomplete inasmuch as we
have not proved the existence and uniqueness (up to isomorphisms) of a system
R with the postulated properties. The student who is not thoroughly familiar
with one of the constructive processes by which real numbers can be
introduced should not fail to fill this gap by consulting any textbook in which a
full axiomatic treatment of real numbers is given.
The equation x  1  0 has no solution in R, for   1 is always
2 2

positive. Suppose now that a field F can be found which contains R as a


subfield, and in which the equation x  1  0 can be solved. Denote a solution
2

x2 1   x  i  x  i
, and the equation x  1  0 has exactly two
2
by i. Then
roots in F, i and –i. Let C be the subset of F consisting of all elements which
can be expressed in the form   i with real  and  . This representation is
  i   '  i' implies    '  i     ' 
2

unique, for , and this is possible


only if    ,    .
' '

6
The subset C is a subset of F. In fact, except for trivial verifications
which the reader is asked to carry out, this is exactly what was shown in Sec.
'
1.1. what is more, the structure of C is independent of F. for if F is another
field containing R and a root i of the equation x  1  0 ,the corresponding
' 2

subset C is formed by all elements   i  . There is a one-to-one


' '

correspondence between C and C which associates   i and   i  , and


' '

this correspondence is evidently a field isomorphism. It is thuds demonstrated


'
that C and C are isomorphic.
We now define the field of complex numbers to be subfield C of an
arbitrarily given F. we have just seen that the choice of F makes no difference,
but we have not yet shown that there exists a field F with the required
properties. In order to given our definition a meaning it remains to exhibit a
field F which contains R (or a subfield isomorphic with R) and in which the
equation x  1  0 has a root.
2

There are many ways in which such a field can be constructed. The
simplest and most direct method is the following: Consider all expressions of
the form   i where ,  are real numbers while the signs + and i are the
pure symbols (+ does not indicate addition, and i is not an element of a field).
These expressions are elements of a field F in which addition and
multiplication are defined by (1) and (2) (observe the two different meanings of
the sign +). The elements of the particular form   i0 are seen to constitute a
subfield isomorphic to R, and the element 0  i1 satisfies the equation
x 2  1  0 ; we obtain in fact  0  i1    1  i0  . The field F has thus the
2

required properties: moreover, it is identical with the corresponding subfield C,


for we can write
  i     i0     0  i1
The existence of the complex number field is now proved, and we can
go back to the simpler notation   i where the + indicate addition in C and i
is a root of the equation x  1  0 .
2

Conjugation, Absolute Value.


A complex number can be denoted either by a single letter a,
representing an element of the field C, or in the form   i with real  and 
. Other standard notations are z  x  iy,     i, w  u  iv, and when used
in this connection it is tacitly understood that x, y, , , u, v are real numbers.
The real and imaginary part of a complex nuber a will also be denoted by Re a,
Im a.

7
In deriving the rules for complex addition and multiplication we used
only the fact that i  1 . Since –I has the same property, all rules shows that
2

this is indeed so. The transformation which replaces   i by   i is called


complex conjugation, and   i is the conjugate of   i . The conjugate of
a is denoted by a . A number is real if and only if it is equal to its conjugate.
The conjugation is an involutory transformation: this means that a  a .

The formulas
aa a a
Re a  , Im a 
2 2i
Express the real and imaginary part in terms of the complex number and
its conjugate. By systematic use of the notations a and a it is hence possible to
dispense with the use of separate letters for the real and imaginary part. It is
more convenient, though, to make free use of both notations.
The fundamental property of conjugation is the one already referred to,
namely, that
ab ab
ab  a.b
The corresponding property for quotients is a consequence : if ax = b,

then
 
ax  b, and hence b  b
a a . More generally, let R  a, b, c,.... stand for
any rational operation applied to the complex numbers a, b, c,…… then


R  a, b, c,....  R a, b, c,...... 
As an application, consider the equation
c 0 z n  c1z n 1  .....  c n 1z  c n  0

If  is a root of this equation, then  is a root of the equation.


c 0 z n  c1z n 1  .....  c n 1z  c n  0

In particular, if the coefficients are real,  and  are roots of the same
equation, and we have the familiar theorem that the nonreal roots of an
equation with real coefficients occur in pairs of conjugate roots.

The product aa     is always positive or zero. Its nonnegative


2 2

square root is called the modulus or absolute value of the complex number a; it

8
a
is denoted by . The terminology and notation are justified by the fact that
the modulus of a real number coincides with its numerical value taken with the
positive sign.
We repeat the definition
2
aa  a

a 0 a a
where , and observe that . For the absolute value of a product we
obtain
2 2 2
ab  ab.ab  abab  aabb  a b
And hence
ab  a b

Since both are  0. in words:


The absolute value of a product is equal to the product of the absolute
values of the factors.
It is clear that this property extends to arbitrary finite products:
a1a 2 ......a n  a1 a 2 ........ a n

The quotient
a , b0
b satisfies
b a a
b , and hence we have also
 
b a a
b or
a a

b b
The formula for the absolute value of a sum is not a simple. We find


a  b   a  b  a  b  aa  ab  ba  bb
2
  
or
2 2 2
(7) a  b  a  b  2 Re ab
The Corresponding formula for the difference is
2 2 2
(7' ) a  b  a  b  2 Re ab
and by addition we obtain the identity
(8)
2 2
a b  a b  2 a  b  2 2

Inequalities.

9
We shall now prove some important inequalities which will be of
constant use. It is perhaps well to point out that there is no order relation in the
complex-number system, and hence all inequalities must be between real
numbers.
 a  Re a  a
 a  Im a  a
(9)
Re a  a
The equality holds if and only if a is real and  0 .
It (9) is applied to (7), we obtain
ab  a  b
2 2

And hence
ab  a  b
(10)
This is called the triangle inequality for reasons which will emerge
later. By induction it can be extended to arbitrary sums:
a1  a 2  .....  a n  a1  a 2  ......  a n
(11)
The absolute value of a sum is at most equal to the sum of the absolute
values of the terms.
The reader is well aware of the importance of the estimate (11) in the
real case, and we shall find it no less important in the theory of complex
numbers.
Let us determine all cases of equality in (11). In (10) the equality holds
if and only if ab  0 (it is convenient to let c>0 indicate that c is real and

positive). If b  0 this condition can be written in the form


b a 0
2
b  
, and it
a 0
is hence equivalent to b . In the general case we proceed as follows:
Suppose that equality holds in (11) ; then
a1  a 2  ..........  a n   a1  a 2   a 3  ......  a n
 a1  a 2  a 3  .......  a n
 a1  a 2  ..........  a n
a1
a 1  a 2  a1  a 2 0
Hence a
, and if 2  0 we conclude that a 2 . But
the numbering of the terms is arbitrary; thus the ratio of any two nonzero terms
must be positive. Suppose conversely that this condition is fulfilled. Assuming
that a1  0 we obtain

10
a2 a
a1  a 2  ..........  a n  a 1 . 1   ....  n
a1 a1
 a a 
 a1 1  2  ....  n 
 a1 a1 
 a a 
 a1 1  2  .....  n 
 a1 a1 
 a1  a 2  ..........  a n
To sum up: the sign of equality holds in (11) if and only if the ratio of
any two nonzero terms is positive.
By (10) we have also
a   a  b  b
 a b  b
or
a  b  ab

ba  a b
For the same reason , and these inequalities can be
combined to
ab  a  b
(12)
ab
Of course the same estimate can be applied to .
A special case of (10) is the inequality
  i    
(13)
which expresses that the absolute value of a complex number is at most equal
to the sum of the absolute values of the real and imaginary part.
Many other inequalities whose proof is less immediate are also of
frequent use. Foremost is Cauchy’s inequality which states that
2
 2
a1b1  .....  a n b n  a1  ....  a n
2
 b
1
2
 ....  b n
2

Or, in shorter notation,
n 2 n n

 a i b i  a i b
2 2
i
(14) i 1 i 1 i 1

To prove it, let  denote an arbitrary complex number. We obtain by


(7)

11
n 2 n n n

a   bi   a i   b  2 Re   a i b i
2 2 2
i i
(15) i 1 i 1 i 1 i 1

This expression is  0 for all  . We can choose


n

a b i i
 1
n

b
2
i
1

For if the denominator should vanish there is nothing to prove. This


choice is not arbitrary, but it is dictated by the desire to make the expression
(15) as small as possible. Substituting (15) we find, after simplifications,
n 2

n a b i i

 ai 
2 1
n
0
b
1 2
i
1

which is equivalent to (14) .


From (15) we conclude further that the sign of equality holds in (14) if
and only if the ai are proportional to the bi .
Cauchy’s inequality can also be proved by means of Lagrange’s
identity.
THE GEOMETRIC REPRESENTATION OF COMPLEX NUMBERS
With respect to a given rectangular coordinate system in a plane, the
complex number a    i can be represented by the point with coordinates
 ,   . This representation is constantly used, and we shall often speak of the
point a as a synonym of a number a. The first coordinate axis (x-axis) takes
the name of real axis, and the second coordinate axis (y-axis) is called the
imaginary axis. The plane itself is referred to as the complex plane.
The geometric representation derives its usefulness from the vivid
mental pictures associated with a geometric language. We take the point of
view, however, that all conclusions in analysis should be derived from the
properties of real number, and not from the axioms of geometry. For this
reason we shall use geometry only for descriptive purposes, and not for valid
proof, unless the language is so thinkly veiled that the analytic interpretation is
self-evident. This attitude relieves us from the exigencies of rigor in
connection with geometric considerations.
Geometric Addition and Multiplication.
The addition of complex numbers can be visualized as vector addition.
To this end we let a complex number be represented not only by a point, but

12
also by a vector pointing from the origin to the point. The number, the point,
and the vector will all be denoted by the same letter a. As usual we identify all
vectors which can be obtained from each other by parallel coincides with the
end.
Place a second vector b so that its initial point coincides with the end
point of a. Then a + b is represented by the vector from the initial point of a to
the end point of b. To construct the difference b – a we draw both vectors a
and b from the same initial point; then b – a we draw both vectors a and b from
the same initial lpoint; then b – a points from the end point of a to the end
points of b. Observe that a + b and a – b are the diagonals in a parallelogram
with the sides a and b (Fig 1-1)
An additional advantage of the vector representation is that the length of
a
the vector a is equal to . Hence the distance between the points a and b is
a b ab  a  b
. With this interpretation the triangle inequality and the

identity
2 2

ab  a b  2 a  b
2 2
 become familiar geometric theorems.

13
b
a
a+b
a-b

b a

Fig. Vector addition

The point a and its conjugate a lie symmetrically with respect to the
real axis. The symmetric point of a with respect to the imaginary axis is - a .
The four points a, - a , -a, a are the vertices of a rectangle which is symmetric
with respect to both axes.
In order to derive a geometric interpretation of the product of two
complex numbers we introduce polar coordinates. If the polar coordinates of
the point
 ,   are  r,   , we know that
  r cos 
  r sin 
a    i  r  cos   i sin  
Hence we can write . In this trigonometric
a
form of a complex number r is always 0 and equal to the modulus . The
polar angle  is called the argument or amplitude of the complex number, and
we denote it by arg a.
a1  r1  cos 1  i sin 1 
Consider two complex numbers and
a 2  r2  cos 2  isin 2 
. Their product can be written in the from
a1a 2  r1r2  cos 1 cos 2  sin 1 sin 2   i  sin 1 cos 2  cos 1 sin  2  
. By
means of the addition theorems of the cosine and the sine this expression can
be simplified to

(16) a1a 2  r1r2 cos  1  2   i sin  1  2  

We recognize that the product has the modulus r1r2 and the argument
1  2 . The latter result is new, and we express it through the equation

14
arg  a1a 2   arg a1  arg a 2
(17)
It is clear that this formula can be extended to arbitrary products, and
we can therefore state:
The argument of a product is equal to the sum of the arguments of the
factors.
This is fundamental. The rule that we have just formulated gives a deep
and unexpected justification of the geometric representation of complex
numbers. We must be fully aware, however, that the manner in which we have
arrived at the formula (17) violates our principles. In the first place the
equation (17) is between angles rather than between numbers, and secondly its
proof rested on the use of trigonometry. Thus it remains to define the argument
in analytic terms and to prove (17) by purely analytic means. For the moment
we postpone this proof and shall be content to discuss the consequences of (17)
from a less critical standpoint.
We remark first that the argument of 0 is not defined, and hence (17)
has a meaning only if a1 and a 2 are  0 . Secondly, the polar angle is
determined only up to multiplies of 360 . For this season, if we must to

interpret (17) numerically, we must agree that multiplies of 360 shall not

count.

By means of (17) a simple geometric construction of the product a1a 2


can be obtained. It follows indeed that the triangle with the vertices 0, 1, a1 , is
similar to the triangle whose vertices are 0, a 2 , a1a 2 . The points 0,1,a 1 , and a 2
being given, this similarity determines the point a1a 2 (Fig. 1-2). In this case of
division (17) is replaced by
a2
arg  arg a 2  arg a1
(18) a1

The geometric construction is the same, except that the similar triangles
are now 0,1, a1 , and 0, a 2 a1 , a 2 .
Remark:
A perfectly acceptable way to define angles and arguments would be to
apply the familiar methods of calculus which permit us to express the length of
a circular arc as a definite integral. This leads to a correct definition of the
trigonometric functions, and to a computational proof of the addition theorems.
a1a2

a1

15
O 1
Fig. Vector multiplication
The reason we do not follow this path is that complex analysis, as
opposed to real analysis, offers a much more direct approach. The clue lies in a
direct connection between the exponential function and the trigonometric
functions. Until we reach this point the reader is asked to subdue his quest for
complete rigor.
The Binomial Equation.
From the preceding results we derive that the powers of
a  r  cos   i sin  
are given by
a n  r n  cos n  i sin n 
(19)
This formula is trivially valid for n =0, and since
a 1  r 1  cos   i sin    r 1 cos     i sin    
It holds also when n is a negative integer.
For r = 1 we obtain de Moivre’s formula
 cos   i sin  
n
 cos n  i sin n
(20)
which provides an extremely simple way to express cos n and sin n in terms
of cos  and sin  .
To find the nth root of a complex number a we have to solve the
equation
(21) zn  a
a  r  cos   i sin  
Supposing that a  0 we write and
z    cos   i sin  
Then (21) takes the form
n  cos n  i sin n   r  cos   i sin  
(22)

This equation is certainly fulfilled if   r and n   .


n
Hence we
obtain the root
  
z  n r  cos  i sin 
 n n

16
n
where r denotes the positive nth root of the positive number r.
But this is not the only solution. In fact, (22) is also fulfilled if n
differs from  by a multiple of the full angle. If angles are expressed in radians
the full angle is 2 , and we find that (22) is satisfied if and only if
 2
  k.
n n
where k is any integer. However, only the values k  0,1,......., n  1 give
different values of z. Hence the complete solution of the equation (21) is given
by
  2   2  
z  n r cos   k   i sin   k  , k  0,1,......, n  1
 n n  n n  
These are nth roots of any complete number  0 . They have the same
modulus, and their arguments are equally spaced.
Geometrically, the nth roots are the vertices of a regular polygon with n
sides.
The case a = 1 is particularly important. The roots of the equation
z  1 are called nth roots of unity, and if we get
n

2 2
  cos  i sin
(23) n n
n 1
All the roots can be expressed by 1, ,  ,....,  . It is also quite
2

n
evident that if a denote any nth root of a, then all the nth roots can be
expressed in the form  . a , k  0,1,....., n  1
k n

Analytic Geometry.
In classical analytic geometry the equation of a locus is expressed as a
relation between x and y. it can just as well be expressed in terms of z and z ,
sometimes to distinct advantage. The thing to remember is that a complex
equation is ordinarily equivalent to two real equations; in order to obtain a
genuine locus these equations should be essentially the same.
za  r
For instance, the equation of a circle is . In algebraic form it

can be rewritten as

 z  a z  a  r 
2

. The fact that this equation is invariant


under complex conjugation is an indication that it represents a single real
equation.
A straight line in the complex plane can be given by a parametric
equation z  a  bt, where a and b are complex numbers and b  0 ; the

17
parameter t runs through all real values. Two equations z  a  bt, and
z  a '  b' t represent the same line if and only if a  a and b are real multiples
' '

'
of b. The lines are parallel whenever b is a real multiple of b, and they are
'
equally directed if b is a positive multiple of b. The direction of a directed
line can be identified with arg b. The angle between z  a  bt and z  a  b t
' '

'
arg b
is b ; observe that it depends on the order in which the lines are named.
b'
The lines are orthogonal to each if b is purely imaginary.
Problems of finding intersections between lines and circles, parallel or
orthogonal lines, tangents, and the like usually become exceedingly simple
when expressed in complex form:
za  r
An inequality describes the inside of a circle. Similarly, a
directed z  a  bt determines a right half plane consisting of all points z with
Im
 z  a  0 Im
 z  a  0
b and a left half plane b . An easy argument
shows that this distinction is independent of the parametric representation.
The Spherical Representation.
For many purpose it is useful to extend the system C of complex
numbers by introduction of a symbol  to represent infinity. Its connection
with the finite numbers is established by setting a      a   for all finite
a, and b.  .b  
For all b  0 , including b =  . It is impossible, however, to define
   and 0. without violating the laws of arithmetic. By special convention
a   for a  0 and b  0 for b  
we shall nevertheless write 0  .
In the plane there is no room for a point corresponding to  , but we can of
course introduce an “ideal” point which we call the point at infinity. The points
in the plane together with the point at infinity form the extended complex
plane. We agree that every straight line shall pass through the point at infinity.
By contrast, no half plane shall contain the ideal point.
It is desirable to introduce a geometric model in which all points of the
extended plane have a concrete representative. To this end we consider the unit
place S whose equation in three-dimensional space is x1  x 2  x 3  1 . With
2 2 2

every point on S, except


 0, 0,1 we can associate a complex number
x1  ix 2
z
(24) 1  x3

18
And this correspondence is one to one. Indeed, from (24) we obtain
2 x12  x 22 1  x3
z  
 1  x3  1  x3
2

and hence
2
z 1
(25) x3  2
z 1
Further computation yields
(26)
zz
x1  2
1 z
zz
x2 

i 1 z
2

The correspondence can be completed by letting the point at infinity
correspond to (0,0,1) and we can thus regard the sphere as a representation of
the extended plane or of the extended number system. We note that the
z 1
hemisphere x 3  0 corresponds to the disk and the hemisphere x 3  0 to
z 1
its outside . In function theory the sphere S is referred to as the Riemann
sphere.

If the complex plane is identified with the


 x1 , x 2  -plane with the
x1  and x 2  axis corresponding to the real and imaginary axis, respectively,
the transformation (24) takes on a simple geometric meaning. Writing
z  x  iy we can verify that

(27) x : y : 1  x1 : x 2 : x 3  1

And this means that the points


 x, y, 0   x1 , x 2 , x 3  and  0, 0,1 are in a
straight line. Hence the correspondence is a central projection from the center
(0,0,1) as shown in Fig. 1-3. It is called a stereographic projection. The
context will make it clear whether the stereographic projection is regarded as a
mapping from S to the extended complex plane, or vice versa.
In the spherical representation there is no simple interpretation of
addition and multiplication. Its advantage lies in the fact that the point at
infinity is no longer distinguished.
It is geometrically evident that the stereographic projection transforms
every straight line in the z-plane into a circle of S which passes through the
pole (0,0,1) and the converse is also true. More generally, any circle on the

19
sphere corresponds to a circle or straight line in the z-plane. To prove this we
observe that a circle on the sphere lies in a plane 1x1   2 x 2   3 x 3   0 ,
where we can assume that 1   2   3  1 and 0 0  1 .
2 2 2
In terms z and z
this equation takes the form

  
1 z  z   2i z  z   3 z  1   0 z  1   2
  2

or
 0  3   x 2  y 2   21x  2 2 y   0   3  0
For  0   3 this is the equation of a circle, and for  0   3 it
represents a straight line. Conversely, the equation of any circle or straight line
can be written in this form. The correspondence is consequently one to one.

Z
z
O
z
Z

Fig. Stereographic projection

It is easy to calculate the distance


d z, z ' 
between the stereographic


projection of z and z . if the points on the sphere are denoted by
 
 x1 , x 2 , x 3  , x1' , x '2 , x 3' we have first
From (35) and (36) we obtain after a short computation
x1x1'  x 2 x '2  x 3 x 3'


 z  z  z  z    z  z  z  z    z
' ' ' ' 2

1 z' 1
2

 z  1  z
2 ' 2
1 

 z 1  z 1  2 z  z
2 ' 2 '

 z  1  z  1
2 ' 2

As a result we find that

20
(28)
2 z  z'
d  z, z  
'

 z  1  z
2 ' 2

1

for z '   the corresponding formula is


2
d  z,   

2
z 1 
ELEMENTARY THEORY OF POWER SERIES
Polynomials and rational functions are very special analytic functions.
The easiest way to achieve greater variety is to form limits. For instance, the
sum of a convergent series is such a limit. If the terms are functions of a
variable, so is the sum, and if the terms are analytic functions, chances are good
that the sum will also be analytic.
Of all series with analytic terms the power series with complex
coefficients are the simplest. In this section we study only the most elementary
properties of power series. A strong motivation for taking up this study when
we are not yet equipped to prove the most general properties (those that depend
on integration ) is that we need power series to construct the exponential
function (sec. 3).
Sequences
a 

The sequences n 1 has the limit A if to every  0 there exists an n 0


a  A  for n n 0
such that n . A sequence with a finite limit is said to be
convergent, and any sequence which does not converge is divergent. If
lim n  a n   , the sequence may be said to diverge to infinity.

Only in rare can the convergence be proved by exhibiting the limit, so it is


extremely important to make use of a method that permits proof of the
existence of a limit even when it cannot be determine explicitly. The test that
serves this purpose bears the name of Cauchy. A sequence will be called
fundamental, or a Cauchy sequence4, if it satisfies the following condition:
given any  0 there exists an n0 such that
a n  a m  whenever n n 0 and mn 0
. The test reads:
A sequence is convergent if and only if it is a Cauchy sequence.

The necessity is immediate. If a n  A we can find n 0 such that


a n  A   for n  n 0
2 . For m, n  n 0 it follows by the triangle inequality
a  a m  a n  A  a m  A 
that n

21
The sufficiency is closely connected with the definition of real numbers,
and one way in which real numbers can be introduced is indeed to postulate the
sufficiency of Cauchy’s condition. However, we wish to use only the property
that every bounded monotone sequence of real numbers has a limit.
The real and imaginary part of a Cauchy sequence are again Cauchy
sequences, and if they converge, so does the original sequence. For this reason
we need to prove the sufficiency only for real sequences. We use the
opportunity to recall the notions of limes superior and limes inferior. Given a
  n  1 we shall set a n  max  1 ,..., n  , that non decreasing;

real sequence
hence it has a limit A1 which is finite or equal to +  . The number A1 is
known as the least upper bound or supremum (l.u.b or sup) of the numbers  n :
indeed, it is the least number which is  all  n . Construct in the same way the
  n  k obtained from the original

least upper bound A k of the sequence


sequence by deleting 1 ,....,  k 1 . It is clear that { A k } is a nonincreasing
sequence, and we denote its limit by A. it may be finite  , or   . In any
case we write
A  lim sup  n
n 0

It is easy to characterize the limes superior by its properties. If A is


finite and  0 there exists an n 0 such that An 0  A   , and it follows that
 n  A   for n  n 0 , then An 0  A   , which is impossible. In other words,

there are arbitrarily large n for which  n  A   . If A   there are


arbitrarily large  n , and A = -  if and only if  n tends to -  . In all
cases there cannot be more than one number A with these properties.
The limes inferior can be defined in the same manner with inequalities
reversed. It is quite clear that the limes inferior and limes superior will be
equal if and only if the sequence converges to a finite limit or diverges to
 to   . The notations are frequently simplified to lim and lim . The
reader should prove the following relations:
lim n  limn lim   n   n  lim n  limn
lim n  limn lim   n   n  lim n  limn
Now we return to the sufficiency of Cauchy’s condition. From
 n   n0  we obtain  n   n 0   for n n 0
, and it follows that
A  lim n and a  lim n are both finite. If a  A choose

22

 A  a
3

and determine a corresponding n 0 . By definition of a and A there exists an


 n  a   and an  m  A  with m, n n 0 . It follows that
A  a   A   m     m  a   3
, contrary to the choice of  . Hence a = A,
and the sequence converges.
Series.
A very simple application of Cauchy’s condition permits us to deduce
the convergence of one sequence from that of another. If it is true that
bm  bn  a m  a n b 
for all pairs of subscripts, the sequence n may be termed
a 
a contraction of the sequence n (this is not a strandard term ). Under this
a  b 
condition, if n is a Cauchy sequence, so is n . Hence convergence of
 a n  implies convergence of  bn  .
An infinite series is a formal infinite sum

(15) a1  a 2  ....  a n  ....

Associated with this series is the sequence of its partial sums


s n  a1  a 2  ....  a n
The series is said to converge if and only if the corresponding sequence
is convergent, and if this is the case the limit of the sequence is the sum of the
series.
Applied to a series Cauchy’s convergence test yields the following condition:
The series (15) converges if and only if to every  0 there exists at n 0 such
a n  a n 1  ....  a n p  for all n n 0 and p0
that . For p = 0 we find in
a 
particular that n . Hence the general term of a convergent series tends to
zero. This condition is necessary, but of course not sufficient.
If a finite number of the terms of the series (15) are omitted, the new
series converges or diverges totgethe5r with (15). In the case of convergence,
let R n be the sum of the series which begins with the term a n 1 . Then the sum
of the whole series is S  s n  R n .
The series (15) can be compared with the series
a1  a 2  ...  a n  ...
(16)
Formed by the absolute values of the terms.

23
The sequence of partial sums of (15) is a contraction of the sequence
a n  a n 1  ....  a n p  a n  a n 1  .....  a n p
corresponding to (16), for .
Therefore, convergence of (16) implies that the original series (15) is
convergent. A series with the property that the series formed by the absolute
values of the terms converges is said to be absolutely convergent.

24
NOTES
…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

25
UNIT - II
ELEMENTARY PROPERTIES OF ANALYTIC FUNCTIONS
2.1 In this chapter and afterwards we shall use some terms in dealing with sets
of complex numbers in the Argand diagram. It is therefore necessary to define
and explain such terms.
2.2 Neighbourhood of point z0.
By a neighbourhood of a point z0 in the Argand diagram. We mean a set
z  z0  
of all points z such that , where  is an arbitrary small positives
number,  is called the radius of this neighbourhood.
Deleted Neighbourhood of z0.
z  z0  
If the neighbourhood of the point z0 defined by excludes the
point z0, then such a neighbourhood is called the deleted neighbourhood of z0.
2.3 Limit point.
A point z0 is said to be a limit point or a limiting point of a set of points
S in the Argand plane, if every neighbourhood of z 0 contains a point S other
than z0. The definition implies that if the neighbourhood of z 0 contains one
point of S, it contains infinite number of points of S.
The limit point z0 may not necessarily be the point of the set S. For
z  r.
example let a set of points be defined by In virtue of definition all the
z  r, z  r.
points on the circle are the limit points of the set but they do not
belong to this set.
z  r,
Also all the points interior to the circle are the limit points of the
z  r,
set of and do belong to the set. So there are both possibilities, the limit
points of a set may or may not belong to the set.
2.4 Interior points and boundary points.
There are two type of limit points, interior points (or inner points) and
boundary points. A limit point z0 of S is an interior point if in the
neighbourhood of z0 there exists entirely the point of S.
If all the points in the neighbourhood of the limit point z 0 do not belong
to S, then z0 is called boundary point. To illustrate, all the points inside the
z  r,
circle are the inner points of the set which all the points on the
z  r,
circumference of the circle are the boundary points of the set.
2.5 Open set and closed set.
A set which consists entirely of inner points is said to be open set.

26
If all the limits (both inner and boundary) of a set belong to the set, then
z  r,
such a set is said to be closed set. To illustrate, a set is an open set
because only those limiting points which are inner points, belong to the set; but
z r
a set is a closed set for its limiting points (both inner and boundary
belong to the set.)
It should be observed that there can be sets which are neither open nor
closed. For example a set, consisting of the point z =1 and all the points for
z r
which , is neither open nor closed;
2.6 Bounded set and unbounded set.
As set of points is said to be bounded if there exists a finite positive
z K
number K such that for all points z of the set. If there exists no such
number K, the set is unbounded.
2.7 Domain (or Region).
The word domain means connex (or connected). A set of points in the
Argand diagram is said to be connex if any two of points can be joined by a
continuous polygonal arc which consists only of the points of the set.
Thus an open connex set of points means an open domain or open
region. If the boundary points of the set are also added to an open domain, it is
then called a closed domain.
2.8 Jordan arc.
The set of points z in the Argand diagram determined by the equation z
= x(t)+iy(t), where x(t) and y(t) are real continuous functions of a real variable
l, defined in the range   t   , is called continuous arc. A point z is a1
multiple pint of the arc if the equation z 1 = x(t)+iy(t) is satisfied by more than
one value of t in the given rang. Particularly the multiple point is called a
double point when the above equation is satisfied by two values of I in the
given range. A Jordan arc is defined as a continuous arc without multiple
points.
Simple closed Jordan curve. If the two end points of a continuous arc
corresponding eto the terminal values  and  of t coincide, the arc has then
only one multiple point (which is the double point corresponding to the two
terminal values of t. then such a continuous curve is termed as a simple closed
Jordan curve.
The Jordan curve theorem states that a simple closed Jordan curve
divides the Argand plane into two open domains which have the curve as
boundary.
Of these two domains one is bounded and it is called the interior
domain ; the other is unbounded and it is called the exterior domain.

27
z 1
For example, the circle divides the Argand diagram into two open
z 1 z 1 z 1
domains and . Of these two domains one is bounded and is
z 1 z 1
the interior of the circle ; the other is unbounded and is the
z 1
exterior of the circle .
2.9 Functions of a Complex Variable.
A complex variable w is said to be a function of the complex variable z,
if, to every value of z in a certain domain D, there correspond one or more
definite values of w.
Thus if w is function of z, it is written as w= f(z).
If w = u+iv and z = x + iy, then u and v both the functions of x,y; so we
write then as u(x,y) and v(x,y). Thus in a more explanatory way we may write.
w= u(x,y) + iv(x,y) where x and y are real.
If w takes only one value for each value of z in the region D, then w is
said to be a uniform or single valued function of z. if there correspond two or
more value of w for some or all values of z in the region D, then w is called a
many-valued or multi-valued function of z.
Path of a complex variable, the independent variable z takes values
corresponding to the points lying on the x-axis. But in the case of the function
f(z) of a complex variable the independent variable z takes values lying in a
plane (called the z plane), so that path of z is a polygonal are joining the initial
and the final point z. This path may be a straight line or a curve. So path of the
complex variable z is either a straight line or a curve.
2.10 Continuity.
In the last article a complex number w = f(z) is defined as a function of
a complex variable z if to every value of z in a certain region correspond one or
more values of w. This definition is on the same model as the usual definition
of a function of a real variable. Therefore definition of continuity for a function
of a complex variable is exactly the same as that for function of a real variable.
Thus,
The function f(z) of a complex variable z is said to be continuous at the
point z0 if, given any positive number  , we can find a number  such that,

 0  
f  z  f z

z  z0  
for all points the domain D satisfying .
The number  depends upon  and
also in general upon z0. If it is
   f  z   f  z 0  
possible to find a number independent of z0 such that

28
z  z0  
for every pair of points z, z0 of the region D of which , then the
function f(z) is said to be uniformly continuous in D.

Note.
In the case of the continuity of the functions of a complex variable z
approaches z0 along any path whatsoever, while for function of a real variable z
approaches z0 along x-axis only.
2.11 Differentiability.
Since the mode of definitions of continuity is the same both in case of
the functions of the real and the complex variables, therefore definition of the
derivative of a complex function f(z) is identical with the definition of the
derivative of the real function. Hence following the suggestion of the real
differential calculus, we write
f  z   f  z0 
f   z 0   lim
z z0 z  z0
and we say that f(z) is differentiable if the limit on the right. These are Cauchy-
Riemann conditions in the polar form.
2.15 Derivative of w in polar form.
dw w

We have dz x

w r w 
 .  .
r x  x
w  u v  sin 
 cos     i 
r     r
w  v u  sin 
 cos     r  i.r  Since w  u  iv
r  r r  r
w  u v 
 cos   i   i  sin  from  1 and  2  of  2.14
r  r r 
w w
 cos   i sin  sin ce w  u  iv
r r
w
  cos   i sin  
r
This is derivative of w.
As we have above

29
dw w r w 
 .  .
dz r x  x

 u v  w sin 
   i  cos   .
 r r   r
 1 v 1 u  w sin 
 i  cos   . sin ce w  u  iv
 r  r    r
i  u v  w sin 
    i  cos   . from  1 and  2  of  2.14
r      r
i w
   cos   i sin   this also is a derivative of w.
r 
Thus in polar form
dw w i w
  cos   i sin      cos   i sin   .
dz r r 
Example.
r n  cos n  i sin n 
If n is real, show that is analytic except possibly
nr n 1
cos  n  1   i sin  n  1   .
when r = 0, and that its derivative is
Solution.
w  r n  cos n  isin n   f  z   u  iv
Let

So here u  r cos n, v  r sin n


n n

u v
 nr n 1 cos n,  nr n 1 sin n
then r r
u v
  nr n sin n,  nr n cos n
 
u 1 v 1 u v
 . and .  
Thus we see that r r  r  r
That is, Cauchy-Riemann conditions are satisfied.
dw
Hence the function w  r (cos n  i sin n) is analytic, if dz exists for
n

all finite values of z.


dw w
  cos   i sin  
We have dz r

30
  cos   i sin   .nr n 1  cos n  i sin n 
 nr n 1  cos  n  1   i sin  n  1 
This exists for all finite values of r including zero, except when r = 0
and n  1.
2.16 Function of a function.
w  F   and  f  z  ,
If then w is called function of a function of z.
dw dw d
 . .
We have dz d dz

dw dw d
Hence dz exists if both d and dz exist, that is
w  F     f  z 
If is an analytic function of and is an analytic
function of z.
Inverse Functions.
Let w = f(z) be an analytic function of z, where corresponding to any
point z0 there is a point w0. Now if the relation z = F(w) is such that the
corresponding to the value w0 of w there is the value z0 of z, then the function z
= F(w) is called the inverse function of w = f(z) where z 0 corresponds to w0.
f  z 0   0
If , then w0 is a regular point of z = F(w), that is z is
analytic in the neighbourhood of w0.
1
F  w 0  
f ' z 0 
Also (on account of functions being inverse)
F  w 0 
Hence if F(w) is to be analytic at w 0, then should be finite. This
f  z 0 
requires that should not be zero. Hence
f  z   0
The function z = F(w) will cease to be analytic where or say
dw
 0.
where dz
It immediately follows that the function w = f(z) where z = F(w) will
dz
 0.
cease to be analytic where dw
It may be seen that when
dw w
 .
z = x + iy and w = f(z) we have dz x

31
dz z
w  u  iv and z  f  w  
Therefore, if we should have dw u
Example. 1.
For what values of z do the function w defined by the following
equations ceases to be analytic ?
z  sin u cosh v  i cos u sinh v w  u  iv

Solution.
z  sin u cosh v  i cos u sin hv   1
Here
z
  cos u cosh v  i sin u sinh v.
u
dz z
  See above 
We have dw u
dz
 cos u cosh v  i sin u sinh v.   2
i.e., dw
Squaring and adding (1) and (2), we have
2
 dz 
z     sin u  cos u   cosh v  sinh v   1,
2 2 2 2 2

 dw 
dz
dw
 1 z  .
2

i.e.,
Hence, w ceases to be analytic
dz
 0, i.e, when z  1.
When dw
Example. 2.
For what values of z the function w defined by the following equations
ceases to be analytic ?
z  e  v  cos u  i sin u  w  u  iv.
Sol.
dz z

We have dw u
z  e  v  cos u  i sin u  .
Here
z
  e  v   sin u  i cos u  ,
dw

32
dz  dz z 
  e  v   sin u  i cos u   Since  
dw  dw u 
dz dz
 iz  0, i.e.,  iz.
Thus, dw dw
dz
 0 i.e., when z  0.
Hence, w ceases to be analytic when dw

Ex. 3.
For what values of z the function defined by the following equation
ceases to be analytic.
z  sinh u cos v  i cosh u sin v. where w  u  iv
Sol.
z  sinh u cos v  i cosh u sin v.   1
Here
dz z
  .
dw u
 cosh u cos v  t sinh u sin v   1
From (1) and (2) after squaring and subtracting, we get

 dz 
  z   cosh u  sinh u   cos v  sin v   1
2 2 2 2 2

 dw  .

 dz  dw 1
  z  1 or 
2
i.e., 
i.e.,
 dw  dz z 2
 1

dz
 0. i.e., when z 2  1  0
Hence w ceases to be analytic when dw .
or z  i.
Ex. 4.
For what values of z the function defined by the following equations
ceases to be analytic.
z  log   i where w    cos   i sin  
Sol.
We know that
w  f  r,   and z  r  cos   isin   ,
If then

33
dw w
  cos   i sin   .
dz r
z  log   i and w    cos   i sin  
Here
dz z
 Here   cos   i sin   u sin g the above result
dw 
1
  cos   i sin  


 cos   i sin    cos   i sin   
1
.
w w
1
 0.
Hence w is analytic so long as w
i.e., so long as w is finite.
Hence w is an analytic function in any finite region.
Ex. 5.
Show that the function f(z) = sinx cosh y + i cos x sinh y is continuous
as well as analytic everywhere.
Sol.
Let f(z) = u (x,y) + iv (x,y) then u (x,y) = sinx cosh y, v(x,y) = cos x
sinh y.
Since u and v both are rational functions of x and y, whose
denominators are non-zero for all values of x and y, therefore u and v are both
continuous everywhere.
Hence f(z) is also continuous everywhere.
u u
 cos x cosh y,  sin x sinh y
Again, x y
v v
  sin x sin x sinh y,  cos x cosh y.
x y
The four partial derivatives are rational functions of x and y with non-
zero denominators for all values of x and y, therefore they are continuous
everywhere.
u v u v
 and 
Also here, x y y x
that is, Cauchy-Riemann conditions are satisfied.
Thus the four partial derivatives being continuous everywhere and
Cauchy-Riemann equations being satisfied, the function f(z) is analytic
everywhere.

34
Ex. 6.
Show that the function f(z) = xy + iy is everywhere continuous but is
not analytic.
Sol.
Let f(z) = u + iv, where u(x,y)=xy, v(x,y) = y
Thus we see that both u and v being polynomials in x and y are
continuous everywhere. Hence f(z) is continuous everywhere.
u u u v
 y,  x,  0,  1.
Here x y x y
u v u v
 and  .
Thus we see that x y y x
That is, Cauchy-Riemann conditions are not satisfied anywhere. Hence
f(z) is not analytic at any point.
Ex. 7.
Show that the functions
f  z   e z  z  0 and f  0   0
4

is not analytic at z = 0, although Cauchy-Riemann equations are satisfied at the


point. How would you explain this.
Sol.
 x  iv   4
Here u  iv  e
 x  iy 
4


x  y2 
2 4

or u  iv  e

1
  x 4  y 4  6x 2 y 2   i4xy  x 2  y 2  
 
x 2
y 
2 4

e
x 4  y 4  6x 2 y 2 4xy  x 2  y 2 
 .cos
x  y2  x  y2 
2 4 2 4

u=e
4xy  x 2  y 2 
4
x 4  y 4  6x 2 y 2
 .sin
x  y2  x  y2 
2 4 2 4

35
v  e .
u u  x, 0   u  0, 0  e  x 4
At z  0,  lim  lim 0
x x 0 x x 0 x
u u  0, y   u  0, 0  e  x 4
 lim  lim 0
y y0 y y 0 y
v v  x, 0   v  0, 0  0
 lim  lim  0
x x 0 x x 0 x

v v  0, y   v  0, 0  0
 lim  lim  0
y y  0 y y  0 y
Hence Cauchy-Riemann conditions are satisfied at z =0.
f  z   f  0 e  z 4
f   0   lim 
But z 0 z z
4
e  r .e
 lim i / 4 ,if z  0 along z  rei / 4
z  0 re



Showing that
f   z  does not exist at z = 0.
Hence f(z) may not be analytic at z = 0.
Note.
f  z
may not be analytic at z = z0 if merely the Cauchy-Riemann

conditions are satisfied there. In addition to this


f  z
must exist and should
be unique at z, and at every point in the neighbourhood of z0.
Ex. 8.
Prove that the function f(z) = u + iv, where
x 3  1  i   y3  1  i 
f  z   z  0 , f  0  0
x 2  y2
is continuous and that Cauchy-Riemann equations are satisfied at the origin,
f  z
yet does not exist there.
Sol.
x 3  y3 x 3  y3
u 2 ,v 2  where z  0 
Here x  y2 x  y2

36
Here we see the both u and v are rational and finite for all values of
z  0 , so u and v are continuous at all those points for which z  0 . Hence f(z)
continuous where z  0 .
At the origin u = 0, v = 0. [Since f(0) = 0]
Hence u and v are both continuous at the origin; consequently f(z) is
continuous at the origin.
Hence f(z) is continuous everywhere.
At the origin
u u  x, 0   u  0, 0  x
 lim  lim    1
x x  0 x x  0
x
u u  0, y   u  0, 0   y 
 lim  lim    1
y y 0 y y0
 y 
v v  x, 0   v  0, 0  x
 lim  lim    1
y x 0 x x 0 x
 
v v  0, y   v  0, 0   y
 lim  lim    1
y y0 y y0 y
 
u v u v
 and 
Thus we see that x y y x .
Hence Cauchy-Riemann conditions are satisfied at z = 0.
f  z   f  0
f   0   lim
Again z 0 z

 x 3  y3  i  x 3  y 3  1 
 lim  . 
z 0
 x 2  y2 x  iy 

Now let z  0 along y = x, then
x3  x3  i  x3  x3  1
f   0   lim .
x 0 x x
2 2
x  ix
2i 1
 lim   1  i .
x 0 2  1  i  2

Again let z  0 along y = 0, then


x3  1  i 
f   0   lim 1 i
x 0 x3 .
f   0 f   0
So we see that is not unique, i.e. the values of are not the
same as z  0 along different curves.

37
f  z
Hence does not exist at the origin.
Ex. 9.
Examine the nature of the function
x 2 y5  x  iy 
f  z  z0
x 4  y10
f  0  0

in the region including the origin.


Sol.
x 2 y5  x  iy 
u  iv  
Here x 4  y10

x 3 y5 x 2 y5
u  4 , v 4
x  y10 x  y10
At the origin,
u  x, 0   u  0, 0  00
u x  lim  lim  0
x 0 x x  0
 x 
4  0, y   4  0, 0  00
u y  lim  lim  0
y 0 y y 0
 y 
v x  0, v y  0
Similarly
Hence Cauchy-Riemann equations are satisfied at the origin.
f  z  f  0  x 2 y5  x  iy   1
lim  lim   0 
 x y  x  iy
4 10
z 0 z x 0
 y0 
But
Since f(0)=0 and z = x + iy
x 2 y5
 lim
x 0 x 4  y10
y 0

x 2 m5 y5
 lim 4
x 0 x  m10 y10

m5 x 3
 lim 0
x 0 1  m10 x 6

If z  0 along the radius vector y = mx

x2 x2
and  lim if z  0 along the curve y5  x 2
x 0 x 4  x 4

38
1

2
f   0
Showing that does not exist.
Hence f(z) is not analytic at origin although Cauchy-Riemann equations
are satisfied there.

39
Ex. 10.
Show that the function
f  z   xy 
is not regular at the origin, although Cauchy-Riemann equations are satisfied at
the point.
Sol.
Let the function be f(z) = u(x,y) + iv (u, y)

Here
u  x, y    xy  and v  x, y   0
.
At the origin.
u u  x, 0   u  0, 0  00
 lim  lim  0
x x  0 x x  0
 x 
u u  0, y   u  0, 0  00
 lim  lim  0
y y  0 y y  0
 y 
v v  x, 0   v  0, 0  00
 lim  lim  0
x x 0 x x 0
 x 
v v  0, y   v  0, 0  00
 lim  lim  0
y y0 y y 0
 y 
Hence Cauchy-Riemann equations are satisfied at the origin.

f  z   f  0  xy 
f   0   lim  lim .
Again,
z 0 z z 0 x  iy

 lim
 mx  2

letting z  0 aloing y  mx
x 0 x  imx
m

1  im
f   0
this depends on m, that is is not unique.
Hence f(z) is not analytic at the origin although Cauch-Riemann
equations are satisfied there.
2.17 Orthogonal System.
Two curves are said to be orthogonal to each other when they intersect
at right angle at each of their points of intersection,
To show that, if w = f(z) = u + iv be analytic function of z = x + iy, then
the family of curves u (x,y) = c1 and v(x,y) = c2 form an orthogonal system.

40
For the curve u(x,y) = c1, we have on differentiating,
v v dy u / x
dx  dy  0, i.e., 
x y dx u / y
And for the curve v(x,y) = c2, we have on differentiating,
v v dy v / x
dx  dy  0, i.e., 
x y dx v / y

The product of the above two


dy / x , s
 u / x   v / x 
   
 u / y   v / y 
v / y  v / x 
    1  Since u x  v y, u y   v x 
y / x  v / y 
Hence the two curves u = c1 and v = c2 are orthogonal, and since c1 and
c2 are parameters therefore we say the two family of curves u = c 1 and u = c2
form an orthogonal system.
2.18. Harmonic functions.
Any function of x,y which possesses continuous partial derivatives of
the first and second orders and satisfies Laplace’s equation is called a harmonic
function.
Theorem.
If f(z) = u + iv is an analytic function, then u and v are both harmonic
functions
Let f(z) = u + iv, be an analytic function, then we gave
u v 
 
x y 
 ........  1 Cauchy  Riemann Equations.
u v 
or 
y x 
Also, because u and v are the real and imaginary parts of an analytic
function, therefore derivatives of a u and v, of all orders, exist and are
continuous functions of x and y
So that we have
2 v 2 v
   2
xy yx
Differentiating equations (1), we have
2u 2 v 2u 2 v
 and  
x 2 xy y 2 yx

41
2u 2u
 2 0
Adding these we have  2
x y by virtue of (2)
2v 2v
 0
Similarly  2 x y 2
Hence both u and v satisfy Laplace’s equations
2V 2 V
 0
x 2 y 2
Therefore both u and v are harmonic functions.
Such functions u and v are called conjugate harmonic functions or
simple conjugate functions.
2.19 Determination of the conjugate function.
If f(z) = u + iv is an analytic function where both u (x,y) and v(x,y) are
conjugate function. Being given one of these say u(x,y), to determine the other
v(x,y).
Since v is a function of x,y therefore
v v
dv  dx  dy
x y
u u
 dx  dy
y x by Cauchy-Riemann equations.
u u
dv   dx  dy
Thus y x  (1)
The right hand side, of equation (1) is of the form
u u
Mdx  Ndy, where M   and N  .
y x
M   u  u 2
      ,
So that y y  y  y 2

N   u  2 u
   2.
and x y  x  x
Since u is a harmonic function, therefore it satisfies Laplace’s equation,
 2 u u 2 2u 2u
  0 or  
i.e., x 2 y 2 y 2 x 2

M N
 .
which makes y x

42
Hence equation(1) satisfies the condition of exact differential equation.
So equation (1) can be integrated and thus v determined.
Ex.1.
1
u log  x 2  y 2 
Show that function 2 is harmonic and find its
harmonic conjugate.
Sol.
1
u log  x 2  y 2 
We have 2
u x  2 u x 2  y 2  2x 2 y2  x 2
  2 ,  
x x  y 2 x 2  x 2  y2   x 2  y2 
2 2

u y  2 u x 2  y 2  2y 2 x 2  y2
 2 ,  
y x  y 2 y 2  x 2  y2   x 2  y2 
2 2

and
2u 2u y2  x 2 y2  x 2
    0.
x 2 y 2  x 2  y 2  2  x 2  y 2  2
so that
Hence u is a harmonic function.
Let u is a harmonic function.
Let v be the conjugate harmonic function.
v v
dv  dx  dy
We have dx y (since v is function of x,y)
u u
 dx  dy
y x by Cauchy-Riemann conditions
y x xdy  ydx
 dx  2 dy  2
x y
2 2
x y 2
x  y2
Integrating it, we get
y
v  tan 1  c,
x where c is real constant.
Ex. 2.
Show that the function u = cos x cosh y is harmonic and find its
harmonic conjugate.
Sol.
We have u = cos x cosh y

43
u u
  sin x cosh y,  cos x sinh y
x y
2u 2u
 2   cos x cosh y  cos x cosh y  0
So that x 2
y

Hence u is a harmonic function.


Let v be its conjugate harmonic function we have
v v u u
dv  dx  dy   dx  dy
x y y x
  cos x sinh y dx  sin x cosh y dy
   cos x sinh ydx  sin x cosh ydy  .
Integrating this, we get
v= - sinx sinh y + c, where c is a real constant.
Ex. 3.

Prove that u  y3  3x 2 y is a harmonic function. Determine its


harmonic conjugate and find the corresponding analytic function f(z) in terms
of z.
Sol.

We have u  y3  3x 2 y .

u u 2  u
2
2u
  6xy;  3y  3x , 2  6y, 2  6y
2

x y x y
2u 2u
 2  6x  6y  0,
So that x y
2

that is, u satisfies Laplace’s equation ; hence u is a harmonic function.


Let v be the harmonic conjugate to u we have
v v
dv  dx  dy
x x
u u
  dx  dy
y x
 Since u x  u y and u y   v x 

   3y 2  3x 2  dx  6xy dy
   3y 2 dx  6xydy   3x 2 dx.

Integrating, v  3xy  x  c, this is harmonic conjugate to u.


2 3

44
f (z)  u  iv
 y3  3x 2 y  i  3xy 2  x 3  c   (1)
 i  x  iy   ic  iz 3  ic
3
 (2)
Note:
Form (2) is suggested by noting that when y = 0, equation (1) becomes
f (z)  ix 3  c .

Ex. 4

If u  x  3xy , show that there exists a function v(x,y) such that


3 2

w  u  iv is analytic in a finite region. (Agra 51)


Sol:

Here u  x  3xy .
3 2

u 2u
  3x 2  3y2 ,  6x.
x x 2
u 2 u
 6xy,  6x.
y y 2
2u 2u
so that   6x  6x  0
x 2 y 2
Hence the given function u satisfies Laplace’s equation and is therefore
a harmonic function.
Since v is function of x and y, therefore

45
v v
dv  dx  dy
x y
u u
  dx  (by Cauchy  Riemann conditions)
y x
 6xydx   3x 2  3y 2  dy
  6xydx  3x 2dy   3y 2dy
Integrating this equation, we get
v  3x 2 y  y 3  c where c is cons tan t
Hence the function f (z)  u  iv
 x 3  3xy 2  i  3x 2 y  y3  c 
  x  iy   ic  z 2  ic,
3

i.e., f (z)  z 3  ic

So that f (z)  3z , which exists for all finite values of z.


' 2

Therefore, f(z) is analytic in any finite region.

Ex 5.

If u  (x  1)  xy  3y , determine v so that u+iv is a regular function


3 2 2

of x+iy.
Sol.

Here u  (x  1)  xy  3y
3 2 2

u u
so that  3(x  1) 2  3y 2 and  6xy  6y
x y
sin ce v is a function of x, y therefore
v u
dv  dx  dy
x x

46
u u
 dx  dy (u sin g Cauchy  Riemann equations)
y x
 (6xy  6y)dx   3(x  1) 2  3y 2  dy
 (6xy  6y)dx   3x 2  6x  6x  3  3y 2  dy
Integrating we have
v   (6xydx  3x 2 dy)   (6ydx  6xdy)    3  3y 2  dy
Ex. 6
If  and  are functions of x and y satisfying Laplace’s equations,
show that
 s  it  is analytic.
   
s  and t  
where y x x y
Sol:
Since  and  both satisfy Laplace’s equation
2 2 2  2
   0, and  0
x 2 y 2 x 2 y 2
   
s  and t  
Given y x x y ,

s  2  2 t  2  2
   and  
x xy x 2 y yx y 2
s t   2  2 
Thus     2  2   0 (sin ce  is harmonic)
x y  x y 
s t
Hence 
x y
s  2   2 t  2   2 
Also   and  
y y 2 yx x x 2 xy
s t  2  2
Thus     0 (sin ce  is harmonic)
y x x 2 y 2
s t
Hence 
y x
s t s t
So it is proved  and 
x y y x
That is Cauchy-Riemann conditions are satisfied.

47
Hence (s+it) is an analytic function.
Ex. 7

u  x 2  y2 , v   y
Prove that if (x 2  y 2 ) both u and v satisfy
Laplace’s equation, but u+iv is not an analytic function of z.
Sol.
We have
y
u  x 2  y2 , v 
(x  y 2 )
2

u u
 2x,  2y
x y
v   x  y   2y
2 2 2
y2  x 2 v 2xy
  , 
y  x 2  y2   x 2  y2  x  x  y 2  2
2 2 2

 2 v 2y  y  3x   2 v 2y  3x  y 
2 2 2 2
 2u  2u
 2,  2,  , 
x 2 y 2 x 2  
2 3   x 2  y2 
2 3
x 2
 y y

2u 2u
so that   22  0
x 2 y 2
 2 v  2 v 2y  y  3x   2y  y  3x 
2 2 2 2

Also   0
x 2 y 2 x y 
2 2 3

Hence both u and v satisfy Laplace’s equation.


But we notice that
u v u v
y  y and y  - x
That is, Cauchy-Riemann conditions are not satisfied.
Hence u+iv is not an analytic function.
2.20 To construct a function f(z) when one conjugate function is given.
Let f(z) = u+iv be analytic function, when u and v are the conjugate
function (i.e. both u and v satisfy Laplace’s equation). If either of them say u is
given, then to form f(z).
Since u is already given, therefore the function f(z) can be constructed
if either only v is found out, or function f(z) as a whole is directly find out.
The one method to construct f(z) when one of the conjugate functions,
say u(x,y) is given, is to determine only v as has been done in the last article
and then the function f(z) so constructed is u+iv.

48
The other method to construct f(z) directly without finding v is due to
Milne Thomson and is an elegand method given below:

Since
x
1
2
 
z  z and y 
1
2i
zz  
z  z z  z  z  z z  z 
 f (z)  u  ,   iv  , 
 2 2i   2 2i 
This relation can be regarded as a formal identity in two independent

variable z and z. Putting z =z,


We have x = z and y = 0 and then f(z)=u(z,0)+iv(z,0),
u v
 f ' (z)  i
we have f(z) = u+iv. x x
u u
 i
x y by Cauchy-Riemann equations
u u
 1 (x, y) and   2 (x, y)
Hence if we write x y

we have f ' (z)  1 (x, y)  i2 (x, y)


 1 (z, 0)  i2 (z, 0)
Integrating it, f (z)    1 (z, 0)  i2 (z, 0) dz  C

where C is an arbitrary constant.


Thus function f(z) is constructed when u(x,y) is given.
Similarly if v(x,y) is given, it can be shown that
f (z)    1 (z, 0)  i 2 (z, 0) dz  C
v v
1 (x, y)  and  2 (x, y) 
where y x
Remark:
In the following examples we find f(z) directly by using Milne’s
method.
Ex. 1.
sin 2x
u
If cosh 2y  cos 2x , find the corresponding analytic function
f(z)=u+iv.
Sol.

49
sin 2x
u
We have cosh 2y  cos 2x

u 2 cos 2x  cosh 2y  cos 2x   2sin 2 2x



x  cosh 2y  cos 2x 
2

2  2 cos 2x cosh 2y
  1 (x, y) say
 cosh 2y  cos 2x 
2

and
u 2sin 2x sinh 2y
  2 (x, y) say
y  cosh 2y  cos 2x 
2

The corresponding function f (z)  u  iv is given by

f (z)    1 (z, 0)  i2 (z, 0) dz  C


2  2 cos 2z 1
 dz  c  2  dz  c
 1  cos 2z  1  cos 2z
2

  sec 2 zdz  c.
i.e, f (z)  tan z  c
Ex. 2.

Prove that the function u  e (x cos y  y sin y) satisfies Laplace’s


x

equation and find the corresponding analytic function f (z)  u  iv .


Sol.

Here u  e (x cos y  y sin y)


x

u
  e x (x cos y  y sin y  cos y)  1 (x, y)
x
u
 e x ( x sin y  y cos y  sin y)   2 (x, y)
y
By Mline 's method we have
f ' (z)  1 (z, 0)  i2 (z, 0)  e z (z  1)
Integrating , we have
f ' (z)   e z (z  1)dz  c  ze z  c.
Thus f (z)  ze z  c.
Ex. 3.
Find the regular function w = u+iv where
u  e x  x 2
 y 2  cos y  2xy sin y 

50
Sol.

Here
u  e x  x 2
 y 2  cos y  2xy sin y ,
u

x
 
 e x   x 2  y2  cos y  2xysin y  e x  2x cos y  2ysin y  1 (x, y) say

u  x
and
y
 
 e   x 2  y 2  sin y  2y cos y  2x sin y  2x cos y   2 (x, y) say

By Mi ln e 's method we have


f ' (z)  1 (z,0)  i2 (z,0)  e  z ( z 2  2z)
Integrating f (z)   e  z ( z 2  2z)dz  c

   e z z 2 dz  2 ze  z dz  c

 e z z 2  2 ze z dz  2  ze z dz  c
 z 2 e z  c
Thus f (z)  z 2 e  z  c where z  x  iy.
Ex. 4.

Show that the function u  sin x cosh y  2 cos x sinh y  x  y  4xy


2 2

satisfies Laplace’s equation and determine the corresponding analytic function


f(z)=u+iv.
Sol.
2 2
Here u sin x cosh y  2cos x sinh y  x  y  4xy
u
  cos x cosh y  2sin x sinh y  2x  4y
x
 1 (x, y) say
u
 sin x sinh y  2 cos y cosh y  2y  4x
y
 2 (x, y) say
 2u
  sin x cosh y  2 cos x sinh y  2  (1)
x 2
2u
 sin x cosh y  2 cos x sinh y  2  (2)
y 2
2u 2u
Adding (1) and (2) we see that  0
x 2 y 2
Hence u satisfies Laplace’s equation. In other words u is a harmonic
function.

51
By Milne’s method we have
f ' (z)  1 (z, 0)  i2 (z, 0)
  cos z  2z   i  2 cos z  4z 
Integrating, we get
f (z)    1  2i   cos z  2z  dz  c
or f (z)   1  2i   sin z  z 2   c.

Ex. 5.

Prove that the function u  x  3xy  3x  3y  1 satisfies Laplace’s


3 2 2 2

equation and determine the corresponding analytic function u+iv.


Sol.

Here u  x  3xy  3x  3y  1
3 2 2 2

u
 3x 2  3y 2  x  1 (x, y) say
x
u
 6xy  6y  2 (x, y) say
y
2 u 2 u
 6x  6,  6x  6
x 2 y 2
2u 2u
so that   6x  6  6x  6  0.
x 2 y 2
i.e. u satisfies Laplace’s equation. Here u is a harmonic function.
By Milne’s method we have
f ' (z)  1 (z, 0)  i2 (z, 0)
  3z 2  6z 

Integrating it, we get


f (z)    3z 2  6z  dz  c
f (z)  z 3  3z 2  c.
Ex. 6.
If f (z)  u  iv is an analytic function of z = x+iv and
u  v  e x  cos y  sin y 
, find f(z) in terms of z.
Sol.
f (z)  u  iv

52
Adding, we have
 u  v   i  u  v    1  i  f (z)  F(z) say
Put u-v = U and u+v = V, then F(z) = U+iV is an analytic function.
Now,
U  e x  cos x  sin y 
U
  e x  cos y  sin y   1 (x, y)
x
U
and  e x   sin y  cos y   2 (x, y)
y
By Milne’s method we have
F' (z)  1 (z, 0)  i2 (z, 0)
  e z  ie z    1  i  e z

Integrating it, we get


F(z)    1  i  e z dz  c
  1  i  ez  c
 1  i  f (z)   1  i  e z  c
or f (z)  e z  c1
Ex. 7.
u  v   x  y   x 2  4xy  y 2  and f (z)  u  iv
If is an analytic
function of z = x+iy, find f(z) in terms of z.
Sol.
u  iv  f (z)

Adding these,
 u  v   i  u  v    1  i  f (z)  F(z), say where F(z) is
analytic because f(z) is so.
Let U = u-v and V = u+v, then F(z) = U+iV is an analytic function.
We have
U  u  v   x  y   x 2  4xy  y 2 
U
  3x 2  6xy  3y 2  1 (x, y)
x
U
 3x 2  6xy  3y 2  2 (x, y)
y
By Milne’s method, we have

53
F(z)    1 (z, 0)  i2 (z, 0) dz  c

   3z 2  i3z 2  dz  c   1  i  3z 2  c
i.e.  1  i  f (z)   1  i  3z 2  c
1 i 3 c
i.e. f (z)  z   iz 3  c1
1 i 1 i
Thus f (z)  iz  c1
3

Ex. 8.
2sin 2x
uv , and f (z)  u  iv
If e  e 2y  2 cos 2x
2y
is an analytic function
of z = x+iy, find f(z) in terms of z.
Sol.
Here u+iv = f(z)  iu  v  if (z)

Adding these
 u  v   i  u  v    1  i  f (z)  F(z) say

Now F(z) is analytic because f(z) is analytic.


Let U = u-v and V = u+v, then F(z) = U+iV is an analytic function.
2sin 2x 2sin 2x
Here Vuv 2 y

e  e  2 cos 2x 2 cosh 2y  2 cos 2x
2y

V 2sin 2x sinh 2y
   1 (x, y)
y  cosh 2y  cos 2x 
2

V 2 cos 2x  cosh 2y  cos 2x   2sin 2x


2

and    2 (x, y)
x  cosh 2y  cos 2x 
2

By Milne’s method, we have


F(z)  1 (z, 0)  i 2 (z, 0)
2 cos 2z  1  cos 2z   2sin 2 2z
 0i
 1  cos 2z 
2

1
 2i  i csc 2 z
1  cos 2z
Integrating it, we get
F(z)  i  csc 2 zdz  c  i cot z  c.
i.e.,  1  i  f (z)  i cot z  c,
i c 1
 f (z)  cot z    1  i  cot z  c1
1 i 1 i 2

54
Ex. 9.
cos x  sin x  e  x
uv
If f (z)  u  iv is an analytic function and 2 cos x  e y  e  y ;

f 0
find f(z) subject to the condition  2 
Sol.
We have f (z)  u  iv
On adding, we have
 u  v   i  u  v    1  i  f (z)  F(z) say,
i.e.,  u  v   i  u  v   F(z)
Let u  v  U and u  v  V, then U  iV  F(z) is an analytic function.
cos x  sin x  e  y 1  sin x  sinh y 
U  1
Now, 2 cos x  2 cosh y 2  cos x  cosh x 
Differentiating,
U 1 1  sin x sinh y  cos x cosh y
  1 (x, y)
x 2  cos x  cosh y 
2

U 1 1  sin x sinh y  cos x cosh y


  2 (x, y)
y 2  cos x  cosh y 
2

By Milne’s method, we have


F' (z)  1 (z, 0)  i2 (z, 0)
1 1 i 1
 .  .
2 1  cos z 2 1  cos z
1
 
  1  i  csc 2 z
4 2
Integrating it we get

F(z)  
1
2
 
 1  i  cot z 2  c
i.e.
1
 
 1  i  f (z)    1  i  cot z 2  c
2
or
1
2
 
f (z)   cot z  c1
2

But given that when


z   ,f   0
2 2    c1 
1
2

55
Hence
f (z) 
1
2
1  cot z
2 
Ex. 10
If f (z)  u  iv is an analytic function of
e  cos x  sin x
y
z  x  iy and u  v  ,
cosh y  cos x find f(z) subject to the conditions

 
f  
2
3  i
2 .
Sol.
We have u  iv  f (z).
On adding, we have
u  v  i  u  v    1  i  f (z)  F(z) say
i.e.,  u  v  i  u  v  F(z)

Let U  u  v, and V  u  v, then U  iV  F(z) is an analytic


function.
e y  cos x  sin x cosh y  sinh y  cos x  sin x
Here U 
cosh y  cos x cosh y  cos x
sinh y  sin x  sin x  sinh y 
 1  1  
cosh y  cos x  cos x  cosh y 
U 1  sin x sinh y  cos x cosh y
   1 (x, y)
x  cos x  cosh y 
2

U 1  sin x sinh y  cos x cosh y


and   2 (x, y)
y  cos x  cosh y 
2

By Milne’s method we have


F' (z)   1 (z, 0)  i2 (z, 0) 
1 1
 i
1  cos z 1  cos z
1 1 z
   1  i    1  i  csc 2
1  cos z 2 2

56
Integrating it, we get
1
F(z)   1  i   csc2 z 2  c   1  i  cot z 2  c
2
 1  i  f (z)   1  i  cot z 2  c1
    3i 3i 1 i
But when z  , f     c1  1 
2  22  2 2 2
1
hence f (z)  cot z   1  i 
2 2
2.21
 2 2  2
 2    4 .
x y 2 z z
To show that  
We have x + iy = z and x – iy = z

So that
x
1
2
 
 z  z , y   i 2  z  z ,
x 1 x 1 y i y i
  ,  ,  , 
z 2 z 2 z 2 z 2
  x  y 1    
 .  .   i 
Now, z x z y z 2  x y 

  x  y 1    
 .  .    i .
and z x  z y  z 2  x y 

 1  2 2 
  2  2 ,
z z 4  x y 
Hence
 2 2  2
i.e.,  2  2   4 .
 x y  z z

57
Ex.1.
If f(z) = u + iv is an analytic function of z = x + iy, and  is any
function of x and y with differential coefficients of the first two orders, then
2
      
2 2 2
     
  f  z
2
       
 x   y   u   v  

2 2  2 2 


 f  z
2
  
x 2 y 2  u 2 v 2 
and
Sol.
  u  u
 .  .   1
We have x u x v x
  u  v
 .  .
and y u y v y
 v  u
 .  .  as u x  v y , u v   v x  .   2
u x v x
Squaring and adding (1) and (2), we have
2
              u   v  
2 2 2 2 2

               
 x   y   u   v    x   x  

   2   2 
  f  z
2
   
 u   v  
u v
as f   z   i
x x
This proves the first required result.
Now we proceed to prove the second result.
Result (1) can be written as
 u  v 
 
x x u x v
and result (2) can be written as
 v  u 
 
y x u x v
   u  v    v u   
 i   .  .   i .  . 
x y  x u x v   x u x v 

58
 u v    u u  
 i   i i 
 x x  u  x x  v
  u v    
i.e.,  i    i    i  .   3
x y  x x   u v 
   u v     
i   i  i    4
Similarly x y  x x   u v 
Multiplying (3) and (4) column wise, we have
    u   v     2
2 2
     2 
  i   i          2  2 
 x y  x y   x   x    u v 

 2 2  2  
2
2 
 2  2   f  z  2  2 

i.e.,  x y   u v 

 2 2  2 
2
2 
 2     f   z   2   ,
or  x y2   u v 2 

 2  2    2  2 
 2   2  2  f  z .
2

x 2
y  u v 
i.e.,
This proves the second required result.
Ex. 2.
Show that an analytic function with constant modules is constant.
Sol.
Let an analytic function be f(z) = u + iv,
f  z   u 2  v2 .
2

f  z   cons tan t  c say,  c  0 


Given

i.e., u 2  v2  c2 ,

u v
u v 0   1
x y
u v
u v 0
and y y
v u
i.e.,  u v 0   2
x y

 sin ce u y   v x and v y  u x .

59
Squaring and adding (1) and (2), we have
 u   v  
2 2

 u  v   x    x    0.
2 2

 
2 2
 u   v 
i.e.,      0
 x   x 
 sin ce u 2
 v2  c2  0

or f   z  sin ce f   z   u x  iv x
2
0
i.e., f   z   0. Hence f  z   cons tan t.
Ex. 3.
If f(z) is an analytic function of z, prove that
 2 2 
 Rf  z   3 f   z  .
2 2
 2 
 x y 2 
Sol.
Let f(z) = u + iv, so R f(z) = u.
 2 u
u  2u .
We have x x
2
2 2  u  2 u
u  2    2u .   1
x 2  x  x 2
2
2 2  u  2 u
si milarly u  2    2u .   2
y 2  y  y 2

 2 2  2  u 2  u 2   2 u 2u 
 2   u  2  
      u  2  2
 x y 2   x   y    x y 
Hence

 u 2  u  2 
 2        sin ce u is a harmonic fucntion 
 x   y  
 u 2  v  2 
 2       sin ce u y   v x
 x   y  
u v
 2 f   z  . sin ce f   z  
2
i
x x

60
 2 2  2
u  2 f  z .
2
i.e.,  2  2 
 x y 
 2 2 
Rf  z   2 f   z  . sin ce Rf  z   u.
2 2
or  2  2 
 x y 
then  z  u  iv 

61
Aliter. Let f(z) = u + iv.
1
u
2
 f  z  f  z  .
So that
 2 2 
 Rf  z 
2
 2 
x y 2
Now  

 2 2  2 2 2
 2  2  u 4 u
 x y  z z
2
2 1
4  f  z  f  z 
z z 2
2
 f  z   f  z    f  z   f  z   sin ce f  z   f  z  f  z 
2

z z
2
 f  z  f  z 
2

z z

 .2  f  z   f  z   f   z 
z
 2f   z  f   z   2 f   z 
2 2
by formula z  zz
Ex. 4.
If f(z) is a regular function z prove that
 2 2 
 f  z  4 f  z .
2 2
 2 
 x y 2 
Sol.
Let the analytic function f(z) be
f  z   u 2  v2 ,
2

f(z) = u + iv, then


2
 2 u 2  u  2u
u  2u .  2 u 2  2    2u 2 .
We have x x x  x  x
2
2 2  u  2 u
u  2    2u .
Similarly y 2  y  y 2
On adding, we have

 2 2  2  u  2  u  2  2 u 2 u 
 2   u  2       2u  2  2 
 x y 2   x   y   x y  

62
 u 2  u 2 
 2       sin ce u is harmonic
 x   y  
 u 2  v 2 
 2        sin ce u y   v x
 x   x  
u v
 2 f   z  , sin cef   z  
2
i .
x x
 2 2  2
 v  2 f  z
2
 2 
Similarly  x y 2 

 2 2  2
 2  2   u  v   4 f  z
2 2

Hence  x y 

 2 2 
 f  z  4 f  z
2 2
 2 
Or  x y 2 
(Since f(z) = u + iv)
 2 2  2
  f  z
2 2
 2   f z  4
z z
Aliter  x y 2 

2 
4 f  z f  z   4  f  z f  z 
z z z

 4f   z 
z
 
f  z   4f  z  f   z 
2

Ex. 5.
If f(z) = u + iv is a regular function of z in any domain prove that
 2 2  p2
 f  z  p2 f  z  f  z .
p 2
 2 
i)  x y 2 

 2 2  p
 2  2  u  p  p  1 u f  z .
p2 2

ii)  x y 

Sol.
 2 2  2
 2  24 .
 x y  z z
We know that
 2 2 
 f  z
p
 2 
i) So  x y 2 

63
2
f  z .
p
4
zz
2
 f  z f  z 
p/2
4
zz
2 
 f  z   . f  z   
p/2 p/2
4
zz  
  p / 2  1
 4  f  z   .p / 2  f  z   f   z  
p/2

z  
 p / 2  1  p / 2  1
 4  p / 2  f  z   f   z   p / 2 f   z   f   z  
 
 p / 2  1
 p2  f  z  f  z   . f   z  f   z  

 
 p / 2 1 / 2
 p2 f  z  f  z  .
2 2

p2
 p2 f  z  f  z .
2

u  iv  f  z 
ii) Since u + iv = f(z) and ;
1
u 
2
 f  z  f  z  .
 2 2  p
 2  u
Now  x y 2 
p
2 2 1
4
z z
p
u 4
z z 2
 f  z  f  z 
4 2 2 p/ 2

2p z z
 f  z  f  z 

4 2
 f  z   f  z    f  z   f  z   
p/2
 p
2 z z
4 2
 f  z  f  z 
p
 p
2 z z
4 2
p f  z  f  z   f  z
p 1
 p
2 z
  4 / 2p  p  p  1  f  z   f  z  
p 2
f  z f  z 
 1/ 2   p  2 
  4 / 2p  p  p  1  f  z   f  z    f  z
2 2

 
 1/ 2   p  2 
  4 / 2p  p  p  1  f  z   f  z    f  z   f  z    f  z
2 2

 

64
 1/ 2   p  2 
  4 / 2p  p  p  1  f  z   f  z  
2
f  z
2

 
 1/ 2  p  2
1 2

 p  p  1   f  z   f  z    f  z
2

 2 
2  1/ 2   p  2 
 p  p  1  u  f  z 2
 
 p  p  1 u f  z .
p2 2

Ex. 6.
If w = f(z) is a regular function of z; prove that
 2 2 
 2  2  log f   z   0.
 x y 
f  z
If is the product of a function of x and function of y, show that

f   z   exp z   z   ,
2
where  is a real  and  are complex constants.
Sol.
We know that
 2 2  2
 2    4 .
 x y 2  z z

 2 2 
 2
 x

y 2 


log f   z  . 
Hence
2
4
z z

log f   z  
2  1 2
4  log f   z  
z z  2 
 2
2  log f   z  f   z   
z z 
2
2  log f   z   log f   z  
z z 
  f   z  
2    0,
z  f   z  

f  z  f   z 
Since and are independent of z.

65
Note .
f  z
When differentiating with regard to z, then is treated as constant
f  z
 f   z 
and when differentiating with regard to z, then and are treated as
constants.
Second Part.
f  z    x    y ,  x   y
Let where is a function of x alone and
is a function of y alone; and either both are positive or both negative.
 x   y
Assume and are both positive.
We have proved above that
 2 2 
 2
 x

y 2 

 log f   z   0 
 2 2 
 2   log    x    y    0,
or  x y2 

 x   y
since and are real
 2 2 
i.e.,  2  2  log   x   log   y    0,
 x y 
d2 d2
i.e., log   x   log   y   0.
dx 2 dy 2
 x   y
[since does not contain y and does not contain x]
 x   y
As and are independent of each other, therefore
d2 1
log   x   
dx 2
2 where  is a real constant,
d2 1
2
log   y    
and dy 2 because sum of the two has to be zero.

log   x   x 2  ax  c
Integrating these equations twice, we get and
  y   y  by  d
2
where  , a, b, c, d are real constants,
  x   exp  x  ax  c    y   exp  y  by  d 
2 2
So that and
 f  z   x    y

66

 exp   x 2  y 2   ax  by  c  d 
 exp    x 2
 y    ax  by  e 
where all constants are real.
Hence we may write

 
f   r   exp   x 2  y    ax  by  e exp i  2xy  ay  bx  f 

 exp    x 2
 y 2  2ixy   a  x  iy   ib  x  iy    e  if  

 exp   x  iy    a  ib   x  iy    e  if 
2

 exp  z 2   z    where  is real and   a  ib,   e  if
 exp  z 2   z    .

 where  is real and ,  are complex cons tan ts  .

67
NOTES

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

68
UNIT - III
CONFORMAL MAPPINGS
3.1 Introduction:
The relation w  f (z) when interpreted geometrically is called a
transformation or mapping. If W is a function of z, the points z in the z-plane
correspond to the points w in the w-plane. The assemblages of the points in a
certain plane forms a curve of a region. Hence by the above relation curve or a
region of z-plane is said to be transformed into or mapped upon or represented
by the corresponding curve or region in the w-plane. The corresponding points
of the two planes are called images of each other. Also we speak of the
corresponding called images of each other. Also we speak of the corresponding
called images of each other. Also we speak of the corresponding curves and
regions of the two planes as images of each other. If curves and regions of the
two planes as images of each other. If the function w  f (z) is not one valued
there are two or more values of w for some or all values of z, so that there is no
one-one correspondence between the corresponding points of the two planes.
To use the terms transformation or mapping or images for such
correspondences is not appropriate. To do justice to for such correspondences
is not appropriate. To do justice to the meaning of these words the
correspondence between z-points and w points must be one-one i.e. to each
value of z must correspond one and only one value of w.
Hence under suitable conditions the relation w  f (z) maps a region D
of the z-plane upon a region D’ of the w-plane and a curve C of the z-plane is
mapped upon a curve  of the w-plane.
Definition: Isogonal and conformal Transformation
If the two curves in the z-plane intersect at the point z 0 at an angle  ,
then if the two corresponding curves in the w-plane intersect at the point w 0
which corresponds to z0 at the same angle  , the transformation is said to be
Isogonal. Thus if only the magnitude of the angle is preserved the
transformation is Isogonal.
If the sense of the rotation as well as the magnitude of the angle is
preserved, the transformation is said to be conformal.
Note:
Some writer do not distinguish between isogonal and conformal. They
regard conformality as the preservation of the magnitude of the angles without
considering the sense.

69
3.2 If f(z) is analytic mapping is conformal.
(Necessary conditions of conformality).
If f(z) is a uniform analytic function of z in a region D of the z plane,
then the relation w  f (z) determine a conformal transformation, provided that
f ' (z 0 )  0, z 0 being an interior point of D.

Let C1 and C2 be the two continuous curves, in the z-plane, intersecting


at the point z and let the tangents at this point make angles 1 and  2 with the
0
real axis. Let z1 and z2 be the points on the curve C 1 and C2 near to z0 and at the
same distance r from z0, so we have
z1  z 0  rei1 ,
z 2  z 0  rei2

Then r  0, 1  1 and 2   2
Suppose that z0 corresponds to point w0 in the w-plane and z1 and z2
correspond to points w and w which describe the curves 1 and  2
1 2

Let w1  w 0  1ei1

w 2  w 0  2ei2
w1  w 0
f ' (z 0 )  lim
z1  z 0 z1  z 0

70
1ei1
or Rei  Lt '
rei1 since f (z 0 ) may be written as Re
i

1 i( 1 1 )
Rei  Lt e
r
 
Lim  1   R  f ' (z 0 )
Hence r  .
Lt  1  1    or Lim 1  Lim 1  
And

Or 1  1   i.e.1   2   s

Similarly it can be shown that 2   2   .

Thus we see that the curves 1 and  2 have definite tangents at w0


making angles 1   and  2   with the real axis and so angle between
1 and  2 at w is 1   2 i.e.  (1   )  ( 2   ) i.e 1   2 which is same as
0

the angle between C1 and C2 at z0. Hence the curves 1 and  2 cut at the same
angle as the curves C2 and C1, also the angle between 1 and  2 has the same
sense as angle between C1 and C2.
So the transformation is conformal.
We have seen that every regular function w  f (z) defined in a domain
'
in which f (z) is not zero maps conformally a region of z-plane upon a region
w-plane. We consider below the converse of it. In other words we discuss
converse of the last article.
3.3. Converse. If the mapping conformal, w  f (z) analytic
(Sufficient conditions of conformal mapping).
Being given the mapping of z-plane on w-plane conformal, to show that
the transformation is of the form w  f (z) where f(z) is regular.
Let u  u(x, y), v  v(x, y) be a pair of differentiable equations defining
conformal transformation from (x,y) plane to (u,v) plane.

Let d and ds be the length elements in the (u,v) plane and (x,y) plane
respectively.

Then ds  dx  dy .........(1)
2 2 2

and d2  du 2  dv 2
Since u and v are functions of x,y, therefore

71
u u v v
du  dx  dy, dv  dx  dy
x y x y
2 2
 u u   v v 
 du  dv   dx  dy    dx  
2 2

 x y   x y 

 u   v   2
2 2
 u 2  2  u u v v 
    
2
or d   dx  2  .  .  dxdy  
     dy .......(2)
 x    x y x y   y   y  
Given that the mapping is conformal, therefore the ratio d : ds is
independent of direction, so that from (1) and (2) on comparing coefficients,
we have
2 2
u u v v  u   v 
2 2
 u   v 
     .  .    
 x   x   x y x y   y   y 
1 0 1
From these, we have
2 2 2 2
 u   v   u   v 
        
 x   x   y   y 
u u v v
.  .
and x y x y =0

u x  v y , v x   u y, .......(i)
The first equation is satisfied if
And second is satisfied if
u x  u y, v x  u y ......(ii)
Equation (i) are the Cauchy-Riemann equations and express that w =
f(z), where w  u  iv and z  x  iy ,
Equations (ii) reduce to equation (i) by writing –v for v, that is by
taking as image figure found by reflection in the real axis of the w-plane. So
equations (ii) correspond to an isogonal, but not conformal transformation.
It follows that if the mapping of z-plane upon w-plane is conformal, the
only form of transformation is w  f (z) where f(z) is a regular function of z.

The case f (z 0 )  0, when f (z)  0,


' '

'
Suppose that f (z 0 ) has a zero of order (n-1) at the point z0 , then first
(n-1) derivatives of f(z) vanish at z but f (z 0 )  0, hence at any point z in the
n
0
neighbourhood of z0, we have by Taylor’s Theorem.
f ' (z)  f (z 0 )  a n (z  z 0 ) n  ......

72
f n (z 0 )
an 
where n! , so that a n  0 .

Hence f (z1 )  f (z 0 )  a n (z1  z 0 )  ......


n

i.e.w1  w 0  a n (z1  z 0 ) n  ....


or 1ei1  a n .r n ei n1    .......

where   amp a n
Lim 1  Lim  n1     n1  .
Hence

Similarly Lim 2  n 2  .

Thus the curves 1 and  2 still have tangents at w0


But the angle between the tangents.
 Lim 2  Lim 1  n   2  1 
So magnitude of the angle is not preserved.
R  Lim  1 / r   0
Also the linear magnification
Hence the conformal property does not hold good at a point where
f (z)  0.
'

R  f ' (z) and 1  1  .


3.5. Geometrical interpretation of
We have seen in 3.2 that
1 i( 1 1 )
Rei  Lim e
r
1
 Lim  R  f ' (z)
r
  Lim  1  1   Lim 1  Lim 1  1  1
and
i.e. 1  1   where   amp f ' (z 0 )
Thus to obtain the figure in the w-plane from the figure in the z-plane,
we rotate it through the angle   amp f (z 0 ) and subject it to a magnification
'

Lim  1   R  1f 2 (z 0 )
 r 

73
'
Since f (z 0 ) is unique, hence it is clear from above that the
magnification R is same in all the directions through the point z 0. Hence shape
of the image of a small figure near z0 is the same as that of the original figure.
If  is a regular function of w and w is a regular function of z, then  is
a regular function of z. Hence if a region of z-plane maps conformally on a
region of w-plane and this in its turn maps conformally on a region of  plane.
Then transformation from z-plane to  plane will be conformal.
3.6. Transformation which are isogonal but not conformal.
These exist transformation which are isogonal but not conformal, that
is, in which magnitude of the angles is conserved but their sign is changed. For
example, consider the transformation
w  x  iy

If z  x  iy , then w  x  iy is the reflection of z in the real axis, so


that the angles are conserved but their signs are changed.
In general this is true for every transformation of the form
 __ 
w fz  (1)
  where f(z) is regular; because such a transformation is
combination of the following transformation.
__
(i)   z (ii) w  f ( ).
In (i) angles are conserved but their signs are changed, in (ii) angle and
their signs conserved. Hence in the resultant transformation (1) angles are
conserved and their signs are changed.
So (1) gives transformations which are isogonal but not conformal.
3.7 Bilinear Transformation. (Mobius Transformation).
az  d
w
cz  b
az  b
w
The relation cz  d
where z, w are complex variables and a,b,c,d are complex constants associates
to a point z of the z-plane to a point w of the w-plane. The above equations
may be written as
cwz  dw  az  b  0
which is linear in w and as well in z; that is why the relation (1) is called
Bilinear transformation. It is also sometimes called linear transformation.
Billinear transformations are also called Mobius transformations after the name
of A.F. Mobius (1790-1868) how first studied them.

74
a zb/a
z . .
The relation (1) is equivalent to c zd/c
b d
 , i.e.if ad  bc  0
From this we notice that if a c then for every
value of z there is the same value of w, And if ad  bc  0 , then there
correspond different values of w to different values of z.
The expression ad-bc is called the determinant of the transformation.
3.8 Bilinear transform. To every value of z – corresponds only one value of
W and vice – versa.
az  b
w  (1)
The bilinear transformation cz  d is equivalent to
a zb/a
w .
c zd/c
It is quite obvious from the above transformation that for each value of
z  d / c there corresponds a value of w.
dw  d d w b/d
z or z   .
cw  a c w a/c
It is quite apparent from this relation that to each value of w  a / c
there corresponds a value of z.
Thus the bilinear transformation (1) associated each z  d / c to a
value of w, and each w  a / c corresponds to a value z, and the
correspondence between w and z is one-one.
Geometrically speaking, the bilinear transformation maps one-one the
entire z-plane, with the exception of the z  a / c ,on the entire w – plane with
the exception of the point w = a/c.
However these points will not remain exceptions it we adjoint a new
point called point at infinity to the complex plane, and the complex plane is
then called the extended complex plane. As usual the point at infinity is
denoted by the symbol  , with this definition of the extended complex plane
the critical point z   corresponds to the point w  a / c .
Thus the bilinear transformation maps one-one the entire extended
complex z-plane upon the entire extended complex w-plane.
3.9 Every bilinear transformation is the resultant of bilinear
transformation with simple geometric imports.
az  b
w  (1)
Let a general bilinear transformation be cz  d
c0
where ad  bc  0 and

75
a bc  ad 1
w  .
c c cz  d
It may be written as
a bc  ad 1
w  .
c c 2
zd/c
Or
This transformation can be effected by combining the three
transformations.
d 1 bc  ad
z1  z  , z 2   z3  z2
c z1 c2

w  a / c  z3
So that which can be effected in the same way as
z1  d / c  z
is effected.
It is clear that the above three auxiliary transformations are of the form
w  z  , w  1/ z, w  z.

Showing that every generally bilinear transformation is the resultant of


bilinear transformations of the form.
w  z  , w  1/ z, w   z.

In the following article we give the geometrical interpretations of these


transformations.
Geometrical interpretations of the transformation
w  z  , w  1/ z, w  z.
w  z
I. (Translation)

Let the numbers z and be represented by the points P and A, Where
A is fixed point and P is variable point.

76

OP 
Since z is represented by the vector and is represented by the

OA z 
vector therefore completing the parallelogram AOPQ, the number is
  
OP OP OQ
represented by the resultant of the two vectors and i.e., . In other
words the point Q represents the number w.

 AO
Since the point is fixed, so the vector is fixed. Also here the

PQ 
vector represents the number (since PQ is equal and parallel by OA)
Thus Q is obtained from P by moving P a distance equal and parallel to OA.
In other words the image Q of any point P is obtained by subjecting the

plane to a translation defined by the number .
w z
II . Magnification and rotation

Let the points P and B represent the numbers z and where z is

variable and is fixed. Then the point Q which represents w is such that
QOP  AOB  are  .
The angle

OQ  OB.OP   OP
And the length
Thus Q is obtained from P by rotating OP through an angle equal to the
 
arg and magnifying OP in the ratio .
w z
Thus we see that the transformation corresponds to a rotation
together with a magnification.

77
 
There will be only rotation if is unimodular (i.e. if modulus of is
 
1.), and only magnification if arg is zero, i.e. if lies on the real axis and
is positive.
w  1/ z
III If . Inversion in the real axis and in the unit circle.
w  1/ z
The transformation is the resultant of two transformations.
__ __
z w  1/ z
and .

The former is the inversion (reflection) in the real axis and the latter is
w  1/ z
the inversion (reflection) in the unit circle. Then the transformation is
the resultant of inversion in the real axis together with inversion in the unit
circle.

3.10 Theorem
 z  z1 
arg     ( real cons tan t)
 z  z2 
The equation represents a family of
circles every members of which passes through two fixed points z1, z2
We know that
 z  z1 
arg  
 z  z2 

is the angle between the two lines joining z1 to z, and z2 to z.


 z  z1 
arg  
 z  z2 
Hence the equation

78

Means that the line joining the points z 1 , z2 subtends an angle at the

point z. if z1 , z2 are fixed points, a real constant and z a movable point then
by plane geometry locus of z is a circle passing through the two fixed points z 1,
z2 (angles between same segment being equal).
3.10 Theorem.
The bilinear transformation
az  b
w
cz  d
Transforms the circle
 z  z1 
arg  
 z  z2 

w  w1
w  w2
into similar circle arg = constant where w1, w2 correspond to z1, z2
respectively.
Sol.
az  b
w
cz  d
We have
Since w1, w2 correspond where to z1,z2 respectively.
az1  b az  b
 w1  and w 2  2
cz1  d cz 2  d
az  b az1  b

w  w1 cz  d cz1  d
so that 
w  w 2 az  b  az 2  b
cz  d cz 2  d

cz 2  d z  z1
 .
cz1  d z  z 2
z  z1 cz  d
 where   2 is complex
z  z2 cz1  d
 w  w1   z  z1 
 arg    arg   
 w  w2   z  z2 
 z  z1 
 arg   arg  
 z  z2 
 arg   .

79
 w  w1 
 k
 w  w2 
Thus arg where k is real which is a circle in w plane
passing through two fixed points w1, w2 which are the images of z1, z2.

3.12 Cross-ratio.
z1 , z 2 , z 3 , z 4
taken in order, then the ratio
If there are four points
 z1  z 2   z3  z 4 
 z 2  z 3   z 4  z1 
z1 , z 2 , z 3 , z 4
is called the cross-ratio of .
3.13 Preservance of cross-ration under the bilinear transformation.
w1 , w 2 , w 3 , w 4
If under the bilinear transformation are the images of
z1 , z 2 , z 3 , z 4 w1 , w 2 , w 3 , w 4
respectively, to show that cross ration of is equal to
z1 , z 2 , z 3 , z 4  w1 , w 2 , w 3 , w 4    z1 , z 2 , z3 , z 4 
the cross ratio of to show that .
az  b
w .
cz  d
Let the bilinear transformation be
w1 , w 2 , w 3 , w 4 z1 , z 2 , z 3 , z 4
Since are the images of respectively, therefore
az1  b az  b
w1  , w2  2
cz1  d cz 2  d

az1  b az 2  b  ad  bc   z1  z 2 
 w1  w 2   
cz1  d cz 2  d  cz1  d   cz 2  d 

w2  w3 
 ad  bc   z3  z 4 
 cz3  d   cz 4  d 
Similarly

w3  w 4 
 ad  bc   z3  z 4 
 cz 4  d   cz 4  d 

w 4  w1 
 ad  bc   z 4  z1 
 cz 4  d   cz1  d 
and
From the last four results, we have

80
 w1  w 2   w 3  w 4    z1  z 2   z3  z 4 
 w 2  w 3   w 4  w1   z 2  z 3   z 4  z1 
 w1 , w 2 , w 3 , w 4    z1 , z 2 , z3 , z 4  .
Briefly
3.14 Find the bilinear transformation which transforms the points z 1, z2, z3
of z plane respectively into the points w1,w2, w3 of w-plane.
Let the bilinear transformation be
az  b
w   1
cz  d
Since w1,w2, w3 are the images of z1, z2, z3 respectively, there fore
az1  b
w1    2
cz1  d
az 2  b
w2    3
cz 2  d
az 3  b
w3    4
cz 3  d

Substituting (2) from (1), we have


az  b az1  b  ad  bc   z  z1 
w  w1   
cz  d cz1  d  cz  d   cz1  d 

w1  w 2 
 ad  bc   z1  z 2 
 cz1  d   cz 2  d 
Similarly

w2  w3 
 ad  bc   z 2  z3 
 cz 2  d   cz3  d 

w2  w 
 ad  bc   z 2  z3 
 cz3  d   cz  d 
and


 w  w1   w 2  w 3    z  z1   z 2  z 3 
 w1  w 2   w 3  w   z1  z 2   z3  z 
This is the required transformation.
If we solve it for w it is of the standard form of the bilinear
transformation, ie., it is of the form

81
z  
w .
z  

Ex. 1.
z
Find the bilinear transformation which maps the points , i, 0 into

points w = 0, i, respectively.

Sol.
We know that the bilinear transformation, mapping
z  z1 , z 2 , z 3 ..... int o w  w 1 , w 2 , w 3....
respectively, is
 w  w1   w 2  w 3    z  z1   z2  z3  .
 w1  w 2   w 3  w   z1  z 2   z3  z 
Substituting in this, we have
 w  0   i  w 3    z  z1   i  0  where z1  , w 3  
 0  i   w 3  w   z1  i   0  z 
 i  z 
w  1   1 i
or  w 3    z1 
 w   i 
i 1   1    z 
 w 3   z1 

w i
or  when z1  , w 3  
i z
or w  1/ z.

This is the required transformation.


Ex.2
Find the bilinear transformation which transforms the points z = 2, 1, 0,
into w = 1,0,i.
Sol:
z  z1 , z2 , z3
We know that the bilinear transformation which transforms
w  w1 , w2 , w3
respectively into is
 w  w1   w2  w3    z  z1   z2  z3 
 w1  w2   w3  w  z1  z2   z3  z 

82
Substituting in this equation we have
 w  1  0  i    z  2   1  0 
 1  0  i  w  2  i   0  z 
i  iw 2  z
or 
iw z
or  i  iw  z   i  w   2  z 
{( 1  i ) z  2}w  2iz  2i

or w
 2iz  2i 
 1 i z  2
This is the required transformation.
3.15 Bilinear transformation of a circle.
az  b
w
cz  d
To show that bilinear transformation transforms a circle of
z-plane into a circle of w-plane and inverse points transform into inverse
points. In the particular case in which the circle in the z plane transform into a
straight line in the w-plane, the inverse points transform into points
symmetrical about this line.
z p
k
zq
We know that represents a circle in the z-plane with inverse
points p,q. If k=1 the equations represent a line which is the right bisector of
the join of the point p,q.
z p
k
zq
Now, let be a circle in the z-plane, whose inverse points are
p,q.
az  b
w  (1)
cz  d
The given transformation is
From this we see that the points p and q in the z-plane correspond
ap  b aq  b
cp  d cq  d
respectively to the points and in the z-plane.
dw  b
z
cw  a
Also we see that
Substituting this value of z in (1) it transforms into

83
ap  b
w
cp  b cq  d
k (2)
aq  a cp  d
w
cq  d

The form of this equation shows that it represents a circle in, the w-
plane, whose inverse points are
ap  b aq  b
and
cp  d cq  d

Thus a circle in the z-plane transforms into a circle in w-plane, and the
inverse points transform into the inverse points.
cq  d
cp  d
Particularly, If =1, then equation (2) represents a line bisecting
ap  b aq  b
cp  d cq  d
at right angles the join of the points and so that these points are
symmetrical about the straight line.
Thus in the particular case a circle in the z plane transforms into a
straight line in the w-plane, and the inverse points transform into points
symmetrical about this line.
Ex. 1.
az  b
w
cz  d
Find the condition that the transformation transforms the
unit circle in the w-plane into straight line in the z-plane.
Sol:
The given transformation is
az  b a z  b / c
w  .
cz  d c z  d / c

w 1
Clearly, under this transformation the unit circle , in w-plane
transforms into
b
z
a a in z  plane
1
c z d
c

84
b
z
a  c
d a
z
c
Or
c
1
a a c
This represents a line if or
b
z
a 1
d
z
c
The corresponding line is then
Note.
z p
k
zq
It is to be recalled that the equation represents a line or a
circle according as k is or is not equal to 1.
Ex.2
I (z)  0
Find all the Mobius transformations which transform the plane
w  1.
into the unit circle

The Mobius transformation is


az  b a z  b / a
w  .
cz  d c z  d / c

I ( z)  0 w 1
This transforms into ,

85
i.e., real axis in z-plane transform into unit circle in w-plane. Therefore
1 __
w
points w, inverse with respect to the unit circle transform respectively
_
z
into the points z, inverse with respect to the real axis in z-plane.
__

  ,
Particularly the points w=0, correspond to the points .
b d __
 , 
a c
Hence from (1) we have and then (1) reduces to
a z 
w . __
 (2)
c z 

w 1
The point z=0 corresponds to the point therein we have

a  a 
1 __
or 1  __
c  c 
a __
or 1  sin ce   
c

a
 e i
c 
Or where is real.
z 
w __
e i
z 
Thence (2) reduces to
I (z)  0 w 1
This is the transformation which maps into Further,
under this transformation, we have

86
__ __
z  i z 
ww  1  __
e , __
e  i  1
z 
z 

 z     z      
__ __ __ __

 z  z     
2
or w  1    1    
 __
 __
 __ 2

 z    z    z 
  
2iI ( z ).2i I ( ) 4 I ( z ) I ( )
 
__ 2 __ 2
z  z 

 I ( )  0
Since w=0 corresponds to , therefore
2
w 1  0 I ( z)  0 w 1 I ( z)  0
Hence for or corresponds to
z 
w __
ei
z 
Hence is the required transformation.
Ex. 3
Find the bilinear transformation which transforms the half plane
Re( z )  0 w 1
into the unit circle .
az  b a z  b / a
w  .  (1)
cz  d c z  d / c
The proposed transformation is
Re( z )  0 w 1
This transforms into , i.e. imaginary axis in z-plane
transforms into the unit circle in w-plane.
1
w, __
w
Hence the points inverse with respect to the unit circle in w-plane
__
z,  z
transform into the points inverse with repect to the imaginary axis in z-
plane.
__
w  0 ,  , 
Particularly, the points correspond to the points .
b d __
 ,  
a c
Hence from (1) it follows that

87
a z 
w . __
 (2)
c z 
So that (1) reduces to
w 1
The point z=0 corresponds to the point ; hence taking mod of (2)
w 1
and substituting and z=0 therein we have
a 
1
c 
a __
or 1  sin ce   
c

a
or  e i
c 
where is real.
z 
w __
e i
z 
Thence (2) reduces to
w 1 Re( z )
This is the transformation which maps into . Further,
 Re( )  0
under this transformation w=0 corresponds to , therefore , and
__ __
__
z  i z 
w w 1  __
e , __
e  i  1
z  z 

 z     z   
__ __

2
or w  1    1
 __
  __ __ 
 z   z  
  
 __
 __

 z  z     
    
__ 2
z 

2 Re( z ).2 Re( )



__ 2
z 

Re( )  0
where
2
w 1  0 Re( z )  0
Hence for

88
w 1 Re( z )  0
Or corresponds to
z 
w __
e i
z 
Hence is the required transformation.
Ex. 4.
z 
Find the bilinear transformation which transforms the circle into
w  '
the circle .
The transformation suggested, is
az  b a z  b / a
w  .  (1)
cz  d c z  d / c

z  w  '
This transforms the circle into the circle . hence the
' 2
w, __
w | w | '
points inverse with respect to the circle transform into the
2
z, __
z z 
points inverse with respect to the circle .
2
, __
w  0,  
Particularly correspond to the points respectively.
b d  2
 ,  __
a c 
Hence from (1) it follows that
a z 
w .
c 2
z  __

So that (1) reduces to
__
a z 
 . __
c z  2

zp w  p'
The point corresponds to the point ; hence taking mod of
wp '
z
(2) and substituting and therein we have

89
__
a
 
'  __
c  2
__
1 a  

 c   __

__
a
or  '  __
c p   p 

since , being real
__
a
  ' ei
c 
Or where is real.
z 
w   'ei __
z  2
Consequently (2) reduces to
z  w  '
This is the transformation which maps into .
Further, under this transformation the point w=0 corresponds the point
 z   
which is the internal point of , therefore .

__ __
__
' i z '  i z
And w w     e'2
__
. e __
 '2
z  2 z   2

 z     z   
__ __

2
or w  '2  2'2    '2
 __ 2  __ 2
 z    z    
  

  '2
  z    
2 2 2 2

__ 2

z  2

2
w   '2  0 z 
Clearly for
i.e. w   ' z 
corresponds to

90
z 
w   'ei __
z  2
Hence is the required transformation.

Ex.5

Find all the Mobius transformations which transform the unit circle
z 1 w 1
into unit circle

The mobius transformation is


b
z
az  b a a
w   (1)
cz  d c z  d
c

z 1 w 1
This transforms the circle into the circle
1
w, w 1
w
Hence the points inverse with respect to the unit circle
1
z, z 1
z
correspond to the points inverse with respect to the circle
1
,
w  0,  
Particularly, the points correspond to the points ; hence
from (1) it follows that
b b 1
 a,  __
a c 

So that (1) reduces to


a z
w .
c z 1

__
a z
 . __  (2)
c z  1

91
w 1
The point z=1 corresponds to the point , therefore taking mod of
w 1
(2) and substituting and z=1 we have
__
a
1
1  __
c  1
__
a 1
 __
c
1

__
a __
or 1 sin ce 1    1  
c

__
a
or  e i
c 
where is real.
z
w  ei  (3)
z  1
Thence (2) reduce to
z 1 w 1
This is the transformation which maps into .
Further, under this transformation
__ __
__
z z
w w  1  ei __
.e  i __
1
z  1 z  1

 z     z      z  1  z   1
__ __ __ __

2
or w  1      
2 __
z  1

 __
 __

 1  z z  1    
  
2
__
z  1

92

 1 z   1  
2 2

2
z  1  1
where
2
w 1  0 z 1
Hence for
w 1 z 1
Or corresponds to
z
w  e i __
z  1
Hence is the required transformation.
Ex.6.
za  R w 1
Show that the region is a mapped conformally on by
R  z  c
w ei
 __

__
R2   z  a  c  a 
 


where is real and z=c is the point which is transformed into the origin.
Let the transformation be

z
z    
w  .
z    z  

w 1 za  R
Here the circle corresponds to the circle .
__
w,1/ w w 1
Therefore the points inversed with respect to the circle
R2
a  __ __
z a za  R
correspond to the points inverse with respect to the circle .
R2
a  __ __
w  0,  c a
Particular correspond z=c, respectively. (because by
hypothesis w=0 and z=c correspond). Hence it follows from (1) that
  R2
  c and   a  __ __
  c a
so that (1) reduces to

93
 zc
w  .
  R2 
z   a  __ __ 
 
 c a 
 __ __ 
  c  a 
  cz
 .........(2)
  __ __ 
R  z a  c a 
2

 

w 1 z  a  R;
The point corresponds to hence taking mod of (2)
and substituting therein we have
 __ __ 
  c  a 
  z c
1 
  __ __ 
R2   z  a  c  a 
 
 __ __ 
  c  a 
  zc  __ __ 
 sin ce R 2
  z  a   z a 
  __ __   __ __   
 z  a  z  a   z a  c a 
   

 __ __ 
  c  a 
  zc


 z  c   z  c 
__ __

 
 __ __ 
  c  a 
  zc

  __ __ 
 z  a  z c 
 

 __ __ 
  c  a 
  1

 R
 __ __   __ __ 
  c  a    c  a 
     Re i ,  real
or  R or
 

94
zc
w  Rei
 __ __ 
R2   z  a  c  a 
 
Thence (2) reduces to
The is the desired transformation.
Below we show that, under this transformation, the interiors of the two
circles also correspond.
zc
w  Rei
 __ __ 
R2   z  a  c  a 
 
zc  __ __ 
 Rei sin ce R 2
  z  a   z a 
 __ __   __ __   
 z  a  z  a  
 z  a  c  a 
   
Re  z  c 
i


 z  a   z  c 
__ __

 

R zc
w  __ __
z a z c

R __ __
 sin ce z  a  z  c
za
 1 for z  a  R.

Ex.7
w 1
(a) Find a Mobius transformation which maps the circle into the
1
z 1  1 z , z0
w  0, w  1 2
circle and maps respectively into show also
that the transformation is uniquely determined.
Let the transformation be
b
z
az  b a a
w  .......(1)
cz  d c d
z
c

95
za  R
Remember that in case of the circle , ( the circle of radius
R2
a  __ __ .
z a
unity with centre at the point a) if one inverse point is z, other is
w 1 z 1  1
Here circle corresponds to .
1
w, __
w w 1
Therefore points inverse with respect to the circle
1
1  __
z 1 z 1  1
corresponds to the points z, inverse with respect to the circle
w  0, 
Particularly correspond to
1 1
z  , 1 ;
2 1
1
2
that is
(since given w=0 when z=1/2)
Hence from (1) it follows that
 b 1 d
 ,  1
a 2 c
So that (1) reduces to
1
z
a 2
w
c z 1
Since w=1 corresponds to z=0, therefore substituting in (2) we have
a a
1 or  2
2c c

Whence (2) becomes


1
z
w  2 2
z 1
2z  1
or w  
z 1

96
This the required transformation.
Ex. 7
z 1
Find the Mobius transformation which transforms the circle onto
w 1
and makes the point z =1,-1 correspond to w =1, -1 respectively.
Sol:
Let the transformation be
b
z
az  b a a
w 
cz  d c d
z
c

[Under this transformation |w| = 1 corresponds to |z| =1] Hence the points
1
w 1 z,
z
inverse with respect to the circle correspond to the points inverse
z 1
with respect to the circle .
1
, __
w  0,  
Particularly, the points corresponds to ; hence from (1) it
follows that
b d 1
 ,  __
a c 

Whence (1) reduces to


__
a z a z
w  ........(2)
c z 1 __
c z  1

w 1
Since the point z=1 corresponds to , therefore taking mod of (2)
w 1
and substituting and z=1 therein we have
__ __ __
a  1  a  1  a 
1  __
 __

c  1 c 1  c

97
__
a
 e i
c 
Or where is real
z
w  ei __
 (3)
z  1
Thence (2) reduces to
Also because w=1, -1 correspond to z=1,-1 respectively therefore
substituting in (3) we have
1  1  
1  ei __
and  1  ei __
 1   1
These two equations give
__
1  1  __
 ,   
1  1 
__

Hence (3) becomes


z
w  e i
z  1
This is the required transformation.
Ex. 8.
Prove that in the general in the bilinear transformation
az  b
w
cz  d
there are two values of z (invariant points) for which w=z, but there is only one
 a  d   4bc  0.
2

if Show that, if there are distinct invariant points p and q, the


transformation may be put in the form
w p zp
k ;
w q zq

and that if there is only one invariant point p, the transformation may be put in
1 1
  K.
wp zp
the form
Sol:
az  b
w
cz  d
The bilinear transformation is

98
az  b
z
cz  d
If w=z, then it gives
cz 2   d  a  z  b  0
i.e
Let p,q be the roots of this equation, then solving the above equation:
we have

a  d  { a  d   4bc}
2

p
2c

a  d  { a  d   4bc}
2

q
2c
and
Thus p and q as expressed above are the two invariant points of the
 a  d   4bc  0,
2

transformation. They reduce to one if, and then


ad
pq ........(1)
2c
when p and q are distinct points.
cp 2   d  a  p  b  0 i.e.b  pd  cp 2  ap
We have
cp 2   d  a  q  b  0 i.e.b  qd  cq 2  aq
and
az  b
 p a  cp z  b  pd
w  p cz  d
 
 
w  q az  b  q  a  cq  z  b  qd
cz  d
Now


 a  cp  z  cp2  ap from (2)
 a  cq  z  cq 2  aq

 a  cq   z  p 
 a  cq   z  q 
zp a  cp
k where k 
zq a  cq
w p zp
i.e. k
w q zq

And when there is only one invariant point say p. Then as shown above
in (1) and (2), we have

99
a d
p and b  pd  cp 2  ap
2c
1 1 cz  d
 
w  p az  b  p  a  cp  z  b  qd
cz  d
Now
cz  d
 from (2)
 a  cp  z  cp 2  ap
cz  d

 a  cp   z  p 
cz  a  2cp
 from (1)
 a  cp   z  p 

 a  cp   c  z  p 
 a  cp   z  p 
1 c
 
z  p a  cp
1 c
  K where K 
zp a  cp
1 1
i.e.  K
w p zp

Ex. 9.
az  b
w
cz  d
Prove that if and ad-bc=1 the linear magnification is
 cz  d  cz  d  1 c  0 
2

, Show that the circle is the complete locus of the


points in the neighbourhood of which lengths and areas are unaltered by
transformation, Prove that the lengths and areas within the circle are increased
and lengths and areas outside the circle are decreased in magnitude by the
transformation.
Sol.
az  b
w
cz  d
We have , which may be written as

100
a ad  bc 1
w   .
c c cz  d
a 1 1
  sin ce ad  bc  1
c c cz  d
dw 1
 
dz  cz  d  2

dw 1

dz  cz  d  2
So that , which gives linear magnification. If
dw
1
cz  d  1, dz
then obviously
This shows that the linear magnification is constant for the points on the
cz  d  1,
circle Hence in the neighbourhood of the points on the circle
cz  d  1,
the lengths and areas are altered by the transformation.
a 1 1 1
w  . i.e.cw  a  
c c cz  d cz  d
1
or cw  a 
cz  d

cw  a  or  1 cz  d  or 
Now according as 1, i.e. inside and
cz  d  1
outside of the circle are respectively mapped upon the outside and
the inside of the circle.
cw  a  1

dw 1
 2
dz cz  d
We have
dw dw
1 1
dz dz
So for the points within the circle |cz+d|=1 and for the
point inside the circle.
Hence by the given transformation, the lengths and areas within the
cz  d  1
circle are increased and length and areas outside this circle are
decreased.

101
Ex.10
2z  3
w
z4
Show that the transformation changes the circle
x  y  4x  0
2 2

and explain why the curve obtained is not a straight line.


Solution.
2z  3
w  (1)
z4
The given transformation is
4w  3
z  (2)
w2
Solving for z, we have
__
__
4 w 3
z __
w 2
So that
x 2  y 2  4x  0
The circle
i.e.  x  iy   x  iy   4x  0
__
 __

or z z  2 z  z   0
 
Substituting in this from (2), we have
__
 __

4w  3 4 w  3  4w  2 4 w 5 
. __ 2  __ 0
w2 w  w2  __
2  w 2  2x  z  z
since .
 __

2 w  w   3  0
 
Or
__
4u  3  0 w  w  2u
Or since
which is a straight line to the w-plane.
Aliter:
2z  3
w
z4
Here

102
2x  2iy  3
or u  iv 
x  iy  4
{ 2x  3  2iy}{(x  4)  iy

{(x  4)  iy}{(x  4)  iy}

to make the denominator real.


 2x 2
 2y 2  5x  12   i11y
 x  4
2
 y2

the denominator is now real


2x 2  2y 2  5x  12
u ...........(1)
 x  4
2
 y2
i11y
v ......................(2)
 x  4
2
 y2

x 2  4x  y 2  0 or x 2  y 2  4x
The z-curve is given as
Substituting from this in (1) we have
3x  12 3 x  4 3
u  
4x  16 4  x  4 4

4u  3  0
Or this is the corresponding w-curve.
Ex.11.
w 1
(a) Show that both the transformations transform into the half
Re(z)  0
plane .
1 z z 1
w and w 
1 z z 1
Solution .
1 z
w
1 z
If then

103
__
__
1 z 1 z
w w1  1
1  z 1  __
z

 z  1  z  1
__

   1
 __ 
 z  1  z  1
 

z 1
w
z 1
which is same if

 z  1  z  1   z  1  z  1
__ __

    
2
z 1
 __

2 z  z 
    4 Re(z)
2 2
z 1 z 1

2 4 Re(z)
w 1  2
z 1
Or for both the transformations.
Hence under both transformations
2
w 1  0 Re(z)  0
according as .
w 1
This implies that the circle corresponds to the imaginary axis of
z-plane and interior of the circle is transformed into the left half z-plane.

Ex. 11.
(b) Show that both the transformations
1 z z 1
w and w 
1 z z 1

w 1 Re(z)  0
Transform into the half plane
Sol:
1 z
w
1 z
If then

104
__
__
1 z 1 z
w w1  . __  1
1 z
1 z
 __ 
 z  1  z  1
   1
 z  1  z  1
__

 

z 1
w
z 1
which is same if

 z  1  z  1   z  1  z  1
__ __

    
2
z 1
 __

2  z  z 
  2 
z 1

4 Re(z)
 2
z 1
the denominator being positive
2 4 Re(z)
w 1  2
z 1
Or for both the transformations
2
w 1  0 Re(z)  0
Hence under both transformations. according as
w 1
This implies that the circle corresponds to the axis of
imaginaries in z-plane, and interior of the circle corresponds to the right half z-
plane.

Ex.12.
Show that both the transformations
z i iz
w and w 
zi iz

105
w 1 I(z)  0
transform into the upper half plane
Sol.
zi
w
zi
If then
__
__
z i zi
w w1  . 1
z  i __z  i
__
i  z i  z
 . 1
i  z i  __
z
i 1
w
iz
Same as if

 z  i   z  i    z  i   z  i 
__ __

    
 z  i   z  i 
__

 
 __

2i  z  z 
  2 
zi

4I(z)
 2
zi
the denominator being positive
2 4I(z)
w 1  2
zi
Or for both the transformations.
2
w 1  0 I(z)  0
Hence for both the transformations according as to .
w 1
This means that the circle corresponds to the real axis and the
region interior to this circle corresponding to the upper half z-plane.
Ex.12
zi iz
w w
z i iz
(b) Show that both the transformations and
w 1 I(z)  0.
transform into the lower half plane

106
Sol.
zi
w
zi
If then
__
__
zi zi
w w1  .
z  i __
zi
__
i  z i  z iz
 . w
i  z i  __
z iz
same if

 z  i   z  i    z  i   z  i 
__ __

    
 z  i   z  i 
__

 
 __

2i  z  z 
  
2
zi

4I(z)
 2
zi
the denominator being positive
2 4I(z)
w 1  2
z i
Or for both the transformations.
2
w 1  0 I(z)  0
Hence under both transformations according as
w 1
This means that the circle transforms into the real axis, and
interior of this circle transforms into the lower half z-plane.
Ex. 13
1 z
w i
1 z
In the transformation show that the interior of the circle
z 1
is represented in the w-plane by the plane above the real axis, the upper
semi-circle into positive half of real axis and lower semi-circle into negative
half of the real axis.
Sol:

107
1 z
w i
1 z
The given transformation is

1  x  iy
u  iv  i
1  x  iy
{y  i  1  x  }{ 1  x   iy}

1 x   y
2 2

Equating real and imaginary parts, we have


2y x 2  y2  1
u .....(1) v ......(2)
1 x  y  1  x   y2
2 2 2

From (2), we see that when v=0.


x 2  y 2  1 i.e. z  1
We have
z 1
This shows that the unit circle in the z-plane corresponding to
the real axis in the w-plane . From (1) it is seen that y is positive when u is
positive, so upper semi-circle in the z-plane corresponds to the positive half of
the real axis on the w-plane, also it is seen from (1) that y is negative, when u is
negative, so lower half of the semi-circle in the z-plane corresponds to the
negative half of the real axis in the w-plane.
Next we proceed to show the correspondence between the regions of the
two planes.
x 2  y2  1 x 2  y2  1
is negative, i.e. if
z 1
i.e. If .

108
z 1
Thus interior of the circle in the z-plane corresponds to the half
of the w-plane above the real axis.
Ex. 14
iw
z
iw
In the transformation , show that the positive half of the w-
v0 z 1
plane given by corresponds to the circle in the z-plane.

Sol:
iw
z
iw
The given transformation is
i   u  iv  u  i  1  v 
z z
i   u  iv  u  i  1 v
i.e. or

u  i  1  v  {u 2   1  v  }
2

z  
u  i  1 u {u 2   1  v  }
2

{u 2   1  v  }
2

i.e. z   (1)
{u 2   1  v  }
2

z 1
Clearly when v=0
z 1
Thus real axis in the w-plane corresponds to the unit circle in the
z-plane.
z 1 0
And when v is (positive)

109
Hence the region upper half of the w-plane corresponds to the region
z 1
interior to the circle in the z-plane.
Ex. 15.
iz  1
w
zi
Discuss the application of the transformer to the areas in the
z-plane which are respectively inside and outside the unit circle with its centre
at the origin.
Sol:
iz  1
w
zi
The given transformation is
i  x  iy   1 1  y  ix
u  iv  
x  iy  i x  i 1 y
i.e.
{ 1  y   ix}{x  i  1  y  }

x2   1 y
2

Equating real imaginary parts, we have


2x x 2  y2  1
u ......(1), v .....(2)
x2  1  y x2   1  y
2 2

From (2) it is seen that


x 2  y2  1  0
when v=0
x 2  y2  1
corresponds to v=0.

110
z 1
Thus the unit circle in the z-plane corresponds to the real axis in
the w-plane.
Also from (1), we see that x is positive, when u is ositive; this shows
that the semi-circular arc to the right of the imaginary axis in the z-plane
corresponds to the positive half of the real axis in the w-plane. Similarly the
semi-circular arc to the left of the imaginary axis in the z-plane corresponds to
the negative half of the real axis in the w-plane.
Next, we proceed to establish the correspondence between the regions
of the planes.
x 2  y2 1
From (2) it is seen that is positive or negative according as v
is positive or negative.
i.e. x 2  y2  1  or  0
according as v > or <0
i.e. x 2  y 2  1 or  1
according as v> or <0
i.e. z  1 or  1
according as v > or < 0
z 1
This shows that the exterior and interior regions of the circle in
the z-plane correspond separately to the upper and lower half of the w-plane
which are above and below the real axis.
Ex.16.

 a b
If is real a,b, are complex numbers such that show that the
az  b
w  ei __ __
a b z
bilinear Transformation

111
z 1 w 1
Maps the inside of the circle on the inside of the circle
Sol:
The given transformation is
az  b
w  ei __ __
  1
a b z

az  b az  b
 ww  1  . 1
__ __
a bz a  bz

 az  b   az  b    a  bz   a  bz 

 a  bz   a  bz 

 aa  ba   zz  1
2
a  bz


a 2
b
2
  z  1 2

2
a  bz

Clearly Ww  1  0 if z  1because a  b

w  1corresponds to z  1.
or
Hence the interiors of the two circles correspond.
Ex. 17.
13i z  75 wa za
w k
3z  5i wb zb
Express the relation in the form where
a,b,k are constant. Show that the circle in the z-place whose centre is z= 0 and
radius 6 is transformed into the circle in the w-plane on the line joining w =a
and w=b as diameter, and the points in the z=place which are exterior to the
former circle are transformed into points in the w-place within the latter circle.
Sol.
13i z  75
w
3z  5i
We have
w  a 13iz  75  a  3z  5i 

w  b 13iz  75  b  3z  5i 
Let

112
 13i  3a  z  75  5ai

 13i  3b  z  75  5bi
75  5ai
z
w  a 13i  3a 3a  13i
i.e.,  .   1
w  b 13i  3b 75  5bi
z
3b  13i

75  5ai
 a,
3a  13i
If we want to make
3a 2  18ai  75  0 or a  3i  4
Then
75  5bi
 b, we get b  3i  4
3b  13i
Similarly if we want to make
a  b,
When then we have
a  3i  4, b  3i  4  a  b  .

13i  3a 13i  3  3i  4 
k 
13i  3b 13i  3  3i  4 
Let substituting for a and b,
i  3 4  3i
 
i3 5

 4  3i  1  3 
k  arg    tan   .
 5   4
So that arg
With these values of a,b, k the relation (1) becomes
w a za
k
wb zb
w a  za 
arg  arg  k 
wb  zb
Or
za
 arg k  arg
zb
 arg k  arg  z  a   arg  z  b    2

z  5, z  5ei
If the point z is on the circle then

113
z  5ei
Hence for a point the relation (2) becomes
w a
arg  arg k  arg  5ei  a   arg  5ei  b 
wb

arg  5ei  (3i  4)  arg  5ei  (3i  4)


= arg k +

 arg k  arg   5cos   4    5sin   3   arg   5cos   4    5sin   3  


3  5sin   3  1  5sin   3 
 tan 1    tan 1    tan  
4  5cos   4   5cos   4 
3 4

1  3  1 4 1 4 3  tan 1   1 .
 tan    tan  tan
4 3 3 4 2
1 .
4 3

w a 1
i.e., arg  .
wb 2
which is a circle n w-plane with join of the points a,b as diameter. This proves
the first part of the question.
z  5,
Next, we proceed to show that outside of the circle corresponds
w a 1
arg  .
wb 2
to the interior of the circle
z  5rei
Any point outside the circle |z| = 5 is where r > 1.
For such a point z we have
za
arg  arg  z  a   arg  z  b 
zb

 arg  5rei   3i  4    arg  5re i   3i  4  


 arg   5r cos   4   i  5r sin   3    arg   5r cos   4   i  5r sin   3  

114
 5r sin   3  1  5r sin   3 
 tan 1    tan  
 5r cos   4   5r cos   4 
 8  3  5r sin   
 tan 1  2 
 25r  7  30r sin  
8  3  5r sin  
 tan 1 sin ce r  1
25  7  30r sin 
8  3  5r sin   4
 tan 1  tan 1 ,
6  3  5r sin   3

wa 4
i.e., arg  tan 1
wb 3
wa 3 4 1
arg  tan 1  tan 1 ,  tan 1   ,
wb 4 3 2
Hence
wa 1
i.e., arg  .
wb 2
w a 1
arg  .
wb 2
This gives points w, which are inside the circle
z 5
Hence point exterior to the corresponds to the points interior to
w a 1
arg  .
wb 2
the circle
Ex. 18.
iz  2
w
4z  i
Show that the relation transforms the real axis in the z-plane
into a circle in the w-place. Find the centre and the radius of the circle and the
point in the z-plane which is mapped on the centre of the circle.
Sol.
iz  2
w   1
4z  i
The given transformation is
2  iw
z   2
4w  i
Solving for z, we have
2  iw 2  iw 4w  i
z  .
4w  i 4w  i 4w  i
Now ,

115
[To make the denominator real multiply above and below by
 4w  i 
]
8w  w  4iww  2i
x  iy  .
16ww  4  w  w  / i  1
Thus
The real axis in the z plane is y = 0. So corresponding to y = 0, we must
have imaginary part on the right side of the above relation also equal to zero,
that is
8v  v  4ww  2  0,

i.e.,  8v  v  4  u 2  v 2   2  0,
i.e., 4  u 2  v 2   7v  2  0

which is a circle in w-plane


7 1
i.e., u 2  v 2  v   0
4 2
Hence the real axis in the z=plane transforms into a circle in the w-
plane.

 7 4 9 1 9
 0,      .
 8 6 4 2 8
The centre of the circle is and radius =
7
w 0  u 0  iv 0   i
8
If we denote the centre by w6, then
If z0 corresponds to w0, then from (2), we have
7
2
2  iw 0 8  1 i.
z0  
4w 0  i 7 4
 ii
2
Ex .19
5  4z
w z  1,
4z  2
Show that the relation transforms the circle into a
circle of radius unity in the w-plane and find the centre of the circle.
Sol.
5  4z
w
4z  2
The given transformation is

116
2w  5
z .
4w  4
Solving this for z, we have
2w  5 2w  5
zz  .
4w  4 4w  4
So
4ww  10  w  w   25
 where w  u  iv
16  ww   w  w   1

2 4  u 2  v 2   20u  25
z 
16  u 2  v2  2u  1
or
sin ce ww  u 2  v 2 and w  w  2u 

which is real since u and v are real.


z 1
Hence transforms into
4  u 2  v 2   20u  25  16  u 2  v 2  2u  1

3
i.e., 4u 2  4v 2  4u  3  0, or u 2  v 2  u  0
4
which represents a circle in w-plane.

 1  1 3
  , 0 .     1.
 2  4 4
The centre of this circle is and radius
Ex. 20.
Find the radius and the centre of the circle in the w-plane which
ze  i
w ,
z  ie
corresponds to the real axis in the z-plane, where being a real
constant.
Sol.
ze  i
w
z  ie
The given transformation is ,
we   i we   1 x  e
zi  i .
w  e w  e w  e
Or

117
[To make the denominator real we multiply above and below by
we  

]
wwe   we 2   w  e 
i.e., x  iy  i where w  u  iv
ww   w  w  e   e 2 

(the denominator is real)


Equating imaginary parts
__
w w e   ue 2  u  e
y  __
 __

w w   w  w  e  e 2 
 

The real axis in the z-plane is the line y=0.


Clearly y=0 corresponds to
__
w w e  ue 2   u  e   0
__
i.e. u 2  v 2  ue  ue   1  0 sin ce w w  u 2  v 2
or u 2  v 2  u  e   e    1  0
i.e. u 2  v 2  2u cosh   1  0

which represents a circle in the w-plane.


 cosh ,0 
The centre of this circle is and radius=
cosh 2   1  sinh .

Hence the real axis in the z-plane maps upon a circle whose radius is
sinh   cosh ,0 
and centre is .
Ex. 21
az  b
 a  b
2
 4bc  0, w
cb  d
Show that if then for the transformation
w  z
k
,  w  z 
there are two unequal numbers such that where k is
constant.

118
Show also that the radius of the circle in the w-plane corresponding to
the circle in the z-plane which has a s diameter the line joining the points
 
z  , z   2 cos  
is , where is the argument of k.
Sol.
We proceed below for the second part.
Let z be point on the circle whose diameter is the line joining the points
z  , z  
.
 
Since the angles between the line joining z to and is a right angle,
therefore
z 1
arg  
z  2

z z
z  z 
So that is purely imaginary, or in other words real part of
is zero.
 __ __

1 z   z  
i.e.  0  (1)
2  z   __ __
z  

1 __

z  z 
2 
[since real part of z is ]
The given transformation is
__ __ __ __
z 1 w 
w  z z 1 w   __ __
k  __ __
w  z  z  k w  z   k w 
or , so that
__ __
1 w   1 w 
 __ __  0.
k w  k w 
Hence (1) transforms into
__
 __ __   __ __ 
w  w     w    k  w    w    0
   
__
__
 e i   e  i e  i  ei e  i   ei
or w w  w w 0
2 cos  2 cos  3cos 

119
__ __ __
w w  B w  B w  C  0.
which is of the form which is a circle in the w-plane
 BB  C 
whose radius is equal to .
Thus radius of this circle (of w-plane)
1/ 2
1  __ i __  i   __ i __ i  i  i 
 
2 cos  
 e   e 

  e  i
  e i
    e    e   e  e  
  

1  __ __  __ __  
           
2 cos    
1
 (  )  ((  )) 
2 cos   
1 2
  
2 cos 
     
 
2 cos  2 cos 

Ex. 22
__ __
a wz  bw  b z  a  0
Prove that the transformation maps the circle
z 1 w 1 b  a
on the circle if
Find the condition that the interior of the first circle may be mapped on
the interior of the second. Show also that in this transformation, the fixed points
either are inverse points with respect to the unit circle or lie on that circle.
Sol:
__ __
a wz  bw  b z  a  0
The given transformation is
bw  a
z __
a wb
Or
__ __ __
__
bw b w a
z z 1  __
. __ __  1
a w  b a w b
Now

120
  __ __ __   __ __

z  1   bw  a   a w  a    a w  b  
2

or     
 __ __
 __ 
 a w  b  a w  b 
  

 __ __  __  __ __  
 a a  b b  w w   a a  b b  
    
__ __ 2
a wb

 __ __  __ 
 a a  b b  w w  1
  
__ __ 2
awb


a 2
b
2
  w  1
2

__ __ 2
awb

2
z 1 w 1 z 1  0 w 1
Clearly corresponds to , And if ,
i.e. z  1 w 1
corresponds to
Hence the interiors of the two circles corresponds.
__ __
a wz  bw  b z  a  0
Last part. The given transformation is
For the fixed points, we have w=z. Hence putting w=z in the above
transformations the fixed points are given by
__ __
a z 2  bz  b z  a  0
__
 __

a z2   b  b  z  a  0
 
Or

 __
  __ 2
 __ 

b b   b  b   4a a 
    
z __
 (1)
2a
Or
Thus the two fixed points are

121
 __
  __ 2
 __ 

 b  b  b  b   4a a
    
z1  __
2a
 __
  __ 2
 __ 

 b  b     b  b   4a a
    
z2  __
2a
Sine algebra of complex numbers is the same as that for real numbers,
therefore from (1)
a
z1z 2  __
1
a z1 z 2  1
or
z1 , z 2
Hence if are distinct they are equal, then the relation to the circle
2
z 1 z1  1, or z1  1
. And if they are equal, then the relation becomes ; and
z1 z  1.
then lines on the circle
3.16. Some Special Transformation.
w  zn
The transformation (To examine it by taking polar form).
dw / dz  nz n 1
Here n is any integer. It is noticed that is non zero
except at z=0, hence the mapping is conformal at all points except at z=0.
w  Rei z  rei w  zn
Let and so that the given transformation
Rei  r n ein
becomes
R  rn   n
Thus ……(1) and …(2)
From relations (1) and (2), the following facts about this mapping are
easily seen to be true.
(i)The circle r=constant about origin in the z-plane maps on the circle
R=constant about the origin in the w-plane.

(ii) The lines =constant emanating from the origin, where z-plane
 
are transformed into the lines =constant, where slope of -lines

is n times the slope of the -lines.

122
(iii) A circular sector (wedge) with its vertex at the origin in the z-
plane is mapped upon a circular sector (wedge) with its vertex at
the origin and with n times the centre angle.
/n
(iv) The interior of a wedge of the central angle is mapped upon
I(w)  0
the entire upper half-plane
2 / n
(v) The interior of a wedge of central angle is conformally
mapped upon the entire w-plane cut along the entire positive real
axis. This cut is often called aslit.
Particular case. When n=2, the above transformation is then w=z 2, which can
be discussed similarly as above if we take polar forms for z and w.
3.17 The transformation w = z2.
(To examine it by considering Cartesian forms for z and w).
dw / dz  2z,
Here which is not equal to zero except at z = 0.
Hence the transformation is conformal except at z = 0.
The transformation is w = z2, where z = x+iy and w = u+iv
u  iv   x  iy   x 2  y 2  2ixy
2

.
Equating real and imaginary parts, we have
u  x 2  y 2 .....(1) v  2xy.....(2)

The following facts about these transformations can be easily seen to be


true.
(1) When x = constant, say x = a.
u  a 2  y 2 , v  2ay
Then from (1) and (2), we have
v 2  4a 2  a 2  u 
Eliminating y between these, we have .
Thus the line x=a in the z-plane corresponds to the parabola

v  4a 2 a 2  u
2
 a2, 0 
whose vertex is at and focus at the origin.

123
The relation v=2ay shows that v is positive when y is positive, and v is
negative when y is negative; so two parts of the line x=a lying above and below
the real axis in the z-plane correspond respectively to the parts of the parabola
lying above and below the real axis in the w-plane.
Writing –a for a in all the above equations we see that the line x=-a in
the z-plane corresponds to the same parabola in the w-plane but now upper and
lower part of the line respectively correspond to the lower and upper part of the
parabola.
x      0
Hence in general the line in the z-plane corresponds to the
2 2

v  4   a
2
 
parabola in the w-plane, where is a parameter.

As changes we have the following correspondence between on the
domain between the region of the two planes.
(i) The strip between the lines x=a, x=b in the z-plane maps on the
domain between two parabolas in the w-plane.

From the last not it follows that as takes the values continually from

a to b, the line x= describes the area lying between the lines x=a, x=b in the

v 2  4 2  2  u
z-plane; and the parabola describe the area lying between the
v  4a a  u
2 2
2
 2

v  4b b  u
2 2

parabolas and .

124
Thus the strip between x=a, x=b in the z-plane corresponds to the
domain between the parabolas
v 2  4a 2  a 2  u  v 2  4b 2  b 2  u 
in w-plane.
Similarly another strip between x=-a, x=-b corresponds to the same
domain interior to the above two parabolas.
(ii)The domain in the z-plane to the right of the line x=a and also to
the left of the line x=-a both map on the exterior of a parabola in
the w-plane.
  
If takes the values continually from a to , the line x= describes
the entire region of z-plane lying to the right of the line x=a, and the parabola
 
v 2  4 2  2  u
describes the whole of w-plane exterior to the parabola
2

v  4a a  u
2 2

.

Similarly the domain in the z plane to the left of the line x=-a is
conformally mapped on the domain in the w-plane exterior to the same

v 2  4a 2 a 2  u 
parabola
(iii) The domain in the z-plane between the lines x=0, x=a maps
on the interior of a parabola, in the w-plane, with a slit along the
negative real axis.

125
 
If takes values continually from a to 0, the line x= describes the
area in the z-plane lying between the line x=a, and x=0, while the parabola

v 2  4 2  2  u
describes the entire domain of w-plane interior to the
2 2

v  4a a 2  u  
parabola with a slit along the negative real axis from 0 to ;
as a matter of fact this slit in the w-plane corresponds to the line x=0 in the z-
plane.
Thus the domain in the z-plane for which 0<R(z)<a corresponds to the

v 2  4a 2 a 2  u 
entire interior of the parabola

With a slit along the negative real axis from 0 to - .
 
Particularly by making a i.e. for 0<R(z)< the half of the z-plane
lying to the left of x=0, corresponds to the entire w-plane with a slit along the

negative real axis from 0 to - .
z w
3.18 The transformation . (Inverse mapping of w=z2).
r  w or z 2  w or  x  iy   u  iv
2

The transformation is
Equating real and imaginary parts we have
u  x 2  y 2 .......(1) or v  2xy......(2)

The following facts of these transformation are easily seen to be true.


1. When u = constant, say u=a
(i) The line u=a, where a>0, in the w plane corresponds to the
x 2  y2  a
rectangular hyperbola in the z-plane.

126
If u=a, ten we have form (1)
x 2  y2  a
which is a rectangular hyperbola in the z-plane
whose transverse and conjugate axes lie along the axes of reals
imaginaries respectively.
(ii) The lines u=a, u=b in the w-plane separately correspond to the
x 2  y 2  a, x 2  y 2  b
hyperbolas .
Next, we proceed to establish conformal property with regard
to the regions of the two planes.

In general a line u= in the w-plane corresponds to the
x 2  y2  
hyperbola in w-plane.


Now, if takes all the values continually from a to b, (b>a),

then area swept out by the line u= is the one included
between the lines u=a, u=a; and the area in the z-plane swept
x 2  y2  
out by the hyperbola is the one included between
x  y  a, x  y 2  b
2 2 2

the hyperbolas . Hence the interiors


correspond.

127
(iii) The line u=a, in w-plane corresponds to the hyperbola
x 2  y2  a
, in the z-plane; and the entire domain of the w-

plane for which a<R(w)< corresponds to the domain of the
x 2  y2  a
z-plane which is interior to the hyperbola .


The line u= in the w-plane corresponds to the hyperbola
x 2  y2  
in the z-plane.
 
Let take all the values continually from a to , then area

described by the line u= will be the entire domain of w-plane
lying to the right of the line u=a, and area described by the
x 2  y2  
hyperbola will be the one which is interior to the
x y a
2 2

hyperbola .
II. When v=constant, say v=a.
(i) The line v=a in the z-plane corresponds to the rectangular
2xy  a
hyperbola in the z-plane.
v  2xy
Since ; therefore putting v=a in the above we get
2xy  a
which is hyperbola.
Thus the line v=a in the w-plane correspond to the hyperbola
2xy  a
, whose asymptotes are the real and imaginary axes of
the z-plane.

128
v  ,    0 
As a matter of fact, in general a line in the w-
2xy  
plane corresponds to the rectangular hyperbola in the

z-plane. As changes the correspondence between the regions
of the planes is as follows:
(ii) The line v=a, v=b, in the w-plane respectively correspond to
2xy  a 2xy  b
the hyperbolas and in the z-plane; the strip
between the lines v=a, v=b corresponds to the region between
the two hyperbolas; b>a.

v
We have shown that a line , in the w-plane corresponds
2xy  
to the hyperbola in the z-plane.
Hence the lines v=a, v=b respectively must correspond to the
2xy  a, 2xy  b
rectangular hyperbolas .

Let take all the value continually from a to b, then area

described by the line v= , in the w-plane, is that which lies
between the lines v=a, v=b; and the area described by the
2xy  
hyperbola , in the z-plane, is that which is included
between the hyperbolas
2xy  a, 2xy  b
.

129
Thus the strip in the w-plane between the lines v=a, v=b,
conformally maps on the region of z-plane included between
the hyperbolas.
(iii) The entire region in the w-plane above the line v=a,
corresponds to the entire domain of the z-plane interior to the
2xy  a
hyperbola .

 
Let take all the values continually from a to , then the area

described by the line v= is the entire region of w-plane above
2xy  
the line v=a: and the area described by the hyperbola
is the entire domain of the z-plane interior to the hyperbola.

Ex. 1.
Discuss the application of the transformation w=z 2 to the area in the
z  a, z  b
first quadrant of the z-plane bounded by the axes and the circles
(a>b>0). Is the transformation conformational.
Sol.
w  z2
The given transformation is .
w  Rei and z  rei ,
Putting we have

130
Rei  r 2 e 2 
 R  r 2 and   2
2
i.e. w  z and   2.

z a z b
Thus the circles and in the z-plane correspond to the
w a 2
w b 2

circles and respectively in the w-plane.


b z a c2  w  a 2
If , then
This shows the region in z-plane included between the circles
z  a, z  b
, corresponds to the region in w-plane included between
w  a , w  b2
2

1
0 
  2 2
Since , therefore the region in the z-plane maps on
0
the region in the w-plane.
Thus area in the first quadrant of the z-plane between the two circles
and the axes corresponds to the area in the w-plane between the two
corresponding circles and the real axis.
dw
 2z
dz
The transformation is conformal, because is not zero at any
point within this area.
Ex.2.

131
z  a  c,
By the transformation w=z2, show that the circles (a,c being
real) in the z-plane correspond to the limacons in the w-plane.
Sol.
z  a  c,
The given transformation w=z2. On the circle we have
z  a  c ci i.e. z  a  c ei
w   a  c e i 
2

or w  a 2  c 2   a  c ei   a 2  c 2
2

 c 2 ei2  2 ac ei  c 2
 c ei  cei  ce  i  2a 
 c ei  2 c cos   2a   2v e i  a  c cos  

a 2
 c2 
Transferring pole in the w-plane to the point , the above
equation becomes
w  2c  a  c cos   ei

Rei  2c  a  c cos   ei w  e i


Or writing
 R  2c  a  c cos   
and
R  2c  a  c cos  
Thus
which is limacon in w-plane.
za  c
Hence the circle in the z-plane correspond to the limacons in
the w-plane

132
Note.
za  a
In case when c=a, the circle in the z-plane is and the limacon
becomes the cardioids.
R  2a 2  1  cos  
.

Ex. 3.
w 1  c
Under the mapping w=z2, show that the family of circles is
z 1 z 1  c
transformed into the family of lemniscates where c is the
parameter.
Sol.
The transformation is
w  z2
 w  1  z2  1
or z2 1  w 2  1
or  z  1  z  1  w  1
or z 1 z 1  w 1

z 1 z 1  c w 1  c
Clearly when
z 1 z 1  c w 1  c
Showing that corresponds to the circle .
z 1 z  1  c
Now we proceed to show that represents a lemniscate.
z 1 z 1  c
We have
or z 2  1  c
(c real)
 z 2  1  cei 
where is real
z 2  cei  1 1  c ei
Or Or being a complex number can be written as
2
a e ix
a, 
where are real.

133
z 2  a 2e 2i
Hence
or r 2 ei  a 2 e 2i
or r 2  a 2 e 2i  

r 2  a 2 cos 2     
Equating real parts we have

which is a lemniscates whose axis is inclined at an angle to the real axis.
1 1
w z 
2 z
3.19. The transformation
dw 1  1  dw
 1  2    0 for z  1.
dz 2  z  dz
Here
z  1
The given transformation is conformal except at .
The given transformation is
1
u  iv  r  cos   i sin     1/ r   cos   i sin   
2
1  1 1  1
or u   r   cos ......(1) v   r   sin ......(2)
2 r 2 r

Now we have the following facts of this transformation.


1. When r=constant. We have from (1) and (2)
u2 v2
2
 2
 1.......(1)
1  1 1  1
r   r  
4 r  4 r

Two case arise, namely


(a)When r>1.
Then for all constant values of r positive and greater than 1, the circle
u2 v2
2
 2
 1.
1  1 1  1
z r r   r  
4 r 4 r
in the z-plane corresponds to the ellipse in
the w-plane.

134
Further, we notice that by writing 1/r for r, this equation of the ellipse
z  1/ r
remains unchanged but the circle becomes another circle . Hence both
z r z  1/ r
the circle , and correspond to the same ellipse.
u2 v2
2
 2
1
1  1 1  1 1  1 1  1
r   r   r  ,  r  
4 r 4 r 2 r  2 r
whose semi-axis are ,
z r
Thus corresponding to two systems of concentric circles and
z  1/ r
, ( r positive and greater than 1) in the z-plane there correspond in w-
u2 v2
2
 2
 1.
1  1 1  1
r   r  
4 r  4 r
plane a system of confocal ellipses the parameter
being r.
1  1 1  1
 r  , r  
2 r  2 r
If r increases, 1/r decreases, and lengths increase.
So when r increases continually from r to R, where R>r, then 1/r
1  1
r  
2 r
continually decreases from 1/r to 1/R, and the semi axes and
1  1 1  1 1  1 1 1
r   r   r   R  
2 r 2 r 2 r 2 R
continually increase from and to
1 1
R  
2 R
and respectively.

135
z  r, z  R
Hence the region between the two concentric circles and
z  1/ r, z  1/ R
also the region between the two concentric circles in the z
plane correspond separately to the same region in the w-plane included
u2 v2
2
 2
1
1  1 1  1
r   r  
4 r 4 r
between the two ellipses and
2 2
u v
2
 2
1
1 1 1 1
R   R   R 
4 R 4 R
. Making , we see that the entire region
z r
exterior to the circle , and also the entire region interior to the circle
z  1/ r
in the z-plane both separately correspond to the entire region in the w-
u2 v2
2
 2
 1.
1  1 1  1
r   r  
4 r 4 r 
plane exterior to the ellipse

136
(b)When r=1 Then:
z 1
The unit circle in the z-plane corresponds to the part of the real
axis between -1 to 1 in the w-plane described twice.
Putting r=1 in (1) and (2) equations to set up correspondence between
u  cos , v  0
the two planes are
  
We see that as varies from 0 to , u varies from 1 to -1; further as
 
varies from to 2 , u varies back from -1 to 1.
Thus the semi-circle above the real axis in the z-plane corresponds to
the part of the real axis from 1 to -1 in the w-plane; and the semi-circle below
the real axis in the z-plane corresponds to the part of the real axis from -1 to 1
in the w-plane.

II. When =constant.

The lines =constant, map on hyperbolas.

137
1 1
w z 
2 z
The transformation , is equivalent to
1  1 1  1
u   r   cos  ....(1) v   r   sin .....(2)
2 r 2 r  
when is constant, we have
u2 v2
 1
cos 2  sin 2 
from (1) and (2) by squaring and subtracting which
represents hyperbola.

Hence radial lines =constant in the z-plane map on hyperbolas in the
w-plane.

(a) Image of ,


When , the corresponding image in the w-plane is given by
1  1 1  1
u   r   cos ......(1) v   r   sin ......(2)
2 r 2 r 
where r varies from 0 to .
u2 v2
i.e. by  1  (3)
cos 2  sin 2 
w  1
which is a hyperbola whose foci are at

138
   
As r varies from 0 to , v varies from - to and u varies from
 
to . This shows that the line in the positive quadrant corresponds to
the branch of the hyperbola (3) lying to the right of the imaginary axis.

As a matter of fact as r varies from 0 to 1, v varies from - to 0 while u

varies from to 0.
Thus lower part of this branch is described as r varies from 0 to 1. Also
  cos  
as r varies from 1 to , v varies from 0 to while u varies from to ,

thus upper part of this branch is described as r varies from 1 to .
This branch to the right of the imaginary axis is also the complete image
  
of the line . As a matter of fact the upper part of this branch is

described as r varies from 0 to 1 and the lower part as r varies from 1 to .
Similarly second branch of the hyperbola (3) lying to the left of the
  
imaginary axis is the complete image of either of the two lines and
  
.
  ,   
(b) Image of the wedge
  ,   
The wedge defined by the lines .
  arg z  
i.e., the region in the z plane corresponds to the region in
the w-plane lying between the hyperbolas.
u2 v2 u2 v2
 1  1
cos 2  sin 2  cos 2  sin 2 
and
It is seen from (3) that the lengths of the semi-transverse and semi-
cos  and sin  
conjugate axes are respectively. If is increased, the length
of the transverse axis decreases and that of the conjugate axis increases. Thus

as the line revolves about O in the z-plane in anti-clock direction, the
corresponding curves in the w-plane are the hyperbolas whose transverse axes
become less and less and conjugate axes become more and more, so these
hyperbolas go on failing outside the preceding ones, but are confocal, the foci
w  1
being at .
   (where    )
Thus the line in the z-plane corresponds to the
2 2
u v
 2 1
cos  sin 
2

hyperbola in the w-plane and this hyperbola lies outside


the hyperbola (3) and is confocal with it.

139

Thus the region in the z-plane traced out by the line constant (for
different values of the constant parameter) corresponds to that region in the w-
plane traced out by the curve.
u2 v2
 1
cos 2  sin 2 

  to (where   )
Hence when varies from , then the region lying
  ,   
between the lines in the z-plane must correspond to the region in
the w-plane which lies between the hyperbolas.
u2 v2 u2 v2
 1  1
cos2  sin 2  cos 2  sin 2 
and
1
  ,   
2
(c) Image of the wedge

140
1
 
2
Putting in equation (1) and (2) we see that the positive
1
u  0, v   r  1 r 
2
imaginary axis in the z-plane corresponds to in the w-

plane where r varies from to ; these equations represent the whole of the
imaginary axis in the w-plane.
1
  arg z  
2
Thus the wedge-shaped region in the z-plane
corresponds to that domain of the w-plane which lies to the right of the
u2 v2
 1
cos 2  sin 2 
imaginary axis and is bounded by the hyperbola and the
imaginary axis.
  arg z    
(d) Image of the wedge
  arg z    
The domain of the wedge corresponds to the domain
between the two branches of the hyperbola.

We have shown that the line in the z-plane corresponds to the
branch of the hyperbola.
u2 v2
 1
cos 2  sin 2 
which is to the right of the imaginary axis, and the domain of the wedge
1
  arg z  
2
, corresponds to the region lying between this branch of the
  
hyperbola and the imaginary axis. Also we have shown that the line ,
in the z-plane corresponds to the other branch of the hyperbola, of domain of
1
  arg z    
2
the wedge in the z-plane must correspond to the domain
lying between the other branch and the imaginary axis.

141
Combining the two results, we see that the domain of the wedge
  arg z    
in the z plane corresponds to the domain between the two
branches of the hyperbola in the w plane.

0
(e) Image of the line
0
Putting in equations (1) and (2), we have
1  1
u   r  , u0
2 r 0 to 
where r varies from .

142
 to 1
As r varies from 0 to 1, u varies from , further when r varies from
 
1 to , u varies back from space 1 to .
0
Thus the line in the z-plane corresponds to the part of the real

axis from 1 to described twice.

  arg z  
We have shown that that domain of the wedge
Corresponds to the domain included between the hyperbolas
u2 v2 u2 v2
 1  1
cos 2  sin 2  cos 2  sin 2 
and
0 0  arg z  
Hence letting , we see that the domain of the wedge
corresponds to the domain interior to the hyperbola.
u2 v2
 1
cos 2  sin 2  
with a slit along the real axis from 1 to .
Ex. 1.
w  z 1 z
Show that the transformation , maps:-
(i) The interior of semicircle r = 1 in the uppr half of the z-plane into
the lower half w-plane and the exterior of the semi-circle r = 1 in the
upper half z-plane into the upper half w plane.
z  1, z  k (k  1)
(ii) The semi circular ring between the circle in the
upper half z-plane into the semi-eillpse.
2 2
 ku   kv 
 2   2  1
 k 1   k 1 
in the upper half w-plane.

Sol.

143
w  z  1 z,
The given transformation is
w  u  iv and z  rei
Let ,
Then the above relation becomes
1 1
u  iv  rei  i
 rei  e  i
re r
 1  1
  r   cos   i  r   sin 
 r  r

Equating real and imaginary parts, we have


 1
u   r   cos   (1)
 r
 1
v   r   sin   (2)
 r

Eliminating between (1) and (2), we have
u2 v2
 1  (3)
   
2 2
r 1 r 1
r r

1
r
which is an ellipse whose semi-major axis = r + and whose semi-minor
1
r
axis = r- .
From (3) it is seen that the circle r=constant maps on the ellipse (3).
1
r 0
r
As r1 the semi-major axis and the semi-minor axis
1
r 0
r
i.e. the ellipse (3) reduces to a slit in the w-plane from -2 to 2.
r0 r  
When or , then both semi-axes ; this shows that the
interior and as will as the exterior of the circle r=1 map separately on the entire
w-plane with a slit along the real axis from -2 to 2. As a matter of the fact the
following are the details of this mapping.
0
If , then we see from (2) that for r<1, v is negative, this shows
that the interior of the semi-circle r=1 it, the upper half of z-plane corresponds
to the entire lower half of the w-plane: and if r>1 then for the same variation in

144

, we have v positive, which shows that the exterior of the semi-circle r=1 in
the upper half-plane maps on the entire upper half w-plane.
    2
If , then we see from (2) that when r<1, v is positive, that is
interior of the circle r=1 maps on the entire upper half w-plane; and if r>1, then
v is negative, that is, exterior of the circle r=1 maps on the entire upper half w-
plane.
This proves first part of the question.
z k z k
(iii)When r=k (i.e when ), then from (3) the circle
transforms in the ellipse
u2 v2
2
 2
 1.........(4)
 1  1
k   k  
 k  k

0
If , then for r>1, we have just proved in part (i) that the
z 1
exterior of the semi circle in the upper half z-plane corresponds to the
0
entire upper half w-plane. Also we see from (2) that and k>1, the
z  k(k  1)
value of v is positive, showing that the interior of the semi circle
in the upper half z-plane corresponds to the interior of the semi-ellipse (4) in
the upper half w-plane. Hence we conclude from the above established
z  1 z  k(k  1)
mapping that the semi-circular ring between the circles . in
9 2
 ku   kv 
 2   2  1
 k  1   k 1 
the upper half z-plane maps into the semi-ellipse in
the upper half w-plane.
Ex. 2
1 1
w z  
2 z
Consider the transformation .
 w 1 
arg  
 w 1
By considering the or otherwise, prove that in general any
w  1
circle through two points corresponds to two circles in the z-plane.

145
w 1
In particular, show that corresponds to the circles
x  y  2y  1
2 2

=0. In find the region in the z-plane which corresponds to


w 1
.
Sol.
 w 1 
arg    2
 w 1 
Let where is real.
w  1.
This represents a circle which passes through the points
 w 1 
arg    2
 w 1  
If where is real.
w  1 2 2i
c e
w 1 
Then where c and both are real.
1 1
 z   1
2 z
i.e.   c 2 e 2i
1 1 1 1
 z   1 w z 
2 z 2 z
since
z 2  2z  1
 c 2 e 2i
z  2z  1
2
Or
2
 z2  1  2 2i z 1
 2  c e or  c ei
 z 1 z 1
Or
z 1 z 1
 c ei and  c ei  c ei 
z 1 z 1 (1)  ei
since
These are two circles whose equations are
z 1
arg   (1)
z 1
z 1
arg   (2)
z 1
w  1
Thus any circle through the points corresponds to the two
circles in the z-plane, which are represented by equations (1) and (2).

146
w  1.
Particularly. The circle transform into
1 1
 v    1,
2 z
2
1 1
 z    1,
4 z

1 1  1  __ 1 
 z   .  z  __   1
2 z 2 z
Or
__
__
1 zz
zz  __  __  4
__
zz z z
2 2
 __   __  __

 z z   1  z   z   4z z
2

   
Or
2 2
 __   __ 
 z z  1   z  z   0
   
Or
2

  __ 

2 2
z 1   z  z   0
 
Or

x  y 2  1  4y 2  0
2 2 __
z  z  2iy
since
x 2
 y 2  2y  1  x 2  y 2  2y  1  0
Or
x 2  y 2  2x  1  0.
Or
w 1
Proving that the circle corresponds to the both the circles
x  y  2y  1  0
2 2

and their interiors also correspond.


Ex. 3
1 1
w z 
2 z z
If , then prove that in general circles =constant and
w  1
lines are z=constant correspond to conics with focil at in the w-plane.
What exceptions are there?

147
Sol.
1 1
w z 
2 z
The transformation is
1  z  1
2
1 1
w 1   z   1 
2 z 2 z
So that

 z  1  z  1
__
2
1 z 1 1  
 w 1  
2 z 2 z

1 1  __ __

w 1    z z 1 z  z   (1)
2 2z 
Thus
 z  1
2
1 1
w 1   z   1  2
2 z z
And
 __ 
1 z 1 1
   z  1
z2 1
w 1    
2 z 2 z

1  __ __

w 1   z z1 z  z   (2)
2z 
Thus
z 
(a)When =const= say.
Adding (1) and (2), we have
1  __ 
w 1  w 1   2z z  2 
2z

__ 2
z z 1 z 1
 
z z

2 1
 z 

since (const)
That is, sum of the distances of w, from two fixed points 1, -1, is
constant.
Hence by property of ellipse the locus of w is an ellipse whose focii are
the fixed points 1, -1.

148

(b) When arg (z)=constant say .
Subtracting (1) from (2) we have
1  __

w  1  w 1   z  z
z  

2R e (z)

z

 2 cos  arg z 

 2 cos  
(where is constant)
w 1  w 1
Thus =constant.
That is , difference of the distances of w from two fixed points -1,1 is
constant. Hence by property of hyperbola locus of w is a hyperbola whose foci
are the fixed points -1,1.
1 a  b
w  a  b  z  
2 z 
3.20 The transformation
1 a  b  i 
i
u  iv    a  b  r e i  e 
z  re w  u  iv 2 r 
If and , then
1 a b 1 a b
  a  b    cos   i  a  b  r   sin .
2 r  2 r 

Equating real and imaginary parts, we have


1 a  b
u  a  b    cos   (1)
2 r 

1 a  b
v  a  b  r   sin   (2)
2 r 

Now, we have following facts of the transformation.



I. When r=const. Then eliminating between (1) and (2), we get
u2 v2
2
 2
1 
1  a  b  1  a  b 
 a  b  r     a  b  r   
4  r  4  r 
(3)
which is an ellipse in the w-plane.

149
z r
Hence the circle in the z-plane correspond to confocal ellipses in
2 a 2
 b2 
the w-plane the distance between the foci being

II. When =constant. Then eliminating r between (1) and (2) we get,
u 2 sec2   v 2 cos ec2   a 2  b 2

u2 v2
  1.......(4)
 a 2  b2  cos 2   a 2  b2  sin 2 
Or which represents
hyperbola in the w-plane
Hence the straight lines through the origin in the z-plane correspond to
confocal hyperbolas in the w-plane, the distance between the foci being
2 a 2
 b2 
.
Note.
1 1
w z  
2 z
If a=1, b=0, then the transformation becomes . This has
been discussed in 3.19.
w  z1/ 
3.21. Transformation , This transforms wedge into semi-plane.
z  rei w  Rei
Let and then the given transformation becomes
Re   re
i

 1/ 
i

Rei  r1/  ei / 


or Equating real and imaginary parts, we have
R  r1/    /   
and or …….(2)
Now following facts of the transformation are easily seen to be true.
0     0
We see from (2) that then correspondingly .

Thus the area bounded by the infinite wedge of central angle with
its vertex at the origin and one arm of the angle along the positive x-axis is
transformed into the upper half of the w-plane.
I. When r=const. We see from (1) that when r is constant R is also
constant; this shows that the circles in the z-plane correspond to
circles in the w-plane.

150
z 1
Particularly when r=1. The circle corresponding to the circle
w 1
.

II. When =constant. We see from (2) that
     ,
When =constant: say = , then so these lines in the two
planes corresponding.
0   
Particularly when and .
0 
Then we see from (2), that we have correspondingly . ,
separately.
0  arg z  
Thus the wedge is mapped on the upper half w-plane.
From the above two particulars we conclude that the sector cut off from
0  arg z   z 1
the wedge by an arc of the unit circle is transformed into
the unit semi-circle in the upper half of the w-plane.
 z 
w  tan 2  
 4 
3.22 The transformation
 z
w  tan 2  
 4 
To show that the relation transforms the interior of the
w 1
unit circle into z-space lying within a parabola.

1 
w  tan 2   z 
4 
The transformation is

151
1 
2sin 2   z 
4 
1 
2 cos 2   z 
4 
=
1 
1  cos   z 
2 
 1 
1  cos   z 
2 
=
1  1 1

1  cos   z   1  cos   re 2 
2  2 
Now
1 1 1 1 
1  cos   r cos   i  r sin  
2 2 2 2 
=
 1 1  1 1  1 1  1 1 
 1  cos   r cos   cosh   r sin     isin   r cos   sinh   r sin  
 2 2  2 2  2 2  2 2 

Similarly
1 1 1 1 1  1 1  1 1 
1  cos  z    r cos )cosh   r sin     isin   r cos   sinh   r sin  
2 2 2 2 2  2 2  2 2 
2
1
1  cos  z
2 2
w  2
1
1  cos  z
2
Now,
2 2
 1 1  1 1   1 1  1 1 
1  cos   r cos   cosh   r sin    sin   r cos   sinh   r sin   
 2 2  2 2   2 2  2 2 
2 2
 1 1  1 1   1 1  1 1 
1  cos   r cos   cosh   r sin    sin   r cos   sinh   r sin   
 2 2  2 2   2 2  2 2 

w 1
Hence according as
1 1  1 1 
cos   r cos   cosh   r sin    0
2 2  2 2 
or according as
1 1 
cos   r cos    0
2 4 
since other factor is positive

152
1 1  1
cos   r cos    
2 2  2
According as
1 1
r cos   1 r  sec 2 
2 2
Or or
1
w 1 r  sec2 
2
showing that the circle corresponds to the parabola and
interior of the circle to the interior of the parabola.
Ex.
1  1
w  tan 2  z  x 
2  2
If , the strip in z-plane between x=0, is
represented on the interior or the unit circle in the w-plane cut along the real
w  1 to w  0
axis from .
Sol.

1 1  cos z
w  tan 2  z w
2 1  cos z
The given transformation is i.e. on the line
1 1
x  , z    iy
2 2
1  i sinh y
w
1  i sinh y
So that

153
w 
1  i sinh y

 1  sinh y   1 i.e. w  1
2

1  sinh y  1  sinh y 
2

Thus
1
x 
2
Hence the line in the z-plane is represented by the unit circle in
the w-plane.
1  cosh y 2  e y  e  y
w 
1  cosh y 2  e y  e  y
On the line x=0, z=iy so that which is
real.
2e y  e 2 y  1
y  , w  L im  1
y  2e y  e 2 y  1
When
1  cosh y 
y  0, w    0
1  cosh y  y 0
And when
2  e y  e y 2e  y  1  e 2y
y  , w  L im  L im 1
y  2  e y  e  y y 2e  y  1  e 2y
When
y   to y  
Hence as z travels along the line x=0, from , we travels
along its real axis from -1 to 0, and back again from 0 to -1. In other words the
line x=0 in z-plane corresponds to a cut from u=-1 to u=0, in the w-plane.
Further we proceed to show that their interiors also correspond.
1  cos z 1  cos  x  iy 
w 
1  cos z 1  cos  x  iy 
We have
1  cos x cosh y  i sin x sin y

1  cos x cosh y  i sin x sin y

1  cos x cosh y  i sin x sin y


w 
1  cos x cosh y  i sin x sin y
So that
 1  cos x cosh y  2  sin 2 x sinh 2 y 
  .
 1  cos x cosh y   sin x sinh y 
2 2 2
 

 w  1when cos x  0
As cosh y is always positive

154
1 
0x  x
2 2
i.e., when showing that the strip between x=0, in the
z-plane corresponds to the interior of the unit circle in the w-plane.

w  ez
3.23 The transformation
w  Rei z  x  iy Rei  e x iy  e x .eiy
Writing and , we have
R  ex  (1)
So that
 y  (2)
and
y
(a) when (constant).

Then from (2).

y 
Thus the line in the z-plane corresponds to the line in w-
plane.
Hence the lines parallel to x-axis in the z-plane correspond to radial
lines in the w-plane.
y y
As the line moves parallel to itself to a position , then
 
correspondingly he line turns about O and takes up the position .
y y
Hence the area in the z-plane traced out by as it moves to .

Corresponds to the area in the w-plane traced out by the line as it moves

to
y  , y  
Thus the lines in the z-plane respectively correspond to the
  ,   
radial lines and their interiors also correspond.
  0,   
Particularly. Taking .

155
y  0, y  
The region in the z-plane between the lines corresponds to
half the w-plane above the real axis.

(b) When x=0


w 1
Then, we see from (1) that R=1 i.e. .

w 1
Thus imaginary axis in the z-plane corresponds to the unit circle
in the w-plane.
(c)It is shown above that
y  0, y  , x  0
The boundarlies in the z-plane respectively
  0,   , w  1
correspond to in the w-plane.
y  0, y  , x  0
And the region between in the z-plane corresponds to
the half w-plane above the real axis.

156
y  0, y  , x  0
Hence the domain bounded by the countour in the z-
  0,   , w  1
plane corresponds to domain bounded by the countour in
the plane.
(d) when y=mx.
R  e
i / m 
Then from (1) and (2), we have which represents equiangular
spiral in the w-plane.
Hence the line y=mx in the z-plane corresponds to an equiangular spiral
in the w-plane.
1.24 The transformation w=cosz
u  iv  cos  x  iy   cos x cosh y  i sin x sinh y
Or
Equating real and imaginary parts we have
u  cos x cosh y....(1)

v   sin x sinh y
…(2)
Then following facts of the transformation are true.
1. When x=constant.
(a)Image of x=a
Putting x=a in (1) and (2) we see that
u  cos u cosh y,.......(A) v   sin a sinh y.......(B)

Eliminating y between these two equations we have


u2 v2
  1.........(3)
cos 2 a sin 2 a
Thus the line x=a in the z-plane corresponds to the hyperbola
u2 v2
 1
cos 2 a sin 2 a
.

157
1
0a 
2
As a mater of fact if , then it is seen from (A) and (B) that as
 to 
y varies from , u remains positive, while v is negative and positive
both. Hence the line x=a maps on the branch of the hyperbola which lies to the
right of the imaginary axis.
The second branch of the hyperbola is the mapping of the line x=-a
(b) Image of the domain between x=a, x=b.
It follows from above that the line x=b maps on the hyperbola
u 2
v2 1
2
 2
1 b 
cos a sin a 2
where .
As x increases from a to b, the length of the transverse axis decreases
while length of the conjugate axis increases. Thus the successive hyperbolas go
on falling outside the preceding ones.
Hence the infinite strip between x=a, x=b maps on the domain between
u2 v2 u2 v2
  1  1
cos 2 a sin 2 a cos 2 b sin 2 b
the branches of the hyperbolas and lying
1
b>a>0 and b< .
2
to the right of the imaginary axis where
1
x  .
2
(c) Image of the domain between x = 0,
1
x 
2
From (1) and (2) we see that when then u = 0 and v = -sinh y.
1
x 
2
Further this second equation shows that y varies on the line from
1
x 
 to   to  2
,v varies from . Thus image of is the complete
imaginary axis in the w – plane.
Also putting x = 0 in (1) and (2), we have u = coshy, v = 0. From these
 
we see that as y varies from to 1 and as y varies from 0 to + , u varies form
  
1 to . Thus as y varies from - to , the part of the positive real axis in the

w-plane from 1 to is described twice. Thus image of the line x = 0 is a slit

along the positive real axis from 1 to .

158
1
b 
2 a 0
By making and we see that domain between x = 0,
1
x 
2
, that is the half the z-plane to the right of imaginary axis maps
conformally on the semi-plane to the right of the imaginary axis with a slit

along the positive real axis form 1 to .
II When y = constant.
(a) Image of y = a.
In (1) and (2) on putting y = a, we have
u = cos x cosh a. v = - sin x sinh a
Eliminating x between these, we have
u2 v2
 1
cos 2 a sin 2 a
which is an ellipse in the w-plane.
Hence image of the line y = a is the ellipse
u2 v2
 1
cosh 2 b sinh 2 a
(b) Image of the strip between y = a, y = b, (b > a)
If follows from above that image of the line y = b is the ellipse
u2 v2
 1
cosh 2 b sinh 2 b
The semi-axes cosh b and sinh b have increased as y goes from a to b.
Hence the infinite strip y = a, y = b maps on the domain between the
confocal ellipses
u2 v2 u2 v2
  1, and   1.
cosh 2 a sinh 2 a cosh 2 b sinh 2 b

Re  z   0.
(c) Image of
Re  z   0
Half the z-plane maps on the entire w-plane.
a  0, and b  
In (b) part make , then the corresponding ellipses
separately tend to v = 0 and to an ellipse whose semi-axes cosh b and sinh b
both tend to infinity.

159
Hence half the z-plane to the right of the imaginary axis [i.e., Re (z) >
0] maps on the entire w-plane.

3.25 The transformation z = c sin w.


To discuss the transformation z = c sin w, (c being real(), and to show
that the z-curve corresponding to the contour of the w-rectangular
1
u   , v  ,
2
is the ellipse
x2 y2
 1
c2 cosh 2  c 2 sinh 2 
with two slits from the extremities of the major axis each to the nearer focus,
the interior of the two corresponding.
Sol.

The given transformation is z = c sin w,


i.e., x + iy = c sin (u + iv) = c (sin u cosh v + i cos u sin hv)
Equating real and imaginary parts, we have
x  csin u cosh v   1
y  c cos u sinh v   2

where v is constant the z-curves are obtained by eliminating u between (1) and
(2), so they are
x2 y2
 1   3
c 2 cosh 2 v c 2 sinh 2 v

x2 y2
 1
v   c 2 cosh 2  c 2 sinh 2 
Thus when , the z=curve is the ellipse
1 1
   u  ,
2 2
In general when then cos u is positive.

160
 

v 2 2
On the side PQ, and v varies between and . Hence, from
(2) we see that y is positive and from (1) we see that x varies form
c cosh  to c cosh .

Thus the side PQ of the rectangle corresponds to upper half of the


ellipse above thre real axis.
 

v   2 2
On the side RK, and u varies between and . So from (2) y
c cosh  to c cosh .
is negative and from (1) x varies from
Thus the side RK of the rectangle corresponds to the lower half of the
ellipse below the real axis.
Hence only two sides of the rectangle which are parallel to the real axis
correspond to the whole arc of the ellipse.
1
u 
2  
On the side PK, and v varies from to
; so from (2), y = 0
 
and x goes form c cosh to c; further as u goes from 0 to then also y = 0 but

x goes back form c to c cosh .
Thus the side PK of the rectangle corresponds to a slit from an
extremity A of the major axis to the nearer focus S.
Similarly the side QR of the rectangle corresponds to a slit from the
other extremity A’ of the major axis to the nearer focus H.
1
u   , v  
2
Thus the contour of the rectangle in the w-plane
corresponds to the contour of the ellipse in the z-plane with two slits from the
extremities of the major axis each to the nearer focus.
Next, we proceed to show that the interiors also correspond.
We see form equations (1), (2) and (3) that y = 0, when v = 0 and
lengths c cosh u and c sinh u of the semi-axes increases as v increases. Hence
the point interior to the rectangle corresponds to a point interior to the ellipse.
Hence the interiors of the two correspond.
Ex. 1.
In the transformation z = sin w, show that rectangle
  u  , v1  v  v 2
maps into the region bounded by two confocal ellipses
v1  0
and if , the elliptic ring has a cut along the negative y – axis.

161
Sol.
The given transformation is
z = sin w
i.e., x + iy = (sin u + iv) = sin u cosh v+ i cos u sinh v,
x  sin u cosh v   1
so that
x  cos u sinh v.   2

When v = const. then eliminating u between (1) and (2), we have


2
x y2
  1,
cosh 2 v sinh 2 v
which represents confocol ellipses for different values
of v, the semi-axes being cosh u and sinh u.

v  v1 , v  v 2
Thus the lines in the w-plane correspond to the ellipses
2 2
x y x2 y2
  1 and  1
c 2 cosh 2 v1 c 2 sinh 2 v1 c 2 cosh 2 v 2 c 2 sinh 2 v 2
in the z-plane.
v  v1 , v  v 2
If v lies between v1 and v2, then a line between
corresponds to an ellipse lying between, and confocal to, the above two ellipse.
v  v1 , v  v 2
Thus the entire region between the lines corresponds to
the entire regions to the entire region between the above two confocal ellipse.

162
u   or u  
Putting (1) and (2), we see that x = 0, y = - sin v. Thus
u   , u  
both the lines in the w-plane correspond to the negative end of
the minor axis of the ellipse in the z-plane provided that v is positive.
Hence the region bounded by the rectangle
v  v1 , v  v 2 ; u  , u  

in the z-plane corresponds to the region bounded by the confocal ellipse


x2 y2 x2 y2
  1 ,  1
cosh 2 v1 sinh 2 v1 cosh 2 v 2 sinh 2 v 2
, with a cut along the negative
y-axis.
Ex. 2.

w  2z  z 2 z 1
If , prove that the circle corresponds to a cordioid in
the plane of w, show that vertices of an equilateral triangle inscribed in the
circle correspond to the points of contact of parallel tangents of the cordioid.
Sol.
w  2z  z 2 or  w  1   z  1
2

The given transformation is


w  1  Rei
Put i.e. shift the pole to the point (-1,0) in the w-plane.
z 1 z  e i .
Any point on the unit circle is
Substituting in the given transformation, we have
1 
Rei   1  ei   ei  ei / 2  ei / 2   4ei cos 2   
3 2

2 

R  4 cos 2  / 2 and   0,
So that

163
R  4 cos 2  / 2  2  1  cos    2  1  cos   .
Hence
i.e., R  2  1  cos   .

z 1
Hence the circle corresponds to a cordioid in w-plane.
Second part.

It can be easily seen in the figure that the vectorial angles of the vertices

of the equilateral triangle ABC inscribed in the circle are ,
  2 / 3,   4 / 3.
respectively.
  0,
Since therefore vectorial angles of the corresponding points on
,   2 / 3,   4 / 3.
the cordioid are also
R  2  1  cos   ,
Now in the cordioid we have
u  R cos   2  1  cos   cos 

 2 cos   2 cos 2 

v  R sin   2  1  sin   sin   2sin   2sin  cos 

dv dv / d 2 cos   2 cos 2
  
du du / d 2 sin   2sin 2

cos   cos 2  3 
   cot  
sin   sin 2  2 

dv 3
i.e.,   cot at any po int  of the cordioid.
du 2
Hence the slopes at the points of the cordioid, corresponding to the
vertices of the equilateral triangle in z-plane, are

164
3 3 2  3 4 
 cot ,  cot     ,  cot    
2 2 3  2 3 

3  3   3 
i.e.,  cot ,  cot     ,  cot   2 
2  2   2 

3 3 3
i.e.,  cot ,  cot ,  cot .
2 2 2
That is, tangents at these points have same slopes, therefore the three
tangents are parallel.
Hence it is proved that the vertices of an equilateral triangle inscribed in
the unit circle in z-plane correspond to the points of contact of parallel tangents
of the cardioid in w-plane.
Ex. 3.
 w  1
2

If = 4/z, then prove that the unit circle, in the w-plane


correspond to a parabola in the z-plane, and the inside of the circle to the
outside of the parabola.
Sol.
4
 w  1
2

z
The given transformation is
1
2 2  
or w  1  e 2
1
z r

2  1 1 
  cos   i sin    1
r 2 2 
 2 1  2 1
 cos   1  i sin 
 r 2  r 2
1
 2 1   2 1 
2
 2
 w   cos   1   sin   
 r 2   r 2  
1
4 4 1 2
  cos   1
r r 2 

165
4 1 4
Clearly w  1according as  cos   1  1.
r r 2
4 4 1
Clearly w  1according as  cos   1  0.
r r 2
1
Clearly w  1according as1  r cos   0.
2
1
Clearly w  1according as r cos 2   1.
2

w 1
Showing that the unit circle in w-plane corresponds to the parabola
1
r cos 2   1
2
in z-plane, and interior of the circle correspond to the exterior of
the parabola.
Ex. 4.
w  z  i  1
2

Show that the transformation maps the inside of the circle


z 1
in the z-plane on the exterior of the parabola.
Sol.
The given transformation is
1
w  z  i   1or z  i 
2

w
1
or z i
w
1
1  
 e 2
 i taking w  Rei
R
1  1 1 
  cos   i sin    i
R 2 2 
1 1  1 1 
 cos   i  sin   1
R 2  R 2 
1
 1 2
1   1 1  2
2

z   cos     sin   1 
 R 2   R 2  
So that

166
1
1 2 1 2
  sin   1
R R 2 

1 2 1
z  1 according as  sin   1  1
R R 2
Clearly
1 2 1
i.e., z  1 according as  sin   1  0
R R 2

1
or z  1 according as 1  2 R sin   1
2
1
or z  1 according as 4R sin 2   1
2

z 1
Showing that the unit circle in z-plane corresponds to the
1
4R sin 2   1
2
parabola in w-plane, and interior of the circle to the exterior of
the parabola.
Ex. 5.
2
 z  ic 
w   c real 
 z  ic 
Show that by means of the transformation
The upper half of w-plane may be made to correspond to the interior of
the certain semi-circle in the z-plane.

2
 z  ic 
w 
 z  ic 
The given transformation is

167
 x  i  y  c  
2 2
 x  iv  ic 
i.e., u  iv     
 x  iy  ic   x  i  y  c  

  x  i  y  c    x  i  y  c  
2

 
  x  i  y  c    x  i  y  c   
  x 2  y 2  c 2   icx 
2

 
 x 2   y  c 
2


x  y 2  c2   4c2 x 2  4icx  x 2  y 2  c 2 
2
2


x 2
  y  c 
2 2

x  y 2  c 2   4c 2 x 2
2
2

so that u   1
x 2
  y  c 
2 2

4cx  x 2  y 2  c 2 
2

v   2
x 2
  y  c 
2 2

x 2  y2  c2
We see from (2) that v is positive when x is positive and
negative. In other words upper half of w-plane corresponds to the right half
x 2  y2  c2 .
interior of the circle

168
Ex. 6.

z z 2
 c2   z  z 2
 c2   z  c  z  c .
Prove that
Hence show that if the relation between z-plane and w-plane be given
z  2wz cos   w 2  1.
2

by and if z describes an ellipse whose foci are the


branch points in the z-plane, w describes an ellipse whose focii are the branch
points in w-plane.
Sol.
The result

z z 2
 c2   z  z 2
 c2   z  c  z  c   1

has already been proved in ex.1, page 9.


The given transformation is
z 2  2wz cos   w 2  1  0

or w 2  2wz cos   z 2  1  0.

Solving for z, we have

z   w cos   i sin  w 2
 cos ec 2     2

and solving for w, we have

w   z cos  sin  z 2
 cos ec 2     3

It can be seen from (3) that the two values of w are equal if
z   cos ec.
z   cos ec.
Thus branch points are in the z-plane.
Similarly from (2), the values of z are equal if
w   cos ec.
w   cos ec.
Thus branch points are in the w-plane.
In an ellipse sum of the focal distances of any point is equal to length of
the major axis.

169
z   cos ec.
Hence taking the points as the focii and length of major
axis 2a, the equation of an ellipse in z-plane is
z  cos ec z  cos ec  2a   4

Using result(1) to the left-hand side of the above, we have

z z 2
 cos ec 2   z  z 2
 cos ec 2    2a.

w  z cos 
 z 2
 cos ec 2   
i sin 
.
From (3), we have so that
z z 2
 cos ec  
2

w  z cos  w  z  cos   i sin   w  zei


 z 
i sin  i sin  i sin 



w   w cos  i sin  w 2
 cos ec 2   ei  from  2  ,
i sin 

e i
we  i  w cos   i sin  w 2
 cos ec 2   e 
i

i sin 
w cos   iw sin   w cos   i sin  w 2
 cos ec 2 
e i

i sin 
 ei   w   w 2  cos ec 2    .
 

z  z 2
 cos ec 2    e i  w 
 w 2
 cos ec 2   


and z z 2
 cos ec 2    e i  w 
 w 2
 cos ec 2   


so that z  z 2
 cos ec 2    w  w 2
 cos ec 2  

sin ce e  i  1

and z  z 2
 cos ec 2    e i  w 

w 2
 cos ec 2   


w w 2
 cos ec 2    w  w 2
 cos ec 2   .

170
 z z 2
 cos ec 2    z  z 2
 cos ec 2   .

Or
w  cos ec  w  cos ec  z  cos ec  z  cos ec by  1

= 2a by (4)
i.e., w  cos ec  w  cos ec  2a.  cons tan t.
This shows that in the w-plane sum of the distances of point w form the
cos ec  cos ec
two points and is constant.
 cos ec
Hence w describes an ellipse whose focii are the points in the
w-plane.
Note.
If w = f(z) is a many valued function of z, then the value of z for which
two values of w are equal is called a branch point in the z plane.
Ex. 7.
4a cot 
z  0     / 4 .
1  2w cot   w 2
If show that when w
describes a unit circle z describes twice over an arc of a certain circle

subtending an angel 4 at the centre.
Sol.
4a cot 
z .
1  2w cot   w 2
The given transformation is
1 4a cot 
or  w  2 cot   .
w z

w 1 w  ei
On a unit circle in w-plane, we have
z  rei
Let in the z-plane.
Substituting these in above transformation, we have
4a cot 
 cos   i sin     cos   i sin    2cot    cos   i sin  
r
4a cot 
i.e.,  2i sin   2 cot    cos   i sin  
r

171
Equating real and imaginary parts, we have
4a cot 
2 cot   cos ,i.e., r  2a cos ,   1
r
4a cot 
2sin   sin .   2
r
and
Equation (1) represents a circle in the z-plane, with pole on its
circumference and the diameter through the pole as fixed line.
w 1 r  2a cos 
Thus the unit circle corresponds to the circle in the z-
plane.
4a cot 
2sin   sin .
2a cos 
From (1) and (2), we have
i.e., tan   sin  tan 

1 1
  to , 
 2 2  to ;
which shows that as varies from varies from further as
1 3
 ,
 2 2  to  ;
varies from to varies back from
Thus as w describes the complete circumference of the unit circle z-
2
describes twice over an arc of a circle subtending an angle at the
4
circumference or say subtending an angle at the centre.
Ex. 8.
z 1
w  log
z 1
In the transformation show that lines parallel to v-axis in
the w-plane correspond to a system of co-axial circles having their centres on
the x-axis in the z-plane. Also that lines parallel to u-axis in the w-plane.

172
Sol.
z 1
w  log
z 1
Here the transformation is .

i.e., u  iv  log
z 1
 log
 x  iy   1
z 1 x  iy  1

 log
 x  1  iy .
 x  1  y
Equating real and imaginary parts, we have
 x  1  y 2
2
1
u  log   1
 x  1  y 2
2
2

y y 2
v  tan 1  tan 1  tan 1 2
x 1 x 1 x  y2 1

2y
i.e., v  tan 1   2
x  y2 1
2

Lines parallel to the v axis are u=c where c is a parameter.


Now, when u = c(constant), then from (1),
 x  1  y2
2

e 2c

 x  1  y2
2

e 2c  1
i.e., x 2  y 2  2 x 1  0
e 2c  1
where c is parameter
x 2  y 2  2gx  1  0
or where g is the parameter
which represents a system of co=axial circles having their centres on the x-axis.
Hence for different values of the parameter c the lines u = c (i.e., the
lines parallel to v=axis) correspond to the system of coaxial circles having their
centres on the x-axis in the plane.
Line parallel to the u-axis are v = d
2y
d  tan 1 .
x  y2  1
2

So when v = d (constant), Then from (2), we have


i.e., x 2  y 2  2y cot d  1  0
where d is the parameter.

173
Hence the line parallel to u-axis correspond the system of co-axial
circles having their centres on the y-axis in the z-plane.
Ex. 9.
1 w2
z
1 w2
Show that the transformation maps the interior of the positive
quadrant of the unit circle in the w-plane conformally on the interior of the
positive quadrant of z-plane. Discuss also the correspondence between the
boundaries of the domains.
Sol.
  w2
The given transformation
1 
z .   2
1 
and
  w2
Now, consider the transformation .
i i
  e and w  Re ,
Let
ei  R 2 ei2  , so that   2.

1

 2  
Here when goes from 0 to , goes from 0 to , i.e., interior of
w 1
the positive quadrant of the unit circle in the w-plane maps on the

interior of the semi-circle in the upper half of the -plane, and the mapping is
d
 2w
dw
conformal except at the origin because is not zero except at w = 0
1  z 1
z or z 
1  z 1
Now, consider the transformation
z 1 z 1 2 z  z 4 Re  z 
1    1  .  
z  1 z  1  2  1  z  1 z 1
2

So that
2 4 Re  z 
1   2
z 1

 1
showing that then Re (z) is positive.

174
 1
It means that interior of circle corresponds to the right half of the
z-plane.
   1  z 1 z 1 
I     . 
2i 2i  z  1 z  1 
Further we see that
1 2 z  z  2I  z 
 
2i  z  1  z  1 z  1 2

I 
showing that is positive if I(z) is positive; which means that semi-circular

region in upper half of -plane corresponds to the positive quadrant of the z-
plane.
w 1
To sum up the interior of the positive quadrant of the circle ,

maps on the interior of the semi-circle in the upper half of the -plane and this
in turn maps on the interior of the positive quadrant of z-plane. Hence interior
w 1
of the positive quadrant of the circle , maps on the interior of the positive
quadrant of the z-plane.
Correspondence between the boundaries of the two domains :
As w goes its plane from 0 to 1 along the real axis, z goes in its plane

from 1 to .
w 1
As w changes on the circle in positive quadrant from 1 to I, z

goes from to 0 along the imaginary axis.
As w changes from i to 0 along the imaginary axis in its plane, z goes
from 0 to 1 along the real axis in its plane.
Thus positive parts of real and imaginary axes in z-plane correspond to
the positive quarter of the circle and the bounding radii.

175
Ex. 10.
2
 zc
w  
 zc
If where c is real and positive, find the areas of the z-plane
of which the upper half of the w=plane is the conformal representation.
Sol.
The given transformation is
2 2 2
 zc  x  iy  c   x  c  iy 
w   or u  iv     
zc  x  iy  c   x  c  iy 

   x  c   iy  x  c   iy 
2

 . 
  x  c   iy  x  c   iy 

 Multiply above and below by   x  c   iy to make the deno min ator real 
  x 2  y 2  c 2   i2cy 
2

 
 x  c   y2 
2

 2 2 
  x  y  c   4c y  i.4cy  x  y  c  
2 2 2 2 2 2

 .

 
xc  y
2 2 2


Equating real imaginary parts, we see

u
x 2
 y2  c2   4c 2 y 2
  1
  x  c2   y2
2

4cy  x 2  y 2  c 2 
v   2
 xc  2 2
 y2 
x 2  y2  c2
From (2) we see that v is positive if y and are either both
negative or both positive.
x 2  y2  c2
It means that the interior of the circle in the lower half and
exterior of this circle in the upper half both separately correspond to the upper
half of the w plane.
Ex. 11.

176
zc
w  
 zc 
If where c is real and positive, find the areas of the z-plane
of which the upper half of the w-plane is the conformal representation.
Sol.
Putting w = u + iv and z = x + iy in the given transformation and
equating only imaginary parts, we have
4cy  x 2  y 2  c 2  4cy  c 2  x 2  y 2 
v 
   
2 2
 x c 
2 2
 x  c2   y2
2
y 2

c2  x 2  y2
Clearly v is positive if y and are both positive or both
negative,
x 2  y2  c2
This implies that interior of the semi-circle in the upper half
and its exterior in the lower half of the z-plane both separately correspond to
the entire upper half of the w-plane.
Ex. 12.
Show that by means of the transformation
2
 z  ic 
w  
 z  ic 
the upper half of w plane may be made to correspond to the interior of a certain
circle in the z-plane.
Sol.
Putting w = u + iv and z = x + iv in the given transformation and
equating only imaginary parts, we get
4cx  x 2  y 2  c 2  3cx  c 2  x 2  y 2 
v 
 x2   y  c 
2 2
 x2   y  c 
2 2

x 2  y2  c2
Clearly v is positive if both x and are negative. It means
that the upper half of the w-plane corresponds to the interior of the semi-circle
x 2  y2  c2
in the z-plane, lying to the left of the imaginary axis.
Ex. 13.

177
 1 z 
3 2
 i  1  z3 
2

w
 1 z 
3 2
 i  1  z3 
2

Show that the transformation maps the domain


1
z  1, 0  arg z   w  1.
3
conformally on Discuss also the correspondence
between the boundaries of two domains.

178
Sol.
The given transformation is
 1  z3 
 1 z 3 2
 i 1 z 3 2  i
 1  z3 
w 
 1 z 3 2
 i  1 z 3 2  1  z3 
 3 
i
 1 z 

This is clearly a combination of the following transformations


2
 1   Zi
z , Z
3
 ,w .
 1   Zi

1
0  arg z  
z , 3
3
Now, maps the wedge conformally on the entire

upper half of -plane.
1  Z 1
Z or  
1  Z 1
Again, maps the interior of unit semi-circle
 1
on the entire upper half Z-plane,
Zi
w w 1
Zi
And, maps the upper half of Z-plane on to the circle .
Combining these three transformations we see that

 1 z  3 2
 i  1  z3 
2

w
 1 z  3 2
 i  1  z3 
2

1
z  1, 0  arg z   on w  1.
3
maps the domain

179
NOTES

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

180
UNIT - IV
COMPLEX INTEGRAL CALCULAS
4.1 Complex Integration.
In order to develop the subject further, it is necessary to consider the
definition of the integral of a function of a complex variable along a place
curve. In the case of a real variable, the integration is considered from two
points of view, namely the indefinite integration as an operation inverse to that
of differentiation and the definite integration as the limit of a sum.
The concept of indefinite integral as the process of inverse
Differentiation in case of a function of a Real variable also extends to a
function of a Real variable also extends to a function of a complex variable if
the complex function f(z) is analytic. Thus in case of complex variable if f(z)
 f  z  dz  F  z 
is an analytic function of z, and if , then differential of F(z) is
equal to f(z),
F  z   f  z 
i.e.,
However the concept of definite integral of a function of a real variable
does not out0right extend to the domain of complex variables. For example in
b

 f  x  dx
a
the case of real variables for the definite integral the path of
integration is always along the x-axis form x = a to x = b; or may be along the
b

 f  y  dy
a
y-axis of integral . But in the case of a complex function f(z) the path
b

 f  z  dz
a
of the definite integral can be along any curve from z = a to z = b; so
that its value depends upon the path (curve) of integration. But this variation in
values can be made to disappear if the different curve (paths) form a a to b are
regular curves.
4.2 Some Definitions.
If a point z on an arc is such that
z    t   i  t    A

then we may write

x   t   1 y   t   2

181
 t  t
If andare real continuous functions of the real variable t
t 
defined in the range , then the arc is called a continuous arc.
If the equation (A) or say the equation (1) and (2) are satisfied by more
than one values of t in the given range, then the point z, or say the point (x,y) is
a multiple point of the arc.
A continuous arc without multiple point is called a Jordan arc.
Thus for a point z on a Jordan arc, z as expressed in equation (A) is one
 t   t   t    t 
valued and , are continuous. In addition if and are also
 t 
continuous in the range , then the arc is called a regular arc of a Jordan
curve.
A continuous Jordan curve consists of a chain of finite number of
continuous arcs.
Contour. Contour means a Jordan curve consisting of continuous chain
of a finite number of regular arcs.
If A be the starting point of the first arc and B the end point of the last
 f  z  dz
AB
arc, then integral along such a curve is written as .
The contour is said to be closed if the starting point A of the arc
coincides with the end point B of the last arc.

 f  z  dz
C
The integral along such closed contour is written as , and is
 f  z  dz
C
read as integral f(z) taken over the closed contour C. Although does
not indicate the direction along the curve, but it is conventional to take the
direction positive which is anti clockwise, unless indicated otherwise.
4.3 Rectifibable curve (and its length).
Let the equations of the arc of a plane curve be
x   t , y    t where   t  .

 ,   t 0 , t1 , t 2 ....., t n
Subdivide the interval by points and let
P0 , P1 , P2 .....Pn
be the points on the curve corresponding to these values of t. The
P0 P1P2 .....Pn ,
length of the polygonal line is the sum of the lengths of the lines
P0 P1 , P1P2 .....Pn 1Pn .

182
z 0 , z1 , z 2 ......z n
Let be the points on the arc which correspond to the
t 0 , t1 , t 2 ....., t n P0 P1P2 .....Pn
values , then length of the polygonal arc
n 1/ 2
   x r  x r 1    y r  yr 1  
2 2
.
r 1
 

The value of this sum depends upon the mode of sub-division and is
called the length of an inscribed polygon.
If the arc is such that this sum (the lengths of all the inscribed polygons)
have a finite upper bound l, for all modes of the subdivisition, the curve said to
be Rectifiable and l is the length of the curve.
4.4 Riemann’s definition of Integration.
Let a function f(z) of a complex variable z be continuous in a domain D
and a, b two points in this domain, then integral of f(z) from a to b is defined as
below.

Let C be any curve joining a to b and lying entirely in the domain D, so


that f(z) is continuous on C.
Let there be any partition P = (a = z 0, z1, z2 …., zr-1, zr, ….. , zn = b) of
the curve C.
n
   z r  z r 1 
r 1
From the sum
f  r  z r 1  r  z r .
where exists uniquely, for any path whatsoever
joining a,b lying in D, and for any mode of division (partition), is called the
b

 f  z dz or
C
 f  z  dz
a
integral of f(z) over C from a to b and is written as
b n

 f  z  dz  lim   z r  z r1  f   r  where z r1   r  z r .


a
n 
r 1
Thus

183
4.5 Evaluation of some integrals, ab-initio (by direct definition.)
It is difficult to find the integrals of all the complex functions by the
direct application of the initial definition of the integral of a complex function.
However by taking some simple functions we find below their integrals ab
initio.
Ex. 1.

 zdz
C
Evaluate (ab-initio).
Sol.
The ab initio integral of f(z) is
 n 
C f  z  dz  lim    f   r   z r  z r 1   
n 
 r 1 

Hence making application of this, we have


 n 
C zdz  lim
n  

 r 1
 r  z r  z r 1   Since f  z   z

r
Where lies on C between zr and zr-1.
 n 
 lim    r  z r  z r 1   taking  r  z r .
n 
 r 1 

And also
 n 
 lim   z r 1  z r  z r 1  
n 
 r 1 

Hence on adding the two we have


 n 
2  zdz  lim    z r  z r  z r 1   z r 1  z r  z r 1   
C
n 
 r 1 

 
 lim    z r 2  z 2 r 1  
n 
 r 1 

 lim  z12  z 0 2    z 2 2  z12   ....   z n 2  z 2 n 1  


n 

 lim  z 2 n  z 0 2 
n 

 b2  a 2

184
Since zn = b and z0 = a, where a an db are the end points
A and B of the curve C.
1 2 2
  zdz 
2
 b a 
C

Cor.
If the curve C is closed curve the end points a and b coincide, hence in
the case of a closed curve C.

 zdz  0
C

Ex. 2.

 dz
C
Evaluate (ab-initio)

185
Sol.

See the figure of


 4.4.
Ab-initio integral of f(z) over C is
 n 
C f  z   lim
n   
 r 1
 r  z r  z r 1   ab  initio

Here f(z) = 1, hence making application of the above, we have


 n 
C n 
dz  lim
r 1
1 z r  z r 1   sin ce f  z   1

 lim  z1  z 0    z 2  z1   ....   z n  z n 1  


n 

 lim  z n  z 0 
n 

 b  a  Since z n  b and z 0  a 

= Chord AB (since vector b-Vector a = vector AB)


Where A and B are the end points of the curve
C).
Cor.

 dz
C
If C is a closed curve then points a and b coincide and then =0.
Ex. 3.

 dz .
C
Evaluate (ab-initio)
Sol.
Ab-initio integral of f(z) over C is
 n 
C f  z   lim
n   
 r 1
 r  z r  z r 1  

 dz
C
We are required to integrate by this method.
dz
Here f(z) =1, and in place of dz we have .
 n 
 dz  lim  1. z r  z r 1 
C
n 
 r 1 

186
 lim  z1  z0  z 2  z1  ....  z n  z n 1 
n 

 lim  Chord z1z 0  chord z 2 z1  ....  chordz n z n 1 


n 

= arc AB
where AB is the length of the curve C between the end points z0 and zn.
4.6 Complex integral as sum of two real line integrals.
Ex. 1. Find the value of the integral
1 i

  x  y  ix  dz
0

(a) Along the straight line form z =0 to z = 1 + i.


(b) Along the real axis form z = 0 to z = 1 and then along a line parallel
to the imaginary axis from z =1 to z = 1 + i.
Sol.

Let A be the point of affix 1 + i and N be the point of affix 1.


(a) Let OA be line from z = 0, to z = 1+i .
On OA, y = x, z = x + i x,
dz = (1+i) dx.

  x  y  ix  dz.
2

OA
Hence
1
  ix 2  1  i  dx
0

2
 x3  1  i
  1  i     .
 3 0 3

(b) The real axis form z = 0 to z = 1 is the line ON and then form z
= 1 to z=1+i a line parallel to imaginary axis is the line NA.
So here the contour of integration consists of the lines ON and NA.
On ON, y = 0, z = x + iy = x, dz = dz.

187
1 1
 x2 x3  1 i
  x  y  ix  dz    
x  ix dx    i    .
2 2

ON 0 2 3 0 2 3
Hence
On the line NA, x = 1, so on it z = 1 + iy, dz = idy.
1

  x  y  ix  dz    1  i   y idy


2

AN 0
Hence
1
 iy 2 
  1  i  y 
 2  0
i i
  1  i   1 
2 2

  x  y  ix  dz
1 i
2
0
Hence along the contour ONA
= Integral along ON + Integral along NA
1 i i 1 5i
  1    .
2 3 2 2 6
Ex. 2.
1 i

0
z 2 dz
Evaluate the integral
Sol.
f  z   z2
is analytic for all finite values of z, therefore its integral along
a curve joining two fixed points will be the same whatever be the path.
Here we have to integrate z2 between two fixed points (0,0) and (1,1).
Choose the path of integration joining these points as a curve made up of
(1) part of the real axis form (0,0) to the point (1,0). On this line z = x,
dz = dx and x goes, from 0 to 1.
(2) Followed by a line parallel to the axis of imaginaries from the point
(1, 0) to the point (1,1). On this line z = 1 + iy, dz = idy, and y goes
from 0 to 1.
1 i 1 1
 z 2dz   x 2 dx    1  iy  idy along the chosen path
2
0 0 0
Thus

188
1 1
1  1 3
  x 3     1  iy  
 3 0  3 0
1 1
   1  i   1
3

3 3 
1
  1 i .
3

3
Ex. 3.
Using the definition of the integral of f(z) on a given path, evaluate
5  3i
2  i
z 3dz.

Sol.
f  z   z3
is analytic for all finite values of z, so its integration. Along a
curve joining two fixed points will be the same, what ever be the path. Here we
have to integrate z3 between two points (-2,1) and (5,3). Let the path of
integration joining these points be along the curve made up of
(1) a line parallel to the real axis from the point (-2,1) to the point
(5,1) . On this line z = x+i, dz = dx and x goes from -2 to 5.
(2) Followed by a line parallel to the axis of imaginaries form the
point (5,1) to the point (5,3). On this line z = 5 + iy, dz = idy and
y goes from 1 to 3.
5  3i 6 3
 z3dz    x  i dx    5  iy  idy
3 3
2 i 2 1
Thus along the chosen path
5 3
1 4 1 4
   x  i      5  iy  
4  2  4 1
1 1
  5  i    2  i     5  3i    5  i  
4 4 4 4

4  4 

Ex. 4.

 z  3z  2  dz.
2

C
Evaluate where C is the arc of the cycloid
x  a    sin   , y  a  1  cos    a, 2a 
between the points (0,0) and .

Sol.

189
f  z   z 2  3z  2
The function is a polynomial and therefore analytic in
 a, 2a 
z-plane; hence its integral between tow points (0,0) and is independent
of the path joining these points.
For the sake of convenience choose the path of integration a curve C
consisting of :
 a, 0 
(1) the part of real axis form the point (0,0) to the point . On
a
this line z = x, dz = dx, and x goes form 0 to .
(2) followed by a line parallels to the axis of imaginaries from the
 a, 0   a, 2a 
point to the point . On this line
z  a  iy, dz  idy
, and y goes from 0 to 2a.
Hence

 z  3z  2  dz   x  3x  2  dx    a  iy   3  a  iy   2  idy
a 2a 2
2 2
0 0  
C

a 2a
1 3  1 3 
  x 3  x 2  2x     a  iy    a  iy   2iy 
2 2

4 2 0  2 2 0
1 3 3  1 3 3 1 3 3 2
   a    a   2a     a  i2a    a  i2a   4ia   a    a  
2 2

3 2  3 2 3 2 
1 4 3
 2a   a  i2a    a  i2a   4ia.
2

3 2
Ex. 5.
1
 z  a dz
Find the value of the integral round a circle whose equation
z  a  .

Sol.
z  a  .
On the circle C,
 z  a  ei .id,  2
where varies from 0 to .
dz  ei .id.
So that,

190
1 2 1 2
C z  a 0 ei 0 d  2i.
i
dz  .e .id.  i
Hence
4.7 Some elementary properties of complex integrals.
We have seen that a complex integral is the combination of tow real
integrals. Hence some elementary properties of real integrals must hold good in
case of the complex integral also. Some essential ones are stated below :

 f  z   F  z   dz   f  z  dz   F  z  dz
C C C
(1)

 f  z  dz   f  z  dz   f  z  dz
C C1 C2
(2)
where C1, C2 are two parts of C.

 f  z  dz    f  z  dz
C C
(3)
where –C indicate the direction opposite to that indicated
by C.
If the contour C when described in the opposite direction be denoted by
C1, then result (3) is equivalent to

 f  z     f  z  dz
C1 C

 kf  z  dz  k  f  z  dz
C C
(4) where k is a constant
4.8 An upper Bound for a complex integral.
Theorem.
If a function f(z) is continuous on a contour C of length l, and if M be
f  z
the upper bound of on C, then

 f  z  dz  Ml.
C

x   t , y    t
Let the equations of the curve C be .
1/ 2
 dx  2  dy 2 
l         dt   1
 dt   dt  

191
z = x + iy, dz = dx + idy

dz  dx  idy   dx    dy  
2 2
 
So that
2 2
 dx   dy 
 dz    dx    dy          dt  lfrom  1
2 2
   dt   dt 

 dz  l   2
C
Thus
We know that the modulus of the sum of n complex numbers cannot
exceed the sum of their moduli, therefore
n n

  f  r   zr  zr 1     f  r   z r  z r 1 
r 1 r 1

n
  f   r  z r  z r 1
r 1

n
By making the above inequality may be written as

 f  z  dz   f  z 
C C
dz

 M  dz Since max f  z   M
C

 MlSince from  2   dz  l
C

 f  z  dz  Ml.
C
Thus

192
4.9 (a) Cauchy’s Theorem. (Original). Riemann Proof.
f  z
If f(z) is a analytic function of z and if is continuous at each point
within and on a closed contour C, then

 f  z  dz  0
C

Let D be the region which consists of all points within and on the
contour C. If
Q P
P  x, y  , Q  x, y   .
x y

are all continuous functions of x and y in the region D, then Green’s theorem
states that
 Q P 
  Pdx  Qdy     x  y  dx dy,  Green 's Theorem 
C D

f  z
Since f(z) = u + iv is continuous on the simple curve C and exits
and is continuous of Green’s theorem are thus satisfied. Hence

 f  z  dz    u  iv   dx  idy 
C C

   udx  vdy   i   udx  udy 


C C

 v u   u v 
      dx dy  i     dx dy
D 
x y  D 
x y 

(by virtue of the above result of Green’s Theorem


= 0 By Cauchy’s Riemann equations.

 f  z  dz  0
C
Hence .
Remark :
Goursat showed that for the truth of the theorem the assumption of the
f  z
continuity of is unnecessary, and that Cauchy’s theorem holds if and only
if f(z) is analytic within and on C.
4.9 (b) Cauchy’s Theorem. (Revised form) or Cauchy Gousrat theorem).
Gousrat’s proof.

193
If a function f(z) is analytic and one valued inside and on a simple
closed contour C, then

 f  z  dz  0
C

Lemma 1.

Given , it is possible to divide the region inside C into finite number
of meshes, either complete squares or parts of squares, such that within each
mesh there exists a point z0, such that
f  z   f  z0 
 f   z 0  ,   1
z  z0

for all values of z in the mesh.


Suppose the lemma is false ; then it fails at least in one mesh.

Subdivide this mesh by lines joining the middle points of the opposite
sides. If there still remain any parts which do no satisfy the condition (1), we
again subdivide them in the same way. The process may end either after a finite
number of steps, when the condition (1) is satisfied for every subdivision, or
the process may go on indefinitely. In the second case we obtain a sequence of
squares (each contained in the preceding one) whose limit point is z0 which lies
inside C and at which the condition (1) is not satisfied. Since the condition (1)
is not satisfied at the point z0, therefore
f  z   f  0
 f   z 0  
 where z  z 0 is small.
z  z0

This shows that f(z) is not differentiable at z0, or in other words f(z) is
not analytic at z0. This contradicts the hypothesis that f(z) analytic at all the
points within and on the contour C.
Hence the lemma is true.

194
f  z   f  0
 f   z 0  
z  z0
i.e,.
f  z   f  0
 f   z 0     z  where   
z  z0
or
n  0 as z  z 0
and .
f  z   f  z 0    z  z 0  f   z 0    z  z 0  .   2
Thus
Lemma 2.
We know that

 dz  0 and also  zdz  0   3

taken over any continuous closed curve.


Proof of the theorem.
The lemma can be written as
f  z   f  z0 
 f   z 0    where   0 as  0.
z  z0

f  z   f  z 0   z 0 f   z 0   zf   z 0    z  z 0  .
or
C1 , C2 , C3 ....C m P1 , P2 ,....Pn
Let be the complete squares, and be the
partial squares, parts of whose boundaries are parts of C. consider
M N


m 1 Cm
f  z  dz    f  z  dz.
n 1 Pn

where integral along each contour is taken in anti-clock-wise direction (in the
sense of description of C).
In the complete sum the integration along each straight side of each
square (whether complete or partial) happens to be taken twice in opposite
directions. Hence all the integrals along straight boundaries of the squares
cancel, and only those integrals remain which are taken along the curved
boundaries of the partial squares because these are described only once. The
integrals which are left behind sum to

 f  z  dz.
C

Thus

195
M N

 f  z  dz  
C

m 1 Cm
f  z  dz    f  z  dz.
n 1 Pn
  4

M N
   z  z  dz     z  z  dz.
0 0   5
m 1 Cm n 1 Pn

where z0 represents any point inside the boundary Cm or Pn and z represents any
point lying on these boundaries, other integrals vanish by lemma (2).

  z  z  dz
Cm
0

Now,
  zz 
Cm
0  dz

l 2  dz sin ce  z  z 0   I 2
Cm

 
(length of the diagonal) and .
 l 2.  .4l Since 
Cm
dz  4l

(perimeter of the square)


l2  A m
Since
 .4 2.A m
(area of the square Cm)

  z  z  dz
Pn
0

Similarly
 l 2.  .  dz
Pn

 l 2.  .  4l  s n 

where sn is that length of the arc C,


which forms curved boundary of Pn.
 . 2  4A n  ls n 
where An is the area of the square Pn.

196
  f  z  dz
C

M N
    z  z  dz     z  z  dz.
m 1 Cm
0
n 1 Pn
0 from  5 

M N
   z  z 0  dz     z  z 0  dz
m 1 Cm n 1 Pn
M N
 .4 2   A m   . 2    4A n  ls n 
m 1 Cm n 1 Pn

M N  N
 .4 2    A m    A n    . 2l. s n
 m 1 Cm n 1 Pn  n 1

 .4 2A   . 2aL.
N
L   sn
n 1
Where
(total length of C), a, A the side and area of the square enclosing C.
 2  4A  aL   0 as  0

 f  z  dz  0
C
Hence , this proves the theorem.
4.10 Connected Region, Simply-Connected and Multi-Connected Region.
Connected Region.
A region is said to be connected region if any two points of the region D
can be connected by a curve which lies entirely within the region.
Simply-Connected.
A connected region is said to a simply-Connected region if all the
interior points of a closed curve C drawn in the region D are the points of the
region D. In other words, if all the points of the area bounded by any single
closed curve C drawn in the region D are the points of the region D, then the
region D is said to be a simply connected.
Multi Connected Region. If all the points of the area bounded by two or
more closed curves drawn in the region D, are the points of the region D, then
the region D is said to be Multi Connected Region.
C, C1 , C 2 , C3 .......
For example let there be a number closed curves all
drawn in the certain region D.( See fig. of the next article).

197
If all the points of the area lying between the closed curves
C, C1 , C 2 , C3 .......
the area which is interior to C and exterior to the other curves
C1 , C 2 , C3 , C4
are points of the region D, then the region D is said multi-
connected region.
Note :
In case of simply-connected region only one curve at a time is
concerned; while in case of multi0connected region more then one curve at a
time are concerned.
4.11 Cross cur (or cur) the lines drawn in a multi connected region, without
intersecting any one of the curves, which make a multi- connected region a
simply-connected one are called cuts or cross cuts.
Thus, let there be a multi-connected region lying between the curves C,
C1, C2, C3, C4 where C1, C2, C3, C4 lie inside C. In this region draw lines
joining C to all the curves C1, C2, C3, C4 then the same region which lies
between several curves C, C1, C2, C3, C4 can also be said to lie between a simple
curve whose boundary consists of the boundaries of C, C1, C2, C3, C4, and the
lines AB, PQ, RS, etc.

Such lines AB, PQ, RS …. etc., are called cuts or Cross-cuts.


By such a manipulation the same region which is a multi connected
region which is a multi connected region is made a simply-connected region.
This device make different curves C, C1, C2, C3, C4, as parts of the continuous
curve consisting of these curves and the cross-cuts.
4.12 Extension of Cauchy’s Theorem to multi-connected region.
Theorem.
If C is closed curve and C1, C2, C3,….. the other closed curves which
lies inside C, and if a function f(z) is analytic in the region between these
curves, and continuous on C, then

 f  z  dz   f  z  dz   f  z  dz   f  z  dz  .....
C C1 C2 C3

where integral along each curve is taken in the anti-clockwise direction.


Connect the curves C1, C2, C3, each to the curve C by means of the
narrow cuts AB, PQ, RS respectively. In this way the curve C is connected C 1

198
by a line AB going from a point A on the curve C to a point B on C 1 and by
another line BA going form the point B on C 1 to the point A on C, thus creating
a narrow channel cut of parallel sides AB and BA.
Similarly, we introduce other cuts PQ, RS. …. Etc., As a result of the
introduction of all these cuts curves to the already existing curves, the sum of
the lengths of these curves is a single closed curves, say F, in the interior of
which f(z) is analytic and on whose boundary f(z) is at least continuous. Hence
by Cauchy’s theorem.

 f  z  dz  0
F

where the integration along F, is taken in such way that while taking a complete
 
round over , the area enclosed by is always to the left of the goer.

When we go round in the above manner the rounding over C is anti-
clockwise, the rounding over each of the curves C 1, C2, C3 is clockwise and
rounding over the each cut is repeated twice in opposite directions, it follows
 f  z  dz
 C
that when going in this manner along , we have integral over C as ,
the integral over C1, C2, C3 as
  f  z  dz,   f  z  dz,   f  z  dz,
C1 C2 C3

 f  z  dz
AB
 f  z  dz
BA
the integral over the cut AB once as and again as with
similar two integrals for each cut cancel being opposite in sense. Hence


 f  z  0
is equivalent to

 f  z  dz   f  z  dz   f  z  dz   f  z  dz  0
C C1 C2 C3

 f  z  dz   f  z  dz   f  z  dz   f  z  dz
C C1 C2 C3
i.e.,
where all integrals are taken in anti-clockwise direction.
Thus the theorem is proved.
Particularly. If there is only one closed curve C1 within C2 then,

 f  z  dz   f  z  dz
C C1

where integration over each curve is in anti-clockwise direction.

199
4.13 Indefinite Integral or (Anti-derivative)
If a function f(z) is analytic in a simply-connected region D and if z 0 is
a fixed point and z any point in this region, then for all the paths from z 0 to z in
the region, the integral,
z
 f  z  dz
z1

is a one-valued function of the upper limit z, and is called an Indefinite Integral


of f(z) or Anti-Derivative of f(z).
Let z0 be a fixed point and z any other point in a simply be any two
curves in this region joining z0 to z.
Since the same tow points z0 and z are the end points of both the curves
C1and C2, therefore these two curves taken together from one simple closed

contour say within and on which f(z) is analytic. Hence by Cauchy’s theorem.

 f  z  dz  0


where integration round is taken in anticlockwise direction; ie.,
Integral of f(z) on C1 (in anti-clock direction) + Integral of f(z) on C2 (in anti-
clock direction) = 0 or Integral of f(z) on C1 (in clockwise direction) = 0

 f  z  dz   f  z  dz  0
C1 C2
i.e,.

 f  z  dz
C1
where integral taken over C1 is anti-clockwise direction and integral
 f  z  dz
C2
taken over C2 is in clockwise direction,
i.e., both integrals are from z0 to z as seen arrow marked in figure.

 f  z  dz   f  z  dz
C2 C2
Hence
When z0 is a fixed point then obviously value of these both integrals is a
function F(z) of the upper limit z, and since z is any point of a regular curve in
a simply connected region, therefore F(z) is one valued.
z
 f  z  dz
z0
 f  z  dz or  f  z  dz
C1 C1
Hence which is equivalent in is a
single-valued function F(z) of the upper limit z3.
4.14 Derivative of F(z).

200
To show that F(z) the indefinite integral of f(z) is an analytic of its
upper limit.
We have shown in the article that
z
F  z    f  z  dz
z0

z z
F  z  z    f  z  dz
z0
It follows that,
z z z
F  z  z   F  z    f  z  dz   f  z  dz
z0 z0
So that,
z z z
 f  z  dz   f  z  dz
z0 z0
z z
 f  Z  dZ
z0

The right hand side is the indefinite integral of f(z) where z is a fixed
z  z
point and Z is any point on the path joining z to .
z  z
Let the path of integration from z to be a straight line, then

 f  Z   f  z   dZ  f (z) 
z z z z
F  z  z   F  z    dZ
z0 z0

 f  Z   f  z   dZ  f  z  z
z z

z0

F  z  z   F  z  1 z z
z
 f  z 
z z0
 f  Z   f  z   dZ
or
 f  Z   f  z   Zz  
Since f(z) is continuous when
Zz  
In particular : when ,we have
F  z  z   F  z  1 z z
 f  z   f  Z   f  z  dZ
z z z

1 z z
  . dZ
z z

1
 .  . z 
z

201
F  z   f  z 
This shows that the derivation of F(z) exists and that at
every point z of the region D; proving there by that F(z) is analytic at every
point z of the region D.
Hence it is proved that F(z) the indefinite integral of f(z) is an analytic
function of its upper limit.
4.15 Cauchy’s Integral formula :
If f(z) is analytic within and on a closed contour C, and a is any point
within C, then
1 f  z  dz
f  a 
2i C z  a


About the point z = a, describe a small circle of radius r lying entirely
within C.
f  z
.
za
Consider the function


This function is analytic in the region between C and ; hence by
Cauchy’s theorem of multi-connected region we have
f  z  dz f  z  dz

C
za


za

f  z  dz f  a  dz f  z  f  a 
or 
C
z  a


z  a


z  a
dz

f  z dz f  z  f  a 
or  z  a dz  f  a   z  a  
C  
za
dz

f  z f  z  f  a  dz
or  z  a dz  f  a  2i  
C 
za
dz sin ce 

za
 2i

f  z f  z  f  a 
or  z  a dz  2i f  a 
C
 

za
dz

202
f  z  f  a 
 dz

za
dz
 

za
s in ce f  z   f  a   because f  z  is continuous at z  a

 dz sin ce z  a  r for z on 
r

 2r  2  0 as  0
r

f  z  dz
  2i f  a   0
C
za
Hence
1 f  z  dz
f  a 
2i C z  a
.
or
This is Cauchy’s integral formula.
The Cauchy’s integral formula established above expresses the value of
an analytic function at any point without the contour C in terms of the value of
the function at the boundary only.
4.16 Extension of Cauchy’ Integral formula to multi-connected regions.
C
If f(z) is analytic in the region bounded by two closed curves C and
and a is a point in the region, then
1 f  z  dz 1 f  z  dz
f  a 
2i C z  a 2i C z  a
 .

where C is the outer contour,


 
Draw a small circle with it centre the point .
f  z
.
za
Now consider the function
C 
It is analytic in the region bounded by the three contours C, and
(because z – a is not zero for any value of z in this annulus) and it is also
analytic on these curves.

203
Hence by Cauchy’s theorem for multi-Connected region
f  z  dz f  z  dz f  z  dz

C
za

C
za


za

where integration round each curve is taken in such a way that the annular
region always lies to the left
f  z  dz f  z  dz
C z  a C c  a  2i f  a 

Thus by Cauchy’s formula
1 f  z  dz 1 f  z  dz
f  a 
2i C z  a 2i C z  a

or
C, C,...
In general, if there be more curves etc., .. then we have
similarly.
1 f  z  dz 1 f  z  dz 1 f  z  dz
f  a 
2i C z  a 2i C z  a 2i C z  a
   ....etc.

4.17 Cauchy’s Integral formula for the Derivative of an Analytic


Function.
If a function f(z) is analytic in a region D, then its derivative at any
point z = a of D is also analytic in d, and is given by
1 f  z  dz
f  a  
2i C  z  a  2

where C is any closed contour in D surrounding the point z = a.


Let a + h be a point in the neighbouhood of the point a, then by
Cauchy’s Integral formula, we have
1 f  z  dz
f  a 
2i C  z  a 

1 f  z  dz
f  a  h 
2i C z  a  h
and

204
1  1 1 
f  a  h   f  a     f  z  dz
2i C  z  a  h z  a 

1 h
  f  z  dz
2i C  z  a  h   z  a 

f  a  h  f  a 1 f  z  dz
or
h
 
2i C  z  a  h   z  a 

f  a  h  f  a 1 f  z  dz
h 0 2i   z  a  h   z  a 
or lim  lim
h 0 h C

1 f  z  dz
or f   a     2
2i C  z  a  2

As a is any point of the region D, we see from the above established


f  a 
result that is analytic in D. Hence it is also established that the derivative
of an analytic function is an analytic function of z.
4.18 Analytic character of the successive derivates of an analytic function.
Theorem.
If a function f(z) is analytic in a domain D, then f(z) has, at any point z
= a of D. derivatives of all orders, all of which are again analytic functions in
D. their values given by
n! f  z  dz
f n  a 
2i C  z  a  n 1

where C is any closed contour in D surrounding the point z = a.


Assume that the theorem is true for n = m.
m! f  z  dz
f m  a 
2i C  z  a  m 1
i.e., is true.
f m  a  h  f m  a
h
So that
1 m!  f  z  dz f  z  dz 
h 2i  C  z  a  h  m 1 C  z  a  m 1 
 .   

205
1 m!  
 m 1
1 
 h  
 f  z  dz
h 2i C   z  a  m 1 
 .   1    1
 z  a  

1 m!  1  h  m  1  m  2  h 2 
 .   m  1   ....terms with higher powers of h  f  z  dz
h 2i C   z  a  m1  za 2!  z  a 
2


h 0
Taking limit as we have
f m  a  h   f m  a   m  1 !  f  z  dz 
2i C   z  a  m  2 
lim   
h 0 h

i.e., f m 1  a  
 m  1 ! f  z  dz
2i   z  a
C
m 2

which shows that theorem is also true for n = m +1.


But we know that the theorem is true for n = 1, therefore must be true
for n = 2 and so on.
Hence it must be true for any positive integral values of n.
n! f  z  dz
f n  a 
2i C  z  a  n 1
Thus
Also we see from this result that fn(a) is an analytic function of z. This
implies that derivatives of f(z) of all order are analytic if f(z) is analytic.
4.19 Morea’s theorem
If f(z) is a continuous function in a region D, and if the integral
 f  z  dz , taken round any simple closed contour in D, is zero then f(z) is an
analytic function inside D.
This is a sort of converse of Cauchy’s theorem.
If z0 be a fixed point and z any variable point in the region D, then value
of the integral
z
 f  z  dz
z0

is independent of the curve joining z0z is a function of the upper limit z; hence
we may write
z
F  z    f  z  dz
z0

206
zh
F z  h   f  z  dz
z
So that
zh z
F z  h   F z   f  z  dz   f  z  dz
z z0

z z h z
  f  z  dz   f  z  dz   f  z  dz
z0 z z0
zh
 f  z  dz
z

Let the points z and z + h be on the closed curve C, then


zh
lim  F  z  h   F  z    lim  f  z  dz
h 0 h 0 z

  f  z  dz
C
where C is the closed curve
h 0
(because when the two points coincide and the curve C becomes closed,
 f  z  dz  0
C
so that
 lim  F  z  h   F  x    0
h 0

F z  h   F z   1 zh 
or lim   f  z    lim  f  z   dz 
h 0
 h  h  0
 h z

 1 
lim  f  z   dz 
 h
h 0
C 
since joining curve C becomes closed
h0

 F z  h  F z 
or lim   f  z    0 sin ce  dz  0
h 0
 h  C

When C is closed
F z  h  F z
or lim  f  z
h 0 h

F  z   f  z  .
or
Thus we see that derivative of F(z) exists for all values of z in D,
therefore F(z) is analytic function is analytic; (see the last article).
4.20 Cauchy’s Inequality.

207
za  R f  z  M
If f(z) is analytic within a circle C, given and if on
C, then
Mn
f n  a 
Rn

We know from
 4.18. that
n! f  z  dz
f n  a 
2i C  z  a  n 1

n! f  z  dz
f n  a 
2i C  z  a  n 1

n! f  z  dz

2i 
C za
n 1

n! M 2 
2 R n 1 
 . Rd
sin ce z  Rei

dz  i.Rei d  Rd

n! M
 . .2R
2 R n 1
M.n!

Rn
4.21 Liouville’s Theorem.
If a function f(z) is analytic for all finite values of z, and is bounded, is
f  z  M
a constant. Since f(z) is bounded so ; where M is a positive
constant.
Let z1 z2 be any two points of z-plane.
Take the contour C to be a large circle, with its centre at origin, and
radius R, enclosing the points z1 and z2,
R  z1 and also  z 2
So that
By Cauchy’s integral formula, we have

208
1 f  z  dz
f  z1  
2i C z  z1

1 f  z  dz
f  z2  
2i C z  z 2
and
f  z1   f  z 2 
so that
1 f  z  dz 1 f  z  dz
2i C z  z1 2i C z  z 2
 

1 z1  z 2
  f  z  dz
2i C  z  z1   z  z 2 

 f  z1   f  z 2 

1  z1  z 2  f  z  dz
2i C  z  z1   z  z 2 

1  z1  z 2  f  z  dz

2i C  z  z1   z  z 2 
1 dz
 z  z 2 M. sin ce f  z   M
C  z  z1   z  z 2 
2
1 z1  z 2 M

 . dz sin ce z  R
2  R  z1   R  z 2 C

1 z1  z 2 M 


 . Rd sin ce z  Rei , dz  Rd.
2  R  z1   R  z 2 0


 z1  z 2  M which  0 as R  ,
2  R  z1   R  z 2 
f  z1   f  z 2  .
Hence
Since this holds for all values of z1 and z2, therefore f(z) is constant.
Note.
A function which is analytic in the whole of the z-place is called an
Integral function or Entire function.
4.22 Poission’s Integral formula.

209
z R
If f(z) be analytic in a region including the circle , prove that for
0 > r > R.
1 2 R 2  r2
F  re i
   2 
f Rei  d
2 0 R  2Rr cos       r
2

a  rei z R
where is any point of the domain .
z  R,
Since f(z) is analytic within and on the circle C for which hence
if a is a point within and this circle, then by Cauchy’s integral formula, we have
1 f  z  dz
f  a    1
2i C z  a

R2
z  R, a
The inverse point of a, with respect to the circle is and lies
f  z
 R2 
z 
 a 
outside the circle, therefore function is analytic within and on C;
hence by Cauchy’s theorem
1 f  z  dz
  0.   2
2i C  zR 2 
 
 a 

Subtracting (2) from (1), we have


R 2  za  a  z  a 
f  a  f  z  dz
 z  a   R 2  za 
1 R 2  aa
f  z  dz
2i C  z  a   R 2  za 

a  rei z  Rei ,
In this result put and then it becomes
1 2   R  r   f  Re  .i.Re d
2 2 i i

f  re  2i 0  Rei  rei   R 2  Rei .re i 


i

210
1 2 R 2  r2
f  Rei  d
2 0 R  rei  R 2  e  i 

  
1 2 R 2  r2
  f  Rei  d proved.
2 0
 R  2Rr cos      r  
2 2

f  rei   u  r,    iv  r,  
Writing
f  Rei   u  R,    iv  R,  
and
and equating real and imaginary parts, we get

1 2 R 2  r2
u  r,    u  R,   d   3
2 0 R 2  2Rr cos       r 2

1 2 R 2  r2
and v  r,    v  R,   d   4
2 0 R 2  2Rr cos       r 2

The formulae (3) and (4) express the real and imaginary parts of the
value of the function at any interior point in terms of the real and imaginary
parts of the values of the function at points on the bounding circle.
4.23 Taylor’s theorem.
If a function f(z) is analytic at all the points inside a circle C, with its
centre at the point a and radius R, then at each point z inside C,

fn  a 
f  z   an  z  a 
n
where a n 
n 0 n!

We have the identity

 z  a   ...   z  a    z  a 
2 n n 1
1 1 za 1
  
  z   a    a 2
   a
3
   a
n 1
   a   z
n 1

   z  a n   z  a n 1
  
 n 0    a  n 1     a  n 1   z
 
to be used soon.
C   a R; z  rei
Let z be a point within the circles and we write , so
za  r
that , where r < R.

211
Since z is a point within C, therefore by Cauchy’s integral formula we
have
1 f    d
f  z 
2i C   z

1    z  a    z  a  1 
n n

      f    d
2i C  n 0    a  n 1     a  n 1   z 

 
1
z
(Substituting for from the identity)

 z  a  f  d  1  z  a  1 f  d
n n

1
 
2i C n 0    a  n 1
 
2i C    a  n 1   z
 

 z  a  f  d  R
n
n
1
  
2i C 0    a  n 1
  n   1

1  z  a
n 1
1
Rn   f    d
2i C    a    z
n 1

where .
If n is finite. Then in result (1) integral and cigma can be interchanged
and so we have
n
1 1
f  z   z  a   f    d  R n
n

0 2i C    a  n 1

n
f n  a
  z  a  R n from  4.18 page155.
n

0 n!
n f n  a
  a n  z  a   R n taking a n 
n

0 n!

This is Taylor’s theorem with remainder Rn.


Further, we see that
1  z  a  f    d
n 1

2i C    a  n 1   z
Rn 
which may be written as
1  z  a f    d
n 1

 
2i C    a      a    z  a  
n 1

212
2  z  a  f    d
n 1

2i C    a  n 1   z
 Rn 

f    d
n 1
1 za

2i  a
C
n 1
 a  za
1 r n 1 M
2 C R n 1 R  r
 d

M  max . f    on C and z  a  r,   a  R
where
n 1 n 1
1 r M 1 r M
  
2  R  
R r C
d   
2  R  R r
2R

n 1
MR  r 
   . which  0 as n   sin ce r  R.
R rR 

 R n  0 as n  .

Hence when n is infinite, Rn vanishes and so we have from (1)


 z  a  f  d
n

1
f  z   
2i C 0    a  n 1
 

Further we see that the series


 z  a
n

0   a n 1 f   
 
n
 za
 n 1
f  
0 a


rn
  n M sin ce f     M on C
0 R
 n
 r r
 M    where  1
0 R  R

which is convergent, being G.P. with common ratio less than 1.

213
 z  a
n

0   a n 1 f   
 
Hence by M-test the series is uniformly convergent;
therefore term by term integration is valid; that is, integral and cigma may be
interchanged, and so we have
 z  a  f  d
n

1
f  z   
2i C 0    a  n 1
 

(integral and cigma interchanged)



1 f    d
  z  a 
n

0 2i C    a  n 1

f n  a
  z  a
n
as before
0 n!

f n  a
  an  z  a
n
where a n 
0 n!

f  z   an  z  a  .
n

0
Thus
This is Taylor’s theorem.
Particularly, If centre of the circle C is at the origin, then writing a = 0
in the above result we have

f n  0
f  z    a n z n where a n 
0 n!

This is Maclaurin’s theorem.

Remark.
For the truth of Taylor’s theorem it is not necessary that f(z) be analytic
on the boundary of the circle C; what is required is that f(z) should be analytic
inside the boundary.
4.20 Laurent’s Theorem.
If a function f(z) is analytic in the annulus (ring-shaped region) between
C
two concentric circles C and with centre at the point z = a and radii R and
R  R  R 
, then at any point z to the annulus

214
 
f  z    a n  z  a    bn  z  a  ,
n n

n 0 n 1

1 f    d
where a n     1
2i C    a  n 1
1 f    d
and bn     2
2i C    a  n 1

We have the identity

 z  a    z  a   ..   z  a    z  a 
2 n n 1
1 1 1
 
  z   z    a 2
   a
3
   a
n 1
   a   z
n 1

 n  z  a  n   z  a  n 1 1
   .
 n 0    a  n 1     a  n 1   z
 

Interchanging and z we have another identity

   a      a   ....     a      a 
2 n 1 n
1 1  1 
   
z   z  a  z  a 2  z  a 2  z  a  z a
n n
 z 

 n    a  n 1     a  n 1
  
 n 0  z  a  n   z  a  n z  
 
both to be used soon.
C
If z is a point in the region bounded by two concentric circles C and
with centre at z = a, then by Cauchy’s integral formula for multi-connected
region we have
1 f    d 1 f    d
f  z   2i C   z

2i C   z

1   n  z  a    z  a  1 
n n 1

   
 

  f    d
2i C  n 0    a      a    z 
n 1 n 1
 


 n    a      a 
n 1 n
1  
      f    d
C 
 n 0  z  a    z  a  z   
n n
  

215
 z  a  f  d  R  1 n    a  f  d  P
n n 1
1

2i C    a  n 1
  n 
2i C 1  z  a  n
  n

1  z  a  f    d
n 1

2i C    a  n 1   z
Rn 
where
1    a  f    d
n

2i C  z  a  n z  
Pn  .
and
Now we see that
1  z  a  f    d
n 1

2i C    a  n 1   z
Rn 

f    d
n 1
1 za

2i   a
C
n 1
a  za
1 r n 1 M d
2 C R n 1 R  r

M  max f    on C, and   a  R, z  a  r
where
1 r n 1 M
 2R sin ce  d  2R
2 R n 1 R  r C

n
Mr  r  r
   which  0 as n   sin ce  1
R r R  R

Pn  0 as n  .
Similarly
n 
Hence when , both Rn and Pn vanish and the last equation reduces
to
 z  a  f  d  1     a  f  d.
n n 1

1
f  z  
2i C 0    a  n 1
  
2i C 0  z  a  n
 

Further we see that the series

216
 z  a f 
n

 n 1  
0    a

n

za
 n 1
f  
0  a

rn
 M where f     M
0 R n 1
 n
M r 
  
0 R R 

r / R  1.
which is convergent, being a G.P. with common ratio
 z  a
n

0   a n 1 f   
 
Hence by M-test the series is uniformly convergent.
   a
n 1

0 z  a n f   
 
Similarly the series is uniformly convergent.
Therefore for both the series term by term integration is valid; that is, in
the above equation integral and eigma may be interchanged at pleasure. Thus
the last equation may be written as
1   z  a 1     a
n n 1

f  z  
2i 0 C    a  n 1
f    d   
2i 0 C  z  a  n
f    d


1 f    d  n 1
  z  a n 1  
z a    a  f    d
n 1
 
n

0 2i C    a  0 2i C

 
  a n  z  a   bn  z  a 
n n

0 1

 
f  z    a n  z  a   b n  z  a 
n n

0 0
Thus
1 f    d 1 f    d
2i C    a  n 1 2i C    a   n 1
an  and b n 
Where
This is Laurent’s theorem.
Remarks.

217
Since the integrals of the integrals in the formulae (1) and (2), which
give values of an and bn are analytic everywhere in the annulus hence by
Cauchy’s theorem path of integration can be taken any other concentric circle
C
say C0, lying between C and , then
1 f    d
   a  f    d  
 n 1
bn  
2i C0 C0    a 
 n 1
 a n .

1 f    d


f  z   an  z  a 
n an  .
2i C0    a  n 1

Hence we have where
4.5 Laurent’s expansion is unique.
Suppose that we have obtained in any manner or as the definition of f
the formula.

f  z   A  z  a
m
m R   z  a  R.
m 

Is the series necessarily identical with the Laurent’s series.


Sol.
If z is a point in the annulus between two concentric circles of radii R,
R  R  R 
with centre at z = a, then Laurents series as

f  z   a  z  a
n
n
n 

1 f  z
an   dz where z  a  r,  R   r  R 
2i C  z  a  n 1
where
Am  z  a 
m

1
  
2i m   z  a  n 1
dz


1
 Am  z  a 
m  n 1
 
2i m 
dz

1 
 Am   z  a 
m  n 1
 dz
2i m 

(cigma and integral are interchanged because term by term integration is


possible since the series is uniformly convergent)


1 

2i m 

A m  r m  n 1.e 
i m  n 1 
.riei .d 

218
sin ce z  a  rei

1  2
 
2 m
A m r m n  ei m n  d
0

1 
  A m r mn .0 when m  n
2 m

Because then all integrals vanish.


= 0.
1 2
an  A n  d when m  n
2 0
and
1
  A n .2   A n
2
 Am  a n .

Hence the given series is identical with the Laurent’s series.

4.6
Ex. 1.
z R
The function f(z) is analytic when and has the Taylor’s

a 0
n zn.
expansion Show that if r < R.

1 2
 
2

 
i 2
f re d   a n r 2n .
2 0
0

f  z   M when z  R.
Hence prove that, if

a
2
n r 2n  M 2 .
0

Sol.
z  r r  R
Since f(z) is analytic within the circle hence f(z) can be
z r
expanded in a Taylor’s series within the circle ; that is,

f  z    a n zn
n 0

219

  a n r n ein where z  rei
n 0


f  z    a m r m e  im
m 0
So that
f  z  f  z f  z
2

We know that
    
   a n r n ein   a m r m e im 
 n 0  m  0 

2 2      
f  z  d     a n r n ein   a m r m e im  d
2

0 0
 n 0  m 0 

2  2

2
a
n 0
n r n ein .d a m r m e im d.
0

(Since the two series on the right are absolutely convergent therefore
their product is uniformly convergent therefore their product is uniformly
0    2,
convergent for so term by term integration is justified) .
2
 0
eikd  0 for k  1
It is easy to see that
Hence all the integrals on the right hand side of the above vanish except
for which m = n.
2  2
f  z  d   a n ar 2n  d
2

0 0
n 0


  a n .r 2n .2.
2

n 0


1 2
 
2
 
i 2 n
or f re d   an r2
2 0
n 0

this proves the first result.



1 2
a f  z  d
2

2 n

n r2 
n 0 2 0
Now,
1 2 2
sin ce f  z   M
2 0
 M d

220
M2 2 M2

2 
0
d 
2
.2  M 2

a
2 n
i.e., n r2  M2.
n 0

Ex. 2.
za  R
If the function f(z) is analytic an done valued in , prove that
for 0<r<R.
1 2
f  a   P    e  i d,
r 0

P    a  re  i

where is the real part of .


Sol.
z  a  R,
Since f(z) is regular in therefore it must be regular in
z  a  r,  r  R 
.
Hence f(z) can be expanded in a Taylor’s series within the circle
za  r
,

f  z   am  z  a 
m

0
i.e.,
 m

  a m r m  ei  sin ce z  a  rei


0


f  z    a m r m  ei 
m
only amp changes.
0
So that
dz
 f  z.  z  a 
C
n 1
.
Now, consider the integral
dz 2  rei .id
 f  z.  z  a  n 1

0
a 0
m
m  im
r e . n 1 i n 1 
r e
C

 2
  a m r m  n .i  e  i m  n  d  0
0
0
for all valuesof n.

221
dz
 f  z  z  a 
C
2
 0.   1
Particularly
We know that
1 f  z  dz
f  a   formula 
2i C  z  a  2

1 f  z  f  z
dz from  1
2i C  z  a  2

1
2

f (a  rei )  f a  re i  re .id

i
 2 i2 
sin ce z  a  re i
2i 0
re
2
1 2Rf (a  rei ) i

2 
0
r 2 e 2i
re .id sin ce z  z  Re(z)

2
1
 P()e
 i
 d, [sin ce P()  Re f (a  re i )]
r 0

This proves the result.


Ex. 3.
z 
If a function f(z) is analytic for all finite values of z and as ,
f (z)  A z k

k
, then f(z) is a polynomial of degree .
Sol.
Since f(z) is analytic in the finite part of plane, therefore by Taylor’s
series

f (z)   a n z n , where z  R.
n 0

f (z)  M(r) z  1, (r  R)
Now, if max. on the circle , then by
M(r)
an  n
r
Cauchy’s inequality we have for all values of n.


 
A z
k

k
rn M(r)  f (z)  A z z 
since when ,

222
Ar k
rn z r
= (since on the circle )
 Ar k  n
where r is large.
r an  0
So when , the right hand side tends to zero if n > k. For
an
value of n > k. all the coefficients for which n > k become zero.
nk
Hence f(z) is a polynomial of degree n become

Ex. 4
If C is a closed contour around origin, prove that
2
a 
n
1 a e n az

   
 n!  2i C n!z
n 1
dz.

2

 an  1 2  2a cos  d
   
n  0  n!  2 0
e .
Hence deduce
Sol.
f (z)  eaz,  f n (z)  a n eaz .
Let
a n  f n (0)
So that putting z=0 in the above relation.
1 f (z)dz
 n!  n 1
2i C z
n
 an  1 1 a n eaz
2i n! C z n 1
or    . dz
 n! 
, multiplying by an on both sides.
2

 an  
1 1 a n eaz
   
n 1  n! 
 .  n 1 dz
n  0 2 i n! C z
Hence

223

1 a n eaz 1 a n dz
 
2i C 0 n!z n 1
dz 
2i C
e az
 n! . z n 1
1    a  n 1  dz
az 
e   
2i C  0  z  n! z
 

1 dz
 
2i C
eaz. ea / z
z
1 dz
  ea  z 1/ z  .
2i C z

1 2  2a iei d
2i 0
 e cos  .
ei

z  1, z  e i
where the circle C is taken as so on C.
1 2  2a cos 
2 0
 e .d

4.27. Some useful Results to remember.


We state and prove some results which may be directly used (where
necessary) while expanding given functions in a Taylor’s or Laurent’s series.

1  z z2  1 z n  1 z n 1 I 
n n 1
1
I.  1    .....   . 
z  a a  a a 2 an an z  a 

1 1  z z2 z n z n 1 1 
  1   2  ....  n  n . 
za a a a a a za

Proof:
They can be easily seen to be true by using elementary knowledge of
partial fractions.
dz dz
II.   0if k  1 and  z  a  2i
 z  a
k
C C
where C is the circle
za  r

dz dz
z  0,  k  1 , and   2i
C
k
C
z za  r
Particularly Where
Proof:

224
2
reii d

0
r k eik za  r
if C is the circle .
2
i 2  i 1 k  d
   ei 1 k   
r k 1 0
 e  k 1 
 0, if k  1.
  1  k  r 0
i
dz 2  re .id 2
C z  a 0 rei 0 d  2i
  i
and
dz dz
z
C
k
 0, k  1 and 
C
z
 2i if z  r.
Particularly
 1 
n
1 1 dz 
C z n 1 . z  a dz  a n 1 2i  C z  a  for n  0
III. (i)
1 1 2i
 z . z  a dz 
C
a
,
Particularly (it is when n=0 in above)
1 1 1  dz 
z C
n 1
.
za
dz  n 1  2i  
a  C
za
 for n  0
(ii)
1 1 2i
 z . z  a dz  
C
a
Particularly (it is when n=0 in above)
1
z
n 1
. dz  0.
C
za
(iii) except when n=0.
1
z
n 1
. dz  0.
C
za
(iv) except when n=0.
Proofs.
The proofs of all the four case of III are given below:

225
1 1
z
C
n 1
.
za
dz

1 1  z z2  1 n  1 z n 1 1 
n n 1

  n 1 . 1   2  .....  n z  .  dz u sin g I
C z
a a a a a n
za
 
 1 dz  1
n n 1
dz

a n 1 C z  a n 1 za
C

The other integral’s vanish, as they are of the form


dz
z
C
k
, k  1.

 1
n
 dz dz 
    
a n 1 C z C z  a 
 1
n
 dz  dz
 n 1  2i    sin ce   2i
a  C
za C
z
Particularly :
1 1 1 1  z z2 
C z z  a C z . a 1  a  a 2 .... dz
. dz 

1 dz 2i
a C z
  .
a

1 1
z
C
n 1
.
za
dz.
(ii)
1 1  z z2 z n z n 1 1 
C z n 1 a  a a 2
. 1    ....   .
a n a n z  a 
dz by I

1 dz 1 dz

a n 1 
C
 n 1 
z a A za

other integrals vanish as before


1  dz 
 n 1 
2i   
a  C
z a
Particularly :
1 1 1  z z2 
C z z  a
. dz    1   2  ...... dz
a  a a 

226
1 dz 2i
 
a z

a
1
z
n 1
. dz.
C
za
(iii)
1  z z2 
  z n 1 . 1   2  .... dz u sin g I
C
a a a 
n0
= 0 except when
1
z
n 1
. dz
C
za
(iv)
1 n 1  z z 2 
  z 1   2  ..... u sin g I
aC  a a 

= 0 except . or n=0
Note:
In (iii) and (iv) when n=0, they become respectively the particular cases
of (i) and (ii).
Ex 1.
z 2  5z  6
f (z) 
z2
If , does cauchy’s theorem apply.
(i) When the path of integration C is a circle of radius 3 with origin
as centre.
(ii) When C is a circle of radius 1 with origin as centre.
Sol.
z 2  5z  6
f (z) 
z2
Clearly f(z) is not analytic at z=2
z 3
(i) When C is the circle , the point z=2 lies inside, so f(z) is
not analytic within this circle; hence in this case Cauchy’s theorem does not
z 2  5z  6
C z  2 dz  0
apply i.e. .

227
z 1
(ii) When C is the circle , the points z=2 lies outside it, so that
f(z) is analytic within and on this circle; hence in this case Cauchy’s theorem
z 2  5z  6
C z  2 dz  0
does apply .
Ex. 2
 z  2  z  2
 z  1  z  4 
Obtain expansions for which are valid,
z 1
(i) When
1 z  4
(ii) When
z 4
(iii) When
Sol.

f (z) 
 z  2  z  2  1
1

4
 z  1  z  4  z 1 z  4
Here
The singularities of f(z) are z=-1 and z=-4
z 1
(i) When .
z 1
Then f(z) is regular within the region , hence by Taylor’s
expansion we have
z2 " zn n
f (z)  f (0)  zf (0)  f (0)  ....  f (0)  ....
'

2! n!
 n
f (0) n
 f (0)   z  (1)
n 1 n!

1 1
f (z)  1  
z 1 1 1 z
4
Now,
 
n  
f (z) 1 1 1
  1  
n 1
 .
  z  1 4  1  
n 1 n n 1
n!
 1  z  
  4  
So that

228
f n (0) n 1  1
   1 1  n  , and f (0)  1.
n!  4 
Hence substituting in (1) we have

f (z)  1    1
n 1
1 4  z
n n

n 1

1 z  4
(ii) When
z  1, z  4
Then f(z) is regular in the annulus , therefore by Laurent’s
series,

we have
 
f (z)   a n z n   b n z  n
n 0 n 1

1 f (z)dz
2i C z n 1
an 
where
1
bn  
2i C
z n 1f (z) dz
and
z r
where C is any circle where 1<r<4.

f (z) 
 z  2  z  2  1
1

4
 z  1  z  4  z 1 z  4
Here,
1 1  1 4 
2i C z n 1  z  1 z  4 
a n  1  dz

1    1  z z 2  1 z n 1 
n 1 n n 1
1 n 1 z n z
     
2i C z n 1  
n n
 1  1  z  z 2
 ....   1 z   1    1    ....   1  .  dz
z  1  4  4 42 4n 4 n z  4 
  
 1
n 1
 1 1 1 1 1 

2i  1  4
C
n 
  n  dz
 z z 1 4 z  4

(only the last two terms of each of the two series within the crooked
dz
C z m
m 1
brackets survive, since all the integrals of the form vanish when )

229
 1
n 1
 1 1 1 1 1 

2i  1  4
C
n 
  n dz
 z z 1 4 z  4 
 1
n 1
 1  dz 1 dz 
  1  n  4i    n 
2i  4  C
z 1 4 C z  4 

1  dz dz dz 
a0     .
2i  C z  1 C z C z  4 
and
dz 1
 z 1  0
C z4
But (by Cauchy’s theorem , because is analytic within
dz dz
C z  4  ' z  4
C
the contour C also [by Cauchy’s theorem where C’ is a small

circle with its centre at z=-1 and small radius ]
2 ei i d

0  ei z  1   or z  1   ei
[since ]
2i
= ,
 1   1
n 1 n 1
 1 
an   1  n  2i  2i  
2i  4   4n
and a 0  0
1  1 4 
And b n  
2i C
z n 1 1    dz
 z 1 z 1 
1  n 1 z n 1 z n 1 
2i  C
  z dz   dz  4  dz 
C z 1 C z4 

dz
z
n 1
dz  0
C
z m
m 1
Now being of the form ,
z n 1 z n 1
C z  4dz  0
z4
also (by Cauchy’s theorem, because is analytic
within the contour C)
z n 1 z n 1
C z  1 ' z  1
dz 
C
and [by Cauchy’s theorem where C’ is a small circle about
z=-1].

230
2  1   e  i n 1

 0 e i
. ei id

 1 e 
2
i n 1
 i  1
n 1
.d
0

i  1
n 1.2 

[ Expanding by binomial; all the integrals except first one


vanish].
 b n   1 .
n

 1
n 1
 
f (z)   z n    1 z  n
n
n
n 0 4 n 1
Hence
z 4
(iii) When
r   1/   ,
Put and the given function transforms into

F  
 1  2   1  2   1 
1

1
.
 1     1  4  1   1  4

1
F  
4
Thus function is analytic within the circle and as such can
be expanded in a Taylor’s series.

F     ann
0
Hence ( Taylors expansion)
Fn (0)
an 
n!
where
1 1
F()  1  
1   1  4
We have
F n ( )  1 4n 
  1 
n
  n 1 
  1     1  4  
n 1
n!

Fn (0)
 (1) n 1  4n  , and F(0)  1
n!
So that

231

F( )  1    1  1  4n   n
n

0 
. Replacing by 1/z, we have the

z  4 as f (z)  1    1
n
 1  4  1/ z .
n n

1
required expansion when
Ex.3
2z3  1
f (z) 
z2  z
For the function find,
(i) a Taylor’s series valid in the neighbourhood of the point z=i.
(ii) a Laurent’s series valid within the annulus of which centre is origin.
Sol.
2z3  8 1 1
f (z)   2  z  1  
z z
2
z 1 z
(i) Here Taylor’s series for f(z) in
f (z)   a n  z  i 
n

the neighbourhood of the point z=I is Taylor’s expansion


f  i
n
an 
n!
where
1 1
f (z)  2  z  1   ,
z 1 z
Now,
 1 1 
 f n (z)   1 n! 
n
 n 1 
  z  1
n 1
z 
except for n=0 and n=1
 1 1 
f n (i)   1 n! 
n
 n 1 
  i  1  i  
n 1

So that except for n=0 and n=1


f n (i)  1 1  i
a1   2   1   2
 3 ,
  i  1
n 1
n! i  2
and
1 1 i 3
a 0  f (i)  2  i  1     .
i 1 i 2 2
Hence
 i 3  i   1
n  1 
f (z)       3    z  i     1  n 1  
z  i
n

  1  i   i  
n 1
2 2  2 n 2

232
(ii) Singularities of f(z) are at z=0 , z=-1; so f(z) is regular in the domain
0  z  1.

0  z  1.
Hence f(z) can be expanded in the annulus in the Laurent’s
series.
 
f (z)   a n z n   b n z  n
0 1
Thus
1 1
an  
2i C z n 1
f (z)dz
where
1
bn  
2i C
z n 1f (z)dz
and
1 1  1 1
n 1  
an   2 z  1   dz
2i z  z 1 z 
so
1 1  1 

2i C z n 1  z  1 
dz
, except when n=0 and n=1, the other integrals
dz
C z m  m  1
vanish because they are of the form
1 1 1  1 
1  z  z 2  ....   1 z n   1 z n 1
n 1

n
 n 1   
2i C z  a  z  1 
 1
n 1
 dz dz 
   
2i  C z  1 C z 

1 dz
 z 1  0
z 1 C
But is analytic within C, so by Cauchy’s theorem.
dz

C
z
 2i
And .
 1
n 1

an  (1)2i  ( 1) n
2i
Hence

233
1 1 1 1
a0    2  z  1   dz
2i C z  z 1 z 
and
1 dz 1
 
2i C z

2i
.2i  1

1 1  1 1
a1 
2i C z 2  2  z  1  z  1  z  dz
and
1 dz 1 1
 2   2 . dz
2i C z C z z  1
other integrals vanish
1  1  1  
  4i   2 1  z  z .
2
 dz 
2i  C
z  z  1 

1  dz dz 
  4i     
2i  C
z C z  1
first integral vanishes
1
  4i  2i   1
2i
dz
 z 1
C
since vanishes as before Also
1 1  1 1 
bn   z n 1f (z)dz   z n 1  2  z  1   dz
2i C 2i C  z z  1 

z n 1
z 1 z 1
But, is analytic within the circle , hence by Cauchy’s
n 1
z
 z  1dz  0
theorem Also, all the other integrals vanish (being of the form
dz
 zk ,  k  1
, except for n=1.
1 dz 1
bn  0
b1   
2i C z 2i
.2i  0
Hence except for n=1 and

f (z)  a 0  a1z   a n z n  b1 / z
n 2
Hence

234

1
 1  z    1 z n 
n

n 1 z

Ex. 4.
z
f (z) 
 z  1  z  3
Represent a function by a series of positive and
0  z 1  2
negative powers of (z-1) which converges to f(z) when .
Solution.
z
f (z) 
 z  1  z  3
is regular except at z=1, z=3. Hence f(z) can be
0  z 1  2
expanded in a Laurent’s series in the annulus. in the positive and
negative power of (z-1)
z 1 3 1
f (z)    . .
 z  1  z  3 2  z  1 2 z  5
We may write
Since we want to expand the function f(z) in powers of (z-1) therefore
1

2  z  1
we need not care for the first term which is already expressed as a
3
2  z  3
negative power of (z-1). So it remains only to expand in the positive
and negative powers of (z-1).
1
(z) 
z 3
Let .
By Laurent’s Theorem, we have
 
1
  z    a n  z  1   b n  z  1 ,
n n

z 3 0 0

1  z
where a n  
2i C  z  1 n 1
dz

1
 z  1   z  dz.
n 1
and bn  
2i C

235
1 1
Now, a n   dz, sin ce   z   .
 z  3  z  1
n 1
C
z3

1 1
 
2i C t  i  2 
n 1
dt Putting z  1  t

1 1 1  t t2 t n t n 1 1 
 
2i C t
 
n 1 1 
 2 2 2
 2
 ...   .  dt
2n 2 n t  2  
1  1 dt 1 1 
   n 1   n 1  dt 
2i  2 C t 2 C t  2 

for all values of n.


the other integrals vanish as they are of the form.
dt
t
C
m
, m 1

1  1 1 1 
so, a n    n 1 .2i  n 1  dt 
2i  2 2 Ct2 

1 dt
t2
is analytic with C, hence  t2 0
C
But
[By Cauchy’s Theorem]
 1  1  1
an      n 1 .2i   n 1 .
 2i   2  2

1
 z  1   z  dz
n 1
bn  
2i C
And
1 1
 z  1
n 1
 
2i C z3
dz

1 1
 
2i C
 t n 1
t2
dt if z  1  t

1  1  t t2 

n 1
 t   1   2  ...... dt
2i C  2 9 2 

= 0 because all integrals are then of the form


dt
t
C
m
, m  1.

236
0  z 1  2
Hence when
1 3 
f  z    a n  z  1 , sin ce b n  0 for all values of n
n

2  z  1 2 0

1 3   1 
    n 1   z  1
n

2  z  1 2 0  2 
n
1 3   z 1
    .
2  z  1 4 0  2 

Ex. 5.
Find the Taylor’s or Laurent’s series which represent the function
1
 1  z   z  2
2

,
z  1. 1  z  2, z 2
i) when ii) when iii) when .
Sol.
We have
1 1 1 z2 
f  z     .
 1  z   z  2  5  z  2 1  z 2 
2

Now, we find below expansions of f(z) under specified values of z :


z  1.
i) When
1 1 1 z2 
f  z     .
 z  2   z  1 5  z  2 1  z 2 
2

We have
1 z
1
z 1
Since binomial expansion of the form is valid if ,
therefore partial fractions of f(z) are to be so arranged that the binomial
z 1
expansions involved may be valid for .
1
1 1 1  2z
 1  z2 
1
f  z   . 1  z  
5 2 2  5

z 1
Obviously binomial expansions are valid for

237
n
1  n z 2z 
  1      1 22n.
n

10 n  0 2 5 n 0

This represents series in positive powers of z, in other words, it is an


z 1
expansion of f(z) in a Taylor’s series within the circle .
1 z  2
ii) When .
1 1 z2 
f  z    , where 1  z  2.
5  z  2 1  z2 
We have
1 x
1

In order that the binomial expansions of the form be valid, it is


x 1
necessary that .
Hence the partial fractions of f(z) are to be so arranged that the binomial
1 z  2
expansions may be valid for .
1 1
1 1 1  z2 1  1
f  z   . 1  z   . 2 1  2 
5 2 2  5 z  z 
Thus
1  z  2.
which are obviously valid for
n n
1  n z 2z  n  1 
   1    2   1  2  .
10 n 0 2 5z n  0 z 
These are the expansions in positive and negative powers of z, that is, it
1  z  2.
is a Laurent’s expansion of f(z) within the annulus
z  2.
iii) When
1 1 1 z2
f  z  . 
5 z  2 5 1  z2
1 1
1 1 2 1 1  1
 . 1     z  2 . 2 1  2 
5 z z 5 z  z 

z  2.
arranged suitably to make the binomial expansions valid for

238
n
1 1 n 2 11 2   n 1
 .  1      2    1 2n
5 z  n  5  z z  n 0 z

2 z R
This is also Laurent’s expansion within the annulus where R
is large.
Ex. 6.
Find the Taylor’s and Laurent’s series which represent the function
z2  1
 z  2   z  3
z 2 2 z 3 z  3.
i) when ii) when iii) when
Sol.
We have
z2  1 3 8
f  z   1 
 z  2   z  3 z2 z3

Now we find below expansions of f(z) for specified values of z.


z  2.
i) When
3 8
f  z  1 
z2 z3

1 1
3 z  8 z
 1  1    1   binomials valid for z  2
2 2 3 3
3  8 
       1  z / 3
n n n
 1  1 z / 2 
2 0 3 0

which is a series in positive powers of z. In other words it is an expansion of


z  2.
f(z) in Taylor’s series within a circle
2  z  3.
ii) When
3 8
f  z  1 
z2 z3

239
1 1
3 z 8 z 
 1   1    1  
2 2 3 3

2 z 3
(arranged suitably to make the binomial expansions valid for )
n n
3  n 2 8 
 1  
n z
 1 
z 0
   
 1 
z 3 0
 3
which on the whole is a series in the positive and negative powers of z
2  z  3.
valid for It is therefore the expansion of f(z) in a Laurent’s series
2  z  3.
within the annulus
z 3
iii) when .
3 8
f  z  1  
z2 z3
1 1
3 z 8 z 
 1   1    1  
2 2 3 3

z 3
(arranged suitably to make the binomial expansions valid for )
n n
3  2 8  n 3
 1    1  z   z   1  z 
n

z 0 0

3 z  R
This is also Laurent’s expansion within the annulus where R
is a large.
Ex. 7 (i)
z3
f  z 
z  z  z  2
2

(a) Expand in powers of z;


i) z  1, ii)1  z  2, iii) z  2.
where
Sol.
z3 3 2 5
f  z     .
z  z  z  2
2
2z 3  1  z  6  z  2 

Expansions of f(z) for specified values of z are as below :

240
0  z  1.
i) When
2 2 5
f  z    
2z 3  1  z  6  z  2 

1
3 2 5  z
    1  z   1  
1

2z 3 12  2 

z  1.
arranged suitably for validity of binomial expansions when
n
3 2  5  z
     1 z n     .
n

2z 3 n 1 12 n 0  2 

This represents a series in positive and negative powers of z. In other


0  z  1.
words it is expansion of f(z) in Laurent’s series in the annulus
1 z  2
ii) When
3 2 5
f  z    
2z 3  1  z  6  z  2 
Then
1 1
3 2  1 5  z
   1    1  
2z 3z  z  12  2 

(arranged suitably to make the binomial expansions valid when


1 z  2
).
n
3 2  n 1 5  z
     1 n     .
2z 3z n 1 z 12 n 0  2 

which being a series in positive and negative powers of z, is a Laurent’s


1  z  2.
expansion in the annulus
z 2
iii) When
1 1
3 2  1 5  2
f  z     1    1  
2z 3z  z  6z  z 
Then
z 2
arranged suitably to make the binomial expansions valid when

241
n
3 2  n 1 5  2
     1 n     .
2z 3z n 1 z 6z n  0  z 

which being a series in positive and negative powers of z is Laurent’s


2 z R
expansion within the annulus where R is large.
Ex. 7 (ii)
4z  3
f  z 
z  z  3  z  2 
(b) Represent the function in Laurent’s series.
z  1.
i) within
z  2 and z  3.
ii) in the annular region between
z  3.
iii) exterior to
Sol.
4z  3 1 1 1
f  z    
z  z  3  z  2  2z 2  z  2  z  3

This function is analytic everywhere except at


z = 0, z = -2, z = 3.
0  z  1.
i) When
Then we write f(z) as
1 1 1
f  z    
2z  z  z
4 1   3 1  
 2  3
1 1
1 1 z 1 z
  1     1  
2z 4  2  3 3
n n
1 1  n z 1  z
     1      
2z 4 0 2 3 0 3

1  n 1 1 1 
    1 n2
 n 1 z n .
2z 0  2 3 

This being a series, in positive and negative powers of z, is Laurent’s


expansion in

242
0  z  1.
the annulus
2  z  3.
ii) When
Then we write f(z) as
1 1 1
f  z    
2z  2  z
2z 1   3 1  
 z  3
1 1
1 1  2 1 z
   1    1  
2z 2z  z 3 3
n n
1 1  n 2 1  z
    1       .
2z 2z 0  z  3 0 3

This being a series, in positive and negative powers of z, is Laurent’s


2  z  3.
expansion in the annulus
z 3
iii) When .
Then we write f(z) as
1 1 1
f  z    
2z  2  3
2z 1   z 1  
 z  z
1 1
1 1  2 1 3
   1    1  
2z 2z  z  z z
n n
1 1  n2 1  3
     1       .
2z 2z 0 z z 0 z

This being a series, in positive and negative powers of z, is Laurent’s


3 z R
expansion in the annulus where R is large.
Ex. 8. (i)
1
 z  1  z  3
(a) Find different developments of in powers of z;
according to the position of the point z in the plane. Expand the function in a
Taylor’s series about z = 2, and indicate the circle of convergence.
Sol.

243
1 1 1 1 
f  z     
 z  1  z  3 2  z  3 z 1
Here
Thus function is regular every-where except at z =1, z = 3. We give
below different expansions of f(Z) according to the positions of z in plane.
0  z  1.
i) When
1 1 1 
f  z    
2  z  3 z 1 
We have
1
1  z  1
 1     1  z 
1

6  3 2

0  z 1
arranged suitably in order to make binomial expansions valid for )
n
1  z 1 
        z
n

6 n 0  3  2 n 0

1 1 
  1  n 1 z n
n 0 2  3 

0  z 1
which is obviously Taylor’s expansion of f(z) valid in .
1  z  3.
ii) When
1 1 1 
f  z  
2  z  3 z  1 

Then
1 1
1 z 1  1
  1     1  
6 3 2z  z 
arranged suitably
n n
1  z 1  1
     
6 0  3  2z 0  z 

which is obviously a Laurent’s series in the positive and the negative powers of
1  z  3.
z within the annulus
z  3.
iii) When

244
1 1 1 
f  z  
4  z  3 z  1 

Then
1 1
1  3 1  1
 1    1  
2z  z  2z  z 
n n
1  3 1  1
    2z 
2z n 0  z 
  
n 0  z 

1 
 .  3n  1 1/ z n 1
2 0

which is obviously a Laurent’s series, in negative power of z, with the annulus


3  z  R.

Last part. If centre of a circle is at z = 2, then distance of both the


singularities z = 1 and z = 3 from this centre is 1. Hence if a circle is drawn
with centre at z =2 and radius 1, then within this circle, i.e., within the circle
z  2  1,
the function f(z) is regular, hence it can be expanded in a Taylor’s
z  2  1,
series within the circle which is therefore the circle of convergence.
1 1 1
f  z   
 z  1  z  3 z 2
 4z  3  z  2
2
1
Now,
z  2  1.
expansion is now valid for

   z  2 
2n

n 0

which is obviously a Taylor’s series in positive powers of (z - 2) within a circle


z  2  1.

Ex. 8. (ii)
1
f  z 
z  3z  2
2
(b) Expand in the series the function in the regions.
 i z  1,  ii  1  z  2,  iii  z 2

Sol.

245
1 1 1 1
f  z    
z  3z  2  z  1  z  2  z  2 z  1
2

clearly f(z) is analytic


everywhere except at z =1 and z =2.
z 1
i) When .
1 1
f  z  
z  2 z 1
Thus we write
1 1
 
1 z  z
2 1  
 2
1
1 z
  1  z
1
 1  
2 2
 n
1  z
 z   
n

0 2 0 2

Using binomial theorem.



 1 
  1  n2  zn
0  2 

This being a series in positive powers of z is Taylor’s expansion within


z  1.
the circle
1  z  2.
ii) When
1 1
f  z  
z  2 z 1
Then we write
1 1
 
 z  1
2 1   z 1  
 2  z
1 1
1 z 1 1
   1    1  
2 2 z z
n n
1  z 1  1
      
2 0 2 z 0 z

This being a series in positive and negative powers of z is Laurent’s


expansion in

246
1  z  2.
the annulus
z  2.
iii) When
1 1
f  z  
z  2 z 1
Then we write
1 1
 
 2  1
z 1   z 1  
 z  z
1 1
1 2 1 1
 1    1  
z z z z
n n
1   2 1  1
     
z 0 z z 0 z

1
   2n  1 n 1
0 z

This being a series in negative powers of z is Laurent’s expansion in the


2 z R
annulus where R is large.
Ex. 9.
1
f  z 
z  z  1  z  2
2 2

Express in a Laurent’s series in the region


5 7
 z  .
4 4
Sol.
1
f  z 
z  z  1  z  2
2 3

1  3 1 3 3 1 
      
z  r  1  z  1 2  z  2   z  2  2  z  2  3 
 

Partial fractions.
It is clear that f(z) is analytic everywhere except at z = 0, z = -1, z = -2.

247
5 7
 z 
4 4
The region is part of the annulus between two concentric
z  1 and z  2
circles within which f(z) is analytic; hence we can express it
in Laurent’s series at any z within the annulus.
As seen above we have

1  3z  2 1 
f  z  
z   z  1 2
  3z 2
 15z  17  
 
3
 z  2 

1  3z  2  1  1 z  
2 3

   1     3z  15z  17   1   
2

z  z  z 8  2  

1  z  2.
both binomials are valid since
 3 2      1  n  1  n  2   z  
n 1 n
1 1 17
     2    1  n  1 n   3z   15  
n
  
  z z  0 z 8 z  0 2n  2  

   1  n  1  n  2  n
n
3 2   1 1 17
   2    1  n  1 n   3z   15  
n 1
z
z z  0 z 16  z  0 2n

This being a series in positive and negative powers of z is Laurent’s


1  z  2;
expansion in the annulus or we may say as desired in the region
5 7
 z  .
4 4
Note:
1 z  2
The region given in the question is a part of the annulus ;
5 7
 z  .
4 4
hence above expansion is true for the region
Ex. 10.
a zb
Show that in the annulus defined by the function
1/ 2
 bz   a n zn 
  S0   Sn  n  n ,
  z  a   b  z   z b 
can be expanded in the form

248
1.3.5......  2l  1 1.3.5...  2l  2n  l   b 
 l

Sn   .  
i0 22l  n  l  !  l  n ! a
where
Solution:
1/ 2
 bz 
f (z)   
  z  a   b  z  
is regular except at z=a, z=b, hence it can be
a z b
expanded in a Laurents series in the annulus .
 
f (z)   a n z n   b n z  n
0 1
So
1 f (z) dz
an 
2i C z n 1
Where
1
bn 
2i C
z n 1f (z) dz
and
1/ 2
 
 bz 
1/ 2
 1 
f (z)     
  z  a   b  z    1  z   1  a  
  b   z  
1/ 2 1/ 2
 z  a
 1   1  
 b  z

The binomials are put in this form so that their expansions may be valid
a zb
for .
1/ 2
 z 1.3.5..(3m  1) z m
1    . m
 b 2 m m! b
The binomial
1 l
 a 1.3.5..(2l  1)  a 
1      
 z 2l l! z
and
1 f (z)dz
an  
2i C z n 1
Now by definition

249
1   1.3.5.....  2m  1 z m  1.3.....  2l  1 a l  dz
 . m 
2i C  m 0
 . l  n 1
2m l! a l0 2l l! z z
1 2    1.3.5.....  2m  1 R m eim    1.3.....  2l  1 a l  Rei id
 . m   
2i 0  m 0
 . il  m 1 i n 1 
2m m! b   l0 2l l!R l e  R e

2  i m l n 
 
0
e d,
The terms on the right are of the form and value of such
2

0
d,
m  ln
a integral is zero except when it becomes i.e. except when .
Therefore collecting only these terms whose integrals are not zero, we
have
1 2    1.3.5...  2l  2n  1 1.3.5...  2l  1 l 

2i 0  l 0 2l  n  l  n  !bl  n
an  . a  d
2l l! 

1 1 2    1.3.5...  2l  2n  1 1.3.5...  2l  1  a  
l


b 2 0  l 0
 n .    d
2l  n  l  n  ! 2l l!  b  

1
 .Sn
bn
by hypothesis
a 0  S0
Particularly (Putting n=0 on both sides).
b n  a n .Sn
Similarly
 
f (z)  a 0   a n z n   b n z  n
1 1

Hence

 a n zn 
 S0   Sn  n  n 
1 z b 

Ex.11.
  1 
sin c  z   
  z 
Show that can be expanded in a series of the type
b1 b 2
a 0  a1z  a 2 z 2  .....    ....
z z2

250
z n and z  n
In which the coefficients both of are
1 2
sin  2c cos   cos n d
2 0
.
Sol.
  1 
f (z)  sin c  z   
  z 
The function is analytic except at z=0. Hence
  1 
sin c  z   
  z  r z R
is analytic in the annulus where r is small and R is
large.
  1   
sin c  z      a n z n   c n z  n
  z  0 1
Hence by Laurent’s theorem
1   1   dz
an 
2i C
sin c  z    n 1
  z  z
Where
1   1 
bn 
2i C
sin c  z    z n 1 dz
  z 

where C is any circle in the annulus with origin as centre.


1   1   dz
Now an 
2i C
sin c  z    n 1
  z  z
i
1 2  i n ie d
 
2i 0
 sin 2c cos  e
ei n 1 

z 1 2  ei
taking the circle C as on which
1 2
  sin  2c cos   e  i n d.
2i 0

1 2
sin  2c cos   {cos n  i sin n d} d
2i 0

1 2
sin  2c cos   cos n d
2 0

2 2
 sin  2c cos   sin n d  F() d
0 0
[The second integral , write it as
F(2  )   F()
vanishes by the property being satisfied]

251
1 2
sin  2c cos   cos( n)d
2 0
bn  a n 
And
1 2
sin  2c cos   cos n d  a n
2 0

  1   
sin c  z      a n z n   b n z  n
  z  0 1
Hence
1 2
sin  2c cos   cos n d.
2 0
a n  bn 
where

252
Ex. 12

 1  1 
cosh  z    a 0   a n  z n  n 
 z 1  z 
Prove that
2
1
cos n cosh  2 cos   d.
2 0
an 
where
Sol.
 1
f (z)  cosh  z  
 z
The function is analytic every where in the finite
r R R
part of the plane except at z=0; i.e. it is analytic in the annulus
where r is small and R is large. Hence f(z) can be expanded in Laurent’s series
r z R
in the annulus .
 1  
cosh  z     a n z n   b n z  n
 z 1 1
So
1  1  dz
an 
2i C
cosh  z   n 1
 zz
where
1  1
bn  
2i C
cosh  z  z n 1dz
 z
and
where C is any circle lying in the annulus with origin as centre.
Now,
1  1  d
an  
2i C
cosh  z   n 1
 zz
1 2 i eid
 
2i 0
 cosh 2cos 
ei n 1 

z 1 z  e i
taking C a circle on which
1 2
  cosh  2 cos   e  ind
2 0

1 2
cos  2 cos    cos n  i sin n  d
2 0

253
1 2
cosh  2 cos   cos n e
2 0

the other term vanishes;
2 2
 cosh  2 cos    sin n d  F() d
0 0
[the other integral write it as , vanishes
by the property
F  2      F()
being satisfied]
1 2
cosh  2 cos   cos( n)d
2 0
bn  a  n 
and
1 2
cosh  2 cos   cos nd
2 0

 an

 1  
cosh  z     a n z n   b n z  n
 z 0 1
Hence
 
  a n zn   a n zn
0 0
bn  a n
since here
 
 a 0   a n zn   a n zn
1 1


 1 
 a 0   a n  zn  n 
1  z 

Ex. 13
1 
c[z  (1/ z)] 1 2
e2  a z
n 
n
n
an 
2 0
cos  n  c sin   d
Show that where
Sol.
1
c[z  (1/ z)]
e 2
z
The function f(z)= is analytic except at z=0 and , i.e.
r z R
f(z) is analytic in the annulus where r is small and R is large Hence
by Lauret’s Theorem.

254
1 
c[z  (1/ z)]
e2   a nzn


 a
n 
n zn

1
1 c[z  (1/ z)] dz
an 
2i C
e2
z n 1
where
where C is any circle lying the annulus with origin as centre.
1
1 c[z  (1/ z)] dz
an 
2i C
e2
z n 1
We have
i
1 ic sin  i d

2i C
 e
ei n 1 

z 1 z  ei 
taking C a circle on which
1 2   i n csin 
2 0
 e d

1 2
2 0
 [cos(n  c sin )  i sin(n  e sin )]d

1 2
2 0
 [cos(n  c sin ) d
the other term vanishes,
2 2
 0
sin(n  c sin ) d  0
F(d)
[the other integral denote it by , vanishes by the
F  2  i   F()
property being satisfied].
Ex.14
b 2 b3
e(u / z)  yz  a 0  a1z  .....    ...
z z
Show that or where
1 2   u  v  cos 
2 0
an  e .cos{(v  u)sin   n}d

1 2   u  v  cos 
2 0
bn  e .cos{(u  v)sin   n}d

Sol.

255
f (z)  e u  v   vz
is regular except at z=0, hence it can be expanded in
r z R
Laurent’s series in the annulus where R is large, and r is small. So
we have
 
f (z)   a n z n   b n z  n
0 1

1 f (z) 1
an  
2i z
C n 1
dz, b n 
2i C
z n 1f (z)dz,
where
where C is any circle whose centre is at the origin and whose radius is
(r<radius<R).
z 1 z  e i
Let us take the contour circle so that on this circle C,
e 
u / z  vz
1
2i C z n 1
an  dz
Now

1 2 e
 ue i ve  i


2i 0
e i  n 1 
ei .id
since
z  e i
1 2   u  v cos  ei  vsin  u sin  n
2 0
e d

1 2
  e u  v  os  cos  vsin   u sin   n   isin  vsin   u sin   n   d
2 0
F(2  )  F()
[The second integral vanishes, since the property of the
definite integral is satisfied].
bn  a  n
1 2   u  v cos 
cos{ v  u  sin   n}d
2 0
 e

1 2   u  v cos 
cos{ u  v  sin   n}d
2 0
 e

Ex. 15.

 
a z
z  c3 / 2z 2 n
e n
n 
Show that if c>0, then

256
e 
 1/ 2 c
2  cos  cos   cos{c sin  1  cos   n}d
2

an   e  
2c n 0
where
Sol.

z  c3 / 2z 2 
f (z)  e
is regular except at z=0; therefore it can be expanded in
the annulus r<z<R in a Laurent’s series, where r is small and R is large.

  1 f (z)
a z
z  c3 / 2z 2

2 C z n 1
n
e n an  dz
n 
Thus where and C is a circle
z c z  ce i
with centre at the origin, and radius c, so that .
Now
1 f (z)dz
an  
2i C z n 1
dz

1 e

z  c3 / 2z 2 
2i C z n 1
 dz

 c 
exp  cei  e 2i 
1 2
 2 ceii d

2c n  0 n 1 i  n 1 
c e
as z  cei

1  c
2 

2cn 
exp  cei  e 2i  in d
 0 2 
1 2  c   c 
n 0
 exp  c cos   cos 2  .exp i  csin   sin 2  n  d
2c  2   2 
1 2  1    1   1 

2cn 
exp  c cos   c cos 2   cos  csin   csin 2  n   isin  csin   csin 2  n  d
0
 2    2   2 
1 2   1   c 
2cn 0
 exp  ccos   c cos 2   cos  csin   sin 2  n  d
 2   2 
t
F  2      F()
he second integral vanishes on account of the property.
being satisfied.
1  1 
exp  c cos   c cos 2   cos  c sin   1  cos    n d
2
an 
2cn  0
 2 
e c / 2
 
exp c  cos   cos 2    cos  c sin 1  cos )  n d
2

2cn  0

Ex. 16.

257
z 1
Determine a function which shall be regular within the circle and
shall have on the circumference of this circle the value
a 2
 1 cos   i  a 2  1 sin 
a 4  2a 2 cos 2  1 a2  1 
Where , and is the vectorial angle at points
on the circumference.

f (z)   a n z n
0
Taylor’s Expansion
1 f (z)dz
an  
2i C z n 1 z 1
where where C is the circle
Given that

f (z) 
a 2
 1 cos   i  a 2  1 sin 
on z  1
a 4  2a 2 cos 2  1
a 2  cos   i sin     cos   i sin  

a 4  a 2  e 2i  e2i   1
a 2 ei  e i

a 4  a 2  e 2i  e 2i  1
1
a 2z 
 z
sin ce on z  1, z  ei
 2 1
a  a  z  2  1
4 2

 z 
 1
z  a2  2 
 z  z
  2 2
 2 1 2 2 a z
a  2  a  z 
 z 

1 z 1 1 1 1
an  
2i a  z z
C 2 2
. n 1 dz  
2i a  z z
C 2 2
. n dz

1 2  1 1

2i 0 a e e
2 i
. in .ei .id sin ce z  e i
1
1 2   i n 1   e 2i 
2i 0
 e 1  a  d
 
 2i  n 1   / 2
1 2  i  n 1 
 e 2i e 4i e 

2a 2 0
e 1  2  4  .....
 a a a n 1
 .... d


258
1 2 1
an 
2a 2 
0 a n 1
d
So if n is odd, all other integrals vanish being of
2

ik
e d k  0.
0
the form
1 1 1
. n 1 2  n 1 an  0
2a a
2
a
and if n is even, because then all the
integrals vanish.

f (z)   a n z n
0
Hence ( n odd.)

1
 n 1
zn
0 a

1  z z3 z5 
   3  5  ......
a a a a 
even values of n do not
contribute
z  z2 z4 
 2 
1  2  4  ......
a  a a 
1
z  z2 
 2 1  a 2 
a  
z 1 z
 2.  2
a z 2
a  z2
1 2
a

Ex. 17
Find the function f(z) which is analytic throughout the circle C and its
interior, whose centre is at the origin and whose radius is unity and has the
value.
a  cos  sin 
i 2
a  2a cos   1 a  2a cos   1
2

where a>1 and is the vectorial,
angle at points on the circumference C.
Sol.
z 1
Since f(z) is analytic inside and on the circle , therefore f(z) may
be expanded in a Taylors series at any point inside and on the circle,

259

f (z)   a n z n
n 0
i.e.
1 f (z)
an  
2i C z n 1
dz z  1, z  ei
where on the circle C,
a  cos   i sin 
f (z)  z 1
a 2  2a cos   1
Given on the circle
a  e  i a   1/ z 
 2
a  a  e  e  1
2 i  i
a  a  z  1/ z   1


 a  1/ z  
1
.
 a  1/ z   a  z  a z
f (z)
an  
C z n 1

1 1
 . dz
C a  z z n 1

1 2 1 1

2i 0 i
a e e
. i n 1  .ei .id sin ce z  e i on C
1
1 2 ei 
0
 in
 e  1   d
2a  a 
1 2   in  1 i 1 2i 1 

2a 0
e 1  e  2 e ...  n ein  ... d
 a a a 

1 2 1
2a 0 a n
d
= , other integrals vanish
1 2 1
. n  n 1
2a a a
 n
1  z
1
f (z)   n 1 z     n

0 a a 0 a
Hence
1
1 z 1 1
1    
a a  z  a z
a 1  
 a

260
Ex.18
1
z  .
2
Expand cos z about the point
Sol.
f (z)  cos z.
Here
1
z 
2
By Taylor’s Theorem expansion of f(z) at is
1 
n f n  
 1 
where a n  
2 
f (z)   a n  z   
 2  n!

f (z)  cos z
Now
 1 
f n (z)  cos  z  n 
 2 
1  1 1 
 f n     cos    n 
2  2 2 
1 
  sin  n 
2 
1 1 
 a n   sin  n 
n!  2 
 a 2  a 4  a 6 ....  0.

a0  1
But
1
 sin n
cos z  1   2
n!
Hence
where n is odd integer.

261
NOTES

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

262
UNIT - V
COMPLEX INTEGRAL CALCULAS
The zeros of analytic function.
A zero of an analytic function f(z) is a value of z such that f(z)=0 .
If f(z) is analytic in the neigbourhood of a point z = a, then by Talyor’s
theorem

f n (a)
f (z)   a n  z  a 
n
where a n 
n 0 n
i.e., f (z)  a 0  a1 (z  a)  a 2 (z  a) 2  .....  a n (z  a) n  ....

a 0  a1  ........a m 1  0, but a m  0
If then f(z) is said to have a zero of
order m at z = a, The zero is said to be simple if m = 1
f n (a)
am  ,
n
It may also be seen that since hence at a zero of order m,
we have
f (a)  f ' (a)  .......  f m 1 (a)  0, but f m (a)  0.

Thus in the neighbourhood of the zero z = a of order m the function f(z)


is
f (z)  a m (z  a) m  a m 1 (z  a) m 1  ....
 (z  a) m  a m  a m 1 (z  a)  .....
 (z  a) m (z)
Where (z)  a m  a m 1 (z  a)  ....

is analytic and non-zero at and in the neighbourhood of z = a.


(z) am
As a matter of fact is equal to at z = a.
Singularities of an analytic function
Define ‘pole’ and ‘singularity’ and discuss different kinds of the latter.
Singularities.
A singularity of a function f(z) is a point at which the function ceases
to be regular (analytic).
Isolated and non-isolated singularity.
If z = a is a singularity of f(z) and if there is no other singularity within
a small circle surrounding the point z = a, then z = a is said to be an isolated
singularity of the function f(z); otherwise it is called non-isolated.

263
z 1
f (z) 
z(z  2)
For example the function is analytic every where
except at z = 0 and z = 2. Thus z = 0, z = 2 are the only singularities of this
function; so there are other singularities of f(z) in the neighbourhood of z = 0, z
= 2; therefore z = 0, z = 2 are the isolated singularities of this function.
1

tan  
2
Consider another function . This function is not analytic at
 
tan    n,
2 2
the points where = 0, i.e. at the points i.e. at the points
1 1 1
z   n  1, 2,3,... z  1, z  , z  ,......., z  0
n 2 3
. Thus are the singularities of
the function all of which are isolated except z = 0 because in the
z 1
n
neighbourhood of z = 0 there are infinite number of other singularities
where n is large. Therefore z = 0 is the non isolated singularity of the function
1

tan  
2

Kinds of singularity.
Consider a function f(z) which is analytic within a domain D, except at
the point z = a, which is an isolated singularity. Now draw at a circle C with its
centre at z = a and radius as small as we please, and another large concentric
circle C of any radius R lying wholly within the domain D. Now in the annulus
between these two circle f(z) is analytic, hence if z is any point of this annulus,
then by Laurent’s theorem, we have
 
f (z)   a n (z  a) n   b n (z  a)  n , 0  z  a  R 
n 0 n 1

 n 
 b n (z  a) 
 n 1 
The term on the right hand side is called the principal
part of f(z) at the isolated singularity z = a.
In relation to the principal part there are three possibilities:-
(i) All the coefficients bn may be zero, i.e. no term in the principal
part
(ii) Infinite number of terms in the principal part.
(iii) Finite number of terms in the principal part.

264
These above three cases give rise to three kinds of singularities:
(i) removable singularity. When all the coefficients bn are zero.

f (z)   a n  z  a 
n

n 0
za  R
We have then which is analytic for , except
(z)
at the point z = a. Let be the sum function of the power series

a  z a
n

n 0
n
(z)
. Now differs from f(z) only for z = a where there is
singularity. To avoid this singularity z = a, we can suitably define f(z) at z = a,
so that, we have
(z)  f (z) for 0  z  a  R
 a 0 for z  a

This type of singularity which can be made to disappear by defining the


function suitably, is called Removable singularity. This is rather artificial sort
of singularity and is of little importance in the theory of functions.
sin(z  a)
za
For example function has removable singularity at z = a
because

1   z a  z  a  z  a 
3 5 7
sin(z  a)
  z  a      ....
za z  a  3 5 7 
 z a  z a  z  a
2 4 6

 1    ...
3 5 7

 z  a
have no term containing negative powers of .
However, this singularity can be removed and the function made
analytic by defining
sin(z  a)
 1 at z  a
za
(iii)Essential singularity (or isolated essential singularity). If there are
infinite number of terms in principal part, that is, if the series
b1 (z  a) 1  b 2 (z  a) 2  .....
does not terminate, then z = a is called an isolated
essential singularity of the function f(z).

265
 1 
sin  
 za 
For example the function has essential singularity at z = -a
because
 1  1 1 1
sin     
 z  a  z  a 3(z  a) 5(z  a)5
3

…….. has infinite number of the


terms in negative powers of (z-a)
(ii) Pole. If there are finite number of terms in the principal
part; that is, if the series
b1 (z  a) 1  b 2 (z  a) 2  ....  b m (z  a)  m

consists of finite number of terms, then z = a is said to be a pole of order m of


the function f(z).
When m = 1, the pole is said to be simple pole.
sin(z  a)
 z  a
4

For example the function has a pole at z = a because

  z  a  z  a  z  a 
3 5 7
sin(z  a) 1
  z  a      ....
 z  a  z  a
4 4
 3 5 7 

 z  a  z  a
3
1 1
     ...
 z  a
3
3 z  a  5 7

has finite number of terms (here first two terms only) in negative powers of (z-
a).
we have seen above that in the neighbourhood of the pole z = a of order m the
function

   b b2 bm
f (z)    a n (z  a) n   1   ...... 
 z  a  z  a  z  a
2 m
 n
1   n  
 m  
a (z  a) n  m    b m  b m 1 (z  a)  ......  b1 (z  a) m  
(z  a)  0  
1
 (z)
(z  a) m
 
Where (z)    a n (z  a) n  m   b m  b m 1 (z  a)  ......  b1 (z  a) m  
 0 

266
(z)  b m as z  a (z)
Obviously ; hence if analytic in the
neighbourhood of the pole z = a.
f (z)   as z  a
If f(z) has a pole at z = a, then .
Suppose z = a is the pole is of order m, then by Laurent’s theorem
 m
f (z)   a n (z  a) n   b n (z  a)  n
n 0 n 1

b1 b2 bm
  a n (z  a) n    .... 
n 0 z  a  z a 2
 z  a
m


1
  a n (z  a) n   b m  b1 (z  a) m 1  b 2 (z  a) m 2  .....  b m 1 (z  a) 
 z  a
m
n 0

The expression within the square brackets on right hand side tends to
b m as z  a
, so that the whole right hand expression tends to infinity.
1
f (z)
If an analytic function f(z) has a pole of order m at z = a, then has
a zero of order m at z = a; and conversely.
If f(z) has a pole of order m at z = a, it can be written as
1
f (z) 
 z  a
m
(z)

(z)
where is analytic and not zero at z = 0.
1
  z  a  (z)
m

f (z)
So that
1
f (z) (a)  0
Obviously has a zero of order m at z = a, since
Zeros are isolated.
Let z = a be a zero of order m of the function f(z) which is analytic in
f (z)   z  a  (z), where (a)  0
m

the neighbourhood of z = a, then


(z)
Now is analytic and non-zero in the neighbourhood of and at z =
 z  a   0 for values of z  a.
m

a. Also

267
So there exists no other point in the neighbourhood of z = a at which
f(z) vanishes. Hence the zero z = a is isolated. The same is true for every zero
of f(z). Hence all the zeros of f(z) are isolated.
Poles are isolated.
Let z = a be a pole, of order m, of the function f(z); then in the
neighbourhood of z = a we have by Laurent’s expansion

b1 b2 bm
f (z)   a n (z  a) n    ...... 
n z  a  z  a 2
 z  a
m

1  
 m  n
a (z  a) n  m   b m  b m1 (z  a)  ......  b1 (z  a) m  
(z  a)  0 
1
 (z)
(z  a) m
   
Where (z)    a n (z  a) n  m    b m  b m 1 (z  a)  ......  b1 (z  a) m  
 0  

(z)
Obviously does not tend to infinity for any finite values of z near z
= a.
In other words there is no pole in the neighbourhood of z = a. Hence
the pole z = a is isolated; proving that every pole is isolated.
Limit point of zeros.
The limit point of the zeros of a function f(z) in an isolated essential
singularity.
Let z1,z2,z3,….be an infinite set, of zeros of f(z), which necessarily must
have at least one limit point. Suppose z0 is their limit point which may or may
not be a point of the set.
If z0 is a point of the set, it must itself be a zero of f(z); and, in
addition, by virtue of being the limit point, should have in its neighbourhood a
cluster of zeros, which goes against the zero around it (zeros are isolated).
Hence z0 cannot be a zero of f(z) unless the function is identically zero in the
domain D.
If the function f(z) does not vanish identically in D, then z 0 is not a zero
of f(z), while being surrounded by many zeros. This shows that z 0 us a
singularity. The singularity is not a pole since f(z) does not tend to infinity in
the neighbourhood of z0 (as a matter of fact it tends to zero). Therefore z 0 is an
essential singularity. But the singularity is isolated since in the neighbourhood
of z0 the function f(z) is analytic (tending to zero every where in the
neighbourhood). Hence z0 is an isolated essential singularity.
Remark:

268
If we are to prove that a certain point z0 is an isolated essential singular
point of f(z), it will do if we prove that z0 is a limiting point of zeros of f(z).
Limit point of poles
The limit point of the poles of a function f(z) is a non-isolated essential
singularity.
Let p1,p2,p3,……be an infinite set, of the poles f(z), which necessary must have
at least one limit point. Suppose p0 is their limit point which may or may not
be a point of the set.
If p0 is a point of the set, it must itself be a pole of f(z), and by virtue of
being the limit point, should have a cluster of poles around it. This goes
against the proved fact that poles are isolated. Hence p0 cannot be pole.
Also, a0 cannot be a zero of f(z) because the function f(z) is not analytic
(has poles) in the neighbourhood of p0. therefore p0 is an essential singularity.
The singularity is not isolated because there are poles around p0. Hence p0 is a
non-isolated essential singularity.
The point at infinity.
In the theory of complex variable it is convenient to regard infinity as a
z  1 
point. We may consider infinity as point by making a substitution in
f 1 
f(z) and then examing . Then behaviour of f(z) at infinity is determined
f 1  0
by the behaviour of at .
Thus f(z) is analytic, or has a zero, or has a simple pole, or has an
f 1 
essential singularity, etc., at infinity according as has the corresponding
0
property at .
f 1  0
For example, suppose has a zero of order m at . Then by
f 1  0
Taylor’s theorem expansion of at , is
f  1    a m  m  a m 1 m 1.....

1 z
Putting = z we get the corresponding expansion of f(z) at as
a m a m 1
f (z)    ....
z m z m 1
This expansion does not contain the constant term and positive powers
of z; so f(z) has a zero of order m at infinity.

269
f 1  0
Again, suppose has a pole of order m at . Then by
Laurent’s theorem
 m
f  1     a nn   bn n
n 0 n 1

1 z
Putting = z we get the corresponding expansion of f(z) at as
m 
f  z    bnz n   bn z n
n 1 n 0

z
Here, we see that principal part of f(z) at , is
m

b zn 1
n
n

z
i.e. principal part of f(z) at is a finite series in positive powers of
z. Hence f(z) has a pole of order m at infinity.
Similiarly the case of essential singularity can be discussed.
If a single-valued function f(z) has no singularities other than poles
either in the finite part of the plane or at infinity then f(z) is a rational function.
Polynomial.
The function f(z) is called a polynomial if it is of the form
f (z)  a 0  a1z  a 2 z 2  ....  a n z n
where n is a finite positive integer.
Rational function:
A function f(z) is called rational function of z if it is of the form
(z)
f (z) 
 (z)

(z) (z)
where and are polynomials.
Suppose that the poles of f(z) in the finite part of the z-plance are a,b,c,
……,k of order p,q,r,……s, respectively, then
(z)
f (z) 
 z  a
p
(z  b)q ......(z  k) s

(z)
where is analytic for all finite values of z.
 z  a
p
(z) (z  b)q ......(z  k)s
Thus = .f(z)

270
Hence by Taylor’s theorem

(z)   a n z n  (1)
n 0

1
Putting for z, we have
n
1 
( 1 )   a n    (2)
 n 0   

(z)
The behaviour of at infinity is same as behaviour of
( 1 ) at   0

(z) at z  
Since singularity of , is a pole, therefore singularity of
( 1 ) at   0

, is also a pole, hence the series (2) must terminate, that is, must
(z)
be finite: and consequently the series (1) must also terminate and as such
is a polynomial. Thence we have
(z)
f (z) 
 z  a
p
(z  b) q ......(z  k)s

a polynomial

another polynomial

Thus f(z) is the quotient of two polynomials, hence f(z) is rational


function.
Conversely. A rational function has no singularities other than poles.
Let f(z) be a rational function, then it must be expressed as
 (z)
f (z) 
(z)

(z) (z)
where and are two polynomials having no common factors.
(z) (z)
Singularities of f(z) are given by = 0 , i.e. at the zeros of .
1
(z) (z)
But zeros of are poles of .

271
1
 (z)
Hence has pole at the zero of f(z).
(z)
Consequently f(z) has poles at the zeros of .
That is a rational function has no other singularities except the poles.
A function which has no singularity in the finite part of the plane or at
infinity is constant.
The function f(z) has no singularity in the finite part of the plane.
Therefore it can be expanded in a Taylor’s series in terms of positive powers of
z R
z in a circle about the origin,

f (z)   a n z n  (1)
0
i.e.
z
Also since f(z) has no singularity at
z
i.e., f(z) is analytic at ; therefore

 z
 n
f (z)   A n 1  (2)
0

R  z R
Now if we let , then on the circle both the expression (1)
and (2) must be the same; that is,
n

 z
 

a
n 1
n z   An
n

n 1
1

This can be possible if either an = An = 0, and then f(z) = 0 or


n  0 and then f (z)  a 0  A 0
.
i.e., f (z)  Cons tan t

A function f(z), which is regular every where except at infinity where it


has a pole of order n, is a polynomial of degree n.
Since f(z) is regular in the finite plane, hence by Taylor’s theorem

f(z) = a0 + a1z + a2z2 + … + anzn + … (1)
1

Writing for z in this relation, we have
1 a a a
f    a 0  1  22  ....  nn  ....   2
   

272
1
f 

Then is analytic because f(z) is analytic.
z
Since the given function f(z) has a pole of order n at . Therefore
1
f 
 0
the function has a pole of order n at , consequently the expansion

(2) must have only finite number of terms in the negative powers of ; hence
1 a a a
f    a 0  1  22  ....  nn  ....
   

f  z   a 0  a1z  a 2 z 2 ...  ...a n z n .


Consequently,
Showing that f(z) is a polynomial of degree n.
Tests for detecting singularities.
Whether a given function f(z) at any point z = a, has a removable
singularity or a pole or an essential singularity, the mention of these has already
been made in the foregoing articles of this chapter. However we re-state those
tests below :
1. Removable Singularities.
lim f  z 
z a
If exists finitely, then z = a is a removable singularity.
z2  a 2
za
For example the function has removable singularity at z = a,
z2  a 2
lim  2a.
z a z  a
because
II. Poles :
lim f  z   
z a
(a) If , then z= a is a pole of f(z).
z2  a 2
za
For example the function has a pole at z = a because
z  a2
2
lim  .
z a z  a

(b) If there are finite number of terms in the principal part of f(z).
For example the function

273
e z  a  1   z  a  z  a 2 
 1    ....
 z  a  z  a
2 2
 1! 2! 

1  z  a  z  a
2
1 1
      ...
 z  a z  a 2!
2
3! 4!

Here are only two terms in the negative powers of (z-a), i.e, there are
only a finite number of terms in principal part of the function. Hence the
function has a pole at z = a (of order 2).
In general, if there are n terms in the principal part, then pole is of order
n.
III. Essential Singularity :
lim f  z 
z a
(a) If does not exist, then z= a is an essential singularity.
1
 z a 
e
For example the function has an essential singularity at z = a
1
 z a 
lim e
z a
because does not exist,
Note – the students can mark form the last two cases the difference
between the limit being infinity and limit not existing.
(b) Limit point of zeros. (is an isolated essential singularity).
1
sin
za
For example consider the function .
Its zero are given by
1 1
sin  0 i.e.,  n
za za
1
or z  a   n  1, 2,....
n

Thus we see that there is a sequence of zeros of f(z0 which has z = a its
1
sin
za
limit point. Hence z = a is an essential singularity of the function .
(c) Limit point of poles. (is a non isolated essential singularity).

274
1
f  z 
1
sin
za
For example consider the function
Putting the denominator equal to zero, the poles of f(z) are given by
1 1
sin  0 i.e., z  a   n  1, 2,....
za n

Thus we have a sequence of poles


1 1 1
a  , a  , a  ..........etc.
 2 3
The limit of these poles is z = a. Hence z = a is a non-isolated essential
singularity of the given function.
(d) If there are infinite number of terms in the principal part of f(z).
1 1 1 1
sin   .  ....
z  a z  a 3!  z  a  2
For example
1
sin
za
Here we see that in the expansion of the function there are
infinite number of terms in the negative powers of z - a, or in other words there
1
sin
za
are infinite number of terms in principal part of the expansion of .
1
sin
za
Hence z = a is an isolated essential singularity of the function .
Ex. 1.
z
Show that the function ez has an isolated essential singularity at .
Sol.
Let f(z) = ez.
z
To know the behaviour of f(Z) at , let us know the behaviour of
1
f   at   0.


1
f   at   0.

Now

275
1
 1 1 
lim e  lim 1   2  ....

 0 0
  2 
= this limit does not exist.
1
  0 e 
is an essential singularity of
z
Here is an essential singularity of ez.
lim f  z 
z a
Not that does not exit z= a is an essential singularity.
1
1 1
e   1   2  ....
 2
Aliter

Hence we have infinite number of terms in the negative powers of
1

e    0.
therefore the function has an isolated essential singularity at
z
If follows that ez has an isolated essential singularity at .

276
Ex. 2.
e1/ z
Show that the function actually takes every value except zero an
infinity of times in the neighbourhood of z = 0.
Sol.
f  z   e1/ z .

What is required to prove is that this function has an isolated essential


singularity at z = 0.
f  z   e1/ z
The function is analytic except at z= 0, that is f(z) is analytic
r  z  R.
for Thus f(z) is analytic in the annulus
r  z  R.

where R is finite and r is small, hence it can be expanded in a Laurent’s series.


Thus
1 1 1
e1/ z  1   2  ...  n  ...
z z 2! z n!
 
1 1 1 1
 . n  . n
0 n! z 1 z! z

Obviously there are infinite number of terms in the principal part.


Hence z = 0 is an isolated essential singularity.
Ex. 3.
What kind of singularity have the following functions :
1
at z  2i
1  ez
(i)
1 
at z  .
sin z  coz z 4
(ii)
cot z
at z  0 and z  
 z  a
2

(iii)
Sol.
1
f  z 
1  ez
(i)

277
Poles of f(z) are given by putting the denominator equal to zero, i.e., by
1  e z or e x  1  e 2ni

 z  2ni  n  0, 1, 2,....


z  2i
Obviously , is a simple pole.
1
f  z  .
sin z  cos z
(ii) Here
Poles of f(z) are given by putting the denominator equal to zero.
sin z  cos z  0 or tan z  1
i.e., by

or z  n   n  0,1, 2....
4

z
4
Obviously is a simple pole.
cot z
f  z  .
 z  a
2

(iii)
Poles of f(z) are given by putting the denominator equal to zero,
by  sin  z    z  a   0
2
i.e.
 0 and  z  a   0
2
or by sin z
Now sin z  0 gives z  n
or z n  n  0, 1, 2,....
z z
Obviously is the limit point of these poles, hence is a non-
isolated essential singularity.
 za   02

And gives z =a repeated twice.


Hence z = a is a double pole.
Ex . 4.
What kind of singularity have the following functions ;
z 2
 4  / ez at z  
i)
cos z  sin z at z  
ii)

278
1
sin at z  1
1 z
iii)
1
sin at z  0.
z
iv)
Sol.
f  z    z 2  4  / ez
i)
z 2
 4  / e z  0,
Zeros of f(z) are given by which gives
z  2i
which are two simple poles and
e  z  0  e 2ni i.e., z  2ni  n  0, 1, 2,....

z
Obviously is the limit point of these zeros.
z
Hence is an isolated essential singularity.
f  z   cos z  sin z
ii)
cos z  sin z
Zeros of f(z) are given by =0

i.e., by tan z  1, z  n   n  0, 1, 2,....
4
z
Obviously is the limit point of these zeros.
z
Hence is an isolated essential singularity.
1
f  z   sin
1 z
iii)
1 1
sin  0 i.e.  n
1 z 1 z
Zeros of f(z) are given by
1
z  1  n  0, 1, 2,....
n
or
Obviously z = 1is the limit point of these zeros.
Hence z =1 is an isolated essential singularity.
1
f  z   sin
z
iv)

279
Zeros of f(z) are given by putting the numerator equal to zero.
1
 n
z
i.e., by sin 1/z =0 or
1
z  n  0, 1, 2,....
n
i.e.,
Obviously z = 0 is the limit point of these zeros
Hence z = 0 is an isolated essential singularity.

280
Ex. 5.
What kind of singularity have the following functions :
1
(i) tan at z  0
z
1
(ii) at z  0
1
cos
z
(iii) cot z at z  
1
(iv)sec at z  0
z
1  ez
(v) at z  
1  ez
(vi)z cos ec zat z  
1
(vii) cos ec at z  0
z
Sol.
1 sin 1/ z
f  z   tan  .
z cos 1/ z
(i)
Poles of f(z) are given by putting the denominator equal to zero,
i.e., by cos1/ z  0 or1/ z  2n   / 2.
1
or z  n  0, 1, 2,....
 1
 2n   
 2

Obviously z = 0 is the limit point of these poles.


Hence z=0 is non-isolated essential singularity.
1
f  z  .
cos1/ z
ii)
Poles of f(z) are given by putting the denominator equal to zero,
i.e., by cos1/ z  0 or1/ z  2n   / 2
1
or z  n  0, 1, 2,....
 1
 2n   
 2

Obviously z = 0 is the limit point of the sequence of these poles


Hence z = 0 is a non-isolated essential singularity.

281
cos z
f  z   cot z 
sin z
iii)
Poles of f(z) are given by putting the denominator equal to poles.
i.e., by sin z  0 or z  n  n  0, 1, 2,.... .

z
Obviously is the limit point of the sequence of these poles.
z
Hence is a non-isolated essential singularity.
1 1
f  z   sec 
z cos 1
z
(iv)
Poles of f(z) are given by putting the denominator equal to zero.
i.e., by cos1/ z  0 or 1/ z  2n   / 2
1
or z   n  0, 1, 2,....
 1
 2n   
 2

Obviously z = 0 is the limit point of the sequence of these poles.


Hence z = 0 is a non-isolated essential singularity.
1  ez
 v f  z 
1  ez
Poles of f(z) are given by putting the denominator equal to zero.
i.e., by1  e z  0or e z  1  ei  e 2ni i
 z   2n  1 i  n  0, 1, 2,.... 
z
Obviously is the limit point of the sequence of these poles.
z
Hence is a non-isolated essential singularity.
(vi) f(z) = z cosec z = z/sin z.
Poles of f(z) are given by putting the denominator equal to zero.
i.e., by sin z  0 or z  n  n  0, 1, 2,.... 

z
Obviously is the limit point of this sequence of these poles.
z
Hence is a non-isolated essential singularity.

282
1 1
f  z   cos ec 
z sin 1
z
(vii)
Poles of f(z) are given by putting the denominator equal to zero.
i.e., by sin1/ z  0 or 1/ z  n
i.e., z  1/ n  n  0, 1, 2,.... 

Obviously z = 0 is the limit point of the sequence of these poles.


Hence z =0 is a non-isolated essential singularity.
Ex. 6.
1

z2
e
Show that the function has no singularities.
Sol.
1

z2
e
Here f(z) = .
1

e 0 z2
Zeros of f(z) are given by or z2 = 0
z  0
is a zero of order two.
So there it no limit point of zeros, hence there is no singularity.
Next, Poles of f(z) are given by putting the denominator equal to zero,
i.e., by
1

e z  0,
2

which is impossible for any value of z, real or complex; therefore no poles.


Hence no singularity anywhere.
Ex. 7.
Find out the zeros, and discuss the nature of the singularities of
z2 1
f  z  sin
z 2
z 1
Sol.
Putting the numerator equal to zero, the zeros of f(z) are given by
1
 z  2  sin 0
z 1

283
1
or by z  2 and sin 0
z 1
1 1
sin  0 gives  n
z 1 z 1
Now ,
1
z  1  n  0, 1, 2,.... 
n
giving thereby
1
z  1  n  0, 1, 2,.... 
n
Hence zeros of f(z) are z = 2 and
Obviously z = 1 is an isolated essential singularity.
Next putting the denominator of f(z) equal to zero the poles of f(z) are
given by
z2=0 or z= 0 is a pole of order two.
Ex. 8.
e  
c / z a

ez / a  1
Find the singularities of the function , indicating the character of
each singularity.
Sol.
ec /  z  a 
f  z  z/a ,
e 1
Here
The singularities of f(z) are obtained by putting the denominator equal
to zero.
Hence the singularities of f(z) are given by
ez / a  0 or e z / a  1or e z / a  e ni
or z  2nia  n  0, 1, 2,.... 

The numerator does not vanish for any one of these values of z; hence
they are the singularities of f(z).
Now, we indicate the character of these singularities. Z= 0 is a simple
pole of the function f(z).
z  2nia  n  0, 1, 2,.... 
Similarly are each a simple pole of the
function f(z).
e   e  
c / z a c / z a
f  z  z/a 
e  1 e1 z a / a   1
Also,

284
2
c  c  1
1    ...
z  a  z  a  2!


 za  za  1
2


e 1     .... 1

 a  a  2! 

The expression on the right hand side involves all positive and negative
powers of (z – a). Therefore it represents a Laurent’s series. There are infinite
number of terms which have negative powers of (z – a); i.e., there are infinite
number of terms in the principal part of the series. Hence z = a is an isolated
essential singularity of f(z).
Thus we have found out all the singularities of f(z) and their nature
indicated.
Ex. 9.
Consider the singularities of the function represented by the series

1 1
 n!
1   2n z 
2
n 0

and obtain expansion by Laurent’s theorem,


Sol.

1 1
f  z  
n  0 n! 1   2 z 
n 2

Let
1   2n z   0
2

Singularities of f(z) are given by


i i
i.e., by z  n
and z   n  n  0, 1, 2,.... 
2 2
Hence f(z) has simple poles at each of these points.
z 1
All these singularities lie within the circle ; and z = 0 is a limit
point of this sequence of singularities, hence z = 0 is an essential singularity of
the function f(z).
The function f(z) is a regular (analytic) on and outside the circle C,
z  r  r  1
, therefore it must be expanded in a Laurent’s series of negative
power of z only; also we see that the function is an even function of z, so its
expansion must not contain odd power terms in z. So we should have

b2 b4 b6 b 2m b 2m
f  z 
z 2

z 4

z 6
 ... 
z 2m
 
m 1 z
2m

285
1
b 2m   z 2m 1f  z  dz as usual
2n C
where
  
1 1 1

2m 1  
2i C
 z
1   2 z  
2
n 0 n! n

2m 1  1 2 2 n 

1

2i C
 z 2n 2 
dz
 n  0 n! 2  z 

1 
2 2 n

2i C
z 2m 3
  1  2 2n
 z  dz
2 1

n  0 n!

1 
2 2 n 
 1  22n  z 2  2 4n z 4   1 2 2n (m 1) z 2 m 1 dz
m 1

2m 3
 z
2i C n  0 n!
 
  nm
1 m 1 2 dz
  
2i C n 0
 1
n! z

2 1
 0 zk
dz k  1.
the other integrals vanish being of the form
z  1,
[ This series converges absolutely because and hence converges
uniformly; so term by term integration is valid]
1 
2 nm dz
 1 
m 1
b 2m 
2i 
n  0 n! C z
Hence

1 2m dz

2 
n 2m

 1   e 2    e2 
2m

m 1
 sin ce
2i C
z n 0 n!
 1
m 1

e  
2
2 2m
 id where z  rei
2i 0

  1  e  2 2m
m 1

z 1
Hence when

 2m 1 
f  z     1  e2 
m 1

m 1  z 2m 


e  2 2


e 
2 4


e 2 6

.....
z2 z4 z6

286
z  1/  f  1/  
Taking in f(z) we see that the function has a zero of

second order at =0; hence the given function has a zero of second order at
infinity.
Ex. 10.

z2
f  z   ,
n 0 4  n z
2 2

If show that f(z) is infinite and continuous for all


real values of z but f(z ) possesses Laurent’s expansion valid in a succession of
the ring-shaped spaces.
Sol.

z2
f  z   ,
n 1 4  n z
2 2
un
When z is real, let denote general term of the
2
z
un 
4  n 2z 2
above series, then which is obviously positive when z is real.
z2 1
 2 2  2.
nz n
1
vn 
n2
Let
1
n 2
u n
But is convergent, therefore is uniformly convergent. Also
1 
 n2  6 .
we know that
1
0  f  z  
6
If follows that showing that f(z) is finite.
z2
4  n 2z2
Further we see that the denominator of does not vanish for any
real value of z, so each term is continuous function of z; consequently the sum
f(z) is continuous.
Hence f(z) is finite and continuous when z is real.
When z is complex variable.

z2
f  z  
n 1 4  n z
2 2

Then is function of complex variable,

287
z2
4  n 2 z2
The general terms is .
Putting its denominator equal to zero, the singularities of this term are
given by
2i
4  nz 2  0 i.e., z     1
n
2i 2i
z and z 
n n
Thus each term has two singularities where n =
1,2,3,....
n 
As these singularities have the limit point z = 0. Thus there are
infinite number of singularities in the neighbourhood of the point z = 0,
howsoever small the neighbourhood may be taken.
It immediately follows that howsoever large or small circle be drawn
with its centre at z = 0, the function is not regular within the circle. Hence f(z)
cannot be expanded in the z plane by Maclaurin’s theorem.
Putting n = 1,2,3,... and so on in (1) we see that the singularities are
2i 2i 2i
 ,  ,  ,....
2 2 3
and so on.
These singularities lie respectively on the boundaries of the circle
2 2
z  2, z  , z 
2 3
and so on.
Thus all singularities lie on the boundaries of these circles and so f(z) is
analytic between the ring shaped region between any two consecutive
boundaries. Thus there is a chain of successive ring shaped regions in every
one of which f(z) is analytic, and hence f(z) can be expanded by Laurent’s
theorem within these ring shaped spaces.

288
Ex. 11.
log  1  z 2 
Prove that near z=1, function may be expanded in a series of
the form

log 2   a n  z  1 ,
n

n 1

and find the values of an; the principal value of the logarithms being taken.
Obtain also the radius of convergence of the series.
Sol.
f  z   log  1  z 2 
Here
1  z2 z  i
The zeros of i.e., are the two branch points of the function

log 1  z 2 
.
Thus f(z) has two singularities at z = i and z = -i.
Distance of the point z =1 from the nearest singularity there both z = i
2
or z = -i are at the same distance form z =1, is equal to . Thus for the region
2
whose centre is at z = 1, the radius of convergence is . In this circular
domain f(z) can be thought of as single valued and analytic, the particular
branch being a value of the function f(z) for any z in the domain. The particular
branch referred to in the question is given as the principal value i.e. the real
value log 2 for z = 1.
z 1  2
Since f(z) is regular within the circle , hence it can be

f  z    a n  z  1
n

n 0
expanded in a Taylor’s series within this circle. So by
f n  1
an 
n!
Taylor’s Theorem where
f  z   log  1  z 2 
we have
2z 1 1
f  z   
1 z 2
z i z i
Differentiating it n times we have

289
 1 1 
f n  z    1  n  1 ! 
n 1
 n 
  z  i   z  i  
n

f n  1   1  n  1 !  1  i   1 i 


n 1 n n

So that
n
 n  1 ! 
n
 2  1
n 1
2 cos .
4

 n  1 ! 
n n
2  1
n 1
2 cos
an 
f n
 1  4
n! n!

2  1
n 1
n
 2
n
 cos .
n 4

a 0  f  0   log 2.
and
 
f  z   a 0   a n  z  1  log 2   a n  z  1
n n

1 1
Hence
2  1
n 1
n
 2
n
an cos .
n 4
where
Ex. 12.
The function f(z) has a double pole z = 0 with residue 2, a simple pole
at z = 1 with residue 2, is analytic at all other finite points of the plane and is
z  f (2)  5 and f (1)  2
bounded as . If find f(z).
Sol.
Since f(z) has a simple pole at z = 1, the residue at which is 2, therefore
2
z 1
is a term in the principal part of f(z).
Also, since there is a double pole at z = 0, the residue at which is 2
2 b

z z2
therefore must be the terms in the principal part of f(z).
2 2 b
f (z)  a 0  a1z  a 2 z 2  ....    2  (1)
z 1 z z
Hence

290
z  z
f(z) is bounded as implies that f(z) has no singularity at .

Hence
   must have no singularity at   0 .
f 1

So there must be no terms in the principal part of


f 1   .
The principal part of
f 1    is obtained by putting 1

for z in the
polynomial part of f(z).

Thus principal part of


   is a  a
f 1 1 2
2
 .......  ...

a1  a 2  ....  a m  0
The terms of this can be zero only if etc.
Then (1) becomes
2 2 b
f (z)  a 0    2  (2)
z 1 z z
f (2)  5 and f (1)  2
But given that
Substituting in (2) we have
5  a0  2 1 b and 2  a 0  1  2  b
4
a 0  1 and b  4
On solving these we get
Then (2) simplifies to
2 2 4
f (z)  1    2
z 1 z z
z 3  3z 2  2z  4
or f (z) 
(z  1)z 2

Thus f(z) is determined.


Aliter. Given that f(z) has a simple pole at z = 1 and residue at it is 2, also f(z)
has a double pole at z = 0 and residue at it is 2. hence the principal part of f(z)
is
2 2 b

z 1 z z2
+

291
Since f(z) is analytic at all other points except at z = 1 and z = 0 and
f (z)
is bounded, therefore
 2  2 b 
f (z)      2 
 z 1   z z 

is analytic and bounded in the whole z-plane.


 2  2 b 
f (z)      2 
 z 1   z z 
Hence by Liouville’s theorem is constant.
 2  2 b 
i.e. f (z)       2   a 0 where a 0 is a cons tan t.
 z 1   z z 
2 2 b
i.e. f (z)  a 0    2
z 1 z z
which is same as relation (2) of the method 1.
The rest of the calculation will be now exactly the same as adopted in
method 1.
Ex. 13.
The only singularities of a single valued function f(z) are poles of order
1 and 2 at z = -1 and z = 2, with residues at these poles 1 and 2 respectively. If
7 5
f (0)  ; f (1) 
4 2
; determine the function and expand it in a Laurent’s series
1  z  2.
valid in
Sol.
z  1
Since f(z) has a simple pole at where residue is 1, therefore
1
z 1
must be a term in principal part f(z).
Also, f(z) has a double pole at z = 2 at which residue is 2, therefore

 2 b  
  2
 z  2  z  2 
 
must be the terms in the principal part of f(z).
Thus under the above facts f(z) can be expressed as
1 2 b
f (z)  a 0  a1z  a 2 z 2  ....     (1)
z 1 z  2  z  2 2

292
z 0
Further, since is not a pole of f(z), therefore is not a pole of

 
f 1

, it means in the principal part of
f 1  

there must be no terms


containing negative powers of . The principal part of
f 1
 
can be obtained
1

by putting for z in the polynomial part of f(z). thus principal part of the

 
 a1 a 2 
f 1   2  ....  .. 
   
expansion of is . The terms of this can zero only if
a1  a 2  a 3  ......  a m  etc,.....  0

Hence (1) simplifies to


1 2 b
f (z)  a 0     (2)
z  1 z  2  z  2 2

7 5
f (0)  ; f (1) 
4 2
But given that
Applying this to the above expansion we have
7 b 5 3
 a 2  and  a 0   b
4 4 2 2
i.e. a 0  1 and b  3
1 2 3
Hence f (z)  1   
z  1 z  2  z  2 2

Thus f(z) is determined.


1  z  2.
Now we see that f(z) is regular in the region Hence it can be
1  z  2.
expanded in a Laurent’s series within the annulus
We may write
1 1 2
1 1  z 3 z
f (z)  1  1    1    1  
z z  2 4 2

arranged suitably for the validity of binominals

293
1 
   
 n 3  n
  1 . 1 z n   z 2   n  1 z 2
n
 1 
z 0 0 4 0
n 

3n  1  z  1
 1    1
n
 
n 0 4 2 0 z n 1

which being a series in the positive and negative powers of z, is the Laurent’s
expansion.
z  1 and z  2
Aliter: Principal parts of f(z) at are
1 2 b
and 
z 1 z  2  z  2 2
respectively.
Since f(z) is analytic and bounded in the whole of the z plane except at
z  1 and z  2
, therefore

 1   2 b 
f (z)     2
 z  1   z  2  z  2  

is analytic and bounded in the whole of the z plane. Hence by Liouvile’s


Theorem

 1   2 b 
f (z)     2
 z  1   z  2  z  2  
is constant.

 1   2 b 
i.e. f (z)      2
 a 0 , where a 0 is cos n tan t.
 z 1   z  2  z  2  
1 2 b
i.e. f (z)  a 0   
z  1 z  2  z  2 2

which is exactly the same as relation (2) of method 1. the rest of the
calculations will be identical with those adopted in method 1.

294
Ex. 14.
The only singularities of a single valued function f(z) are poles of order
2 and 1 at z = 1 and z = 2, with residues of these poles 1 and 3 respectively.
3
f (0)  , f (1)  1
2
, determine the function.
Sol.
Since there is a single pole at z = 2 at which residue is 3, therefore
3
z2
must be a term of the principal part of f(z).
Also, z = 1 is a pole of order 2 at which the residue is 1, therefore
1 b

z  1  z  1 2
must also be the terms of the principal part of f(z).

Under these above facts we should have


3 1 b
f (z)  a 0  a1z  a 2z 2  ...     (1)
z  2 z  1  z  1 2

Since
z
is not a pole of f(z), hence
0
is not a pole of
  ;
f 1

that is, there are no terms in the principal part of


  .
f 1
The principal part of

 
f 1

is
a1 a 2
  ....
 2

a1  a 2  a 3  ....  0
The terms of this expression will be zero if
3 1 b
f (z)  a 0     (2)
z  2 z  1  z  1 2
Consequently
3
f (0)  ,f ( 1)  1
2
But given ; hence substituting these in (2) we get
respectively
3 3 b
 a 0   1  b and a 0  1  1 
2 2 4
On solving these we get a0 = 2 and b = 2

295
Hence (2) simplifies to
3 1 2 2z 3  4z 2  3z  3
f (z)  2    
z  2 z  1  z  1 2  z  2   z  1
2

2z 3  4z 2  3z  3
i.e., f (z) 
 z  2   z  1
2

Thus f(z) is determined.


Aliter:

3 1 b

z2 z  1  z  1 2
Principal parts of f(z) at z = 2 and at z = 1are and
respectively.
Since f(z) is analytic and bounded in the whole of the z-plane except at
z = 2 and z = 1, therefore

 3   1 b 
f (z)     2
is cons tan t.
 z  2   z  1  z  1 

 3   1 b 
i.e. f (z)     2
 a 0 where a 0 is cons tan t.
 z  2   z  1  z  1 
3 1 b
or f (z)  a 0   
z  2 z  1  z  1 2

This is exactly the same as relation (2) of method I.


Now, the rest of calculations will be identical with those adopted in
method I.
THE CALCULUS OF RESIDUES
Definition of the Residue at a Pole.
za
Let be a pole of the order m of a one-valued function
f  z  , and  za
any circle of radius r with centre at , which does not contain
za f  z
any other singularities except ; then is regular within the annulus
0  za  r
, hence it can be expanded within this annulus in a Laurent’s
series in the form

296
 m
f  z    a n  z  a    bu  z  a 
n n

n 0 n 1

where
1
 z  a  f  z  dz  formula 
n 1
bn  
2 
1
f  z  dz.
2i 
P articularly b1 

b1 za
The coefficient is called the residue of f(z) at the pole .

The circle is arbitrary and may therefore be replaced by any closed
contour C (small or large) containing within it no other singularities except
za
. Thus
1
f  z  dz.
2i C
b1 

za
where C is a closed contour containing the only singularity , and the
integration along C being taken in anti-clockwise direction.
Definition of the Residue at infinity.
f  z za
In the last article we defined residue of at a pole as
1
f  z  dz
2i C

where C is a closed curve such that f(z) is analytic everywhere within it except
za
at .
Now we proceed to define the residue f(z) at infinity.
If f(z) has an isolated singularity at infinity, or is analytic there, and if C
, is a large circle which encloses all the finite singularities of f(z); then residue
z
of f(z) at is defined by
1
f  z  dz
2i C

the integration taken round C in clockwise direction (which is negative with


respect to the interior of C, bnt is positive with respect to exterior of C),
provided that this integral has a definite value.
z
Or we may define residue of f(z) at as

297
1
f  z  dz
2i C
-
the integration taken round C in anti-clockwise direction.
z
To show that a function f(z) may be analytic at but yet has a residue
z
at .
1
f 
z  0
If f(z) is regular (analytic) at , then must be regular
1
f 
 0
Hence by Taylor’s theorem expansion of at is
1 
f     a nn
 0

1
z
 z
Putting we get expansion of f(z) at as

an
f  z   n
0 z

a a
 a 0  1  22  ....
z z
z
By definition residue of f(z) at is
1
f  z  dz
2i C



1 an
  
2i C 0 z n
dz by the above

1 a1
2i C z
 dz,

dz
the other int egal vanish as they are of the form   k  1
C
z
a1 dz

2i
.2i as 
C
z
 2i

 a1 .

which is finite, being the negative of the coefficient of 1/z in the expansion of
z
f(z) at as seen from the above.

298
z
Hence it is shown that a function may be analytic at , but yet has a
residue there.
The Theorem of Residues. (Cauchy’s ResidueTheorem).
Let f (z) be one-valued and analytic within and on a closed contour C,
z1 , z 2 , z 3 ,....., z n ; R1 , R 2 , R 3 ,....., R n
except at a finite number of poles and let
be respectively the residues of f(z) at these poles, then to prove that

299
za
Residue at a pole of any order (simple or of order m).
za
The formula given below for the residue of f(z) at the pole holds
for the poles of all orders including one.
za
We have seen that the expansion of f(z) about in a Laurent’s

b b2 bm
f  z   an  z  a   1 
n
 ...  .
0 z a  z a  z  a
m

series is When b1 is said to


za
be residue of f(z) at the pole ,
Whatever be the order of the pole, simple or order m.
z  a  t or z  a  t
If we put where t is small, then we have from
above

b1 b 2 b
f  a  t   antn    ...  m .
0 t1 t 2 tm

We see that b1 is the coefficient of 1/t in the above expansion. This


establishes a rule for finding the residue of function f(z) at a pole z = a of any
order (simple or order m).
The formal statement of the rule is as below:
Put z = a+t in the function f(z), expand it in powers of t where t is
sufficiently small. Then
Coefficient of 1/t is the residue of (z) at z = a .
Ex .1.
z2
at z  ia
z2  a 2
Find the residue of
Sol.
z2 z2
f  z   .
z 2  a 2  z  ia   z  ia 
We have
z  ia
So is a simple pole, hence
lim  z  ia  f  z 
z  ia za
Residue at is

300
z2
 lim  z  ia  .
z ia  z  ia   z  ia 
2
z
 lim
z ia  z  ia 

a 2 1
  ia.
2ia 2

z2    z  
f  z  form is 
z2  a 2    z  
Aliter.
  z   z 2 and   z   z 2  a 2
where
So residue at
  ia 
z  ia 
  ia 
 ia   1 ia.
2
 z2 
  
 2z  z ia 2ia 2

Ex.2.
z3
at z  1.
 z  1  z  2   z  3
4

Find the residue of


Sol.
Here z=1 is a pole of order 4.
z3
f  z 
 z  1  z  2   z  3
4

1   1 t  
3

 4  putting z  1  t where t is small.


t   t  1  t  2  

1  1  
1
1 
   
3
  1  t 1  t  1  t  
2t 4   2  
1   1 1 2 1 3 
 (1  3t  3t  t )  1  t  t  t ... 1  2 t  4 t  8 t  
2 3 2 3

2t 4   
1  3  3 7 2 51 3 
 (1  3t  3t  t ) 1  2 t  4 t  8 t  ....  
2

2t 4   

301
1 1 9 21 15 
Coefficient of  1    
t 2 2 4 8 
101

16
101
Hence the residue at z  1 is .
16

Ex.3.
1
at z  ia.
z  a2 
2 2

Find the residue of


Although the residue can be found by the method suggested in case II of
z  ia
the article (because here is a pole of order more than unity), but we
demonstrate below the method suggested in case III applicable for finding
residue whether the pole is simple or of any order.
z  ia ia  t
To find the residue at , we write
for z in the given function
z  ia
f(z), then coefficient of 1/t will be the residue at .
Here
1
f  z 
Z  a2 
2 2

1

 z  ia   z  ia 
2 2

1
 putting z  ia  t, where t is small.
t 2  t  2ia 
2

2

1 t  1  t 
 1  4ia    4a 2 t 2 1  ia  ....
4a 2 t 2
1 1
Here coefficient of is
t 4ia 3
1
Hence residue at z  ia is .
4ia 3
Here
Ex.4.
z4
.
 c2  z2 
4

Find the residues at the poles of the function


Sol.

302
z4
Here f (z) 
c  z2 
2 4

z4

 z  ic   z  ic 
4 4

z  ic and z  ic
Obviously are the two poles each of order 4.
z  ic z  ic  t
To find the residue at the pole , put
 t  ic 
4

then f (z)  4
t  2ic  t 
4

4

 t  ic  1  
4 t

  2ic 
t 4  2ic 
4

1 1
4 
t  ic   4  t 4  4ict 3  6c 2 t 2  4ic3 t  c 4 
4
Now
t t
4ic 6c 2 4ic3 c 4
 1  2  3  4
t t t t

1  4t 
4
1  t  10 2 20 3
and 4 
1     1   t  t  .... 
 2ic   2ic  16c  2ic  2ic   2ic  
4 2 3

coefficient of 1/ t in f (z) is
1  5 
4 
4ic  12ic  10ic  ic 
16c  2 
i
 .
32c3
This is residue at the pole z  ic.
i
Writing  c for c in this result we have residue at the other pole z  ic as .
32c3
Ex.5.
cot z
 z  a
2

Find the residues at the function .


Sol.

303
cot z cos z
Here f (z)  
 z  a
2
 z  a  sin z

 z  a
2
sin z  0
Poles of f(z) are given by
 z  a
2
0 za
i.e., by or repeated twice,
so z = a is a pole of order 2.
Also
sin z  0 gives z  n
z  n  n  0, 1, 2,.......

These are simple poles if n is finite. The limit point of these poles is
z
which is therefore a non-isolated essential singularity.
za
Further we proceed to evaluate the residues at the poles of order 2
zn
and at (n finite) we have
cot z  z
f  z   form
 z  a  z  a
2 m

za
Residue at (Pole of order 2) is
d 
 dz  cot z    z
m 1

  by formula
 1!  m  1!
 z  a
  cos ec 2 a
cos z /  z  a   z
2

Again, f  z   form
sin z   z

 Residue at the simple pole


zn
is

304
  z 
   
 cos z /  z  a    d  z 
2

  by formula  
 d   dz 
sin z
 dz z n  
 z  n
 cos z /  z  a  2 
 
  cos z z  n
1

 n  a
2

Computation of Residue at infinity.


lim  zf  z  
z z 
(a) To establish that residue of f(z) at is
1
f  z  dz
z 2i C
By definition the residue of the function f(z) at is
taken clockwise round a large circle C which encloses in its interior all other
finite singularities it any.
1
f   at   0 is

Therefore residue of
1  1   d  1
 f   .   2  putting z 
2i y       
1 1 1
i.e.,   2 f   d
2i y  


taken round a small circle , centre origin, in anti-clockwise direction.
1
f  z  dz
z0 2i C
Since residue of f(z) at , is , and also equals to
1
lim z f  z  f 
z0  
, therefore by implication residue of at =0 which is shown
1 1 1
  2
2i y 
f   d

to be equal to must be equal to

305
 1  1 
lim   2 f   
0
    
 1  1 
or lim  2 f   
0
    

1
z

Writing we have

f i  z  at z   lim   zf  z   .
z 
Residue of is the
(b) To establish that residue of f(z) at infinity is the negative of the
1
z
coefficient of in the expansion of f(z) for values of z in neighbourhood
z
of .
1
f 
 0
By Laurent’s theorem expansion of , at the pole of order
m, is
1  m
f     a n  n   b n   n ...........  1
 0 1

1
f z
 z
Putting we get the expansion of f(z), at the pole of
order m, as
 m
f  z    a n z  n   b n z n ........  2 
0 1

306
By definition, residue of f(z) at infinity is
1
f  z  dz
2i C

1  m

   
2i C  0
a n z n
 
n 1
b n z n dz from (2)

1 a1
2i C z
 dz;

dz
the other int egrals vanish sin ce they are each of the form z
C
k
, k  1.

1 dz

2i
.a 2 .2i sin ce z
C
 2i

1
z
which is [see relation (2)] the coefficient of with sign changed in the
z
expansion of f(z) in the neighbourhood of .
Ex.1.
z3
at z  .
z2  1
Find the residue of
Sol.
z3
f (z)  .......  1
z2  1
Let
1 1
then f  .........  2 
    1  
2

1
f 
 0
It is seen from (2) that has a simple pole at , hence f(z) has
z
a simple pole at .
z
Residue of f(z) at .
 lim  zf  z   formula
z 

 z3 
 lim  z. 2  which does not exist.
z 
 z  1
 Alternatively we write f(z) as

307
1
z3  1
1  2 
z2  z 
 1 1 
 z 1      ........
 z z 
1 1
 z   2  ......
z z
z
It is seen that the coefficient of 1/z is 1, hence residues of f(z) at
is equal to 1.
Ex.2.
z2
 z  a   z  b  z  c
Find the residue of at infinity.
z2
f (z) 
 z  a   z  b  z  c z
Sol. The function is analytic at
1
f 
 0
because is analytic at .
 z2 
 lim  z.   1.
  z  a   z  b   z  c  
z 

Residue of f(z) at infinity


This verifies the above theorem that a function although analytic at
infinity, yet has residue there.
Evaluation of Real Definite Integrals by Contour Integration.
A large number of real definite integrals, whose evaluation by usual
methods becomes sometimes very tedius, can be easily evaluated by using
Cauchy’s theorem of residues. To do this we take a closed curve C, find the
poles of the function f(z) and calculate residue at those poles which lie within
C. then using Cauchy’s theorem of residues we have

 f (z)dz  2i  (sum of


C
the residues of f (z) at the poles within C)

The curve is usually called a contour and the process of integrating


along a contour is called contour-integration.
Integration Round the Unit Circle.
Evaluation by contour integration of the integrals which are of the type
2

   cos ,sin   d
0

308
  cos ,sin   cos  sin 
where is a rational function of and .
ei
To evaluate this type of integrals, write z = .
1 1 1 1 dz
cos    z   and sin    z   , d 
2 z 2i  2 iz
So that
2
Then    cos ,sin   d
0

1  1 1  1   dz
     z   ,  z  
2  z  2i  z   iz
  f (z)dz
C

z 1
where f(z) is a rational function of z, and the contour C is the unit circle .
Find the poles of f(z) and the calculate the residues at those poles only
which lie within C, then using Cauchy’s residue theorem we have
2

   cos ,sin   d
0

  f (z)dz
C

2i
= (sum of the residues of f(z) at its poles within C)
Ex. 1.
2
2
e
cos 
cos  sin   n  d 
0
n
Apply calculus of residues to prove that
where n is a positive integer.

309
Sol.
Consider the integral

e
cos 
cos  sin   n   i sin  sin   n  d
2
cos  i  sin  n 
 e
0
e d

2
 cos  i sin 
 e
0
e  in d

2

e
i  in
 e d
0

1 dz
 C
ez .
z n iz
dz
writing ei  z, d 
iz
ez
 i  dz
C z n 1

ie z
 C
f (z)dz where f (z) 
z n 1

z 1
And C is the unit circle
Pole of f(z) are given by putting the denominator equal to zero: that is,
n 1
z  0 or z  0
by .
Thus z = 0 is a pole of order (n+1) and lies within C.
The residue at the pole z = 0 of order n+1.
1  dn z  m 1 (a)
  n  ie   by formula
n  dz z  0  m  1
i z i
 e  
n z 0 n

Hence by Cauchy’s theorem of residues

310
 f (z)dz  2i  (sum of residues within C)
C

 i 
 f (z)dz  2i   n 
C

ez  i 
or  i dz  2i   
C z n 1
 n
2
2
or e
cos 
cos  sin   n   i sin  sin   n  d 
0
n

Equating real parts we have


2
2
e
cos 
cos  sin   n  d 
0
n

Ex. 2.
2

e
 cos 
cos  n  sin   d
0
Evaluate where n is a positive integer.
Sol.
Consider the integral
2
I  e
 cos 
cos  n  sin    i sin  n  sin   d
0
2
 e
 cos 
.e  i nsin  d
0
2
 cos  i sin  
 e
0
.e  in d

2

e
 ei
 .e  in d
0

 1  dz dz
   e z .  writing ei  z, d 
C
 2  iz iz
z
1 e
  n 1 dz
i Cz

z 1
where C denotes the unit circle .
Obviously the only pole within the contour is z = 0 of order n+1.

311
 1
n
1  d z  e z  
 n  
n  dz  i  z 0 i (n)
At z = 0, the residue =
Hence by Cauchy’s residue theorem, we have
I  2i 
(the sum of the residues within the contour)
 1
n
2
 1
n
2i  
i (n) n
=
2
i.e  e
 cos 
 cos  n  sin    i sin  n  sin     d  2n  1 n

Equating real parts, we have


2
2
e cos  n  sin   d   1
 cos  n

0
n

Ex. 3
2
d 2
 2  cos  
0 3
Show that

Sol.
2
d
 2  cos 
0
Let I=
2
2d
  4e
0
i
 e  i
1 2dz
i C

 1
z4  z  
 z
2 dz
Thus I   2
i z  4z  1

312
dz
ei  z, d  z 1
iz
Writing where C is the unit circle
Poles of the integrand are given by
4   16  4 
z 2  4z  1  0 or z   2  3
2
only z  2  3 lies inside C,
2 1
the residue at which  .  2  3
i d 2 
 dz  z  4z  1  z
2  1  1
 .   2  3 
i  2z  4  i 3

Hence by Cauchy’s residue theorem


I  2i 
sum of residues within the contour
1
 2i 
i 3
2

3

Ex. 4
(a) Use method of contour integration to prove that
2
d 2
 1 a
0
2

 2a cos  1  a 2
, 0  a  1.

Sol.

313
2
d
Let I   1 a
0
2
 a  ei  e i 
1 dz

C  1  iz
1 a2  a  z  
 z
dz
writing ei  z, d 
iz
where C is the unit circle z  1
1 dz
 
i C  1  a  z  az 2  a
2

1 dz
ia C

 z  a   z  
1
 a
1
  f (z)dz where f (z) 
C  1
ia  z  a   z  
 a

1
z  a, z 
a
Poles of f(z) are
As 0<a<1, therefore z = a only lies inside C.
Residue at the simple pole z = a is
 lim  z  a  f (z)
z 0

 
 1 
 lim  z  a   
 ia  z  a   z  1  
z 0

  a  
1 1
 lim  
ia  z   i  1  a 
z a  1 2

 a

Hence by Cauchy’s residue theorem we have


I  2i 
sum of residues within C
1 2
2i. 
i  1 a  1 a
2 2

314
Ex. 5.

  if a  b  0
2
sin 2 d 2
 
a  b cos  b 2
a a 2
 b2 
0
(b) Prove that
Sol.
z  ei , then
If we write
1 1 1 1 dz
cos    z   , sin    z   , d 
2 z 2i  z iz
2
sin 2 d
and so I 
0 a  b cos 

z  1
2 2
1
2b C
 dz
 2 2a 
z z  z  1
2

 b 
where C denotes the unit circle z  1

 z 2  1 dz
2
1
2b C z 2  Z     Z   

a  a 2
 b2  a  a 2
 b2 
where   , 
b b

  1,so     1,
Obviously lies outside C, and since
   1, so  lies inside C.

z
So inside the contour C there is a simple pole and a pole z = 0 of
order two.
z 
Residue at the simple pole is

315
lim  z    f (z)
z 

 z 2  1
2
i
 lim  z    .
z  2b z 2  z     z   
2
 1
 
i    i  
 2
 2
 1

2b       2b   
2

i    
2
 1 
  because   1, so   
2b      

i i 2 a 2
 b2  i a 2
 b2 
      . 
2b 2b b b2

i
2b
and residue at the double pole z = 0 is the coefficient of 1/z in
z  1
2 2

 2a 
z2  z2  z  1
 b 
, where z is small.

z  1
2 2
i
Now,
2b  2a 
z2  z2  z  1
 b 
1
i  1  2a 1 1 
 1  2  1  
2b  z  b z z 2 

1 i  2a  ia
  i.e. 2
z 2b  b  b
The coefficient of is easily seen to be
I  2i 
Hence (sum of the residues at the poles within C)
i a  b2  
 
2
 ia  2
 2i  
b2
 2 2 a
b  b
a 2
 b2 
 

316
Ex. 5(b)

ad
a
0
2
 cos 2 
(a) Evaluate where a is positive.
Sol.
 
ad 2ad
0 a 2  cos 2   2a
0
2
 1  cos 2
Let I = =

Clearly   1 , therefore  lies outside C.


and as    1,    1, so  lies inside C.
Re sidue at z   is
2ai
 lim  z   
z   z     z  
2ai

 
2ai i
 
4a 1 a  2
2 1 a 
2

2
ad
  1  2a
0
2
 cos 
where   2

1 1 dz
put z  ei , then cos    z   , d 
2 z iz
2ai
and I   2 dz
C
z  2z  1  2a 2   1
where C is the unit circle z  1
2ai
 dz
C
 z     z  
where     1  2a 2   2a  1 a  2
and     1  2a 2   2a 1 a 
2

Hence by Cauchy’s residue theorem we have


I  2i 
Hence the sum of the residues within C.

317
i 
 2i  
2 1 a 2
 1 a  2

Ex. 5.(c)

ad
a
0
2
 sin 2 
(b) Evaluate

ad
Let I 
0
a  sin 2 
2


2ad

0
2a  1  cos 2
2

2
ad
  2a
0
2
 1  cos 
putting 2  

2
2ad
  4a
0
2
 2   ei  e  i 
2a dz dz
 writing ei  z, d 
C
4a  2   z  1  iz
2
 z    iz

where C is unit circle z  1


dz
 2ai 
C
z  2z  1  2a 2   1
2

2aidz


C z
  z  

where    1  2a 2   2a  1  a  ,    1  2a   2a  1  a 
2 2 2

Clearly   1 and as   1,    1.
 only  lies inside the contour C.
Re sidue at z   is
2ai
 lim  z   
z   z     z  
2ai


2ai i
 
4a  1 a  2
2  1 a2 

Hence by Cauchy’s residue theorem

318
I  2i 
the sum of the residues within C.
i 
 2i  
2a 1 a 
2
 1 a  2

Ex. 5.
2
cos 2
 5  4 cos d
0
Evaluate
Sol.

2
cos 2
Let I  5  4 cos d
0
2
1 ei2   e 2i dz

2 
0 5  2 e  e 
i  i
d write z  ei , d 
iz
 2 1
1  z  2  dz
  
z 
2 C  1  iz
5  2 z  
 z

1 z4  1
2i C z 2  2z 2  5z  2 
 dz

1 z2  1
2i C z 2  2z  1  z  2 
 dz

where C denotes the unit circle z  2

1
z
2
The poles within the contour C are : a simple pole at and a pole
of order two at z = 0.
Now we calculate the residues at these poles.

319
1
Residue at z   is
2
1 z4  1
lim
1 1 
z   z   2i z 2 (2z  1)  z  2 
2 2 
4
 1  1
  1 1
1 1  2  1 16 17
 .  
 4i 1 . 3
2
2 2i  1   1 24i
    2 2 2
 2   2 

1
z
And residues at z=0 is the coefficient of in
1 z4  1
where z is small.
2i z 2  2z  1  z  2 
1 1
1 z4  1 1 1  1   2
Now  1  4   1   1  
2i z  2z  1  z  2  4i  z   2z  
2
z
1 1  1  2 

1  4  1   ....  1   ... 
4i  z   2z  z 
1 1  5  5
the coefficient of is easily seen to be   , i.e.
z 4i  2  8i

Hence by Cauchy’s residue theorem we have


I  2i 
(sum of the residues within the contour)
 17  5   
 2i      
 24i  8i   6
2
cos 2 
 5  4 cos d  6
0
Thus
Ex. 7.
Use the residue theory to show that
2
d 2a
  a  b cos    where a  0, b  0, a  b
a  b2 
2 3
2 2
0

320
Sol.
2
d
  a  b cos 
0
2

I=
2
d dz
  2
write ei  z, d 
zi
1  i 
a  b  e  e  
0 i

 2 
1 dz
 2
C
 1  1   zi
a  b  z   
 2  z 
4izdz
 where C is the unit circle z  1
 bz  2az  b 
C 2 2

4i zdz
2 C

  z     z   
2
b

a  a 2
 b2  a  a 2
 b2 
where   , 
b b
4iz
 C
f (z)dz where f (z) 
b 2
 z     z  
2 2

 z     z   = 0
2 2

Poles of f(z) are given by


z   and z   are the poles each of order 2. We see from
Therefore
their values above that numerically  is greater than unity while  is less
hence only  lies within the contour C.
Re sidues at the double pole z  
 d  4iz 
 2 
 dz b  z     z 
2

4i   z     2z  z    
2

  
 z  
4
bz   z 

 m 1  n  
 u sin g the formula, Re sidues  
  m  1 
4i    ai
 2. 
b    
 a 2  b2  2
2 3

Hence by Cauchy’s residues theorem we have


C
f (z)dz  2i 
sum of the residues within the Contour C

321
ai 2a
 2i  
a  b2  a  b2 
3 3
2 2 2 2

2a
or I
a  b2 
3
2 2

Ex. 8.
Use the method of contour integration to prove that
  
a cos   a 
 a  cos d  2a 1   , where a  1
  a 2
 1 
.
Sol.
 2
a cos  2a cos 
I  d   2a  2 cos d
 a  cos  0
2
2aei
I  real part of 
0 2a   ei  e i 
d

dz 2az dz
 real part of  1
writing ei  z, d 
.
2a  z  iz iz
C

z
where C is the unit circle z  1

2aiz
 real part of 
C z  2az  1
dz 2

2aizdz
 real part of 

C z
  z  
where   a  a 2
 1 ,   a  a 2
 1
2aiz
 real part of  C
f (z)dz where f (z) 
 z     z  
 z     z  
Poles of f(z) are given by =0
z   and z  
i.e. are the two simple poles.
 
The value of is obviously greater than unity while that of is less.

Thus only the pole lies within the contour C.
z
Residue of f(z) at the simple pole is

322
 lim  z  a  f (z)
z 

2aiz 2aiz
 lim  z  a  
z   z     z     



2ia a  a 2
 1   ai  a

 1
2 a 2
 1
 
 a 2 1
 
Hence by Cauchy’s residue theorem we have
 
 a 
 2i.ai   1
  a 2  1 
 
 a 
 2a 1   which is purely real.
  a  1 
2

 C
f (z)dz
Hence I = real part of
 
 a 
 2a 1  
  a  1 
2

Ex. 9.
Show by the method of contour integration that

1  2 cos 
 5  4 cos d  0
0

Sol.

1  2 cos 
 5  4 cos d
0
Let I=

323
2
1 1  2 cos 

2  5  4 cos d
0
2
1 1  2ei
 real part of
2 
0 5  4 cos 
d

2
1 1  2ei
 real part of
2 
0
5  4 cos 
d

2
1 1  2ei
 real part of
2 
0 5  2  e i  e  i 
d

1 1  2z dz dz
2 C
 real part of . writing e i  z, d 
 1  iz iz
5  2 z  
 z
where C is the unit circle z  1
1 i  1  2z 
 real part of 
2 2z 2  5z  2
C
dz

1 i  1  2z 
 real part of  dz
2 C  2z  1  z  2 
1  1  2z 
 real part of C
f (z)dz where f (z)   i.
2  2z  1  z  2 

 2z  1  z  2 
Poles of f(z) are given by =0
1
z and z  2
2
Thus are the two simple poles.
1
z
2
But only lies inside C.

1  1
z lim  z   f (z)
z   2
1
2 2
Residues at is

324
 1   1  1  2z  
 lim1  z    i. 0
z   2   2  2z  1  z  2  
2

 by Cauchy 's residue theorem, we have

C
f (z)dz  2i  sum of residues within C.

C
f (z)dz 2i  0  0

Hence I  real part of  C


f (z)dz 0

Ex. 10.
By the method of contour integration, prove that

cos 2d a 2
0 1  2a cos   a 2 1  a 2
  1  a  1

Sol.
 2
cos 2d 1 cos 2d
I    1 a  e
0 1  2a cos   a  e  i   a 2
2 i
2 0

2
1 e 2i
 real part of
2   1  ae   1  ae  d
0
i  i

2
1 z2 dz
 real part of 0 .
 1  az  1  
2 a iz
 z
dz
writing ei  z, d 
iz
1 iz 2
 real part of  dz
2 
C 1  az
  z  a
1 iz 2
 real part of  f (z)dz where f (z) 
2 C
 1  az   z  a 
 1  az   z  a 
Poles of f(z) are given by =0

325
z 1 and z  a
a
Thus are the simple poles.
Only z = a lies within the unit circle C as a<1.
The residue of f(z) at z = a.
 lim  z  a  f (z)
za

iz 2 ia 2
 lim  z  a  
za  1  az   z  a  1  a 2
Hence by Cauchy’s residue theorem we have

C
f (z)dz  2i  sum of residues within the contour.
ia 2 2a 2
 2i   which is purely real.
1 a2 1 a2
1
Hence I  real part of  f (z)dz
2 C

1 2a 2
a 2
 . 
2 1 a 2
1 a2

326
Ex. 11
2
cos 2 3 1  p  p2
0 1  2p cos 2  p2  .
1 p
,0  p 1
Prove that

Sol.

2
cos 2 3
Let I 0 1  2p cos 2  p 2
2
1 cos 6
  d
2 0 1  2p cos 2  p 2
2
1 1  e i

2
real part of  1 p e
0
2i
 e 2i   p 2
d

1 1  z6 dz dz

2
real part of C  2 1 
.
iz
write ei  z, d 
iz
1 pz  2 p
2

 z 
where C denotes the unit circle z  1

1 1 z  1  z6 
 real part of  dz
2 i C  1  pz 2   z 2  p 

1 1 z  1  z6 
 real part of  f (z)dz where f (z) 
2 i C  1  pz 2   z 2  p 
 1  pz   z
2 2
 p
Poles of f(z) are given by =0
1
z   p and z  
p
Thus are the simple poles.
z p
The only poles which lie within C are , as p<1.
z  1  z6  1 1  p2
Re sidues at z  p is lim z  p .
z p
   1  pz   z
2 2
 p
 .
2 1  p2

z  1  z6  1 1  p2
and residues at z   p is lim z  p .
z  p
   .
 1  pz 2   z2  p  2 1  p2

327
1  p2
1  p2
So that sum of the residues =
Hence by Cauchy’s residue theorem we have

C
f (z)dz  2i  sum of residues within the contour
1  p2
 2i 
1  p2
1 1
 I  real part of  f (z)dz
2 i C
1 1 1  p2
 real part of .2i 
2 i 1  p2
1 1  p2
 real part of 2.
2 1  p2
1  p2 1  p  p2
  
1  p2 1 p

Ex. 12.
Use the method of contour integration to prove that
 1  2 cos  
n
2
cos n 2
 
n

0
3  2cos 
d 
5
3 5

Sol.
 1  2 cos  
2 n
cos n
Let I 
0 3  2 cos 
d

 1 e  e  i  ein
i n
2
 real part of 
0
3  ei  e  i
d

n
 1 n
1  z   z dz
 real part of  
z
C 1 iz
3 z 
z
dz
writing ei  z, d 
iz

328
1 1 z  z 
2 n

 real part of  dz
i C 1  3z  z 2
where the contour C is the unit circle z  1

1
 real part of  f (z)dz where f (z) 
1 z  z  2 n

i C 1  3z  z 2

3  5
Poles of f (z) are given by 1  3z  z 2  0 i.e., by z 
2
3  5 3  5
let   and  
2 2
3  5
The only pole which lies within C is
2

The residues at  is lim  z    .


 1 z  z  2 n

z   z     z  
 1       3  5 
n
2


  5

Hence by Cauchy’s residue theorem we have


C
f (z)dz  2i  sum of residues within the contour C

 3 5
n

 2i 
5
1
i C
Hence I  real part of f (z)dz

 
n
1 3 5
 real part of .2i 
i 5
2
 
n
 3 5
5

 f (x)dx

5.10 Evaluation of (where no poles on the real axis)
If the function f(z) is such that it has no poles on the real axis and
possibily has some pole in the upper half of the z-plane then we can e valuate

 f (x)dx


329
C
f (z)dz
Consider the integral taken round a closed curve C, consisting
z R
of the upper half CR of the circle and part of the real axis from –R to R.

z R
If there are no poles of f(z) on the real axis, the circle which is
arbitrary can be taken such that there is no singularity on its circumference C R
in the upper half of the plane, but possibly some poles inside the contour C
specified above.
Using Cauchy theorem of residue we have

 f (z)  2i   sum of the residues of f (z) at the poles within C


C
R
i.e.,  f (x)dx   f (z)dz  2i   sum of residues within C 
CR
R
R
Or  f (x)dx   f (z)dz  2i   sum of residues within C 
CR
R
R
 lim  f (x)dx   lim  f (z)dz 2i   sum of residues within C 
R  R  CR
R

Or  f (x)dx   lim  f (z)dz  2i   sum of residues within C   (1)
R  CR


All the poles of f(z) can be found by putting the denominator of f(z)
equal to zero. Then the using the method for calculating the residue at a pole,
we compute the residues only at those poles which lie inside C, and their sum is
taken.
lim  f (z)dz
R  CR
Now if exists, then right hand side in equation (1) above
is perfectly known, and this

 f (x)dx

is determined.
5.11 Jordan’s Inequality.
y  cos 
Consider the relation
 cos 
As increases decreases and therefore the ordinate y decreases.

330
  0 to 
The mean ordinate between
1 
 0
 cos  d

sin 
 .

1
  ,
  0, 2
when the ordinate is clearly cos 0, i.e., equal to 1; and when the
1
sin 
2 ,
1 2
 .
2 
mean ordinate is equal to i.e., equal to

1 2
0  .
2 
Hence when the mean ordinate lies between 1 and
2 sin 
i.e.,  1
 
This is known as Jordan’s Inequality.

5.12 Jordan’s Lemma.


If f  z   0 uniformly as z  ,
then
 e f  z  dz  0,  m  0 
imz
lim
R 
CR
where CR denotes the semi-circle
z  R, I  z   0.

Here R is large and can be made larger so as to include within it all the
singularities and none on its boundary.
lim f  z   0
R 
Since
 f  z  
for z on the large circle.

331
 e f  z  dz
imz

CR
Now,
  eimz f  z  dz
CR

  eimz f  z  dz sin ce f (z) 


CR

  eim  R cos iR sin  Reiid sin ce z  Rei
0

  e  mR sin  Rd
0
/ 2
 2R e  mR sin d
0

/2
 2R e 2mR /  d
0
by Jordan’s inequality
n

m
 1  e mR  which  0 when R   and  0

 e f  z  dz  0
imz
lim
R 
CR
Hence lemma proved.
Ex. 1.
(a) Show by contour integration that
 dx 
0 1 x 2
 .
2
Sol.

1
 f  z  dz
C
where f  z  
1  z2
Consider the integral taken round a
z  R,
closed contour C consisting of upper half of a large circle and real axis
from – R to R.
1  z 2  0 i.e,. z  i,
Poles of f(z) are given by are two simple poles,
Only z = i lies within the contour, the residue there at

332
1
 lim  z  i  .
z i 1  z2
z i 1
 lim 
z i  z  i   z  i  2i

Hence by Cauchy’s residue theorem, we have

 f  z  dz  2i  sum of the residues within C


C

R 1
i.e.,  f  x  dx   f  z  dz  2i  2i
R
CR

R 1 1
or  dx   dz     1
R 1  x 2
1  z2
CR

1
 1 z
CR
2
dz
Now
dz dz
 
CR 1  z2

CR z 1
2

Rd 
 sin ce z  Rei , dz  t Rei d  Rd
0 R 2 1

R
 2  0 as R  
R 1
R  ,
Hence making relation (1) becomes
1

 1  x 2
dx  
 1 1
 dx  .
0 1 x 2
2
Ex. 1. (a)
(b) Apply the calculus of residues to prove that

1 1
  1 x 

2
dx 
4
.

Sol.
See the contour of the last exercise.

333
1
f  z 
 f  z  dz
C
 1 z 2 2

Consider the integral where taken round the


CR
closed contour C consisting of the semi-circle which is upper half of a
z R
large circle , and the real axis from –R to R.

 1 z 
2 2
0
Poles of f(z) are given by
1  z2  0 z  i
i.e. by or each repeated twice.
The only pole within the contour is z=i, and is of the order 2.
Here

1
f  z 
 z  i  z  i
2 2

 z 1
 where   z  
 z  i  z  i
2 2

2
so that  ' z   .
 z  i
3

Re sidue at the double pole z  i,


 ' i n 1  i 
 by .
1!  n  1 !
2 i
 
 2i 
3
4

Hence by Cauchy’s residue theorem we have

334
 f  z  dz  2i  sum
C
of the residues within C

R
 1 
i.e.  f  x  dx   f  z  dz  2i    4 i 
R CR
R
1 1 1
or  dx   dz  . ...  1
R  1 x 
2 2
CR  1 z 2 2
 2

1 dz
Now,  dz  
1 z 2 2 2
CR CR 1  z2

dz Rd
   sin ce z  Rei
CR  z  1
2

2
0 R 2
 1
2

R
  0 as R  .
 R 2  1
2

R 
Hence making ; relation (1) becomes

1 1
 dx  
  1 x  2 2 2

1 1
or  dx  
0  1 x  2 2 4

Ex.1 (c)
(c) Apply the calculus of residues to prove that

dx 3
  .
x  1
3

2 8

Sol.

335
1
 f  z  dz, where f  z  
z  1
2 3
C
Consider taken round the closed
CR
contour C consisting of real-axis and the upper half of a large semi-circle
z  R.

z  1  0, i.e.  z  i   z  i
3 3 3
2
0
Poles of f(z) are given by
z  i
i.e. are the poles each of order 3.
The only pole which lies within C is z=i of order 3.
 z n 1  a 
at  z  a  is
 z a n  1!
n

We know that residue of


1 1
 residue of at  z  i 
 z  i  z  i
3 3

1  d2 1  3
  2 3

2  dz  z  i   16i
z i

Hence by Cauchy’s residue theorem, we have

 f  z  dz  2i  sum of
C
residue within C

R
3
i.e.  f  x  dx   f  z  dz  2i  16i
R CR
R
1 1 3
or  dx   dz  ...  1
 x 2  1  z 2  1
3 3
R CR
8

Now,

336
1 1
 dz   dz
z  1 z  1
2 3 2 3
CR CR

dz
 
 z  1
2 3
CR


Rd
 sin ce z  Rei , dz  Rd
R  1
2 3
0

R
 which  0 as R  
R  1
2 3

R 
Hence making , relation (1) become

1 3
 dx  .
x  1
3

2 8

Ex.1. (d)

x2 1
 dx  
x  1
2
0
2 2
(d) Apply the calculus of residues to prove that
Sol.
See the contour of the last exercise.
z2
 f  z  where f  z   ,
 z2  1
2
C
Consider the integral taken round a
CR z R
closed contour C consisting of the upper half of a large circle and
the real axis from –R to R.

z  1  0
2 2

Poles of f(z) are given by


 z  i and z  i
are the two poles each of order 2. But only z=i lies
within the contour.
z it
To get residue at z=i, put in f(z) then it becomes

337
 i  t
2


 1  2it  t  1  2
t
2

 2i 
  i  t  2  1
2
4t 2

1 i
is
t 2
Here coefficient of which is therefore the residue at z=i.
Hence by Cauchy’s residue theorem, we have

 f  z  dz  2i  sum of
C
the residues within C

R
 i
i.e.  f  x  dx   f  z  dz  2i   
R CR  2
R
x2 z2
or  dx   dz   ..  1
x  1 z  1
2 2 2 2
R CR

Now,

z2
 dz
z  1
2 2
CR

2
z dz
 
CR z2  1
2


R2
  Rd sin ce z  Rei , dz  Rd
R  1
2 2
0

R3
 which  0 as R  .
R  1
2 2

R 
Hence by making , relation (1) becomes

x2
 dx  
x  1
2 2



x2
2 dx  
x  1
2 2
0
Or

x2 1
 dx  
x  1
2
0
2 2
Or

338
Ex.2.
Prove by contour integration that
 1.3.5...  2n  3

dx 1
  . . 1/ 2 2n 1
1.2.3...  n  1 a
 a  bx  2 n n1/ 2
0
2 b

Sol.
1
 f  z  dz where f  z   ,
 a  bz 2 
n
C
Consider the integral taken round a
CR z R
closed contour C consisting of the upper half of a large circle .

 a  bz  2 n
0
Poles of f(z) are given by
a
z  i repeated n times
b
i.e.,
a
zi ,
b
The only pole within the contour is a pole of order n at the
residue at which
 n 1  

1 d  1  a
 dz n 1  n 
 
n
b  n  1 !    bz  i a    b
    z i
 1 . 1 . n  n  1  n  2  ...  2n  2 
n 1


 n  1 ! b n  a
2n 1

 2i 
 b

Hence by Cauchy’s residue theorem we have

 f  z  dz  2i 
C
sum of the residues within C.
R
dx dz
or     2i 
R  a  bx 
2 n
CR  a  bz  2 n

sum of the residues …..(1)

339
dz dz
  
 a  bz 2 n n
CR CR a  bz 2
Now,
dz
 
b z 
2 n
CR a


1
  Rd
 bR a
2 n
0

R
 which  0 as R  
 bR 2  a 
n

Hence when R  , relation  1 becomes

 1 . 1 . n  n  1  n  2  ....  2n  2 
 n 1
dx
  2i 
 n  1 ! b n
ax  2 n 2n 1
  a
 2i 
 b

 1 . 2n  2! .
n 1
1
 2i 
 n  1 !  n  1 !  
2n 1
b 2i a
 1.3.5...  2n  3
 .
2n 1  n  1 ! b1/ 2a n 1/ 2
 1.3.5...  2n  3

dx 1
or   . . n  1/ 2 
2 b 1.2.3...  n  1 ! a
 a  bx  2 n n 1/ 2
0

Particular Case.
Taking a=1, b=1 and proceeding as above step by step we can prove
that
 1.3.5...  2n  3

dx
  .
2n 1.2.3...  n  1 !
0  a  bx  2 n

Ex.3
Use the method of Contour integration to prove that

dx   2n  !
  .
x  b2   2b   n!
2 n2 2n  2 2
0

Sol.

340
1
f  z  ,
 f  z  dz
C
z 2
b 
2 n 1

Consider the integral where taken


z R
round the closed contour C consisting of the upper half of a large circle
and part of the real axis from –R to R.

z  b2 
n 1
0 2

Poles f(z) are given by


z  ib
i.e. are the two poles each of order (n+1).
The only pole which lies which the contour is z=ib and is of order
z  ib z  ib  t
(n+1). To find the residue at put in f(z), then it becomes
1
n 1
 ib  t  2  b 2 
 
 n 1
1 1  t 
  1
 2ib  t   2ib  , t  2ib 
n 1 
2n 1 n 1


1
.
1 
   1
r  n  1  n  2   n  3 ...  n  r  tr
 2ib   2ib 
n 1
t n 1 r 0 r! r

where in coefficient of 1/t is


 1  n  1  n  2  ...  2n 
n
1
 2ib  n! 2ib 
n 1 n

2n!
i.e.
i  2b   n!
2n 1 2

z  ib
which is therefore the residue at
Hence by Cauchy’s residue theorem, we have

 f  z  dz  2i 
C
sum of the residues within the contour.
R
 2n  ! 1
 f  x
R
dx   f  z  dz  2i. i  2b 
CR
2n 1
.
 n!
2

i.e.
R
1 1  2n  ! 2 .... 1
 dx   dz   
x 
2 n 1
z 
2 n 1
 n!  2b 
2 2n 1
R
2
b CR
2
b
or
Now,

341
1 dz
 dz  
z 
2 n 1 n 1
CR
2
b CR z 2  b2
dz
 
z 
2 n 1
CR  b2

1
  Rd sin ce z  Rei
R 2
b 
2 n 1
0

R
 which  0 as R  .
R  b2 
2 n 1

R 
Hence by making , relation (1) becomes

dx 2 (2n)!
 
x 
2 n 1 2n 1

2
b (2b) (n!) 2

dx 2 (2n)!
or  
0 x 2
b 
2 n 1 (2b) n 1 (n!) 2

Ex. 4
Apply the calculus of residues to evaluate

cos x dx

 
x  a 2 x 2  b2
2
, a  b  0.

Sol.
eiz
 f  z  dz where f  z  
C z 2
 a 2   z2  b2 
,
Consider the integral
taken round the closed contour C consisting of the upper half of a large circle
z R
and the real axis from –R to R.

Poles of f(z) are given by


z 2
 a 2   z2  b2   0
i.e.  z  ia   z  ia   z  ib   z  ib   0
i.e., z = ia, - ia, ib, - ib.

342
The poles which lie within the contour are at z = ia and z = ib both
simple.
The residue of f(z) at z = ia
 lim   z  ia  f  z  
z ia

 z  ia  eiz e a
lim 
z ia
 z  ia   z  ia   z 2  b 2  2ia  a 2  b 2 

e b
z  ib,  .
2ib  a 2  b 2 
Similarly residues off f(z) at
1  e b e a 
Sum of residues    
2i  a 2  b 2   b a 

Hence by Cauchy’s residue Theorem we have

 f  z  dz  2i  sum of residues within C


C

R
i.e.,  f  x  dx   f  z  dz  2 i  sum of residues
R
CR

R e ix
ezi dz 1  e  b e a 
or    1
 x 2  a   x 2  b  CR  z2  a 2   z 2  b2 
dx   2i    
R
2i  a 2  b 2   b a 
zi
eiz e dz
Now,  2 dz   2
CR  z  a   z  b  CR z  a z  b2
2 2 2

e zi dz
 
CR z 2
 a2 z 2
 b2 
 e  R sin  .Rd
 sin ce z  Rei .
0
R a R b 
2 2 2 2

2R / 2


 e 2R  /  d,
R 2
a 2
R 2
b 2 0

By Jordan’s inequality

343
2R 
 .  1  eR 
R a R b 
2 2 22
6R

  1  eR 
 which  0 as R 
R 2
 a 2   R 2  b2 

R 
Hence by making , relation (1) becomes
 cix 1  e b ea 
 x
 2
 a 2   x 2  b2 
dx  2i  
2i  a 2  b 2   b

a 

  eb ea 
   
a 2  b2  b a 

Equating real parts, we have


 cos x   eb ea 
 x
 2
 a 2   x 2  b2 
dx 
a  b2
2 
 b

a 

.

344
Ex. 5.
 x 2 sin x
 dx,  a  0, b  0 
0
 x 2  a 2   x 2  b2 
Calculate
Sol.
See the contour of the last exercise.
z 3 eiz
f  z 
 f  z  dz,
C
 x 2  a 2   x 2  b2 
Consider the integral where taken
z  R,
round the contour, consisting of the upper half of the large circle and
the real axis from – R to R.
z 2
 a 2   z 2  b2   0
Poles of f(z) are given by ,
i.e., z  ia and z  ib
are the four simple poles.
The poles which the within the contour are z = ia and z =ib both simple.
lim
z ia
Residue of f(z) at z = ia, is (z-ia) f(z)
ze 3 iz
a 2 e a
lim  z  ia  
z ia
 z  ia   z  ia   x 2  b2  2  b2  a 2 

b2 ea

2  a 2  b2 
Similarly residue at z = ib is equal to
So that sum of these two residues
1
  b 2 e  b  a 2 e  a 
2 a  b 2 2

Hence by Cauchy’s residue theorem, we have

345
 f  z  dz  2i  sum of the residues within the contour
C

R
i.e.,  f  x  dx   f  z  dz  2i  sum of residues
R
CR

x 3eix z 3eiz 1
 a 2e  a  b 2e  b 
R
or  dx   dz  2i    1
R
 x  a   x  b  CR  z  a   z  b 
2 2 2 2 2 2
2i  a  b 2 
2

z 3 eiz
Now, 
Cr  z2  a 2   z2  b2 
dz

z3eiz dz
 
CR z 2  a 2 z2  b2
3
z eiz dz
 z
CR
2
 a2  z 2
 b2 
2
R 


 e  R sin .Rd
R 2
a 2
R 2
b 2 0

R4 / 2

 R a   R b 
2 2 2 2
.2 0
e 2R / d

By Jordan’s inequality,
R 3
 2 
1  e  R  which  0 as R 
R a R b 
2 2 2

R 
Hence by making , relation (1) becomes
x 3 eix i
  x 2  a 2   x 2  b2  dx   a 2  b2   a z  b e 

2 a 2 b

Equating imaginary parts, we have


x 3 sin x 
  x 2  a 2   x 2  b2  dx   a 2  b2   a z  b e 

2 a 2 b

x 3 sin x 
2 
a 2 z a  b2e b 

or 0
 x a   x b 
2 2 2 2
dx 
2 a  b 
2

Ex. 6.
Use the method of contour integration to prove that

346
 dx   b  2c 
  .
x x 
2 2
2bc 2  b  c 
 2
2
b 2 2
c

where b > 0 , c > 0.


Sol.
See the contour of the last exercise.
1
 f  z  dz where f  z    z
C
2
b 2
z 2
 c2 
Consider the integral taken
z  R,
round a closed contour C, consisting of the upper half of a large circle
and the part of the real axis from – R to R.
z 2
 b 2   z 2  c2   0
Poles of f(z) are given by
i.e., z  ib and z  ic
are the two simple poles are the two double
poles.
The poles which lie within the contour are: a simple pole at z = ib and a
double pole at z = ic.
lim  z  ib  f  z 
z ib
Residue at the simple pole z = ib is
1
lim  z  ib  .
 z  ib   z  ib   z 2  c2 
z ib 2

1 i
 
2ib  c  b 
2 2
2b  b  c 2 
2 2 2

z  ic 4
To find the residue at the double pole put z = ic + t in f(z), then
it becomes
1

  ic  t     ic  t  
2 2 2
 b2  c2

347
1

b  c  2ict  t 2   2ict  t 2 
2 2 2

2 1
 2ict  t 2   t 
1  2 2  1  
b c 
  .
2ic 
b2  c2 4c 2 t 2
 2ict  t 2  t 
1  2  ..... 1  
b c 2
 2ic 
 
4c 2  b 2  c 2  t 2

in which the coefficient of 1 / t is


1 i 2ic 
2 
 2 2
4c  b  c   c b  c 
2 2

i.e.,
 3c  b  i 2 2

4c  b  b 
2 2 2 2

which is therefore the residue at z = ic.

So sum of the residue  


i

 3c  b 
2 2

2b  b 2  c 2  4c  b  c 
2 3 2 2 2

i  b  2c 
 .
4bc3  b  c 
2

Hence by Cauchy’s residue theorem, we have

 f  z  dz  2i  sum of the residues within the contour


C

R
i.e.,  f  x  dx   f  z  dz  2i  sum of residues
R
CR

R 1 1
or  dx   dz
x x  z z  c2 
R 2 2
2
b 2 2
c 2
Cr
2
b 2 2

i  b  2c 
 2i.   1
4bc 2  b  c 
2

1
Now,  z
Cr
2
b 2
z 2
 c2 
dz

348
dz

CR
 z 2  b2 z 2  c2
2

dz
 
z  z 
2
Cr
2
 b2 2
 c2
1 
  Rd sin ce z  Reia
R 2
 b2   R 2  c 
2 2 0

R
 which  0 as R .
 R 2  b2   R 2  c2 
2

R 
Hence by making , relations (1) becomes
 dx   b  2c 
 
x  b2   x 2  c 
2 2
2bc3  b  c 
 2 2

Ex. 7.
Using the calculus of residues prove that
 x sin ax 1  ak
0 x2  k2
dx  e
2
where a  0

Sol.

 f  z  dz, f  z  dz,
zeiaz
,
C z2  k 2
Consider the integral where taken round the
closed the contour C consisting of real axis from –R to R, and the upper half of
z R
the large circle .
z 2  k 2  0 ie. z  ik.
Poles of f(z) are given by
z  ik
The only pole within the contour is the simple pole at . The
residue at which is
zeiaz e ka
lim  z  ik  i.e.
z ik  z  ik   z  ik  2

Hence, by Cauchy’s residue Theorem we have

349
 f  z  dz  2i 
C
sum of residues within the contour
R

R
 f  x  dx   f  z  dz  2i 
CR
i.e. sum of the residues
R
xeiax zeiaz e  ka
 x 2  k 2 C z 2  k 2
 dz  2 i  ...  1
R R
2
or
Now,

zeiaz

CR z2  k 2
dz

z eiaz
  dz
CR  z  k2
2


R2
2 
 2 e  aR sin d as z  Rei
R k 0

2R 2 2 2aR / 
R  k 2 0
 2 e d by Jordan 's inequality

R / 2
  e 2a R  /  
a R k  0
2 2 0

R  1  e0aR 
 which  0 as R  ,  sin ce a  0 
a  R 2  k2 

R 
Hence by making , relation (1) redues to

xeiax dx e  ka
 2 2
 x  k
 2 i 
2

Equating imaginary parts we get



x sin ax
x

2
k 2
dx  e  ka

or

x sin ax 1
x
0
2
k 2
dx  e  ka
2

350

x sin ax 
x dx  a ,  a  0  .
0
2
k 2
2e
Particular Case.
Ex. 8.
Use the method of contour integration to prove that

x 3 sin x 
 dx    a  2  ea ,  a  2
0 x 2
a 2 2
 4

Sol.
z 3 eiz
 f  z  dz, where f (z) 
z  a2 
2 2
C
Consider the integral taken round
z R
the closed contour C, consisting of the upper half of a large circle and
the real axis from -R to R.

z  a2   0
2 2

Poles of f(z) are given by


z  ia and z  ia
i.e. are the two poles each of order 2.
z  ia
The only pole which lies within the contour is and is of order
two.
 ia  t  e
i ia  t 
2

z  ia  t in f  z  it becomes
  ia  t  
2 2
 a2
Putting

1
t
in which coefficient of is
ea  3  1   2  a a
2 
a 1    3a 2  ie. e
4a   a   4
which is therefore residue
z  ia
at .
Hence by Cauchy’s residue theorem we have

351
 f  z  dz  2i 
C
sum of the residues within C
R
2  a a
 f  x  dx   f  z  dz  2i.
R CR
4
e . ...  1
i.e.
Now,

z3eiz

CR  z  a 
2 2 2
dz

z 3eiz dz
 CR 2
z2  a 2

R3
   e  R sin  Rd
R 2
a 
2 2
0


R4
 .2  e 2 R / d by jordan 's ineuality.
R 2
a 
2 2

R 3
  1  e  which
R
 0 as R  .
 R a
2

2 3

R
Hence making , relation (1) becomes

x 3 sin x 2  a a
 dx  2i e
0 x 2
a 2 2
 4

Equating imaginary parts, we



x 3 sin x 2  a a
 dx  2 e
0 x 2
a 
2 2 4
or
 3
x sin x 1
 dx     a  2  e  a
0 x 2
a 
2 2 4

352
Ex. 9.
Show by the method for contour integration that

cos m x 
2 
 dx  1  ma  e  ma ,  a  0, m  0  ,
0 a 2
x 2 2
 4a

Sol.
eimz
 f  z  dz, where f  z   ,
 a 2  z2 
2
C
Consider the integral taken
z R
round the closed contour C consisting of the upper half of a large circle
and the real axis from –R to R.

a  z2   0
2 2

Pole of f(z) are given by


z  ia and a  ia
i.e. are the two poles each of order two.
z  ia
The only pole which lies within the contour is at of order 2.
z  ia z  ia
Since the pole is of order 2, hence to know the residue at ,
z  ia  t
put in f(z) , then it becomes
eim ia  t 

a 2
  ia  t  
2 2
.
2
e  ma .eimt e  ma .eimt  t 
  1  2ia 
 2iat  t  2 2 4a 2 t 2

e  ma  2t 
2 2 
 1  imt  ... 1   ...
4a t  2ia 

1 ie  ma  1  ma 
,
t 4a 3
in which coefficient of is easily seen to be which is
therefore residue theorem, we have

 f  z  dz  2i 
C
sum of the residues within C

353
R
ie  ma  1  ma 
 f  x  dx   f  z  dz  2i 
R CR
4a 3
R
eimz eimz 2  ma
or  dx   dz  e  1  ma  .....  1
a  x2  a 
2 2 2
R
2
CR
2
z 4a 3

354
Now,

eimz

CR  a  z 
2 2 2
dz

eimz dz eimz dz
   
z 
2 2
CR a 2  z2 CR
2
 a2

e  mR sin  Rd

R  a2 
2 2
0

/2
R
 .2  e 2mR /  d
R  a2 
2 2
0


 1 e   mR
which  0 as R  
m R  a
2 2 2

R  ,
Hence by making relation (1) becomes
 eixm 2
 dx   1  ma  e ma

a 2
x 2 2
 4a 3

Equating real parts, we have


 cos mx 2
 dx   1  ma  e ma

a 2
x 2 2
 4a 3

 cos mx 
3 
or  dx  1  ma  e  ma

a 2
x 2 2
 4a

Particular cases.
 cos mx 
 i  0 dx  3 
1  ma  e  ma
a 2
x 
2 2 4a
 cos mx 
 ii  0 dx 
1 x  2 2 2e

355
Ex. 10.
Use the Method of contour integration to prove that
 cos mx   ma
 dx  e ,  m  0 .
0 a x
2 2
2a

Sol.
eimz
 f  z  dz,
C
where f  z  
a 2  z2
.
Consider the integral taken round
z R
the contour C, consisting of the upper half of a large circle and real axis
from – R to R.
a 2  z 2  0.
Poles of (z) are given by
i.e., z = ia, z = -ia are the two simple poles, only z = ia lies within the
contour.
at z  ia,  lim  z  ia  f  z 
z ia
Residue
eimz e  ma
 lim  z  ia  
z ia  z  ia   z  ia  2ia
Hence by Cauchy’s residue theorem, we have

 f  z  dz  2i  sum of the residue within the contour.


C

R
e  ma
 f  x  dx   f  z  dz  2i.   1
R CR
2ia
i.e,
Reimx eimz 1
or  2 dx   2 dz  e  ma
R a  x 2
CR a  z
2
a

eimz
Now, 
CR a 2  z2
dz

eimz dz
 
CR a 2  z2

356
 e  mR sin  d

0
z 2
 a 
R 
2 0
 e 2mR sin  d
R a 2

R /2
 2 2 e2mR sin  d
R a 2 0


 2 
1  e mR  which  0 as R .
M R  a 
2

R 
Hence by making relation (1) becomes
 e imx
e  ma
 a 2  x 2 dx  a .
 cos mx e ma
or 0 a  x2
dx 
2a
.

 cos mx 
 0 1 x 2
dx 
2e
Particular Case.
Ex. 11.
Use the method of contour integration to prove that
 x 2 dx 
 

x 2
a 
2 3 8a 3

Sol.
See the contour of the last exercise.
z2
 f  z  dz, where f  z  
z  a2 
2 3
C
Consider the integral taken round
z R
the closed contour C consisting of the upper half of a large circle and
the real axis from -R to R.

z  a 2   0.
2 3

Poles of f(z) are given by


i.e., by z  ai
both of which are the poles of order 3. The only pole
which lies within the contour is z= ia and is of order 3. To find the residue at z
= ia.

357
Put z = ia + t in f(z) then it becomes
 ia  t 
2

  ia  t  
2 3
 a2

3
a 2  2iat  t 2 a 2  2iat  t 2  t 
  1  2ia 
 2iat  t  2 3 8ia 3 t 3

a 2  2iat  t 2  3t 6t 2 
  2ia  4a 2  ...
1 
8ia 3 t 3  

1 1  3  1
is 1 1   3 , i.e,.
3 
t 8ia  2  16ia 3
in which coefficient of which is therefore
residue at z = ia.
Hence by Cauchy’s residue theorem, we have

 f  z  dz  2i  sum of the residue within the contour C.


C
R
1
i.e.  f  x  dx   f  z  dz  2i. 16ia
R CR
3

R x2 z2 1
or  dx   dz     1
x  a2  z 
3 2 3
R 2
CR
2
a 8a 3

z2
 dz
z  a2 
2 3
CR

Now,
z 2 dz
 
CR z2  a 2
3
.

z 2 dz
 
z a  2 3
2
CR

R2 
  Rd
R 2
a 2 3
 0

sin ce z  Rei
dz  R i .id  Rd

358
R 3
 which  0 as R  
 R2  a2 
3

R
Hence by making , relation (1) becomes
 x2 
 dx 

x 2
a 
2 3 8a 3

Ex. 12.
 dx
  a  0 .
0 x  a4
4
Evaluate
Sol.
1
 f  z  dz,
C
where f  z  
z  a4
4
,
Consider the integral taken round a
z R
closed contour C consisting of the upper half of a large circle and the
real axis from – R to R.

z4  a 4  0
The poles of f(z) are given by
i.e., z 4  a 4  a 4 e i  a 4 e 2ni i
1
 2n 1 i
or z  ae 4

1 3
i 
z  ae 4
and z  ae 4
 for n  0, and 1
are the only two poles which lie within the contour.
 a 4   4 .
Let denote any one of these poles than
 
1
i  1 
ae 4
is  
 d  z4  a 4   1
 dz  z ae 4 i
Residue of f(z) at
1
1 4 i
 e
4a 3

359
e  i / 4
ae3i / 4  .
4a 3
Similarly residue at
1 ei / 4  ei / 4
 Sum of residues  
2a 3 2

1  i
 3
.i sin  
2a 4 2 2a 3

Hence by Cauchy’s residue theorem, we have

 f  z  dz  2i  sum of the residue within the contour C.


C
R
 i 
i.e.  f  x  dx 
 f  z  dz  2 i    
3
R CR  2 2a 
R dx dz 
or   4    1
R x  a
CR z  a
4 4 4
2 2a 3

dz
Now, z
CR
4
 a4

dz

CR
 z  a4
4

dz

CR
 4
z a4
Rd


0 R4  a4

R
 4 4 which  0 as R  .
R a

R  ,
Hence by making relation (1) becomes
 dx 
  x a44

2 2a 3

 dx   2
or 
0 R a
4 4

2 2a 3

4a 3

 dx  2

0 x 1
4

4
Particularly,

360
5.11 Theorem.
    of the circle z  a  r.
Let AB be the arc
lim  z  a  f  z   k,
z 0
If where k is a constant; then
lim
r 0  f  z  dz  i     k
AB

the integration along AB taken in anti-clock directions.


lim  z  a  f  z   k,
z a
Since
 lim   z  a  f  z   k  0
z a
since k is constant,
Or
 z  a  f (z)  k  where z  a is small and  0 as z  a or as r  a
 z  a  f (z)  k  (z) where (z) 
Or
k  (z)
f (z) 
za
So that
Hence
k  (z)
AB
f (z)dz 
AB za
dz

dz (z)
 k  dz
AB z  a AB z  a

i
dz  ire
But  AC za

 rei
d

z  a  rei on AB, dz  irei d


Since

 i  d  i     

(z)
Hence  f (z)dz  i      k   dz
AB AB z  a

(z)
or AB f (z)dz  i      k  AB z  a dz
(z)
so that  f (z)dz  i      k   dz
AB AB z  a

361
(z)
 dz
AB z  a


  rd sin ce (z) 
 r


  d

       0 as  0,
and  0 as z  a as r  0

Thus when r  0 and so we have


lim
r 0

AB

f (z)dz  i      k  0

or lim  f (z)dz  i      k
r 0 AB

362
5.12 Theorem
    z R
Let AB be the arc of the circle
lim zf (z)  R
z 
If where k is constant; then
lim  f (z)dz  i      k
R  AB

The integration along AB taken in anti-clock direction.


lim zf (z)  k
z 

Since

363
lim  zf (z)  k  0 as k is cons tan t.
z 

or  zf (z)  k  for l arg e values of z.


or zf (z)  k  (z) where (z)  
k  (z)
or f (z) 
z
k  (z)
AB f (z)dz  AB z dz
dz (z)
 k  dz
AB z AB z

i
dz  i Re d
AB z  Rei sin ce z  Re on AB, dz  i Re d
i i
But 

 i  d  i(   )

(z)
Hence  f (z)dz  i      k   dz
AB BA z

(z)
or  f (z)dz  i      k   dz
AB AB z

(z)
  f (z)dz  i      k   dz
AB AB z

(z)
 dz
AB z
1
  dz sin ce (z) 
AB z
 1
  Rd sin ce z  R, dz  Rd
 R

  d

       0 as  0 and  0 as z  

R   , 0 and so we have


Thus when
lim
R 
AB

f (z)dz  i      k  0

or lim  f (z)dz  i      k
R  AB

5.13 Indented Semi-Circular Contour.

364
When the poles of f(z) lie on the real axis and also within the semi-
circular region, then those which lie on the real axis can be avoided by drawing
C r , C'r
small semi-circles , etc. about those poles as centres and small radii r and
r’ in the upper half of the plane.
This method is said to be “indenting the semi-circular contour” . when
the semi-circle of radius R has been indented, then f(z) is regular along this
C f (z)dz
modified contour C, and the integral can be evaluated by Cauchy’s
residue theorem, and by using theorems stated and proved in the last two

articles (
 5.11 and
 5.12)

 f (x)dx

Consequently the value of the integral immediately follows.
The following examples illustrates the method.
Ex. 1.

cos 2ax  cos 2bx
 dx    b  a 
ab0 0
x2
Show that, if , then
Sol.
e 2iaz  e 2ibz
 f (z)dz,
C z2
Consider the integral where f(z) = ,
Taken round the contour C consisting of the upper half of the large
z R
circle and the bounding diameter indented at z = 0, and let r be the
radius of indentation. The modified contour is the upper half of the large
z R z r
circle the upper half of the small circle , and the two lines along
the real axis joining the ends of these two circle.

e 2iaz  e 2ibz
z2
Poles of f(z) are given by =0
z 2  0 or z  0
i.e. thus we see that there are no poles within the
modified contour C.
Hence by Cauchy’s residue theorem we have

365
 f (z)dz  2i  sum of
C
the residues within the contour

r R
i.e.  f (x)dx   f (z)dz   f (x)dx   f (z)dz  0  (1)
R Cr r CR

e 2iaz  e 2ibz
Now, CR
f (z)dz  
CR
z
2
dz

e 2iaz  e 2ibz
 2
dz
CR
z
 e 2aR sin   e 2bR sin 
 2
Rd sin ce z  Rei
0 R
2  2 4aR   4bR  
R 0 
 e e d

By Jordan’s inequality

2   
 
R  4aR
 1  e 2aR  
4bR
 1  e 2bR  

which   as R  

 e 2iaz  e 2ibz 
we have lim  zf (z)  lim z 
z 0 z 0
 z2 
 2i  a  b  z  2  a 2  b 2  z 2  ..... 
 lim  
z0 z
 
 2i  a  b 
 lim  f (z)dz  i    0   2i  a  b 
r 0 C r

 2  b  a 

366
R  r0
Hence by making and , relation (1) becomes
0 
 f (x)dx  2  b  a    f (x)dx  0  2i  0
 0

0 e 2iax
e 2ibx
 e 2ibx
 e 2iax

 dx   dx  2  b  a 
x2 0 x2
2iax
e  e2ibx e
2iax
 e 2ibx
or 
0 x2
dx  0 x 2 dx  2  b  a 
2iax
e  e2ibx e
2iax
 e 2ibx
or 
0 x2
dx  0 x 2 dx  2  b  a  (writing  x for x in first term)
 e  e2iax e2ibx  e2ibx  1
2 iax
or  
2

2  x 2 dx    b  a 
0
 

cos 2ax  cos 2bx
or  dx   b  a 
0
x2

Ex. 2.
 sin mx 
0 x

2
By contour integration, prove that
Sol.
eimz
 f (z)dz,
C
f (z) 
z
Consider where
The singularity of f(z) is at z = 0.

z R
Take the contour C consisting of a large semi-circle indented at z
= 0, let r be the radius of indentation. There is no singularity within the given
contour.
Hence by Cauchy’s residue theorem we have

 f (z)dz
C
=0

367
r R
i.e. R
f (x)dx   f (z)dz   f (x)dx   f (z)dz  0  (6)
Cr r CR

 eimz 
Since lim zf (z)  lim z  1
z 0 z 0
 z 
 lim  f (z)dz  i    0  .1  i
r 0 Cr

The negative sign is taken because the contour C r is taken in clockwise


direction.

Ex. 3
(a) By integrating round a suitable contour, prove that
 log x 
2
 3
0 1 x2

8
Sol.

368
 log z 
2

f (z) 
 f (z)dz,
C 1  z2
Consider the integral where
z R
taken round the contour C, consisting of the upper half of a large circle ,
z r
the upper half of a small circle , and the lines joining their ends.

1  z 2  0 i.e. z  i
The singularities of f(z) are given by
zi
Thus the only singularity within the contour is a simple pole at the
residue at which is
  log z  
n

lim  z  i  
r i
 1  z 2 
  log z  n 
 lim  
 z  i 
z i

 log e 
2
i
2  
 log i 
n 2
i  2
2
    
2i 2i 2i 8i
Hence by Cauchy’s residue theorem we have

C
f (z)dz  2i  sum of the residue within C.
r R  2 
 R xdx  Cr f (z)dz  r f (x)dx  CR f (z)dz  2 i  
 8i 

Now, we see that


z  log z 
2

lim zf (z)  lim


z  z  1  z2
2 2
 log   log z 
z3    
z z  log z
 lim  2   lim  0 Since lim  0.
z  1 z z  1 1 z  z

z3 z
 lim  f (z)dz  0
R  0 CR

369
z  log z 
2

Also, lim zf (z)  lim


z 0 z 0 1  z2
  log  
2
1
 lim Putting z 
0 1  2

 0 as before
Hence making R  , r  0, relation (1) becomes
 log x   log x 
2 2
0   2 
  1 x2
dx  
0 1 x2
dx  2i 
 8i 

 log  x  
2
 log x 
2
  2
or  0 1 x2
dx  
0 1 x2
dx  
4
putting  x for x in first int egral.

 log  xe  
2
 log x 
i 2
  2
or 
0 1 x2
dx  
0 1 x2
dx  
4
Since ei  1

 log x  i  dx    log x 
2
 2
or  0 1 x2 0 1 x2
dx  
4
2  log x   2  i2 log x
2
 2
or  
0 1 x2 4
Equating real parts we have
 log x 
2
  dx 3
2 dx  2  
0 1 x2 0 1 x2 4
 log x 
2
  3
i.e 2 dx  2 
0 1 x2 2 4
 log x 
2
 3  3 3
or 2 dx    
0 1 x2 4 2 4
 log x 
2
 3
0 1 x2
dx 
8

Ex. 3 (a)
(b) Use the method of contour integration to prove that
 log x 1
 dx  
0
 1 x 
2 2 4

Sol.

370
log z
 f (z)dz,
C
 1 z 
2 2

Consider the integral where f(z) =


Taken round a closed contour C consisting of the upper half of a large
z R
circle , and the real axis from –R of R indented at z = 0 and let r be the
radius of indentation.

 1 z 
2 2
0
Poles of f(z) are given by
z  i
i.e. are the two poles each of order 2.
But only the pole z = i lies within the contour.
To know the residues of f(z) at the double pole z = i, put z = i+t in f(z)
then it becomes
log  i  t 

 1  i  t 
2 2

 t
log i  log 1  
  i
 2it  t 2 
2

 t t2 
log i    2  .......  2
 i 2i   t
 1  2i 
4t 2
 t t2  t 
log i     .....  1   .... 
 i 2  i 
4t 2

1 i
is
t 4
In which coefficient of which is therefore residues within C.
Hence by Cauchy’s residue theorem we have

371
 C
f (z)dz  2i  sum of the residue within C.
r R i
R
f (x)dx   f (z)dz   f (x)dx   f (z)dz  2 i
Cr r CR 4
 (1)

log z
Now, lim zf (z)  lim z.
z  z 
 1 z  2 2

 
z2 log z 
 lim  .
z  
 1  z 2  z 
2

 
 1 3
1 log z  log z 3
 lim  .   0 because lim 0
z  
 1  1  3 z 
3
z2  z3
  z 2  
 lim  f (z)dz  i    0  0  0
R  CR

log z
Also, lim zf (z)  lim z.
z 0 z 0
 1 z  2 2

1
 lim .  z log z   0 because lim z log z  0
z 0
 1 z  2 2 z 0

 lim  f (z)dz  l    0  0  0
r 0 Cr

R   and r  0,
Hence making relation (1) becomes
0  i
 
f (x)dx   f (x)dx
0
 2i
4
0 log x  log x 1
 dx   dx   
 1 x   1 x 
2 2 2
 2 0 2

Putting –x for x in the first integral we have


 log   x   log x 1
 dx   dx   
 1 x 
0
 1 x 
2 2 0 2 2 2

log  xe 

i
log x  1
or  1  x dx   1  x dx    sin ce ei  1
0
   2 2
 0 2 2 2
 i  log x  log x 1
or  dx   dx  
0
 1 x  2 2 0
 1 x 
2 2 2

Equating real parts, we have

372
 log x 1  log x 1
2 dx    or  dx   
0
 1 x 
2 2 2 0
 1 x 
2 2 4

Ex. 4
Use the method of contour integration, evaluate
 log  1  x 
2

0 x1 dx, 0    1.
Sol.
log  1  x 2 
f (z) 
 f (z)dz,
C x1
Consider the integral where
z1  0 or by z  0
Singularities of f(z) are given by
To avoid the singularity z = 0, take the contour C made up of the upper
z R
half of the large and the real axis from –R to R indented at z = 0, and let
r be the radius of indentation. The function f(z) is now analytic within and on
C; hence by Cauchy’s theorem we have

 f (z)dz  0
C

r R
R
f (x)dx   f (z)dz   f (x)dx   f (z)dz 
Cr r CR
0 ---------- (1)
Now, we see that

373
log  1  z 2 
lim zf (z)  lim z where z  1
z0 z0 z1
1
z 2  z 4  ...
 lim 2 where 0    1
z 0 z
 0 Since 0    1
 lim  f (z)dz  i    0  0  0
r 0 Cr

log  1  z 2 
Also lim zf (z)  lim z
z  z  z1
 1 
log z 2  log  1  2 
 z 
 lim
z  z
 log z 1  1 1 
 lim  2z1    2  4  ....  
z 
 z z z 2z 
log z
 0 sin ce lim 0
z z 

 lim  f (z)dz  i    0  0  0
R  CR

r  0 and R  
Hence letting , the relation (1) becomes
log  1  x 
2
 
 
f (x)dx  0 or  x1 dx  0
log  1  x   log  1  x 
2 2
0
or  1
dx   dx  0
 x 0 x1
log  1  x   log  1  x 
2 2
0
But  x1 dx  0 x1  1 1 writing  x for x

 log  1  x 2 
 dx Since e  i  1
 e  i 
 1
0
.x

 log  1  x 2 

i
 e dx
0 x1
 log  1  x 2   log  1  x 
2

Hence  ei  dx   0
0 x1 0 x1
 log  1  x 
2

or 1 e   i
0 x1
dx  0

 log  1  x 
2

or  0 x1
dx  0

Ex. 5.

374
(a) Use the method of contour integration to prove that
 xb  b
 dx  sec ,  1  b  1
0 1 x 2
2 2

Sol.
zb

C
f (z)dz, f (z) 
1  z2
Consider the integral where
Taken round a closed contour C consisting of the upper half of a large
z R
circle and the real axis –R to R indented at the origin. Let r be the
radius of indentation.
1  z 2  0 or z  i
Poles of f(z) are given by
The only pole z = i lies inside the contour.
lim  z  i  f (z)
z i
Residue at z = i is
 i
b
zb
 lim  z  i  . 
z i  z  i  z  i 2i

 
ib
b
1 i 2 e 2
 e 
2i 2i

Hence by Cauchy’s residue theorem we have

 f (z)dz  2i  sum of residues


C

ib
r R e 2

R
f (x)dx   f (z)dz   f (x)dx   f (z)dz  2i.
Cr r CR 2i
z b dz
Now, CR
f (z)dz  
CR
f (z) dz  
CR
z2  1
zb
 2
dz
CR
z 1
Rb 
R 2  1 0
 Rd sin ce dz  Rd

R b 1
 2 which  0 when R  , sin ce b  1
R 1

375
b
z
Also,  CR
f (z)dz  
Cr
z2  1
dz

b
z
 2
dz
Cr
1 z
rb
 dz
Cr 1  r 2

r b 1
  which  0 as r  0 sin ce b  1  0
1  r2
R   and r  0, (1)
Hence by making becomes
0  2i ib 2
 
f (x)dx   f (x)dx 
0 2i
e
b b
0 x x x ib
or  1  x 2 0 1  x 2 dx  e 2
dx 

Putting  x for x in the first int egral we have


 x 
b
  xb ib
or 0 1 x 2
dx  
0 1 x2
dx  e 2

  xe  i b
 xb ib
or 1 x2
0 0
dx  
1 x 2
dx  e 2 Since ei  1
b
 x
 1  eib  dx
ib
or   e 2
0 1 x2

Equating real parts we have


xb b
0 1  x 2   1  cos b  dx   cos 2
1
 cos b
 x
b
2  b
or  dx   sec
0 1 x2 1  cos b 2 2
Ex. 5
(b) Use the method of contour integration to prove that
 x a 1 1  a 
0 1 x 2
dx   csc   , 0  a  2
2  2 

z a 1
f (z)  where 0  a  2
1  z2
Let

376
Then z = 0 is a singularity of f(z0, for 0<a<1. and poles are given by
1  z 2  0, i.e., z  i
are two simple poles.
Choose the closed contour C, consisting of the real axis from z = 0; let r
be the radius of indentation.
The pole z = i only lies within C.
Residue at the simple pole z = i is
lim  z  i  f  z 
z i

z a 1 z a 1
 lim  z  i   lim  z  i 
z i z 2  1 z i  z  i  z  i
 i
 
a 1 a
1 a 1 i 1 i a
   i   e 2  e 2
2i 2 2 2
Hence by Cauchy’s theorem of residues we have

 f (z)dz  2i  sum of residues


C

r R  1 i a 
R 
f (x)dx   f (z)dz   f (x)dx   f (z)dz  2 i    e 2 
Cr r CR
 2 
 (1)
a 1 a
z z
Now, lim zf (z)  lim z   0, Since 0  a  2
z  z  1  z 2
1  z2
 lim  f (z)da  0
r 0 C r

R  , r  0
Hence making , relation (1) becomes
0  i a
 
f (x)dx   f (x)dx  ie
0
2

0 x a 1  x
a 1
i a
or  1  x 2 dx  0 1  x 2 dx   ie 2
  x 
a 1
x a 1
0
But  dx   dx writing  x for x
 1  x 2 0 1 x2

 1  1
a
 x a 1
 dx
0 1 x2

 
 e  i a
x a 1
dx Since ei  1
0 1 x2
a 1
 x
 eia  dx
0 1 x2

377
 x a 1  x
a 1
i a
Hence  e  ia  dx   dx  ie 2
0 1 x 2 0 1 x 2

a 1
 x
  0 1  x 2 dx  iei 2
a
ia
or 1  e

e i a
2
e
 i a
x a 1
2
 1 
or,
2i 0 1 x 2
dx  
2
a 1
 a   x 1
or sin    dx  
 2  1 x
2
0 2
 x a 1 1  a 
or 
0 1 x2
dx   csc  
2  2 

Ex. 5
(c) Use the method of contour integration to prove that
 xa   1 a   a 
 dx  sec  ,  1  a  3;
0
 1 x  2 2 4  2 

Sol.
za
f (z) 
 1 z  2 2

Let
Since -1 < a < 3, therefore z = 0 is a singularity of f(z). Poles of f(z) are
 1 z 2 2
z  i
given by = 0, i.e. are two poles each of order 2.
za
  i and   i, then f (z) 
 z     z  
2 2

Let .
Choose the closed contour C consisting of the real axis from –R to R
z R
and the upper half of the large circle indented at z = 0 and let r be the
radius of indentation.
z
The only pole which lies within C is at .
z
The residue at the double pole is

378
 za 
d 
  z   n 1  a 
2
 by formula
 dz   n  1
 
  z 
 az a 1  z    2  2z a  z    
 
 z  
4
  z 
a a 1       2 a     
2


   
4

1  a  2  n
         
   
2
  

 
a
1 i
  1 a  e 2
4

 
a
1 i
  1 a  e 2
4i
1 ia 
  1 a  e 2
4i

Hence by Cauchy’s residue theorem we have

 f (z)dz  2i   sum of residues within C 


C
r R 1 ia 
or  f (x)dx   f (z)dz   f (x)dx   f (z)dz  2 i   1  a  e 2  (1)
R Cr r CR 4i
za z a 1
Now, lim zf (z)  lim z.  lim  0, Since  1  a  3
z  z 
 1 z  2 2 z 
 1 z  2 2

 lim  f (z)dz  i    0   0, int egration anti  clock.


R  CR

za z a 1
Also, lim zf (z)  lim z.  lim 0 Since a  1 is positive.
z 0 z 0
 1  z 2  z 0  1  z  2
 lim  f (z)dz  i    0  0  0, int egration clockwise.
r 0 CR

Hence making R  , r  0, relation (1) becomes


0  1 ia 
 f (x)dx  0 f (x)dx  2   1  a  e 2
0 xa  xa 1 ia 
or  dx   dx    1  a  e 2
 1 x2   1 x2 
2 2
 0 2

379
 x 
a
0 xa 
But  dx   dx writing  x for x

1 x  2 2 0
 1 x  2 2

 1
a
 xa
 dx
0
1 x  2 2

 xa

ia
e dx Since ei  1
0
 1 x  2 2

 xa  xa 1 ia 
 eia  dx   dx    1 a  e 2
 1 x   1 x 
2 2 2
0 2 0 2

xa 1
e  1 
 ia 
or ia
dx   1 a  e 2
1 x 
2
0 2 4

or
2 
1 ia  2 ia  2
e e  0
 xa
 1 x  2 2
dx 
1
4
1 a 

 a   x
a
1
or cos   . dx    1  a 
 2   1 x2 
2
0 4
 xa 1  a 
or  dx    1  a  sec  
0
 1 x  2 2 4  2 

Ex. 6.
eiz log  iz 
z2  4
By integrating , prove that
 2 cos x log x   sin x 1 2

0 x2  4
 e log 2
2
Sol.
eiz log  iz 
C
f (z)dz, where f (z) 
z2  4
Consider the integral , taken round a
z R
closed contour C, consisting of the upper half of a large circle and the
real axis from –R to R indented at the origin, and let r be the radius of
indentation.
z 4  4  0 i.e. z  2i
Poles of f(z) are given by both simple, of which
only z = 2i lies within the contour.

380
lim  z  2i  f (z)
z  2i
Residue at z = 2i is
eiz log  iz 
 lim  z  2i 
z  2i  z  2i   z  2i 
1 2
 e log 2.
4i
Hence by Cauchy’s residue theorem we have


C
f (z)dz  2i  Sum of residues within the contour
r R e 2
i.e.  f (x)dx   f (z)dz   f (x)dx   f (z)dz  2 i  log 2.
R Cr r CR 4i
eiz log  iz 
Now, lim zf (z)  lim z
z  z  z2  4
zeiz
 lim 2 log  i   log z 
z  z  4 

zeiz 
log e 2  log z 
 i
 lim 2
z  z  4   
iz
ze  i
 lim 2   log z 
z  z  4  2 
 
 i z 1 log z 
 lim  cos z  i sin z     . 
2 z 2
 4 4
1 2 z 
z 

 z 
  
  cos z  i sin z  i  .0  1.0 
 2 
0
log z
Since lim 0
z  z

381
Hence lim  f (z)dz  i    0  0  0
z  CR

zeiz   
Also, lim zf (z)  lim  i  log z  see above
z  z 2  4
z 0
 2 
   1  
 eiz   log   
 lim  2  z  
z 0  z  4
 i z  1 
 2 z 
  

1  
  i .0  0   0
4 2 
Hence lim  f (z)dz  i    0  .0  0
r 0 Cr

R  , r  0
Hence by making , relation (1) becomes
0  e2
 f (x)dx  0 f (x)dx  2i 4i log 2
0 e log   x   e log  ix 
ix ix
1
i.e.,  dx   dx  e2 log 2.
 x 4
2 0 x 4
2
4
Putting  x for x in the first int egral we have
 e log  ix 
ix
 eix logix 1 2
0 x 2  4 0 x 2  4 dx  2 e log 2
dx 

        

    cos x  isin x   i  log x     cos x  isin x   i  log x  
or     2     2  dx  1 e2 log 2
  
0
 x 4
2
  x 4
2
 2
    

Equating real parts we have
 2cos x log x   sin x 1
0 x 4
2
 e2 log 2
2
Ex. 7.
Apply the calculus of residue to prove that
 sin mx  e  ma  2

0
x  x2  a2 
2
dx  4 
2a 3 
4a 
m   , where m  0, a  0
a

Sol.

382
eimz
f (z) 
 f (z)dz z  z2  a 2 
C f (z)dz
2

C
Consider , where , Evaluate ,
taken round a closed contour C consisting of the upper half of a large circle
z R
indented at z = 0, let r be the radius of indentation at the origin.

z  a2   0
2 2

The poles of f(z) are given by


 z  ia   z  ia 
2 2
0
i.e.
Thus only singularity of f(z) within the contour is a pole z = ia of order
two.
The singularity z = 0 is avoided by indentation.
To calculate residue at z = ia write z = ia + t in f(z), then it becomes
eim  i  t 
, where t is small.
 i  t  t 2  2i  t 
2

1 2
e  am eimz t   t 
 . 1   1  
4a 2i t 2  ai   2ai 
ie  am  t  t 
 3 2  1  imt  ... 1   ....  1   .... 
4a t  ai   ai 
ie  am  2t 
3 2 
 1  imt  1   as t is small
4a t  ai 
ie  am   2  2m 2 
 1   im   t  t  ...
4a 3 t 2   ai  a 

1 ie  am   m 2 
  
t 4a 3  i ai 
Obviously coefficient of is
e  am
  ma  2 
4a 3
i.e.
Hence by Cauchy residue theorem we have
r R e am
i.e.  f (x)dx   f (z)dz   f (x)dx   f (z)dz  2 i   ma  2   (1)
R Cr r CR 4a 4

383
eimz 1
Now, lim zf (z)  lim z 
z 0 z 0
z  z2  a 2 2
 a4
1 
 lim  f (z)dz  i    0  . 
r 0 Cr a4 a4
Also, 
CR
f (z)dz 
CR
f (z) dz

eimz dz eimz dz
 
z  z2  a 2  z  z2  a 2 
CR 2 CR 2

eimz dz  e  mR sin 
  Rd
R  R2  a2 
z 
CR 2 0
2
z  a2
2mR 

 e
 2 2
d by Jordan 's inequality.
R  a2 
0 2 2

 mR  R 2  a 2  which  1  e  mR   0 as R  
2

R  , r  0,
Hence making we have
0 i  e  am
  am  2 
a 4 0
f (x)dx   f (x)dx  2  i 
 4a 4
 i ie  am
i.e.,  f (x)dx  a 4  2a 4  am  2 
 eix i ie  am
4 
i.e  x x 2  a 2 2 dx   am  2 
 
4
a 2a

Equating imaginary parts we have


 sin mxdx  e  am
    am  2 
x  x2  a2 
2
 a4 2a 4
 sin mxdx  e am
or     am  2 
x  x2  a2 
2
0 2a 4 4a 4

Ex. 8
Apply the calculus of residues to prove that
sin 2 mxdx  2ma
3 
 1  2ma  ,  m  0, a  0 

 0

x 2  a 2  x 2  4a
e

384
Sol.
 sin 2 mxdx 1  1  cos 2mx
0
x 2  a 2  x 2  2 0 x 2  a 2  x 2 
 dx
We have
1 1  eimz
f (z)  ,
 f (z)dz 2 z2  a 2  z 2 
C
Consider the integral , where taken
z R
round a closed contour C consisting of the upper of a large circle , the
real axis from –R to R, indented at z = 0, and let r be the radius of indentation.
The singularity z = 0 is avoided by indentation.
a 2
 z 2   0 or by z  0, z  ia, z  ia
Poles of f(z) are given by
z  ia, sin ce a  0
The only pole within the contour is at
lim  z  ia  f (z)
z ia
The residue at z = ia is
1  e 2imz
 lim  z  ia 
z ia z 2  z  ia   z  ia 
e 2ma  1

4ia 3
Hence by Cauchy residue theorem we have

385
 f (z)dz  2i  Sum of residues within the contour
C

r R e2ma  1
i.e.  f (x)dx   f (z)dz   f (x)dx   f (z)dz  2i   (1)
R Cr r CR 4ia 3
Now, 
CR
f (z)dz   f (z) dz
CR

1 1  e 2imz 1 e 2imr

 dz   dz
CR 2 z a z 
2 2 2 CR 2 2
z z  a2  
1
 1 e  Rd

2mR sin 
 by Jordan 's inequality
R  R2  a2 
2 0

 
  1  e 2mR 
R  R  a  2mR  R  a 
22 2 2 2 

both the terms  0 as R  


1  e2imz
lim zf (z)  lim z
z 0 z0 2z 2  a 2  z 2 
2imz  2m 2 z 2
 lim
z 0 2z  a 2  z 2 

im

a2
 im  m
 lim  f (z)dz  i    0    2    2
r 0 C r
 a  a

K   and r  0
Hence making relation (1) becomes
0  m e 2ma  1
 f (x)dx  0 f (x)dx  a 2  2i  4ia 3
m  2ma
 f (x)dx  a 2  2a 3  e  1

or

 1 1  e 2imx  2ma
or  2 x 2  a 2  x 2  dx  2a 3  e  1  2ma 
.

Equating real parts we have

386
1 1  cos 2mx 
dx  3  e 2ma  1  2ma 

 2 x a x 
2 2 2
2a
sin 2 mxdx  2ma
 x 2  a 2  x 2   2a 3  e  1  2ma 

i.e.

sin 2 mxdx  2ma


 e  1  2ma 

or  0
x2  a2  x2 

4a 3

Ex. 9
(b) Use the method of contour integration to prove that
 x a 1 2
csc  a  cos
 2a    , 0  a  1.
0 x  x 1
2
dx 
3 6

Sol.
z a 1
f (z)  2
z  z 1
Let
z a 1  0 or z  0 when a  1  0
Singularities of f(z) are given by and
1 i 3
z 2  z  1  0,i.e. z 
2
poles by are two simple poles.
1 i 3 i 1 i 3  i
  e 3 and    e 3
2 2 2 2
Let

But only lies within the contour.

Residue at the simple pole z = is
lim  z    f (z)
z 

 
a 1
i 3
a 1
  
a 1 e
z
 lim  z    
z   z     z      i 3
z a 1 za
Also lim zf (z)  lim z  lim 0
z 0 z 0 z 2  z  1 z 0 z 2  z  1
Since a is positive
 lim  f (z)dz  i    0  .0  0
z 0 Cr

r  0, R  
Hence making , relation (1) becomes

387
 
a 1
0  1 i


f (x)dx   f (x)dx  2i 
0
i 3
e 3

x a 1 2i i  3
 
 a 1
or  x 2  x  1 i 3 e
dx 

2i
  x a 1 x a 1
a 1
i  0
or  e 3  dx   dx
i 3 0 x2  x 1  x 2  x  1

  x  dx
a 1
 x a 1 0
 dx  
0 x  x 1
2  x 2  x  1

Writing –x for x in the second integral

  1  x  dx
a 1 a 1
 x a 1
 dx  0 x 2  x  1
0 x2  x 1
  1  e i   x 
a 1 a a 1
 x a 1
 2 dx   dx Since  1  ei
0 x  x 1 0 x2  x 1
 x a 1  x a 1
 2 dx  eia  2 dx
0 x  x 1 0 x  x 1

Equating imaginary parts we have


 x a 1 2 1
sin a dx   sin  a  1
0 x  x 1
2
3 3
2  1 
 cos    a  1  
3 2 2 
2 1
 cos  2a  1 
3 6

Ex. 10
Prove by contour integration.
 x a 1  x a 1 
 0 1 x
dx   cot a and 
0 1 x
dx 
sin a
,0  a  1

Sol.
z a 1
C f (z)dz f (z) 
1 z
Consider the integral , where taken round the
closed contour C consisting of real axis from –R to R, and upper half of a large
z R
circle indented at z = 0, z = 1, the radii of indentations being r and r ’
respectively.

388
The singularities of f(z) are z = 0,z = 1 which have been avoided by the
indentation, so there are no singularities within the contour.
Hence by Cauchy residue theorem we have
r 1 r ' R
i.e.  f (x)dx   f (z)dz   f (x)dx   f (x)dx   f (z)dz  0  (1)
R Cr r r CR
a 1 a
z z
Since lim zf (z)  lim z  lim  0, 0  a  1.
z  z  1 z z  1 z
 lim  f (z)dz  i    0  .0  0
R  0 CR

 z.z a 1   za 
Again, lim  zf (z)  lim    lim    0, a  0
z 0 1  z
z 0
  z 0 1  z 
 lim  f (z)dz  i    0  .0  0
r  0 Cr

 z a 1 
Also, lim   z  1 f (z)  lim   z 1   1
z 1 z 1
 1 z 
 lim
'  f (z)dz  i    0  .  1  i
r  0 Cr'

Hence making R  , r  0, r '  0, we have (1)


0 1 
 
f (x)dx   f (x)dx   f (x)dx  0
0 1

or 

f (x)dx i  0
x a 1

or  1  x dx i  0
a 1 a 1
0 x  x
or  1  x 0 1  x dx  i
dx 

Putting –x for x in the first integral we have

389
 1
a 1
 x a 1  x
a 1

 0 1 x
dx  
0 1 x
dx  i

 e 
i a 1
x a 1 x a 1

or 0 1  x dx  0 1  xdx  i
 i ai a 1 a 1
e e x  x
or 0 1  x dx  0 1  xdx  i
a 1 a 1
 x  x
 e 
 ai
dx   dx  i Since e  i  1
0 1 x 0 1 x

Equating imaginary and real parts we have


x a 1
  x
a 1

sin a dx   or  dx 
0 1 x 0 1 x sin a
a 1 a 1
 x  x 
and  dx  cos a  dx  cos a .
0 1 x 0 1 x sin a
 x a 1
Thus 
0 1 x
dx   cot a

Ex. 11
Apply the calculus of residues to prove that
 sin x
 x  1  x  dx  
0 2

Sol.
eiz
f (z) 
 f (z)dz
C
z  1  z2 
Consider , where taken round a contour C
z R
consisting of the upper half of a large circle and the real axis from –R to
1 ,  2 and  3
R indented at z = -1, 0 and 1. The small circles are denoted by ,
r1 , r2 and r3
their centres are the points z = -1, z = 0, z = 1 and small radii
respectively.

Within the computer thus modified f(z) has no singularities.


Hence by Cauchy residue theorem we have

390
 1 r1   r2 1 r3 R

R
f (x)dx   f (z)dz  
1   1 r1 
f (x)dx   f (z)dz   f (x)dx   f (z)dz   f (x)dx   f (z)dz  0
2 r2 3 1 r3 CR

Now,  CR
f (z)dz  
CR
f (z) dz

eiz
 dz
CR
z  1  z2 

eiz
 dz
CR

z 1 z
2


 1 e  R

which  0 as R  
R  1 R  2

eiz 1
Also, lim  z  1 
z 1 z  1  z2  2
zeiz
lim 1
z 0 z  1  z2 

and lim
 z  1 eiz 
1
z 1 z.  1  z 2
 2

1 1
 lim  f (z)dz  i  0      i
r1 0 1 2 2
lim  f (z)dz  i  0    .1  i
r2 0  2

1 1
lim  f (z)dz  i  0      i
r3 0 3 2 2
R  , r1  0, r2  0, r3  0
Making , we have from (1)
1 1 0 1 1 
 f (x)dx 
2
i  1 f (x)dx  i  0 f (x)dx 
2
i  1 f (x)dx  0
 
i.e. 
f (x)dx  2i  0 or 

f (x)dx  2i
 eix
i.e.  x  1  x 2  dx  2i
Equating imaginary parts we have
sin x  sin x

x  1 x2 
dx  2 or  x  1  x  dx  
0 2

5.14 Integral involving Many Valued Functions.

391

 
x n 1Q(x)dx
Integrals of the type , where n is not an integer and as
n 1
x
such is a many-valued function can also be evaluated by the method of
z R
contour integration by taking the integral round a large circle a small
z r
circle enclosing the branch point z = 0,and a cut along the real axis
joining the ends of the two circles. This is illustrated by means of the
following examples.
Ex.1
Use the method of contour integration to prove that
 x
1/ 6
log x 3
0  1  x  2 dx  2  
3
.

Sol.
z1/ 6 log z
f (z) 
 f (z) dz
C  1 z
2

Consider the integral where , the integration is


z R z r
taken round a large circle , a small circle , and cut along the real
axis joining their arcs.

 1 z
2
0
Pole of f(z) is given by
i.e. z=-1 is a pole of order 2.
(z)
f (z) 
 1 z
2
(z)  z1/ 6 log z
The function where so that
1 1 6  log z
' (z)  z 5/ 6 log z  z1/ 6 . 
6 z 6z5 / 6

392
 (1) 
' 6  log  1

 6  log e  i

6  1 6 e 
5/ 6 i 5/ 6

 6  i  ei / 6  1
  6  i   3 i 
6 12

'  1
 1!
Residue at the double pole z = -1 is


1
12
 6  i   3 i 
Hence by Cauchy’s residue theorem, we have

 f (z)dz  2i 
C
Sum of residuces within the contour C

f (z)dz   f  xe 2 i    f (z) dz   f (x) dz


r R

CR
R
Cr
r

 6  i   3 i 
 2i  (1)
12
Note:
That while integrating from r to R, z=x (because arg z is zero); but
z  xe 2 i 2
when integrating from R to r, (because arg z is ).
z1/ 6 log z
Lt zf (z)  Lt z.
 1 z
z  z  2

Now since
log z  
Lt  from  
1 z
z  7 / 6 2
z

1/ z
Lt  (A)
7
 z 13/ 6  1  z   z 7 / 6 .2  1  z 
z  2

6
1
Lt 2
0
z 
7  1  1
 z 5 / 6 1    2z 5/ 6 1  
6  z  z
 Lt
R   f (z)  i  2  0  .0  0
CR

393
z1/ 6 log z
Lt zf (z)  Lt z
 1 z
z0 z 0 2

Also, since indeterminate.


7
log z  z1/ 6
Lt 6 0
z 0 z 1 z
 Lt
r 0
Cr
 f (z)  i  2  0  .0  0
R  r0
Hence by making and relation (1) becomes
6  i
 f  xe  e  
0 
2 i 2i
dx   f (x)dx  2i 3i
 0 12

 xe 
2i 1/ 6
log  xe 2 i  .e 2 i dx i
x1/ 6 log x
 
0 
  dx    6  i  3 i
 1  xe  2 i 2
1 x
2
 0 6
Or
Or
1/ 6 i / 3
. log x  2i i
x .e
 
1/ 6
  x log x
 dx   dx    6  i  3i
1 x  1 x 
2 2
0 0 6

Equating real and imaginary parts, we have


1 1/ 6
x log x   3x1/ 6

2 2 3
 0
1 x
2
dx   
6

 3
Multiplying second by and then adding it to first we get
 x1/ 6 log x 2 3
 dx   2
 1 x
2
0 3

 x1/ 6 log x 2
 1 x
0 2
dx  2 
3
Or
Ex.2.
1  p  1     
If and , show by contour integration that

394
 x  p dx  sin p
 0 1  2x cos   x 2
 .
sin p sin 


explaining clearly the necessity of conditions on p on .
Sol.
zp
 f (z) dz f (z) 
C 1  2z cos   z 2
Consider the integral , where , taken
z R
round the contour C, consisting of the large circle , the small circle
z r
, and the lines (cross cuts) joining the arcs of the two circles along the
real axis.

The singularities of f(z) are given by


1  2z cos   z 2  0

i.e.  z  ei   z  e  i   0
z  e  i  ei   and z  e  i  ei   .

 
Thus we see that amplitudes of these points are and and
  
since lies between and , therefore the two amplitudes lie between 0
2
and .
z  ei z  e  i
So and are the two simple poles within the contour.
f (z) z   e i
Residue of at is

Lim
 z  e  z   1 e
i p p ip

 z  e   z  e  2i sin 
z  ei i  i

e  ip
 e  pi
2i sin 

z  ei
Also, Residue at is

395
Lim
 z  e    1 e  i p ip

 z  e   z  e  2i sin 
z e i i  i


Sum of these two residues
e  pi eip  e  ip sin p
.  e  pi
sin  2i sin 

Hence by Cauchy’stwo residue theorem we have
R r
 r
f (x)dx   f (z)dz  
CR
R
f (xe 2 i )d(xe 2 i )   f (z)dz
Cr

 2i 
sum residues ........(1)
p
z dz

CR
f (z)dz  
CR z  ei z  e i
Now,
p
z z
   z  e  z  e 
CR
i  i

R  p dz
   R  1  R  1
CR

R  p .2R 2R 1p


 
 R  1  R  1
2 2
0 R 
which as , since 1-
p>0
2r1 p
 f (z)dz   1 r 
2
Cr  r0
Similarly which 0 as since 1-p>0
1  p  1
So the condition is necessary.
R  , r  0
Hence by making , the relation (1) reduces to

 f (x)dx   f  xe e
0
2 i 2 i
dx  2 i 

0
sum of residues

396
 xe2i  .e2idx  2i  eip sin p
p
 x  p dx 0
i.e. 
1  2x cos   x 2  1  2xe 2 i cos    xe 2 i  2

0 sin 
 x  p dx  x  p e 2pi dx e  ip sin p
i.e. 
0 1  2x cos   x 2 0 1  2x cos   x 2
  2 i 
sin 
p  ip

 1  e2pi   0 1  2xxcosdx  x 2  2i  e sinsin p



or

 x  p dx 2i sin p
i.e. 0 1  2x cos   x 2  pi  pi .
e e sin 
 sin p
 .
sin P sin 
This gives value of the given integral.
Note.
If in the above result we write –p for p we have
 x p dx  sin   p 
0

1  2x cos   x 2 sin   p 
.
sin 
 sin p
 .
sin p sin 

which is same as result of the previous integral. Hence


 x  p dx  x p dx
0 1  2x cos   x 2 0 1  2x cos   x 2

 sin p
 .
sin p sin 

5.15 A Quadrant or a sector of a circle as the contour.


Ex.1.
2
eiz
z
By integrating along a suitable paths show that

sin x 2 1
0 x dx  4 

Sol.

397
2
eiz
 f (z)dz f (z) 
C z
Consider the integral , where
taken round a closed
z R
contour C, consisting of the positive quadrant of a large circle , and the
positive real and imaginary axes indented at z=0; let r be the radius of
indentation. Since there are no singularities, within the contour, hence by
Cauchy’s residue theorem we have

 f (z)dz  2i 
C
sum of residues within C

R r
i.e.  f (x)dx   f (z)dz   f (iy)idy   f (z)dz  0
r R
CR Cr
2 2 2 2
 iy
R eix eiz r e eiz
i.e. r x
dx  
z
dz  
R iy
idy  
z
dz  0  (1)
CR Cl
2
eiz
Now, 
CR
z
dz

2
eiz
 
CR
z
dz

2
/ 2 e  R sin 2
 Rd sin ce z  Re i
0 R
1   R 2 sin  d
2 0
 e Putting 2  
/ 2

2
e 2R sin  /  d
0

   R 2 /    / 2
 e
2R 2  0



2R 2 
1  e R
2


which  0 as R  
2
eiz
Also lim zf (z)  lim z 1
z 0 z0 z
1  1
 lim  f (z)dz  i    0  .1   i.
z 0
Cr 2  2

398
R  , and r  0
Hence making , the relation (1) becomes
2 2
 iy
 eix 0e 
0 x
dx  
 y
dy  i.  0
2
2 2
 ix
 eix e 
i.e.  dx   dx i
0 x 0 x 2
2 2
eix  e ix
 
or  dx i
0 x 2
 sin x
2

i.e.  dx  .
0 x 4

399
Ex. 2.
By contour integration prove that
 cos x 
 0
x
dx   
2

Sol.
eiz
 f  z  dz, where f  z  
C z
Consider
The only singularity of f(z) is at z = 0. Take the contour C consisting of
z  R,
a positive quadrant of a large circle its bounding radii (taken a two
axes), indented at z = 0, and let r be the radius of indentation at z = 0.
Thus, there is no singularity of f(z) inside the given contour. Hence by
Cauchy’s theorem we have

R r
 f  x  dx   f  z  dz   f  iy  dy   f  z  dz  0
r
CR
R
Cr
  1

Now,
CR
 f  z  dz   f  z 
CR
dz

eiz dz
 
CR z

/2 e R sin 
 Rd sin ce z  Rd
0
R
R / 2

R 0
e 2R /  d by Jordan 's inequality

R 
 1  e R 
/2
 e 2R /   
2R R 0
2 R

which  0 as R  .

eiz
lim zf  z   lim z 0
z 0 z 0
z
Also,

400
 
 lim  f  z  dz  i   0  .0  0
r 0
Cr 2 

the negative sign is taken because integral along Cr, is in clockwise direction.
Hence making R  , r  0,
the relation (1) reduces to
 0
 f  x  dx   f  iy  idy  0
0 

y
 eix 0 e
i.e.  0
x
dx  

 iy 
idy  0

y
 eix  e
 0
x
dx  
0
 iy 
idy

  
  s tan dard value 
2

2 i  e  t dt 2 i.
2
sin ce e  x dx 
0 2 0 2
=
  e1/ 2i      e1/ 4i 
1/ 2

 cos x 1
0
x
dx   cos 
4
Equating real parts we have
 cos x 
or 
0
x
dx   
2

 sin x 

0
x
dx   
2
Also equating imaginary parts we have
Ex. 3.
By contour integration prove that
 a
0
x a 1 cos xdx  a cos
3
where 0  a  1.

Sol.

 f  z  dz.
C
where f  z   eiz z a 1.
Consider

401
The singularity of f(z) is at z= 0. Take the contour C as the quadrant of
z R
a large circle bounded by the two radii and indented at z= 0; let r be the
radius of indentation where r is small.
Within the given contour f(z) has no singularity, hence by Cauchy’s
residue theorem we have
R r
 f  x  dx   f  z  dz   f  iy  idy   f  z  dz  0
r
CR
R
Cr
  1

Now,  f  z  dz   f  z 
CR CR
dz

1/ 2 

CR
 eiz z a 1 dz  
0
e R sin  .R a 1.Rd

1/ 2 
 Ra  e 2R /  d
0
By Jordan’s inequality
R a
1/ 2  
 e 2R /    1  e  R 
1 a 
2 R 0 2R

which  0 as R  , sin ce 0  a  1.

Also, lim zf  z   lim zeiz z a 1  lim z a eiz  0 sin ce a  0


z 0 z 0 z 0

 
lim  f  z  dz  i   0  .0  0
r 0
Cr 2 

The negative sign is taken because Cr is described in clockwise


direction.
R  , r  0
Hence making the relation (1) reduces to
 0
 f  x  dx   f  iy  idy  0
0 

 
i.e., f  x  dx   f  iy  idy
0 0
 
or  e x dx   e  iy  idy
ix a 1 y a 1
0 0

e  y y a 1dy   ei / 2 
 
  i
a
 
a
e  y y a 1dy
0 0

 eia / 2 a
by an important integral

402
Equating real parts we have
 a
0
x a 1 cos xdx a cos
2
.

Note :
If we equate imaginary parts we may have
 a
0
x a 1 sin xdx a sin
2
.

403
Ex. 4.
3
e Z
Integrating over a contour consisting of
i) the real axis from 0 to R,
z  R, from   0 to   , where    / 4,
ii) the circle
 , from z  R
iii) the line to the origin,
Prove that

e  x cos 2 cos  x 2 sin 2  


2
cos .
0 2
(a)

e  x cos  sin  x 2 sin 2  


2
sin .
0 2
(b)
Sol.

 f  z  dz.
C
f  z   e z
2

Consider the integral where taken round the


given contour (see the fig). Since f(z) is analytic within and on the contour (i.e.
there is no singularity within the contour) hence by Cauchy’s residue theorem
we have

 f  z  dz  0
C

  2
i.e.,        e  z dz  0
 GA AB BO 
R 0
d  rei   0
 rei 
2

  e  z dx   e
2 2
e  x dx   (1)
0 R
CR
Or
 because on OB, z  re  i

e
 z2
Now, dx
CR

404

 e  z dz   e  R e i i Re i d sin ce z  Rei
2 2 2

0
CR

 
  eR
2
cos 2 
.Rd Put 20 

0 2
1 1
 
 R  2 e  R sin d where    2
2

2  2
1
1 
  2 e 2R  /  d
2
by Jordan 's inequality
2 


4R
e 
 2  R 2 /   R 2
e which  0 as R   
R 
Hence making , (1) reduces to
 
 e  x dx   e  r e i ei dr  0
2 2 2

0 0
  r 2  cos 2  isin 2   i 
 e dr   e  x dx
2
e
0 0

  
i.e.  e x dx  e i  e  x dx  e i
2 2 2
cos 2 
.e ix sin 2 
0 0 2
 
sin ce  e  x dx 
2

0 2
Equating real and imaginary parts we have

.cos  x 2 sin 2  dx 


2
e x cos 2 
cos 
0 2

.sin  x 2 sin 2  dx 

0
 x 2 cos 2 
and e sin 
2
Ex.5.
  1 
0
cos x 2dx   sin x 2dx  .
0 2 2
Prove that
 /4 
Therefore are particular cases obtained by taking for , in the last
exercise. However to prove these proceed exactly, like the last exercise.
5.16 Rectangular contour.
The contours we have been using so far were either circles or semi-
circles, with or without indentation, for evaluating the integrals of the type


f (x)dx

405
There is however no special merit in a semi-circle. The integrals of
above type can also be evaluated by using a rectangle with or without
indentation, as the contour of integration. The fallowing examples illustrate the
method.
Ex.1.
2
e x
By integrating round the rectangle whose vertices are 0, R, R+ia, ia
show that
2
 ea
 i  0 e  ix 2
cos 2ax dx  
2
 a
 ii  0 
2 2 2
and e  x sin 2ax dx  e  a e y dy
0

Sol.

 f (z) dz
C f (z)  e  z
2

Consider the integral where taken round closed


contour C which is the perimeter of the given rectangle OABD.

Since f(z) is analytic within and on the contour C(i.e. there is no


singularity within the contour) hence by Cauchy’s residue theorem we have

 f (z)dz  0i.e.  e
 z2
dz  0
C

   z2
         e dz  0
 OA AB BD Do 

Since
 On OA, z  x, dz  dx; on AB, z  R  iy, dz  idy 
 On BD, z  x  ia, dz  dx; On DO, z  iy,dz  idy 
 
Hence the above equation becomes
R a 0 0
e x dx   e  R iy  i dy e x  ia    e iy .idy  0
2 2 2

 
2 dx
 (1)
0 0 R a
+

406
a   R  iy 
2 a   R  iy 
2
Now,  0
e .idy   e
0
idy
a
  e R
2
 y2
dy
0

0
  e R
2
a 2
dy sin ce y  a on AB
a

e

 R 2 a 2  a which  0 as R  .

R 
Hence by making , equation (1) becomes
  a
e  x ia  dx   e  x dx  i  e y .dy
2


2 2

0 0 0

i.e.  e
  x 2 2
    i a e y dy sin ce  e  x dx  
 a  2iax dx
 
2 2

0 2 0 0 2

e   cos 2ax  i sin 2ax dx  


 a
  i  e y dy
 x2 a 2 2

0 2 0
Or
Equating real and imaginary parts, we have
2
 e a

 x2
e cos 2ax dx  
0 2
 a
and  e x sin 2ax  e  a 
2 2 2
e y dy
0 0

Ex.2.
eiz
za
Integrating along the boundary of the square defined by x=0,
x=R, y=0, y=R Prove that
 ax
cos x
  xe
0 x  a dx  0 1  x 2 dx, a  0
 ax
 sin x  e
0 x  a 0 1  x 2 dx.
dx 

Sol.
eiz
 f (z) dz f (z) 
C za
Consider the integral where taken round the
enclosed contour C, consisting of the permeter of the given square.

407
Poles of f(z) are given by z+a=0 i.e. z=-a is the simple pole which too is
outside the given contour
Hence by Cauchy’s theorem, we have

 f (z)dz  0
C

 
i.e.          f (z) dz......(1)
 AB BD DO OA 

Now, on AB, z=R+iy, dz=idy and y goes from 0 to R.


So
eiz
 f (z)dz  
AB AB
f (z) dz  
AB
za
dz

eiz
 
AB
z a
dz

R ei R iy  R e  y dy
 idy  
0 R  iy  a 0 R  iy  a
R e y
 dy Put R  y  a  t
0 R ya
ei  a  R
a
  dt
R a t
ea R a e t
 R dt which  0 as R  
e a t
 R a e
t



sin ce a t
dt is finite 

  f (z) dz  0 when R  
AB

Also, on DO, z=iy, dz=idy; and on OA, z=x, dz=dx.


R 
Hence by making , relation (1) becomes

408
DO
 f (iy) idy   f (x)dx  0
OA
i  iy  ix
e  e
i.e.  idy   dx  0
iy  a 0 xa

y
 eix  e
or 0 xa
dx  i 
0 a  iy
dy

 i
 i  x  .adx Putting y  ax

0 a  1  ix 
  i  x e
 ax

 dx.
0 1 x2
Equating real and imaginary parts we have
 ax
 cos x  xe
0 x  a 0 1  x 2 dx
dx 
 ax
 sin x  e
and 0 1  a dx  0 1  x 2 dx.
Ex.3
By contour integration, prove that
 x sin x dx 
  log  1  a  if 0  a  1
0 1  a  2a cos x
2
a
 1 a
 log if a  1.
a a

Sol.
z
 f (z)dz f (z) 
C a  e iz
Consider where
The singularities of f(z) are given by
a  e iz  0 or e  iz  a  e log a  e log a   2ni

iz  log a  2ni or z  i log a  2n where n  0


or any integer
positive or negative.
Taken the contour C a rectangle whose side are
x  , y  0 and y  R

When a>1 then the only singularity within the contour is a simple pole
at z=I loga, at which the residue of

409
 z 
f (z)   
  d / dz   a  e   z i log a
 iz

i log a log a
 
i elog a a

Hence by Cauchy’s residue theorem we have


 R  0 log a
 f (x)dx   f    iy  idy   f  x  iR  dx   f    iy  idy  2 i   (1)
 0  R a
  x  iR
Now,  f  x  iR  dx  
 
a e  
 i x  iR
dx

 x  iR
 dx

a  e i x iR 

 x  iR
 dx

a  e  ix  R
 xR
 dx which  0 as R  .
 a  e R

 xdx
Also,  
f (x)dx  
 a  e  iz

0 xdx  xdx
 

 a  e ix 0 a  e  ix

 xdx  xdx
 
0 a e ix

0 a  e  ix
 writing  x for x in first int egral 
 x e e  dx    2ix sin xdx
ix  ix

 
0
 a  eix   a  eix  0 1  2a cos x  a 2
R 
Hence making relation (1) reduces to
 x sin x  0 2i log a
2i  dx   f    iy  idy   f    iy idy 
0 1  2a cos x  a 2 0  a
 x sin x     iy   iy  2i log a
i.e.  2i  dx  i    dy 
0 1  2a cos x  a 2 0
a  e
y
a  e 
y
a
 x sin xdx  dy 2i log a
i.e.  2i   2i  
0 1  2a cos x  a 2 0 ae y
a

410
y
x sin xdx  e dy 2i log a
or 2i   2 i  

0 1  2a cos x  a 2 0 ae y
1 a
x sin xdx 2i  2 i log a
log  1  ae  y  

i.e. 2i  
0 1  2a cos x  a 2
a  0 a
 x sin xdx 2i 5i log a
or  2i   log  1  a  
0 1  2a cos x  a 2
a a
 x sin xdx 
or 0 1  2a cos x  a 2  log  1  a   log a 
a
 1 a
 log
a a
And when 0<a<1. then there is no singularity within the contour, hence
in that case we should have
 x sin xdx 2 i
 2i   log  1  a   0
0 1  2a cos x  a 2
a
 x sin xdx 
  log  1  a 
0 1  2a cos x  a 2
a

Ex. 4.
Apply the calculus of residues to prove that
 cosh ax 1
 dx  sec  a / 2  where    a  
0 cosh x 2

Sol.
2e 
a  z
eaz
 f (z) dz
C
f (z)  
cosh z e 2 z  1
Consider where taken round the
x  R
rectangular contour consisting of the real axis, the lines and the line
y=1.

e 2 z  1  0,
The poles of f(z) are given by i.e.

411
e 2 z  1  ei  2n   
i.e.2z  i  2n  1

where n is zero of any integer.


1
z i
2
The only pole which lies within the contour is a simple pole at
The residue of f(z) at (z=1/2 i)
 eaz  eai / 2
  
  d / dz  cosh z  z i / 2 i

Hence by Cauchy’s residue theorem we have


  eia / 2
C f (z)dz  0or    
 
AB BC CD DA 
 f (z) dz  2 i 
i
On AB, z  x, dz  dx. on Bc, z  R  iy, dz  idy
On CD, z  x  i, dz  dx. on DA, z  R  iy, dz  idy
R 1 R 0 eai / 2
  f (x)dx   f (R  iy) idy   f  x  i  dx   f  R  iy idy  2i  ......(1)
R 0 R 1 i
1
Now  f  R  iy  idy
0

1 1 2e a   R iy 


  f  R  iy  i dy   dy
0 0
e 2  R iy  1

1 e a   R iy  1 e a   R
 2 dy  2  dy
0
e 2  R iy   1 0
e 2 R iy   1

e a  R
1 2e a  R
 2  2 R dy  2 R
0 e 1 e 1
2
  a  R
 0 as R   sin ce    a  
e  e  a   R
0 2e a   R
Similarly  f  R  iy  idy 
1 1  e 2 R
2
  a  R  a  R Which  0 as R  
e e
Hence from (1)

412
 
e1/ 2ai
 f (x  i) dx 

 f (x  i) dx  2i 

i

e  
ai x  i 1
  ai
 f  x  i  dx  
  cosh x
dx  2e 2
1
 eax  eiax eax ai
 cosh x dx    cosh x dx  2e 2
sin ce cosh   x  i   cos i  x  i   cos  xi      cos xi   cosh x 
1
eax
 1  e   cosh xdx  2e 2
 ai
ai
i.e.

 eax 1
or  dx   sec  a / 2 
 cosh x
1  2 ai  12 ai 
1

e  e 
2 

 e  ax
 cosh x dx  sec  a / 2 
Putting –a for a in the result we have
Adding the last two results we have
 eax  e  ax
 cosh x dx  2sec  a / 2 
 cosh ax  cosh ax 1
 cosh x dx  sec a / 2 or   sec  a / 2 
0 cosh x 2

413
Ex.5
eaz
 e2iz  1
(a) By integrating round a suitable contour prove that
 sin ax 1 1  1
 0 e 1
2x
dx   cosh  a  
4 2  a

Sol.
eaz
 f (z) dz f (z) 
C e 2iz  1
Consider where
The singularities of f(z) are given by
e 2iz  1  0 i.e.e 2iz  1  e 2ni
i.e.  2iz  2ni
z   n
Or where n=0 or any integer.

 f (z) dz
C
The integral , is taken round the rectangular contour C
x
consisting of x-axis. Y-axis, the line and the line y=R; intented at z=0
z r '
z
and . Let r and be the radii of indentation at z=0 and .
Within the contour so formed f(z) has no singularities.
Hence by Cauchy’s theorem, we have

414
 r ' R 0 r
 f (z)dz   f (x)dx   f (z)dz   f    iy idy   f  x  iR dx   f  iy idy  0.....(1)
r r  R
Cr C'
r

eaz  0 
lim zf (z)  lim z 2iz  from 
z 0 z 0 e 1  0 
eaz  zeaz  1 
lim   
z 0 2ie 2iz
 2
1  1  
 Lt  f (z) dz  i    0    
r 0
Cr 2  2i  4

negative sign is taken because the arc Cr is described clockwise.


eaz  0 
lim  z    2iz  form 
z  e 1  0 
eaz  aeaz  z    ea 
 lim 
z  2ie 2iz 2i
  e  1 a
a
 Lt  f (z)dz  i   / 2  0     e
r 0
C'
r
 2i  4
0 
Also,  f  x  Ri  dx

  f  x  iR  dx
0

 e a  x iR  dx

0
e 2i x iR   1
 e ax  e ax
 dx   2R dx
0
e 2i x iR  1 0 e 1


e a
 1
which  0 as R  
ae 2R
 1

R   r  0, r  0
'

Hence making , in (1), we have

415
1  1  0
   f (x)dx  ea   f    iy  idy   f  iy  dy  0  (1)
4 0 4 0 

  eax  e e
ax ix
But  f (x)dx   2ix dx   ix ix dx
0 0 e 1 0 e e
1  e  cos x  i sin x 
ax
1  1 
  dx  i  e ax cot xdx   e ax dx
2i 0 sin x 2 0 2 0
1  1
 i  eax cos xdx   ea  1
2 0 2a
a   iy 
  e
and  f    iy  idy   idy
0 0
e  
2i   iy 
1
eax .eiay
 i dy
0 e 2y  1

iay
0  e
also  f  iy  idy  i 0 e2y  1dy
Hence from (2) we have
1 1  ea  1 1 a  e
iay
 e
iay
  i  eax cot xdx   e  iea  2y dy  i  2 y dy  0
4 2 0 2a 4 0 e 1 0 e 1

1 a ea  1 1  ax  cos ay  i sin ay


or
4
 e  1 
2a
 i  e cot xdx  i  e a  1 
2 0 0 e2 y 1
dy  0

Equating real parts we have


1 a e a  1  sin ay

4
 e  1 
2a
  ea  1  2 y dy  0
0 e 1
a
 sin ay 1 e 1 1
i.e.  2y dy  . ay 
0 e 1 4 e  1 2a
1 1
a  a
1 e2  e 2 1
  1 1

4 2 a  2 a 2a
e e
1 1 1
  coth a 
4 2 2a
Ex. 6
eiaz
e 2 x  1
Integrate , where a is real, round the rectangle of sides x = 0 , x
= R, y = 0, y = 1 indented at 0 and i, and show that

416
 sin ax 1 a 1
 0 2 z
e 1
dx  coth   
4  2  2a

Sol.
eiaz
f  z 
Consider the integral
 f (z)dz , where taken round the
e 2 z  1
given ABCDEFA indented at z = 0 and z = i; and let r and r ’ be the radii of
indentation.

Since f(z) has no singularities within the contour, hence by Cauchy’s


residue theorem

 AB BC CD EF   


  f z dz  f  z  dz  f  z  dz  0
   Cr, Cr
On AB, z = x, dz = dx and on BC, z = R+iy , dz = idy
On CD, z=x+I, dz=dx: and on EF, z=iy, dz=idy
Hence the above relation can be written as
R 1 r' r
 f (x)dx    R  iy  .idy   f  x  i dx   f  iy  .idy   f  z dz   f  z  dz  0........(1)
r 0 R 1 r
Cr C'
r

iaz
e 0 
Now,lim z  form 
2 z
z 0 e 1  0 
e  zaieiaz 1
iaz
 lim 
z 0 2e2 z 2
1  1 i
lim  f  z  dz  i    0  . 
r 0
Cr 2  2 4

And lim
 z  i  eiaz  0
form

 
z i e2 z  1  0 
e  aieiaz  z  i  e a
iaz

 lim 
z i 2e2 z 2

417
a
1 e ie  a
r 0 
lim f (z)dz  i    0  
2  2 4
'
C'
r

1
also,  f  R  iy  .idy
0

1 1 eia  R iy  idy


  f  R  iy  idy  
0 0
e 2  R iy   1
1 e  ay dy 1 1
    e  ay dy
0
e 2  R iy   1 e 2 R  1 0

1 1  e a
 .  0as R  
e 2 R 1 a
Hence relation (1) becomes
 0 0 i ie  a
0 f (x)dx   f  x  i  dx  1 f  iy  .idy   0
4 4
 eiax  ei  x  i  1 e
 ay
i ie  a
or  2 x dx   2  x i  dx  i  iiy dz   0
0 e 1 0
e 1 0 e 1 4 4
a
 eiax
i.e.  1  e   2 x dx  i
1  ea

1e
 ay
 cos y  i sin y  dy
0 e 1 4 0 2sin y

Equating imaginary parts we have


sin ax 1  e  a 1 1  ay
 1 e  

a
dx    e dy
0 e 2x  1 4 2 0
1  ea 1  ea
 
2a 2a
 sin ax 1 1  ea 1 1 ea / 2  ea / 2 1
i.e. 
0 e 2 x  1
dx    
4 1  e  a 2a 4 ea / 2  e  a / 2 2a
1 a 1
 coth   
4  2  2a

Ex.7
A contour C is formed by indenting at 0 and I the rectangle whose
e z
sinh z
vertices are R, R+I, -R+I, -R. Prove, by integrating round this contour,
that
 sinh x 1 
0 sinh x
dx  tan  
2 2

418
Sol.
Consider the integral
e z
 f (z)dz where f (z) 
C
sinh z

taken round the perimeter of the given rectangle indented at z=0, and z=i.
poles of f(z) are given by
sinh z  0 i.e.by e z  e z  0
or e 2 z  1  e 2ni or z  ni  n  0, 1, 2,.....

But none of these poles lie within the contour; hence by Cauchy’s
 f (z)dz  0
C
residue theorem we have

 

i.e.                  f (z)dz  0. ......(1)

 AB BP Cr QC CD DR Cr' SA 

e z
Now sin ce, Lt zf (z)  Lt z.
z 0 z 0 sinh z
zez 1
Lt 
 z  
z 0 3

z   .....
3!
1
 Lt
'
r 0
 f (z)dz  i    0  .   i
C'
r

Also sin ce Lt  z  i  f (z)


z i

e z  0 
Lt  z  i  . form 
z i sinh z  0 

Lt
 z  i  zez  ez

e i

2ei 1
  e i
z i  cosh z  cos i   e  e 
i i

1  i
 Lt
r 0  f (z)dz  i    0 
Cr 
e  ie i

(the negative sign is taken because Cr is described in clock wise


direction)
On AB, z=R+iy, dz=idy and y goes from 0 to 1.
So

419
1
 f (z)dz   f (z) idy   f  R  iy 
AB AB
0
idy

1 e R iy  idy



0 sinh   R  iy 
1 dy
 e R   R  iy 
0
e  e R iy 
2
1 dy
 2e R 
0
e R iy   e  R iy 
2eR 1 2e R

e R  e R  0
dx 
e R  eR
 0 as R  ,sin ce 0    .

so  f (z)dz  0 when R  .
AB

 f (z)dz  0 when R  
CD
Similarly .
R  , r  0, r  0 '

Hence when , then relation (1) becomes

 f (z)dz  ie  f (z)dz   f (z)dz  i   f (z)dz  0


i

BP QC DR SA

i.e.  f (z)dz   f (z)dz   f (z)dz   f (z)dz  i  1  e  .


i
....(3)
BP QC DR DR

But on BP or QC, z=x+I, dz=dx and on DR or SA, z=x, dz=dx.


Hence relation (2) may be written as

420
f  x  i  dx   f (x)dx   f (x)dx  i  1  e Ai 
0  0 
 f  x  i  dx  
 0  0

 i 1 e 
 
or  f (x)dx   f  x  i  dx i
 

e x e x  i  dx
 i  1  e i 
 
or  dx  2  x i   x  i 
 sinh x 
e e
e x e x i 
 i  1  e i 
 
or  dx  2 x x dx
 sinh x  e e
ex
 1  e i  .  i  1  e i 

or  sinh x
dx

 ex 1  ei 


or  sinh x dx  i
1  e i
 tan  
2
0 ex  e x 
or  sinh xdx  0 sinh x dx  tan  2 
 e x  e x 
or  dx   dx  tan  
0 sinh x 0 sinh x 2
Writing –x for x in first integral
 e x  e x 
or 0 sinh x
dx  tan  
2
 sinh x 
or 2 dx  tan  
0 sinh x 2
 sinh x 1 
or 0 sinh x
dx 
2
tan  
2

5.17 To find the residue by knowing the integral first.

 f (z)dz
C
So far we evalualed integral by knowing first the residue of
the function f(z) and then using Cauchy’s residue theorem. We now illustrate
 f (z)dz C
below that residue of f(z) can be found out by integrating first and
then using Cauchy’s theorem. The following example illustrates it.
Ex: 1.
1
tan n 1  z  at z 
2
If n be an even positive integer, find the residue of
Sol.

421
n 1
sin z 
f (z)  tan n 1
 z    
 cos z 
Let
The poles of f(z) are given by
cos z  0
eiz  e iz  0
or eiz  1  e 2mi i
or 2inz   2m  1 i
1
or z  , taking m  0.
4
y   R;
Take the rectangular contour C whose sides are x=0, x=1 any
then z=1/2 is the only pole which lies inside the contour.
Hence by Cauchy’s residue theorem, we have

 f (z)dz  2i 
C
sum of residuces inside C,
Or
  n 1
          tan z dz
 AB BC CD DA 
 1
 2i  residue at  z   .  (1)
 2

Now, since on BC, z=1+iy and dz=idz

422
R
  tan n 1 z dz   tan n 1   1  iy  .idy
R
BC
R
 tan n 1  iy  .idy
R

And on DA, z  iy, so that dz  idy


R
  tan n 1 z dz   tan n 1  iy  .idy
R
DA
R
   tan n 1  iy  .idy
R

 tan  tan
n 1 n 1
so that z dz  z dz  0
BC DA

Also, since on AB, z=x-iR, dz=dx


and on CD,z=x+iR, dz=dx
1
  tan n 1  z  dz   tan n 1   x  iR  dx
0
AB
n 1
1 ei x iR   e i x iR  
1
   i x iR   dx
0 l
 e  e i x iR  
n 1 n 1
1 1 1  e 2 ix .e 2 R 
0  l  
 1 e
2ix 2 R 
.e 
dx

n 1
1 1 
   dx as R  
0 l
 
n 1
1
 
i
 tan  z  dx
n 1
Similarly
CD
n 1
0 0  1 
  tan n 1
  x  iR  dx     dx as above
1 1
 i 
n 1 n 1
1 1
  1  
n
   as n is even
i i
 1  n 1  1  n 1   1
 Hence from (1),        2i  Re sidue at  z  
 i   i    2

 1
z  
 2
Hence residue at is

423
1  1   1
n 1 n n/2
2 1
      
2i  i   i  
since n is even.
5.18 expansion of a Meromorphic Function.
A function f(z) which is analytic in a region D except at a finite
number of poles is said to be meromorphic in the region D.
Let f(z) be a function whose only singularities, except at infinity, are
z  a1 , z  a 2 , z  a 3 ,....a n
simple poles at where
0  a1  a 2  a 3  ....  a n
and let b1, b2, b3,….bn be respectively the
residuces a these poles. Let there be a sequence of closed contours, either
circles or squares C1, C2, C3…. Such that
(i) The contour Cn encloses the poles a1,a2,…an but no other poles.
(ii) The contour Cn must be such that minimum distance Rn of Cn
from origin tends to infinity as n tends to infinity.
(iii) Length Ln of the contour Cn is O(Rn),
 n 
Ln= Rn when .
f (z)
 0 when n  
Rn
(iv) On Cn,f(z)=0 (Rn), or
F(z) were bounded on the whose system of contours Cn.

 1 1 
f (z)  f (0)   b n   
n 1  z  an an 
If these conditions are satisfied, then
for all z except at poles.
Proof:
f (z)
j      z  d
Cn
Let where z is a point inside Cn.

The poles of the integrand are at the value of for which
f      at   a1 , a 2 ,...a n  by hypothesis 

f  
    z
Hence poles of inside Cn are
  a m  m  1, 2,3,....n    z and at   0.
and at

424
f  
  a m is Lt    a m  .
a m     z
Residue at
bm

am  am  z

f     am
[since bm being residue of at is given by
b m  Lt
a m
    a  f  
m

f   f (z)
Lt    z  
  z     z z
Also, residue at =z is
f    f (0)
  0is Lt 
 0     z z
and residue at
Hence by Cauchy’s residue theorem we have
j  2i
. Sum of residues inside whole system Cn.
  n bm  f  0  f  z  
 2i     
  m 1 a m  a m  z   z z 
 n bm f  0 f  z  
 2i     
 m 1 a m  a m  z  z z 

j  0as n  
If now we can prove that , the theorem is proved.
Now from (1),
  
j       z  d
Cn

Mn
 R R d where max f     M n on C n
 z  Cn

Cn n n

Mn
 Ln
Rn  Rn  z 
Ln
 Mn [sin ce as n   , R n  ]
Rn  Rn  z 

Hence relation (2) becomes

425
 n bm  f  0  f  z 
0  Lt   
n 
 m 1 a m  a m  z   z z

 1 1 
i.e.f  z   f  0    b n   
n 1  2  an an 

Ex.1.

1 1
cot z   2z  2
1 z n 
2 2
z
Prove that
Sol.
cos z
f  z   cot z 
sin z
Consider the function
sin z  0 or z  m  m  1, 2,...
Poles of f(z) are given by are the
simple poles.
Let the contour Cn be the square ABCD with centre at the origin, the
 1
 n    1, i  .
  2
length of each side (2n+1) , so that its vertices are the point
z  m  m  1, 2,...  n 
Hence poles of f(z) inside Kn are
z  n z   n
In this sequence of poles there are two poles and
whose absolute value is greatest compared to those of other poles.
 cos z 
z   n is Lt  
z  n  d / dz   sin z 
 
Residue at the poles
 cos z 
Lt  1
z  n cos z 
 

Now, the following conditions are satisfied:


z  m  m  1, 2, 3,....  n 
(i) Cn encloses the poles no other
pole.
 1
n  
 2
(ii) The minimum distance Rn of Cn from the origin is
n
which tends to infinity at .

426
(iii) The length Ln of Cn (the perimeter of the square ABCD) is
 1
Rn   n    :
8n  4  2 L n  8R n i.e.L n  O  R n 
and and so
(iv) f(z) is bounded on Cn i.e. f(z) is bounded along the permeter of
the square ABCD as shown below:

we have
eiz  e  iz e 2iz  1 1  e 2iz
cot z i  or
eiz  e  iz e 2iz  1 1  e 2iz
e 2i(x iy)  1 1  e 2i( x iy)
 2i(x iy) or
e 1 1  e 2i(x iy)
1  e 2y  2ix 1  e 2y  2ix
 or
1  e 2 y  2ix 1  e 2y 2ix
1  e 2y 1  e 2y
 or
1  e 2y 1  e 2y
to be chosen according as y is positive or negative.
 1  1
z  n   y  n  
 2  2
Now, on AB, and on CD

on AB and CD both we have
 1
2 n   
1 e  2
cot z   1
2 n   
1 e  2
1
which is finite when n is finite and as
n
, showing that cotz is bounded on AB and CD. And on AD and BC, y is
1  e 2y 1  e 2y
 cot z  or
1  e 2 y 1  e 2 y
positive and negative both, on AD and BC which is
1
finite when y is finite chosen according as y is positive or negative and as
y  y n

427
y  y n
or . As a matter of fact which , showing that cot z is
bounded on AD and BC.

cot z is bounded on the contour Cn.
Hence by theorem on expansion of meromorphic function we have

 1 1 
cot z  f (0)   b n   
1  z  an an  b n  1and a m   n
where

 1 1 
  cot z  z 0   b n   
1  z  an an 
 
1 z   1 1   1 1 
 . cos z    1.     1  
 z sin z  z  0 1   z  a n n   z  n  n   
1   1 1 
     
z 1  z  n z  n  
1  2z
  2
z 1 z  n 2 2

Ex.2.
 1  2n  1
n

sec z  4
1  2n  1   4z
2 2

Prove that
Sol.
1
f (z)  sec z 
cos z
Let
Poles of f(z) are given by cos z=0
1
z  m    m  0,  1, 2...
2
Or by all these are simple poles, and
are arranged in the ascending order with regard to their absolute values.
Let Cn be the square ABCD with centre at the origin and length of
2  n  1 
each side , with vertices at the points
z   n  1  1, i  .

1
z  m    m  0,  1, ,...  n 
2
Hence the poles of f(z) inside Cn are

428
 1  1
z  n   z  n  
 2  2
There are two pole and whose absolute
values are greatest (both being the same) in the sequence of the poles.
 1 
  1
n 1
 1   
z   n      d / dz   cos z  z n 1 
 2  4
Residue at is
 1 
z    n    is  1 .
n

 4 
Similarly residue at

Now the following conditions are satisfied


 1 
z    m    , m  0,1, 2,..., n
 2 
(i) Cn enclosed the poles but no
other poles.

(ii) The minimum distance Rn of Cn from the origin is (n+1)
 as n  
which tends to .
8(n  1)
(iii) The length Ln of the perimeter of the square ABCDis
R n   n  1  Ln  0  R n 
and , so .
(iv) f(z) is bounded on Cn (i.e. on the perimeter of the square ) as
shown below:

429
2 2
sec z   i  x  iy 
 ei x iy 
 iz
e e iz
e
2
  y  ix
e  e y ix
2
  y  ix
the terms in the deno min ator may be int erchanged.
e  e y ix
2 2
 y
or y  y
e e y
e e

y   n  1 
to be chosen according as y is positive or negative. On AB, and
y    n  1 
on CD,
2
 sec   n 1 
e  e n 1 
on AB and CD both which tends to 0 as
n
, showing that secz is bounded on AB and CD.
And on AD and BC
2
sec z  y
e  ey
which is finite for finite values of y and tends to zero as y   ; as a matter of
fact ,y  as . n  Thus sec z is bounded on AD and BC.
Therefore, f(z) is bounded on the contour Cn,
Hence by theorem on the expansion of meromorphic functions we
have

n
 1 1 
sec z  f  0    b n   
0  z  an an 
 1
where b n   1
n 1
and a n    n   
 2

430
     

 
n 1  1 1 
 
n  1 1  
 sec 0    1       1    
0   z   n  1    n  1     z   n  1    n  1    
   2  2     4  2   
 
   
 
n 
 1 1 
 
n  1 1 
 1    1       1   
0  z   n  1    n  1    z    1  1  
n   n   
   2
 
 2
    2  2  

 
4  2n  1 4 
 1    1  
n n
  1 
0 
  2n  1   4z 2
2 2
 2n  1  
 1  2n  1  4   1
n n

 1  4  0  2n  1 
0  2n  1   4z
2 2 2

   1 n  2n  1  4  1 1 
 1   4  2
 1   .... 
 0  2n  1   4z    3 5 
2 2

   1 n  2n  1  4  1 1 1
 1   4  2
 . Since 1   .....  x
 0  2n  1   4z   4
2 2
3 5 4

 1  2n  1
n

 4
0  2n  1   4z
2 2 2

Ex. 3.
    
Prove that, if ,
cos z 1 2z   1 cos n
n

  
sin z z  n 1 z 2  n 2

Sol.
cos z
f (z) 
sin z
Here
sin z  0 or z  m
Poles of f(z) are given by
z  m  m  0, 1, 2...
Or

431
Let the contour Cn be the square ABCD with centre at the origin and
 1
2 n  
 2
length of each side equal to .
The poles which lie with the contour Cn are
z  m  m  0, 1, 2......  n 

In this sequence of poles these are two poles z = n and z = -n whose


absolute value is greater than those of the other poles .

Residues at the poles z = n is


 
 cos z 
d 
  sin z  
 dz z  n
 cos z  cos n
  
  cos z z n  cos n
 1
n
cos n cos n
 
  1 
n

Also retidue at the poles z = -n is


 
 cos z 
d 
 1
n
cos n
  sin z  
 dz z  n 
=
Now we observe that
z  m  m  0, 1, 2......  n 
(i) Cn encloses the poles and no other
poles.

432
 1
n  
 2
(ii) The minimum distance of Rn of Cn from the centre is
n
which tends to infinity as
 1  1
8  n   and R n   n  
L n of C n  2  2
(iii) The length is , so that
L n  8R n Ln  0  R n 
; that is .
(iv) f(z) is bounded on the contour Cn as shown below:

cos z eiz  e iz ei x iy  e  i x iy 


 i iz  iz  i i x iy 
sin z e e e  e  i x  iy 
e y ix  ey ix
 i y ix y ix
e e
ey ix  eiy
 The terms is deno min ator can be
e y ix  e  y ix
int erchanged if desired.
y y
e e e  e y
 y
 or
e y  e y e y  e y
interchanging terms in deno min ator.
ey  1  ey  e y  1  e 2 y 
 or
e y  1  e 2 y  e y  1  e 2 y 
taken according as z is positive or negative.
2  y
1 e 1  e 2 y
 or Since      
1  e 2  1  e 2 y
Taken according as y is positive or negative.
 1  1
y  n   y  n  
 2  2
On AB, ; and on CD
 1
1  e 2   n  
cos z  2

sin z  1
1  e2  n  
  2
on AB and CD both which
 1 as n  

433
and on AD and BC we have y positive above the real axis and negative below
it.

on AD and BC we have
cos z 1  e 2 y 1  e2 y
 any one of or
sin z 1  e 2 y 1  e 2 y
according as y is positive
y   or y  
or negative, which tend to 1 as .

By theorem on the expansion of meromorphic functions we have
cos z 
 1 1 
 f  0     
sin z n 1  z  a n an 
 1 z 
 cos z 
 z sin z z  0
   1 n cos n  1 1   1 cos n  1
n
1 
       
n 1 
   z  n n    z  n n  

 1 cos n  1  1 
n

1
   
z n 1  zn zn
1 2z   1 cos n
n

  
z  n 1 z 2  n 2

Ex. 4

2z
tan z   2
n 1  1 2 2
n    z
 2
Prove that
Sol.
sin z
f (z)  tan z 
cos z

z  m   / 2  m  0, 1, 2,....
Poles of f(z) are given by cosz=0
Let the contour Cr be the square ABCD with centre at the origin and
2  n  1
length of its each side , so that its vertices are the points
 n  1  1, i  
.

434
The poles which lie within the contour Cn are
z  m   / 2  m  0, 1, 2,....  n 

 1
z  n  
 2
In this sequence of poles there are poles and
 1
z  n  
 2
whose absolute value is greatest compared to those of the other
poles.
 1
z  n  
 2
Residue at the pole is
 
 sin z   sin z 
d     1
  cos z     sin z  z   n 1/ 2  
 dz z  n 1/ 2  

 1  sin z 
z   n      1
 2   sin z z  n 1/ 2 
Similarly residues at is
Now we observe that,
(i) The contour, Cn enclosed the poles
1
z  m    m  0, 1, 2,....  n 
2
and no other poles.
 n  1 
(ii) The minimum distance Rn of Cn from the origin is
n
which tends to infinity as .

435
(iii) The length Lm, of Cn (that is the perimeter of the square ABCD)
 R n   n  1 
is 8(n+1) , and ; so that ]
Lm=8Rm i.e. Ln=O(Rn)
(iv) f(z) is bounded on the contour Cn as shown below:
1 eiz  e  iz 1 ei x  iy   e  i x iy 
tan z  
i eiz  e  iz i ei x iy   e  i x iy 
e  y ix  e y ix

e  y ix  e y ix

e y ix  e  y ix

e y ix  e  y ix
the terms in the denominator my be interchanged if
desired.
e y  e y e y  e y
 y  y or  y y
e e e e
terms interchanged in denominator.
1  e2 y 1  e 2 y
 or
1  e2 y 1  e2 y
taken according as y is positive or negative, we
y   n  1  y    n  1 
have on AB, and on CD.

On AB and CD both we have
1 e  
 n2 
tan z  which  1as n  
1 e  
 n 1 
.
And for points on AD and BC we have y positive above the real axis
and y negative below it.

on AD and BC we have
1  e 2y
1  e 2y for positive value of y
tan z  
1  e for negative value of y
2y

1  e 2y
y 
Which are finite for finite value of y and tend to 1 as or .
y   as n  
As a metter of fact .
Thus tanz is bounded on the contour Cn.

436
Hence by theorem of the expansion of meromorphic functions we
have

 1 1 
tan z  f (0)   b n   
n 1  z  an an 
where bn and an have usual meaning
    
   1 1 
 
 1 1 
  tan z z  0    1       1   
n 1   z   n  1   n  1     z     1    n  1   
   2  2     2  2  

 
  1 1 
 0   
n 1   1  1 
 n     z  n     z
  2  2 

2z
 2
.
n 1  1 2 2
n     z
 2

437
HARMONIC FUNCTIONS
The real and imaginary parts of an analytic function are conjugate
harmonic function. Therefore, all the theorems on analytic functions are also
theorems on pairs of conjugate harmonic functions. However, harmonic
functions are important in their own right, and their treatment is not always
simplified by the use of complex methods. This is particularly true when the
conjugate harmonic function is not single-valued.
We assemble in this section some facts about harmonic functions that
are infinitely connected with Cauchy’s theorem. The more delicate properties
of harmonic functions are postponed to a later chapter.
Definition and Basic Properties.
u(z) u  x, y 
A real-valued function or , defined and single-valued in a

region or a potential function, if it is continuous together with its partial
derivatives of the first two orders and satisfies Laplace’s equation.
2u 2u
u   0
x 2 y 2
(54)
We shall see later that the regularity conditions can be weekened, but
this is a point of relatively minor importance.
The sum of two harmonic functions and a contact multiple of a
harmonic function are again harmonic; this is due to the linear character of
Laplace’s equation. The simplest harmonic functions are the linear functions
 r,  
ax+by. In polar coordinates equation (54) takes the form
  u   2 u
r r  0
r  r  2
.
This shows that logr is a harmonic function and that any harmonic
function which depends only on r must be of the form alor + b. the argument

is harmonic whenever it can be uniquely defined.

If u is harmonic in , then
u u
f (z)  i
x y
(55)
u u
U ,V  
x y
is analytic, for writing we have

438
U  2 u  2 u V
 2  2 
x x y y
U  2 u V
 
y xy x

That, it should be remembered, is the most natural way of passing from


harmonic to analytic functions.
From (55) we pass to the differential
 u u   u u 
fdz   dx  dy   i   dx  dy 
 x y   y x 
(56)
In this expression the real part is the differential of u,
u u
du  dx  dy
x y

If u has a conjugate harmonic function v, then the imaginary part can be


written as
v v u u
dv  dx  dy   dx  dy
x y y x

In general, however, there is no single-valued conjugate function, and in


these circumstances it is better not to use the notation dv. Instead we write
u u
*du  dx  dy
y x

and call *du the conjugate differential of du. We have by (56)


(57) fdz = du + i*du.
By cauchy’s theorem the integral of fdz vanishes along any cycle which

is homologous to zero in . On the other hand, the integral of the exact
differential du vanishes along all cycles. It follows by (57) that
u u
 *du    y dx  xdy  0
 
(58)
 
For all cycles which are homologous to zero in .
The integral in (58) has an important interpretation which cannot be left
 z  z(t),
unmentioned. If is a regular curve with the equation the direction
  arg z (t)
'

of the tangent is determined by the angle , and we can write


dx  dz cos ,dy  dz sin .

439
dx  dz cos ,dy  dz sin .
the normal which points to the right of the
 cos    sin ,sin   cos 
2
tangent has the direction , and thus .
The expression
u u u
 cos   sin 
n x y

is a directional derivative of u, the right-hand normal derivative with respect to


 *du   u n  dz
the curve . We obtain , and (58) can be written in the form
u
 n dz  0

(59)
u n
This is the classical notation. Its main advantage is that actually

represents a rate of change in the direction perpendicular to . For instance, if
 z r u n
is the circle , described in the positive sense, can be replaced by
u r
the partial derivate . It has the disadvantage that (59) is not expressed as
an ordinary line integral, but as an integral with respect to are length. For this
reason the classical notation is less natural in connection with homology
theory, and we prefer to use the notation *du.
In a simply connected region the integral of *du vanishes over all
cycles, and u has a single-value conjugate function v which is determined up to
an additive constant. In the multiply connected case the conjugate function has
periods
u
i
*du  
 i n
dz

Corresponding to the cycles in a harmology basis.


There is an important generalization of (58) which deals with a pair of
u1 and u 2 
harmonic functions. If are harmonic in , we claim that

u
 1 *du 2  u 2 *du1  0
(60)
 
For every cycle which is homologous to zero in . According to
  R
Theorem 16, Sec 4.6, it is sufficient to prove (60) for , where R is a

440
 R, u1 and u 2
rectangle contained in . In have single-valued conjugate
v1 , v 2
functions and we can write
u1 *du 2  u 2 * du1  u1dv 2  u 2dv1  u1dv 2  v1du 2  d(u 2 v1 )

d(u 2 v1 ) u1dv 2  v1du 2


Here is an exact differential, and is the
imaginary part of
 u1  iv1   du 2  idv2 
F1f 2 dz F1 (z)
The last differential can be written in the form where and
f 2 (z) F1f 2 dz
are analytic on R. the integral of vanishes by Cauchy’s theorem,
and so does therefore the integral of its imaginary part. We conclude that (60)
  R
holds for , and we have proved.

441
Theorem.
u1 and u 2 
If are harmonic in a region , then

u
 1 *du 2  u 2 *du1  0
(60)
 
For every cycle which is homologous to zero in .
u1  1, u 2  u
For the formula reduces to (58). In the classical notation
(60) would be written as
 u 2 u 
  u
 1
n
 u 2 1  dz  0
n 

The Mean-value Property.


u1  log r and u 2
Let us apply theorem 19 with equal to a function u,
z   z  
harmonic in . For we choose the punctured disk 0< , and for
C1  C2 z  ri  
we take the cycle where Ci is a circle described in the
z r *du  r u d
r  
positive sense. On a circle we have and hence (60)
yields
du du
log r1  r1 d   ud  log r2  r2 d   ud
C1 dr C1 C 2 dr C2

In other words, the expression


u

z r
ud  log r  r r d
z r

is constant, and this is true even if u is only known to be harmonic is an


annulus. By (58) we find in the same way that
u

z r
r
r
d

is constant in the case of an annulus and zero if u is harmonic in the whole disk.
Combining these results we obtain:
Theorem:
The arithmetic mean of a harmonic function over concentric circles
z  r
is a linear function of logr.

442
1
2 zr
ud   log r  ,
(61)
0
and if u is a harmonic in a disk and the arithmetic mean is constant.
  u  0
In the latter case , by continuity, and changing to a new origin we find
1 2
u  z0    u  z 0  rei  d
2 0
(62)
It is clear that (62) could also have been derived from the corresponding
formula for analytic functions, Sec. 3.4, (34). It leads directly to the maximum
principle for harmonic function:
Theorem.
A nonconstant harmonic function has either a maximum nor a minimum
in its region of definition. Consequently, the maximum and the minimum on a
closed bounded set E are taken on the boundary of E.
The proof is the same as for the maximum principle of analytic
functions and will not be repeated. It applies also to the minimum for the
reason that –u is harmonic together with u. in the case of analytic functions the
corresponding procedure would have been to apply the maximum principle to
1
f (z) f (z)  0
which is illegitimate unless . Observe that the maximum
principle for analytic functions follows from the maximum principle for
f  z
harmonic functions by applying the latter to log which is harmonic when
f (z)  0
.
Poisson’s Formula.
The maximum principle has the following important consequence: if
u(z) is continuous on a closed bounded set E and harmonic on the interior of E,
then it is uniquely determined by its values on the boundary of E. indeed, if
u1 and u 2 u1  u 2
are two such functions with the same boundary values, then
is harmonic with the boundary values 0. by the maximum and minimum
u1  u 2
principle the difference must then be indentically zero on E.
These arises the problem of finding u when its boundary values are
given. At this point we shall solve the problem only in the simples case,
namely for a closed disk.
Formula (62) determines the value of u at the center of the disk. But this is all
we need, for these exists a linear transformation which carries any point to the

443
z R
center. To be explicit, suppose that u(z) is harmonic in the closed disk .
The linear transformation
R  R  a 
z  S  
R  a

 1 onto z  R with   0
Maps corresponding to z = a. the function
u  S    1
is harmonic in , and by (62) we obtain
1
u a  u  S     d arg 
2 1
]
R(z  a)

R 2  az
From
We compute
d  1 a   z az 
d arg   i  i   2  dz    2  d
  z  a R  az   z  a R  az 

R 2  zz 
On substituting the coefficient of d in the last expression can
be rewritten as
2
z a R2  a
 
za za za
2

Or, equivantely, as
1za za  za
    Re
2 z a za  za

We obtain the two forms


2
1 R2  a 1 za
u  a   u  z  d   Re u  z  d
2 z  R z  a 2
2 z R za
(63)
Of poisson’s formula. In polar coordinates,
1 2 R 2  r2
u  rei   u  Rei  d
2 0 R 2  2rR cos       r 2

In the derivation we have assumed that u(z) is harmonic in the closed


disk. However, the result remains true under the weaker condition that u(z) is

444
harmonic in the open disk and continuous in the closed disk. Indeed, if 0 < r <
1, then u(rz) is harmonic in the closed disk, and we obtain
2
1 R2  a
u  ra   u  rz  d
2 z R z  a 2

Now all we need to do is to let r tend to 1. Because u(z) is uniformly


z R u  rz   u  z  z R
contionous on it is true that uniformly for , and we
conclude that (63) remains valid.
We shall formulate the result as a theorem.
Theorem.
z R z R
Suppose that u(z) is harmonic for , continuous for . Then
2
1 R2  a
u  a  u  z  d
2 z  R z  a 2
(64)
a R
For all .
The theorem leads at once to an explicit expression for the conjugate
function of u. indeed, formula (63) gives
 1 z d 
u(z)  Re   z u    
 2i  R
 
(65)
The bracketed
z R
expression is an analytic function of z for . It follows that u(z) is the real
part of
1 z d
f (z)   u    iC
2i  R
z 
(66)
where C is an arbitrary real constant. This formula is known as Schwarz’s
formula.
As a special case of (64), note that u = 1 yields
2
R2  z

z R za
2
d  2
(67)
a R
For all .

445
Schwarz’s Theorem.
Known Theorem serves to express a given harmonic function through
its values on a circle. But the right-hand side of formula (64) has a meaning as
z R
soon as u is defined on , provided it is sufficiently regular, for instance
piecewise continuous. As in (65) the integral can again be written as the real
part of an analytic function, and consequently it is a harmonic function. The
z R
question is, does it have the boundary value u(z) on ?
There is reason to clarify the notations. Choosing R = 1 we define, for
U    in 0 2
any piecewise continuous function ,
1 2  ei  z
U    d
2 0
PU (z)  Re i
e z

PU (z)
and call this the Poisson integral of U. observe that is not only a
function of z, but also a function of the function U; as such it is called a
functional. The functional is linear in as much as
PU  V  PU  PV
and
PcU  cPU

U 0 PU (z)0
For constant c. moreover, implies ; because of this
property PU is said to be a positive linear functional.
We deduce from (67) that Pc = c. from this property, together with the
linear and positive character of the functional, it follows that any inequality
mPU M.

The question of boundary values is settled by the following


fundamental theorem that was first proved by H.A. Schwarz:
Theorem.
PU (z) z 1
The function is harmonic for , and
lim PU (z)  U  0 
z  e i
(68)
0
provided that U is continuous at .
PU
We have already remarked that is harmonic. To study the boundary
C1 and C 2
behavious, let be complementary arcs of the unit circle, and denote

446
U1 C1 C 2 , by U 2
by the function which coincides with U on and vanishes on
C2 PU  PU1  PU 2
the corresponding function for . Clearly,
PU1 C1
Since can be regarded as a line integral over
it is, by the same
C1
reasoning as before, harmonic everywhere except on the closed are , the
expression
2
e i  z 1  z
Re i 
e  z e i  z 2

z 1 z  e i PU
Vanishes on for . It follows that is zero on the open
i
C2 PU1 (z)  0 as z  e  C 2
are , and since it is continuous .
U  0   0
In proving (68) we may suppose that , for if this is not the
U  U  0   0 C1
case we need only replace U by . Given we can find and
U     i
for e  C 2
C2 ei0 C2 2
such that is an interior point of and .
U2     for  and hence PU2 (z)   for all z  1
2 2
Under the condition .
i0
e 
On the other hand, since U1 is continuous and vanishes at , there exists a
PU1 (z)   for all z  ei0  
2
such that . It follows that
i0
PU (z)  PU1  PU 2  as soon as z  1 ze  
and , which is precisely
what we had to prove.

447
ei0

ei0 *
There is an interesting geometric interpretation of Poisson’s formula,
also due to Schwarz. Given a fixed z inside the unit circle we determine for
ei the po int ei0 ei ei the po int ei*
each which is such that , z and are in a
straight line. It is clear geometrically, or by simple calculation, that
2
1  z  ei  z ei*  z
(69)
e i
 z
e i*
 z
But the ratio is negative, so we must have
1  z    ei  z  ei*  z
2
 
* 
We regard as a function of and differentiate. Since z is constant
we obtain
eid ei*d *

ei  z e  i*  z

And, on taking absolute values,


d * ei*  z
  i*
d e z
(70)
It follows by (69) and (70) that

448
2
1 z d *

ei  z
2
d
and hence
1 2 1 2
  U   *  d
2 0 2 0
PU (z)  U  d  * 

PU (z) U  
In other words, to find , replace each value of by the value
of the point opposite to z, and take the average over the circle.
The Reflection Principle.
An elementary aspect of the symmetry principle, or reflection principle,
has been discussed already in connection with linear transformations (Chap. 3,
Sec 3.3 ). There are many more general variants first formulated by H.A.
Schwarz.
The principle of reflection is based on the observation that is u(z) is a
 
u z
harmonic function, then is likewise harmonic, and if f(z) is an analytic
f (z)
function, then is also analytic. More precisely, if u(z) is harmonic and
u z   f (z)
f(z) analytic in a region then is harmonic and analytic as functions
* 
of z in the region obtained by reflecting in the real axis; that is,
z   * if and only if z  
. The proofs of these statements consist in trivial
verifications.
*   
Consider the case of a symmetric region: . Because is
connected it must intersect the real axis along at least one open interval.

Assume now that f(z) is analytic in and real on at least one interval of the
f (z)  f (z)
real axis. Since is analytic and vanishes on an interval it must be
f (z)  f (z) 
identically zero, and we conclude that in . With the notation
f  u  iv  
u  z   u z , v  z   v z  
we have thus .
This is important, but it is rather weak result, for we are assuming that

f(z) is already known to be analytic in all of . Let us denote the intersection
  
of with the upper half plane by , and the intersection of with the real

   
axis by . Suppose that f(z) is defined on , analytic in ,

449

continuous and real on . Under these conditions we want to show that f(z) is
 
the restriction to of a function which is analytic in all of and satisfies
f (z)  f (z)
the symmetry condition .
In other words, part of our theorem

asserts that f(z) has an analytic continuation to .
Even in this formulation the assumptions are too strong. Indeed, the

main thing is that the imaginary part v(z) vanishes on , and nothing at all
need to be assumed about the real part. In the definitive statement of the
reflection principle the emphasis should therefore be on harmonic functions.
Theorem.
 
Let be the part in the upper half plane of a symmetric region ,
 
and let be the part of the real axis in . Suppose that v(x) is continuous in

   
, harmonic in , and zero on . Then v has a harmonic extension to

  
v z  v  z 
which satisfies the symmetry relation . In the same situation,

if v is the imaginary part of an analytic function f(z) in , then f(z) has an
f (z)  f (z)
analytic extension which satisfies .
For the proof we construct the function V(z) which is equalt to v(z) in

  v z  
, 0 on , and equal to - in the mirror image of . For a point
x0   x0  PV
consider a disk with center contained in , and let denote the
Poisson integral with respect to this disk formed with the boundary values V.
V  PV
The difference is harmonic in the upper half of the disk. It vanishes on
the half circle, by Theorem 23, and also on the diameter, because V tends to
PV
zero by definition and vanishes by obvious symmetry. The maximum and
V  PV
minimum principle implies that in the upper half disk, and the same
proof can be repeated for the lower half. We conclude that V is harmonic in
x0
the whole disk, and in particular at .
For the remaining part of the theorem, let us again consider a disk with

center on . We have already extended v to the whole disk, and v has a
u 0
conjugate harmonic function in the same disk which we may normalize so
u 0  Re f (z)
that in the upper half. Consider

450
 
U0  z   u 0  z   u 0 z

U 0
0
x
On the real diameter it is clear that and also
U 0 u v
 2 0  2 0
y y x

U 0 U 0
i
x y
It follows that the analytic function vanishes on the
U0
real axis, and hence identically. Therefore is a constant, and this constant
u0  z  u0 z  
is evidently zero. We have proved that .
The construction can be repeated for arbitrary disks. It is clear that the
u0 
coincide in overlapping disks. The definition can be extended to all of ,
and the theorem follows.

The theorem has obvious generalizations. The domain can be taken
to be symmetric with respect to a circle C rather than with respect approaches
C'
another circle . Under such conditions f(z) has an analytic continuation
which maps symmetric points with respect to C onto symmetric points with
C'
respect to .

NOTES

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

451
…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

…………………………………………………………………………………..

452

You might also like