Complex Analysis
Complex Analysis
Complex Analysis
i i
i i
This page is intentionally left blank
i i
Complex Analysis
R. Roopkumar
Department of Mathematics
Alagappa University
Karaikudi
Tamilnadu
Chennai • Delhi
i i
i i
i i
This eBook may or may not include all assets that were part of the print
version. The publisher reserves the right to remove any material in this
eBook at any time.
ISBN 978-93-325-3761-3
eISBN 978-93-325-4159-7
Head Office: A-8 (A), 7th Floor, Knowledge Boulevard, Sector 62, Noida
201 309, Uttar Pradesh, India.
Registered Office: Module G4, Ground Floor, Elnet Software City, TS-140,
Block 2 & 9, Rajiv Gandhi Salai, Taramani, Chennai 600 113, Tamil Nadu,
India.
Fax: 080-30461003, Phone: 080-30461060
www.pearson.co.in, Email: [email protected]
i i
i i
i i
Contents
Preface vii
1. Preliminaries 1
1.1 Introduction and Brief Prerequisites 1
1.2 Complex Numbers and Geometrical Representations 9
1.3 Sequences and Series of Complex Numbers 24
1.4 Some Topological Properties of the Complex Plane 36
1.5 Extended Complex Numbers and Stereographic
Projection 52
1.6 Limit and Continuity 55
2. Analytic Functions 71
2.1 Differentiability 71
2.2 Cauchy–Riemann Equations 79
2.3 Power Series and Abel’s Theorems 94
2.4 Exponential and Trigonometric Functions 102
2.5 Hyperbolic Functions 119
i i
i i
i i
vi Contents
Bibliography 455
Index 457
i i
i i
i i
Preface
There are many good text books on complex analysis, some of which are
listed in the bibliography. In my opinion, most of the books are written with
the assumption that the reader can understand the intricacies of the proofs
by filling the gaps in the arguments. However, based on my personal experi-
ence of offering several courses on complex analysis to the present generation
of students, I sensed the necessity to write a book on complex analysis by
explaining each and every argument in any proof in a lucid manner so that the
book would be an ideal self study material for the students. The present book
has been written to address this need. Since many concepts in complex analy-
sis are geometrical in nature, more geometrical arguments are given, without
any compromise in rigor. While the detailed proofs presented in the book
may appear to be self-evident to the experts in complex analysis, beginner
students who try to learn the subject with rigor and without any assumptions
will find such treatment helpful. At the same time, this book may also be used
as a hand book by a young teacher who needs explanations for some tedious
theorems in complex analysis.
This text book is intended for both under graduate and post graduate
courses in complex analysis. The first chapter consists of the basic concepts
of complex numbers, operations on the complex numbers, topological prop-
erties and limiting concepts with more details. All the results given in this
chapter are used at least once in the later part of this book. For an under-
graduate course, the chapters 2,3,4,5,6 may be prescribed by omitting big
theorems. For a postgraduate course, chapters 1 to 6 may be prescribed as
a first course on complex analysis. For those who have the desired level of
exposure to complex analysis in an under graduate course, directly chapters
4,5,6,7,8 may be prescribed in the post graduate course. In the last chapter, the
properties of Jacobian elliptic functions are having very long proofs. Indeed,
these proofs have been presented to motivate the young students and urge
them persevere instead of being bewildered by the subtle nuances and com-
plications involved. However, my personal opinion is that it is not advisable
to ask the proofs of those results in any examination.
i i
i i
i i
viii Preface
The content of the book is subdivided into 8 chapters with two objectives.
The first objective is to club similar topics under a chapter title and the second
one is to maintain the size of the chapters as uniform as possible. I have made
sure that each chapter has a sufficient number of results and problems of
varied lengths, so that it becomes suitable for setting questions of all types.
For doing this, I realized that I either had to compromise on at least one of
the above said objectives or I had to use two results from the later chapters in
earlier chapters. I have chosen the latter option. However, I have confirmed
that there is no begging of question by using a result before it is proved.
There are some concepts such as the Jordan curve Theorem, Hardamard’s
Theorem regarding the estimate of genus of an infinite product, introduction
to univalent and star-like functions that are not discussed here, since these
are to be taken up at an advanced level beyond the scope of this book. I have
taken stringent efforts to make this book error-free. Nonetheless, if any error
is found, please write to me so that they may be eliminated in future editions.
My e-mail address is roopkumar [email protected]
I am grateful to Dr R. Vembu for his constant encouragement, fruitful dis-
cussions and valuable suggestions while preparing the content for this book.
My thanks are also due to the editors at Pearson Education who worked on
this book, for their professionalism and diligence.
R. Roopkumar
i i
i i
i i
Complex Analysis
i i
i i
This page is intentionally left blank
i i
1
Preliminaries
i i
i i
i i
x x0
z0
x0 x
z
z
z
z0
z0
i i
i i
i i
Preliminaries 3
i i
i i
i i
and
n
∩ Aα = ∩ Aα = A1 ∩ A2 ∩ A3 ∩ · · · ∩ An ,
α∈I α=1
7. Let {Xi }ni=1 be a finite collection of sets, then the Cartesian product
n
X1 × X2 × · · · × Xn (or) Xi
i=1
i i
i i
i i
Preliminaries 5
x ≤ α, ∀x ∈ A.
α ≤ x, ∀x ∈ A.
i i
i i
i i
2. Let f : A → B be a function.
(a) Then A and B are called the domain and codomain of f , respec-
tively.
(b) (a, b) ∈ f is equivalently denoted by f (a) = b.
(c) If f : A → A is defined by f (x) = x, ∀x ∈ A, then f is called the
identity function on A.
1 Archimedian property: Given x > 0 and y ∈ R, there exists n ∈ N such that nx > y.
i i
i i
i i
Preliminaries 7
i i
i i
i i
i i
i i
i i
Preliminaries 9
THEOREM 1.2.1 The set C of all complex numbers is a field with respect to
addition and multiplication as defined above.
i i
i i
i i
6. a · (b · c)
= (a1 , a2 ) · [(b1 , b2 ) · (c1 , c2 )]
= (a1 , a2 ) · [(b1 c1 − b2 c2 , b1 c2 + b2 c1 )]
= (a1 (b1 c1 − b2 c2 ) − a2 (b1 c2 + b2 c1 ), a1 (b1 c2 + b2 c1 )
+ a2 (b1 c1 − b2 c2 ))
= (a1 b1 c1 − a1 b2 c2 − a2 b1 c2 − a2 b2 c1 , a1 b1 c2 + a1 b2 c1
+ a2 b1 c1 − a2 b2 c2 )
= ((a1 b1 − a2 b2 )c1 − (a1 b2 + a2 b1 )c2 , (a1 b1 − a2 b2 )c2
+ (a1 b2 + a2 b1 )c1 )
= ((a1 b1 − a2 b2 ), (a1 b2 + a2 b1 )) · (c1 , c2 )
= ((a1 , b1 ) · (a2 , b2 )) · (c1 , c2 )
= (a · b) · c.
i i
i i
i i
Preliminaries 11
9. a · (b + c)
= (a1 , a2 ) · ((b1 , b2 ) + (c1 , c2 ))
= (a1 , a2 ) · (b1 + c1 , b2 + c2 )
= ((a1 (b1 + c1 ) − a2 (b2 + c2 )), a1 (b2 + c2 ) + a2 (b1 + c1 ))
= (a1 b1 + a1 c1 − a2 b2 − a2 c2 , a1 b2 + a1 c2 + a2 b1 + a2 c1 )
= ((a1 b1 − a2 b2 ) + (a1 c1 − a2 c2 ), (a1 b2 + a2 b1 )+(a1 c2 +a2 c1 ))
= (a1 b1 − a2 b2 , a1 b2 + a2 b1 ) + (a1 c1 − a2 c2 , a1 c2 + a2 c1 )
= ((a1 , a2 ) · (b1 , b2 )) + ((a1 , a2 ) · (c1 , c2 ))
= (a · b) + (a · c).
Remark 1.2.3: C is not an ordered field with respect to any order relation.
1. (3 + i5)(4 − i2).
(3 + i)2
2. .
1 − i2
3. sin(1 + i).
i i
i i
i i
Solution:
1. (3 + i5)(4 − i2) = (12 + 10) + i(−6 + 20) = 22 + i14.
(3 + i)2 (9 − 1 + i6)(1 + i2) (8 − 12) + i(16 + 6)
2. = = =
1 − i2 5 5
−4 + i22
.
5
3. sin(1 + i) = sin(1) cos(i) + cos(1) sin(i) = sin(1) cosh(1) +
i cos(1) sinh(1).
Answers:
4 − i33
1. .
17
2. −527 − i336.
3. cos(1) cosh(2) + i sin(1) sinh(2).
4. exp(−1) cos(1) + i exp(−1) sin(1).
5. cosh(1) exp(2) cos(3) + i cosh(1) exp(2) sin(3).
i i
i i
i i
Preliminaries 13
3. z1 + z2 = z1 + z2 .
4. z1 · z2 = z1 · z2 .
z1 z1
5. = , when z2 (0, 0).
z2 z2
6. If z = x + iy then z · z = x2 + y2 .
7. z = z if and only if z is a real number.
2. z = x + iy = x − iy = x + iy = z.
3. (z1 + z2 ) = (x1 + iy1 ) + (x2 + iy2 )
= (x1 + x2 ) + i(y1 + y2 )
= (x1 + x2 ) − i(y1 + y2 )
= (x1 − iy1 ) + (x2 − iy2 )
= z1 + z2 .
i i
i i
i i
x y
= +i 2
x2 + y2 x + y2
x + iy
=
x2 + y2
x + iy
=
(x + iy)(x − iy)
1
=
x − iy
1
= .
z
Then, applying (4), we get
z1 1 1 1 z1
= z1 · = z1 · = z1 · = .
z2 z2 z2 z2 z2
here, we take the positive square root. This unary operator satisfies the
following properties:
i i
i i
i i
Preliminaries 15
i i
i i
i i
i i
i i
i i
Preliminaries 17
n
n
1 n
1
2 2
Proof: Let c = xk yk , x = |xk |2 and y = |yk |2 .
k=1 k=1 k=1
Consider
n 2
0 ≤ (y2 xk − cyk )
k=1
n
= (y2 xk − cyk )(y2 xk − c yk )
k=1
n
= (y4 |xk |2 − y2 cxk yk − y2 cyk xk + ccyk yk )
k=1
= y4 · x2 − y2 · |c|2 − y2 · |c|2 + y2 · |c|2
= y2 (x2 y2 − |c|2 )
If y2 = 0, then equality holds in (1.1); otherwise, we get |c|2 ≤ x2 y2 .
This implies (1.1).
Proof:
n
x + y2 = |xk + yk |2
k=1
n
= (xk + yk )(xk + yk )
k=1
n
= (xk xk + xk yk + yk xk + yk yk )
k=1
n n
= x2 + y2 + xk yk + yk xk
k=1 k=1
n n
(note that xk yk + yk xk is real)
k=1 k=1
i i
i i
i i
2 At present, treat r exp(iθ ) as just a symbol. The reason for using this notation is explained
in Corollary 2.4.25.
i i
i i
i i
Preliminaries 19
To justify this remark, if we use the above formula, then we will get arg (1 + i)
π
= arg (−1 − i) = , but the first point is in the first quadrant and the second
4
point is in the third quadrant, and hence obviously, their arguments should be
different. See the following diagram.
1+i
arg(−1 − i) arg(1 + i)
−1 − i
π −π
• Case (i): If a = 0, b > 0, then θ = ; If a = 0, b < 0, then θ = ;
2 2
If a < 0, b = 0, then θ = π .
b
• Case (ii): If a > 0, b ∈ R, then θ = arctan .
a
b
• Case (iii): If a < 0, b > 0, then θ = π − arctan .
|a|
|b|
• Case (iv): If a < 0, b < 0, then θ = −π + arctan .
|a|
3 For arctan and its properties, we refer the reader to Section 2.4.
i i
i i
i i
(a, b) (a, b)
p q
q
q
q q
−q
−p + q
(a, b) (a, b)
Example 1.2.17 Write the given numbers in the polar form with principal
1 √
arguments: (1) 2 + i3 and (2) (1 − 3).
2
√ √ 4
1. |3 + i4| = 9 + 16 = 25 = 5 and arg 3 + i4 = arctan
3
√ iarctan 43
⇒ 3 + i4 = 25 e .
1 √ 1√ 1 √
2. (1 − 3) =
1 + 3 = 1 and arg (1 − 3) = − arctan
2 2 2
√ π 1 √ −π
3 = − ⇒ (1 − 3) = e 3 .
3 2
i i
i i
i i
Preliminaries 21
The sum of two complex numbers (a, b) and (c, d) can be viewed geomet-
rically as follows. Construct a parallelogram using the line segment joining
(0, 0), (a, b) and the line segment joining (0, 0), (c, d), as the adjacent sides.
Then the fourth vertex of the parallelogram is (a, b)+(c, d). See the following
figure.
(a + c, b + d)
(a, b)
(c, d )
(0, 0)
i i
i i
i i
r2 exp(iq 2)
r1r2 exp(i(q 1 + q 2))
−z
i i
i i
i i
Preliminaries 23
Now, we find the equation of the straight line passing through the given
two points ζ and ξ . Geometrically, the slope of the line segment joining ξ
and ζ is same as the slope of the line segment joining 0 and ζ − ξ .
z+x
−x
Therefore, the straight line passing through ξ and ζ is same as the straight
line passing through ξ with the slope as that of the line segment joining 0 and
ζ − ξ . Note that if t moves from 0 towards ∞, then t(ζ − ξ ) moves from 0
to ∞, and it passes through ζ − ξ when t = 1. If t moves from 0 to −∞,
then t(ζ − ξ ) moves from 0 to ∞ through −(ζ − ξ ) when t = −1. Hence, the
equation of the straight line passing through 0 and ζ − ξ is t(ζ − ξ ), t ∈ R.
If the straight line passing through 0 and ζ − ξ is translated by ξ , then the
resulting straight line passes through ξ and ζ . Hence, its equation becomes
ξ + t(ζ − ξ ), t ∈ R. Therefore, the equation of a straight line passing through
a and having slope as the slope of the line segment joining 0 and b, for some
a, b ∈ C with b 0, is given by
z = a + tb, t ∈ R. (1.2)
i i
i i
i i
The circle with centre a ∈ C and radius r > 0 can be defined as the set of all
points that are at distance r units from a. More explicitly {z : |z − a| = r}.
We shall prove latter (in Corollary 2.4.25) that z − a = r(exp(iθ )) for some
θ ∈ [0, 2π]. Thus, an equation of a circle can be given by z = a + r exp(iθ ),
θ ∈ [0, 2π].
i i
i i
i i
Preliminaries 25
Proof: Let (an ) be a convergent sequence. Then, there exists a ∈ C such that
given > 0, there exists m ∈ N such that
|an − a| < , ∀n ≥ m.
2
For the same m ∈ N, if j, k ≥ m, then
Then, put
Then, K > 0 and for every n ≤ m − 1, obviously we have |an | ≤ K and for
n ≥ m,
|an | ≤ |an − am | + |am | < 1 + |am | ≤ K,
which implies that (an ) is a bounded sequence.
From the above two results, we obtain the following corollary.
i i
i i
i i
1
Given > 0, applying Archimedian property, we choose N > . If n ≥ N ,
then
n 1 1 1
n + 1 − 1 = n + 1 = n + 1 < n < .
n
Thus, → 1 as n → ∞.
n+1
i i
i i
i i
Preliminaries 27
Proof:
|Re an − Re a| < √ , ∀n ≥ m1
2
and
|Im an − Im a| < √ , ∀n ≥ m2 .
2
For n ≥ max{m1 , m2 }, we have
2 2
|an − a| = |Re an − Re a|2 + |Im an − Im a|2 < + = .
2 2
Hence, an → a as n → ∞.
i i
i i
i i
If n ≥ max{m1 , m2 }, then
LEMMA 1.3.10 Let (an ) and (bn ) be real sequences such that an ≤ bn ∀n ∈ N.
If an → a and bn → b as n → ∞, then a ≤ b.
i i
i i
i i
Preliminaries 29
The half of the proof of the above theorem is already proved and the remain-
ing half follows from the fact that C is a complete metric space. (See Theorem
1.4.46.)
LEMMA 1.3.16 For a real sequence (xn ), lim inf xn ≤ lim sup xn .
n→∞ n→∞
Proof: Let > 0 be given. If α = lim inf xn and β = lim sup xn , then there
n→∞ n→∞
exist N1 , N2 ∈ N such that
i i
i i
i i
LEMMA 1.3.18 If (xn ) and (yn ) be sequences of real numbers such that xn ≤
yn , ∀n ∈ N, then lim sup xn ≤ lim sup yn and lim inf xn ≤ lim inf yn .
n→∞ n→∞ n→∞ n→∞
Proof: Let x and y be the limit superior of (xn ) and (yn ), respectively, then by
definition, given > 0, there exists N ∈ N such that yn < y + , ∀n ≥ N .
From the hypothesis, we get
xn ≤ yn < y + , ∀n ≥ N.
n1 ≥ N and xnk → x as k → ∞.
THEOREM 1.3.19 Let (xn ) and (yn ) be sequences of positive real numbers.
If (xn ) converges to a positive real number, then lim sup(xn · yn ) = lim xn ·
n→∞ n→∞
lim sup yn .
n→∞
and
yn < y + , ∀n ≥ m2 .
1+x+y
i i
i i
i i
Preliminaries 31
∀n ≥ m. Hence,
x x
x − xn ≤ |xn − x| < ⇒ xn > , ∀n ≥ m.
2 2
Next, choose p ∈ N such that np ≥ m. Now, for k ≥ max{j, p}, we get
x 2K
xnk · ynk > · = K.
2 x
Hence, xnk · ynk → ∞ as k → ∞. As we have
xn · yn < ∞ + , ∀ > 0, ∀n ∈ N.
we get lim sup(xn · yn ) = ∞ = x · y.
n→∞
√ xn+1
Proof: Let α = lim sup n xn and β = lim sup . If β = +∞, then
n→∞ n→∞ xn
obviously we have α ≤ β . Therefore, we assume that β < +∞. Then by
definition of limit superior, for a given > 0, there exists m ∈ N such that
xn+1
< β + , ∀n ≥ m. For n ≥ m, we have
xn
xn xn−1 xm+1
xn = × × ··· × × xm < xm × (β + )n−m .
xn−1 xn−2 xm
i i
i i
i i
√ xn+1
lim sup n
xn = α ≤ β = lim sup .
n→∞ n→∞ xn
xn+1 √
lim inf ≤ lim inf n xn .
n→∞ xn n→∞
i i
i i
i i
Preliminaries 33
∞
It is customary to define series of complex numbers by a formal sum an .
n=1
Before talking about its convergence, the sum has no meaning other than in
∞
the expression an , first term is a1 , second term is a2 , third term is a3 , and
n=1
so on. In the sequence (an ) of complex numbers also, we say that first term
is a1 , second term is a2 , third term is a3 , and so on. So what is the difference
∞
between (an ) and an ? To find the difference, we can redefine the series of
n=1
complex numbers as follows.
∞
Definition 1.3.24 A series an of complex numbers is said to be conver-
n=0
m
gent if sm = ak , ∀m ∈ N and sm → s as m → ∞ for some complex
k=0
∞
number s. In this case, we call ‘s’ the sum of the series an and is also
n=0
denoted by the series itself.
∞
Definition 1.3.25 A series an of complex numbers is said to be absolutely
n=0
∞
convergent if the series |an | converges.
n=0
∞ ∞
RESULT 1.3.26 A series an converges iff lim an = 0.
n=0 m→∞ n=m
i i
i i
i i
∞
Proof: an converges to s iff sm → s as m → ∞, where sm = a0 + a1 +
n=0
∞
a2 + · · · + am , ∀m ∈ N and s = an iff sm−1 − s → 0 as m → ∞ iff
n=0
∞
lim an = 0.
m→∞ n=m
The following result gives a necessary condition for a convergent series
of complex numbers.
∞
RESULT 1.3.27 If an converges, then lim an = 0.
n=0 n→∞
∞
Proof: Let an converge. If
n=0
sm = a0 + a1 + a2 + · · · + am , ∀m ∈ N,
then (sm ) is a convergent sequence and by Result 1.3.2, (sm ) is a Cauchy
sequence. Let > 0 be given. Then, there exists m ∈ N such that
|sj − sk | < , ∀j, k ≥ m.
If n ≥ m + 1, then n, n − 1 ≥ m, and hence, |an | = |sn − sn−1 | < . Hence,
(an ) converges to 0.
∞
RESULT 1.3.29 If an converges to S absolutely, then every rearrangement
n=1
∞
of an converges to the same S .
n=1
∞
Proof: Let af (n) be a rearrangement of an , where f : N → N is a
n=1 n=1
m m
bijection. Let Sm = |an | and Tm = af (n) , ∀m ∈ N. Since (Sm ) con-
n=1 n=1
verges, by Cauchy criterion (Theorem 1.3.11), we have (Sm ) is a Cauchy
i i
i i
i i
Preliminaries 35
sequence and hence for a given > 0, there exists N0 ∈ N such that
q
|Sp − Sq | = |an | < ∀q ≥ p ≥ N0 .
n=p
2
∞
Since an convergs to s, there exists N ∈ N such that N ≥ N0 and
n=1
p
an − S ≤ ∀p ≥ N. (1.3)
2
n=1
Since f (N) = N we can choose M ∈ N such that {1, 2, . . . , N} ⊆ {f (n) : 1 ≤
n ≤ M}. Now for all p ∈ N with p ≥ M , we have
p q
q
Tp − S = af (n) −
an +
an − S (where q ≥ p.)
n=1 n=1 n=1
< + = ,
2 2
∞
because by the choice of M , we have N ≤ K ≤ J . Thus af (n) converges
n=1
to S .
i i
i i
i i
∞
Example 1.3.31 If z ∈ C with |z| < 1, then the geometric series zn
n=0
converges.
sn (1 − z) = (1 + z + z2 + · · · + zn )(1 − z) = 1 − zn+1 , ∀n ∈ N.
1 − zn+1 1
sn = → as n → ∞
1−z 1−z
∞ 1
and hence, zn = .
n=0 1−z
i i
i i
i i
Preliminaries 37
1. |z − w| ≥ 0, ∀z, w ∈ C,
2. |z − w| = |w − z|, ∀z, w ∈ C,
3. |z − w| = 0 iff z = w,
4. |z − w| ≤ |z − ζ | + |ζ − w|, ∀z, ζ , w ∈ C.
i i
i i
i i
THEOREM 1.4.6 Let Aα , Aj ⊆ C for all α ∈ and for all j ∈ {1, 2, . . . , n},
where is an arbitrary index set.
(i) If Aα is open ∀α ∈ , then Aα is open.
α∈
n
(ii) If Aj is open ∀j ∈ {1, 2, . . . , n}, then Aj is open.
j=1
Proof: To prove Aα is open, let x ∈ Aα be arbitrary. Then x ∈ Aβ
α∈ α∈
some β ∈
for . As Aβ is open, there exists r > 0 such that B(x, r) ⊆ Aβ ⊆
Aα . Thus, Aα is open.
α∈ α∈
n
Next, if x ∈ Aj is arbitrary, then x ∈ Aj , ∀j ∈ {1, 2, . . . , n}. Since each
j=1
Aj is open, there exists rj > 0 such that B(x, rj ) ⊆ Aj , ∀j ∈ {1, 2, . . . , n}. If
s = min{rj : 1 ≤ j ≤ n}, then s > 0, and hence, B(x, s) ⊆ B(x, rj ) ⊆ Aj ,
n
n
∀j ∈ {1, 2, . . . , n} ⇒ B(x, s) ⊆ Aj . Thus, Aj is open.
j=1 j=1
THEOREM 1.4.7 Let Bα , Bj ⊆ C for all α ∈ and for all j ∈ {1, 2, . . . , n},
where is an arbitrary index set.
(i) If Bα is closed ∀α ∈ , then Bα is closed.
α∈
n
(ii) If Bj is closed ∀j ∈ {1, 2, . . . , n}, then Bj is closed.
j=1
Proof: Proof of this theorem follows from Theorem 1.4.5, De-Morgan’s laws,
and Theorem 1.4.6.
i i
i i
i i
Preliminaries 39
1. A ⊆ B ⇒ Cl A ⊆ Cl B.
2. A is closed iff A = Cl A.
4. Cl Cl A = Cl A.
Proof:
x ∈ Cl A ⇒ B(x, r) ∩ A ∅, ∀r > 0
⇒ B(x, r) ∩ B ∅, ∀r > 0
⇒ x ∈ Cl B.
2. A = Cl A ⇔ A = A ∪ A ⇔ A ⊆ A ⇔ A is closed.
4. Using the property (3), Cl A is a closed set, and by using the property
(2), we get Cl Cl A = Cl A.
i i
i i
i i
6. Since A∩B ⊆ A and A∩B ⊆ B, using the property (1), we have Cl (A∩
B) ⊆ Cl A and Cl (A ∩ B) ⊆ Cl B. Thus, Cl (A ∩ B) ⊆ (Cl A) ∩ (Cl B).
r ≤ y − t y − a − a
2 |y − a|
t
= |y − a| 1 −
2|y − a|
t
= |y − a| 1 −
2|y − a|
< |y − a|.
Exercise 1.4.13 Prove that x ∈ Cl A iff there exists a sequence (xn ) in A such
that xn → x as n → ∞.
i i
i i
i i
Preliminaries 41
i i
i i
i i
S ⊆ A ∪ B, A ∩ S ∅ and B ∩ S ∅.
As
Definition 1.4.24 (Line segment) Let a, b be two points in C. We call the set
La,b = {(1 − t)a + tb : t ∈ [0, 1]} the line segment joining a and b. The point
a is called the initial point of the line segment and b is called the end point of
the line segment.
i i
i i
i i
Preliminaries 43
A1 = {t ∈ [0, 1] : (1 − t)c + td ∈ A ∩ E}
A2 = {t ∈ [0, 1] : (1 − t)c + td ∈ B ∩ E}
As the line segment l ⊆ E ⊆ A ∪ B, we get [0, 1] = A1 ∪ A2 by an easy
verification. We note that
c ∈ A ∩ E ⇒ 0 ∈ A1 ⇒ A1 ∅
and
d ∈ B ∩ E ⇒ 1 ∈ A2 ⇒ A2 ∅
Claim. (Cl A1 ) ∩ A2 = A1 ∩ (Cl A2 ) = ∅.
Suppose t ∈ (Cl A1 ) ∩ A2 . Then, there exists a sequence {tn } from A1
such that tn → t as n → ∞. Therefore, (1 − tn )c + tn d ∈ A, ∀n ∈ N,
and so,
(1 − t)c + td = lim (1 − tn )c + tn d ∈ Cl A
n→∞
i i
i i
i i
Proof: Let x, y ∈ B(a, r). We shall show that Lx,y = {(1 − t)x + ty : t ∈
[0, 1]} ⊆ B(a, r).
Now, if t ∈ [0, 1], then
and hence, we get (1 − t)x + ty ∈ B(a, r). Hence, the corollary follows from
the above theorem.
Claim. (Cl A) ∩ B = ∅.
If x ∈ (Cl A) ∩ B, then there exists r > 0 such that B(x, r) ⊆ E (as x ∈ E
and E is open). Then, there exists y ∈ A ∩ B(x, r), using x ∈ Cl A. If
x = (x1 , x2 ) and y = (y1 , y2 ), then let z = (x1 , y2 ). From
i i
i i
i i
Preliminaries 45
(x1, y2)
( y1, y2)
(x1, x2)
Claim. A ∩ (Cl B) = ∅.
If a ∈ A ∩ (Cl B), then choose s > 0 such that B(a, s) ⊆ E. As a ∈
Cl B there exists b ∈ B ∩ B(a, s). By a similar argument used in the
justification of previous claim, we can find a polygon Pa,b ⊆ E. This
implies that b ∈ A, which is a contradiction.
i i
i i
i i
Proof: For each, x ∈ S , let CS [x] be the union of all connected subsets of S
containing x. Clearly, this union is non-empty as {x} is a connected subset of
S containing x. We claim that CS [x] is connected. Suppose, CS [x] ⊆ A ∪ B,
with CS [x] ∩ A ∅, CS [x] ∩ B ∅ and A ∩ Cl B = ∅ and B ∩ Cl A = ∅.
Let a ∈ A and b ∈ B. Then, there exist connected subsets Ca and Cb of S
such that a, x ∈ Ca ⊆ S and b, x ∈ Cb ⊆ S . Therefore, Ca ⊆ A ∪ B and
Cb ⊆ A ∪ B. Since Ca is a connected set either Ca ⊆ A or Ca ⊆ B, otherwise
Ca ∩ A ∅ and Ca ∩ B ∅, and hence, Ca becomes a disconnected set,
which is a contradiction. As a ∈ A ∩ Ca , we conclude that Ca ⊆ A. Similarly,
we conclude that Cb ⊆ B. Therefore, x ∈ Ca ∩ Cb ⊆ A ∩ B ⊆ A ∩ Cl B = ∅,
which is a contradiction. Hence, CS [x] is a connected set. Next we observe
that CS [x] cannot be properly contained in a connected subset of S because
if K is a connected subset of S containing CS [x], then x ∈ K , and hence by
definition of CS [x], we have K ⊆ CS [x] which implies CS [x] = K . Thus,
CS [x] is a component of S .
As for every x ∈ S , there exists a component CS [x] of S containing x;
to conclude this theorem, we show that any two component of S are either
identical or disjoint. If CS [x] and CS [y] are two components of S such that
CS [x] ∩ CS [y] ∅, then let z ∈ CS [x] ∩ CS [y]. If CS [z] is the component of
S containing z, then CS [x] ⊆ CS [z], since CS [x] is a connected subset of S
containing z. Now, x ∈ CS [x] ⊆ CS [z] imply that CS [z] is a connected subset
of S containing x, and hence, CS [z] ⊆ CS [x] ⇒ CS [x] = CS [z]. Similarly, we
show that CS [y] = CS [z]. Thus, CS [x] = CS [z] = CS [y].
i i
i i
i i
Preliminaries 47
|x − y|
Proof: Let x K . Then, for every y ∈ K , we have y x. If ry = ,
2
then ry > 0 and B(x, ry ) ∩ B(y, ry ) = ∅, ∀y ∈ K . As {B(y, ry ) : y ∈ K}
is a collection of open sets satisfying K ⊂ ∪ B(y, ry ). Hence, there exist
y∈K
m
y1 , y2 , . . . , ym ∈ K such that K ⊂ ∪ B(yk , ryk ). If
k=1
r = min{ryk : 1 ≤ k ≤ m}
then
m m
B(x, r) ∩ K ⊆ B(x, r) ∩ ∪ B(yk , ryk ) ⊆ ∪ B(x, ryk ) ∩ B(yk , ryk ) = ∅
k=1 k=1
i i
i i
i i
THEOREM 1.4.39 Every infinite subset of a compact set K has a limit point
in K .
K ⊆ C = F ∪ F c ⊆ ∪ Eα ∪ F c .
α∈I
rwj
|z − wj | ≤ |z − w| + |w − wj | < μ + ≤ rwj
2
i i
i i
i i
Preliminaries 49
which implies that z ∈ B wj , rwj . This is a contradiction to B wj , rwj ∩ A =
∅. Hence, the theorem follows.
∞ ∞
Proof: First, we show that ∩ Kn has at most one point. If x, y ∈ ∩ Kn , then
n=1 n=1
x, y ∈ Kn , ∀n ∈ N. Then, 0 ≤ |x − y| ≤ D(Kn ) → 0 as n → ∞, and hence,
x = y. c
∞ ∞ ∞
Suppose that ∩ Kn = ∅. Then K1 ∩ ∩ Kn = ∅ ⇒ K1 ⊆ ∩ Kn =
n=1 n=2 n=2
∞
∪ Knc .
As each Kn is a closed set (Theorem 1.4.38), is open, ∀n ≥ 2.
Knc
n=2 c
m m
Hence, by the compactness of K1 , we have K1 ⊆ ∪ Knj = ∩ Knj for
c
j=1
j=1
m
some 2 ≤ n1 < n2 < · · · < nm . Thus, Knm = K1 ∩ ∩ Knj = ∅, which is
j=1
a contradiction. Hence, the theorem follows.
1. In+1 ⊆ In , ∀n ∈ N.
3. D(In+1 ) = 12 D(In ) = 1
2n D(I), ∀n ∈ N.
i i
i i
i i
Hint: First prove that [a, b] × [c, d] is compact as in the proof of Heine–
Borel theorem. Next find positive real numbers M1 and M2 such that K ⊂
[−M1 , M1 ] × [−M2 , M2 ] and complete the remaining proof as in the proof of
previous corollary.
Therefore, (xn ) and (yn ) are Cauchy sequences of real numbers. Then by
Result 1.3.3, (xn ) and (yn ) are bounded sequences. Therefore, we can find
M > 0 such that xn ∈ [−M, M], ∀n ∈ N. If (xn ) has a subsequence as a
constant sequence, then obviously the subsequence is convergent (Example
1.3.5), and by Lemma 1.3.8, (xn ) itself is convergent. If (xn ) has no such
subsequence, then {xn : n ∈ N} is an infinite set of [−M, M], which is a
i i
i i
i i
Preliminaries 51
Proof: We know that the set of all rational numbers is countable and dense
in R. Therefore, the collection
B = B(wj , q) : wj ∈ Q × Q, q ∈ Q ∩ (0, ∞)
i i
i i
i i
⇒ ζ ∈ B (z, r) .
Therefore, z ∈ B w, 2r ⊆ B (z, r) ⊆ G. We note that B w, 2r ∈ B . Hence,
the theorem follows.
1. a ± ∞ = ∞ ± a = ∞, for all a ∈ C,
3. ∞ = ∞,
a
4. = ∞, for all a ∈ C∞ \ {0},
0
a
5. = 0, ∀a ∈ C.
∞
∞
∞ − ∞, and 0 · ∞ are not defined.
∞
2. A is an open subset of C.
Proof: Given a point (x1 , x2 , x3 ) ∈ U other than the north pole (0, 0, 1), we
define the image of φ((x1 , x2 , x3 ))) by the intersection of the line passing
through (x1 , x2 , x3 ) and (0, 0, 1) with the complex plane C.
i i
i i
i i
Preliminaries 53
To find φ((x1 , x2 , x3 ))) explicitly, first we consider the equation of the line
passing through (x1 , x2 , x3 ) and (0, 0, 1), which is given by
(1 − t)(0, 0, 1) + t(x1 , x2 , x3 ) = (tx1 , tx2 , (1 − t) + tx3 ), t ∈ R.
This line meets the x1 x2 -plane if 1 − t + tx3 = 0 ⇒ t = 1−x1
3
. Therefore,
⎧
⎨ x1 x2
, ,0 if x3 1
φ((x1 , x2 , x3 )) = 1 − x3 1 − x3 .
⎩
∞ if x3 = 1
(0, 0, 1)
(0, 0, 0)
z
i i
i i
i i
RESULT 1.5.3 The image of a circle on the sphere U under the map φ is
either a circle or a straight line in C∞ .
Definition 1.5.4 Given any two points z, w ∈ C∞ , define the distance D(z, w)
by the actual distance between φ −1 (z) and φ −1 (w) in R3 .
Proof:
Case 1: z, w ∈ C.
Let φ −1 (z) = (x1 , x2 , x3 ) and φ −1 (w) = (y1 , y2 , y3 ). Then,
i i
i i
i i
Preliminaries 55
i i
i i
i i
1. lim Re (f (z)) = Re l,
z→z0
2. lim Im (f (z)) = Im l,
z→z0
3. lim f (z) = l,
z→z0
5. lim (f + c)(z) = l + c,
z→z0
Proof: Let > 0. Then, there exists δ > 0 such that |f (z) − l| < whenever
0 < |z − z0 | < δ . We choose the same δ for proving (1), (2), (3), (4), (5). If
0 < |z − z0 | < δ , then
6. For proving the last assertion, we first choose δ > 0 such that
1. lim (f + g)(z) = l1 + l2 ,
z→z0
2. lim (fg)(z) = l1 l2 ,
z→z0
i i
i i
i i
Preliminaries 57
f l1
3. lim (z) = , provided l2 0.
z→z0 g l2
2. Consider
Hence, |f (z)| < |l1 | + 1, whenever 0 < |z − z0 | < δ1 . For the same ,
find δ2 > 0 such that
i i
i i
i i
1 1
3. First we prove that lim = if l2 0. Given > 0, choose δ > 0
z→z0 g(z) l2
such that
|l2 | |l2 |2
0 < |z − z0 | < δ ⇒ |g(z) − l2 | < min , .
2 2
Thus, if 0 < |z − z0 | < δ , then
|l2 | |l2 | 1 2
| |l2 | − |g(z)| | ≤ |g(z) − l2 | < ⇒ |g(z)| > ⇒ < .
2 2 |g(z)| |l2 |
We also have
1 1 |g(z) − l2 | |l2 |2 2
g(z) − l = |g(z)l | < 2 · |l |2 = .
2 2 2
Hence, our claim follows. Next by using (2) of this theorem, we get
f (z) 1 1 l1
lim = lim f (z) = lim f (z) lim = .
z→z0 g(z) z→z0 g(z) z→z0 z→z0 g(z) l2
2. lim (f + g)(z) = ∞,
z→a
3. lim (f · g)(z) = ∞ if A 0,
z→a
f (z)
4. lim = ∞,
z→a g(z)
i i
i i
i i
Preliminaries 59
g(z)
5. lim = 0.
z→a f (z)
M
If 0 < |z − a| < δ , then |cf (z)| = |c||f (z)| > |c| = M.
|c|
2. We choose δ1 > 0 and δ2 > 0 such that
and
and
|A| |A|
0 < |z − a| < δ2 ⇒ |g(z) − A| < ⇒ |g(z)| ≥ .
2 2
If 0 < |z − a| < min{δ1 , δ2 }, then
2M |A|
|f (z)g(z)| > = M.
|A| 2
i i
i i
i i
f (z)
Hence, lim = ∞.
z→a g(z)
5. Given > 0 we choose δ1 > 0 and δ2 > 0 such that
1 + |A|
0 < |z − a| < δ1 ⇒ |f (z)| >
and
0 < |z − a| < δ2 ⇒ |g(z) − A| < 1 ⇒ |g(z)| ≤ 1 + |A|.
If 0 < |z − a| < min{δ1 , δ2 }, then
g(z)
f (z) < 1 + |A| (1 + |A|) = .
g(z)
Hence, lim = 0.
z→a f (z)
i i
i i
i i
Preliminaries 61
Proof: First we choose r > 0 such that B((a, b), r) ⊆ . For the given > 0,
there exists 0 < δ < r such that
and hence, |f (x, y) − f (x, b)| < . In other words, we have proved that
lim f (x, y) = f (x, b) for every x ∈ R with |x − a| < δ . Using the definition of
y→b
lim f (x, b) = A , we get
(x,y)→(a,b)
Thus, lim lim f (x, y) = lim f (x, b) = A. Similarly, we can prove that
x→a y→b x→a
lim lim f (x, y) = A.
y→b x→a
i i
i i
i i
2. lim f (y) = l exists for some l ∈ R if given > 0, then there exists
y→x+
δ > 0 such that (x, x + δ) ⊂ I and x < y < x + δ ⇒ |f (y) − l| < .
3. lim f (y) = +∞ exists if given M > 0, then there exists δ > 0 such
y→x−
that (x − δ, x) ⊂ I and x − δ < y < x ⇒ f (y) > M .
4. lim f (y) = +∞ exists if given K > 0, then there exists δ > 0 such
y→x+
that (x, x + δ) ⊂ I and x < y < x + δ ⇒ f (y) > K .
Similarly, we can define lim f (y) = −∞ and lim f (y) = −∞.
y→x− y→x+
lim (f + g)(z) = c ∞.
z→0
f
THEOREM 1.6.14 If f and g are continuous at z0 ∈ , then f + g, fg, and
g
are continuous at z0 .
i i
i i
i i
Preliminaries 63
Proof: Let f be continuous at z0 . For a given > 0, there exists δ > 0 such
that |f (z) − f (z0 )| < , whenever |z − z0 | < δ . For the same δ ,
i i
i i
i i
f (). Using the continuity of g at f (z0 ) for a given > 0, choose r > 0
such that
|g(w) − g(f (z0 ))| < if |w − f (z0 )| < r.
Next, using the continuity of f at z0 , for this r > 0, there exists δ > 0 such
that
|f (z) − f (z0 )| < r if |z − z0 | < δ.
Therefore, combining these two statements, we get
if |z − z0 | < δ, then |h(z) − h(z0 )| = |g(f (z)) − g(f (z0 ))| < .
Hence, h is continuous at z0 .
Proof: Assume that f is continuous at x. For a given > 0, we can find δ > 0
such that
|y − x| < δ ⇒ |f (y) − f (x)| < .
If xn → x0 as n → ∞, for the δ > 0, there exists N ∈ N such that
|xn − x| < δ, ∀n ≥ N.
Therefore, if n ≥ N , then
|xn − x| < δ ⇒ |f (xn ) − f (x)| < .
Conversely, assume that f is not continuous at x. Then, there exists >
0 such that for every n ∈ N, there exists xn ∈ A with |xn − x| < 1n and
|f (x) − f (xn )| ≥ . Hence, there exists a sequence (xn ) such that xn → x as
n → ∞ but f (xn ) → f (x) as n → ∞. Hence, the result follows.
whenever
n
(t1 , t2 , . . . , tn ) ∈ Rn and |tj − xj |2 < δ.
j=1
i i
i i
i i
Preliminaries 65
φ −1 (z) → (0, 0, 1) as z → ∞.
x1 x2
Proof: If φ1 (x1 , x2 , x3 ) = and φ2 (x1 , x2 , x3 ) = ∀(x1 , x2 , x3 ) ∈
1 − x3 1 − x3
U \ {(0, 0, 1)}, then φ = (φ1 , φ2 ). (Theorem 1.5.2.)
x1 (ν) x1
→ as ν → ∞.
1 − x3 (ν) 1 − x3
and
x2 (ν) x2
→ as ν → ∞.
1 − x3 (ν) 1 − x3
Therefore, φ is continuous on U \ {(0, 0, 1)}.
(b) If xj (ν) → 0 for j = 1, 2 and x3 (ν) → 1 as ν → ∞, then
i i
i i
i i
Proof: Let > 0 be given. For every w ∈ K , there exists δw > 0 such that
i i
i i
i i
Preliminaries 67
δw
If δ = min 2 : j
j = 1, 2, . . . , n , then δ > 0. If x, y ∈ K with |x − y| < δ ,
δwj
then x ∈ B wj , 2 for some 1 ≤ j ≤ n. We note that y ∈ B wj , δwj for the
same j, because
i i
i i
i i
Now we claim that f −1 (A) and f −1 (B) are separated sets in C. Suppose
f −1 (A) ∩ (Clf −1 (B)) ∅, then there exists a point x ∈ f −1 (A) and a sequence
(xn ) in f −1 (B) such that xn → x as n → ∞, indeed choose
1
xn ∈ B x, ∩ f −1 (B), ∀n ∈ N.
n
As f is continuous, using Result 1.6.21, we have
f (xn ) → f (x) as n → ∞.
i i
i i
i i
Preliminaries 69
|f (z)−f (z0 )| ≤ |f (z)−fN (z)|+|fN (z)−fN (z0 )|+|fN (z0 )−f (z0 )| < + + = .
3 3 3
Therefore, f is continuous on E.
i i
i i
This page is intentionally left blank
i i
2
Analytic Functions
2.1 DIFFERENTIABILITY
Definition 2.1.1 Let be a region, z0 ∈ and f : → C. f is said to be dif-
ferentiable at z0 if lim f (z)−f
z−z0
(z0 )
exists in C or equivalently, lim f (z0 +h)−f
h
(z0 )
z→z0 h→0
exists in C, and this limit is called the derivative of f at z0 , which is denoted
by f (z0 ).
71
i i
i i
i i
72 Differentiability
i i
i i
i i
Analytic Functions 73
= nz0n−1 .
Thus, if we set
f (z) − f (z0 )
R(z) = − f (z0 ) in B(z0 , r) \ {z0 }
z − z0
and
g(w) − g( f (z0 ))
S(w) = − g ( f (z0 )) in B( f (z0 ), r) \ {f (z0 )}
w − f (z0 )
i i
i i
i i
74 Differentiability
i i
i i
i i
Analytic Functions 75
k × (k − 1) × · · · × (k − j + 1)
k ∈ N, where kCj =
1 × 2 × ··· × j
k × (k − 1) × · · · × (k − j + 1)
=
1 × 2 × ··· × j
k × (k − 1) × · · · × (k − j + 2)
×
1 × 2 × ··· × j − 1
k × (k − 1) × · · · × (k − j + 2) (k − j + 1)
= × +1
1 × 2 × ··· × j − 1 j
k × (k − 1) × · · · × (k − j + 2) (k + 1)
= ×
1 × 2 × ··· × j − 1 j
= (k + 1)Cj .
Again applying Theorem 2.1.6 and by using nC0 = nCn = 1, ∀n ∈ N, we get
( f · g)(k+1) = (( f · g)(k) )
⎛ ⎞
k
= ⎝ kCj f ( j) · g(k−j) ⎠
j=0
k
= kCj f ( j+1) · g(k−j) + f ( j) · g(k−j+1)
j=0
k+1 k
= kCj−1 f ( j) · g(k−j+1) + kCj f ( j) · g(k−j+1)
j=1 j=0
i i
i i
i i
76 Differentiability
k
= (k + 1)C0 f · g(k+1) + (k + 1)Cj f ( j) · g(k+1−j)
j=1
+ (k + 1)Ck+1 f (k+1)
·g
(k+1)
= (k + 1)Cj f ( j) · g(k+1−j) .
j=0
i i
i i
i i
Analytic Functions 77
Example 2.1.17 In general, though fxy and fyx exist at a point, they need not
be equal to each other.
⎧ 3
⎨ x y
(x, y) (0, 0)
Consider f (x, y) = x2 + y2 .
⎩
0 (x, y) = 0
Now for s, t ∈ R \ {0}, we have
f (0 + s, 0) − f (0, 0) 0−0
fx (0, 0) = lim = lim =0
s→0 s s→0 s
f (0 + s, t) − f (0, t) s2 t
fx (0, t) = lim = lim 2 =0
s→0 s s→0 s + t2
f (0, 0 + t) − f (0, 0) 0−0
fy (0, 0) = lim = lim =0
t→0 t t→0 t
f (s, 0 + t) − f (s, 0) s3
fy (s, 0) = lim = lim 2 = s.
t→0 t t→0 s + t2
Therefore, we have
fy (s, 0) − fy (0, 0) s
fxy (0, 0) = lim = lim = 1
s→0 s s→0 s
and
fx (0, t) − fx (0, 0) 0−0
fyx (0, 0) = lim = lim = 0.
t→0 t t→0 t
Thus, fxy (0, 0) fyx (0, 0).
Next we recall mean-value theorem from real analysis, which will be applied
in the following sequel.
i i
i i
i i
78 Differentiability
Define φ : [a, b] → R by
Therefore,
|(x, y)|2 = x2 + y2
= x(u(b) − u(a)) + y(v(b) − v(a))
= xu(b) + yv(b) − (xu(a) + yv(a))
= φ(b) − φ(a)
= φ (t)(b − a)
= (xu (t) + yv (t))(b − a)
≤ x2 + y2 (u (t))2 + (v (t))2 (b − a)
(by Cauchy–Schwarz inequality. )
= |(x, y)|| f (t)|(b − a).
If |(x, y)| 0, then we have |(x, y)| ≤ | f (t)|(b − a). If |(x, y)| = 0, then
|(x, y)| = 0 ≤ | f (t)|(b − a).
Proof: Let fx , fy , and fxy exist on B ((a, b), r). Fix (s, t) ∈ C such that (a +
s, b + t) ∈ B ((a, b), r). If F : [b, b + t] → R is defined by
i i
i i
i i
Analytic Functions 79
= F(b + t) − F(b)
= tF (b + λt), for some λ ∈ (0, 1)
= t fy (a + s, b + λt) − fy (a, b + λt)
= stfxy (a + μs, b + λt), for some μ ∈ (0, 1).
f (z0 + h) − f (z0 )
Proof: As f is differentiable at z0 , we have f (z0 ) = lim
h→0 h
exists or equivalently,
i i
i i
i i
80 Cauchy–Riemann Equations
Equating the real and imaginary parts on both sides, we get ux = vy and
vx = −uy . Thus, u and v satisfy C–R equations.
Remark 2.2.3: From the above theorem, we conclude that a necessary con-
dition on f has to be differentiable at (x, y) such that f should satisfy the C–R
equations at (x, y) (i.e., fx = −ify ).
∂u ∂u
= 2x, = 0
∂x ∂y
∂v ∂v
= 0, = −3y2 .
∂x ∂y
∂u ∂v
Therefore, . Thus, f does not satisfy C–R equations at any (x, y) ∈
∂x ∂y
C\{(0, 0)}, and hence, it is not differentiable on C.
i i
i i
i i
Analytic Functions 81
Therefore, f does not satisfy C–R equations at any (x, y) ∈ C, and hence, it is
not differentiable on C.
Proof: (At present, assume the fact that partial derivatives of u and v of all
orders exist and they are continuous. This will be obtained as a consequence
of Theorem 4.3.3.) By Theorem 2.2.2, if f is differentiable on and u and
v satisfy C–R equations, then ux = vy and uy = −vx . Now applying Young’s
theorem (Theorem 2.1.20), we get
∂ux ∂uy ∂vy −∂vx
1. uxx + uyy = + = + = vyx − vxy = 0. Hence, u is
∂x ∂y ∂x ∂y
harmonic.
∂vx ∂vy −∂uy ∂ux
2. vxx + vyy = + = + = uxy − uyx = 0. Hence, v is
∂x ∂y ∂x ∂y
harmonic.
i i
i i
i i
82 Cauchy–Riemann Equations
i i
i i
i i
Analytic Functions 83
vy = ux = 3x2 − 3y2 + 4x + 1
implies
v = vy dy = (3x2 − 3y2 + 4x + 1) dy = 3x2 y − y3 + 4xy + y + φ(x)
v = 3x2 y − y3 + 4xy + y + c
i i
i i
i i
84 Cauchy–Riemann Equations
2(cos(2z) − 1)
=
(1 − cos(2z))2
2
=
cos(2z) − 1
2
=
cos2 (z) − sin2 (z) − cos2 (z) − sin2 (z)
2
=
−2 sin2 (z)
= − csc2 (z)
i i
i i
i i
Analytic Functions 85
As
Now,
f (z) = exp(z)(z + 1) dz
= (z + 1) exp(z) − exp(z) dz
= (z + 1) exp(z) − exp(z) = z exp(z).
i i
i i
i i
86 Cauchy–Riemann Equations
Example 2.2.16 Find an analytic function f whose real and imaginary parts
are u and v, respectively, if u + v = exp(−y) (cos(x) − sin(x)).
We find
ux + vx = exp(−y) (− sin(x) − cos(x)) , (2.2)
uy + vy = − exp(−y) (cos(x) − sin(x)) ,
−vx + ux = exp(−y) (sin(x) − cos(x)) . (2.3)
Now, equations (2.2) + (2.3) and (2.2) − (2.3) imply that
ux = − exp(−y) cos(x),
vx = − exp(−y) sin(x).
Therefore,
f (z) = ux (z, 0) + ivx (z, 0)
= −(cos(z) + i sin(z)) = − exp(iz)
and hence,
f (z) = i exp(iz) + c = exp(iz) + c
for some c ∈ C.
Exercise 2.2.17
1. Prove that the following functions are harmonic and find the harmonic
conjugates.
(a) u(x, y) = x3 − 3xy2 .
(b) u(x, y) = x4 + y4 − 6x2 y2 + 2xy − x.
(c) u(x, y) = exp(−x)(x sin(y) − y cos(y)).
(d) u(x, y) = cosh(x) sin(y).
(e) u(x, y) = exp(−x)(x cos(y) + y sin(y)).
y
2. Verify that u(x, y) = 2 , ∀(x, y) ∈ C \ {(0, 0)} is a harmonic
x + y2
function and find an analytic function whose real part is u.
3. Verify that v(x, y) = −3 exp(2x) sin(2y), ∀(x, y) ∈ C is a harmonic
function and find an anlytic function whose imaginary part is v.
4. Verify that u(x, y) = − exp(x2 − y2 ) sin(2xy), ∀(x, y) ∈ C and find an
analytic function whose real part is u. Find also the harmonic conjugate
of u.
i i
i i
i i
Analytic Functions 87
i i
i i
i i
88 Cauchy–Riemann Equations
(x1, y2)
(x2, y2)
(x1, y1)
As (x1 , y1 ) and (x1 , y2 ) lie on a vertical line segment in B(z0 , r), we have
u(x1 , y1 ) = u(x1 , y2 ). As (x1 , y2 ) and (x2 , y2 ) lie on a horizontal line segment
in B(z0 , r), we have u(x1 , y2 ) = u(x2 , y2 ). Thus, we get u(z1 ) = u(z2 ), and
hence, u is constant on B(z0 , r).
By the same argument, from vx = vy = 0, we can prove that v is constant
on B(z0 , r), and so f is a constant function on B(z0 , r).
If a, b are any two points in , then by Theorem 1.4.31, we can find a
polygon such that
1. it passes through the vertices a = a0 , a1 , a2 , . . . an = b,
i i
i i
i i
Analytic Functions 89
i i
i i
i i
90 Cauchy–Riemann Equations
∂ 2 u2 ∂ 2 u2
RESULT 2.2.22 If f = u + iv is analytic, then + = 2| f |2 .
∂x2 ∂y2
∂u ∂v
Proof: First we note that u and v satisfy the C–R equations, = and
∂x ∂y
∂u ∂v
=− .
∂y ∂x
Now,
2
∂ 2 u2 ∂ ∂u ∂u ∂ 2u
= 2u = 2 + 2u .
∂x2 ∂x ∂x ∂x ∂x2
2
∂ 2 u2 ∂u ∂ 2u
Similarly, we get = 2 + 2u . Therefore, using the fact that
∂y2 ∂y ∂y2
u is harmonic, we get
2 2
∂ 2 u2 ∂ 2 u2 ∂u ∂u ∂ 2u ∂ 2u
+ = 2 + + 2u + 2
∂ 2 x2 ∂y2 ∂x ∂y ∂x2 ∂y
2 2
∂u ∂u
= 2 +
∂x ∂y
2 2
∂u ∂v
= 2 + (using C–R equations)
∂x ∂x
∂f ∂u ∂v
= 2| f |2 as f = = +i .
∂x ∂x ∂x
i i
i i
i i
Analytic Functions 91
∂u ∂v
Proof: We know that C–R equations in Cartesian form are given by =
∂x ∂y
∂u ∂v
and = − . From the relation x = r cos(θ ) and y = r sin(θ ) between the
∂y ∂x
Cartesian coordinates and polar coordinates, we get,
∂u ∂u ∂x ∂u ∂y
r = r +
∂r ∂x ∂r ∂y ∂r
∂v ∂v
= − r sin(θ ) + r cos(θ )
∂x ∂y
∂v ∂x ∂v ∂y
= +
∂x ∂θ ∂y ∂θ
∂v
= ,
∂θ
∂v ∂v ∂x ∂v ∂y
r = r +
∂r ∂x ∂r ∂y ∂r
∂u ∂u
= r sin(θ) − r cos(θ )
∂x ∂y
∂u ∂x ∂u ∂y
= − −
∂x ∂θ ∂y ∂θ
∂u
= − .
∂θ
1 1
Proof: If z = x + iy, then we have x =(z + z) and y = (z − z). Therefore,
2 i2
∂f ∂f ∂x ∂f ∂y 1 ∂f ∂f
= + = +i = 0.
∂z ∂x ∂z ∂y ∂z 2 ∂x ∂y
Example 2.2.26 There exists a function f on C such that fx and fy exist and
satisfy C–R equation at a point, but it is not differentiable at the same point.
i i
i i
i i
92 Cauchy–Riemann Equations
Define
⎧
⎨ xy(x + iy) if (x, y) (0, 0)
f (x, y) = x2 + y2 , ∀(x, y) ∈ C
⎩
0 if (x, y) = (0, 0)
As f (x, 0) = 0, ∀x ∈ R, we obtain fx (0, 0) = 0, and as f (0, y) = 0, ∀y ∈ R,
we get fy (0, 0) = 0. Hence, the C–R equation fx = −ify is satisfied at (0, 0).
However,
f (2h, h) − f (0, 0) 2h2 (2h + ih)
lim = lim
h→0 2h + ih h→0 5h2 (2h + ih)
2
=
5
1
2
h2 (h + ih)
= lim 2
h→0 2h (h + ih)
f (h, h) − f (0, 0)
= lim
h→0 h + ih
Hence, f is not differentiable at (0, 0).
Example 2.2.27 There exists a function f on C such that fx and fy exist and
are continuous at a point, but it is not differentiable at the same point.
Consider the function
f (x, y) = x − iy, ∀(x, y) ∈ C
Now, fx = 1 and fy = −i, ∀(x, y) ∈ C. Clearly, fx and fy are continuous
functions. However, it is not differentiable at (0, 0) as f does not satisfy C–R
equations at (0, 0).
i i
i i
i i
Analytic Functions 93
f (x0 + s, y0 + t) − f (x0 , y0 )
= s(ux + r1 ) + t(uy + r2 ) + is(vx + r3 ) + it(vy + r4 )
= s(ux + r1 ) + t(−vx + r2 ) + is(vx + r3 ) + it(ux + r4 )
= (s + it)ux + i(s + it)vx + s(r1 + ir3 ) + t(r2 + ir4 )
= (s + it)(ux + ivx ) + s(r1 + ir3 ) + t(r2 + ir4 )
= (s + it)fx + s(r1 + ir3 ) + t(r2 + ir4 ).
Hence,
f (x0 + s, y0 + t) − f (x0 , y0 )
0
≤ − fx
s + it
s t
≤ (r1 + ir3 ) + (r2 + ir4 )
s + it s + it
≤ |r1 | + |r3 | + |r2 | + |r4 | → 0 as (s, t) → (0, 0).
f ((x0 + iy0 ) + (s + it)) − f (x0 + iy0 )
Thus, f (z0 ) = lim exists and is
s+it→0 s + it
equal to fx (z0 ).
we get
and
ψx (x, y) = −vx (x, −y), ψy (x, y) = vy (x, −y), ∀(x, y) ∈ .
i i
i i
i i
and
ϕy (x, y) = −uy (x, −y) = vx (x, −y) = −ψx (x, y).
Similarly, we can prove that ϕx = ψy and ϕy = −ψx imply ux = vy and uy =
−vx . Hence, using Theorems 2.2.2 and 2.2.28, we get that f is differentiable
iff ux , uy , vx and vy are continuous and satisfy C–R equations iff ϕx , ϕy , ψx
and ψy are continuous and satisfy C–R equations iff g is differentiable.
The convergence of a power series depends on the sequence (an ) and the
point z.
∞
2. If 0 < S < R, then an (z − a)n converges absolutely and uniformly
n=0
on {z : |z − a| ≤ S}.
∞
3. If |z − a| > R, then an (z − a)n does not converge.
n=0
−1
1
Proof: Define R = lim sup |an | n ∈ [−∞, ∞] with the convention that
n→∞
1 1
= ∞ and = 0. We claim that this R has the desired properties.
0 ∞
i i
i i
i i
Analytic Functions 95
k R R t
i i
i i
i i
1 1
Therefore, we have |ank | nk > , ∀k ≥ m, which implies
t
1
|ank | > , ∀k ≥ m.
t nk
Hence, for |z − a| > t and k ≥ m, we have
nk
|z − a|
|ank (z − a) | >
nk
> 1, ∀k ≥ m.
t
Thus, the subsequence (ank (z−a)nk ) of (an (z−a)n ) does not converge to
∞
0. Hence, (an (z −a)n ) itself does not converge to 0. Thus, an (z −a)n
n=0
diverges by using Result 1.3.27.
∞
THEOREM 2.3.3 Let the radius of convergence of an (z − a)n be R. Then
n=0
∞
1. the radius of convergence of nan (z − a)n−1 is also R,
n=1
∞
2. if f (z) = an (z − a)n on |z − a| < R, then f is analytic on |z − a| < R
n=0
∞
and f (z) = nan (z − a)n−1 .
n=1
1 1
Proof: Recall that = lim sup |an | n .
R n→∞
1 1
1. First, we prove that n n → ∞ as n → ∞. If rn = n n − 1, ∀n ∈ N, then
we have for each n ≥ 2,
n(n − 1) 2
n = (rn + 1)n = 1 + nC1 rn + nC2 rn2 + · · · + rnn > rn ,
2
2 1
which implies that rn < → 0 as n → ∞. Therefore, n n −1 →
n−1
0 as n → ∞. Hence, our claim follows. Using Theorem 1.3.19, we get
1 1 1 1
lim sup n n |an | n = lim n n lim sup |an | n = R.
n→∞ n→∞ n→∞
i i
i i
i i
Analytic Functions 97
∞
m
2. Put g(z) = nan (z − a)n−1 , Sm (z) = an (z − a)n , and Rm (z) =
n=1 n=0
∞
an (z − a)n , ∀z ∈ B(a, R) and ∀m ∈ N. Obviously, we have the
n=m+1
following:
(a) f (z) = Sm (z) + Rm (z), ∀m ∈ N on B(a, R).
(z) → g(z) as m → ∞ on B(a, R).
(b) Sm
∞
(c) n|an |rn−1 → 0 as m → ∞ for 0 ≤ r < R, using Result
n=m+1
1.3.26.
For a fixed z0 ∈ B(a, R) and any z ∈ B(a, R), choose r ∈ R such that
From the definition of Sm (z ), given > 0, there exists δ > 0 such that
0
Sm (z) − Sm (z0 )
− Sm (z0 ) < whenever 0 < |z − z0 | < δ. (2.7)
z − z0 3
i i
i i
i i
∞
COROLLARY 2.3.4 Let the power series an (z−a)n converge in |z−a| < R
n=0
∞
f (n) (a)
for some R > 0. If f (z) = an (z − a)n , ∀z ∈ B(a, R), then an = n! ,
n=0
∀n = 0, 1, 2, 3, . . .
Proof: Applying the previous theorem repeatedly, first we note that f (m)
exists for every m = 0, 1, 2, 3, . . . and
m−1 ∞
d m
f (m) (z) = an (z − a)n + an (z − a)n
dzm n=m
n=0
∞
= n(n − 1) · · · (n − m + 1)an (z − a)n−m
n=m
∞
= am m! + n(n − 1) · · · (n − m + 1)an (z − a)n−m
n=m+1
∞
∞
COROLLARY 2.3.5 an (z − a)n = bn (z − a)n , ∀z ∈ B(a, R) iff an =
n=0 n=0
bn , ∀n = 0, 1, 2, 3, . . .
i i
i i
i i
Analytic Functions 99
∞
To prove the converse, let f (z) = an (z − a)n , ∀z ∈ B(a, R). If an =
n=0
f (n) (a)
bn , ∀n = 0, 1, 2, 3, . . . then by using previous corollary, we get an = =
n!
bn , ∀n = 0, 1, 2, 3, . . .
Example 2.3.7 Find the radius of convergence of the following power series:
∞
1. nn zn ;
n=0
∞
2. n2 zn ;
n=0
n2
∞ 1
3. 1− zn ;
n=0 n
∞
4. (4 + i3)n zn .
n=0
Solution:
1. Let an = nn , ∀n = 0, 1, 2, . . . . As |an |1/n = n → ∞ as n → ∞, the
radius of convergence of this power series is 0.
2. Let an = n2 , ∀n = 0, 1, 2, . . . . As |an |1/n = n2/n → ∞ as n → ∞, the
radius of convergence of this power series is 0.
2 1n
1 n 1 n 1
3. As lim 1− = lim 1 − = , the radius of
n→∞ n n→∞ n e
convergence of this power series is e.
i i
i i
i i
1
4. As lim (|4 + i3|n ) n = |4 + i3| = 5, the radius of convergence of
n→∞
∞ 1
(4 + i3)n zn is .
n=0 5
Exercise 2.3.8
∞ 2n
1. zn ;
n=0 n!
∞ √
n zn ;
2. 2
n=0
∞ n+1 an+1 √
3. zn (Hint. Find lim which is same as lim n an .);
n=0 (n + 2)(n + 3) n→∞ an n→∞
∞ 1
4. 2
zn ;
n=0 n
∞ (n!)2
5. zn .
n=0 (2n)!
Answer: 1. ∞; 2. 1; 3. 1; 4. 1; 5. 4.
∞
Proof: First we note that, as an converges at 1, the radius R of conver-
n=0
∞
gence of an is at least 1 by Abel’s theorem. If R > 1, then f is continuous
n=0
at 1, and hence, the theorem follows. Therefore, assume that R = 1.
∞
Case 1: f (1) = an = 0.
n=0
n
n
If αn = ak and Sn (z) = ak zk , ∀z ∈ B(0, 1) and ∀n = 0, 1, 2, . . .,
k=0 k=0
i i
i i
i i
|1 − z|
then for z ∈ B(0, 1) with ≤ K , for some K > 0, we get
1 − |z|
n
Sn (z) = ak zk
k=0
n
= α0 + (αk − αk−1 )zk
k=1
n n
= α0 + αk zk − αk−1 zk
k=1 k=1
n−1 n−1
= α0 + αk zk + αn zn − α0 z − αk zk+1
k=1 k=1
n−1
= α0 (1 − z) + αk (zk − zk+1 ) + αn zn
k=1
n−1
= α0 (1 − z) + αk zk (1 − z) + αn zn
k=1
n−1
= (1 − z) αk zk + αn zn .
k=0
∞ ∞
f (z) = (1 − z) αk zk + lim αn zn = (1 − z) αk zk ,
n→∞
k=0 k=0
0<δ< .
m−1
2 1+ |αn |
k=1
i i
i i
i i
|1 − z|
If 0 < |z − 1| < δ and ≤ K , then
1 − |z|
m−1 ∞
| f (z)| ≤ |1 − z| |αk zk | + |αk zk |
k=0 k=m
m−1 ∞
< |1 − z| |αk | + |z |
k
2K
k=0 k=m
m−1 ∞
< δ |αk | + |zm ||1 − z| |zk |
2K
k=0 k=0
|1 − z|
< +
2 2K 1 − |z|
≤ + = .
2 2
Hence, the theorem follows in this case.
Case 2: f (1) 0
∞
If g(z) = f (z) − f (1) = a0 − f (1) + an zn , then g(1) = 0. Then, by
n=1
Case 1, we obtain
|1 − z|
g(z) → g(1) as z → 1 with is bounded,
1 − |z|
and hence,
|1 − z|
f (z) − f (1) → 0 as z → 1 with is bounded.
1 − |z|
∞ zn
LEMMA 2.4.2 exp(z) = for z ∈ C at which the series converges.
n=0 n!
i i
i i
i i
∞
Proof: Let exp(z) = an zn . As it is the solution of the above initial value
n=0
problem, we have
∞ ∞ ∞
an zn = an−1 zn−1 = nan zn−1 , and a0 = f (0) = 1.
n=0 n=1 n=1
an−1
Hence, an = , n ∈ {0, 1, 2, . . .}, and a0 = f (0) = 1. We claim that
n
1
an = , ∀n = 0, 1, 2, 3, . . .
n!
We prove our claim by induction on n. For n = 0, we have a0 = 1. Assume
that the claim holds for some n ∈ N. Now,
an 1 1
an+1 = = = .
n+1 (n + 1)n! (n + 1)!
∞ zn
Thus, our claim follows. Therefore, exp(z) = , ∀z ∈ B(0, R), where R
n=0 n!
∞ zn
is the radius of convergence of .
n=0 n!
∞ zn
LEMMA 2.4.3 The radius of convergence of exp(z) = is ∞.
n=0 n!
i i
i i
i i
1
(n + 1)!
≤ lim sup
n→∞ 1
n!
= 0
1
1 n
and hence, lim sup = 0.
n→∞ n!
∞ zn
Therefore, the radius of convergence of is ∞.
n=0 n!
Proof: As
d
(exp(z) · exp(w − z)) = exp(z) · exp(w − z) − exp(z) · exp(w − z) = 0,
dz
we get exp(z) · exp(w − z) as a constant, say K . Putting z = 0 and using
exp(0) = 1, we get K = exp(w). Thus, we obtain
∞ xn
Proof: First, note that exp(0) = 1, for x > 0, we have exp(x) = >
n=0 n!
1 and by previous corollary, for x < 0, exp(x) = exp(−x)
1
∈ (0, 1), since
−x > 0 ⇒ exp(−x) > 1. Therefore, we have exp(x) ≥ 1 if x ≥ 0 and
0 < exp(x) < 1 if x < 0. So, exp is a mapping from R into (0, ∞). To prove
exp : R → (0, ∞) is onto, let y > 0 be arbitrary.
i i
i i
i i
By definition, both cos and sin are differentiable functions on C, and the
∞ (−1)n z2n
power series representations of these two functions are cos(z) =
n=0 (2n)!
∞ (−1)n z(2n+1)
and sin(z) = .
n=0 (2n + 1)!
i i
i i
i i
Proof: By direct computation, we get (1) and the Euler’s formula as given
below.
exp(−iz) − exp(iz) exp(iz) − exp(−iz)
sin(−z) = =− = − sin(z)
i2 i2
exp(−iz) + exp(iz)
cos(−z) = = cos(z)
2
i i
i i
i i
Proof: As
i2 sin(z + w) = exp(i(z + w)) − exp(−i(z + w))
= exp(iz) exp(iw) − exp(−iz) exp(−iw)
= (cos(z) + i sin(z))(cos(w) + i sin(w))
− (cos(z) − i sin(z))(cos(w) − i sin(w))
= (cos(z) cos(w) − sin(z) sin(w)) + i(cos(z) sin(w)
i i
i i
i i
d d
LEMMA 2.4.14 (1) sin(z) = cos(z) and (2) cos(z) = − sin(z).
dz dz
i i
i i
i i
r3 r2 r4 r2 r4
sin(r) ≥ r − ⇒ − cos(r) + 1 ≥ − ⇒ cos(r) ≤ 1 − + .
6 2 24 2 24
√ 3 9
Therefore, cos( 3) ≤ 1 − + < 0, and from the definition, we obtain
2 24
cos(0) = 1. Therefore,
√ by intermediate value theorem (Theorem 1.6.30),
there exists r0 ∈ (0, 3) such that cos(r
√0 ) = 0.
Next, we claim that if 0 < r < r0 < 3, then cos(r) 0. From one of the
above inequalities, we get
√
r3 r2 ( 3)2 r
sin(r) ≥ r − =r 1− >r 1− = > 0.
6 6 6 2
d
Therefore, cos(r) = − sin(r) < 0, and hence, cos is strictly decreasing on
dr
(0, r0 ), applying Lemma 2.2.18. Since sin(r) > 0, ∀r ∈ (0, r0 ) and sin2 (r) +
cos2 (r) = 1, we get sin, and it is strictly increasing on (0, r0 ). Therefore,
0 < sin(r) < sin(r0 ) = 1 − cos2 (r0 ) = 1.
Definition 2.4.16 Define π = 2r0 , where r0 is the least positive real number
such that cos(r0 ) = 0.
π π
Example 2.4.17 Prove that cos = 0, sin = 1, cos(π ) = −1,
2 2
sin(π ) = 0, cos(2π ) = 1, and sin(2π) = 0.
i i
i i
i i
By definition, π = 2r0 , where r0 is the least positive real number such that
cos(r0 ) = 0. Therefore,
π π π
cos = cos(r0 ) = 0, sin = ± 1 − cos2 = ±1,
2 2 2
π π π
but sin = sin(r0 ) > 0, we have sin = 1. Therefore, exp i = i,
2 2 2
which implies that
exp(iπ) = i2 = −1 and exp(i2π ) = (−1)2 = 1.
Equating real and imaginary parts in the above two equations, we can get the
required identities.
Proof: If 0 < t < 2π and exp(it/4) = u + iv, then using the following:
t π
1. 0 < < ,
4 2
2. exp(it/4) = cos(t/4) + i sin(t/4),
π π
3. is the least positive real number such that cos = 0,
2 2
π
4. sin > 0 on 0, ,
2
i i
i i
i i
we get 0 < u < 1 and 0 < v < 1. As u + iv is not equal to any of 1, −1, i,
and −i, which are the solutions of z4 = 1, we obtain that exp(it/4) = u + iv
is not a solution of z4 = 1. Therefore, exp(it) 1.
1 = exp(iy)
= exp(i(y − 2nπ + 2nπ))
= exp(i(y − 2nπ)) exp(i2nπ)
= exp(i(y − 2nπ)),
Remark 2.4.22: From addition theorem of exp function and Result 2.4.20,
we conclude that i2kπ are periods of exp.
Now, we prove a simple but useful lemma on the integrals of a periodic
function.
i i
i i
i i
= f (y − mp) dy + f (z − (m + 1)p) dz
a−mp 0
(by putting y = x − mp and z = x − (m + 1)p)
p
a−mp
i i
i i
i i
and hence,
Case 3: −1 < Re z < 0, 0 < Im z < 1. Then, −Re z = cos(θ ) for some
θ ∈ 0, π2 . As sin(θ) > 0 in 0, π2 , we get
Im z = 1 − (Re z)2 = 1 − cos2 (θ ) = sin(θ ),
and hence,
Case 4: −1 < Re z < 0, −1 < Im z < 0. Then, −Re z = cos(θ ) for some
θ ∈ 0, π2 . As sin(θ) > 0 in 0, π2 , we get
−Im z = 1 − (Re z)2 = 1 − cos2 (θ ) = sin(θ ),
and hence,
and hence,
i i
i i
i i
z z
Proof: As z 0, we write z = |z| , where = 1. Therefore, from the
|z| |z|
z
proof of previous result, we get = exp(iy) for some y ∈ (−π , π ]. Hence,
|z|
the corollary follows.
r
r sin (q )
q
A = (0,0) r cos (q ) B
i i
i i
i i
Example 2.4.28 Find the domains of tan, csc, sec, and cot.
π
We have already proved that is the least positive real number at which cos
2
vanishes. See Result 2.4.15. Therefore, for any k ∈ Z,
π π
π exp i + kπ + exp −i + kπ
cos + kπ = 2 2
2 2
i exp (ikπ ) − i exp (−ikπ )
=
2
i(−1)k − i(−1)−k
=
2
= 0 (since ± 1 = (±1)−1 ).
i i
i i
i i
π
⇒ z = (2k + 1) , k ∈ Z
2
π
⇒ z = kπ + , k ∈ Z.
2
Therefore,
π
{z ∈ C : cos(z) = 0} = kπ + : k ∈ Z .
2
Similarly, we get
sin(z) = 0 ⇔ cos(z) = ±1
⇔ exp(iz) = ±1 = exp(0) or exp(iπ )
⇔ exp(iz) = 1 or exp(i(z − π )) = 1
⇔ z = 2nπ or z = 2nπ + π , n ∈ Z
⇔ z = kπ , k ∈ Z.
Therefore,
{z ∈ C : sin(z) = 0} = {kπ : k ∈ Z} .
π
Therefore, the domain of tan and sec is same as C \ kπ + : k ∈ Z and
2
that of cot and csc is C \ {kπ : k ∈ Z}.
THEOREM 2.4.29 lim tan(x) = +∞, lim tan(x) = −∞, and tan :
x→ π2 − x→− π2 +
π π
− , → R is a bijection.
2 2
i i
i i
i i
and
π 1
0 < x − < δ2 ⇒ | cos(x)| < .
2 2M
If δ = min{δ1 , δ2 , π2 }, then δ > 0, and
π π | sin(x)| sin(x) 1
−δ <x< ⇒ = > > M.
2 2 | cos(x)| cos(x) 2 cos(x)
Therefore, lim tan(x) = +∞.
z→ π2 −
π
Similarly, using the continuity of cos and sin at , we obtain that for a given
2
K < 0, there exists δ3 > 0 such that
π
1 1
0 < x − < δ3 ⇒ | cos(x)| < ⇒ 2 cos x <
2 −2K −k
1 1
⇒ > −k ⇒ < k.
2 cos x −2 cos x
π π
If r = min{δ1 , δ3 , π2 }, then r > 0, and − < x < − + r implies
2 2
| sin(x)| − sin(x) 1 sin(x) 1
= > ⇒ < <K
| cos(x)| cos(x) 2 cos(x) cos(x) −2 cos(x)
π
since sin < 0 and cos > 0 on − , 0 . Therefore, lim tan(x) = −∞.
2 z→− π2 +
sin(0) 0 π π
We know that tan(0) = = = 0 and is continuous on − , .
cos(0) 1 π 2 2
Given any y > 0, using lim tan(x) = +∞, we find t ∈ 0, such that
x→ π2 − 2
tan(t) > y. Then by using intermediate value theorem, there exists x ∈ (0, t)
such that tan(x) = y. Similarly, for a given y < 0, using lim tan(x), we can
x→− π +
π π π2
find x ∈ − , 0 such that tan(x) = y. Thus, tan : − , → R is onto.
2 2 2
cos(x) cos(x) + sin(x) sin(x) 1
Since, tan (x) = 2 (x)
= 2
and cos(x) 0, ∀x ∈
π π cos cos
(x)
π π
− , , we get that tan is strictly increasing on − , . Therefore, tan :
π 2 2
π
2 2
− , → R is one-to-one. Thus, the theorem follows.
2 2
i i
i i
i i
Exercise 2.4.31 Prove that 1 + tan2 (z) = sec2 (z), 1 + cot2 (z) = csc2 (z), and
tan(z) + tan(w)
tan(z + w) = , ∀z, w ∈ C.
1 − tan(z) tan(w)
π π
Exercise 2.4.32 Prove that arctan x → as x → +∞ and arctan x → −
2 2
as x → −∞.
Exercise 2.4.33
i i
i i
i i
π
(e) sec z − 2 = csc(z)
π
(f) csc z − 2 = − sec(z)
π
(g) sec z + 2 = − csc(z)
π
(h) csc z + 2 = sec(z)
i i
i i
i i
LEMMA 2.5.4 sinh and cosh are differentiable and sinh = cosh and cosh =
sinh.
Exercise 2.5.5 Prove that the power series representations of cosh(z) and
∞ z2n ∞ z2n+1
sinh(z) are and .
n=0 (2n)! n=0 (2n + 1)!
1. sinh(−z) = − sinh(z).
2. cosh(−z) = cosh(z).
i i
i i
i i
Proof: As
sinh(z + w) sinh(z) cosh(w) + sinh(z) cosh(w)
tanh(z + w) = = ,
cosh(z + w) cosh(z) cosh(w) + sinh(z) sinh(w)
sinh(z) sinh(w)
+
cosh(z) cosh(w) tanh(z) + tanh(w)
tanh(z + w) = = . (2.8)
sinh(z) sinh(w) 1 + tanh(z) tanh(w)
1+
cosh(z) cosh(w)
Since
sinh(−w) sinh(w)
tanh(−w) = =− = − tanh(w),
cosh(−w) cosh(w)
by replacing w by −w in equation (2.8), we get
tanh(z) − tanh(w)
tanh(z − w) =
1 − tanh(z) tanh(w)
Exercise 2.5.9 Prove that for every z ∈ C, 1 − tanh2 (z) = sech2 (z) and
coth2 (z) − 1 = csch2 (z).
i i
i i
i i
The proof of the above result is easy, and hence, it is left to the reader as
a simple exercise.
i i
i i
i i
3
Rational Functions
and Multivalued
Functions
123
i i
i i
i i
Proof: By division algorithm, we can find polynomials Q(z) and R(z) such
that
P(z) = (z − z0 )Q(z) + R(z), ∀z ∈ C
with 0 < deg R(z) < deg (z − z0 ) or R(z) = 0, ∀z ∈ C. As deg (z − z0 ) is 1,
R(z) must be a constant and is equal to
P(z) = (z − z0 )m Q(z), ∀z ∈ C
for some polynomial Q such that Q(z0 ) 0, then m is called the order of the
zero z0 of P.
P(z) = (z − z0 )m Q(z), ∀z ∈ C
dk k
dj d k−j
k
(z − z0 ) m
Q(z) = kCj j (z − z0 )m · k−j Q(z), ∀k ∈ N.
dz dz dz
j=0
dj
(z − z0 )m = m(m − 1) · · · (m − j + 1)(z − z0 )m−j and m − j > 0
dzj
and hence, we get P(k) (z0 ) = 0, 1 ≤ k ≤ m − 1. Furthermore, using the same
observation, we obtain
m
d
P (z0 ) =
(m)
(z − z0 ) Q(z) = m!Q(z0 ) 0.
m
dzm
Thus, the result follows.
Converse of the above result is also true and is given by the following result.
i i
i i
i i
P(z) = (z − z0 )2 P2 (z), ∀z ∈ C.
Using
P(z) = (z − z0 )m Pm (z), ∀z ∈ C
i i
i i
i i
Next we discuss a theorem that states a relation between the zeroes of a poly-
nomial P with the zeroes of its derivative. For that purpose, now we discuss
how to represent a half plane described by a straight line. We know that the
equation of a straight line, which has the slope as the argument of a non-zero
complex number b and it passes through the complex number a, is given by
z = a + tb, t ∈ R. Hence, we have
z−a
z lies on the straight line iff is a real number
b
z−a
iff Im = 0.
b
z−a
Therefore, z is not on the straight line iff Im 0. Hence,
b
z−a z−a
z ∈ C : Im >0 and z ∈ C : Im <0
b b
are the two half planes determined by the straight line z = a + tb, t ∈ R.
Closed polygon
i i
i i
i i
are the zeroes of P including multiplicities. Then, using Lemma 3.1.3 and
fundamental theorem of algebra (Theorem 4.3.12), we get
d
Therefore, either by direct calculation or by finding log P(z) (with less
dz
rigor), we get
P (z) 1 1 1
= + + ··· + , ∀z ∈ C. (3.1)
P(z) z − α1 z − α2 z − αn
z − aj
Let 1 ≤ j ≤ m and let Hj = z ∈ C : Im > 0 be a half plane
bj
described by the straight line z = aj + tbj , t ∈ R. Now assume that αk ∈ Hj ,
∀k ∈ {1, 2, . . . , n}. Therefore, we have
αk − aj
Im > 0, ∀1 ≤ k ≤ n.
bj
z0 − aj
If z0 Hj , then Im ≤ 0, and hence,
bj
z0 − αk z0 − aj αk − aj
Im = Im − Im < 0.
bj bj bj
bj P (z)
n
bj
Im = Im > 0 ⇒ bj P (z0 ) 0 ⇒ P (z0 ) 0.
P(z) z0 − bk
k=1
i i
i i
i i
P
Proof: Let R = , where
Q
m
n
P(z) = ak z and Q(z) =
k
bj z j , ∀z ∈ C
k=0 j=0
Note that if R = Q P
is written in the reduced form, then every zero of P is a
finite zero of R and every zero of Q is a finite pole of R. Furthermore, ∞ is a
zero of R if deg P < deg Q and is a pole if deg P > deg Q.
P
Definition 3.1.13 Let R = be a rational function, z0 ∈ C∞ be a zero of R,
Q
and k ∈ N.
i i
i i
i i
P
Note that if R = , then R has a zero of order k at ∞ if deg P = k + deg Q
Q
and has a pole of order k at ∞ if deg Q = k + deg P.
i i
i i
i i
n
1
R(z) = G(z) + Fj , ∀z ∈ C
z − zj
j=1
P
Proof: Let R = be in reduced form. Hence, by division algorithm for
Q
polynomials (Result 3.1.2), we can write P = AQ + B, where A and B are
B
polynomials such that either B = 0 or deg B < deg Q. Therefore, R = A + .
Q
B
Let C = A − A(0) and D = A(0) + . Hence, we can write R = C + D,
Q
where C is a polynomial with C(0) = 0 and D is a rational function such that
A(0)Q + B
D(∞) ∞, as D = and degree of (A(0)Q + B) ≤ degree of Q.
Q
We note that if R(∞) is finite, then C = 0, and if R(∞) = ∞, then degree of
P > degree of Q ⇒ C is a non-constant polynomial. . . . . . . . . . . . . (I)
For every j = 1, 2, . . . , n, let Rj (w) = R zj + w1 , ∀w ∈ C∞ , where
z1 , z2 , . . . , zn are distinct finite poles of R. Note that Rj (∞) = ∞; then by
previous argument, there exist non-constant polynomial Cj with Cj (0) = 0
and rational function Dj with Dj (∞) ∞ such that
i i
i i
i i
1
Applying the change of variable w = , we get
z − zj
R(z) = R zj + w1
= Rj (w)
+ Dj(w)
= Cj (w)
1 1
= Cj + Dj
z − zj z − zj
= Ej (z) + Fj (z), (say) ∀z ∈ C∞
and Fj (zj ) ∞, j = 1, 2, . . . , n. Now the rational function
n
R−C− Ej (3.2)
j=1
n
n
1
R = (C + α) + Ej ⇒ R(z) = G(z) + Fj , ∀z ∈ C
z − zj
j=1 j=1
where G = C + α .
i i
i i
i i
i i
i i
i i
THEOREM 3.2.4 Every linear fractional transform maps circles and lines to
circles and lines.
Proof: To prove this theorem, we shall first show that the elementary linear
fractional transforms map circles and lines to circles and lines.
1. The translation operator w = z + c maps the line z = a + tb into the
line w = (a + c) + tb, and it maps the circle z = a + reit into the circle
w = (a + c) + reit . Geometrically,
(a) the line passing through a with slope m is mapped into the line
passing through a + c with slope m.
(b) the circle with centre a and radius r is mapped to the circle with
centre a + c and radius r.
2. The rotation operator w = eiθ z maps the line z = a + tb into the line
w = (eiθ a) + t(eiθ b), and it maps the circle z = a + reit into the circle
z = (eiθ a) + reit . Geometrically,
(a) the line passing through a with slope m is mapped to the line
passing through eiθ a with slope m + θ .
(b) the circle with centre a and radius r is mapped to the circle with
centre eiθ a and radius r.
i i
i i
i i
(a) the line passing through a with slope m is mapped to the line
passing through ρa with slope m.
(b) the circle with centre a and radius r is mapped to the circle with
centre ρa and radius rρ .
(b) a line not passing through 0 is mapped onto a circle (with non-zero
centre) passing through 0.
1
z = r exp(iθ ), θ ∈ [−π, π] ⇒ w = exp(−iθ ), θ ∈ [−π , π ].
r
i i
i i
i i
a1 z + b1 a2 z + b2
Proof: Let T(z) = and S(z) = , where ai , bi , ci , and
c1 z + d1 c2 z + d2
di ∈ C with ai di − bi ci 0, for i = 1, 2. As
a2 z + b2
a1 + b1
c2 z + d2 (a1 a2 + b1 c2 )z + (a1 b2 + b1 d2 )
(T ◦ S)(z) = = ,
a2 z + b2 (c1 a2 + c2 d1 )z + (c1 b2 + d1 d2 )
c1 + d1
c2 z + d2
Az + B
which is of the form , and
Cz + D
AD − BC = (a1 a2 + b1 c2 )(c1 b2 + d1 d2 ) − (a1 b2 + b1 d2 )(c1 a2 + c2 d1 )
= a1 a2 c1 b2 + a1 a2 d1 d2 + b1 c2 c1 b2 + b1 c2 d1 d2
− a1 b2 c1 a2 − a1 b2 c2 d1 − b1 d2 c1 a2 − b1 d2 c2 d1
= a1 a2 d1 d2 + b1 c2 c1 b2 − a1 b2 c2 d1 − b1 d2 c1 a2
= a1 d1 (a2 d2 − b2 c2 ) + b1 c1 (b2 c2 − a2 d2 )
= (a1 d1 − b1 c1 )(a2 d2 − b2 c2 ) 0.
az + b
Proof: Let T(z) = , ∀z ∈ C∞ with ad − bc 0. The fixed points of
cz + d
az + b
T are the solutions of z = ⇔ cz2 + (d − a)z − b = 0.
cz + d
If c 0, then the equation has at most two solutions.
i i
i i
i i
THEOREM 3.2.8 Every linear fractional transform w = T(z), which has only
w−α z−α
two distinct fixed points α, β ∈ C, satisfies the equation =γ
w−β z−β
for some γ ∈ C.
az + b
Proof: Let w = T(z) = , ∀z ∈ C∞ . As α and β are the fixed points of
cz + d
aα + b aβ + b
T , we have α = and β = . Therefore,
cα + d cβ + d
az + b aα + b (ad − bc)(z − α)
w−α = − =
cz + d cα + d (cz + d)(cα + d)
az + b aβ + b (ad − bc)(z − β)
w−β = − =
cz + d cβ + d (cz + d)(cβ + d)
Therefore,
w−α z−α
=γ
w−β z−β
cβ + d
where γ = .
cα + d
i i
i i
i i
THEOREM 3.2.9 Every linear fractional transform w = T(z), which has only
1 1
one fixed point α ∈ C, satisfies the equation = + γ for some
w−α z−α
γ ∈ C.
az + b
Proof: As in the proof of previous theorem, if w = T(z) = , ∀z ∈ C∞ ,
cz + d
then we have
aα + b
α= ⇒ α(cα − a) = b − dα.
cα + d
Since,
az + b
w−α = −α
cz + d
az + b − αcz − αd
=
cz + d
(a − cα)z + α(cα − a)
=
cz + d
(a − cα)(z − α)
=
cz + d
we have
1 1 1 cz + d 1 cz + d − a + cα
− = −1 = .
w−α z−α z−α a − cα z−α a − cα
PROBLEM 3.2.10 Find the image of |z − 2| < 2 under the linear fractional
transform T(z) = 2z−8
z
.
i i
i i
i i
Solution: We first find the images of the real and imaginary axes under z+iz−i .
x+i
As |T(x)| = x−i = 1, for all x ∈ R, the image of the real axis is the
unit circle. As T(i) = ∞, the image of the imaginary axis under the linear
fractional transform is a straight line. Seeing T(2i) = 3 and T(3i) = 2, we
conclude that the imaginary axis is mapped onto the real axis. As −1 + i lies
on the second quadrant {z : Re z < 0 and Im z > 0} and T(−1 + i) = 1 − 2i
(which lies outside the circle and below the real axis), we conclude that the
image of the second quadrant under z+iz−i is {z ∈ C : |z| > 1 and Im z < 0}.
i i
i i
i i
⎧
⎪ z 1 − z3 z2 − z4
⎪
⎪ × if z1 , z2 , z3 , z4 ∈ C
⎪
⎪ z − z z 2 − z3
⎪
⎪
1 4
⎪
⎪ z2 − z 4
⎪
⎪
⎪
⎪ if z1 = ∞, z2 , z3 , z4 ∈ C
⎪
⎪ z 2 − z3
⎪
⎪
⎪
⎨ z1 − z3
(z1 , z2 , z3 , z4 ) = if z2 = ∞, z1 , z3 , z4 ∈ C
⎪
⎪ z 1 − z4
⎪
⎪
⎪
⎪ z2 − z4
⎪
⎪
⎪
⎪ if z3 = ∞, z1 , z2 , z4 ∈ C
⎪
⎪ z 1 − z4
⎪
⎪
⎪
⎪
⎪
⎪ z1 − z3
⎩ if z4 = ∞, z1 , z2 , z3 ∈ C.
z 2 − z3
1. (1 + i, 2 − i, 3, −i).
2. (2 + i, i, 5 − i2, ∞).
4. (2, ∞, 1 − i, 3 + i).
5. (∞, 1 − i, 1 + i, i).
Solution:
1+i−3 2−i+i
1. (1 + i, 2 − i, 3, −i) = ×
1+i+i 2−i−3
−4 + i2
= = −1 − i.
1 − i3
i i
i i
i i
2 + i − 5 + i2
2. (2 + i, i, 5 − i2, ∞) =
i − 5 + i2
−3 − i3 3
= = (4 − i).
−5 + i3 17
2 + i3 − 1
3. (1 + i2, 2 + i3, ∞, 1) =
1 + i2 − 1
1 + i3 1
= = (3 − i).
i2 2
2−1+i
4. (2, ∞, 1 − i, 3 + i) =
2−3−i
1+i
= = −1.
−1 − i
1−i−i
5. (∞, 1 − i, 1 + i, i) =
1−i−1−i
1 − i2 1
= = (2 + i).
−i2 2
1. (2 + i3, 1 + i, 3 − i, 1 − i4).
Answers: 1. 19
20 − i 15 , 2. 25 (21 + i22),
1
3. 4
3 − i, 4. 15 (9 + i3), 5. 13 (9 + i7).
1
1. (z1 , z2 , z3 , z4 ) = (z1 , z2 , z3 , z4 ).
i i
i i
i i
T1 (z2 ) = T(z2 ) = 1 = 1,
T1 (z3 ) = T(z3 ) = 0 = 0,
T1 (z4 ) = T(z4 ) = ∞ = ∞.
LEMMA 3.2.17 Let C be a circle and z and w be any two points not on C .
If (w, z1 , z2 , z3 ) = (z, z1 , z2 , z3 ) for some three distinct points z1 , z2 , and z3
i i
i i
i i
By previous lemma, the symmetry of the points does not depend on the choice
of the three points on C .
which imply geometrically that z and w lie on different half planes determined
by the line C , and they are at equal distance from an arbitrary point on C .
i i
i i
i i
z
w
C
Geometrically, one of the two points z and w lies inside the circle and the
other point lies outside the circle satisfying |z − a||w − a| = r2 . Furthermore,
a, w, and z are collinear, such that both z and w lie on a same ray starting
from a.
w
z
C a
i i
i i
i i
Algorithm 3.2.22 (To find the LFT that maps a, b , and c to p, q , and r ,
respectively.)
Step 1: Let T be the linear fractional transform that maps a, b, and c into
(z − b) (a − c)
1, 0, and ∞, respectively. That is, T(z) = .
(z − c) (a − b)
Step 2: Let S be the linear fractional transform that maps p, q, and r into
(w − q) (p − r)
1, 0, and ∞, respectively. That is, S(w) = .
(w − r) (p − q)
Example 3.2.23 Find the linear fractional transform that maps 1, 2, and 3 to
0, i, and − i, respectively.
Solution:
The linear fractional transform that maps 1, 2, and 3 into 1, 0, and ∞,
(z − 2) (1 − 3)
respectively, is given by , and the linear fractional trans-
(z − 3) (1 − 2)
form that maps 0, i, and − i into 1, 0, and ∞, respectively, is given by
(w − i) (0 + i)
.
(w + i) (0 − i)
(w − i) (0 + i) (z − 2) (1 − 3)
=
(w + i) (0 − i) (z − 3) (1 − 2)
(w − i) (z − 2)
⇒ − =2
(w + i) (z − 3)
⇒ −(w − i)(z − 3) = 2(z − 2)(w + i)
⇒ −w(z − 3) − 2w(z − 2) = i2(z − 2) − i(z − 3)
⇒ w(−3z + 7) = i(z − 1)
i(z − 1)
⇒ w= .
−3z + 7
i i
i i
i i
i(z − 1)
Hence, R(z) = , ∀z ∈ C∞ , is the required linear fractional
−3z + 7
transform.
Exercise 3.2.24 Find the linear fractional transform that maps a, b, and c to
p, q, and r, respectively, in the following:
1
6. 1, ∞, and − 2 to −1, 2, and
2
7. −i, 0 and ∞ to ∞, −i2, and 1
3 + i4 −1 + i 2 + i3
10. 1 + i, 1, and i to , , and .
5 2 2
Answers:
(1 + i2)z + 1 (i − 1)z + (3 + i)
1. , 2. ,
z + (1 + i2) (1 + i3)z + (1 − i)
−iz − 1 z−1
3. , 4. ,
−iz + 1 z+1
1 2z + 1
5. , 6. ,
−z + 1 z−4
z+2 3z + 4
7. , 8. ,
z+i 2z + 1
(2 + i3)z − 1 2z − 3
9. , 10. .
z + (1 − i) z+i
i i
i i
i i
Example 3.2.25 Check the symmetry of given z, and w with respect to the
given C in the following:
Solution:
1. We choose the three points ∞, 0, and 1 + 2i on y = 2x. As
11 − i3 11 + i3
(−4 + i7, ∞, 0, 1 + 2i) = and (8 + i, ∞, 0, 1 + 2i) = ,
10 10
−4 + i7 and 8 + i are symmetric with respect to y = 2x.
2. We choose the three points ∞, i, and 5 + 4i on 3x − 4y + 1 = 0. As
1 + 7i 1 + 7i
(2 + i5, ∞, i, 5 + i4) = and (3, ∞, i, 5 + 4i) = ,
5 10
2 + i5 and 3 are not symmetric with respect to 3x − 4y + 1 = 0.
3. We choose the three points 15 + 3i, −5 + 3i, and 5 + 13i on |z −
7 + 5i
5 − i3| = 10. As (8 + 7i, 15 + i3, −5 + i3, 5 + i13) = and
6
7 − 5i
(17 + i19, 15 + i3, −5 + i3, 5 + i13) = , 8 + 7i and 17 + 19i are
6
symmetric with respect to |z − 5 − i3| = 10.
r(z − a)
Example 3.2.26 Prove that if w = T(z) = , then T maps |z − a| = r
r2 − az
onto |w| = 1 and T(B(0, r)) = B(0, 1).
i i
i i
i i
Using the observation T(a) = 0, it follows that T(B(0, r)) ⊆ B(0, 1). Since,
T −1 is also a linear fractional transform, such that T −1 maps |w| = 1 onto
|z − a| = r and T −1 (0) = a, by a similar argument, we obtain T −1 (B(0, 1)) ⊆
B(0, r). Thus, we get T(B(0, r)) = B(0, 1).
PROBLEM 3.2.27 Find all linear fractional transform that maps |z| = r onto
|w| = ρ .
Case 2: α = 0
In this case, immediately, we have β = ∞ using the symmetry of α
and β with respect to |z| = r. Now, T(0) = 0 and T(∞) = ∞ imply
b = 0 and c = 0, and hence, a 0 d. As T maps |z| = r onto
|w| = ρ , we have ρ = |w| = |T(z)| = da r. Therefore, there exists
ϕ ∈ R such that da = ρr exp(iϕ). Thus, T(z) = ρr exp(iϕ)z.
Case 3: α = ∞.
Then, we have β = 0. Now, T(0) = ∞ and T(∞) = 0 imply that
d = 0 and a = 0. As T maps |z| = r onto |w| = ρ , we have
ρ = |w| = |T(z)| = bc 1r . Therefore, there exists ψ ∈ R such that
b
c = rρ exp(iψ). Thus, T(z) = rρ exp(iψ) 1z .
i i
i i
i i
transforms obtained in Case 3 of the previous problem map |z| < r onto
|w| > ρ .
PROBLEM 3.2.29 Find all linear fractional transform that maps the upper
half-plane onto the unit circle.
exp(x + y) = and
exp(x) exp(y)
exp(x)
exp(x − y) = exp(x) exp(−y) = .
exp(y)
Therefore,
i i
i i
i i
and
a
exp(x)
ln a − ln b = x − y = ln(exp(x − y)) = ln = ln .
exp(y) b
1
RESULT 3.3.3 ln : (0, ∞) → R is differentiable and ln a = .
a
|a − b| = | exp(x) − exp(y)|
∞
xn − y
n
=
n!
n=1
∞ xk yn−1−k
n−1
k=0
= |x − y|
n!
n=1
n−1
∞ xk yn−1−k
k=0
= |x − y|
n!
n=1
≥ |x − y|
= | ln a − ln b|.
ln b − ln a y−x
lim = lim
b→a b−a y→x exp(y) − exp(x)
i i
i i
i i
1
= lim
y→x exp(y) − exp(x)
y−x
1
=
exp(y) − exp(x)
lim
y→x y−x
1 1 1
=
= = .
exp (x) exp(x) a
1
Thus, ln is differentiable on (0, ∞) and ln a =
.
a
We would like to define complex logarithm as the inverse function of
exp on C. However, unfortunately, exp : C → C is neither one-to-one nor
onto. However, we can define the complex logarithm of a non-zero complex
number as a multivalued function as follows.
Note that in the above definition, the imaginary part of log is not single
valued.
Proof: Using the continuity of arctan, we get that arg is continuous on each
of the following regions.
1. {(a, b) : a > 0, b ∈ R}
2. {(a, b) : a < 0, b > 0}
3. {(a, b) : a < 0, b < 0}
Hence to conclude the proof of this lemma, we shall show that arg is
continuous on the positive imaginary axis and on the negative imaginary axis.
i i
i i
i i
Case 1: an > 0, ∀n ∈ N
bn bn π
As → +∞ as n → ∞, we have arctan → as n → ∞.
an an 2
Therefore,
π bn
arg ((0, b)) = = lim arctan = lim arg ((an , bn )).
2 n→∞ an n→∞
Case 2: an < 0, ∀n ∈ N
bn π
Therefore, using the fact that arctan → as n → ∞, we
|an | 2
obtain
π π bn
arg ((0, b)) = = π − = π − lim arctan
2 2 n→∞ |an |
= lim arg ((an , bn )).
n→∞
Case 3: an ∈ R, ∀n ∈ N.
Given > 0 using Case 1 and Case 2, we find N1 , N2 ∈ N such that
N1 > M , N2 > M , and
π
|arg ((an , bn )) − arg ((0, b))| = arg ((an , bn )) − < ,
2
∀n ≥ N1 if an > 0
π
|arg ((an , bn )) − arg ((0, b))| = arg ((an , bn )) − <
2
∀n ≥ N2 if an < 0.
As
π π
|arg ((an , bn )) − arg ((0, b))| = − =0< ,
2 2
∀n ∈ N if an = 0,
π
it follows that arg ((an , bn )) − < , ∀n ≥ max{N1 , N2 } if
2
an ∈ R.
i i
i i
i i
Remark 3.3.7: The principal branch of arg is not continuous at any point of
negative real axis.
Indeed, if (x, 0) ∈ C with x < 0, we have
1 1
x + , 0 → (x, 0) and x − , 0 → (x, 0) as n → ∞
n n
but
1 π 1 π
arg x + , 0 → and arg x − , 0 → − as n → ∞.
n 2 n 2
log(w) − log(w0 ) z − z0
lim = lim ,
w→w0 w − w0 z→z0 exp(z) − exp(z0 )
1
= lim
z→z0 exp(z) − exp(z0 )
z − z0
1
=
exp(z) − exp(z0 )
lim
z→z0 z − z0
i i
i i
i i
1 1 1
=
= = .
exp (z0 ) exp(z0 ) z0
1
Thus, log is differentiable on and log (a) = , ∀a ∈ .
a
√
Definition 3.3.10 For a non-zero complex number z, we define w = z by
√ √ θ
the solution of the equation w2 = z. More explicitly, z = ± r exp i ,
2
where z = r exp(iθ ).
√
RESULT 3.3.11 (Analytic branch for√ ·) √
If = C \ {(x, 0) : x ≤ 0} and √ if z is defined by the unique value of z
whose real part is positive, then · : → C is analytic.
|z − z0 | = |w2 − w20 | = |w − w0 | · |w + w0 | ≥ |w − w0 | · Re (w + w0 )
> |w − w0 | · Re w0
√ 1
we get · is differentiable and its derivative is √.
2 ·
√
Example 3.3.12 Define an analytic branch for 1 + z and justify that it is
analytic. √
We know that z is defined on√ 1 = C \ {(x, 0) : x ≤ 0} as an analytic
function by the unique value of z, whose real part is positive. As the range
i i
i i
i i
exp(iz) + exp(−iz)
Proof: Let z = arccos w. Then w = cos z = . Hence, it
2
follows that
exp(iz) − exp(−iz)
Proof: Let z = arcsin(w). Then w = sin(z) = . Hence, it
i2
follows that
i i
i i
i i
exp(z) − exp(−z)
If z = tanh−1 w, then w = tanh z = ,
exp(z) − exp(−z)
i i
i i
i i
i i
i i
i i
( f ◦ ϕ)(t) − ( f ◦ ϕ)(t0 )
lim arg
t→t0 ϕ(t) − ϕ(t0 )
f (z) − f (z0 )
lim 0.
z→z0 z − z0
Let γ be any curve with a parametric equation ϕ(t), t ∈ [a, b] such that
ϕ(t0 ) = z0 and ϕ (t0 ) 0. By chain rule, we have
and as f (z0 ) 0, then arg is continuous on a region that contains f (z0 ) with
respect to a suitable branch (Lemma 3.3.8), and hence, we get
( f ◦ ϕ)(t) − ( f ◦ ϕ)(t0 )
lim arg
t→t0 ϕ(t) − ϕ(t0 )
( f ◦ ϕ)(t) − ( f ◦ ϕ)(t0 )
= arg lim
t→t0 ϕ(t) − ϕ(t0 )
( f ◦ ϕ)(t) − ( f ◦ ϕ)(t0 ) t − t0
= arg lim ·
t→t0 t − t0 ϕ(t) − ϕ(t0 )
( f ◦ ϕ)(t) − ( f ◦ ϕ)(t0 ) t − t0
= arg lim ·
t→t0 t − t0 ϕ(t) − ϕ(t0 )
( f ◦ ϕ) (t0 )
= arg
ϕ (t0 )
= arg f (z0 )
i i
i i
i i
which is independent of the curve γ . Again using the hypothesis and the
continuity of the modulus function (Example 1.6.18), we get
| f (z) − f (z0 )|
lim 0.
z→z0 |z − z0 |
Thus, f is conformal at z0 .
zn → z0 as n → ∞ with zn z0 , ∀n ∈ N
then
f (zn ) − f (z0 )
→ A exp(iθ ) as n → ∞.
z n − z0
First, we choose a curve γ with a parametric equation ϕ(t), t ∈ [a, b] with
the following properties:
1. ϕ(t0 ) = z0 , for some t0 ∈ [a, b] and ϕ(tn ) = zn , ∀n ∈ N for some
sequence (tn ) of distinct points of [a, b] such that tn → t0 as n → ∞.
2. ϕ (t0 ) 0.
Now, using the assumptions and Lemma 3.3.8,
f (zn ) − f (z0 )
lim
n→∞ z n − z0
i i
i i
i i
Proof: Let γ be a curve with a parametric equation ϕ(t), t ∈ [a, b] such that
ϕ(t0 ) = z0 and ϕ (t0 ) 0. If ψ(t) = f (ϕ(t)), ∀t ∈ [a, b] and if x (t0 ) and y (t0 )
are the real and imaginary parts of ϕ (t0 ), respectively, then ψ is differentiable
at t0 and
ψ (t0 ) = fx (ϕ(t0 ))x (t0 ) + fy (ϕ(t0 ))y (t0 )
ϕ (t0 ) + ϕ (t0 ) ϕ (t0 ) − ϕ (t0 )
= fx (z0 ) + fy (z0 )
2 i2
fx (z0 ) − ify (z0 ) fx (z0 ) + ify (z0 )
= ϕ (t0 ) + ϕ (t0 ) .
2 2
Thus, we have
ψ (t0 ) fx (z0 ) − ify (z0 ) fx (z0 ) + ify (z0 ) ϕ (t0 )
= + . (3.3)
ϕ (t0 ) 2 2 ϕ (t0 )
fx (z0 ) − ify (z0 )
which represents an equation of a circle with centre and radius
2
fx (z0 ) + ify (z0 )
, as ϕ varies through the continuous function on [a, b] with
2
ϕ(t0 ) = z0 for some t0 and ϕ (t0 ) 0.
1. Assume that f is conformal of first type at z0 .
Then, we have
ψ (t0 ) ( f ◦ ϕ) (t0 ) ( f ◦ ϕ)(t) − ( f ◦ ϕ)(t0 )
arg = arg = lim arg
ϕ (t0 ) ϕ(t0 ) t→t0 ϕ(t) − ϕ(t0 )
i i
i i
i i
Example 3.4.6 The function exp maps the rectangle {z ∈ C : 1 < Re z <
π
2, < Im z < π3 } onto the part of angular sector {w ∈ C : a < |w| <
6
b, π6 < arg w < π3 }.
Let z = x + iy. Under the exponential function,
i i
i i
i i
Thus, the image of the rectangle under exp is shown as in the following
diagram.
Exercise 3.4.7
1. Prove that the map f (z) = z2 , ∀z ∈ C maps the first quadrant {(x, y) ∈
C : x > 0, y > 0} onto the upper half-plane {(x, y) ∈ C : y > 0}.
z−1
2. Prove that the map f (z) = , ∀z ∈ C \ {−1} maps the upper
z+1
half-plane {(x, y) ∈ C : y > 0} onto the unit circle {z ∈ C : |z| < 1}.
i i
i i
i i
i i
i i
i i
i i
i i
i i
called sheet-k , ∀k ∈ Z.
• Paste upper edge on the slit on the left side of sheet-k with lower edge
on the slit on the left side of sheet-(k + 1) and lower edge on the slit
on the left side of sheet-k with upper edge on the slit on the left side of
sheet-(k + 1).
• Paste upper edge on the slit on the right side of sheet-k with lower edge
on the slit on the right side of sheet-(k + 1) and lower edge on the slit
on the right side of sheet-k with upper edge on the slit on the right side
of sheet-(k + 1).
• By gluing the line segment {(u, 0) : |u| ≤ 1} in every sheet altogether,
identify it as a single line segment on the surface.
Then, cos becomes a continuous bijection from C onto the surface.
i i
i i
i i
4
Complex Integration
165
i i
i i
i i
b b
2. (αf )(t) dt = α f (t) dt, ∀α ∈ R,
a a
b b
3. f1 (t) dt ≤ f2 (t) dt, whenever f1 ≤ f2 on [a, b],
a a
b b c
4. f (t) dt = f (t) dt + f (t) dt, provided a < c < b.
a c a
b b
where u(t) dt and v(t) dt are Riemann integrals of u and v, respectively, on
a a
[a, b].
( f + g)(t) dt
a
b b
i i
i i
i i
=c f (t) dt.
a
i i
i i
i i
b b
Proof: If f (t) dt = 0, then obviously the inequality follows. If f (t) dt 0,
a a
b
then we can write it as f (t) dt = reiθ . Then, using the previous theorem, we
a
get
b
f (t) dt = r
a
b
−iθ
= e f (t) dt
a
b
= e−iθ f (t) dt
a
⎛ b
⎞
= Re (e−iθ f (t)) dt
a
b
−iθ
≤ e f (t) dt
a
b
= | f (t)| dt.
a
RESULT 4.1.9 If f : [a, b] → C is Riemann integrable and | f (t)| ≤ M ,
b
∀t ∈ [a, b], then | f (t)| dt ≤ M(b − a).
a
i i
i i
i i
Proof: Let ε > 0 be given. Since fn → f uniformly on [a, b], there exists
N ∈ N such that | fn (x) − f (x)| < /(b − a) ∀ n ≥ N and ∀ x ∈ [a, b].
Now for n ≥ N ,
b
b b b
fn (x)dx − f (x)dx ≤ | fn (x) − f (x)|dx < dx = .
b−a
a a a a
b b
Therefore, fn (x) dx → fdx as n → ∞.
a a
t
THEOREM 4.1.11 If f : [a, b] → C is continuous and g(t) = f (τ ) dτ , ∀t ∈
a
[a, b], then g is differentiable on [a, b] and g = f.
i i
i i
i i
Proof: First, we prove this theorem for the case that f is a real-valued
Riemann integrable function on [a, b]. By Theorem 4.1.2, given > 0,
there exists a partition P = {t0 , t1 , t2 , . . . , tn } of [a, b] such that U(P, f ) −
L(P, f ) < . By mean value theorem, there exists sk ∈ (tk−1 , tk ) such
that
f (tk ) − f (tk−1 ) = f (sk )(tk − tk−1 ), ∀k = 1, 2, . . . , n. (4.1)
n
L(P, f ) = mk (tk − tk−1 )
k=1
n
≤ f (sk )(tk − tk−1 )
k=1
n
≤ Mk (tk − tk−1 )
k=1
= U(P, f ).
b
n
f (t) dt − f (sk )(tk − tk−1 ) ≤ U(P, f ) − L(P, f ).
k=1
a
i i
i i
i i
b
Since > 0 is arbitrary, we get f (t) dt = f (b) − f (a).
a
If f is a complex-valued function, let f = u + iv, where u and v are real-
valued Riemann integrable functions. Hence, using this theorem for u and v,
we obtain
b b b
f (t) dt = u (t) dt + i v (t)) dt
a a a
= (u(b) − u(a)) + i(v(b) − v(a))
= f (b) − f (a).
2
Example 4.1.13 Find [(1 + 9t2 ) + i(4t + 2)] dt.
0
2 2 2
i i
i i
i i
= (ϕ(t) + 1)(1 + i) dt
0
1
= ((1 + t) + it)(1 + i) dt
0
1
i i
i i
i i
= (1 + i(2t + 1)) dt
0
1 1
= dt + i (2t + 1) dt
0 0
= 1 + i2.
ϕ (t) = 2 + i2t
f (ϕ(t)) = (2t + i(t2 + 1))2
= (4t2 − (t2 + 1)2 ) + i2t(t2 + 1)
= (−t4 + 2t2 − 1) + i(2t3 + 2t)
f (ϕ(t))ϕ (t) = 2(−t4 + 2t2 − 1) − 2t(2t3 + 2t)
+ i(4t3 + 4t − 2t5 + 4t3 − 2t)
= (−6t4 − 2) + i(−2t5 + 8t3 + 2t).
1
= (−6t − 2) dt + i
4
(−2t5 + 8t3 + 2t) dt
0 0
6 1
= − −2 +i − +2+1
5 3
16 8
= − +i .
5 3
Example 4.1.20 Find (z + 3) dz, where γ is the circular arc given by ϕ(t) =
γ
i + 2eit , t ∈ [0, π ].
i i
i i
i i
Solution:
π
Exercise 4.1.21
1. Evaluate f (z) dz, where f and the parametric equation ‘ϕ(t),
γ
t ∈ [a, b]’ of γ are given as follows:
i i
i i
i i
Definition 4.1.24 If a curve γ is given by ϕ(t), t ∈ [a, b], then the opposite
curve −γ is defined by ψ(t) = ϕ(a + b − t), t ∈ [a, b].
g −g
LEMMA
4.1.25 If f is continuous on and γ is a curve in , then
f (z) dz = − f (z) dz.
−γ γ
i i
i i
i i
= − f (z) dz.
γ
Proof: For each fixed t ∈ [a, b], let γa,t be the part of the curve γ with
parametric equation z(τ ) = u(τ ) + iv(τ ), τ ∈ [a, t], where u and v are
real-valued differentiable function on [a, b]. Define S : [a, b] → R by
S(t) = L(γa,t ), ∀t ∈ [a, b]. We shall show that S is a differentiable function
and its S (t) = |z (t)|, ∀t ∈ [a, b]. Therefore, using the uniform continuity of
z , given > 0, there exists δ > 0 such that
|τ − t| < δ ⇒ |z (τ ) − z (t)| < √ (4.2)
2
⇒ |u (τ ) − u (t)| < √ and |v (τ ) − v (t)| < √ .
2 2
Therefore, for |τ − t| < δ and |σ − t| < δ , we have
|(u (τ ), v (σ ))| − |z (t)| = |(u (τ ), v (σ ))| − |(u (t), v (t))|
≤ |(u (τ ), v (σ )) − (u (t), v (t))|
= (u (τ ) − u (t))2 + (v (σ ) − v (t))2
2 2
< +
2 2
= .
i i
i i
i i
|z (t)| − < |(u (τ ), v (σ ))| < |z (t)| + , ∀σ , τ ∈ (t − δ, t + δ). (4.3)
Let 0 < s < δ and let P = {t0 , t1 , . . . , tn } be an arbitrary partition of [t, t + s].
Then, applying mean-value theorem (Theorem 2.1.18) for u and v, we get
n n
|z(tk ) − z(tk−1 )| = (u(tk ) − u(tk−1 ))2 + (v(tk ) − v(tk−1 ))2
k=1 k=1
n
= (tk − tk−1 )|(u (xk ), v (yk ))|, (4.4)
k=1
where P varies over all partitions of [t, t + s]. Therefore, if γt, t+s is the part of
gamma, whose parametric equation is z(τ ), τ ∈ [t, t + s], then it follows that
i i
i i
i i
b
In particular, L(γ ) = S(b) = S (τ ) dτ .
a
= h (t) dt
a
= h(b) − h(a)
= F(ϕ(b)) − F(ϕ(a)).
Hence, the lemma follows.
i i
i i
i i
RESULT
4.1.30 If f is a continuous function on a region such that
f (z) dz = 0, for every closed curve γ in , then f is the derivative of an
γ
analytic function F on .
3. Pz0 ⊂ .
(x, y0)
(x0, y0)
(u, u)
i i
i i
i i
i i
i i
i i
h
1
< |ds|
|h|
0
|h|
= = ,
|h|
as |(x0 + s, y0 ) − (x0 , y0 )| = |t| < |h| < δ and by using equation (4.5).
Thus,
our claim is proved. Next for every z0 ∈ , we define G(z0 ) = f (z) dz,
Qz0
where Qz0 is a polygon joining w and z0 such that
1. the line segments of Qz0 are parallel to the coordinate axes.
2. the line incident with z0 is vertical and completely contained in B(z0 , δ).
3. Qz0 ⊆ .
Next we claim that ∂G
∂y (x0 , y0 ) = if (x0 , y0 ). We fix > 0, δ > 0 and h ∈ R as
before. As the parametric equation of L(x0 , v),(x0 , y0 ) is
ϕ(t) = x0 + it = (x0 , t), t ∈ [v, y0 ],
by a similar argument, we get
G(x0 , y0 + h) − G(x0 , y0 )
− if (x0 , y0 )
h
h
1
= f (x0 , y0 + t) idt − if (x0 , y0 )
h
0
h
1
≤ | f (x0 , y0 + t) − f (x0 , y0 )| |dt| < .
|h|
0
Since
i i
i i
i i
b
1 dz 1 ϕ (t) dt
= .
i2π z − z0 i2π ϕ(t) − z0
γ a
s ϕ (t) dt
If F(s) = ϕ(t)−z0 , ∀s ∈ [a, b], then F is a differentiable function on
a
ϕ (s)
[a, b] and F (s)
= ϕ(s)−z 0
, ∀s ∈ [a, b], by Theorem 4.1.11. To conclude
the theorem, we shall show that F(b) is an integral multiple of i2π . If
ψ(s) = exp(−F(s))(ϕ(s) − z0 ), ∀s ∈ [a, b], then
exp(−F(a))(ϕ(a) − z0 ) = exp(−F(b))(ϕ(b) − z0 ),
i i
i i
i i
S3
R2 S4
S1
S2
R1
g Γ
Remark 4.2.4: For a given closed curve, among the regions determined by
γ , there must be only one unbounded region determined by γ . That region is
treated as the region containing ∞ in C∞ .
In the above diagrams, R2 and S4 are the unbounded regions determined by
γ and , respectively.
1. WN (−γ , z0 ) = −WN(γ , z0 ),
Proof: Let z0 γ .
1 dz 1 dz
WN (−γ , z0 ) = =− = −WN(γ , z0 ).
i2π z − z0 i2π z − z0
−γ γ
2. Case 1. First, we show that if the line segment Lz0 ,w0 joining z0 and
w0 is completely contained
in a same region determined by γ . We
z − z0 1 1
know that log = − , provided the expression
z − w0 z − z0 z − w0
z − z0
belongs to the domain of log, where it is analytic. From Result
z − w0
i i
i i
i i
Case 2. Next, let z0 and w0 be two points in the same region determined
n
by γ . Then, we can join z0 and w0 by a polygon ∪ Laj−1 ,aj contained in
j=1
the same region determined by γ , where a0 = z0 , an = w0 , and Laj−1 ,aj
is the line segment joining ai−1 and ai . Applying case 1, repeatedly, we
get
WN (γ , z0 ) = WN (γ , a0 )= WN(γ , a1 ) = · · · = WN(γ , an )
= WN (γ , w0 ).
1 dz
WN (γ , z0 ) = WN(γ , w0 ) = = 0.
i2π z − w0
γ
1 Though Theorem 4.2.14 is proved later, and its proof does not depend on the present
theorem.
i i
i i
i i
From the above properties. WN(γ , z0 ) is called the winding number of γ with
respect to z0 .
1 if |z − a| < r
Example 4.2.6 If C is |ζ − a| = r, then WN (C, z) = .
0 if |z − a| > r
Since the parametric equation of the circle |ζ − a| = r is given by ζ =
a + r exp(iθ ), θ ∈ [0, 2π], we get
2π
1 dζ 1 ir exp(iθ )dθ
= = 1.
i2π ζ −a i2π r exp(iθ )
|ζ −a|=r 0
If |z − a| < r, then a and z are lying in the same region determined by C , and
using Theorem 4.2.5(2), we get WN(C, z) = WN(C, a) = 1. If |z − a| > r,
then by using Theorem 4.2.5(3), we get WN(C, z) = 0.
There is an interesting theorem, namely, Jordan curve theorem, which
states that the complement of a simple closed curve has exactly to regions.
Although it can be realized geometrically, its proof is too lengthy. Hence, we
prefer to omit this theorem.
i i
i i
i i
sd
Ω B
i i
i i
i i
i i
i i
i i
4
4
I(R) = I(Sk ) ⇒ |I(R)| ≤ |I(Sk )|.
k=1 k=1
This implies that there exists at least one k ∈ {1, 2, 3, 4} such that |I(Sk )| ≥
1
|I(R)|. Denote that rectangle Sk by R1 . Next, we subdivide R1 into four
4
equal subrectangles of same size as before and choose the one say R2
1
such that |I(R2 )| ≥ |I(R1 )|. Proceeding further, we get a sequence of
4
rectangles R ⊇ R1 ⊇ R2 ⊇ · · · such that
1 1
|I(Rn+1 )| ≥ |I(Rn )| ⇒ |I(Rn )| ≥ n |I(R)|. (4.7)
4 4
1 1
L(Rn+1 ) = L(Rn ) ⇒ L(Rn ) = n L(R). (4.8)
2 2
1 1
D(Rn+1 ) = D(Rn ) ⇒ D(Rn ) = n D(R). (4.9)
2 2
S3 = R1
S4
R3
R2
S1 S2
i i
i i
i i
Therefore, we have
0 < |z − z0 | < δ ⇒ f (z) − f (z0 ) − (z − z0 )f (z0 ) < |z − z0 |.
D(R)L(R)
Using equations (4.9) and (4.10), we choose m ∈ N such that D(Rm ) < δ , and
hence, Rm ⊆ B(z0 , δ). If f1 (z) = 1 and f2 (z) = z, ∀z ∈ C, then f1 and f2 are
the derivatives of the analytic functions F1 and F2 defined by F1 (z) = z and
z2
F2 (z) = , ∀z ∈ C, respectively. Therefore, by Corollary 4.1.29, we have
2
dz = 0 and z dz = 0. Now,
∂Rm ∂Rm
i i
i i
i i
i i
i i
i i
Therefore,
m
m
f (z) dz ≤ f (z) dz < = .
k=1
k=1
m
∂R ∂Sk
Thus, f (z) dz = 0.
∂R
Proof: First we show that there exists an analytic function on such that
= f . This part of this theorem is almost similar to that of the proof of
Result 4.1.30. Let (a, b) ∈ \F be arbitrarily fixed. For every (x0 , y0 ) ∈ \F ,
let γ(x0 ,y0 ) ⊆ \F be a polygon joining (a, b) and (x0 , y0 ) consisting of the
line segments parallel to the coordinate axes and the line segment incident
with (x0 , y0 ), which is horizontal (such a polygon exists by Theorem 1.4.31).
Furthermore, as F is a finite set, we can choose thepolygon such that it does
not pass through any a ∈ F . Define (x0 , y0 ) = f (z) dz. First, we show
γ(x0 ,y0 )
that the definition of (x0 , y0 ) is independent of the choice of the polygon
γ(x0 ,y0 ) .
(a,b) (x0,y0)
If σ(x0 ,y0 ) is another polygon joining (a, b) and (x0 , y0 ), then γ(x0 ,y0 ) ∪
(−σ(x0 ,y0 ) ) is a finite union of boundary of rectangles. As is simply con-
nected, all rectangles enclosed by γ(x0 ,y0 ) ∪ (−σ(x0 ,y0 ) ) are contained in .
2 Note that empty set is a finite set, and hence, this theorem is also true for the case F = ∅.
i i
i i
i i
Let h ∈ R with 0 < |h| < δ . Then, the line segment L(x0 ,y0 ),(x0 +h,y0 ) ⊆ and
is horizontal. Therefore, for |h| < δ , we have
and hence,
(x0 + h, y0 ) − (x0 , y0 )
− f (x , y )
h
0 0
1
= f (z) dz − f (x0 , y0 )
h
L(x ,y ),(x +h,y )
0 0 0 0
h
1
= f (x0 + t, y0 ) dt − f (x0 , y0 )
h
0
h
1
≤ | f (x0 + t, y0 ) − f (x0 , y0 )| |dt|
|h|
0
1
< |h| = .
|h|
∂
Thus, (x0 , y0 ) = f (x0 , y0 ).
∂x
∂
Similarly, we can prove that (x0 , y0 ) = if (x0 , y0 ). Furthermore, we get
∂y
∂ ∂ ∂ ∂
(x0 , y0 ) = −i (x0 , y0 ) = f (x0 , y0 ) (the C–R equation), and ,
∂x ∂y ∂x ∂y
are continuous at (x0 , y0 ). Therefore, by Theorem 2.2.28, we get that is
i i
i i
i i
f (z) − f (z0 )
Proof: Let F(z) = , ∀z ∈ \{z0 }. Then, F is analytic on \{z0 }.
z − z0
Since
f (z) − f (z0 )
lim F(z) = lim = f (z0 ),
z→z0 z→z0 z − z0
i i
i i
i i
Therefore,
by Cauchy’s theorem for simply connected region, we get
F(z) dz = 0. Hence, from the definition of WN(γ , z0 ), it follows that
γ
then lim Gm,n (z) = Gm,n (z0 ). We choose δ > 0 such that B(z0 , δ) ∩ γ = ∅.
z→z0
δ δ
Then, we have |ζ − z| ≥ and |ζ − z0 | ≥ δ, ∀z ∈ B z0 , and ∀ζ ∈ γ . If
2 2
M = |φ(ζ )| |dζ |, then M < ∞ and
γ
i i
i i
i i
m−1
(by using am − bm = (a − b) am−1−k bk )
k=0
|z − z0 |
m−1
|φ(ζ )| |dζ |
≤
δn |ζ − z|m−k |ζ − z0 |k+1
k=0 γ
|z − z0 |
m−1
2m−k
≤ |φ(ζ )| |dζ |
δn δ m−k δ k+1
k=0 γ
m−1
M 2m−k
k=0
≤ |z − z0 | → 0 as z → z0 .
δ m+n+1
For z ∈ B z0 , 2δ , we have
Fn (z) − Fn (z0 )
z − z0
1 1 1
= − φ(ζ ) dζ
z − z0 (ζ − z)n (ζ − z0 )n
γ
n−1
1 1 1 1 1
= − φ(ζ ) dζ
z − z0 (ζ − z) (ζ − z0 ) (ζ − z)n−1−k (ζ − z0 )k
γ k=0
n−1
1 z − z0 1 1
= φ(ζ ) dζ
z − z0 (ζ − z)(ζ − z0 ) (ζ − z)n−1−k (ζ − z0 )k
γ k=0
n−1
1 1
= φ(ζ ) dζ
(ζ − z)n−k (ζ − z0 )k+1
k=0 γ
n−1
n−1
n−1
φ(ζ )
= Gn−k,k+1 (z) → Gn−k,k+1 (z0 ) = dζ
(ζ − z0 )n+1
k=0 k=0 k=0 γ
= nFn+1 (z0 ), as z → z0 .
i i
i i
i i
n! f (z)
WN (γ , z0 )f
(n)
(z0 ) = dz,
i2π (z − z0 )n+1
γ
for every closed curve not passing through z0 in , and for every n ∈ N.
1 f (ζ )
WN (γ , z0 )f (z0 ) = dζ ,
i2π ζ − z0
γ
..
.
⎛ ⎞
dn ⎝ 1 f (ζ ) n! f (ζ )
dζ ⎠ = dζ .
dzn i2π (ζ − z) i2π (ζ − z)n+1
γ γ
f (z)
Therefore, f (n) (z0 ) = n!
dz, ∀n ∈ N.
i2π
γ (z − z0 )n+1
i i
i i
i i
and
1 f (ζ )
f (w) = dζ , ∀w ∈ B(z, r),
i2π (ζ − w)2
|ζ |=r
r
For all w ∈ Cl B z, , and for all n ≥ N ,
2
| fn (ζ ) − f (ζ )|
f (w) − f (w) ≤ 1 |dζ | ≤
1
|dζ | =
4
.
n
2π |ζ − w| 2 2π r2 r
|ζ |=r |ζ |=r 4
Therefore, for every z ∈ , there exists ρz > 0 such that fn → f uniformly
on Cl B(z, ρz ). Let K be a compact subset of . Then, {B(z, ρz ) : z ∈ K} is
i i
i i
i i
f (z) = (z − z0 )m φ(z),
i i
i i
i i
1
(d) Split (z−z1 )m1 (z−z2 )m2 ···(z−zn )mn into partial fractions as follows.
m1
A1,k 2 m
A2,k n
An,k
m
+ + · · · + .
(z − z1 ) k (z − z2 ) k (z − zn )k
k=1 k=1 k=1
(f) Doing Step 3(b) repeatedly, we can find the given integral.
z2 + 1 exp(z)(3z2 + 4)
(2) dz. (7) dz.
|z−3|=1 z−1 |z−(1−i)|=3 (z − 1)3
z4 − 4z − 6 z2 + 5z + 6
(3) dz, where γ is (8) dz.
γ z2 − 6z + 5 |z+1|=2 (z2 + 1)(z − 1 + i)
exp(z)
the square with vertices (3, 3), (9) π
dz.
(−3, 3), (−3, −3), and (3, −3). |z|=4 z z − 2 (z − π )
2z + 3 z+i
(4) dz. (10) dz, where γ
|z|=5 z 2 − 2z − 3
γ (z + 4)(z−1−i)
z2 − 3
(5) dz. is the boundary of [0, 2]×[0, 3].
|z−1|=4 (z2 + 3z + 2)(z + 6)
Solution:
z3 + 1
(1) Let φ(z) = . As z2 + 3z − 10 = 0 ⇒ z = −5, z = 2, the
+ 3z − 10
z2
poles of φ are −5 and 2. We know that 2 is enclosed by |z| = 3, and
−5 is not enclosed by |z| = 3 because |2| = 2 < 3 and | − 5| = 5 > 3.
i i
i i
i i
z3 + 1
We put f (z) = . Therefore,
z+5
z3 + 1 f (z) 23 + 1 18iπ
dz = dz = i2π f (2) = i2π = .
z2 + 3z − 10 z−2 2+5 7
|z|=3 |z|=3
z2 + 1
(2) Let φ(z) = . Then, clearly 1 is the only pole of φ and is not
z−1
enclosed by |z − 3| = 1 (because |1 − 3| = 2 > 1.) Therefore,
z2 + 1
dz = 0.
|z−3|=1 z − 1
z4 − 4z − 6
(3) Let φ(z) = . Since z2 − 6z + 5 = 0 ⇒ z = 5, z = 1,
z2 − 6z + 5
the poles of φ are 5 and 1. Obviously, 1 is enclosed by γ and 5 is
z4 − 4z − 6
not enclosed by γ (check geometrically). We put f (z) = .
z−5
Therefore,
2z + 3
(4) Let φ(z) = . Now, z2 − 2z − 3 = 0 implies that z = −1,
z2 − 2z − 3
and z = 3 are the poles of φ , and both are enclosed by |z| = 5, because
| − 1| = 1 < 5 and |3| = 3 < 5. Therefore, we put f (z) = 2z + 3. Now,
1
we split into partial fractions as follows.
(z + 1)(z − 3)
1 A B
Let = + ⇒ 1 = A(z − 3) + B(z + 1).
(z + 1)(z − 3) z+1 z−3
1
Putting z = 3, we get 4B = 1 ⇒ B = .
4
−1
Putting z = −1, we get −4A = 1 ⇒ A = .
4
1 1 1 1
Hence, = − . Therefore,
(z + 1)(z − 3) 4 z−3 z+1
2z + 3 f (z)
dz = dz
z2 − 2z − 3 (z + 1)(z − 3)
|z|=5 |z|=5
i i
i i
i i
i2π iπ
= ( f (3) − f (−1)) = (9 − 1) = 4iπ .
4 2
z2 − 3
(5) Let φ(z) = . As (z2 + 3z + 2)(z + 6) = 0, it follows
(z2 + 3z + 2)(z + 6)
that z = −1, z = −2, and z = −6 are the poles of φ . Furthermore,
−1 and −2 are enclosed by |z − 1| = 4, and −6 is not enclosed by
|z − 1| = 4 because | − 1 − 1| = 2 < 4, | − 2 − 1| = 3 < 4, and
z2 − 3
| − 6 − 1| = 7 > 4. Hence, we take f (z) = . Now, we split
z+6
1
into partial fractions as follows.
(z + 1)(z + 2)
1 A B
= + ⇒ 1 = A(z + 2) + B(z + 1).
(z + 1)(z + 2) z+1 z+2
Putting z = −2, we get −B = 1 ⇒ B = −1.
Putting z = −1, we get A = 1.
1 1 1
Hence, = − . Therefore,
(z + 1)(z + 2) z+1 z+2
z2 − 3 f (z)
dz = dz
(z2 + 3z + 2)(z + 6) (z + 1)(z + 3)
|z−1|=4 |z−1|=4
f (z) f (z)
= − dz + dz
z+2 z+1
|z−1|=4 |z−1|=4
3z + 1
(6) If φ(z) = , then z = 1 and 2 are the poles of φ , and
(z − 1)(z − 2)2
both are enclosed by |z| = 5. Let f (z) = 3z + 1. Now, we split
i i
i i
i i
1
into partial fractions.
(z − 1)(z − 2)2
1 A B C
= + +
(z − 1)(z − 2)2 z − 1 z − 2 (z − 2)2
exp(z)(3z2 + 4)
(7) Let φ(z) = . Then, 1 is the only pole of order 3 for φ ,
(z − 1)3
which is enclosed by |z − (1 + i)| = 3, as |1 − (1 − i)| = 1 < 3. If
f (z) = exp(z)(3z2 + 4), then
z2 + 5z + 6
(8) Let φ(z) = . Then, i, −i, and 1 − i are the poles
+ 1)(z − 1 + i)
(z2
for φ , among
√ which i and −i√ are enclosed by |z − 1| √ = 2, as
|i + 1| = 2 < 2, | − i + 1| = 2 < 2, and |1 − i + 1| = 5 > 2. Let
z2 + 5z + 6 1
f (z) = . Now, we split into partial fractions
z−1+i (z − i)(z + i)
as follows.
1 A B
= + ⇒ 1 = A(z + i) + B(z − i).
(z − i)(z + i) z−i z+i
i i
i i
i i
1 1
Putting z = i and z = −i we get, respectively, A = and B = − .
i2 i2
Therefore,
z2 + 5z + 6
dz
(z2 + 1)(z − 1 + i)
|z−1|=2
⎛ ⎞
1 ⎜ f (z) f (z) ⎟
= ⎝ dz − dz⎠
i2 z−i z+i
|z−1|=2 |z−1|=2
= π( f (i) − f (−i))
2
i + i5 + 6 i2 − i5 + 6
=π −
i−1+i −i − 1 + i
i5 + 5 5 − i5
=π −
i2 − 1 −1
= π(1 − i3 + 5 − i5) = π (6 − i8).
exp(z)
(9) Let φ(z) = . Then, 0, π , and π/2 are the poles for
z(z − π/2)(z − π)
φ , and all are enclosed by |z − 1| = 2, as |0| < 5, |π/2| < 5, and
1
|π | < 5. Let f (z) = exp(z). Now, we split into
z(z − π/2)(z − π )
partial fractions as follows.
1 A B C
= + + ⇒ 1 =
z(z − π/2)(z − π ) z z − π/2 z−π
A(z − π/2)(z − π) + Bz(z − π ) + Cz(z − π/2).
exp(z)
dz
z(z − π/2)(z − π )
|z|=5
i i
i i
i i
2
= i2π (f (0) − 2f (π/2) + f (π ))
π2
i4
= (1 − 2 exp(π/2) + exp(π )).
π
z+i
(10) Let φ(z) = . Then, −4 and 1 + i are the poles for φ .
(z + 4)(z − 1 − i)
Since the given curve γ is the boundary of the rectangle with vertices
(0, 3), (2, 0), (2, 3), and (0, 3), 1 + i is enclosed by γ , and −4 is not
z+i
enclosed by γ . Therefore, let f (z) = . Hence,
z+4
z+i f (z)
dz = dz
(z + 4)(z − 1 − i) z−1−i
γ γ
1 + i2 π
= i2πf (1+i) = i2π = (−9 + i7).
5+i 13
dz
(2) dz.
|z+i2|=2 z2 +1
sin(z)
(3) dz.
|z|=2 z+i
exp(z)
(4) dz.
|z|=1 (z − 2)3
i i
i i
i i
z2 exp(z)
(5) dz.
|z+ 12 |=1
(z + 1)(z − 2)
z+1
(6) dz.
|z−1|=2 z2 − 9z + 20
z exp(z)
(7) dz.
|z|=3 z−1
z
(8) dz.
|z−(2+i)|=2 (z − (1 + i))
exp(z)
(9) dz.
|z|=4 (z + 1)(z − 3)2
z2 + 1
(10) dz.
|z|=3 (z2 − 1)(z − 2)
z
(11) dz.
|z|=2 (z − 1)2
2z + 3
(12) dz.
|z−i2|=4 (z − (1 − i))2
ez
(13) dz.
|z|=4 (z2 + 1)3
z+i
(14) dz.
|z−1|=4 (z − 1)2 (z − 2)(z − 6)
exp(z + 1)
(15) dz.
|z−2|=3 (z + 6)4
i i
i i
i i
i i
i i
i i
Clearly, every polynomial, exp, sin, and cos are entire functions. Obviously,
every constant function is a bounded entire function but the interesting fact
is that the converse of this statement is also true, which follows.
Proof: Let f be a bounded entire function such that sup | f (z)| ≤ M for some
z∈C
M > 0. For an arbitrary z ∈ C and for an arbitrary r > 0, from Cauchy’s
estimate, we get
Mr
| f (a)| ≤ , where Mr = sup | f (z)|.
r |z|=r
As Mr ≤ M, ∀r > 0, we have
M
| f (a)| ≤ → 0 as r → ∞.
r
Thus, f (z) = 0, ∀z ∈ C. Hence, by Theorem 2.2.19, we get that f is a constant
function.
n−1
n k
≥ an z − ak z
k=0
n
n−1
≥ an z − |z|n−1 |ak |
k=0
i i
i i
i i
n−1
≥ |an z| − |ak |
k=0
1
n−1
≥ 1 if |z| ≥ 1+ |ak | .
|an |
k=0
1
As is a continuous real-valued function on the compact set
P
1
n−1
z ∈ C : |z| ≤ max 1, 1+ |ak | ,
|an |
k=0
1
there exists M > 0 such that ≤ M on the compact set. Therefore,
P(z)
1
≤ M + 1 < ∞ on C,
P
1
and hence, is a bounded entire function. Therefore, by Liouville’s theorem
P
1
(Theorem 4.3.11), we get that is a constant function. This implies that P is
P
a constant function. This is a contradiction. Therefore, P has a zero in C. If
α1 is a zero of P, then using Lemma 3.1.3, we can write
P(z) = (z − α1 )Q1 (z), ∀z ∈ C.
Certainly, degree of Q1 is n − 1. If n − 1 > 0, then Q1 has a zero say α2 , and
hence, we can write
P(z) = (z − α1 )(z − α2 )Q2 (z), ∀z ∈ C.
Proceeding further, at the nth stage, we get
P(z) = (z − α1 )(z − α2 ) · · · (z − αn )Qn (z), ∀z ∈ C,
where α1 , α2 , . . . , αn are the zeroes of P, and Qn is a polynomial of degree 0,
which means that Qn is a constant. Therefore, P has exactly n zeroes in C.
i i
i i
i i
r = d(γ , c
) = inf{|z − w| : z ∈ γ , w ∈ c
},
n
γ = ∪ γk and γk ⊂ B(φ(tk ), r), 1 ≤ k ≤ n.
k=1
Moreover, by the choice of r, we have B(φ(tk ), r), which does not intersect c ,
as φ(tk ) ∈ γ ⊂ . Since φ(tk ), φ(tk−1 ) ∈ γk ⊆ B(φ(tk ), r), and B(φ(tk ), r) is a
convex set,3 we can connect φ(tk ) and φ(tk−1 ) by a polygon λk ⊆ B(φ(tk ), r),
which consists of line segments parallel to coordinate axes. Hence, γk ∪(−λk )
is a closed curve contained in the simply connected region B(φ(tk ), r), ∀1 ≤
k ≤ n.
3 A set C is said to be a convex set if for every pair of points of C , the line segment joining
the two points is contained in C .
i i
i i
i i
f(tk) l
r
f(tk−1)
n
Therefore, if λ = ∪ λk , then
k=1
Now, we subdivide the entire complex plane into finite number of bounded
rectangles Bp , 1 ≤ p ≤ μ, and finite number of unbounded rectangles Uq ,
1 ≤ q ≤ ν , by extending each line segment involved in the construction of λ
into straight lines. Now, for each p ∈ {1, 2, . . . , μ} and q ∈ {1, 2, . . . , ν}, we
μ
fix one point bp ∈ Int Bq and uq ∈ Int Uq . If we denote ∪ WN (λ, bp )∂Bp by
p=1
χ , then we claim that after removing the line segments, which are common
sides of two rectangles in χ , we get χ = λ. First, we note that for 1 ≤ j ≤ μ
and 1 ≤ l ≤ ν , we have
μ
WN (χ , bj ) = WN ∪ WN(λ, bp )∂Bp , bj
p=1
μ
= WN (λ, bp ) WN(∂Bp , bj )
p=1
i i
i i
i i
and
μ
WN (χ , ul ) = WN ∪ WN(λ, bp )∂Bp , ul
p=1
μ
= WN (λ, bp ) WN (∂Bp , ul )
p=1
= 0, since WN(∂Bp , ul ) = 0. (4.14)
Next, we claim that the multiplicity m of any line segment Lp, j in λ ∪ (−χ ),
which is common to two bounded rectangles Bp and Bj , is zero for some
1 ≤ p, j ≤ μ. Since λ ∪ (−χ ) ∪ (−m∂Bp ) does not have the line Lp, j , the
points bp and bj belong to the same region determined by λ∪(−χ )∪(−m∂Bp ).
Therefore,
WN (λ ∪ (−χ ) ∪ (−m∂Bp ), bp ) = WN(λ ∪ (−χ ) ∪ (−m∂Bp ), bj ).
Hence, by using equation(4.13), we get
WN (λ, bp ) − WN (λ, bj ) − m = WN(λ, bp ) − WN(λ, bj ) − 0 ⇒ m = 0.
Similarly, if Lp,q is the common side of a bounded rectangle Bp and an
unbounded rectangle Uq , then arguing as before by applying equation(4.14),
we get the multiplicity of the line segment Lp,q in λ ∪ (−χ ) as zero. Hence,
our claim holds.
By Cauchy’s theorem for rectangle, we have f (z) dz = 0, provided
∂Bp
Bp ⊆ . If Bp , then there exists a ∈ Int Bp and a . Since γ is
homologous to zero in , we get WN(γ , a) = 0. As γ ∪(−λ) does not enclose
a, we get
WN (γ ∪ (−λ), a) = 0 ⇒ WN(λ, a) = WN(γ , a) = 0.
Since a and bp belong to the same region determined by λ, we get
WN (λ, bp ) = WN(λ, a) = 0.
This implies that for each p = 1, 2, . . . , μ,
i i
i i
i i
μ
Hence, it follows that f (z) dz = f (z) dz = WN (λ, bp ) f (z) dz = 0.
λ χ p=1 ∂Bp
m
Definition 4.4.2 A cycle is a finite union of closed curves ∪ γj .
j=1
m
m
We define f (z) dz = f (z) dz, so that we get WN(, a) = WN (γj , a).
j=1 γj j=1
Then, the general version of Cauchy’s theorem can be further extended to the
following form easily.
1 f (z) n
dz = WN (γ , zj ).
i2π f (z)
γ j=1
i i
i i
i i
1 f (z) 1
n
dz 1 g (z) n
dz = + dz = WN (γ , zj ),
i2π f (z) i2π z − zj i2π g(z)
γ j=1 γ γ j=1
g (z)
since we have dz = 0, by general version of Cauchy’s theorem.
γ g(z)
Note that if we take γ as a simple closed curve in the above theorem, then
1 f (z)
dz counts the number of zeroes of f inside γ .
i2π γ f (z)
Proof: Since f is non-constant, gw0 is not identically zero, and hence, by the
principle of analytic continuation, it follows that z0 is isolated. Therefore,
there exists r > 0 such that
1. Cl B(z0 , r) ⊆ .
2. gw0 0 on B(z0 , r)\{z0 }, since z0 is an isolated zero of gw0 .
3. f 0 on B(z0 , r)\{z0 }. (This is possible because of the following rea-
son. Since f is non-constant, f is not identically 0. If f (z0 ) = 0, then
z0 must be an isolated zero of f . Then, there exists r > 0. If f (z0 ) 0,
using the continuity of f , we can find such an r > 0.)
Let γ be the circle |z − z0 | = r. Since z0 γ , we have w0 = f (z0 ) f (γ ).
Hence, we can choose ρ > 0 such that B(w0 , ρ) ∩ f (γ ) = ∅. Then, for each
w ∈ B(w0 , ρ), we get that w and w0 lie in a same region determined by f (γ ).
Therefore, from a property of winding number (Thereom 4.2.5), it follows
that
WN ( f (γ ), w) = WN ( f (γ ), w0 ). (4.15)
i i
i i
i i
the only one zero of gw0 inside γ is z0 and its multiplicity is n, we have
Since
WN (γ , zj ) = n. Therefore, from equation (4.15), we get
j∈Jw0
n= WN (γ , zj ) = WN( f (γ ), w0 ) = WN( f (γ ), w) = WN (γ , zj ).
j∈Jw0 j∈Jw
i i
i i
i i
g(w) − g(w0 ) z − z0 1 1
lim = lim = lim = .
w→w0 w − w0 z→z0 f (z) − f (z0 ) z→z0 f (z) − f (z0 ) f (z0 )
z − z0
Proof: To prove this theorem, we show that for each z0 ∈ , there exists
z1 ∈ such that | f (z1 )| > | f (z0 )|. Since f is a non-constant analytic function
on , then by open mapping theorem, f is an open map. Therefore, f ( ) itself
is an open subset of C. If w0 = f (z0 ), then w0 ∈ f ( ), and hence, there exists
> 0 such that B(w0 , ) ⊆ f ( ). If |w0 | < r < and if
r exp(iarg w0 ) if w0 0
w1 = ,
r if w0 = 0
then w1 ∈ B(w0 , ) ⊆ f ( ) and |w1 | > |w0 |. Thus, there exists z1 ∈ such
that f (z1 ) = w1 , and hence, | f (z1 )| = |w1 | > |w0 | = | f (z0 )|.
i i
i i
i i
1
Proof: Since f is nowhere zero on , is also an analytic function on .
f
1
Hence, by maximum modulus principle, has no maximum in . That is,
f
1 1
for every z ∈ , there exists z1 ∈ such that < , equivalently,
f (z) f (z1 )
|f (z1 )| < |f (z)|. Thus, |f | has no minimum on .
f (z)
Proof: If g(z) = for z 0, then g is analytic on B(0, 1)\{0}, and 0 is a
z
removable singularity for g since
f (z) − f (0)
lim g(z) = lim = f (0)
z→0 z→0 z−0
exists in C. Hence, if we define g(0) = f (0), then g becomes analytic on
B(0, 1), by Lemma 5.1.2. This implies that g is analytic on Cl B(0, r) for every
0 < r < 1. Since |g| is a continuous function on the compact set Cl B(0, r),
there exists zr with |zr | ≤ 1 such that |g(zr )| = sup |g(z)|. If g is a
z∈Cl B(0,r)
constant function on B(0, 1), then we can choose this zr with |zr | = 1, and if
g is non-constant, then by maximum modulus principle, |g| does not attain its
maximum on |z| < r, and hence, |zr | = r. Since 0 < |zr | < 1, we have
f (zr ) | f (zr )| 1
sup |g(z)| = |g(zr )| = = ≤ ,
z r r
z∈Cl B(0,r) r
i i
i i
i i
Hence, by Schwarz lemma, we get | F(w)| ≤ |w|, ∀w ∈ B(0, 1). If z ∈ B(0, R),
then |T(z)| < 1, and hence,
S( f (z) − b) R(z − a)
| F(T(z))| ≤ |T(z)| ⇒ |( f (z))| ≤ |T(z)| ⇒ ≤ .
S 2 − bf (z) R2 − az
i i
i i
i i
z−α
Proof: If Tα (z) = , ∀z ∈ C, then we claim that Tα : B(0, 1) → B(0, 1)
1 − αz
is an analytic bijection such that Tα (α) = 0 and Tα−1 is also analytic.
i i
i i
i i
for every closed curve γ is . Then by Result 4.1.30, there exists an analytic
f
function g on , such that g = . If log( f ) is defined, then its derivative
f
f
must be . Hence, we expect that log( f ) would be defined by g + c for some
f
suitable constant c. We find c from
log( f ) = g + c ⇔ exp c = f exp(−g).
To conclude this proof, first we show that f exp(−g) is a constant. This can
be obtained since
d
( f exp(−g)) = f exp(−g) − f exp(−g)g
dz
f
= f exp(−g) − f exp(−g)
f
= 0
and by using Theorem 2.2.19. Hence, let f exp(−g) = exp(c) for some con-
stant c, and this is possible since f exp(−g) is nowhere zero. Then, for a fixed
z0 ∈ ,
f (z0 ) exp(−g(z0 )) = exp(c) ⇒ c = log( f (z0 )) − g(z0 ),
where log( f (z0 )) is any one of the infinite number of its values. Thus, if log( f )
is defined by g + log( f (z0 )) − g(z0 ), then log( f ) is differentiable.
COROLLARY 4.5.13 If f and are as in the previous theorem, then for every
n ∈ N, there exists ananalytic function g on such that gn = f . In other
words, we can define n f as an analytic function on .
i i
i i
i i
i i
i i
i i
and
|μ|2
|fn (ζ ) − f (ζ )| < , ∀n ≥ N2 .
4M1
|μ|
Therefore, for n ≥ N1 , we have |f | − |fn | ≤ |fn − f | < 2 , and hence,
1 f (ζ ) 1 fn (ζ )
dζ = lim dζ
i2π f (ζ ) n→∞ i2π fn (ζ )
|ζ −z|=δ |ζ −z|=δ
1 fn (ζ )
= lim dζ
i2π n→∞ fn (ζ )
|ζ −z|=δ
= 0,
i i
i i
This page is intentionally left blank
i i
5
Series Developments
and Infinite Products
223
i i
i i
i i
1 f (ζ )
Since f is continuous on |z − a| = s, if f1 (z) = dζ , then f1
i2π |ζ −a|=s ζ − z
is analytic on B(a, r) (including at a) and f1 (z) = f (z), ∀z ∈ B(a, r). Hence, if
g : → C is defined by
f (z) if z a
g(z) = ,
f1 (a) if z = a
n−1 (k)
f (z0 )
f (z) = (z − z0 )k + φn (z)(z − z0 )n , ∀z ∈ .
k!
k=1
Furthermore,
1 f (ζ )
φn (z) = , dζ , ∀z inside C,
i2π (ζ − z)(ζ − z0 )n
C
Proof: If we define
f (z) − f (z0 )
F(z) = , ∀z ∈ \{z0 },
z − z0
f (z) − f (z0 )
lim F(z) = lim = f (z0 ).
z→z0 z→z0 z − z0
i i
i i
i i
n
dj (n−j)
f (n)
(z) = nCj ((z − z0 )n )φn (z), ∀z ∈ .
dzj
j=0
dj
For each j = 0, 1, 2, . . . , n − 1, − z0 )n ) has the factor (z − z0 ), and
dz j
((z
dn f (n) (z0 )
dzn ((z − z0 ) ) = n!, we get f (z0 ) = n!φn (z0 ). Hence, φn (z0 ) =
n (n) ,
n!
∀n ∈ N. Therefore, we have
n−1 (k)
f (z0 )
f (z) = f (z0 ) + (z − z0 )k + φn (z)(z − z0 )n , ∀z ∈ .
k!
k=1
Next, we choose r > 0 such that ClB(z0 , r) ⊆ . If C is the circle with centre
z0 and radius r > 0, then by Cauchy’s integral formula, we have
1 φn (ζ )
φn (z) = dζ , ∀z ∈ B(z0 , r).
i2π ζ −z
C
i i
i i
i i
Since
1
n−1 (k)
f (z0 )
φn (z) = f (z) − (z − z0 ) k
,
(z − z0 )n k!
k=0
we have
n−1
1 f (ζ ) f (k) (z0 )
φn (z) = − dζ . (5.1)
i2π (ζ − z0 )n (ζ − z) k !(ζ − z0 )n−k (ζ − z)
C k=0
dζ
Now, we claim that = 0, ∀j ∈ N and ∀z ∈ B(z0 , r). For
C (ζ − z0 ) (ζ − z)
j
each j ∈ R and z ∈ B(z0 , r), define
dζ
Fj,z (w) = , ∀w ∈ B(z0 , r).
(ζ − w)j (ζ − z)
C
1 1 1 1
= − ,
(ζ − w)(ζ − z) z−w ζ −z ζ −w
and hence, using Example 4.2.6, for every w ∈ B(z0 , r), we have
dζ 1 1 1
F1,z (w) = = − dζ
(ζ − w)(ζ − z) z−w ζ −z ζ −w
C C
i2π
= (WN(C, z) − WN(C, w)) = 1 − 1 = 0.
z−w
Using Theorem 4.3.2, we get
(1) (2)
Fj−1,z (w) Fj−2,z (w)
Fj,z (w) = =
j j(j − 1)
(3) (j−1)
Fj−3,z (w) F1,z (w)
= = ··· = = 0,
j(j − 1)(j − 2) j!
for all j > 1. Thus, Fj,z (w) = 0, ∀j > 0 and ∀w ∈ B(z0 , r). In particular,
dζ
= Fj,z (z0 ) = 0, ∀j = 1, 2, 3, . . . , n − 1.
(ζ − z0 )j (ζ − z)
C
i i
i i
i i
n−1 (k)
f (z0 )
f (z) = (z − z0 )k + φn (z)(z − z0 )n , ∀z ∈ ,
k!
k=0
where
1 f (ζ )
φn (z) = dζ , ∀z ∈ B(z0 , r),
i2π (ζ − z)(ζ − z0 )n
C
where C is the circle |ζ − z0 | = r. If M = sup | f (ζ )|, then for z ∈ B(z0 , r),
ζ ∈C
n−1 (k)
f (z0 )
f (z) − (z − z0 )k = |z − z0 |n |φn (z)|
k!
k=0
|z − z0 |n | f (ζ )|
≤ |dζ |
2π |ζ − z||ζ − z0 |n
C
i i
i i
i i
Remark 5.1.6: At this juncture, it should be noted that finite Taylor series
n−1 f (k) (z )
0
expansion f (z) = (z − z0 )k + φn (z)(z − z0 )n of an analytic function
k=0 k !
f on is valid on the entire region , but the infinite Taylor series expan-
∞ f (k) (z )
0
sion (z − z0 )k is valid only in a largest possible neighbourhood
k=0 k!
B(z0 , r) ⊆ .
∞ ∞
Definition 5.1.7 Let the power series an (z − a)n and bn (z − a)n con-
n=0 n=0
verge on B(a, R). Then, the Cauchy product of the two power series is defined
by
∞
n
ak bn−k (z − a)n , ∀z ∈ B(a, R).
n=0 k=0
∞ ∞
THEOREM 5.1.8 If f (z) = an (z − a)n and g(z) = bn (z − a)n ,∀z ∈
n=0 n=0
∞ n
B(a, R), then ( f · g)(z) = ak bn−k (z − a)n , ∀z ∈ B(a, R).
n=0 k=0
Proof: We know that f and g are analytic on B(a, R), and hence, f · g is also
analytic on B(a, R). By Taylor’s theorem, we have
∞
( f · g)(n) (a)
( f · g)(z) = (z − a)n , ∀z ∈ B(a, R).
n!
n=0
i i
i i
i i
Now using Leibniz rule (Result 2.1.13) and Corollary 2.3.4, we have
1
n
( f · g)(n) (a)
= nCk f (k) (a) · g(n−k) (a)
n! n!
k=0
1
n
= nCk f (k) (a) · g(n−k) (a)
n!
k=0
1
n
n!
= (ak k !) · (bn−k (n − k)!)
n! k !(n − k)!
k=0
n
= ak bn−k .
k=0
Hence, for every z ∈ B(a, R), using Corollary 2.3.5, we get ( f · g)(z) =
∞ ( f · g)(n) (a)
(z − a)n .
n=0 n!
∞ ∞
COROLLARY 5.1.9 If an (z − a)n and bn (z − a)n converges on B(a, R),
n=0 n=0
then the Cauchy product of the two power series converge on B(a, R).
f
of the power series is defined as the Taylor series of , where
g
∞
∞
f (z) = an (z − a)n and 0 g(z) = bn (z − a)n , ∀z ∈ B(a, R).
n=0 n=0
i i
i i
i i
Finding the coefficients of the Taylor series of gf in terms of an and bn are not
easy as that of f · g. However, the first few coefficients of the Taylor series of
f
g can be calculated successively.
In the following sequel, we use the notation ANsr (a) to denote the annulus
{z ∈ C : r < |z − a| < s}.
∞ ∞ ∞
Definition 5.1.11 We say that an is convergent if an and a−n
n=−∞ n=0 n=1
∞ ∞ ∞
are convergent; in this case, we write an = a−n + an .
n=−∞ n=1 n=0
∞
THEOREM 5.1.12 For a given an (z − z0 )n , if r = lim sup |a−n |1/n , s =
n=−∞ n→∞
lim sup |an |1/n in [−∞, +∞], and r < s, then
n→∞
∞
1. an zn converges on r < |z − z0 | < s,
n=−∞
∞
2. an zn converges uniformly on r1 ≤ |z − z0 | ≤ s1 , if r < r1 <
n=−∞
s1 < s,
∞
3. an zn diverges on |z − z0 | < r or |z − z0 | > s.
n=−∞
Proof: Choose positive real numbers r1 and r2 such that R1 < r1 < r2 < R2 .
If Cj is the circle with centre z0 and radius rj , for j = 1, 2, then we claim that
C2 ∪ (−C1 ) ∼ 0 in ANRR21 (z0 ). Let w ANRR21 (z0 ).
i i
i i
i i
Case 1: |w − z0 | > R2 .
Then, |w − z0 | > r1 and |w − z0 | > r2 , which implies WN (C2 ∪
(−C1 ), w) = WN(C2 , w) − WN(C1 , w) = 0.
Case 2: |w − z0 | < R1 . In this case, |w − z0 | < r1 , |w − z0 | < r2 , and hence,
WN (C2 ∪ (−C1 ), w) = WN (C2 , w) − WN (C1 , w) = 1 − 1 = 0.
Hence, our claim holds. Then, by the general version of Cauchy’s integral
formula, we have
1 f (ζ )
WN (C2 ∪ (−C1 ), z)f (z) = dζ ∀z C2 ∪ (−C1 ).
i2π ζ −z
C2 ∪(−C1 )
|ζ − z0 | r1
Since = < 1, ∀ζ ∈ C1 , we get
|z − z0 | |z − z0 |
1 1
=
ζ −z (ζ − z0 ) −(z − z0 )
∞
k
−1 1 −1 ζ − z0
= = .
z − z0 1 − ζ −z0
z−z0
z − z0 z − z0
k=0
We note that the power series on the right-hand side converges uniformly on
|z − z0 |
C1 . Similarly, using < 1, ∀ζ ∈ C2 , we get
|ζ − z0 |
1 1
=
ζ −z (ζ − z0 ) −(z − z0 )
∞
k
1 1 1 z − z0
= = ,
ζ − z0 1 − ζz−z
−z
0 ζ − z0 ζ − z0
0 k=0
i i
i i
i i
∞
(z − z0 )−k−1 f (ζ )
+ dζ
i2π (ζ − z0 )−k
k=0 C1
∞
(z − z0 )k f (ζ )
= dζ
i2π (ζ − z0 )k+1
k=0 C2
∞
(z − z0 )−j f (ζ )
+ dζ
i2π (ζ − z0 )−j+1
j=1 C1
∞
= ak (z − z0 )k ,
k=−∞
⎧ 1 f (ζ )
⎪
⎪ dζ if k ≥ 0
⎨ i2π (ζ − z0 )k+1
C2
where ak = 1 , ∀k ∈ Z.
⎪
⎪
f (ζ )
if k < 0
⎩ dζ
i2π C1 (ζ − z0 )k+1
i i
i i
i i
∞
2π
1
= ck (r2 exp(it))k−n−1 ir2 exp(it) dt
i2π
k=−∞ 0
(as C2 is given by ζ (t) = z0 + r2 exp(it), t ∈ [0, 2π])
∞ 2π
k−n 1
= ck ir2 exp(i(k − n)t) dt
i2π
k=−∞ 0
= cn ,
1 2π 1 k=n
since exp(i(k − n)t) dt = . By a similar argument, we get
2π 0 0 kn
an = cn for every negative integer also. Hence, the result follows.
1
Step 2(b). To expand , in |z − z0 | > R for some R > 0 and for some
(z ± a)
a ∈ C with |b| < R, where b = −(z0 ± a).
i i
i i
i i
Then write
1 1
=
(z ± a)m ((z − z0 ) − b)m
1 1
=
(z − z0 )m 1 − b m
(z−z0 )
∞
k−m
1 1 b
= k(k − 1) · · · (k−m+1) .
(z − z0 )m m! (z − z0 )
k=m
Step 3. Finally, replace all the summands by the corresponding series and
simplify.
From the following examples, one can understand the algorithm still better.
Solution:
1. Applying partial fraction technique, we write
1 1/3 1/3
= − .
z2 − 5z + 4 z−4 z−1
(a) If 0 < |z| < 1, then we have1
1 z k
∞
1/3 1
=− z =− (5.2)
z−4 12 1 − 12 4
k=0
4
1 Here note that |z| < 1 < 4 so we have taken −4 and −1 as the common factors in equations
(5.2) and (5.3), respectively.
i i
i i
i i
and
∞
−1/3 1 1 k
= = z . (5.3)
z−1 3 (1 − z) 3
k=0
Thus, we get
1 z k 1 k
∞ ∞
1
= − + z
z − 5z + 4
2 12 4 3
k=0 k=0
∞
1 1
= 1− zk , 0 < |z| < 1.
3 4k+1
k=0
1 z k
∞
1/3 1
=− z =− (5.4)
z−4 12 1 − 12 4
k=0
4
and
∞ ∞
−1/3 1 1 1 k
1 1 k+1
=− =− =− .
z−1 1 3z z 3 z
3z 1 − k=0 k=0
z
(5.5)
Thus, we get
1 z k 1 1 k+1
∞ ∞
1
= − −
z − 5z + 4
2 12 4 3 z
k=0 k=0
∞ ∞
1 1 k −k
= − z + z , 1 < |z| < 4.
3 4k+1
k=0 k=1
i i
i i
i i
and
∞ ∞
−1/3 1 1 1 k
1 1 k+1
=− =− =− .
z−1 1 3z z 3 z
3z 1 − k=0 k=0
z
Thus, we get
∞ ∞
1 1 4k 1 1 k+1
= −
z − 5z + 4
2 3 z k+1 3 z
k=0 k=0
∞
1 k
= − (4 − 1)z−(k+1) , 0 < |z| < 1.
3
k=1
and
−(z + 1) −(z + 1)
= z
5(z + 2) 10 1 +
2
∞
z+1 z k
= −
10 −2
k=0
i i
i i
i i
Thus, we get
4 z k+1 −1
∞ ∞
z+1 1 1 z k+1
= − + + +
z2 − z − 6 15 15 3 10 10 −2
k=0 k=0
∞
−1 4 −k 1
= + 3 + (−2)k zk , 0 < |z| < 2.
6 15 10
k=1
4 z k+1
∞
z+1 z+1 1
=− z = − +
5 (z − 3) 15 1 − 15 15 3
k=0
3
and
z+1 z+1
− = −
5(z + 2) 2
5z 1 +
z
∞
z+1 2 k
= − −
5z z
k=0
∞ ∞
1 2 k 1 2 k+1
= − − − −
5 z 2 z
k=0 k=0
∞ ∞
1 2 k+1 1 2 k+1
= − 1+ − − −
5 z 2 z
k=0 k=0
∞
k+1
1 1 2
= − + − .
5 10 z
k=0
i i
i i
i i
Thus, we get
4 z k+1 1
∞ ∞
z+1 1 1 2 k+1
= − + − + −
z −z−6
2 15 15 3 5 10 z
k=0 k=0
∞
4 4
= − + 3−k zk
15 15
k=1
∞
1
+ (−2)k z−k , 2 < |z| < 3.
10
k=1
i i
i i
i i
and
z z
− = −
2(z − 3) 2(z − 2 − 1)
z−2+2 1
=
2 (1 − (z − 2))
z−2 1
= +1
2 (1 − (z − 2))
∞
z−2
= +1 (z − 2)k
2
k=0
∞ ∞
1
= (z − 2)k+1 + (z − 2)k
2
k=0 k=0
∞ ∞
1
= (z − 2)k + 1 + (z − 2)k
2
k=1 k=1
i i
i i
i i
∞
3
= 1+ (z − 2)k .
2
k=1
Therefore,
∞
1 1 5 −k
= − − 3 (z − 2)k + 1
z2 − 8z + 15 3 6
k=1
∞
3
+ (z − 2)k
2
k=1
∞
2 3 5 −k
= + − 3 (z − 2)k .
3 2 6
k=1
and
z z
− = −
2(z − 3) 2(z − 2 − 1)
z−2+2 1
= −
2(z − 2) 1 − 1
z−2
i i
i i
i i
1 1 1
= − −
2 z−2 1− 1
z−2
∞ ∞
−1
= (z − 2)−k + (z − 2)−k
2
k=0 k=1
∞
1 1
= − + (z − 2)−k .
2 2
k=1
Therefore,
∞ ∞
1 1 5 −k 1 1
= − − 3 (z − 2)k − + (z − 2)−k
z − 8z + 15
2 3 6 2 2
k=1 k=1
∞ ∞
5 5 −k 1
= − − 3 (z − 2)k + (z − 2)−k .
6 6 2
k=1 k=1
z z
=
2(z − 5) 2(z − 2 − 3)
z−2+2 1
=
2(z − 2) 1 − 3
z−2
1 1 1
= +
2 z−2 1− 3
z−2
∞
k
1 1 3
= +
2 z−2 z−2
k=0
∞ ∞
1 3 k
1 3 k
= +
2 z−2 3(z − 2) z−2
k=0 k=0
∞
k ∞
k
1 1 3 1 3
= + +
2 2 z−2 3 z−2
k=1 k=1
∞
1 5
= + 3k (z − 2)−k
2 6
k=1
i i
i i
i i
and
z z
− = −
2(z − 3) 2(z − 2 − 1)
z−2+2 1
= −
2(z − 2) 1 − 1
z−2
1 1 1
= − −
2 z−2 1− 1
z−2
∞
1 1
= − + (z − 2)−k .
2 2
k=1
Therefore,
∞ ∞
1 1 5 k 1 1
= + 3 (z − 2)−k − + (z − 2)−k
z − 8z + 15
2 2 6 2 2
k=1 k=1
∞
∞
5 1
= 3−k (z − 2)k + (z − 2)−k .
6 2
k=1 k=1
i i
i i
i i
∞ ∞
1 z+1 k
2 2 z+1 k
= − − −
2 5 5 5 5
k=1 k=1
∞
2 7 −k
= − − 5 (z + 1)k
5 10
k=1
and
z+5 z+5
− = −
2(z + 2) 2(z + 1 + 1)
z+1+4
= −
1
2(z + 1) 1 +
z+1
∞
k
1 2 −1
= − +
2 z+1 z+1
k=0
∞ ∞
1 −1 k −1 k+1
= − +2
2 z+1 z+1
k=0 k=0
∞
k ∞
k
1 1 −1 −1
= − − +2
2 2 z+1 z+1
k=1 k=1
∞
1 3
= − + (−1)k (z + 1)−k .
2 2
k=1
Thus, we get
∞
1 1 7 −k
= − − 5 (z + 1)k + 2
z2 − 5z + 4 2 10
k=1
∞
1
− (−1)k (z + 1)−k
2
k=1
∞
7 −k
= −1 − 5 (z + 1)k
10
k=1
∞
3
+ (−1)k (z + 1)−k , ∀z with 1<|z + 1|<3.
2
k=1
i i
i i
i i
Exercise 5.1.17 Find the Laurent series expansions of the given functions in
the specified regions.
1
1. f (z) = in 0 < |z| < 1.
z2 (1 − z)
1 √ √
2. f (z) = in (i) 0 < |z − i| < 2, (ii) |z − i| > 2.
1+z
1
3. f (z) = 2 in 0 < |z| < 2.
z − 5z + 6
(1 − z)3
4. f (z) = in |z − 1| > 1.
z−2
1
5. f (z) = in (i) 0 < |z − 1| < 2, (ii) 0 < |z − 3| < 2.
(z − 1)2 (z − 3)
Answers:
∞
1
1. .
zn
n=−2
∞
n ∞
n
z−i 1+i
2. (i) (−1)n , (ii) (−1)n+1 .
1+i z−i
n=0 n=1
∞
1 1
3. − n zn−1 .
2n 3
n=1
∞
1
4. −(z − 1)3 .
(z − 1)n
n=1
∞
n ∞
n−1
1 z−1 1 z−3
5. (i) − , (ii) n − .
2(z − 1) 2 2 4(z − 3) 2
n=0 n=1
i i
i i
i i
n−1 (k)
f (a)
f (z) = (z − a)k + φn (z)(z − a)n = φn (z)(z − a)n , ∀z ∈ ,
k!
k=0
where
1 f (ζ )
φn (z) = dζ , ∀z inside C,
i2π (ζ − a)n (ζ − z)
|ζ −a|=r
where r > 0 is such that ClB(a, r) ⊆ . If M = sup | f (ζ )|, then M < ∞ and
ζ ∈C
for every z ∈ B(a, r),
|z − a|n | f (ζ )|
| f (z)| ≤ |dζ |
2π |ζ − a|n |ζ − z|
|ζ −a|=r
M|z − a|n |dζ |
≤
2π rn (r − |z − a|)
|ζ −a|=r
(as |ζ − a| ≤ |z − ζ | + |z − a| ⇒ r − |z − a| ≤ |z − ζ |)
M|z − a|n 2πr
≤
2π r (r − |z − a|)
n
|z − a| n Mr
≤ → 0 as n → ∞.
r r − |z − a|
i i
i i
i i
B = ∅ ⇒ = A ⇒ f = 0 on ,
k−1 (j)
f (z0 )
f (z) = (z − z0 ) j + g(z)(z − z0 )k = g(z)(z − z0 )k , ∀z ∈ .
j!
j=0
dν
k
f (k)
(z) = kCν g(k−ν) (z) (z − a)k
dzν
ν=0
and
dν 0 ν<k
ν
(z − a)k = ,
dz z=a k! ν=k
we obtain f (k) (a) = g(a)k ! = 0, which contradicts the fact that the order of
the zero of f at z0 is k . Thus, g(a) 0.
i i
i i
i i
i i
i i
i i
Proof: We note that the regions enclosed by a closed curve are bounded sets.
Suppose, there are infinitely many zeroes of f , which are enclosed by γ , then
the set {z ∈ : f (z) = 0} has a limit point, by Bolzano–Weierstrass property
(Theorem 1.4.47). Hence, by the principle of analytic continuation (Theorem
5.2.4), it would follow that f is identically zero. Hence, there must be only
finitely many zeroes of f enclosed by γ .
i i
i i
i i
In particular, maximum of |F| is attained at all points of B(0, 1); hence, again
by using maximum modulus principle, we get F = c on Cl B(0, 1) for some
c ∈ C with
|c| = 1. Then, by principle of analytic continuation, we get F = c
1
on C \ : k = 1, 2, . . . n with zk 0 . Therefore,
zk
n
z − zk 1
f (z) = c , ∀z ∈ C \ : k = 1, 2, . . . n with zk 0 .
1 − zk z zk
k=1
Proof:
(1) ⇒ (2) Since a is a removable singularity of f , we have lim f (z) = for
z→a
some ∈ C. Therefore, for = 1, there exists r > 0 such that
0 < |z − a| < r ⇒ | f (z) − | < 1.
Therefore, | f (z)| ≤ 1 + | |, ∀z ∈ B(a, r)\{a}. Hence, (2) follows.
i i
i i
i i
(2) ⇒ (3) By assumption, there exists r > 0 and M > 0 such that
Proof: Since lim f (z) = ∞, there exists > 0 such that B(a, ) ⊆ and
z→a
i i
i i
i i
i i
i i
i i
Hence, we have
∞
∞
−m
f (z) = (z − a) bn (z − a) =
n
bn (z − a)n−m , ∀z ∈ B(a, r)\{a}.
n=0 n=0
i i
i i
i i
Since
∞
ψ(z) = (z − a)m cn (z − a)n
n=−m
∞
= cn (z − a)n+m
n=−m
∞
= cn−m (z − a)n , |z − a| < r
n=0
∞
and the power series cn−m (z − a)n is analytic inside B(a, r), we get
n=0
that ψ is an analytic function in B(a, r) with
f (z) = (z − a)−m ψ(z), ∀z ∈ B(a, r)\{a} and ψ(a) = c−m 0.
Hence, f has a pole at a of order m.
3. a is essential singularity iff a is neither a removable singularity nor a
pole iff cn 0 for some n < 0 and for every m ∈ N, there exists n ∈ Z
such that n < −m and cn 0 iff cn 0 for infinitely many n < 0.
Hence, the result follows.
i i
i i
i i
Hence, φ(a) = 0. Since φ is not identically zero3 using Result 5.2.3, we have
for some ν ∈ N, for some r > 0 and for an analytic function φ1 on B(a, r)
with φ1 (a) 0. Now, let m = n − ν . Then,
lim |z − a|k | f (z)| = lim |z − a|k−m lim |z − a|m | f (z)| = ∞ · |φ1 (a)| = ∞.
z→a z→a z→a
Next, assume that (2) holds. Then, choose an integer l such that l < s. Then,
lim |z − a|l | f (z)| = lim |z − a|l−s lim |z − a|s | f (z)| = ∞,
z→a z→a z→a
Hence, by using Result 5.2.11, there exist a negative integer λ, δ > 0 and an
analytic function ψ1 on B(a, δ) ⊆ such that
ψ(z) = (z − a)λ ψ1 (z), ∀z ∈ B(a, δ)\{a}.
3 As f has a singularity, f should not be identically zero, and hence, φ is not identically zero.
i i
i i
i i
Therefore,
f (z) = (z − a)−l ψ(z) = (z − a)λ−l ψ1 (z), ∀z ∈ B(a, δ) \ {a}.
If m = l − λ, then
lim |z − a|m | f (z)| = lim |ψ1 (z)| = |ψ1 (a)|,
z→a z→a
and
lim |z − a|k | f (z)| = ∞, ∀k < mi , for i = 1, 2,
z→a
then obviously, m1 = m2 .
i i
i i
i i
Proof: Suppose that this theorem is not true. That is, there exist c ∈ C, > 0
and δ > 0 such that
|z − a| < δ ⇒ | f (z) − c| ≥ .
i i
i i
i i
Clearly,
lim |z − a|−1 | f (z) − c| ≥ lim |z − a|−1 = ∞.
z→a z→a
Hence, by Result 5.2.15, there exists r > 0, such that
Therefore,
Exercise 5.2.20 Prove that the algebraic order of f at a is the least integer
∞
m ∈ Z such that cm 0 in the Laurent series expansion of f (z) = ck (z−
k=−∞
z0 )k , 0 < |z − z0 | < δ for some δ > 0.
1
In the partial fraction of f , Pn is called the singular part of f at an .
z − an
i i
i i
i i
for some suitable polynomial Qn such that the above series converges
uniformly on every compact subset of C\{an : n = 0, 1, 2, . . .}.
Let M > 0 be arbitrary. Since an → ∞ as n → ∞, we choose N ∈ N
1
such that 4|an | > M , ∀n ≥ N . For n ≥ N , if gn (z) = Pn , ∀z ∈
z − an
B(0, |an |), then gn is analytic on B(0, |an |). Then, the finite Taylor series
expansion of gn is given by
1
mn
gn (z) = Pn = ck zk + φmn +1 (z)zmn +1 , ∀z ∈ B(0, |an |),
z − an
k=0
1
2mn −n > Mn = sup Pn < +∞
|z|≤ |a2n |
z − an
and
1
Pn
1 z − an |an |
φmn +1 (z) = +1
dζ , ∀z ∈ B 0, .
i2π ζ mn (ζ − z) 2
|ζ |= |a2n |
mn
Let Qn (z) = ck zk , ∀z ∈ C. Now for z ∈ B 0, |a4n | ,
k=0
1
Pn − Qn (z) = φmn +1 (z)zmn +1
z − an
Mn |dζ |
≤ |z|mn +1 mn +1
2π |an |
|ζ − z|
|ζ |= |a2n | 2
Mn |dζ |
≤ |z|mn +1 mn +1
2π |an | |an | |an |
−
|ζ |= |a2n | 2 2 4
i i
i i
i i
|an | mn +1 Mn |an | 2 mn +1
4
≤ 2π
4 2π 2 |an | |an |
= 2−mn Mn < 2−n .
∞ 1
Hence, by comparison test, Pn − Qn (z) converges uni-
n=N z − an
formly on |z| ≤ M . As every compact subset of C\{an : n =
0, 1, 2, . . .} is contained in B (0, M) for some M > 0, we conclude that
∞ 1
Pn − Qn (z) converges uniformly on every compact subset
n=N z − an
of C\{an : n = 0, 1, 2, . . .}. Thus, there exists a meromorphic function
∞
1
f (z) = Pn − Qn (z) , +φ(z), ∀z ∈ C\{an : n = 0, 1, 2, . . .},
z − an
n=0
1
whose singular part at an is Pn , ∀n = 0, 1, 2, . . ., where φ is an
z − an
arbitrary entire function.
π n (−1)k
3. = lim , ∀z ∈ C\Z.
sin(π z) n→∞ k=−n z − k
π2
Let f (z) = , ∀z ∈ C\N. Clearly, f has double pole at every integer.
sin2 (π z)
Therefore, the Laurent series expansion of f around 0 is of the form
∞
A2 A1
f (z) = 2
+ + An zn , 0 < |z| < δ for some δ > 0,
z z
n=0
A2 A1
and hence, the singular part of f at 0 is 2
+ , where
z z
π2
A−2 = lim z2 = 1 and A−1 = 0,
z→0 sin2 (πz)
i i
i i
i i
as f is an even function, and hence, all odd coefficients in the Laurent series
1
expansion of f around 0 are zeroes. Thus, the singular part of f at 0 is 2 .
z
1
Since sin2 (π (z − n)) = sin2 (πz), the singular part of sin2 (π z) at n is .
(z − n)2
Therefore, by Mittag-Leffler theorem, we have
∞
1
f (z) = + φ(z), ∀z ∈ C\N,
n=−∞
(z − n)2
∞1
for some entire function φ . (As itself is convergent, we have
(z − n)2
n=−∞
taken Qn (z) = 0, ∀z ∈ C and n ∈ N.) We claim that φ is identically zero.
π2 ∞ 1
First, we note that as and are periodic functions of
2
sin (πz) n=−∞ (z − n)2
period 1, φ is also a periodic function of period 1. Now, writing z = x + iy,
we get
| sin(π z)|2 = | sin(π(x + iy))|2
= | sin(πx) cosh(π y) + i sinh(πy) cos(π x)|2
= sin2 (πx) cosh2 (πy) + sinh2 (π y) cos2 (π x)
= (1 − cos2 (π x)) cosh2 (π y) + sinh2 (π y) cos2 (π x)
= cosh2 (πy) − cos2 (π x)(cosh2 (π y) − sinh2 (π y))
= cosh2 (πy) − cos2 (π x) → ∞ as |y| → +∞ uniformly in x.
Therefore,
g(x + iy) → 0 as |y| → +∞ uniformly on C\Z. (5.6)
In particular, g is bounded in the strip {(x, y) : 0 ≤ x ≤ 1, y ∈ R}. Since
g is a periodic function of period 1, g is bounded on C. Thus, by Liouville’s
theorem (Theorem 4.3.11), g is a constant function. Moreover, using equation
(5.6), we conclude that g is identically 0. Thus,
∞
n
π2 1 1
= = lim .
sin2 (π z) (z − k)2 n→∞ (z − k)2
k=−∞ k=−n
− cot(π z)
n
−1
π2 = lim ,
π n→∞ z−k
k=−n
i i
i i
i i
which implies the required series representation of π cot(π z). Next, consider
n
(−1)k
(2j+1)
(−1)k
lim = lim
n→∞ z−k j→∞ z−k
k=−n k=−(2j+1)
⎛ ⎞
j
1 j
1
= lim ⎝ − ⎠
j→∞ z − 2k z − (2k + 1)
k=−j k=−j−1
⎛ ⎞
1 j
1 j
1 ⎠
= lim ⎝ −
2 j→∞ z
− k z−1
−k
k=−j 2 k=−j−1 2
π πz π (z − 1)
= cot − cot
2 2 2
⎛ πz ⎞
π (z − 1)
cos cos
π⎜ ⎜ 2 ⎟
⎟
= ⎝ π2z −
2 sin π (z − 1) ⎠
2 sin
⎛ 2
πz πz ⎞
π ⎜ cos 2 sin
⎟
= ⎝ πz + π2z ⎠
2 sin cos
2 2
π 1
=
2 sin π z cos π z
2 2
π
= .
sin(πz)
PROBLEM 5.3.4 Find the partial fraction expansion for sec and deduce the
∞ (−1)n+1
Gregory–Leibniz–Madhava series, which is given by .
n=0 2n + 1
Solution:
From Example 5.3.3(3), we have
∞
π π (−1)n
$π %= = .
cos 2 − πz sin(πz) n=−∞ z − n
π
By substituting w = 2 − π z, we get
∞
π (−1)n
= ,
cos(w) n=−∞ − πw + 12 − n
i i
i i
i i
∞
∞
(−1)n (−1)n
= − −
n=1 w + n − 12 π n=0 w + −n − 12
∞
∞
(−1)n (−1)m
= − +
n=1 w + n − 12 π m=1 w − m − 12
(using the change of variable m = n + 1 in the second sum)
⎡ ⎤
∞
1 1
= (−1)n ⎣− + ⎦
n=1 w + n − 1
2 π w − n − 2 π
1
∞ 2 n − 12 π
= (−1)n 2
n=1 w2 − n − 12 π 2
∞
(2n − 1)π
= 4 (−1)n .
4w2 − (2n − 1)2 π 2
n=1
Substituting w = 0 in the above series representation of sec(w), we get
∞
∞
(−1)n+1 (−1)n+1 π
1=4 ⇒ = .
(2n − 1)π (2n − 1) 4
n=1 n=1
Exercise 5.3.5
$ % ∞ 1
1. Prove that π tan πz2 = 4z . Hint: Use the identity
n=1 (2n − 1) − 4z
2 2
$ πz % $ πz %
tan 2 = cot 2 − 2 cot(πz) and the partial fraction expansion of
π cot(π z).
2. Obtain the value of Gregory–Leibniz–Madhava series from the partial
π 1
fraction expansions of πz cot(π z) and . Hint: Substitute z =
sin(π z) 4
in both expansions.
∞ 1 π2
PROBLEM 5.3.6 Prove that 2
= .
k=1 k 6
i i
i i
i i
Solution:
From Example 5.3.3, we have
∞
π2 1
=
sin2 (π z) (z − k)2
k=−∞
1 1 1
= $ % + 2
k2 1 − z 2 z
0k∈Z k
1 ∞ z m−1 1
= 2
m + 2. (5.7)
k k z
0k∈Z m=1
1
Since the Laurent series of 2
is given by
sin (πz)
⎡ ⎤2
⎢ 1 ⎥
1 ⎢ 1 ⎥
= ⎢ ⎥
sin2 (πz) ⎣ πz (πz)2 (π z)4 ⎦
1− + − ···
3! 5!
∞ n 2
1 (πz)2 (π z)4
= + − ···
πz 3 5!
n=0
∞ n 2
1 (πz)2 (π z)4
= + − ···
πz 3! 5!
n=0
∞
1 2
= + + Am zm ,
(πz)2 3!
m=1
π2
Equating the constant terms of the expansions of in equations
sin2 (π z)
(5.7) and (5.8), we get
1 ∞
∞
2π 2 1 1 π2
= = 2 ⇒ = .
6 k2 k2 k2 6
0k∈Z k=1 k=1
i i
i i
i i
n
Definition 5.4.2 (Customary definition) ζk is said to be convergent if
k=1
n
lim ζk exists in C\{0}.
n→∞ k=1
∞ 1
n 1
Note that according to this definition, does not converge, as =
k=1 2 k=1 2
1
→ 0 as n → ∞.
2n
n
LEMMA 5.4.3 If ζk converges, then ζk → 1 as k → ∞.
k=1
n
Proof: Let n = ζk , ∀n ∈ N. Then, by assumption, we have lim n =
k=1 n→∞
k
A for some A ∈ C\{0}. As ζk = , ∀k > 1, allowing k → ∞, we get
k−1
A
lim ζk = = 1.
k→∞ A
∞ ∞
LEMMA 5.4.4 ζk converges iff log(ζk ) converges, for a suitable
k=1 k=1
branch of log.
Proof: Using Corollary 2.4.25, for every ζk , there exists unique θk ∈ (0, 2π ]
such that arg ζk = θk . Since {θk : k ∈ N} is countable, there exists θ ∈ (0, 2π )
such that arg ζk θ , ∀k ∈ N. Now, we define a branch of log as follows. Let
θ = C\ ({0} ∪ {r exp(iθ ) : r > 0}) and define
i i
i i
i i
so that θ contains positive real axis and arg 1 = 0. Then, as in the proof
of Result 3.3.9, we can prove that log is analytic on θ . Now, assume that
∞
log(ζk ) converges. Then, using the continuity of exp, we get
k=1
∞
n
exp log(ζk ) = exp lim log(ζk )
n→∞
k=1 k=1
n
= lim exp log(ζk )
n→∞
k=1
n
= lim exp (log(ζk ))
n→∞
k=1
n
= lim ζk .
n→∞
k=1
∞
Therefore, ζk converges.
k=1
∞
n
Conversely, assume that ζk converges to A. If n = ζk , ∀n ∈ N, then
k=1 k=1
we have
n
n
log = log(ζk ) − log A + i2π νn , where νn ∈ Z, ∀n ∈ N.
A
k=1
Using the continuity of log at 1, we get log An → log(1) = 0 as n → ∞.
Therefore,
n+1 n
0 = lim log − log
n→∞ A A
n+1
n
= lim log(ζk ) − log(ζk ) + i2π (νn+1 − νn )
n→∞
k=1 k=1
= lim [log(ζn+1 ) + i2π(νn+1 − νn )]
n→∞
= lim i2π(νn+1 − νn ) (by Lemma 5.4.3).
n→∞
n
Using → 1 and ζn → 1 as n → ∞, we obtain
A
n+1 n
arg − arg → 0 and arg ζn+1 → 0 as n → ∞.
A A
i i
i i
i i
Since
n+1 n
(νn+1 − νn )2π = arg − arg + arg (1 + ζn+1 ),
A A
we get νn+1 − νn → 0 as n → ∞. Therefore, by definition of a convergent
1 1
sequence, for = > 0, there exists N ∈ N such that |νn+1 − νn | < .
2 2
Since νn+1 − νn is an integer, for each n ∈ N, we have
νn = νn+1 = ν (say), ∀n ≥ N.
∞
Thus, log(ζk ) converges.
k=1
It is interesting to note that we can use any branch of log which is analytic
on the positive real axis, in particular the principal logarithm.
∞ ∞
Indeed, ζk → 1 iff log(ζk ) → 0 implies that both ζk and log(ζk )
k=1 k=1
do not converge whenever ζk → 1 (or equivalently log(ζk ) → 0). In the other
case, as ζk → 1 as k → ∞, we can very well define principal logarithm
of ζk by log(ζk ) = In(|ζk |) + i arg(ζk ) where ζk ∈ (−π , π ]. However, we
can find N ∈ N such that ζk ∈ C\{(x, 0) : x ≤ 0} ∀k ≥ N , using ζk → 1 as
k → ∞. This is sufficient to get the proof of the previous lemma for principal
logarithm.
∞
Definition 5.4.5 An infinite product ζk is said to be absolutely convergent
k=1
∞
if log(ζk ) converges absolutely.
k=1
As in the case of infinite series, we can define the concept of rearrange-
ment of an infinite product as follows.
∞
Definition 5.4.6 If f : N → N is a bijection, then ζf (k) is called an
k=1
∞
rearrangement of ζk .
k=1
i i
i i
i i
∞
RESULT 5.4.7 If ζk converges absolutely, then every rearrangement of
k=1
∞
ζk converges.
k=1
∞ ∞
Proof: By definition, ζk converges absolutely means that log(ζk ) con-
k=1 k=1
verges absolutely. Then, using Result 1.3.29, we get every rearrangement of
∞
∞
log(ζk ) converges, and hence, every rearrangement of ζk converges.
k=1 k=1
∞ ∞
THEOREM 5.4.8 ζk converges absolutely iff (ζk − 1) converges
k=1 k=1
absolutely.
∞ ∞
Proof: First, we note that if ζk converges absolutely or (ζk − 1) con-
k=1 k=1
verges absolutely, then by Lemma 5.4.3 or by Result 1.3.27, we have ζk → 1
as k → ∞. ∞ ∞
We claim that (ζk − 1) converges absolutely iff log(ζk ) converges
k=1 k=1
absolutely. We note that Taylor series expansion of log(ζk ) around 1 is given
by
∞
∞
log(m) (1) (−1)m+1
log(ζk ) = (ζk − 1)m = (ζk − 1)m .
m! m
m=0 m=1
Therefore,
∞
log(ζk ) (−1)m+1
lim = lim (ζk − 1)m−1 = 1.
k→∞ ζk − 1 k→∞ m
m=1
i i
i i
i i
Solution:
m 1
Let Pm = 1− , ∀m ∈ N with m ≥ 2. Therefore,
n=2 n2
m
1 1
Pm = 1− 1+
n n
n=2
m
n−1 n+1
=
n n
n=2
1 3 2 4 m−1 m+1
= ···
2 2 3 3 m m
1 m+1 1
= → 0 as m → ∞.
2 m 2
∞ 1 1
Hence, 1− converges to .
n=2 n2 2
∞ n 1
PROBLEM 5.4.10 Prove that (1 + z2 ) = , when |z| < 1.
n=1 1−z
Solution:
m−1 n
2m −1
If Pm (z) = (1 + z2 ), ∀m ∈ N, then we claim that Pm (z) = zk ,
n=1 k=0
∀m ∈ N. For m = 1, our claim is true. Assume that our claim is true for some
m ∈ N, then for
m
n
Pm+1 (z) = (1 + z2 )
n=1
m
= Pm (z) (1 + z2 )
i i
i i
i i
Solution: ,
∞ z - −z
To prove this statement, we shall obtain that 1+
e n converges
log
n=1 n
absolutely and uniformly on {z ∈ C : |z| ≤ R}, ∀R > 0 with respect to
the principal branch of log. As principal logarithm is analytic at 1, we can
∞ 1
expand the Taylor series expansion of log(1+w) = − wk , ∀w ∈ B(0, 1).
k=1 k
Therefore, for each |z| ≤ R and each n > R, we have
, z - −z z z
log 1 + en = log 1 − −
n n n
(−1)k z k z
∞
= −
k n n
k=1
(−1)k z k
∞
=
k n
k=2
i i
i i
i i
∞
k
1 R
≤
k n
k=2
∞
1 R2 R k
≤
2 n2 n
k=0
⎛ ⎞
1R2 ⎜ 1⎟
≤ ⎝
2n2 R⎠
1−
n
R2
= ,
2n(n − R)
which is the nth term of a convergent series of real numbers (i.e., this
∞ , z - −z
term is independent of z). Hence, by comparison test, log 1 + en
n=1 n
converges absolutely and uniformly on {z ∈ C : |z| ≤ R}.
∞ 2n + 1
2. 1− = 3.
n=2 n(n + 2)
∞ (−1)n
3. 1+ = 1.
n=2 n
Definition 5.4.13 Let f be an entire function with zeroes {an : n ∈ N}. Then,
the representation
∞
z
f (z) = zm exp(g(z)) 1− exp(Pn (z)), ∀z ∈ C
ak
n=1
i i
i i
i i
Proof:
d
(f (z) exp(−g(z))) = f (z) exp(−g(z)) − exp(−g(z))g (z)f (z)
dz
f (z)
= f (z) exp(−g(z)) − exp(−g(z)) f (z)
f (z)
= f (z) exp(−g(z)) − exp(−g(z))f (z)
= 0.
Case (ii) If f has finite number of zeroes, say for example, and
a1 , a2 , a3 , . . . , an ∈ C\{0} including multiplicities and 0 is a zero
of order m (where m ≥ 0), then applying Result 5.2.3 repeatedly,
we get an entire function h, which has no zeroes in C, such that
n
(−1)n ak h(z) = exp(g(z)), ∀z ∈ C.
k=1
i i
i i
i i
Case (iii) Let f has countably infinite number of zeroes {an : n ∈ N} such
that an → ∞ as n → ∞. In this case, with the motivation from
Case (ii), we shall find an entire function f such that
∞
z
f (z) = z exp(g(z))
m
1− , ∀z ∈ C.
ak
n=1
∞ z
However, the convergence of 1−
is not guaranteed
n=1 a k
for an arbitrary sequence of complex numbers with an →
∞ as n → ∞. Thus, we have to find a suitable sequence
∞ z
of polynomials (Pn ) such that 1− exp(Pn (z)) con-
n=1 a k
∞ z
verges. We know that 1− exp(Pn (z)) converges iff
ak
∞
n=1
z
log 1 − + Pn (z) converges with respect to a suitable
n=1 a k
branch of log. Let R > 0 be given. Choose n0 ∈ N such that
|an | > 2R, ∀n ≥ n0 . If |z| ≤ R and n ≥ n0 , then log 1 − azn
is analytic and its Taylor series is given by
∞
k
z 1 z
log 1 − = −
an k an
k=1
n k ∞
k
1 z 1 z
= − − .
k an k an
k=1 k=n+1
n k
1 z
Therefore, for n ≥ n0 , if Pn (z) = , ∀z ∈ B(0, R) then
k=1 k an
∞
k
z 1 z
log 1 − + Pn (z) ≤
an k an
k=n+1
∞
k
1 R
≤
n+1 |an |
k=n+1
i i
i i
i i
∞
n+1 k
R R 1
≤ since ≤1
|an | |an | n+1
⎛
k=0 ⎞
n+1 ⎜ ⎟
R ⎜ 1 ⎟
= ⎝
|an | R ⎠
1−
|an |
n+1
1
≤ M for some M > 0.
2
⎛ ⎞
⎜ 1 ⎟
Since ⎜
⎝
⎟ converges to 1, such an M > 0 exists, by
R ⎠
1−
|an |
∞
Corollary 1.3.4. Hence, by comparison test, log 1 − azn +
n=1
Pn (z) converges absolutely on B(0, R). Therefore, a most general
entire function that has the zeroes {an : n ∈ N}, is of the form
∞
z
f (z) = zm exp(g(z)) 1− exp(Pn (z)), ∀z ∈ C,
ak
n=1
for some entire function g.
i i
i i
i i
sin(πz)
Using the easy fact that → 1 as z → 0, we observe that
πz
sin(πz)
1 = lim
z→0 πz
∞ z $ %
z exp(c) 1− exp nz
n=−∞ n
n0
= lim =1
z→0 πz
exp(c)
⇒ =1
π
⇒ exp(c) = π
∞
z z
⇒ sin(πz) = π z 1− exp
n=−∞
n n
n0
∞
z z z z
⇒ sin(π z) = πz 1− 1+ exp exp −
n n n n
n=1
∞
z2
⇒ sin(πz) = π z 1− .
n2
n=1
i i
i i
i i
∞ 4z2
COROLLARY 5.4.17 cos(π z) = 1−
n=1 (2n − 1)2
1 2j
n
√
π = lim √ .
n→∞ n 2j − 1
j=1
i i
i i
i i
1
Proof: Substituting z =in the Euler’s product for sin, we get
2
π π ∞
1
∞
π 4n2 − 1
sin = 1− 2 ⇒1= ,
2 2 4n 2 4n2
n=1 n=1
which implies the first formula.
We rewrite the first formula as
∞
π 4n2
=
2 4n2 − 1
n=1
n
(2j)2 2j − 1
= lim
n→∞ (2j − 1)2 2j + 1
j=1
⎡ ⎤
n
(2j)2
1 3 2n − 1 ⎦
= lim ⎣ ×
· ···
n→∞ (2j − 1)2
3 5 2n + 1
j=1
⎡ ⎤
n
(2j)2 1 ⎦
= lim ⎣ ×
n→∞ (2j − 1) 2 2n + 1
j=1
1 (2j)2
n
1
= lim as lim =1 .
n→∞ 2n (2j − 1)2 n→∞ 1 + 1
2n
j=1
√ 1 n 2j
This implies that π = lim √ .
n→∞ n j=1 2j − 1
Exercise 5.4.19
1. Prove the following:
∞ z2
(a) sinh(π z) = πz 1+ .
n=1 n2
∞ z2
(b) cosh(πz) = 1+ .
n=1 (2n − 1)2
√
∞ (−1)n+1 π
2. Prove that 2= 1+ . Hint: Substitute z = in the
n=1 2n − 1 4
expansion of cos(π z).
π
∞ (2n)2 1
3. Prove that = . Hint: Substitute z = in the
2
n=1 (2n − 1)(2n + 1) 2
sin(πz)
expansion of .
πz
i i
i i
i i
∞ 1 π2 ∞ 1 π4
RESULT 5.4.20 2
= and 4
= .
n=1 n 6 n=1 n 90
∞ x2
If g(x) = log 1 − , then from the proof of Lemma 5.4.4, we see
n=1 n2
sin(π x)
that = exp(g(x)). Therefore,
πx
2 k
∞ ∞
∞ x
sin(πx) x2 n2
log = g(x) = log 1 − = , ∀x ∈ [0, 1).
πx n2 k
n=1 n=1 k=1
Since
2 k
∞
∞ x ∞
n2 |x|2
= log 1 − = g(|x|) < +∞,
k n2
n=1 k=1 n=1
i i
i i
i i
1 1 1
+ − + (π x)6 + · · · . (5.11)
7! 3! × 5! 3 × (3!)3
π6
Exercise 5.4.21 Prove that . Hint: Compare the coefficients of x6 from
945
equations (5.10) and (5.11).
1 1 1
where γ = lim 1+ + + · · · + − ln(n) , which is called the
n→∞ 2 3 n
Euler’s constant.
∞
z z
First, we note that the infinite product 1+ exp − converges
n=1 n n
uniformly on the compact subsets of C. If |z| ≤ r and n > 2r, then
z z z z2 z3 z
log 1 + − = − 2 + 3 − ··· −
n n n 2n 3n n
∞
rk
≤
knk
k=2
∞ k
r
<
nk
k=2
∞
r2 rk
=
n2 nk
k=0
i i
i i
i i
r2 r −1 r 1
= 1 − (as 0 < < < 1)
n2 n n 2
r2 r −1
= 2 2 (as 1 − < 2).
n n
∞
Hence, log 1 + 1n − z
n converges uniformly on compact subsets
n=1
∞
z z
of C. Equivalently, the infinite product 1+ exp − converges
n=1 n n
∞
z z
uniformly on the compact subsets of C. If G(z) = 1+ exp − ,
n=1 n n
then (z) = z exp(γ1z)G(z) , and hence, is well defined on C \ {−n : n ∈ N}.
1. (z + 1) = z(z), ∀z ∈ C,
2. (n) = (n − 1)!, ∀n ∈ N,
√
3. 12 = π ,
Proof:
∞
z z
1. Let G(z) = 1+ exp − , ∀z ∈ C. Hence, the zeroes of
n=1 n n
G are −1, −2, −3, . . . . This implies that the zeroes of G(z − 1) are
0, −1, −2, −3, . . . . Therefore, from Theorem 5.4.14, we have
G(z − 1) = z exp(φ(z))G(z), ∀z ∈ C,
i i
i i
i i
Now,
∞
∞
1 1 1 1 1
− = −1+ −
z−1+n n z z−1+n n
n=1 n=2
∞
1 1 1
= −1+ −
z z+m m+1
m=1
∞
1 1 1
= −1+ −
z z+n n
n=1
∞
1 1
+ −
n n+1
n=1
∞
1 1 1
= −1+ −
z z+n n
n=1
1
+ lim 1−
n→∞ n+1
∞
1 1 1
= + − .
z z+n n
n=1
i i
i i
i i
1 1 1
= lim + + ··· +
ln (N + 1) − 1 +
N→∞ 2 3 N
N +1 1 1 1
= lim ln (N) + ln − 1 + + + ··· +
N→∞ N 2 3 N
1 1 1
= lim ln (N) − 1 + + + · · · + ,
N→∞ 2 3 N
as lim ln N+1
N = ln(1) = 0. Thus, C = γ , the Euler’s constant.
N→∞
Thus, we have
G(z − 1) = z exp(γ )G(z), ∀z ∈ C.
From the definition of and G, we can easily write that
1
(z) =
z exp(γ z)G(z)
1
(z + 1) =
(z + 1) exp(γ (z + 1))G(z + 1)
1
=
(z + 1) exp(γ (z + 1))G(z)(z + 1)−1 exp(−γ )
1
=
exp(γ z)G(z)
= z(z).
i i
i i
i i
1
Substituting z = , we get
2
1 1 π 1 √
1− = ⇒ = π.
2 2 1 2
we get
−1 ∞
(z) −1 1
= −γ + + .
(z) z z+n n
n=1
Therefore,
⎛ ⎞
1
z +
d (z) d ⎜
⎜ 2 ⎟ ⎟
+
dz (z) dz ⎝ 1 ⎠
z+
2
∞ ∞
1 1
= +
(z + n)2 2
n=0 n=0 z + 1 + n
2
∞ ∞
1 1
=4 +
(2z + 2n)2 (2z + 2n + 1)2
n=0 n=0
∞
1
=4
(2z + m)2
m=0
1 d (2z)
=4×
2 dz (2z)
d (2z)
=2 .
dz (2z)
i i
i i
i i
1
z +
(z) 2 (2z)
+ =2 + a,
(z) 1 (2z)
z+
2
1
log ((z)) + log z + = log ((2z)) + az + b,
2
1
(z) × z + = exp(az + b)(2z). (5.13)
2
1
Substituting z = and z = 1 in equation (5.13), we get
2
1 a a √
(1) = exp + b (1) ⇒ + b = log( π ), (5.14)
2 2 2
√
3 π
(1) = exp(a + b)(2) ⇒ a + b = log . (5.15)
2 2
√
a π √
Equation (5.15)−equation (5.14) ⇒ = log − log( π )
2 2
= − log(2) ⇒ a = −2 log 2.
1
Using this value in equation (5.15), we get b = log(2) + log(π).
2
Therefore, from equation (5.13), we get
1 1
(z) z + = exp (−2 log(2))z + log(2) + log(π) (2z)
2 2
√
= 2−2z+1 π (2z).
Thus, we get
√ 1
π (2z) = 22z−1 (z) z + .
2
i i
i i
i i
N zN !
Example 5.5.3 Prove that (z) = lim , ∀z ∈
N→∞ z(z + 1) · · · (z + N)
C\{−1, −2, −3, . . .}.
Solution:
For z ∈ C\{0, −1, −2, −3, . . .}, we have
N z exp (z ln(N)) N
n z
= lim exp − · exp
N→∞ n z z+n n
n=1 n=1
Nz
N
n
= lim ·
N→∞ z z+n
n=1
N zN !
= lim .
N→∞ z(z + 1) · · · (z + N)
In the following example, we justify that has some other integral rep-
resentation. However, this justification depends on dominated convergence
theorem. Hence, the following example is for those who are familiar with the
dominated convergence theorem.
∞
Example 5.5.4 Prove that (z) = tz−1 exp(−t) dt, ∀z = (x, y) ∈ C with
0
x > 0.
Solution:
First, we claim that the above integral exists for all z = (x, y) ∈ C with x > 0.
Since exp(−t) ≤ 1, ∀t ∈ [0, ∞], we have
i i
i i
i i
i i
i i
i i
n n
t
Gn (z) = tz−1 1 − dt
n
0
1
= n z
sz−1 (1 − s)n ds
0
(by using the change of variable t = ns)
⎧
⎨ 1 n z+1
Gn+1 (z + 1), n ≥ 2
= z n−1 .
⎩ 1
n=1
z(z+1) ,
(by applying integration by parts)
and hence,
nz n!
G(z) = lim , ∀z ∈ C with Re z > 0.
n→∞ z(z + 1)(z + 2) · · · (z + n)
∞
Remark 5.5.5: The integral tz−1 exp(−t) dt does not exist for z ∈ C with
0
Re z < 0.
Indeed, if z = x + iy with x < 0 and > 0, then
∞ 1
1− x
|t z−1
exp(−t)| dt ≥ tx−1 exp(−t) dt = exp(−1) → ∞ as → 0.
x
0
i i
i i
i i
6
Residue Calculus
6.1 RESIDUE
Definition 6.1.1 Let f be an analytic function except at a singularity ‘a’.
The residue of f at ‘a’ is defined to be a complex number R, which makes
R
f (z) − as the derivative of an analytic function on B(a, δ)\{a} for some
z−a
δ > 0.
Proof: Since a is an isolated singularity, then there exists δ > 0 such that f is
analytic on (Cl B(a, δ))\{a}. For 0 < r < δ circle |z − a| = r
, let C denote the
1 R
and R = f (z) dz. We claim that f (z) − dz = 0 for every
i2π C γ z−a
closed curve γ in B(a, δ)\{a}.
287
i i
i i
i i
288 Residue
which implies
⎛ ⎞
R dz ⎠
f (z) − dz = m⎝ f (z) dz − R
z−a z−a
γ
⎛C C ⎞
dz ⎠
= m ⎝i2π R − R
z−a
C
= m (i2π R − Ri2π ) = 0.
Hence, by Result 4.1.30, F is the derivative of an analytic function.
1
Thus, residue of f at a exists, and (Res f )(a) = R = f (z) dz.
i2π C
R S
If there are R, S ∈ C such that f (z) − and f (z) −
z−a z−a
are derivatives of the analytic functions and on B(a, δ)\{a}
and B(a, )\{a}, respectively, then obviously, R − S ∈ C and
S−R
( − ) (z) = in B(a, min{δ, }) \ {a}. Hence, by Corollary
z−a
4.1.29, we get
S−R
dz = 0 ⇒ (R − S)i2π = 0 ⇒ R = S.
z−a
C
Thus, the lemma follows.
i i
i i
i i
(z − a)n+1
Proof: We know that (z − a)n is the derivative of inside
n+1
B(a, r)\{a}. Using the fact that
1 1
cn n 1 c−n n 1
lim sup = lim sup |cn | n and lim sup = lim sup |c−n | n ,
n→∞ n+1 n→∞ n→∞ −n + 1 n→∞
∞ cn
we conclude that (z − a)n converges in the neighbourhood of a
n=−∞ n +1
n−1
∞
(except at a) on which cn (z − a)n+1 converges (cf. the proof of Theorem
n=−∞
n−1
2.3.2). Hence,
⎛ ⎞
∞ ∞
c1 d ⎜ cn ⎟
f (z) − = cn (z − a)n = ⎝ (z − a)n+1 ⎠ ,
z − a n=−∞ dz n=−∞ n + 1
n−1 n−1
1 d m−1
lim m−1 (z − a)m f (z)
(m − 1)! z→a dz
∞
1 d m−1
= lim (z − a) m
cn (z − a) n
(m − 1)! z→a dzm−1 n=−m
i i
i i
i i
290 Residue
∞
1 d m−1
= lim m−1 cn−m (z − a) n
which is equal to c−1 if m = 1.
(m−1)! z→a dz
n=0
∞
m−2
1
= lim cn−m (n − ν) (z − a)n−m+1 if m ≥ 2
(m − 1)! z→a
n=m−1 ν=0
1
= (m − 1)!c−1
(m − 1)!
= c−1
= (Res f )(a).
z
2. If f (z) = , then −1 and 2 are the simple poles of f . Then,
(z + 1)(z − 2)
z 1
(Res f )(−1) = lim (z + 1)f (z) = lim =
z→−1 z→−1 z − 2 3
z 2
and (Res f )(2) = lim (z − 2)f (z) = lim = .
z→2 z→2 z + 1 3
2z + 3
3. If f (z) = , then −2 is a double pole of f , and 3 is a
(z + 2)2 (z − 3)
simple pole of f . Then,
d d 2z + 3
(Res f )(−2) = lim [(z + 2) f (z)] = lim
2
z→−2 dz z→−2 dz z−3
−9 −9
= lim =
z→−2 (z − 3)2 25
2z + 3 9
and (Res f )(3) = lim (z − 3)f (z) = lim = .
z→3 z→3 (z + 2)2 25
i i
i i
i i
z2 + 2
4. If f (z) = , then −2 is the only pole of f of order 3. Then,
(z + 2)3
1 d2
(Res f )(−2) = lim [(z + 2)3 f (z)]
2 z→−2 dz2
1 d2 2
= lim 2
z +2
2 z→−2 dz
1
= × 2 = 1.
2
i i
i i
i i
292 Residue
d z+1
= lim
z→i dz (z + i)2
(z + i)2 − (z + 1)2(z + i)
= lim
z→i (z + i)4
(i2)2 − (i + 1)2(i2)
=
(i2)4
−4 − i4 + 4
=
16
i
= − .
4
1 d z+1
(Res f ) (−i) = lim (z + i)2 2
(2 − 1)! z→−i dz (z + 1)2
d z+1
= lim
z→−i dz (z − i)2
(z − i)2 − (z + 1)2(z − i)
= lim
z→−i (z − i)4
i
= .
4
exp(z)
3. Let f (z) = , ∀z ∈ C. z = −3 is a simple pole, and
(z + 3)(z − 2)2
z = 2 is a pole of the order 3 for f .
exp(z)
(Res f ) (−3) = lim (z + 3)
z→−3 (z + 3)(z − 2)2
exp(z)
= lim
z→−3 (z + 3)(z − 2)2
exp(−3)
= .
25
1 d2 exp(z)
(Res f ) (2) = lim (z − 2)3
(3 − 1)! z→2 dz2 (z + 3)(z − 2)2
2
1 d exp(z)
= lim 2
2 z→2 dz z+3
1 d (z + 2) exp(z)
= lim
2 z→2 dz (z + 3)2
i i
i i
i i
1 π π
4. Let f (z) = , ∀z (2k − 1) , k ∈ Z. Clearly, z =
cos(z) 2 2
π 1
is a simple pole. Hence, (Res f )(0) = lim z − =
π 2 cos(z)
z→
2
π −1
lim z− = −1.
π 2 sin(z − π )
z→
2 2
i i
i i
i i
RESULT 6.1.9 If f and g are analytic functions on a region such that
f f (z0 )
f (z0 ) 0 and g has a simple zero at z0 , then Res (z0 ) = .
g g (z0 )
f
Proof: By assumption, it is clear that has a simple pole at z0 . Hence, by
g
Lemma 6.1.4,
f f (z)
Res (z0 ) = lim (z − z0 )
g z→z0 g(z)
f (z)
= lim
z→z0 g(z)
z−z0
f (z)
= lim
z→z0 g(z)−g(z0 )
z−z0
lim f (z)
z→z0
=
lim g(z)−g(z
z−z0
0)
z→z0
f (z0 )
= .
g (z0 )
Hence, the result follows.
i i
i i
i i
n
which implies that f (z) dz = i2π WN (γ , aj )(Res f )(aj ).
γ j=1
Example 6.2.2 Evaluate the following integrals using Cauchy’s residue the-
orem:
z+1
1. dz.
|z|=3 z + 7z + 10
2
3z2 − 2z + 1
2. dz.
|z|=4 z2 − 2z
z+3
3. dz.
|z|=3 (z2 + 6z + 8)(z2 + 1)
i i
i i
i i
z2 + 2
4. dz.
|z−1|=3 (z − 2)2 (z + 4)
z
5. dz.
|z+1−i|=6 (z2 − 2z − 15)(z2 + 1)
sin(z)
6. 2 + 1)2
dz, where γ is the rectangle with vertices (−1, 0), (1, 0),
γ (z
(1, 2), and (−1, 2).
4z + 1
7. dz.
|z|=1 (z + 1)(z + 2)
2 2
z3 + 1
8. dz.
|z|=4 z(z + 1)(z + 2)
exp(z) − 1
9. dz.
|z−1|=5 sin(z)
exp(z)
10. dz.
|z|=1 sin(z)
Solution:
z+1
1. Let f (z) = , ∀z ∈ C\{−2, −5}. Clearly, −2 and −5 are
z2 + 7z + 10
the simple poles of f . We note that −2 is enclosed by |z| = 3 and −5
is not enclosed by |z| = 3. Therefore, by Cauchy’s residue theorem, we
get
z+1
dz = i2π (Res f )(−2)
z2 + 7z + 10
|z|=3
= i2π lim (z + 2)f (z)
z→−2
z+1
= i2π lim
z→−2 z+5
−i2π
= .
3
3z2 − 2z + 1
2. Let f (z) = dz, ∀z ∈ C\{0, 2}. Clearly, 0 and 2 are the
z2 − 2z
simple poles of f and both are enclosed by |z| = 4. Therefore, by
Cauchy’s residue theorem, we get
i i
i i
i i
Since
z+3
3. Let f (z) = , ∀z ∈ C\{−2, −4, i, −i}. Clearly,
+ 6z + 8)(z2 + 1)
(z2
−2, −4, i, and −i are simple poles of f , and −2, i, and −i are enclosed
by |z| = 3, but −4 is not enclosed by |z| = 3. Therefore, by Cauchy’s
residue theorem,
z+3
dz
(z2 + 6z + 8)(z2 + 1)
|z|=3
= i2π [(Res f )(−2) + (Res f )(−i) + (Res f )(i)] .
Since
z+3 −1
(Res f )(−2) = lim (z + 2)f (z) = lim = ,
z→−2 (z + 4)(z2 + 1)
z→−2 10
z+3
(Res f )(−i) = lim (z + i)f (z) = lim 2
z→−i z→−i (z + 6z + 8)(z − i)
−i + 3 −11 + i27
= = ,
(7 − i6)(−i2) 170
i i
i i
i i
and
z+3
(Res f )(i) = lim(z − i)f (z) = lim
z→i (z2 + 6z + 8)(z + i)
z→i
i+3 −11 − i27
= = ,
(7 + i6)(i2) 170
we get
z+3
dz
(z2 + 6z + 8)(z2 + 1)
|z|=3
−1 −11 + i27 −11 − i27
= i2π + +
10 170 170
−39
= i2π
170
−39πi
= .
85
z2 + 2
4. Let f (z) = , ∀z ∈ C\{2, −4}. Obviously, 2 is a double
(z − 2)2 (z + 4)
pole and −4 is a simple pole for f . 2 is enclosed by |z − 1| = 3, and
−4 is not enclosed by |z − 1| = 3. Since
d z2 + 2 (z + 4)2z − (z2 + 2) 1
(Res f )(2) = lim = = ,
z→2 dz z+4 (z + 4)2 2
by Cauchy’s residue theorem, we get
z2 + 2 1
dz = i2π [(Res f )(2)] = i2π = iπ .
(z − 2) (z + 4)
2 2
|z−1|=3
z
5. Let f (z) = , ∀z ∈ C\{5, −3, i, −i}. Clearly,
− 2z − 15)(z2 + 1)
(z2
5, −3, i, and −i are poles of f , and |z + 1 − i| = 6 encloses −3, i,
and −i but not 5. By Cauchy’s residue theorem, we have
z
dz
(z − 2z − 15)(z2 + 1)
2
|z+1−i|=6
= i2π [(Res f )(−3) + (Res f )(i) + (Res f )(−i)] .
i i
i i
i i
Since
z 3
(Res f )(−3) = lim = ,
z→−3 (z − 5)(z + 1)
2 80
z
(Res f )(i) = lim 2
z→i (z − 2z − 15)(z + i)
i −8 + i
= = ,
(−16 − i2)(i2) 260
and
z
(Res f )(−i) = lim
z→−i(z2 − 2z − 15)(z − i)
1 −8 − i
= = ,
2(−16 + i2) 260
we have
z
dz
(z2− 2z − 15)(z2 + 1)
|z+1−i|=6
3 −8 + i −8 − i
= i2π + +
80 260 260
3 −8 + i −8 − i
= i2π + +
80 260 260
iπ
= .
52
sin(z)
6. Let f (z) = , ∀z ∈ C\{i, −i}. Clearly, i and −i are double
(z2 + 1)2
poles of f and both are enclosed by γ . Therefore, by Cauchy’s residue
theorem,
sin(z)
dz = i2π [(Res f )(i) + (Res f )(−i)] .
(z2 + 1)2
γ
Since
d sin(z)
(Res f )(i) = lim
z→i dz (z + i)2
(z + i) cos(z) − 2 sin(z)
= lim
z→i (z + i)3
exp(i)
=
−4
i i
i i
i i
and
d sin(z)
(Res f )(−i) = lim
z→−i dz (z − i)2
(z − i) cos(z) − 2 sin(z)
= lim
z→−i (z − i)3
exp(i)
= ,
−4
we have
sin(z) 2 exp(i)
dz = i2π = −iπ exp(i).
(z + 1)
2 2 −4
γ
4z + 1 √ √ √
7. Let , ∀z ∈ C\{i, −i, −i 2, i 2} . Clearly, i, −i, i 2,
+ 1)(z2 + 2)
(z2 √
and −i 2 are simple poles of f , and i and −i are enclosed by |z| = 1.
By Cauchy’s residue theorem, we have
4z + 1
dz = i2π [(Res f )(i) + (Res f )(−i)] .
(z2 + 1)(z2 + 2)
|z|=1
Since
4z + 1
(Res f )(i) = lim
z→i (z + i)(z2 + 2)
1 + i4
=
i2
4−i
=
2
and
4z + 1
(Res f )(−i) = lim
z→−i (z − i)(z2 + 2)
1 − i4
=
−i2
4+i
= ,
2
we get
4z + 1 4−i 4+i
dz = i2π + = i8π .
(z2 + 1)(z2 + 2) 2 2
|z|=1
i i
i i
i i
z3 + 1
8. Let f (z) = , ∀z ∈ C\{0, −1, −2}. Clearly, 0, −1, and
z(z + 1)(z + 2)
−2 are simple poles of f and all are enclosed by |z| = 4. Therefore, by
Cauchy’s Residue theorem,
z3 + 1
dz = i2π [(Res f )(0) + (Res f )(−1) + (Res f )(−2)] .
z(z + 1)(z + 2)
|z|=4
Since
z3 + 1 1
(Res f )(0) = lim = ,
z→0 (z + 1)(z + 2) 2
z3 + 1
(Res f )(−1) = lim = 0,
z→−1 z(z + 2)
and
z3 + 1 −7
(Res f )(−2) = lim = ,
z→−2 z(z + 1) 2
we get
z3 + 1 1 −7
= i2π +0+ = −3iπ .
z(z + 1)(z + 2) 2 2
|z|=4
exp(z) − 1
9. Let f (z) = , ∀z ∈ C\{nπ : n ∈ Z}. Clearly, 0 is a removable
sin(z)
singularity for f as
∞ zk
exp(z) − 1 k=1 k !
lim f (z) = lim = lim = 0,
z→0 z→0 sin(z) z→0 ∞ (−1)k−1 z2k−1
k=1 (2k − 1)!
d
lim f (z) = ∞ and sin(z) = cos(nπ) = (−1)n 0.
z→nπ dz nπ
i i
i i
i i
(Res f )(0) = 0.
(z − π )(exp(z) − 1)
(Res f )(π ) = lim
z→π sin(z)
= (−1)(exp(π) − 1).
Therefore,
exp(z) − 1
dz = i2π (1 − exp(π )).
sin(z)
|z−1|=4
exp(z)
10. Let f (z) = , ∀z ∈ C\{0}. Clearly, 0 is the simple pole of f and is
sin(z)
enclosed by |z| = 1. Therefore, by Cauchy’s residue theorem, we have
exp(z)
dz = i2π (Res f )(0)
sin(z)
|z|=1
z
= i2π lim
z→0 sin(z)
= i2π .
Note that so far we have evaluated integrals over simple curves. Now, we
evaluate an integral over a closed curve, which is not simple.
2z + 3
Example 6.2.3 Evaluate dz, where γ is given in the
γ (z4 − 1)(z2 − 5z + 6)
following diagram.
i i
i i
i i
−2 −1 −0 1
−i
2z + 3
Let f (z) = , ∀z ∈ C\{−1, 0, i, −i, 1, 2, 3}.
− 1)(z3 − 5z2 + 6z)
(z4
Clearly, 0, 1, −1, i, −i, 2, and 3 are poles of f . From the diagram, we find1
that WN(γ , −1) = −3, WN(γ , 0) = 2, WN(γ , i) = −1, WN(γ , −i) = 0,
WN (γ , 1) = −1, WN (γ , 2) = 0, and WN (γ , 3) = 0.
Therefore, by Cauchy’s residue theorem, we get
2z + 3
dz
(z4 − 1)(z3 − 5z2 + 6z)
γ
= i2π [WN(γ , −1) (Res f )(−1)
+ WN(γ , 0) (Res f )(0) + WN(γ , i) (Res f )(i)
+ WN(γ , −i) (Res f )(−i) + WN(γ , 1) (Res f )(1)
+ WN(γ , 2) (Res f )(2) + WN(γ , 3) (Res f )(3)]
= i2π [−3 (Res f )(−1) + 2 (Res f )(0) − (Res f )(i) − (Res f )(1)] .
Since
2z + 3 1
(Res f )(−1) = lim = ,
z→−1 (z − 1)(z2 + 1)(z3 − 5z2 + 6z) 48
1 WN (γ , a) is found by counting the number of times the curve γ winds around a, and the sign
is given as + (−) if it winds in anti-clockwise sense (clockwise sense).
i i
i i
i i
2z + 3 −1
(Res f )(0) = lim = ,
z→0 (z4− 1)(z − 5z + 6)
2 2
2z + 3 1 + i5
(Res f )(i) = lim 2 = ,
z→i (z − 1)(z + i)(z − 5z + 6z)
3 2 40
and
2z + 3 5
(Res f )(1) = lim = ,
z→1 (z + 1)(z2 + 1)(z3 − 5z2 + 6z) 8
we get
2z + 3 1 −1 1 + i5 5
dz = i2π −3 × +2× − −
(z − 1)(z2 − 5z + 6)
4 48 2 40 8
γ
1 −1 1 + i5
= i π −8− −5
4 2 5
π i137
= 1− .
4 10
Example 6.2.4 Evaluate tan(z) dz using Cauchy’s residue theorem.
|z−1|=2
Solution:
We note that the only pole of tan inside |z−1| = 2 is π2 , which is a simple pole.
π
Hence, by Cauchy’s residue theorem, tan(z) dz = i2π (Res tan) . As
|z|=2 2
sin(z) π π
tan(z) = , sin = 1 0, and cos = 0, then by applying
cos(z) 2 π 2
π sin 1
Result 6.1.9, we get (Res tan) = 2π = = −1. Therefore,
2 cos −1
2
tan(z) dz = −i2π .
|z|=2
i i
i i
i i
−iπ
Answer:
2 exp(1)
z2 − 1
2. dz.
|z−1|=2 z2 − 6z + 8
Answer: −i3π
z2 + 3z
3. dz.
|z−1|=3 (z − 1)2 (z − 2)
Answer: i2π
z−1
4. dz.
|z|=2 (z + 1) (z − 3)
2
−iπ
Answer:
4
z−1
5. dz.
|z+1|=5 (z + 2)2 (z + 3)2
Answer: 0
exp(z)
6. dz.
|z|=2 (z + 1) (z − 3)
2
−i5π
Answer:
8 exp(1)
z+4
7. dz.
|z−1−i|=3 (z − 7)(z − 2)
−i12π
Answer:
5
sin(z)
8. dz.
|z|=1 z2
Answer: i2π
z3
9. dz.
|z|=4 z2 − 2z + 2
Answer: i4π
i i
i i
i i
exp(z)
10. dz.
|z|=3 z(z + 1 + i)
2
Answer: πi e−(1+i) (−1 + 2i) − i
f
Proof: We know that f is also analytic wherever f is analytic. Hence, is
f
analytic on except at zj ’s and wk ’s. By Cauchy’s residue theorem, we have
1 f (ζ ) f
dζ = WN (γ , z) Res (z)
i2π f (ζ ) z
f
γ
f
+ WN (γ , w) Res (w), (6.1)
w
f
where z varies through the distinct zeroes of f and w varies through the dis-
tinct poles of f .
f
Claim. If f has a zero of order pz at z, then Res (z) = pz .
f
Since z is a zero for f of order pz , by Result 5.2.3, there exists an analytic
function g on a neighbourhood B(z, r) of z such that
Hence, by applying Theorem 4.5.12, we get that log(g) is analytic in B(z, r),
and hence,
i i
i i
i i
pz g (ζ )
= +
g(ζ ) g(ζ )
pz d
= + [log(g(ζ ))] .
g(ζ ) dζ
f pz
This implies that − is the derivative of an analytic function on B(z, r),
f ζ − z
f
and hence, Res (z) = pz .
f
f
Claim. If f has a pole of order qw at w, then Res (w) = qw .
f
Since w is a pole for f of order qw , by Result 5.2.11, there exists an analytic
function g on a neighbourhood B(w, r) of w such that
f (ζ ) = (ζ − w)−qw g(ζ ) and h(ζ ) 0, ∀ζ ∈ B(w, r) \ {w}.
Hence, log(h) is analytic in B(z, r) (by Theorem 4.5.12) and hence,
f (ζ ) (−qw )(ζ − w)−qw −1 g(ζ ) + (ζ − w)−qw h (ζ )
=
f (ζ ) (ζ − w)−qw h(ζ )
−qw
h (ζ )
= +
h(ζ ) h(ζ )
−qw d
= + [(log(h(ζ )))] .
h(ζ ) dζ
f
This implies that Res (w) = −qw . Using these observations in equation
f
(6.1), we get
1 f (ζ ) f f
dζ = WN (γ , z) Res (z) + WN (γ , w) Res (w)
i2π f (ζ ) z
f w
f
γ
(where z varies through distinct zeroes of f and w
varies through distinct poles of f .)
= WN (γ , z)pz + WN (γ , w)qw
z w
n m
= WN (γ , zj ) − WN (γ , wk ).
j=1 k=1
i i
i i
i i
Proof: Let z1 , z2 , . . . , zn be the distinct zeroes of f such that the order of the
zero of f at zj is 2pj for some pj ∈ N, ∀j = 1, 2, . . . , , n, and let w1 , w2 , . . . , wm
be the distinct poles of f such that the order of the pole of f at wk is 2qk for
some qk ∈ N, ∀k = 1, 2, . . . , m. By argument principle, for every closed
curve γ in ∗ := \ {zj , wk : 1 ≤ j ≤ n; 1 ≤ k ≤ m}, we have
n m
1 f (ζ )
dζ = 2pj WN(γ , zj ) − 2qj WN(γ , wk ).
2π i f (ζ )
γ j=1 k=1
1
f (ζ )
Therefore, 2 f (ζ ) dζ is an integral multiple of 2π i. For an arbitrarily fixed
γ
z0 ∈ ∗ , we define
⎛ ⎞
1 f (ζ ) ⎠
g(z) = exp ⎝ f (z0 ) + dζ , ∀z ∈ ∗ , (6.2)
2 f (ζ )
σz
where σz is a curve in ∗ joining z0 and z, and f (z0 ) is any one value of
square root of f (z0 ) 0. The definition of g is well defined. Indeed, if z0 and
z are connected by two curves σz and τz , respectively, in ∗ , then σz ∪ (−τz )
f (ζ )
is a closed curve in ∗ . Therefore, 12 f (ζ ) dζ is an integral multiple
σz ∪(−τz )
f (ζ )
f (ζ ) dζ = 1, which implies that
1
of 2π i, and hence, exp 2
σz ∪(−τz )
⎛ ⎞ ⎛ ⎞
1 f (ζ ) ⎠ 1 f (ζ )
exp ⎝ dζ = exp ⎝ dζ ⎠ .
2 f (ζ ) 2 f (ζ )
σz τz
Thus, the definition of g is independent of the curve σz and hence well de-
fined. Next, we claim that g is analytic on ∗ and g = gf2f on ∗ . Let z ∈ ∗
be arbitrary. If r > 0 is such that B(z, r) ⊂ ∗ and if w ∈ B(z, 3r ), then
z ∈ B(w, 2r ∗
3 ) ⊂ B(z, r) ⊂ . Hence, we can write
⎛ ⎞
⎜ 1
f (ζ ) 1
f (ζ ) ⎟
g(z) = exp ⎝ f (z0 ) + dζ + dζ ⎠
2 f (ζ ) 2 f (ζ )
σw Lw,z
i i
i i
i i
(where Lw,z
⎛ is the line segment
⎞ joining w and z)
⎜1 f (ζ ) ⎟
= g(w) exp ⎝ dζ ⎠ .
2 f (ζ )
Lw,z
f (z) 2r
Since f (z) is analytic on B(w, 3 ), from the proof of Cauchy’s theorem for
(z) f (ζ )
disc, ff (z) is the derivative of the analytic function f (ζ ) dζ in B(w, 2r
3 ).
Lw,z
Therefore, g is analytic on ∗ and
⎛ ⎞
⎜1 f (ζ )
⎟ 1 f (z) g(z)f (z)
g (z) = g(w) exp ⎝ dζ ⎠ = , ∀z ∈ ∗ .
2 f (ζ ) 2 f (z) 2f (z)
Lw,z
(g(z0 ))2
Since f (z0 ) = 1 and
d (g(z))2 f (z)2g(z)g (z) − (g(z))2 f (z)
=
dz f (z) (f (z))2
g(z)f (z)
f (z)2g(z) 2f (z) − (g(z))2 f (z)
=
(f (z))2
= 0,
g2
f is a constant on ∗ and the constant value is 1. Thus, f = g2 on ∗ . It
is also easy to see that lim g2 (z) = lim f (z) = f (zj ) = 0 and lim g2 (z) =
z→zj z→zj z→wk
lim f (z) = ∞, where g is the required meromorphic function on such that
z→wk
g2 = f.
i i
i i
i i
Proof: First, we note that f 2 is analytic at 0 (since lim f 2 (z) = c2 ), and hence,
z→0
we can write
∞
f 2 (z) = an zn , ∀z ∈ B(0, r) for some r > 0. (6.3)
n=0
i i
i i
i i
and
g(z) = (z − α1 )(z − α2 ) . . . (z − αs )ψ(z),
where φ and ψ are analytic functions on with the following
properties:
i i
i i
i i
n−m = (n − s) − (m − s)
ψ (z)
φ
= dz
ψ
γ φ (z)
= 0 (by applying Case(1)).
whence we can apply Rouche’s theorem and we get that f and g have same
number of zeroes enclosed by |z| = 1. As three zeroes of f are enclosed by
|z| = 1, including multiplicites, g also has three zeroes enclosed by |z| = 1.
3
Example 6.3.7 Prove that one zero of z5 + 15z + 1 belongs to B 0, and
2
four zeroes of z5 + 15z + 1 belong to AN23/2 (0).
Solution: Let f (z) = z5 and g(z) = z5 + 15z + 1, ∀z ∈ C. Then we note that
on |z| = 2. Hence, we can apply the Rouche’s theorem (Theorem 6.5) and
we get that f (z) and g(z) have same number of zeroes in B(0, 2) including
multiplicities. We know that f has five zeroes (which are obviously 0, 0, 0, 0,
and 0) in B(0, 2) including multiplicities, whence g has five zeroes in B(0, 2).
As g has exactly five zeroes, every zero of g belongs to B(0, 2). On the other
3
hand, if we put f1 (z) = 15z, then on the circle |z| = , we have
2
243 45
| f1 (z) − g(z)| = |z5 + 1| ≤ |z|5 + 1 = +1< = |15z| = | f1 (z)|,
32 2
i i
i i
i i
Example 6.3.8 Prove that z6 + 5z4 + 3 has exactly 2 zeroes in AN31 (0).
Solution: Let f (z) = z6 and g(z) = z6 + 5z4 + 3, ∀z ∈ C. As for |z| = 3,
we can apply Rouche’s theorem, and we conclude that f and g have same
number of zeroes enclosed by |z| = 3, including multiplicities. As f has
six zeroes inside B(0, 3), g also has six zeroes inside B(0, 3). If f1 (z) = 5z4 ,
∀z ∈ C, then
Therefore, again using Rouche’s theorem, we get that the number of zeroes
of f1 enclosed by |z| = 1 (including multiplicities) is same as that of g. Since
f1 has four zeroes enclosed by |z| = 1, g has two zeroes enclosed by |z| = 1.
Thus, g has two zeroes inside the annulus AN31 (0).
Exercise 6.3.9
2. Prove that one zero of z4 − 7z − 1 lies in B(0, 1) and three zeroes are in
AN21 (0).
3. Find the number of zeroes of z8 − 3z5 + 7z3 − 2 in the annulus AN21 (0).
Answer: 5.
4. Prove that all zeroes of z7 − 5z3 + 12 lie in the annulus AN21 (0).
i i
i i
i i
and when θ varies along the interval [0, 2π], z varies along the circle |z| = 1.
Therefore, by Cauchy’s residue theorem,
2π
1 1 1 1 i
S(cos(θ), sin(θ )) dθ = S z+ , z− − dz
2 z i2 z z
0 |z|=1
= G(z) dz
|z|=1
= i2π × sum of the residues of G
at the poles enclosed by |z| = 1.
i i
i i
i i
g (r)
−r O r x
i i
i i
i i
g (r)
−r O r x
i i
i i
i i
M
R(z) exp(iz) dz ≤ |R(z)| exp(−Im z) |dz| ≤ |dz|
γ (r) γ (r) r2 γ (r)
(using Theorem 4.1.27)
M Mπ
= 2
πr = → 0 as r → ∞.
r r
Hence, allowing r → ∞ in equation (6.6), we get
∞
lim R(z) dz = R(x) exp(ix) dx.
r→∞
C(r) −∞
∞
Thus, R(x) exp(ix) dx = lim R(z) exp(iz) dz = i2π × sum of the
−∞ r→∞
C(r)
residues of H at the poles in the upper half-plane.
g1
−X1 X2
i i
i i
i i
Since R has a simple zero at ∞, there exist M > 0 and K > 0 such that
|zR(z)| ≤ M , ∀|z| > K . Therefore, for X2 > K we get
Y
R(z) exp(iz) dz = R(z) exp(iz) dy where z = X2 + iy
γ2 0
Y
≤ |R(z)| exp(−y) dy
0
Y
1
≤ M exp(−y) dy
|z|
0
Y
M
≤ exp(−y) dy (since |z| ≥ X2 )
X2
0
M
= (− exp(−Y ) + 1)
X2
M
< → 0 as X2 → ∞.
X2
i i
i i
i i
−X1
R(z) exp(iz) dz = R(z) exp(iz) dx where z = x + iY
γ3 X2
X2
≤ |R(z)| exp(−Y ) dx
−X1
X2
1
≤ M exp(−Y ) dx
|z|
−X1
X2
M
≤ exp(−Y ) dx (as |z| ≥ Y )
Y
−X1
M
= (X2 + X1 ) → 0 as Y → ∞
Y
for each fixed X1 > 0 and X2 > 0.
Hence, applying lim , lim , and lim in equation (6.8) and using claims
X1 →∞ X2 →∞ Y →∞
(1), (2), and (3), the theorem follows.
i i
i i
i i
−X1 + iY X2 + iY
g3
g4 g2
g6
g5 g1
−d 0 d
−X1 X2
i i
i i
i i
If 0 < δ < δ0 , then |S(z)| ≤ M for all z with |z| ≤ δ , for some M > 0.
Therefore, using Theorem 4.1.27, we get
S(z) dz ≤ |S(z)| |dz| ≤ M |dz| = Mπ δ → 0 as δ → 0.
γ6 γ6 γ6
Hence, applying lim , lim , lim , and lim in equation (6.9) and using
δ→0 X1 →∞ X2 →∞ Y →∞
−δ ∞
claims (1), (2), (3), and (4), we get lim + R(x) exp(ix) dx − iβπ =
δ→0 −∞ δ
i2π× sum of the residues of H at poles in the upper half-plane. Since
⎛ −δ ∞⎞
∞
lim ⎝ + ⎠ R(x) exp(ix) dx = P.V. R(x) exp(ix) dx,
δ→0
−∞ δ −∞
i i
i i
i i
1 1 1
The simple poles of G(z) = are −3 and − . Since − is
+ 10z + 33z2 3 3
enclosed by |z| = 1, by Cauchy’s residue theorem, we get
1 i2 1
−i2 dz = − dz
3z + 10z + 3
2 3 1
|z|=1 |z|=1 (z + 3) z +
3
1
= −i2i2π (Res G) −
3
1
= 4π lim z + G(z)
1 3
z→−
3
4π 1
= lim
3 1 z+3
z→−
3
4π 3 π
= × = .
3 8 2
1
Since is an even function,
5 + 3 cos(θ )
π π
1 1 1 1 π π
dθ = dθ = × = .
5 + 3 cos(θ ) 2 5 + 3 cos(θ ) 2 2 4
0 −π
2π dθ
Example 6.4.7 Evaluate .
0 2 + cos(θ )
1
Solution: Let S(cos(θ ), sin(θ )) = . This integral is of Type 1.
2 + cos(θ )
Hence,
2π
dθ
= i2π × sum of the residues of G
2 + cos(θ )
0
at the poles enclosed by |z| = 1,
i i
i i
i i
where
i 1 1 1 1
G(z) = − S z+ , z−
z 2 z i2 z
i 1
= − ×
z 1 1
2+ z+
2 z
−i2
= .
z2 + 4z + 1
√
The poles of√G are −2 ± 3, which are the roots of z2 + 4z + 1 = 0. We note
that −2 + 3 lies inside |z| = 1 and the other root lies out side the circle.
Hence,
2π √
dθ
= i2π × (Res G)(−2 + 3)
2 + cos(θ )
0
√
= i2π lim √ (z + 2 − 3)G(z)
z→−2+ 3
−i2
= i2π lim √ √
z→−2+ 3 z+2+ 3
2π
= √ .
3
∞ cos(3x)
Example 6.4.8 Evaluate dx.
−∞ (x + 1)
2 2
∞ exp(i3x)
Solution: First, we find dx. As this integral is of Type 3(a), we
−∞ (x + 1)
2 2
have
∞
cos(3x)
dx = i2π (Res H)(i),
(x2 + 1)2
−∞
exp(i3z)
where H(z) = , ∀z ∈ C. Now,
(z2 + 1)2
d 2 exp(i3z)
(Res H)(i) = lim (z − i) 2
z→i dz (z + 1)2
d exp(i3z)
= lim
z→i dz (z + i)2
i i
i i
i i
∞ sin(x)
Example 6.4.9 Evaluate dx.
−∞ x
∞ exp(ix)
Solution: First, we evaluate dx. This integral comes under Type
−∞ x
3(c). Hence,
∞
exp(ix)
dx = i2π × sum of the residues of H at the poles
x
−∞
1
in the upper half-plane + (Res H)(0),
2
exp(iz)
where H(z) = , ∀z ∈ C. We note that H has only one simple pole
z
exp(iz)
at 0. Hence, (Res H)(0) = lim z = lim exp(iz) = 1. Therefore,
z→0 z z→0
∞ exp(ix) ∞ sin(x) ∞ exp(ix)
dx = iπ ⇒ dx = Im dx = π .
−∞ x −∞ x −∞ x
Answer: π
i i
i i
i i
2π cos(2θ ) dθ
2. .
0 5 + 4 sin(θ)
Answer: − π6
π dθ
3. .
0 4 + cos2 (θ )
π
Answer: √
5
2π dθ
4. .
0 (2 + cos(θ ))2
4π
Answer: √
3 3
2π cos(3θ )dθ
5. .
0 5 − cos(2θ)
Answer: 0
∞ dx
6. .
−∞ x2 + 2x + 6
π
Answer: √
2 5
∞ x2
7. dx.
−∞ (x2 + 4)3
π
Answer:
32
∞ dx
8. .
−∞ x4 + 1
π
Answer: √
2
∞ x sin(x)
9. dx.
0 x2 + 9
π
Answer: exp(−3)
2
i i
i i
i i
∞ x exp(i2x)
10. dx.
−∞ 1 + x
2
Answer: iπ exp(−2)
∞ cos(x)
11. dx.
0 (1 + x )
2 3
7π
Answer:
16 exp(1)
∞ sin2 (x)
12. dx.
−∞ (x + 4)
2
π
Answer: (1 − exp(−4))
4
∞
Proof: First, we evaluate x2α+1 R(x2 ) dx. Using the branch of log defined
−∞
on 2 = C\{(0,
y) : y ≤ 0} by the unique value of log whose imaginary
π 3π
part lies in − , , if z2α is defined on 2 by exp(2α log(z)), then z2α is
2 2
analytic on 2 . Consider the closed curve
γ = L(−ρ,0),(−δ,0) ∪ γ1 ∪ L(δ,0),(ρ,0) ∪ γ2
i i
i i
i i
g2
g1
−r −d O d r x
|z2α+1 R(z2 )| ≤ M1 , ∀z ∈ A.
i i
i i
i i
≤ M1 π δ 2 → 0 as δ → 0.
z2α+1 R(z) dz ≤ |z2α+1 ||R(z2 )| |dz|
γ2 γ1
|dz|
≤ ρ 2α+1
|z4 R(z2 )|
|z4 |
γ2
ρ 2α+1
≤ M2 |dz|
ρ4
γ2
ρ 2α+1
= M2 πρ
ρ4
= M2 πρ 2α−2 → 0 as ρ → ∞,
as 2α − 2 < 0.
Allowing ρ → ∞ and δ → 0 on both sides of equation (6.10), we get
∞ 2α+1 2
z R(z ) dz = i2π × sum of residues of at poles in the upper half-
−∞
plane. Now,
∞ 0 ∞
z 2α+1 2
R(z ) dz = z 2α+1
R(z ) dz +
2
z2α+1 R(z2 ) dz
−∞ −∞ 0
0 ∞
= (−t) 2α+1
R(t ) (−dt) +
2
z2α+1 R(z2 ) dz
∞ 0
∞
= 1 + (−1)2α+1 t2α+1 R(t2 ) dt.
0
i i
i i
i i
∞
1
=2× z2α+1 R(z2 ) dz
1 − exp(i2απ )
−∞
i4π
= sum of residues of
1 − exp(i2απ )
at poles in the upper half-plane.
π
THEOREM 6.5.2 (Type (5)): log(sin(x)) dx = −π log(2).
0
π
Proof: First, we show that log(1−exp(i2z)) dz = 0. We know that principal
0
branch of log which is defined on 1 = C\{(x, 0) : x ≤ 0} by the unique value
of log whose imaginary part lies in (−π , π), so that log is analytic on 1 . If
z = x + iy, then
1 − exp(i2z) is real and ≤ 0
γ = L(δ,0),(π−δ,0) ∪ γ1 ∪ γ2 ∪ γ3 ∪ γ4 ∪ γ5
i i
i i
i i
g3
iY p + iY
g4 g2
i g5 g1 p + id
0 p
d p−d
(i2z)n ∞
1 − exp(i2z) ∞ (i2z)zn−1
k=1 n!
1. We know that =− =− → −i2
z z k=1 n!
as z → 0. Since modulus function is continuous on C and log is con-
1 − exp(i2z)
tinuous at 2, we get log → log 2 as z → 0. Therefore,
z
there exists M > 0 such that
1 − exp(i2z)
log ≤ M, ∀z ∈ B(0, r) for some r > 0.
z
i i
i i
i i
i i
i i
i i
Y
= − log(1 − exp(i2(π + iy))) dz
δ
= − log(1 − exp(i2z)) dz.
γ2
π
log(1 − exp(i2z)) dz = 0.
0
Since
we have
π π
log(sin(x)) dx = log(1 − exp(i2x)) dx
0 0
π
− [log(2) + log(−i) + log(exp(ix))] dx
0
π π2
= 0 − π log(2) − i π + i
2 2
= −π log(2).
∞ dx
Example 6.5.3 Evaluate √ .
0 x(x + 1)
1
Solution: Let f (z) = √ . Then, z = −1 is a pole for f . Note that if
√ z(z + 1)
g(z) = z, then g is not a single-valued function. Hence, first we define a
branch of g as follows so that it is analytic at −1.
√ θ
We take = C \ {(x, 0) : x ≥ 0} and define z = r exp i , 1/2
2
where θ ∈ (0, 2π). As there is a unique argument of z in this range (0, 2π),
i i
i i
i i
CR
−1 C D
B A
Cr
Although it seems that there is a gap between the line segments AB and CD,
actually they are lying on the positive real axis. However, they are given dif-
ferent notation because their parametric equations are given in different ways
as AB = {t exp(i2π) : R ≤ t ≤ r} and CD = {t exp(i0) : r ≤ t ≤ R}. √
(The difference between AB and CD is that the image of AB under z is on
the negative real axis, whereas the same for CD is on the positive real axis.)
Since −1 is the only pole of f inside γ , by Cauchy’s residue theorem, we
have
dz
√ = i2π (Res f )(−1) = i2π (−i) = 2π ,
z(z + 1)
γ
as π
dz
(Res f )(−1) = lim √ = (−1)1/2 = exp −i = −i.
z→−1 z 2
Since
1 dz
2π = √ = + + + √ , (6.12)
z(z + 1) z(z + 1)
γ CR AB Cr CD
we find
r r
dz exp(i2π) dt dt
√ = =−
z(z + 1) t1/2 exp(iπ)(t exp(i2π ) + 1) t1/2 (t + 1)
AB R R
i i
i i
i i
and
R R
dz exp(i0) dt dt
√ = = .
z(z + 1) t1/2 exp(i0)(t exp(i0) + 1) t1/2 (t + 1)
CD r r
2π
dz Ri exp(iθ ) dθ
√ = θ
z(z + 1) R1/2 exp i 2 (R exp(iθ ) + 1)
CR 0
2π
R dθ
≤
R1/2 |R exp(iθ ) + 1|
0
2π
−1/2 dθ
≤ R
| exp(iθ ) + R1 |
0
1
≤ R−1/2 2π → 0 as R → ∞
1− 1
R
0
dz ri exp(iθ ) dθ
√ = θ
z(z + 1) r1/2 exp i 2 (r exp(iθ ) + 1)
Cr 2π
2π
dθ
≤ r 1/2
1−r
0
1
= r 1/2
2π → 0 as r → ∞.
1−r
Using these observations in equation (6.12), we get
R R
dt dt dz
2 = 2 + lim √
t (t + 1)
1/2 t (t + 1) r→0
1/2 z(z + 1)
r r Cr
i i
i i
i i
R dt
Hence, = π.
r t1/2 (t+ 1)
∞ exp(−t2 )
LEMMA 6.5.4 For any c ∈ C with Re c > 0, dt < ∞.
−∞ 1 + exp(−2ct)
∞
exp(−t2 )
≤ dt
|1 + exp(−2ct)|
−∞
∞
1
≤ exp(−t2 ) dt
μ
−∞
∞
2
= exp(−t2 ) dt
μ
0
⎛ ⎞
1 ∞
2⎝
= exp(−t2 ) dt + exp(−t2 ) dt⎠
μ
⎛0 ⎞
1
∞
2⎝
< 1 + exp(−t) dt⎠
μ
1
(as t < 1 ⇒ exp(−t2 ) < 1 and t > 1 ⇒ exp(−t2 ) < exp(−t))
i i
i i
i i
√ π exp(−z2 )
Solution: Let c = iπ = (1 + i). If f (z) = , then f has
2 1 + exp(−2cz)
poles at every w ∈ {w ∈ C : exp(−2cw)+1 = 0}. Then, 2cw = i(2k −1)π ⇒
i(2k − 1) i(2k − 1)
w= for some k ∈ Z. Moreover, each is a simple pole
2c 2c
d
of f because it is a simple zero of 1 + exp(−2cz) as (1 + exp(−2cz)) 0.
dz
Now, we consider the following rectangle R with vertices −ρ, r, r + i π2 ,
and −ρ + i π2 , which passes through the point c = π
2 (1 + i), where
r > 1 + π2 and ρ > 0 (later, both of these positive real numbers r and ρ
will be tending to +∞).
−r + i p r+i p
2 c 2
c
2
−r r
0
π
Since Im z ∈ 0, , ∀z ∈ Int R, the only pole of f in Int R is 2c . In other
2
words, 2c is the only pole of f enclosed by the boundary ∂R of R. Therefore,
by Cauchy’s residue theorem, we have
c
exp(−z2 ) √
dz = i2π(Res f ) = π, (6.13)
1 + exp(−2cz) 2
∂R
i i
i i
i i
1
√ (1 − i)
2 −i
= √ = √ .
2π(1 + i) 2 π
By direct calculation, we also get
√π
r 2
f (z) dz = f (t) dt + i f (r + it) dt
∂R −ρ 0
−ρ 0
π
+ f t+i dt + i f (−ρ + it) dt. (6.14)
2 √
r π
2
We note that
√π √π
2 2
| exp(−(r2 − t2 + i2rt))|
f (r + it) dt ≤ dt
|1 + exp(−2a(r + it))|
0 0
√π
2
exp(−(r2 − t2 ))
≤ √ dt
1 − exp(− 2π (r − t))
0
[ as |1 − exp(−2a(r + it))| ≥ 1 − | exp(−2a(r + it))|
√
= 1 − Re exp(−2a(r + it)) = 1 − exp(− 2π (r − t)) ]
√π
2
exp(t2 )
≤ exp(−r2 ) √ dt
1 − exp(− 2π )
0
π π
[ as r > 1 + and t ≤ ⇒r−t >1
2 2
1 1
⇒ √ ≤ √ ]
1 − exp(− 2π(r − t)) 1 − exp(− 2π )
√
π π2
exp −r2 +
≤ √2 dt
1 − exp(− 2π)
0
exp −r2 + π
π
≤ √ 2 → 0 as r → +∞.
2 1 − exp(− 2π )
i i
i i
i i
Therefore, applying lim and lim on both sides of equation (6.14) and
r→+∞ ρ→+∞
using equation (6.13), we get
∞
√ π
π = lim lim f (z) dz = f (t) − f t+i dt. (6.15)
r→+∞ ρ→+∞ 2
∂R −∞
∞
= f (s + c) ds
−∞
and
exp(−(t + c)2 )
f (t + c) =
1 + exp(−2c(t + c))
exp(−(t2 + c2 + 2ct)
=
1 + exp(−2ct − 2c2 )
− exp(−t2 ) exp(−2ct)
=
1 + exp(−2ct)
(as c2 = π i, exp(−iπ ) = −1, exp(−i2π ) = 1.)
− exp(−t2 )(1 + exp(−2ct) exp(−t2 )
= +
1 + exp(−2ct) 1 + exp(−2ct)
= − exp(−t2 ) + f (t),
we obtain
∞ ∞
√
π = f (t) dt − (− exp(−t2 ) + f (t)) dt
−∞ −∞
i i
i i
i i
∞ ∞ ∞
= f (t) dt + exp(−t ) dt −
2
f (t) dt (using Lemma 6.5.4)
−∞ −∞ −∞
∞
= exp(−t2 ) dt.
−∞
i i
i i
This page is intentionally left blank
i i
7
Some Interesting
Theorems
Proof: We know that the Cartesian coordinates and polar coordinates are
related by x = r cos(θ ) and y = r sin(θ ). As
∂u ∂u ∂x ∂u ∂y ∂u ∂u
= + = cos(θ ) + sin(θ ),
∂r ∂x ∂r ∂y ∂r ∂x ∂y
we get
∂u ∂u ∂u ∂u ∂u
r = r cos(θ ) + r sin(θ ) = x +y ,
∂r ∂x ∂y ∂x ∂y
which implies
∂ ∂ ∂ ∂u ∂u
r r = r x +y
∂r ∂r ∂r ∂x ∂y
∂ ∂u ∂u ∂ ∂u ∂u
= x x +y +y x +y
∂x ∂x ∂y ∂y ∂x ∂y
341
i i
i i
i i
i i
i i
i i
Using
Im ( f1 f2 dz) = Im ( f1 d( f2 )) = Im ((u1 + iv1 )(du2 + idv2 ))
= u1 dv2 + v1 du2
= u1 dv2 + d(u2 v1 ) − u2 dv1
and using Lemma 7.1.2, we obtain d(u2 v1 ) = 0, and hence,
γ
(u1 dv2 − u2 dv1 ) = 0.
γ
1 2π
Hence, we get u(z) = u(z + r exp(iθ )) dθ by equating the real parts on
2π 0
both sides.
i i
i i
i i
Solution:
Let u1 = log(r) and u2 = u. Applying Theorem
7.1.1, u1 is a harmonic
∂ ∂u1 ∂ 2 u1 ∂ 1
function, as r r + = r r + 0 = 0. Therefore, by
∂r ∂r ∂θ 2 ∂r r
Theorem 7.1.3, we have (u1 dv2 − u2 dv1 ) = 0 for every cycle γ ∼ 0 in
γ
, where v1 and v2 are harmonic conjugates of u1 and u2 , respectively. For a
harmonic function u and its harmonic conjugate v, we first convert u dv into
the polar form (on |z| = r) as follows. Using C–R equations, we have
∂v ∂v ∂u ∂u
dv = dx + dy = − dx + dy.
dx dy dy dx
As in the proof of Theorem 7.1.1, from x = r cos(θ ) and y = r sin(θ ), we
have
∂u ∂u ∂u ∂u ∂u
r dθ = r cos(θ ) + r sin(θ ) dθ = x dθ + y dθ,
∂r ∂x ∂y ∂x ∂y
and using the fact that r is constant on |z| = r, we also have
i i
i i
i i
∂u
Therefore, u dθ − log(r) dθ is a constant, say b. (Reason: After
r
|z|=r |z|=r ∂r
integrating these functions with respect to θ , we get only a function of r and
by previous equation, it is also independent of r.)
By a same argument using dv = 0, whenever γ ∼ 0 in B(0, r)\{0} (see
γ
∂u
Lemma 7.1.2.), we get r dθ , which is a constant, say a.
|z|=r ∂r
1 a b
Thus, we get u dθ = A log(r) + B, where A = and B = .
2π |z|=r 2π 2π
2π
1
u(z) = u(z + r exp(iθ )) dθ.
2π
0
i i
i i
i i
2π
1
u(z) = u(z + r exp(iθ )) dθ
2π
0
⎛ a
b
1 ⎝
= u(z + r exp(iθ)) dθ + u(z + r exp(iθ )) dθ
2π
0 a
⎞
2π
+ u(z + r exp(iθ)) dθ ⎠
b
⎛ a ⎞
b 2π
1 ⎝ u(z) + u(w0 )
≤ u(z) dθ + dθ + u(z) dθ ⎠
2π 2
0 a b
u(z)
< (a + b − a + 2π − b) (since u(z) + u(w0 ) < 2u(z))
2π
= u(z),
i i
i i
i i
−1 r(rη + z)
w= (η) = ,
r + zη
i i
i i
i i
i i
i i
i i
2π
1 r2 − |z|2
u(z) = lim u(sz) = u (r exp(iθ )) dθ .
s→1 2π |r exp(iθ ) − z|2
0
Remark 7.1.11: The Poisson’s formula can also be written in the following
ways.
a+2π
1 r2 − |z|2
u(z) = u (r exp(iθ )) dθ, (a ∈ R). (7.1)
2π |r exp(iθ) − z|2
a
2π
1 r exp(iθ ) + z
= Re u(r exp(iθ )) dθ . (7.2)
2π r exp(iθ) − z
0
⎛ ⎞
1 ⎜ w + z u(w) ⎟
= Re ⎝ dw⎠ , ∀z ∈ B(0, r). (7.3)
2π w − z iw
|w|=r
2π
log(| f (0)|) = log(| f (r exp(iθ ))|) dθ.
0
n 2π
r
log(| f (0)|) = − log + log(| f (r exp(iθ ))|) dθ .
|aj |
j=1 0
i i
i i
i i
n 2π
r
log(|cm |) + m log(r)=− log + log(| f (r exp(iθ ))|) dθ,
|aj |
j=1 0
Proof:
2π
1
log(| f (0)|) = log( f (r exp(iθ ))) dθ .
2π
0
⎛ − ν log r
log(| f (0)|) ⎞
⎜ | f (0)| ⎟
= log ⎜
⎝
⎟
⎠
ν
|0 − r exp(iθk )|
k=1
= log(| f1 (0)|)
2π
1
= log( f1 (r exp(iθ))) dθ
2π
0 ⎛ ⎞
2π
1 ⎜ | f (r exp(iθ ))| ⎟
= log ⎜
⎝
⎟ dθ
⎠
2π
ν
0 |r exp(iθ ) − r exp(iθk )|
k=1
i i
i i
i i
2π
1
= log(| f (r exp(iθ ))|)
2π
ν
0
− [log(r) + log(| exp(iθ ) − exp(iθk )|)] dθ.
k=1
Thus, we have
2π
1
log(| f (0)|) = log(| f (r exp(iθ ))|)
2π
ν
0
− log(| exp(iθ ) − exp(iθk )|) dθ . (7.4)
k=1
2π
We claim that log(| exp(iθ) − exp(iθk )|) = 0, ∀k ∈ {1, 2, . . . , ν}.
0
Using the fact that exp(it) is a periodic function on R of period 2π
(Result 2.4.20) and using Lemma 2.4.23, we obtian
2π
log(| exp(iθ) − exp(iθk )|)
0
2π
= log(|1 − exp(i(θ − θk )|))
0
θk+2π
= log(|1 − exp(it)|)
θk
(using the change of variable t = θ − θk )
2π
= log(|1 − exp(it)|) dt. (7.5)
0
Since
1
log(|1 − exp(it)|) = log((1 − cos(t))2 + sin2 (t))
2
1
= log(1 + cos2 (t) − 2 cos(t) + sin2 (t))
2
1
= log (2(1 − cos(t)))
2
i i
i i
i i
1
= [log(2) + log(1 − cos(t))]
2
1
= [log(2) + log(2 sin2 (t/2))]
2
= log(2) + log(sin(t/2)),
the right-hand side of equation (7.5) is equal to
2π π
[log(2) + log(sin(t/2))] dt = 2π log(2) + 2 log(sin(u)) du
0 0
(by the change of variable u = t/2)
= 2π log(2) − 2π log(2)
(by Theorem 6.5)
= 0.
Using these observations in equation (7.4), we get
2π
1
log(| f (0)|) = log(| f (r exp(iθ ))|) dθ.
2π
0
n r2 − a z
j
Case 2. Let f2 (z) = f (z) , ∀z ∈ Cl B(0, r). Clearly, f2 becomes
j=1 r(z − aj )
an analytic function on Cl B(0, r), and it has no zeroes in B(0, r). If
|z| = r, then let z = r exp(iθ ) for some θ ∈ [0, 2π], and hence,
r2 − a z r exp(iθ )(r exp(−iθ ) − a ) r exp(−iθ ) − a
j
=
j j
r(z − aj ) r(r exp(iθ ) − aj ) = r exp(iθ ) − a = 1.
j
2π
1
= log(| f2 (r exp(iθ ))|) dθ
2π
0
2π
1
= log(| f (r exp(iθ ))|) dθ .
2π
0
i i
i i
i i
n 2π
r 1
log(| f3 (0)|) = − log + log |f3 (r exp(iθ ))| dθ .
|aj | 2π
j=1 0
n 2π
r 1
= − log + log | f (r exp(iθ ))| dθ ,
|aj | 2π
j=1 0
since | f3 | = |f | on | z| = r. If the coefficient of zm in the Taylor’s se-
ries expansion of f is cm , then we have f3 (0) = cm rm , and hence, the
right-hand side of the above equation becomes log(|cm |) + m log(r).
i i
i i
i i
2π
1 r exp(iθ ) + z
= Re log |F(r exp(iθ ))| dθ
2π r exp(iθ ) − z
0
2π
1 r exp(iθ ) + z
= Re log | f (r exp(iθ ))| dθ .
2π r exp(iθ ) − z
0
Proof:
1. If
1 u(ζ ) 1 u(ζ )
f1 (z) = dζ , f2 (z) = dζ , ∀z ∈ γ c ,
2πi ζ −z 2π i ζ (ζ − z)
γ γ
then as in the proof of Theorem 4.3.2, we get that f1 and f2 are analytic
on γ c .
(Note that the only difference between the current situation and Theo-
rem 4.3.2 is that here, we have a piecewise continuous function u on γ ,
whereas φ is a continuous function on γ in Theorem 4.3.2. However,
to follow the proof of Theorem 4.3.2, we need continuity of φ to get
i i
i i
i i
i i
i i
i i
i i
i i
i i
which is called the Schwarz’s formula. The following theorem is also due to
Schwarz.
i i
i i
i i
Proof: Note that uz : [0, 2π) → [0, 2π) is a bijection, with uz = u−1
z .
B = exp(iq )
A*z
z
0
B*z = exp(iq *z )
By the intersecting chord theorem1 , we obtain that the area of the rectangle
with sides |z − A| and |z − A∗z | is equal to the area of the rectangle with sides
|z−B| and |z−B∗z |. From the diagram, |z−A| is the length of the line segment
joining z and A, which is equal to |z|+|A| = |z|+1 and |z−A∗ | = |A∗ |−|z| =
1 − |z|, and hence, we have
As z lies between the points exp(iθ) and exp(iθz∗ ), we get that z − exp(iθ ) and
z − exp(iθz∗ ) lie on the same line but in opposite directions from z. Therefore,
z − exp(iθ )
it follows that < 0. If arg (z − exp(iθ )) = ϕ and arg (z −
z − exp(iθz∗ )
exp(iθz∗ )) = ψ , then we have ϕ − ψ = π (if the arguments are written in
[0, 2π)). Moreover, we have
i i
i i
i i
Using the fact that dθ and dθz∗ are the usual Lebesgue measures on [0, 2π ]
(positive measures) and using equation (7.6), we have
dθ dθ z − exp(iθ ) 1 − |z|2
= = =
dθz∗ dθ ∗ z − exp(−iθ ∗ ) |z − exp(−iθ ∗ )|2 .
z z z
i i
i i
i i
i i
i i
i i
ρ−r ρ 2 − r2 ρ 2 − r2 ρ 2 − r2 ρ+r
= ≤ ≤ = .
ρ+r (ρ + r) 2 |a + ρ exp(it) − z0 | 2 (ρ − r) 2 ρ−r
ρ−r ρ 2 − r2
u (a + ρ exp(it)) ≤ u (a + ρ exp(it))
ρ+r |a + ρ exp(it) − z0 |2
ρ+r
≤ u (a + ρ exp(it)) ,
ρ−r
2π
ρ−r 1
u (a + ρ exp(it)) dt
ρ + r 2π
0
2π
1 ρ 2 − r2
≤ u (a + ρ exp(it)) dt
2π |a + ρ exp(it) − z0 |2
0
2π
ρ+r 1
≤ u (a + ρ exp(it)) dt, ∀t ∈ [0, 2π).
ρ − r 2π
0
2π
ρ−r 1 ρ 2 − r2 ρ+r
u(a) ≤ u (a + ρ exp(it)) dt ≤ u(a).
ρ+r 2π |a + ρ exp(it) − z0 |2 ρ−r
0
If v(w) = u(a + w), ∀w ∈ B(0, ρ), then using the Poisson’s formula for v as
in equation (7.1), we obtain that the integral in the above inequality is equal
to v(z0 ) = u(a + z0 ), and hence, we have
ρ−r ρ+r
u(a) ≤ u(a + z0 ) ≤ u(a), ∀z0 ∈ C with |z0 − a| = r.
ρ+r ρ−r
i i
i i
i i
i i
i i
i i
2π
1
un (z) = un (z + r exp(it)) dt, ∀n ≥ n0 .
2π
0
2π
1
u(z) = u(z + r exp(it)) dt.
2π
0
v(z) if z ∈ + ∪ σ
Proof: Define V : → R by V (z) = ∀z ∈ .
−v(z) if z ∈ −
Immediately, it follows that V (z) = −V (z), ∀z ∈ ; indeed,
i i
i i
i i
i i
i i
i i
π
1 r2 − |z − x0 |2
= V (x0 + r exp(iθ )) idθ
2π |x0 + r exp(iθ ) − z|2
−π
π
1 r2 − |z − x0 |2
= V (x0 + r exp(−iϕ)) idϕ
2π |x0 + r exp(−iϕ) − z|2
−π
(using x0 ∈ R)
π
1 r2 − |z − x0 |2
= [−V (x0 + r exp(iϕ))] idϕ
2π |x0 + r exp(iϕ) − z|2
−π
1 r2 − |z − x0 |2 V (ω)
= − dω
2π |ω − z|2 (ω − x0 )
|ω−x0 |=r
= −(Pϕ ◦ T)(z).
i i
i i
i i
i i
i i
i i
i i
i i
i i
Proof: Fix w ∈ . Choose δ(w) > 0 such that Cl B (w, δ(w)) ⊆ . First,
we show that
we can find a subsequence that converges uniformly on
δ(w)
B w, . Now applying Theorem 5.1.5, we can expand each fm as
2
∞ (n)
fm (w) δ(w)
fm (z) = (z − w) , ∀z ∈ Cl B w,
n
.
n! 2
n=0
δ(w)
If M = sup | fm (z)| : z ∈ Cl B w, , m ∈ N , then M < +∞ and by
2
Cauchy’s estimate, we have
f (n) (w) M2n
m
≤ , ∀m ∈ N.
n! (δ(w))n
i i
i i
i i
n!2n−1
1 (n) (n)
< and (g − g )(w) < , ∀k, j ≥ N2 , ∀1 ≤ n ≤ N1 .
2N1 4M j k (N1 + 1)(δ(w))n
δ(w)
For j, k ≥ N2 and z ∈ B w, , we have
2
∞
(g − g )(n)
j k
|gj (z) − gk (z)| = (z − w)n
n!
n=0
N1
∞ (g − g )(n)
(gj − gk )
(n)
n j k n
≤ (z − w) + (z − w)
n! n!
n=0 n=N1 +1
N1 ∞
2n−1 2M
≤ |z − w| +n
|z − w|n
(N1 + 1)(δ(w))n (δ(w))n
n=0 n=N1 +1
∞
2M 1
< + N +1
2 2 1 2n
n=0
i i
i i
i i
4M
= + N +1
2 2 1
< + = .
2 2
Since C is second countable (Theorem 1.4.48), we can choose a sequence
∞ δ(wj )
of points wj ∈ such that = ∪ B wj , 2 . Given a sequence of
j=1
functions ( fm ), which is uniformly bounded on compact sets, we find a
subsequence (φn, 1 ) of ( fm ), which converges uniformly on B w1 , δ(w2 1 ) .
Next, we find a subsequence (φn, 2 ) of (φn, 1 ), which converges uniformly on
B w2 , δ(w2 2 ) , and so on. The diagonal subsequence (φn, n ) converges uni-
δ(w )
formly on B wj , 2 j , ∀j ∈ N. Thus, (φn, n ) converges uniformly on each
compact subset of because every compact subset of is contained in a
δ(w )
finite union of the B wj , 2 j .
i i
i i
i i
f2 (z) = z − a, ∀z ∈ ,
1
f3 (z) = , ∀z ∈ ,
( f1 ◦ f2 )(z) + ξ0
then f3 : → Cl B 0, −1 is analytic and injective on . For a fixed z0 ∈ ,
if
( f3 (z) − f3 (z0 ))
f (z) = , ∀z ∈ ,
3
then f is analytic on , and f () is a region containing 0, by open mapping
theorem (Theorem 4.5.4) and by Theorem 1.6.29. Since f3 is injective on ,
f is also injective on . As f3 : → Cl B 0, −1 , we have f : →
Cl B 0, 23 r . Thus, f : → G is an analytic bijective map, where G =
f () B(0, 1) is a region containing 0.
Let H = {h : G → B(0, 1) : h be analytic injective with h(0) = 0}.
Let a ∈ G\{0} be arbitrarily fixed. Using f ∈ H , we have H ∅. If
i i
i i
i i
i i
i i
i i
Hence, by Corollary 4.5.10, we get (F1 ◦ F2−1 )(z) = cz, ∀z ∈ B(0, 1) for some
c ∈ C with |c| = 1. For a given z ∈ , let w = F2 (z) so that F1 (z) = (F1 ◦
F2−1 )(w) = cw = cF2 (z). Using arg (F1 (z0 )) = arg (F2 (z0 )) and F1 (z0 ) =
cF2 (z0 ), we get arg (c) = 0. Thus, c = 1. Hence the result follows.
i i
i i
i i
Thus, zn K , ∀n ≥ NK .
Conversely, if for each compact set K ⊆ , there exists NK ∈ N such
that zn K , ∀n ≥ NK , then for a given z ∈ , we choose r > 0 such
that Cl B(z, r) ⊆ . Clearly, Cl B(z, r) is a compact subset of , then by
assumption, there exists > 0 and N ∈ N such that zn Cl B(z, r), ∀n ≥ N ,
and hence, |z − zn | > . Thus, zn → ∂ as n → ∞.
Case 1. H1 ⊆ and H2 ⊆ c .
Case 2. H1 ⊆ and H2 ⊆ .
i i
i i
i i
H1 g
H2
H1 H2
g
Ω Ω
Proof: As the first statement of this theorem is equivalent to the second state-
ment, we prove only the first statement. Obviously, it follows that if γ is a
one-sided free boundary arc, then γ has a one-sided free boundary point.
Conversely, assume that γ has a one-sided free boundary point ζ . By defi-
nition of free boundary arc, there exists r > 0 such that B(ζ , r) ∩ ∂ is a
diameter of B(ζ , r), and by the definition of one-sided free boundary point
of , if H1 and H2 are the two half-open discs determined by the diameter,
i i
i i
i i
then let H1 ⊆ , H2 ⊆ (Cl )c . Then, for every ξ ∈ B(ζ , r) ∩ ∂ and for
every s ∈ (0, r − d(ζ , ξ )), we have B(ξ , s) ∩ ∂, which is a diameter of B(ξ , s).
Furthermore, one of the two half-open discs determined by this diameter lies
in H1 ⊆ and the other half-open disc lies in H2 ⊆ (Cl )c .
H1
z
x
H2
Proof:
i i
i i
i i
Re log( f (zn )) → 0 as n → ∞.
Ω
F(m)
F(z) F(z1) = F(z2)
R1
Ω2
m (0, 0)
R2
Ω1
z2 g z1
i i
i i
i i
1. S maps the straight line passing along γ onto the real axis,
2. the entire region is mapped into the upper half-plane,
3. its unique finite pole lies in the complement of Cl .
Remark 7.5.11: We point out that according to the notations in the proof of
previous theorem, |F0 (z)| > 1, ∀z ∈ 0 \ Cl .
We justify the above statement for the case γ ⊆ R, and the general case
follows by applying a similar technique used in Case 2 of proof of previous
i i
i i
i i
theorem. If z ∈ B(x, rx ) ∩ (Cl )c for some x ∈ γ , then F0 (z) = Fx (z), and
by the construction of Fx ( from the previous proof), Fx (z) = exp(−iG(z)) for
some analytic function G on B(x, rx ), thereby satisfying
|F(z)| = exp(−iG(z))
= | exp(−iG(z1 ))|
= | exp(−iG(z1 ))|
= | exp(−i[i log( f (z1 ))])|
= | exp(− log(f (z1 )))|
= exp(Re [− log(f (z1 ))])
= exp(log(|f (z1 )|)−1 )
= | f (z1 )|−1 > 1, as f : → B(0, 1) and z1 ∈ .
i i
i i
i i
Proof: If γj = {(1 − t)zj + tzρ(j) : (0, 1)} and ρ(j) is as in the statement of
the theorem, ∀j = 1, 2, . . . , n, then γj is a one-sided free boundary arc. (Note
that γj does not include the end points zj and zρ(j) , whereas Lzj ,zρ(j) includes
the end points zj and zρ(j) .) Then, applying the previous theorem repeatedly,
f can be extended as an analytic function G on a region containing and
γj , ∀j = 1, 2, . . . , n and G(γj ) ⊆ ∂B(0, 1). Note that G is not defined at zj ,
∀j = 1, 2, . . . , n. Fix j ∈ {1, 2, . . . , n} arbitrarily. Now, find rj > 0 such that
zk B(zj , rj ) if j k ∈ {1, 2, . . . , n}. Denote B(zj , rj ) ∩ by Sj . As z − zj 0,
∀z ∈ Sj , log(z − zj ) is defined as an analytic function on Sj , and hence, if
w3
w4
F w5 0 w2
1
w6 w1
z5
a5 z4 y3
z6
a6 a4
f Y3
a1 Ω
z1
a3
a2 S3 1/a3
z3 0
r3 r3
z2
i i
i i
i i
i i
i i
i i
wr ( j)
0
1
wj
(In the above diagram, actually there is a single circular arc from wj and wρ(j) ,
which is traversed in both directions. To make it clear, we have drawn the arc
using double lines.)
We claim that the second case is not possible because in the second case,
we have
0 = WN (F(∂), 0)
1 dw
=
i2π w
F(∂)
1 f (z)
= dz
i2π f (z)
∂
(by the change of variable w = f (z))
i i
i i
i i
where the integral is defined over any curve joining 0 and w in B(0, 1), βj =
1 − αj , wj = lim f (z), ∀j = 1, 2, . . . , n.
z→zj
Proof: From geometry, we first note that sum of the exterior angles of a
n
closed polygon is 2π . That is, βj = 2. We shall use the notations as in
j=1
the proof of previous theorem. By applying Corollary 7.5.12, repeatedly, for
the the circular arc μj joining wj and wρ(j) (which excludes the end points),
j = 1, 2, . . . , n, there exists a simply connected region U0 containing Cl
B(0, 1) \ {wj : 1 ≤ j ≤ n} and an analytic function F on U0 such that
F = F −1 on B(0, 1).
Next, we prove that the function j (0) 0. Suppose that j (0) = 0.
1/α
Then, by local correspondence theorem, there exist 0 < δ < rj j and 0 <
< 1 such that for each w ∈ B(w, ), there exist at least two distinct points
a, b ∈ B(0, δ) such that j (a) = w = j (b). By Theorem 7.5.10 and by
Remark 7.5.11, we see that
i i
i i
i i
1/αj
|j | = 1 on the diameter of the disc B(0, rj )
and
|j | > 1 on {z ∈ C : Im z < 0}.
where (j−1 (w))αj , and (w−wj )αj are analytic functions on B(wj , tj )\Lwj ,wj −tj ,
α
and Pj j is an analytic function on B(wj , tj ), each of which is defined by using
exp and principal branch of log.
By the previous theorem, F : Cl → Cl B(0, 1) is a homeomorphism
such that F(z) = f (z), ∀z ∈ . Therefore, for each w ∈ B(wj , tj ), first we
choose ζ ∈ B(0, δj ) such that j (ζ ) = w, and hence,
Therefore, we have
i i
i i
i i
We note that the function on the right-hand side of equation (7.10) is an an-
alytic function on B(wj , tj ), whereas the function on the left-hand side is an
analytic function on B(wj , tj ) \ Lwj ,wj −tj . Define Ij : B(wj , tj ) → C by
n
H(w) = F (w) (w − wj )βj , ∀w ∈ V0 . (7.13)
j=1
i i
i i
i i
sp sp sp s p
|w − up | ≤ |w − u| + |u − up | < |w| − 1 + < 1 + r − 1 + ≤ + = sp
2 2 2 2
w ∈ B up , sp , and hence, F is analytic and non-zero at w. Therefore, F is
analytic and nowhere zero on B(0, 1 + s), which is a simply connected region.
Therefore, by Theorem 4.5.12, we can find a suitable branch for log(H), so
that it is analytic on B(0, 1 + s). In particular, with respect to this branch,
arg (H(w)) =Im H(w) is continuous on B(0, 1 + s). If j ∈ {1, 2, . . . , m} and
w ∈ {exp(iθ ) : θj < θ < θρ(j) }, where arg (wj ) = θj and arg (wρ(j) ) = θρ(j) ,
then from equation (7.13), we have
n
arg H(exp(iθ)) = arg (F (exp(iθ ))) + arg (exp(iθ ) − exp(iθj )). (7.14)
j=1
Since
1. arg F (exp(iθ)) is the angle between the tangent of the line segment γj
joining zj and zρ(j) at F(exp(iθ)) and the tangent of the circular arc μj
joining wj = exp(iθj ), wρ(j) = exp(iθρ(j) ) at w = exp(iθ ),
we have
π
arg (F (exp(iθ ))) = ϕj − θ + .
2
As
exp(iθ ) − exp(iθj )
θ + θj θ − θj θ − θj
= exp i exp i − exp −i
2 2 2
θ + θj θ − θj
= exp i 2i sin
2 2
θ − θj θ + θj + π
= 2 sin exp i
2 2
i i
i i
i i
θ−θ θ+θj +π
and 2 sin 2 j > 0, we have arg (exp(iθ ) − exp(iθj )) = 2 .
Consolidating the above observations, we get
π θ + θj + π
n
arg H(exp(iθ)) = ϕj − θ + + βj
2 2
j=1
θ +π
n n
π θj
= ϕj − θ + + βj + βj
2 2 2
j=1 j=1
⎛ ⎞
π n
θj
n
= ϕj − θ + +θ +π + βj ⎝ using βj = 2⎠
2 2
j=1 j=1
π
n
θj
= ϕj + + βj ,
2 2
j=1
n
F (w) = ( f −1 ) (w) = c1 (w − wj )−βj , ∀w ∈ B(0, 1),
j=1
for some c1 ∈ C. Integrating on both sides of the above equation over a curve
joining 0 and w in B(0, 1), we get
w
n
−1
f (w) = c1 (w − wj )−βj dw + c2 , ∀w ∈ B(0, 1),
0 j=1
for some c1 , c2 ∈ C.
i i
i i
i i
mapping from onto the upper half-plane H + , then there exist constants C1
and C2 such that
ζ n−1
−1
f (ζ ) = C1 (ζ − ζj )−βj dζ + C2 ,
i j=1
where the integral is defined over any curve joining i and ζ in H + and βj =
1 − αj , ζj = lim f (z), ∀j = 1, 2, . . . , n, with ζn = ∞.
z→zj
Proof: To find a linear fractional transform, which maps the unit disc onto
the the upper half-plane, we find a linear fractional transform T , which maps
1, −i, i to 1, 0, ∞, respectively, by using the definition of cross ratio as given
below.
w+i1−i w+i
T(w) = (w, 1, −i, i) = = −i .
w−i1+i w−i
i
As T maps the unit circle onto the real axis and T(0) = −i = i, it maps
−i
+
B(0, 1) onto H . We find wj ∈ ∂B(0, 1) such that T(wj ) = ζj , j = 1, . . . , n−1,
and we know that if wn = i, then T(wn ) = ∞, and T −1 ◦ f : → B(0, 1)
is a bijective analytic map such that lim (T −1 ◦ f )(z) = wj , j = 1, 2, . . . , n.
z→zj
Hence, by applying the proof of previous theorem, we obtain
n
c1 (w − wj )−βj = ( f −1 ◦ T) (w) = ( f −1 ) (T(w))T (w),
j=1
c1
n
( f −1 ) (T(w)) =
(w − wj )−βj .
T (w)
j=1
Now, using the change of variable ζ = T(w) in the last equation, we also get
n−1
( f −1 ) (ζ ) = c1 (T −1 ) (ζ )(T −1 (ζ ) − T −1 (ζn ))−βn (T −1 (ζ ) − T −1 (ζj ))−βj .
j=1
(7.15)
i i
i i
i i
ζ
n
−1
f (ζ ) = C1 (ζ − ζj )−βj dζ + C2 , ∀ζ ∈ H + ,
i j=1
for some C1 , C2 ∈ C.
i i
i i
i i
Example 7.5.17 Find a Reimann mapping that maps the upper half-plane to
the infinite strip {(x, y) : x ≥ 0, −1 ≤ y ≤ 1}.
p
a1 =
2
−i
1
We take z1 = −i, z2 = i; α1 = α2 = and ζ1 = −1, ζ2 = 1. (The choice
2
of ζ1 and ζ2 are arbitrary but distinct.) Then, by Schwarz–Christoffel formula
(Theorem 7.5.15), the required conformal mapping g is given by
g(ζ ) = C1 (ζ +1)−1/2 (ζ −1)−1/2 dζ +C2 = C1 (ζ 2 −1)−1/2 +C2 , ∀ζ ∈ H + ,
i i
i i
i i
8
Elliptic Functions
391
i i
i i
i i
φ
Hence, the integrals of the form 1 − k 2 sin2 (u) du are called elliptic inte-
0
grals. More specifically, they are called the elliptic integrals of second kind.
There are two more kinds of elliptic integrals, namely, elliptic integrals of
first kind and elliptic integrals of third kind.
φ du
Elliptic integrals of first kind : .
0 1 − k 2 sin2 (u)
φ du
Elliptic integrals of third kind: .
0 1 + n sin2 (u) 1 − k 2 sin2 (u)
By making the change of variable x = sin u, we get
x dx
Elliptic integrals of first kind : √ √ ,
0 1√− x2 1 − k 2 x2
x 1 − k 2 x2 dx
Elliptic integrals of second kind : √ , and
0 1 − x2
x dx
Elliptic integrals of third kind: √ √ √ .
0 1 + nx 1 −√x2 1 − k 2 x2
2
A general elliptic integral is of the form R(x, P(x)) dx, where R is a ra-
tional function in two variables, and P is a cubic or quadratic polynomial.
It is interesting to see from the literature on elliptic integrals that every el-
liptic integral can be expressed by using those standard forms, together with
elementary integrals.
An elliptic curve is a curve in C, which is given by the following equation;
and hence, ℘a,b can be viewed as the inverse of an elliptic integral. We shall
show that every elliptic function is a rational function of ℘a,b and its deriva-
tive, for some suitable choices of a’s and b’s. From the literature of elliptic
curves, we see that every elliptic curve can be parameterized by using ℘a,b .
This is the history behind the similarity in the nomenclature. It should be
noted that ellipse is not an elliptic curve. Now, we are going to discuss the
elliptic functions in this chapter.
i i
i i
i i
Proof:
1. First, we prove that if a ∈ Pf and m ∈ Z, then ma ∈ Pf . For m = 0,
clearly ma = 0 ∈ Pf . For m > 0, using the condition a ∈ Pf repeatedly,
we get
f (z + ma) = f (z + (m − 1)a + a)
= f (z + (m − 1)a)
= f (z + (m − 2)a)
..
.
= f (z + a)
= f (z), ∀z ∈ C
f (z + ma) = f (z + ma + a)
= f (z + (m + 1)a)
i i
i i
i i
= f (z + (m + 2)a)
..
.
= f (z − a)
= f (z − a + a)
= f (z), ∀z ∈ C.
f = {z ∈ C : f is analytic at z} = C \ Sf ,
i i
i i
i i
i i
i i
i i
Proof:
1. As f and g are meromorphic functions, for each w ∈ C, the Laurent
series expansions of f and g around w have at most finite number of
non-zero terms in their singular parts. Hence, the Laurent series ex-
pansion of f + g also has at most finite number of terms in its singular
part. If w is a pole of exactly one of f and g, then w is a pole of f + g. If
w is a pole of both f and g, then either w is a removable singularity of
f + g (and hence, it may be assumed as a regular point after removing
its singularity) or w is a pole of f + g. Hence, Sf +g ⊆ Sf ∪ Sg . Thus,
f + g is a meromorphic function on C. As f and g have same periods,
for each a ∈ Pf = Pg ,
( f + g)(z + a) = f (z + a) + g(z + a) = f (z) + g(z), ∀z ∈ C \ Sf +g .
We know that for each w ∈ C, there exists r > 0 such that f is analytic
on B(w, r) \ {w}. Therefore
lim ( f + g)(z) = lim ( f + g)(z + a) = lim ( f + g)(ζ ), ∀w ∈ C.
z→w z→w ζ →w+a
i i
i i
i i
1 1 1 1
(z + a) = = = (z).
g g(z + a) g(z) g
1
Thus, g is an elliptic function. Finally, using this with (3), we conclude
f
that g is also an elliptic function.
Exercise 8.1.8 If f and g are elliptic functions such that Pf Pg , then prove
that Pf +g = Pfg = P f = Pg .
g
i i
i i
i i
f (z) − f (z0 )
f (z0 ) = lim
z→z0 z − z0
f (z + a) − f (z0 + a)
= lim (using a ∈ Pf )
z→z0 (z + a) − (z0 + a)
f (w) − f (z0 + a)
= lim (by substituting w = z + a)
w→z0 +a w − (z0 + a)
= f (z0 + a).
i i
i i
i i
z0 + a + b
r
z0 + a
r
r
z0 + b
r
z0
A fundamental parallelogram
Proof:
1. As in the proof of Lemma 8.1.3(3), for the given z ∈ C, we get z = ra+
sb for some r, s ∈ R. Next, we choose m, n ∈ Z such that 0 ≤ r+m < 1
and 0 ≤ s + n < 1 so that
i i
i i
i i
Case 2: w0 0.
Then, we assume that ζ0 ∂Da,b (z0 ), and hence, we can
4
partition Da,b (z0 ) as ∪ Rj , where
j=1
w0 + b R3
R4
a + b R2
w0 + a
b R1
z0 + a + b S2 w
Da,b(0)
Da,b(z0) p 0
+ b− S4
z0 + b R3 a S3 a
ζ0
R4 ζ0
R2 z0 + a w0 0
R1 +p
z 0
z0
i i
i i
i i
Definition 8.2.3 Let f be an elliptic function. Then, the number of poles (in-
cluding multiplicities) of f in a fundamental parallelogram of f is called the
order of the elliptic function f .
i i
i i
i i
where
γ1 (t) = z0 + ta, t ∈ [0, 1]
γ2 (t) = z0 + a + tb, t ∈ [0, 1]
γ3 (t) = z0 + (1 − t)a + b, t ∈ [0, 1]
γ4 (t) = z0 + (1 − t)b, t ∈ [0, 1]
z0 + a + b
γ3
γ2
z0 + b
z0 + a
γ4
γ1
z0
i i
i i
i i
Therefore, using a, b ∈ Pf ,
1
f (z) dz = f (z0 + ta) adt
γ1 0
0
= f (z0 + (1 − s)a) a(−ds)
1
1
= − f (z0 + (1 − s)a) a(−ds)
0
= − f (z) dz.
γ3
Similarly, we can prove that f (z) dz = − f (z) dz, and hence, f (z) dz = 0.
γ2 γ4 γ
Thus, we get the sum of the residues of f at poles in Da,b (z0 ) = 0.
f
Proof: Using Theorem 8.1.7 and Lemma 8.1.9, we get that is also an
f
elliptic function. Therefore, as in the proof of the previous lemma, we can
f (z)
prove that dz = 0. However, using argument principle, we get
∂Da,b (z0 ) f (z)
f (z)
1
i2π dz = number of zeroes of f (including multiplicities) en-
∂Da,b (z0 ) f (z)
closed by ∂Da,b (z0 )− number of poles of f (including multiplicities) enclosed
by ∂Da,b (z0 ). This completes the proof of the lemma.
LEMMA 8.2.6 If {z1 , z2 , . . . , zn } and {w1 , w2 , . . . , wn } are the zeroes and poles
of an elliptic function f including multiplicities in a fundamental parallelo-
gram Da,b of f , respectively, then z1 + z2 + · · · + zn − w1 − w2 − · · · −
wn ∈ Pf .
i i
i i
i i
φ
As φ is analytic, by applying Cauchy’s theorem and Cauchy’s integral
formula, we get
⎛ ⎞
n
1 zf (z) ⎜ 1 z 1 z ⎟
dz = ⎝ dz − dz⎠
i2π f (z) i2π (z − zj ) i2π (z − wj )
∂Da,b j=1 ∂Da,b ∂Da,b
1 φ (z)
+ dz
i2π φ(z)
∂Da,b
= z1 + z2 + · · · + zn − w1 − w2 − · · · − wn .
Next, using the parametric equation of ∂Da,b (from Lemma 8.2.4), we get
4
1 zf (z) 1 zf (z)
dz = dz,
i2π f (z) i2π f (z)
∂Da,b j=1 γj
i i
i i
i i
1
zf (z) (ζ + ta)f (ζ + ta)
dz = adt
f (z) f (ζ + ta)
γ1 0
1
(ζ + b + ta)f (ζ + b + ta)
= adt
f (ζ + b + ta)
0
1
bf (ζ + b + ta)
− adt
f (ζ + b + ta)
0
0
(ζ + b + (1 − s)a)f (ζ + b + (1 − s)a)
= a(−ds)
f (ζ + b + (1 − s)a)
1
1
bf (ζ + ta)
− adt
f (ζ + ta)
0
1
(ζ + b + (1 − s)a)f (ζ + b + (1 − s)a)
= − a(−ds)
f (ζ + b + (1 − s)a)
0
f (z)
−b dz
f (z)
γ1
zf (z) f (z)
= − dz − b dz.
f (z) f (z)
γ3 γ1
Now, using a is a period of f and f has no zero or pole on ∂Da,b , we get that
f (γ1 ) is a closed curve not passing through zero. Then, by using the change
of variable w = f (z), we get
f (z) dw
dz = dw = WN( f (γ1 ), 0) ∈ Z.
f (z) w−0
γ1 f (γ1 )
i i
i i
i i
zf (z)
Hence, dz = nb for some n ∈ Z. Similarly, we can prove that
γ1 +γ3 f (z)
zf (z)
dz = ma, for some m ∈ Z. Thus,
γ2 +γ4 f (z)
1 zf (z)
z1 + z2 + · · · + zn − w1 − w2 − · · · − wn = dz = ma + nb
i2π f (z)
∂Da,b
which belongs to Pf .
i i
i i
i i
∀z ∈ Z \ {ma + nb : m, n ∈ Z}.
The above infinite series means lim SJ , whenever the limit exists,
J →∞
J
where SJ = 1
(z−(ma+nb))2
− 1
(ma+nb)2
, ∀J ∈ N. For notational
N=1 m,n∈Z
|m|+|n|=N
convenience, we write ℘a,b as follows.
1 1 1
℘a,b (z) = 2 + − 2 , ∀z ∈ Z \ Qa,b ,
z (z − p)2 p
p∈Qa,b \{0}
Proof: If φ : R × R → R is defined by
i i
i i
i i
Case 1: r0 0.
−s0
Then, |r0 a + s0 b| = 0 ⇒ r0 a + s0 b = 0 ⇒ a
b = r0 ∈ R, which is
a contradiction.
Case 2: s0 0.
−r0
Here, |r0 a + s0 b| = 0 ⇒ r0 a + s0 b = 0 ⇒ a = s0
b
∈ R, which is a
contradiction to b
a R. (Note that b
a R ⇔ ab R.)
LEMMA 8.3.3 If a and b are independent complex numbers, then the series
1 1
− 2
(z − p)2 p
p∈Qa,b \{0}
i i
i i
i i
For a given z ∈ K and p as in equation (8.1), we have |p| > 2C ≥ 2|z|, and
|p| |p|
hence, |z − p| ≥ |p| − |z| > |p| − = . Therefore,
2 2
1 1 2pz − z2
− =
(z − p)2 p2 p2 (z − p)2
pz z(p − z)
≤ 2 +
p (z − p)2 p2 (z − p)2
|z| |z|
< +
|p| 2 |p|
|p| |p|2
4 2
6C
≤ . (8.2)
|p|3
Next, we note that for every N ∈ N, there exist exactly 4N ordered pairs
(m, n) such that |m| + |n| = N . Indeed, the required ordered pairs are given
as follows.
If N is even, then
i i
i i
i i
i i
i i
i i
we get
∞
1 1 1
℘a,b (−z) = + −
(−z)2 (−z − (ma + nb))2 (ma + nb)2
N=1 m,n∈Z
|m|+|n|=N
∞
1 1 1
= + −
z2 (z + (ma + nb))2 (ma + nb)2
N=1 m,n∈Z
|m|+|n|=N
∞
1 1 1
= + −
z2 (z − (ua + vb))2 (ua + vb)2
N=1 u,v∈Z
|u|+|v|=N
= ℘a,b (z).
2 2 −2 −2
℘a,b (z) = − − = = .
z3 (z − p)3 (z − p)3 (z − (ma + nb))3
p∈Qa,b \{0} p∈Qa,b m,n∈Z
−2
℘a,b (z + ja + kb) =
(z + ja + kb − (ma + nb))3
m,n∈Z
−2
=
(z + (j − m)a + (k − n)b)3
m,n∈Z
−2
=
(z + ra + sb)3
r,s∈Z
= ℘a,b (z).
is a doubly periodic with P = Q . If
Therefore, ℘a,b ℘ a,b a,b
then
F (z) = ℘a,b
(z + a) − ℘a,b (z) = 0, ∀z ∈ C \ Qa,b .
i i
i i
i i
LEMMA 8.3.5 If f and g are elliptic functions such that both have same ze-
roes and poles of the same orders, then prove that f = cg for some constant
c ∈ C.
Proof: As f and g are elliptic functions with same zeroes and poles, it is ob-
vious that they have same fundamental periods. Then, gf is an entire function
after removing the singularities at the zeroes and poles of f . Thus, by Lemma
8.2.7, gf is a constant, and hence, the lemma follows.
THEOREM 8.3.6 Every even elliptic function with fundamental periods a and
b is a rational function of ℘a,b .
i i
i i
i i
i i
i i
i i
for some c ∈ C.
Case 2: f can have a zero or a pole at 0.
If f has a zero of order n at 0, then n ∈ 2N, as f is even. We note that
n
℘a,b has a double pole at 0, and hence, ℘a,b
2
has a pole at 0 of order
n
n. Therefore, f ℘a,b
2
has neither a pole nor a zero at 0 after removing
the singularity at 0.
If f has a pole of order m, then m ∈ 2N, as f is even. Therefore,
−m
f ℘a,b2 has neither a pole or a zero at 0 after removing the singularity
at 0.
Hence, the theorem follows.
),
THEOREM 8.3.7 Every elliptic function can be written as R(℘a,b , ℘a,b
where R is a rational function in two variables.
RESULT 8.3.8 The Weierstrass ℘a,b function satisfies the following differen-
tial equation.
(℘a,b (z))2 = 4(℘a,b (z))3 − 60G4 ℘a,b (z) − 140G6 .
i i
i i
i i
Proof: We first find the Laurent series expansion of ℘a,b in B(0, r) \ {0}, for
some r > 0.
1 1 1
℘a,b (z) = + −
z2 (z − p)2 p2
p∈Qa,b \{0}
1 d 1 1
= + −
z2 dz (p − z) p2
p∈Qa,b \{0}
!∞ " #
1 1 d z k 1
= + − 2
z2 p dz p p
p∈Qa,b \{0} k=0
∞ k−1 #
1 1 z 1
= + k − 2
z2 p2 p p
p∈Qa,b \{0} k=1
! ∞ k−1 " #
1 1 z 1
= + 1+ k − 2
z2 p2 p p
p∈Qa,b \{0} k=2
∞ k−1
1 1 z
= + k
z2 p2 p
p∈Qa,b \{0} k=2
∞
1 1 k
= + (k + 1) z
z2 pk+2
k=1 p∈Qa,b \{0}
∞
1
= + (k + 1)Gk+2 zk
z2
k=1
∞
1
= + (2n + 1)G2n+2 z2n
z2
n=1
∞
−2
℘a,b (z) = + 2n(2n + 1)G2n+2 z2n−1 .
z3
n=1
i i
i i
i i
∞
(z))2 , then we first
If An zn is the Laurent series expansion of (℘a,b
n=−∞
0
find An zn (the singular part along with A0 ) as
n=−∞
2
2
(℘a,b (z))2 = − 3 + 6G4 z + 20G6 z + · · ·
3
z
4 1
= − 24G4 2 − 80G6 + A2 z2 + A3 z3 + · · ·
z6 z
(In the first line of the above expression, we have written the series expansion
(z) up to the term corresponding to z3 explicitly. The reason is that
of ℘a,b
only these three terms which contribute to get the terms corresponding to zn,
n ≤ 0 in the square of this expansion.
Similarly, we write only the terms corresponding to zn , n ≤ 0, in the
expansion of ℘a,b (z))3 as follows.
3
1
(℘a,b (z))3 = + 3G 4 z 2
+ 5G 6 z 6
+ · · ·
z2
1 9G4
= + + 15G6 + · · · .
z6 z2
4 24G4 1 9G4
− 2 − 80G6 + · · · − 4 6 + 2 + 15G6 + · · ·
z6 z z z
1
+60G4 2 + 140G6 + B1 z + B2 z2 + B3 z3 + · · · ,
z
i i
i i
i i
= 6℘
Exercise 8.3.9 Prove that ℘a,b a,b − 30G4 .
RESULT 8.3.10 The Weierstrass function ℘a,b satisfies the following differ-
ential equation
2
℘a,b = 4 ℘a,b − e1 ℘a,b − e2 ℘a,b − e3 ,
a
where e1 = ℘a,b 2 , e2 = ℘a,b b
2 and e3 = ℘a,b a+b
2 .
Proof: From Case 1 of proof Theorem 8.3.6, we have the following observa-
tions:
in the fundamental parallelogram D (0) are a ,
1. The only zeroes of ℘a,b a,b 2
b a+b 2
2 and 2 , and each is a simple zero. Thus, the only zeroes of (℘a,b )
are double zeroes at a2 , b2 and a+b
2 .
2. The only zeroes of ℘a,b − e1 ℘a,b − e2 ℘a,b − e3 in Da,b (0) are
a b a+b
2 , 2 and 2 , and each zero has multiplicity 2.
has a pole at 0 in Da,b (0) of order 6. Similarly, ℘a,b has the only pole at
2
0 in Da,b (0) of order 3, and hence (℘a,b ) has a pole at 0 in Da,b (0) which
)2 and ℘
is of order 6. Therefore, (℘a,b a,b − e1 ℘a,b − e2 ℘a,b − e3
have same number of zeroes and poles of same orders in C. Therefore,
)2
(℘a,b
2
is a constant, say c . Thus, we have ℘ =
(℘a,b −e1 )(℘a,b−e
2 )(℘a,b −e3) a,b
c ℘a,b − e1 ℘a,b − e2 ℘a,b − e3 . Next, we find the value of c as follows.
As
⎛ ⎞2
2
lim z6 (℘a,b
(z))2 = lim z6 ⎝ ⎠ = 22 = 4
z→0 z→0 (z − (ma + nb))3
m,n∈Z
and
⎛ ⎞
1 1 1
lim z 2
℘a,b (z) − e1 = lim z2 ⎝ + − 2 ⎠ = 1,
z→0 z→0 z2 (z − p)2 p
p∈Qa,b \{0}
i i
i i
i i
we get
)2
(℘a,b
c =
℘a,b − e1 ℘a,b − e2 ℘a,b − e3
(z))2
(℘a,b
= lim
z→0 ℘a,b (z) − e1 ℘a,b (z) − e2 ℘a,b (z) − e3
(z))2
lim z6 (℘a,b
z→0
=
lim z6 ℘a,b (z) − e1 ℘a,b (z) − e2 ℘a,b (z) − e3
z→0
= 4.
Proof: From the definition of ℘a,b , clearly we have that ℘a,b is an elliptic
function of order 2 (as the only pole of ℘a,b in Da,b (0) is at 0 and it is a
double pole) and
1 1 1
lim ℘a,b (z) − 2 = lim − = 0.
z→0 z z→0 (z − p)2 p2
p∈Qa,b \{0}
i i
i i
i i
Proof: Since ℘a,b is doubly periodic and even if u ± v ∈ Qa,b , then ℘a,b (u) =
(u) = ℘ (v), and hence, the determinant is 0. Therefore, we
℘a,b (v) and ℘a,b a,b
assume that u ± v Qa,b . Now, we consider the following system of linear
equations in two variables α and β .
℘a,b (u) = α℘a,b (u) + β and ℘a,b (v) = α℘a,b (v) + β.
the above system of linear equations has a unique solution, say (α0 , β0 ). If
f (z) = ℘a,b (z) − α0 ℘a,b (z) − β0
i i
i i
i i
is an odd
Here, we have used the fact that ℘a,b is an even function and ℘a,b
function.
(u) − ℘ (v)
#2
1 ℘a,b a,b
℘a,b (u + v) = − ℘a,b (u) − ℘a,b (v).
4 ℘a,b (u) − ℘a,b (v)
(u)
#2
1 ℘a,b
℘a,b (2u) = (u) − 2℘a,b (u).
4 ℘a,b
4(℘a,b (z))3 − 60G4 ℘a,b (z) − 140G6 − (α02 (℘a,b (z))2 + 2α0 β0 ℘a,b (z) + β02 ) = 0
i i
i i
i i
and hence,
4(℘a,b (z))3 −α02 (℘a,b (z))2 −(2α0 β0 +60G4 )℘a,b (z)−(140G6 +β02 ) = 0, ∀z ∈ S.
By the choice of u and v, the three complex numbers ℘a,b (u), ℘a,b (v), and
℘a,b (u + v) are pairwise distinct, and hence, they are the only solutions of the
cubic polynomial equation
4X 3 − α02 X 2 − (2α0 β0 + 60G4 )X − (140G6 + β02 ) = 0.
Therefore, the sum of the roots of the equation ℘a,b (u) + ℘a,b (v) + ℘a,b (u + v)
(u)−℘ (v)
℘a,b
equals 14 α02 . Now, we find α0 = a,b
℘a,b (u)−℘a,b (v) by solving from the system of
linear equations
℘a,b (u) = α0 ℘a,b (u) + β0 and ℘a,b (v) = α0 ℘a,b (v) + β0 .
Thus, we get
(u) − ℘ (v)
#2
1 ℘a,b a,b
℘a,b (u + v) = − ℘a,b (u) − ℘a,b (v). (8.6)
4 ℘a,b (u) − ℘a,b (v)
i i
i i
i i
Proof: Applying Result 8.3.8 in equation (8.6) and using the oddity of ℘a,b
and the evenness of ℘a,b , we get
℘a,b (u + v) + ℘a,b (u − v)
#2
1 ℘a,b (u) − ℘a,b (v)
= − ℘a,b (u) − ℘a,b (v)
4 ℘a,b (u) − ℘a,b (v)
#2
1 ℘a,b (u) + ℘a,b (v)
+ − ℘a,b (u) − ℘a,b (v)
4 ℘a,b (u) − ℘a,b (v)
1 (℘a,b (u)) + (℘a,b (v))
2 2
= − 2(℘a,b (u) + ℘a,b (v))
2 ℘a,b (u) − ℘a,b (v) 2
(u))2 + (℘ (v))2 − 4(℘ (u) + ℘ (v)) ℘ (u) − ℘ (v) 2
(℘a,b a,b a,b a,b a,b a,b
= 2
2 ℘a,b (u) − ℘a,b (v)
1 $
= 2 4(℘a,b (u)) − 60G4 ℘a,b (u) − 140G6
3
2 ℘a,b (u) − ℘a,b (v)
%
+ 4(℘a,b (v))3 − 60G4 ℘a,b (v) − 140G6
$ %
− 4 (℘a,b (u))3 + (℘a,b (v))3 − ℘a,b (u)(℘a,b (v))2 − (℘a,b (u))2 ℘a,b (v)
1
= 2 −60G4 (℘a,b (u) + ℘a,b (v)) − 280G6
2 ℘a,b (u) − ℘a,b (v)
+ 4℘a,b (u)℘a,b (v)(℘a,b (u) + ℘a,b (v))
(2℘a,b (u)℘a,b (v) − 30G4 )(℘a,b (u) + ℘a,b (v)) − 140G6
= 2 .
℘a,b (u) − ℘a,b (v)
Hence, the corollary follows.
i i
i i
i i
i i
i i
i i
Proof: We shall show that the series converges uniformly on Cl B(0, r) for
every r > 0. Using Lemma 8.3.2, first we find μ > 0 such that |ma + nb| ≥
2r
μ(|m| + |n|), ∀(m, n) ∈ Z. Now, choose N0 ∈ N such that N0 > so that
μ
|ma + nb| ≥ μ(|m| + |n|) ≥ μN0 > 2r, ∀(m, n) ∈ Z2 with |m| + |n| ≥ N0 .
a
(Note that as R, we have a 0 b.)
b
z
Now, for |z| ≤ r and for (m, n) ∈ Z2 with |m| + |n| ≥ N0 , we have ma+nb <
1
2 , and hence,
1 1 z
z − (ma + nb) + ma + nb + (ma + nb)2
! "
−1
1 1 z
= + +
ma + nb 1 − ma+nb
z
ma + nb (ma + nb)2
! "
−1 ∞
zk
z 1 z
= 1+ + + +
ma + nb ma + nb |ma + nb|k ma + nb (ma + nb)2
k=2
−1 ∞
zk
=
ma + nb (ma + nb)
k
k=2
∞
|z|2 |z|k
≤
|ma + nb|3 |ma + nb|k
k=0
|z|2 1
≤
|ma + nb| 1 − |z|
3
|ma+nb|
|z|2 1
≤
|ma + nb|3 1 − 1
2
2|z|2 1
≤ ≤ 4r2 ,
|ma + nb| 3 |ma + nb|3
i i
i i
i i
RESULT 8.4.3 There exist constants α, β ∈ C such that ζa,b (z+a)−ζa,b (z) =
α and ζa,b (z + b) − ζa,b (z) = β , ∀z ∈ C \ Qa,b . Furthermore, they satisfy
αb − βa = 2πi.
= −℘ . As a is a period of
Proof: From the previous lemma, we have ζa,b a,b
℘a,b , we get
ζa,b (z + a) − ζa,b (z) = 0 ⇒ ℘a,b (z + a) − ℘a,b (z) = 0, ∀z ∈ C \ Qa,b .
Thus, there exists a constant α ∈ C such that
ζa,b (z + a) − ζa,b (z) = α, ∀z ∈ C \ Qa,b .
Similarly, we can prove that there exists β ∈ C such that
ζa,b (z + b) − ζa,b (z) = β, ∀z ∈ C \ Qa,b
Let Da,b (z0 ) be a fundamental parallelogram such that ∂Da,b (z0 ) does not have
any pole of ζa,b . Since the set of all poles of ζa,b is same as Qa,b , Da,b (z0 )
should contain exactly one pole of ζa,b . We also know that Res (ζa,b , p) = 1,
∀p ∈ Qa,b , as the coefficient of z−p
1
in the Laurent series expansion of ζa,b in
B(a, r) \ {a} is 1, for some r > 0, ∀p ∈ Qa,b . Note that a parametric equation
of ∂Da,b (z0 ) is given as follows:
z0 + a + b
i i
i i
i i
1 1
= ζa,b (z0 + ta) adt + ζa,b (z0 + a + tb) bdt
0 0
0 0
− ζa,b (z0 − ta + b) adt − ζa,b (z0 − tb) bdt
−1 −1
1 1
& '
= a ζa,b (z0 + ta) dt + b ζa,b (z0 + tb) + α dt
0 0
1 1
& '
−a ζa,b (z0 + sa) + β ds − ζa,b (z0 + sb) ds
0 0
1 1
= −a β dt + b α dt
0 0
= αb − βa.
Remark 8.4.4: From the previous proposition, we note that ζa,b is not a
periodic function.
i i
i i
i i
i i
i i
i i
z ( f2 (z) − ρa,b (z)) has a removable singularity at 0. Hence, after removing the
2
z
+ 12 z2
LEMMA 8.4.7 The infinite product 1− z
p ep p2 converges uni-
p∈Qa,b \{0}
formly on every compact subset of C.
i i
i i
i i
(tw)2
ϕ(t) = (1 − tw)etw+ 2 − 1, t ∈ [0, 1]
(z)
σa,b
RESULT 8.4.9 The function σa,b is an odd function and it satisfies σa,b (z) =
ζa,b (z), ∀z ∈ C \ Qa,b .
i i
i i
i i
Proof: For z ∈ C, using the fact that Qa,b \ {0} = −(Qa,b \ {0}), we get
(−z) 1 (−z)2
(−z) +
σa,b (−z) = (−z) 1−
e p 2 p2
p
p∈Qa,b \{0}
z
z 1 z2
+
= −z 1− e (−p) 2 p2
(−p)
p∈Qa,b \{0}
z
z 1 z2
+
= −z 1− e p 2 p2
p
p∈Qa,b \{0}
= −σa,b (z).
1. Pf = Qa,b ,
n
σa,b (z − zj )
f (z) = c , ∀z ∈ C \ Sf ,
σa,b (z − wj )
j=1
i i
i i
i i
σa,b (z + a)σa,b (z) − σa,b (z)σa,b (z + a) − ασa,b (z + a)σa,b (z) = 0. (8.7)
σa,b (z + b)σa,b (z) − σa,b (z)σa,b (z + b) − βσa,b (z + b)σa,b (z) = 0. (8.8)
For$ z ∈ C \ Sf , consider
d σa,b (z+a)
%
dz σa,b (z) exp −α z + a
2
! "
a σa,b (z + a) σa,b (z + a)σa,b (z)
= exp −α z + −
2 σa,b (z) σa,b
2 (z)
a σa,b (z + a)
−α exp −α z +
2 σa,b (z)
$
exp −α z + 2 a
= σa,b (z + a)σa,b (z) − σa,b (z + a)σa,b (z)
σa,b
2 (z)
'
−ασa,b (z + a)σa,b (z)
= 0(by applying equation (8.7)).
σ (z+a)
σa,b (z) exp −α z + 2 is a constant function on C \ Qa,b , and
a
Therefore, a,b
hence, it is a constant function (say C ) on C. By substituting z = − a2 , we get
σ (a)
C = σ a,b −2a = −1 because σa,b is an odd function and σa,b a2 0. Thus,
a,b ( 2 )
we have a
σa,b (z + a) = −σa,b (z) exp α z + (8.9)
2
i i
i i
i i
σa,b (z − zj + a) σa,b (z − zj )
=
σa,b (z − wj + a) σa,b (z − wj )
We start this section with the motivation for introducing elliptic integrals
from the problem of simple pendulum.
A simple pendulum consists of a weightless thread of length l, whose one
end is attached to a sphere of mass m and the other end is attached to a fixed
point on the roof. By drawing the sphere aside from the equilibrium position,
we assume that the sphere oscillates in a same vertical plane. The problem
is to determine the angle θ between the current position of thread with its
equilibrium position, as a function of time t.
l cos (q )
q
l
l (1−cos (q ))
)
(q
sin
mg
1 This statement can be obtained from the fact that a and b are fundamental periods of ℘
a,b
and by the relations among ℘a,b , ζa,b and σa,b .
i i
i i
i i
Applying the Newton’s second law, the force acting upon the sphere due
to this movement is same as the tangential component of force acting upon
the sphere due to gravity. Thus, we get
d2 d2θ g
m 2
(lθ) = −mg sin(θ), ⇒ 2
+ sin(θ ) = 0,
dt dt l
where lθ is the arc length of the path of the sphere from equilibrium position
to the present time. Multiplying on both sides of the above equation by ml2 dθ
dt ,
we get
! "
2 dθ d θ
2 dθ d 1 dθ 2
ml +mgl sin θ =0⇒ m l + mgl(1 − cos(θ )) = 0.
dt dt2 dt dt 2 dt
Thus, we get the total energy (kinetic energy + potential energy) acting upon
the sphere at time t, which is constant. As the energy required to move the
sphere from the equilibrium position (corresponding to θ = 0) to the highest
position (corresponding to θ = π ) is k 2 (2mgl) for some 0 < k < 1 (because
of the oscillatory motion), we have
1 dθ 2
l + gl(1 − cos(θ )) = k 2 (2gl). (8.10)
2 dt
If the initial velocity of the
sphere is v0 , then by the law of conservation of
2
energy, then we have 12 m v20 − l dθdt = mgl(1 − cos(θ )) and so we have
2
dθ
l = v20 − 2gl(1 − cos(θ )). (8.11)
dt
v20
Comparing equations (8.10) and (8.11), we find that k 2 = 4gl . Thus, using
the change of variable x = gl t in equation (8.11), we get
2
dθ θ
= 4 k 2 − sin2 . (8.12)
dx 2
To obtain the Euler’s normal form and Jacobi’s normal form of
equation
(8.12), we consider the change of variable k sin (φ) = sin θ2 , which is
θ dθ
introduced by Euler and Jacobi. As k cos(φ) dφ
dx = 2 cos 2 dx , we have
1
2
4k 2 cos2 (φ) dφ 2 dθ 2 θ
θ = = 4 k 2
− sin = 4k 2 (1 − sin2 (φ))
2
cos 2 dx dx 2
i i
i i
i i
The integral equation (8.17) is called the elliptic integral of first kind (See
first section of this chapter). The constant k is called the modulus, and x is
called the argument. If Fk is the function denoting φ → x, then Fk is an odd
and increasing function. Indeed,
i i
i i
i i
Definition 8.5.1 The basic Jacobian elliptic functions are defined as follows:
1. snk (x) = sin(amk (x)) = sin(φ),
φ
For the case k = 0, as x = F0 (φ) = du = φ , we have, by definition,
0
π
sn0 (x) = sin(x), cn0 (x) = cos(x), dn0 (x) = 1, and K0 = .
2
When k = 1, as
φ
x = F1 (φ) = sec(u)du = ln(sec(φ) + tan(φ))
0
we have
1 + sin(φ)
exp(x) = sec(φ) + tan(φ) =
cos(φ)
i i
i i
i i
and hence,
(1 + sin(φ))2 (1 + sin(φ))2 1 + sin(φ)
exp(2x) = = = .
2
cos (φ) (1 − sin (φ))
2 1 − sin(φ)
Solving for sin(φ) from the above two equations, we get
exp(2x) − 1
(1 − sin(φ)) exp(2x) = 1 + sin(φ) ⇒ sin(φ) = = tanh(x).
exp(2x) + 1
Hence,
sn1 (x) = sin(φ) = tanh(x),
cn1 (x) = cos(φ) = 1 − sin2 (φ) = 1 − tanh2 (x) = sech(x),
dn1 (x) = 1 − sin2 (φ) = 1 − tanh2 (x) = sech(x),
π π
2 2
du
K1 = = sec2 (u) du = ∞.
0 1 − sin (u)
2
0
Proof: Directly from the definition of snk , cnk , and dnk , we get the follow-
ing:
1. From the definition of Fk , we have Fk (0) = 0, and hence, amk (0) =
0. Therefore, snk (0) = sin(amk (0)) = sin(0) = 0, cnk (0) =
cos(amk (0)) = cos(0) = 1, and dnk (0) = 1 − k 2 sn2k (0) = 1.
2. We note that Fk π2 = Kk , and hence, amk (Kk ) = π2 . Therefore,
π
snk (Kk ) = sin(amk (Kk )) = sin = 1,
2
π
cnk (Kk ) = cos(amk (Kk )) = cos = 0,
2
dnk (Kk ) = 1 − k 2 sn2k (Kk ) = 1 − k 2 .
i i
i i
i i
3. We know that amk and sin are odd functions, and hence,
snk (−x) = sin(amk (−x)) = sin(−amk (x)) = − sin(amk (x)) = −snk (x),
sn2k (x) + cn2k (x) = sin2 (amk (x)) + cos2 (amk (x)) = 1
and
dn2k (x) = 1 − k 2 sn2k (x) ⇒ dn2k (x) + k 2 sn2k (x) = 1.
snk (x + 2Kk ) = −snk (x), cnk (x + 2Kk ) = −cnk (x), dnk (x + 2Kk ) = dnk (x).
1
Proof: First, we note that is a periodic function of period π
1 − k 2 sin2 (u)
(as sin2 is a periodic function of period π ). Let φ ∈ R be arbitrary. Applying
Lemma 2.4.23, twice (for a = − π2 and for a = φ ), we get
π
2 φ+π
du du
= . (8.19)
− π2 1 − k 2 sin (u)
2
φ 1 − k 2 sin2 (u)
i i
i i
i i
φ φ+π
du du
= +
0 1 − k 2 sin (u)
2
φ 1 − k 2 sin2 (u)
(Using equation (8.19))
φ+π
du
= .
0 1 − k 2 sin2 (u)
Fk (φ + π ) = x + 2Kk . (8.20)
Similarly, we also get cnk (x + 4Kk ) = cnk (x), and hence, the the following
corollary holds.
COROLLARY 8.5.4 snk and cnk are periodic functions of period 4Kk , and dnk
is a periodic function of period 2Kk .
RESULT 8.5.5 snk = cnk dnk , cnk = −snk dnk , and dnk = −k 2 snk cnk .
snk (x) = d
$dx snk (x) % = d
dx (sin(φ))
dφ
= d
dφ sin(φ) = cos(φ) dam k
dx dx
i i
i i
i i
1 1 1
nsk = , nck = , ndk = ,
snk cnk dnk
snk snk cnk
sck = , sdk = , csk = ,
cnk dnk snk
cnk dnk dnk
cdk = , dsk = , dck = .
dnk snk cnk
So far the Jacobi’s elliptic functions are defined from R into R. To extend
them as functions from C into C, we use the Jacobi’s imaginary transforma-
tion which is given as follows:
Proof: Using the change of variable sin(v) = i tan(u), we get the following
identities;
1. cos(v) = 1 − sin2 (v) = 1 + tan2 (u) = sec(u), and hence,
cos(u) = sec(v).
2. sin(u) = 1 − cos2 (u) = 1 − sec2 (v) = i tan(v).
i i
i i
i i
i i
i i
i i
snk (ix + i4Kj ) = snk (ix); cnk (ix + i4Kj ) = cnk (ix); dnk (ix + i4Kj ) = dnk (ix),
π
2 du √
where Kj = and j = 1 − k2.
0 1 − j2 sin2 (u)
THEOREM 8.5.9
cnk (x)cnk (y) − snk (x)snk (y)dnk (x)dnk (y)
cnk (x + y) =
1 − k 2 sn2k (x)sn2k (y)
snk (x)cnk (y)dnk (y) + snk (y)cnk (x)dnk (x)
snk (x + y) =
1 − k 2 sn2k (x)sn2k (y)
i i
i i
i i
Proof: Let Fk (φ) = x, Fk (ψ) = y, and Fk (μ) = x + y. Then, using the fact
that Fk is an odd function, we have
Dividing on both sides of equation (8.22) by sin(φ) sin(ψ) and then differen-
tiating, we get
cos(φ) cos(ψ) cos(μ)
d =d .
sin(φ) sin(ψ) sin(φ) sin(ψ)
As
cos(φ) cos(ψ) 1
d = [sin(φ) sin(ψ)(− sin(φ)dφ cos(ψ)
sin(φ) sin(ψ) sin (φ) sin2 (ψ)
2
i i
i i
i i
we have
dφ dψ
2
[cos(ψ) − cos(μ) cos(φ)] +
sin (φ) sin(ψ) sin(φ) sin2 (ψ)
[cos(φ) − cos(μ) cos(ψ)] = 0
which implies that
cos(ψ) − cos(μ) cos(φ) cos(φ) − cos(μ) cos(ψ)
dφ + dψ = 0.
sin(φ) sin(ψ)
(8.24)
The expression in equation (8.22) can be rewritten as
(cos(φ) cos(ψ) − cos(μ))2 = sin2 (φ) sin2 (ψ)(1 − k 2 sin2 (μ)) (8.25)
⇒ cos2 (φ) cos2 (ψ) + cos2 (μ) − 2 cos(φ) cos(ψ) cos(μ)
= (1 − cos2 (φ))(1 − cos2 (ψ)) − k 2 sin2 (φ) sin2 (ψ) sin2 (μ)
= 1 − cos2 (φ) − cos2 (ψ) + cos2 (φ) cos2 (ψ)
− k 2 sin2 (φ) sin2 (ψ) sin2 (μ)
⇒ cos (φ) + cos2 (ψ) + cos2 (μ) − 2 cos(φ) cos(ψ) cos(μ)
2
i i
i i
i i
⇒ cos2 (φ) cos2 (ψ) + ζ 2 − 2 cos(φ) cos(ψ)ζ = (1 − k 2 ) sin2 (φ) sin2 (ψ)
+k 2 sin2 (φ) sin2 (ψ)ζ 2
⇒ (1 − k 2 sin2 (φ) sin2 (ψ))ζ 2 − 2 cos(φ) cos(ψ)ζ + cos2 (φ) cos2 (ψ)
+(k 2 − 1) sin2 (φ) sin2 (ψ) = 0.
Solving for ζ from √ the above quadratic equation, we get ζ =
2 cos(φ) cos(ψ) ± D
, where
2(1 − k 2 sin2 (φ) sin2 (ψ))
D = 4 cos2 (φ) cos2 (ψ) − 4(1 − k 2 sin2 (φ) sin2 (ψ))[cos2 (φ) cos2 (ψ)
+(k 2 − 1) sin2 (φ) sin2 (ψ)]
= 4[(1 − k 2 ) sin2 (φ) sin2 (ψ) + k 2 sin2 (φ) sin2 (ψ) cos2 (φ) cos2 (ψ)
+k 2 (k 2 − 1) sin4 (φ) sin4 (ψ)]
= 4[sin2 (φ) sin2 (ψ) − k 2 sin2 (φ) sin2 (ψ)
+k 2 sin2 (φ) sin2 (ψ)(1 − sin2 (φ) − sin2 (ψ) + sin2 (φ) sin2 (ψ))
+k 4 sin4 (φ) sin4 (ψ) − k 2 sin4 (φ) sin4 (ψ)]
= 4[sin2 (φ) sin2 (ψ) − k 2 sin2 (φ) sin4 (ψ) − k 2 sin4 (φ) sin2 (ψ)
+k 4 sin4 (φ) sin4 (ψ)]
= 4 sin2 (φ) sin2 (ψ)(1 − k 2 sin2 (ψ) − k 2 sin2 (φ) + k 4 sin2 (φ) sin2 (ψ))
= 4 sin2 (φ) sin2 (ψ)(1 − k 2 sin2 (φ))(1 − k 2 sin2 (ψ)).
Therefore,
cos(φ + ψ) = ζ
cos(φ) cos(ψ) ± sin(φ) sin(ψ) (1 − k 2 sin2 (ψ))
(1 − k 2 sin2 (φ))
=
1 − k 2 sin2 (φ) sin2 (ψ)
cnk (x)cnk (y) ± snk (x)snk (y)dnk (x)dnk (y)
⇒ cnk (x + y) = .
1 − k 2 sn2k (x)sn2k (y)
By observing
cn0 (x + y) = cos(x + y) = cos(x) cos(y) − sin(x) sin(y)
= cn0 (x)cn0 (y) − sn0 (x)sn0 (y)
we get
cnk (x)cnk (y) − snk (x)snk (y)dnk (x)dnk (y)
cnk (x + y) = .
1 − k 2 sn2k (x)sn2k (y)
i i
i i
i i
Using the identity sn2k + cn2k = 1 in the following straightforward but lengthy
manipulation, we get
sn2k (x + y)
= 1 − cn2k (x + y)2
! "2
cnk (x)cnk (y) − snk (x)snk (y)dnk (x)dnk (y)
= 1−
1 − k 2 sn2k (x)sn2k (y)
[1 − k snk (x)snk (y)]2 − [cnk (x)cnk (y) − snk (x)snk (y)dnk (x)dnk (y)]2
2 2 2
=
& Dr'2
( where Dr = 1 − k 2 sn2k (x)sn2k (y) )
= Dr 1
[1 + k 4 sn4k (x)sn4k (y) − 2k 2 sn2k (x)sn2k (y) − cn2k (x)cn2k (y)
−sn2k (x)sn2k (y)dn2k (x)dn2k (y) + 2snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)]
= Dr 1
[1 + k 4 sn4k (x)sn4k (y) − 2k 2 sn2k (x)sn2k (y)
−(1 − sn2k (x) − sn2k (y) + sn2k (x)sn2k (y))
−sn2k (x)sn2k (y)(1 − k 2 sn2k (x) − k 2 sn2k (y) + k 4 sn2k (x)sn2k (y))
+2snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)]
= Dr 1
[sn2k (x) + sn2k (y) + k 2 sn2k (x)sn4k (y) + k 2 sn4k (x)sn2k (y)
−2sn2k (x)sn2k (y) − 2k 2 sn2k (x)sn2k (y)
+2snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)]
= Dr 1
[(sn2k (x) − sn2k (x)sn2k (y))(1 − k 2 sn2k (y))
+(sn2k (y) − sn2k (x)sn2k (y))(1 − k 2 sn2k (x))
+2snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)]
= Dr 1
[sn2k (x)(1 − sn2k (y))(1 − k 2 sn2k (y))
+sn2k (y)(1 − sn2k (x))(1 − k 2 sn2k (x))
+2snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)]
= Dr 1
[sn2k (x)cn2k (y)dn2k (y) + sn2k (y)cn2k (x)dn2k (x)
+2snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)]
= Dr 1
(snk (x)cnk (y)dnk (y) + snk (y)cnk (x)dnk (x))2 .
Hence,
i i
i i
i i
Next,
dn2k (x + y)
= 1 − k 2 sn2k (x + y)
(snk (x)cnk (y)dnk (y) + snk (y)cnk (x)dnk (x))2
= 1 − k2
(1 − k 2 sn2k (x)sn2k (y))2
Nr
= ,
(1 − k sn2k (x)sn2k (y))2
2
where
Nr = 1 + k 4 sn4k (x)sn4k (y) − 2k 2 sn2k (x)sn2k (y)
−k 2 [sn2k (x)cn2k (y)dn2k (y) + sn2k (y)cn2k (x)dn2k (x)
+2snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)]
= 1 + k 4 sn4k (x)sn4k (y) − 2k 2 sn2k (x)sn2k (y)
−k 2 [sn2k (x)(1 − sn2k (y))(1 − k 2 sn2k (y)) + sn2k (y)
(1 − sn2k (x))(1 − k 2 sn2k (x))
+2snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)]
= 1 + k 4 sn4k (x)sn4k (y) − 2k 2 sn2k (x)sn2k (y)
−k 2 [(sn2k (x) − sn2k (x)sn2k (y))(1 − k 2 sn2k (y))
+(sn2k (y) − sn2k (x)sn2k (y))(1 − k 2 sn2k (x))
+2snk (x)snk (y)cnk (x)cnk (y)]dnk (x)dnk (y)]
= 1 + k 4 sn4k (x)sn4k (y) − 2k 2 sn2k (x)sn2k (y)
−k 2 [sn2k (x) + sn2k (y) − 2sn2k (x)sn2k (y) − 2k 2 sn2k (x)sn2k (y)
+2snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)
= (1 + k 4 sn2k (x)sn2k (y) − k 2 sn2k (x) − k 2 sn2k (y))
+k 4 sn2k (x)sn2k (y) − k 4 sn4k (x)sn2k (y) − k 4 sn2k (x)sn4k (y)
+k 4 sn4k (x)sn4k (y) − 2k 2 snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)
= (1 − k 2 sn2k (x))(1 − k 2 sn2k (y)) + k 4 sn2k (x)sn2k (y)
+k 4 sn2k (x)sn2k (y)[1 − sn2k (x) − sn2k (y) + sn2k (x)sn2k (y)]
−2k 2 snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)
= dn2k (x)dn2k (y) + k 4 sn2k (x)sn2k (y)(1 − sn2k (x))(1 − sn2k (y))
−2k 2 snk (x)snk (y)cnk (x)cnk (y)dnk (x)dnk (y)
= (dnk (x)dnk (y) − k 2 snk (x)snk (y)cnk (x)cnk (y))2 .
Therefore,
i i
i i
i i
In the last expression, we have chosen the sign by considering the particular
case k = 0 as before.
sck (x)dnk (x) + sck (y)dnk (y)
Exercise 8.5.10 Prove that sck (x + y) = .
1 − sck (x)sck (y)dnk (x)dnk (y)
snk (x + y)
Hint: Using the previous theorem, find and simplify.
cnk (x + y)
i i
i i
i i
x
cnk (x)+dnk (x)
PROBLEM 8.5.11 Prove that cnk 2 = 1+dnk (x) .
Solution: As
x x
cn2k −sn2k 2 dnk 2
2 x dn2k 2x −k 2 sn2k 2x cn2k 2x
2
+
cnk (x) + dnk (x) 1 − k 2 sn4k 2x 1 − k 2 sn4k 2x
=
1 + dnk (x) dn2k 2x − k 2 sn2k 2x cn2k 2x
1+
1 − k 2 sn4k 2x
cn2k 2x −sn2k 2x dn2k 2x +dn2k 2x − k 2 sn2k 2x cn2k 2x
=
1 − k 2 sn4k 2x + dn2k 2x − k 2 sn2k 2x cn2k 2x
cn2k 2x 1 − k 2 sn2k 2x + dn2k 2x 1 − sn2k 2x
=
1 − k 2 sn2k 2x sn2k 2x + cn2k 2x + dn2k 2x
2cn2k 2x dn2k 2x
=
1 − k 2 sn2k 2x + dn2k 2x
2cn2k 2x dn2k 2x
=
2dn2k 2x
x
= cn2k .
2
x
(x)+dnk (x)
Thus, we get cnk 2 = cnk1+dn k (x)
.
By adopting the same technique, the reader can prove the following
identities.
By seeing the Theorem 8.5.9, we extend the Jacobi’s functions snk , sck ,
and dnk as follows.
i i
i i
i i
PROBLEM 8.5.14 Prove that if 0 < k < 1, then cnk , snk , and dnk are elliptic
functions.
Solution: First, we note that the values of the three functions, namely, cnk ,
snk , and dnk at x+iy have the same denominator cn2j (y)+k 2 sn2k (x)sn2j (y), and
it is zero iff cnj (y) = 0 and ksnk (x)snj (y) = 0 iff cnj (y) = 0 and snk (x) = 0,
as sn2j (y) = 1.
From Theorems 8.5.2 and 8.5.3, we have
i i
i i
i i
Similarly, using the facts (i) cos 0 on (0, π2 ) and (ii) Fk maps (0, π2 ) onto
(0, Kk ), we obtain cnk 0 on (0, Kk ). Therefore,
Thus, i2Kj and 4Kk are periods of snk ; 2Kk + i2Kj and 4Kk are periods of
cnk ; and i4Kj and 2Kk are periods of dnk .
Next, we show that snk is analytic on . We denote the real and imaginary
parts of snk (x + iy) (see Definition 8.5.13.), respectively, by
i i
i i
i i
∂U
(x, y)
∂x
1 $ 2
= (cnj (y) + k 2 sn2k (x)sn2j (y))(cnk (x)dnk (x)dnj (y))
Dr %
−snk (x)dnj (y)(2k 2 snk (x)cnk (x)dnk (x)sn2j (y))
1 $
= cnk (x)dnk (x)cn2j (y)dnj (y) + k 2 sn2k (x)cnk (x)dnk (x)sn2j (y)dnj (y)
Dr %
−2k 2 sn2k (x)cnk (x)dnk (x)sn2j (y)dnj (y)
1 $ %
= cnk (x)dnk (x)cn2j (y)dnj (y) − k 2 sn2k (x)cnk (x)dnk (x)sn2j (y)dnj (y)
Dr
and
∂V
(x, y)
∂y
1 $ 2
= (cnj (y) + k 2 sn2k (x)sn2j (y))(cnk (x)dnk (x)(cn2j (y)dnj (y)
Dr
−sn2j (y)dnj (y))) − snj (y)cnj (y)cnk (x)dnk (x)(−2cnj (y)snj (y)dnj (y)
'
+2k 2 sn2k (x)snj (y)cnj (y)dnj (y)
1 $
= cnk (x)dnk (x)cn4j (y)dnj (y) − cnk (x)dnk (x)sn2j (y)cn2j (y)dnj (y)
Dr
+k 2 sn2k (x)cnk (x)dnk (x)sn2j (y)cn2j (y)dnj (y) − k 2 sn2k (x)cnk (x)dnk (x)
sn4j (y)dnj (y) + 2cnk (x)dnk (x)sn2j (y)cn2j (y)dnj (y)
%
−2k 2 sn2k (x)cnk (x)dnk (x)sn2j (y)cn2j (y)dnj (y)
1 $
= cnk (x)dnk (x)cn2j (y)dnj (y)(cn2j (y) + sn2j (y))
Dr %
−k 2 sn2k (x)cnk (x)dnk (x)sn2j (y)dnj (y)(sn2j (y) + cn2j (y))
1 $ %
= cnk (x)dnk (x)cn2j (y)dnj (y) − k 2 sn2k (x)cnk (x)dnk (x)sn2j (y)dnj (y)
Dr
∂U
= (x, y).
∂x
Next, we verify the other C–R equation. Although it not as simple as the
previous one, we verify it as follows:
∂U
Dr (x, y)
∂y
= (cn2j (y) + k 2 sn2k (x)sn2j (y))(−j2 snk (x)snj (y)cnj (y))
−(snk (x)dnj (y))(−2cnj (y)snj (y)dnj (y) + 2k 2 sn2k (x)snj (y)cnj (y)dnj (y))
= −j2 snk (x)snj (y)cn3j (y) − j2 k 2 sn3k (x)sn3j (y)cnj (y)
i i
i i
i i
+2snk (x)snj (y)cnj (y)dn2j (y) − 2k 2 sn3k (x)snj (y)cnj (y)dn2j (y)
= (k 2 − 1)snk (x)snj (y)cn3j (y) − (1 − k 2 )k 2 sn3k (x)sn3j (y)cnj (y)
+(2snk (x)snj (y)cnj (y) − 2k 2 sn3k (x)snj (y)cnj (y))(1 − (1 − k 2 )sn2j (y))
= −snk (x)snj (y)cn3j (y) + k 2 snk (x)snj (y)cn3j (y) − k 2 sn3k (x)sn3j (y)cnj (y)
+k 4 sn3k (x)sn3j (y)cnj (y) + 2snk (x)snj (y)cnj (y) − 2snk (x)sn3j (y)cnj (y)
+2k 2 snk (x)sn3j (y)cnj (y) − 2k 2 sn3k (x)snj (y)cnj (y)
+2k 2 sn3k (x)sn3j (y)cnj (y) − 2k 4 sn3k (x)sn3j (y)cnj (y)
= −snk (x)snj (y)cnj (y)[cn2j (y) + sn2j (y)] − snk (x)sn3j (y)cnj (y)
+k 2 snk (x)snj (y)cn3j (y) + k 2 sn3k (x)sn3j (y)cnj (y)−k 4 sn3k (x)sn3j (y)cnj (y)
+2snk (x)snj (y)cnj (y) + 2k 2 snk (x)sn3j (y)cnj (y)−2k 2 sn3k (x)snj (y)cnj (y)
= −snk (x)snj (y)cnj (y) − snk (x)sn3j (y)cnj (y) + k 2 snk (x)snj (y)cnj (y)
[cn2j (y) + sn2j (y)] + k 2 snk (x)sn3j (y)cnj (y) + k 2 sn3k (x)sn3j (y)cnj (y)
+2snk (x)snj (y)cnj (y) − 2k 2 sn3k (x)snj (y)cnj (y) − k 4 sn3k (x)sn3j (y)cnj (y)
= snk (x)snj (y)cnj (y) − snk (x)sn3j (y)cnj (y) + k 2 snk (x)snj (y)cnj (y)
+k 2 sn3k (x)sn3j (y)cnj (y) − 2k 2 sn3k (x)snj (y)cnj (y)
+k 2 snk (x)sn3j (y)cnj (y) − k 4 sn3k (x)sn3j (y)cnj (y)
= snk (x)snj (y)cnj (y)(1 − sn2j (y)) − k 2 sn3k (x)snj (y)cnj (y)(1 − sn2j (y))
+k 2 snk (x)snj (y)cnj (y) + k 2 snk (x)sn3j (y)cnj (y) − k 4 sn3k (x)sn3j (y)cnj (y)
−k 2 sn3k (x)snj (y)cnj (y)
= snk (x)snj (y)cn3j (y) − k 2 sn3k (x)snj (y)cn3j (y) + k 2 snk (x)snj (y)cnj (y)
+k 2 snk (x)sn3j (y)cnj (y) − k 4 sn3k (x)sn3j (y)cnj (y) − k 2 sn3k (x)snj (y)cnj (y)
and
∂V
Dr (x, y)
∂x
= (cn2j (y) + k 2 sn2k (x)sn2j (y))snj (y)cnj (y)[−k 2 snk (x)cn2k (x) − snk (x)dn2k (x)]
−(snj (y)cnj (y)cnk (x)dnk (x))(2k 2 sn2j (y)snk (x)cnk (x)dnk (x))
= −k 2 snk (x)cn2k (x)snj (y)cn3j (y) − snk (x)dn2k (x)snj (y)cn3j (y)
−k 4 sn3k (x)cn2k (x)sn3j (y)cnj (y) − k 2 sn3k (x)dn2k (x)sn3j (y)cnj (y)
−2k 2 snk (x)cn2k (x)dn2k (x)sn3j (y)cnj (y)
$
= −k 2 snk (x)snj (y)cn3j (y) − k 4 sn3k (x)sn3j (y)cnj (y)
%
−2k 2 snk (x)(1 − k 2 sn2k (x))sn3j (y)cnj (y) (1 − sn2k (x))
$ %
+ −snk (x)snj (y)cn3j (y) − k 2 sn3k (x)sn3j (y)cnj (y) (1 − k 2 sn2k (x))
= −k 2 snk (x)snj (y)cn3j (y) − k 4 sn3k (x)sn3j (y)cnj (y) − 2k 2 snk (x)sn3j (y)cnj (y)
i i
i i
i i
+2k 4 sn3k (x)sn3j (y)cnj (y) + k 2 sn3k (x)snj (y)cn3j (y) + k 4 sn5k (x)sn3j (y)cnj (y)
+2k 2 sn3k (x)sn3j (y)cnj (y) − 2k 4 sn5k (x)sn3j (y)cnj (y) − snk (x)snj (y)cn3j (y)
−k 2 sn3k (x)sn3j (y)cnj (y) + k 2 sn3k (x)snj (y)cn3j (y) + k 4 sn5k (x)sn3j (y)cnj (y)
= −k 2 snk (x)snj (y)cn3j (y) − 2k 2 snk (x)sn3j (y)cnj (y) + k 4 sn3k (x)sn3j (y)cnj (y)
+k 2 sn3k (x)sn3j (y)cnj (y) + 2k 2 sn3k (x)snj (y)cn3j (y) − snk (x)snj (y)cn3j (y)
= −k 2 snk (x)snj (y)cnj (y)[cn2j (y) + sn2j (y)] − k 2 snk (x)sn3j (y)cnj (y)
+k 4 sn3k (x)sn3j (y)cnj (y) + k 2 sn3k (x)snj (y)cnj (y)[sn2j (y) + cn2j (y)]
+k 2 sn3k (x)snj (y)cn3j (y) − snk (x)snj (y)cn3j (y)
= −k 2 snk (x)snj (y)cnj (y) − k 2 snk (x)sn3j (y)cnj (y) + k 4 sn3k (x)sn3j (y)cnj (y)
+k 2 sn3k (x)snj (y)cnj (y) + k 2 sn3k (x)snj (y)cn3j (y) − snk (x)snj (y)cn3j (y).
Therefore, the real and imaginary parts of snk satisfy the C–R equations
and obviously the partial derivatives are continuous on , and hence, snk is
analytic on and meromorphic on C, as snk has poles on C \ . Thus, snk
is an elliptic function. Using Theorem 6.3.2, we have cnk and dnk , which are
meromorphic functions on , and hence, they are also elliptic functions.
Remark 8.5.16: All the identities involving the Jacobian elliptic functions
holds for all z ∈ also, using the principle of analytic continuation.
i i
i i
This page is intentionally left blank
i i
Bibliography
455
i i
i i
This page is intentionally left blank
i i
Index
A C
Abel’s limit theorem, 100 canonical product, 270
Abel’s theorem on convergence of a Cantor’s intersection theorem, 49
power series, 94 Cartesian form, 18
Absolutely, 33 Cartesian product, 4
abelian group, 8 Cauchy criterion, 28
addition, 9 Cauchy product, 228
addition theorem for exp, 104 Cauchy sequence, 25
addition theorems for the Jacobi’s elliptic Cauchy’s estimate, 206
functions, 441 Cauchy’s integral formula, 193
adjacent side, 114 Cauchy’s integral formula for
Algorithm, 82, 144 derivatives, 196
analytic, 74 Cauchy’s residue theorem, 294
Analytic branch, 150 Cauchy’s theorem for a simply connected
√ region, 191
analytic branch for ·, 153
arc length, 176 Cauchy’s theorem for rectangle, 187
Archimedian property, 6 Cauchy-Goursat theorem, 193
Arccos, 154 Cauchy-Riemann equations, 79
arcsin(w), 154 Cauchy-Riemann equations in polar
arctan, 117 form, 90
Cauchy-Schwarz inequality, 16
B Chain rule, 73
Circle, 24
basic Jacobian elliptic functions, 435 closed curve, 172
bijective, 7 Closed set, 39
Binary operator, 7 closure, 38
Bolzano-Weierstrass property, 51 codomain, 6
Bounded, 25 Complete metric space, 51
bounded above, 5, 6 compact set, 48
bounded below, 5, 6 comparison test, 35
bounded sequence, 25, 50, 368 complex conjugate, 12
branch, 150 complex logarithm, 150
457
i i
i i
i i
458 Index
Complex Numbers, 9 E
complex version of mean-value
elementary linear fractional
theorem, 77
transforms, 132
component, 46
elliptic curve, 392
conformal, 157 elliptic function, 395
Conformal mapping, 156 elliptic integral, 392
conformal of first type, 156 Elliptic integrals of first kind, 392
conformal of second type, 157 elliptic integrals of second kind, 392
Continuity, 55 Elliptic integrals of third kind, 392
Continuous, 62 Equation of the straight line, 23
continuous at z0 , 62 Equation of a circle, 24
continuous on , 62 entire functions, 207
convergent, 24 essential singularity, 249
convergent sequence, 25 Euler’s formula, 106
Connected, 41 Euler’s normal form, 434
cos(z), 105 evaluation of real integrals, 314
cot(z), 115 exponential function, 102
coth(z), 121 Extended complex numbers, 52
cosh(z), 119
F
cross ratio, 139
csc(z), 115 field, 8
csch(z), 121 First addition theorem for ℘a,b , 419
curve, 156 fixed points, 135
cycle, 212 function, 6
Functions za,b and sa,b , 426
D fundamental parallelogram, 398
fundamental period, 393
de Moivre’s formula, 106, 107 fundamental periods of cnk , 453
De-Morgan’s laws, 4 fundamental periods of dnk , 453
decreasing function, 7 fundamental periods of snk , 453
Define π , 109 Fundamental theorem of algebra, 207
degree, 123 Fundamental theorem of calculus, 170
Dense subset, 41
derivative, 73 G
Degree of the polynomial, 123 general version of Cauchy’s
differentiable, 73 theorem, 209
disconnected, 42 generalized Schwarz lemma, 217
Division algorithm for polynomials, 123 geometric series, 36
domain, 6 greatest lower bound property, 6
i i
i i
i i
Index 459
i i
i i
i i
460 Index
i i
i i
i i
Index 461
U Z
Unary operator, 7 zero, 212, 246, 328
unbounded region determined by γ , 183 Zero of order k , 129, 397
i i
i i
This page is intentionally left blank