Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

MA547 Lecture Summary

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

MA 547 (Complex Analysis)

Summary of Lectures

Motivation for introducing complex numbers: The quadratic equation x2 + 1 = 0 has no


root p
in R. Using Cardano’s method, the roots of the cubic equation x3 = 15x + 4 are given by
√ p √
x = 3 2 + 11 −1 + 3 2 − 11 −1. Since x = 4 is clearly a root of the equation x3 = 15x + 4,

there was a need to give a meaning of −1 by enlarging the number system R.

The field of complex numbers: The field C of complex numbers is defined as the field having
the following properties:
(a) C contains (an isomorphic image of) the field R.
(b) C contains a root of the equation x2 + 1 = 0.
(c) If K is a subfield of C containing R and a root of the equation x2 + 1 = 0, then K = C.

Existence of C : The existence of the field C is provided by any of the following constructions
(models) of the field C.
(a) C = {(a, b) : a, b ∈ R}, where addition and multiplication are defined as follows:
(a, b) + (c, d) = (a + c, b + d) and (a, b).(c, d) = (ac − bd, ad + bc) for all (a, b), (c, d) ∈ C.
(b) C = {a+ib : a, b ∈ R}, where +, i are symbols and the addition and multiplication are defined
as follows:
(a + ib) + (c + id) = (a + c) + i(b + d) and (a + ib).(c + id) = (ac − bd) + i(ad + bc) for all
c + id ∈ C.
a + ib,  
a b
(c) C = : a, b ∈ R , with the usual addition and multiplication of matrices.
−b a
(d) C is the quotient ring R[x]hx2 + 1i .

Uniqueness of C : The uniqueness (up to isomorphism) of the field C can also be proved. In
fact, if C and C0 are fields satisfying the properties (a), (b) and (c) given in the definition of the
field of complex numbers, then there exists an isomorphism f : C → C0 which maps reals to reals
and also maps a root of the equation x2 + 1 = 0 to a root of the equation x2 + 1 = 0.

Real and imaginary parts: If i ∈ C is a root of the equation x2 + 1 = 0, then each z ∈ C can
be uniquely expressed as z = x + iy, where x, y ∈ R. We say that x = Re(z) and y = Im(z) are
respectively the real part and the imaginary part of z.
If z, w ∈ C, then Re(z + w) = Re(z) + Re(w) and Im(z + w) = Im(z) + Im(w), although Re(zw)
need not be equal to Re(z)Re(w) and Im(zw) need not be equal to Im(z)Im(w).

Order: A field F is said to be an ordered field if there exists a nonempty subset P of F satisfying
the following properties:
(a) P is closed under addition and multiplication, i.e. x + y ∈ P and xy ∈ P for all x, y ∈ P .
(b) (Law of trichotomy) For each x ∈ F , exactly one of x ∈ P or x = 0 or −x ∈ P holds.
In such case, we define a relation < on F as follows: For all x, y ∈ F , let x < y (equivalently,
y > x) if y − x ∈ P . We also write x ≤ y (equivalently, y ≥ x) if x < y or x = y. Note that ≤ is
a partial order on F .
Although partial orders (like, dictionary order) can be defined on the set C, the field C cannot be
made an ordered field.
C as a vector space: C is a vector space of dimension 1 over the field C. A basis for this vector
space is {1}.
C is a vector space of dimension 2 over the field R. A basis for this vector space is {1, i}.

Conjugate: For each z = x + iy ∈ C, the conjugate of z is defined as z = x − iy.


If z, w ∈ C, then
(a) Re(z) = z+z 2
and Im(z) = z−z
2i
.
(b) z = z.
(c) z + w = z + w and zw = z w.
More generally, if n ∈ N and z1 , . . . , zn ∈ C, then z1 + · · · + zn = z 1 + · · · + z n and
z1 . . . zn = z 1 . . . z n .
(d) ( wz ) = wz , provided that w 6= 0.
If p(z) = a0 + a1 z + · · · + an z n , where a0 , . . . , an ∈ R, and if z0 ∈ C is a root of the equation
p(z) = 0, then z 0 is also a root of the equation p(z) = 0.
The conjugation map z 7→ z from C to C is an automorphism of the field C whose restriction to R
is the identity map on R. Also, if f : C → C is an automorphism of the field C such that f (R) ⊂ R,
then f (z) = z for all z ∈ C or f (z) = z for all z ∈ C. However, there exist automorphisms of the
field C which are neither the identity map nor the conjugation map.

Linear map: A map T : C → C is C-linear iff there exists a ∈ C such that T (z) = az for all
z ∈ C.
A map T : C → C is R-linear iff there exist a, b ∈ C such that T (z) = az + bz for all z ∈ C.
Clearly, every C-linear map T : C → C is R-linear but an R-linear map T : C → C need not be
C-linear. For example, take T (z) = z for all z ∈ C.
However, an R-linear map T : C → C is C-linear iff T (i) = iT (1).
With respect to the ordered
 basis {1, i} of the vector space C (over R), for all a, b, c, d ∈ R, the
a b
2 × 2 matrix represents an R-linear map T : C → C and conversely every R-linear map
c d
 
a b
T : C → C has a matrix representation , where a, b, c, d ∈ R.
c d
An R-linear map T : C → C with above matrix representation is a C-linear map iff a = d and
b = −c.

Complex plane: By identifying z = x + iy ∈ C with the point (x, y) in the xy-plane, we get a
one-one correspondence between C and the points in the xy-plane.
The complex numbers can also be represented as vectors in the xy-plane, where 0 ∈ C is identified
−→
with the zero vector and z = x + iy ∈ C \ {0} is identified with the vector OP , where O = (0, 0)
and P = (x, y).
When the complex numbers are represented as points (or, vectors) in the xy-plane, the xy-plane
is called the complex plane or the z-plane or the Gaussian plane and the diagram is referred to as
the Argand diagram. Also, in this case, the x-axis is called the real axis and the y-axis is called
the imaginary axis.

Modulus:
p For each z = x + iy ∈ C, the modulus or the absolute value of z is defined as
|z| = x + y 2 .
2

If z, w ∈ C, then
(a) |z| ≥ 0 and |z| = 0 ⇔ z = 0.
(b) |z| = |z|.
(c) |Re(z)| ≤ |z| and |Im(z)| ≤ |z|.
(d) |zw| = |z||w|.
= |z| , provided that w 6= 0.
z
(e) w |w|
(f) (Triangle inequality) |z + w| ≤ |z| + |w| with equality iff either w = 0 or else z = tw for some
t ∈ R with t ≥ 0.
More generally, if n ∈ N and z1 , . . . , zn ∈ C, then |z1 + · · · + zn | ≤ |z1 | + · · · + |zn |.

(g) |z| − |w| ≤ |z − w| with equality iff either w = 0 or else z = tw for some t ∈ R with t ≥ 0.
(h) (Parallelogram law) |z + w|2 + |z − w|2 = 2(|z|2 + |w|2 ).

Equation of a circle: If z0 = x0 + iy0 ∈ C and r > 0, then the equation |z − z0 | = r represents


the circle in the complex plane with centre (x0 , y0 ) and radius r.

Equation of a straight line: The general equation of a straight line in the complex plane is
az + az + c = 0, where a ∈ C \ {0} and c ∈ R.
−→
Polar representation: Let z = x + iy ∈ C \ {0} be represented by the vector OP in the complex
p −→
plane. If r = x2 + y 2 and θ is an angle which OP makes with the positive real axis, then
x = r cos θ, y = r sin θ and so z = r(cos θ + i sin θ). We say that θ is an argument or amplitude of
z and we define arg(z) = {θ ∈ R : z = |z|(cos θ + i sin θ)}.
The unique θ ∈ arg(z) satisfying θ ∈ (−π, π] is called the principal argument of z and it is denoted
by Arg(z). Thus arg(z) = {Arg(z) + 2nπ : n ∈ Z}.
If z, w ∈ C \ {0}, then arg(zw) = arg(z) + arg(w) and arg wz = arg(z) − arg(w).


However, if z, w ∈ C \ {0}, then Arg(zw) need not be equal to Arg(z) + Arg(w) (for example, take
z = −1, w = i) and Arg wz need not be equal to Arg(z) − Arg(w) (for example, take z = −1 and


w = −i).

Square roots: Let w = a + ib ∈ C. In order to find all z = x + iy ∈ C such that z 2 = w, we


solve the simultaneous equations x2 − y 2 = a and 2xy = b.

If b = 0 and a ≥ 0, then z = ± a.

z = ±i −a.
If b = 0 and a < 0, thenp
1
√ b
p √ 
If b 6= 0, then z = ± 2
√ 2 2 2
a + a + b + i |b| −a + a + b . 2

Thus for each w ∈ C \ {0}, there exist exactly two distinct z ∈ C such that z 2 = w.
p√ p√ 
Example: The square roots of 1 + i are ± √12 2+1+i 2−1 .

de Moivre’s formula: (cos θ + i sin θ)n = cos nθ + i sin nθ for all θ ∈ R, n ∈ Z.


√ √
Using de Moivre’s formula, we get (−1 + i 3)100 = 2100 cos 200π 200π
 99
3
+ i sin 3
= 2 (−1 + i 3).

nth roots: Let w = a + ib ∈ C \ {0}. The method of finding square roots of w, as described
above, is not helpful in finding nth roots of w for n > 2. However, using de Moivre’s formula, we
1
find that there exist exactly n distinct nth roots of w and these are |w| n cos 2kπ+α 2kπ+α

n
+ i sin n
,
where α = Arg(w) and k ∈ {0, 1, . . . , n − 1}. All these nth roots of w are the vertices of a regular
1
n-gon which is inscribed in the circle |z| = |w| n in the complex plane.
1
Example: The seventh roots of 1 + i are 2 14 cos (8n+1)π + i sin (8n+1)π

28 28
for n = 0, 1, . . . , 6.

C as a metric space: If du (z, w) = |z − w| for all z, w ∈ C, then du is a metric on C, which is


called the usual metric on C. Thus (C, du ) is a metric space.
If E(6= ∅) ⊂ C, then E is also a metric space with respect to the (restriction to E × E of the)
metric du .
In the following, we give definitions and results in the metric space (C, du ). However, they can
also be formulated in the metric space (E, du ).

Open and closed balls: Let z0 ∈ C and r > 0. Then Br (z0 ) = {z ∈ C : |z − z0 | < r} is the
open ball (or, open disk) in C with centre z0 and radius r. Also, Br [z0 ] = {z ∈ C : |z − z0 | ≤ r}
is the closed ball (or, closed disk) in C with centre z0 and radius r.
We denote the open unit disk B1 (0) = {z ∈ C : |z| < 1} by D.

Open and closed sets: A subset G of C is called an open set in C if for each z ∈ G, there exists
r > 0 such that Br (z) ⊂ G.
A subset F of C is called a closed set in C if C \ F is an open set in C.
Let E(6= ∅) ⊂ C. Then S ⊂ E is open (respectively, closed) in E iff S = G ∩ E for some open
(respectively, closed) set G in C.

Examples:
(a) Every open ball in C is an open set in C and every closed ball in C is a closed set in C.
(b) {z ∈ C : 1 < Re(z) < 2} is an open set in C and {z ∈ C : Im(z) ≥ 0} is a closed set in C.
(c) {z ∈ C : 1 < |z| ≤ 2} is neither open nor closed in C.
(d) E = {z ∈ C : |z| ≤ 1} is both open and closed in E ∪ {z ∈ C : |z − 3| < 1}.

Equivalent metrics on C : Let d1 (z, w) = |Re(z) − Re(w)| + |Im(z) − Im(w)| and


d∞ (z, w) = max{|Re(z) − Re(w)|, |Im(z) − Im(w)|} for all z, w ∈ C. Then d1 and d∞ are metrics
on C and the three metrics du , d1 , d∞ on C are equivalent, i.e. each of the three metric spaces
(C, du ), (C, d1 ) and (C, d∞ ) has the same class of open sets.

Bounded set and bounded function: A subset E of C is said to be bounded if there exists
r > 0 such that E ⊂ Br [0], i.e. |z| ≤ r for all z ∈ E. A set in C which is not bounded is called
unbounded.
For example, {z ∈ C : (Re(z))2 + |Im(z)| < 5} is a bounded set in C and
{z ∈ C : |Re(z)| + Im(z) < 1} is an unbounded set in C
If Ω is a nonempty subset of C, then a function f : Ω → C is said to be bounded if f (Ω) is a
bounded subset of C, i.e. if there exists r > 0 such that |f (z)| ≤ r for all z ∈ Ω.

Convergence of sequence: A sequence in C is a function f : N → C, which we usually denote


by (zn ) by writing zn = f (n) for each n ∈ N.
A sequence (zn ) in C is called convergent if there exists z ∈ C such that for every ε > 0, there
exists n0 ∈ N satisfying |zn − z| < ε for all n ≥ n0 .
Such a z (if it exists) is unique and is called the limit of (zn ), which is denoted by zn → z or
lim zn = z.
n→∞
We note that every convergent sequence in C is bounded.
 
(1+i)n
Example: The sequence n!
is convergent with limit 0.
 
Proposition: A sequence (zn ) in C is convergent iff both the sequences Re(zn ) and Im(zn )
in R are convergent.
Also, in such case lim zn = lim Re(zn ) + i lim Im(zn ).
n→∞ n→∞ n→∞

Cauchy sequence: A sequence (zn ) in C is called a Cauchy sequence if for every ε > 0, there
exists n0 ∈ N such that |zm − zn | < ε for all m, n ≥ n0 .
Proposition: The metric space (C, du ) is complete, i.e. every Cauchy sequence in C is convergent.

Bolzano-Weierstrass theorem: Every bounded sequence in C has a convergent subsequence.


For example, the sequence (cos n + i sin n) has a convergent subsequence although the sequence
itself is not convergent.
 n
P
Series in C : A series in C is a pair (zn ), (sn ) , where (zn ) is a sequence in C and sn = zk
k=1

P
for all n ∈ N. A series in C is usually denoted by zn .
n=1

P
The series zn is said to be convergent if the sequence (sn ) is convergent. In such case the sum
n=1

P ∞
P
of the series zn is defined as zn = lim sn .
n=1 n=1 n→∞

P ∞
P
The series zn is said to be absolutely convergent if the series |zn | is convergent.
n=1 n=1
Every absolutely convergent series is convergent but the converse is not true.

Examples:
1
(a) The geometric series 1 + z + z 2 + · · · converges (absolutely) with sum 1−z for z ∈ D and
diverges (i.e. does not converge) for z ∈ C \ D.
∞ ∞
1
{z ∈ C : z n 6= −1}, then the series
T P
(b) If z ∈ 1+z n
is convergent iff |z| > 1.
n=1 n=1

Limit point: Let E ⊂ C. We say that z ∈ C is a limit point of E in C if for every r > 0,
Br (z) ∩ (E \ {z}) 6= ∅.
Using Bolzano-Weierstrass theorem for sequences, it follows that every infinite bounded subset of
C has a limit point in C.

Closure: Let E ⊂ C. The closure of E in C is defined as


T
E = {F : F is a closed set in C, E ⊂ F }.
Note that E is the smallest closed set in C containing E and also E = E ∪ E 0 , where E 0 is the set
of all limit points of E in C.
It can be shown that E = {z ∈ C : Br (z) ∩ E 6= ∅ for all r > 0}
= {z ∈ C : there exists a sequence (zn ) in E such that zn → z}.
If E = C, then E is said to be dense in C. For example, {x + iy : x, y ∈ Q} is dense in C.

Sequential criterion of closed set: A set E ⊂ C is closed in C iff for every z ∈ C and for every
sequence (zn ) in E converging to z, z ∈ E.
Using this criterion, it can be easily shown that {z ∈ C : Re(z) ≥ 0} is a closed set in C and
{z ∈ C : Im(z) < 0} is not closed in C.

Interior: Let E ⊂ C. The interior of E in C is defined as


E 0 = {G : G is an open set in C, G ⊂ E}.
S

Note that E 0 is the largest open set in C contained in E. It can be seen that
E 0 = {z ∈ E : Br (z) ⊂ E for some r > 0}.

Boundary: The boundary of a set E ⊂ C is defined as ∂E = E ∩ C \ E = E \ E 0 .


Thus ∂E = {z ∈ C : for every r > 0, Br (z) ∩ E 6= ∅ and Br (z) ∩ (C \ E) 6= ∅}.

Examples: Let z0 ∈ C and r > 0. Then Br (z0 ) = Br [z0 ], Br [z0 ]0 = Br (z0 ) and
∂Br (z0 ) = ∂Br [z0 ] = {z ∈ C : |z − z0 | = r}.
In particular, D = {z ∈ C : |z| ≤ 1} and ∂D = ∂D = {z ∈ C : |z| = 1}.
Function on C : Let f : Ω ⊂ C → C. Considering Ω0 = {(x, y) ∈ R2 : x+iy ∈ Ω} ⊂ R2 , we define
u : Ω0 → R and v : Ω0 → R respectively by u(x, y) = Re(f (x + iy)) and v(x, y) = Im(f (x + iy))
for all (x, y) ∈ Ω0 . We say that u and v are the real part and the imaginary part of f . Note that
f (z) = u(x, y) + iv(x, y) for all z = x + iy ∈ Ω.

Continuity: A function f : Ω ⊂ C → C is said to be continuous at z0 ∈ Ω if for every ε > 0,


there exists δ > 0 such that |f (z) − f (z0 )| < ε for all z ∈ Ω satisfying |z − z0 | < δ.
We say that f : Ω → C is continuous if f is continuous at each point of Ω.

Examples: Each of the the functions z 7→ Re(z), z 7→ Im(z), z 7→ z and z 7→ |z| from C to C is
continuous.

Proposition: Let u and v be respectively the real part and the imaginary part of f : Ω ⊂ C → C.
Then f is continuous at z0 = x0 + iy0 ∈ Ω iff both u and v are continuous at (x0 , y0 ).

Sequential criterion of continuity: A function f : Ω ⊂ C → C is continuous at z0 ∈ Ω iff for


every sequence (zn ) in Ω converging to z0 , f (zn ) → f (z0 ).
Re(z)
(
Example: The function f : C → C, defined by f (z) = if z 6= 0,
z
1 if z = 0,
is not continuous at 0.

Combinations of continuous functions: Let f : Ω ⊂ C → C and g : Ω → C be continuous at


z0 ∈ Ω. Then the functions f + g : Ω → C and f g : Ω → C are continuous at z0 . Moreover, if
g(z0 ) 6= 0, then there exists δ > 0 such that the function fg : Ω ∩ Bδ (z0 ) → C is continuous at z0 .
It follows that every polynomial function on C is continuous and every rational function
p
q
: {z ∈ C : q(z) 6= 0} → C is continuous.
Again, let f : Ω1 ⊂ C → C and g : Ω2 ⊂ C → C such that f (Ω1 ) ⊂ Ω2 . If f is continuous at
z0 ∈ Ω1 and g is continuous at f (z0 ), then the composite function g ◦ f : Ω1 → C is continuous at
z0 .

Examples:
(a) If f : Ω ⊂ C → C is continuous at z0 ∈ Ω, then the functions z 7→ f (z) and z 7→ |f (z)| from
Ω to C are continuous
 z+1 at z0 .
z 2 +4
if z ∈ C \ {2i, −2i},
(b) Let f (z) =
1 if z ∈ {2i, −2i}.
Then f : C → C is continuous on C \ {2i, −2i} and discontinuous at 2i and −2i.

Proposition: For a function f : Ω ⊂ C → C, the following statements are equivalent.


(a) f is continuous.
(b) For every open set G in C, f −1 (G) is open in Ω.
(c) For every closed set F in C, f −1 (F ) is closed in Ω.

Examples: {z ∈ C : 1 < |z + 2| < 3} is an open set in C and {z ∈ C : Re(z 2 + 3iz) = 6} is a


closed set in C.

Limit: Let z0 ∈ C be a limit point of Ω ⊂ C and let f : Ω → C. We say that ` ∈ C is a limit


of f at z0 if for every ε > 0, there exists δ > 0 such that |f (z) − `| < ε for all z ∈ Ω satisfying
0 < |z − z0 | < δ.
Such an `, when it exists, is unique and we write lim f (z) = `.
z→z0
Remark: A function f : Ω ⊂ C → C is continuous at z0 ∈ Ω iff either z0 is an isolated point of
Ω (i.e. there exists δ > 0 such that Bδ (z0 ) ∩ Ω = {z0 }) or else lim f (z) = f (z0 ).
z→z0

Proposition: Let z0 ∈ C be a limit point of Ω ⊂ C and let u and v be respectively the real part
and the imaginary part of f : Ω → C. Then lim f (z) exists (in C) iff both lim u(x, y) and
z→z0 (x,y)→(x0 ,y0 )
lim v(x, y) exist (in R).
(x,y)→(x0 ,y0 )
Further, in such case lim f (z) = lim u(x, y) + i lim v(x, y).
z→z0 (x,y)→(x0 ,y0 ) (x,y)→(x0 ,y0 )

Sequential criterion of limit: Let z0 ∈ C be a limit point of Ω ⊂ C and let f : Ω → C. Then


lim f (z) = ` ∈ C iff for every sequence (zn ) in Ω \ {z0 } converging to z0 , f (zn ) → `.
z→z0

zRe(z) z
Examples: lim = 0 but lim does not exist (in C).
z→0 |z| z→0 z

Compactness: A subset E of C is called compact if for every class {Gα : α ∈ Λ} of open sets in
S
C satisfying E ⊂ Gα , there exist α1 , . . . , αn ∈ Λ of such that E ⊂ Gα1 ∪ · · · ∪ Gαn .
α∈Λ
It can be shown that E ⊂ C is compact iff every sequence in E has a subsequence converging to
a point of E.

Example: C and D are not compact.

Heine-Borel theorem: A subset E of C is compact if E is closed and bounded in C.


Also, every compact set in C is closed and bounded in C.

Example: D and ∂D are compact.

Continuity and compactness: Let Ω be a nonempty compact set in C. If f : Ω → C is


continuous, then f (Ω) is compact (and hence f (Ω) is closed and bounded in C) and there exist
z1 , z2 ∈ Ω such that |f (z1 )| = sup{|f (z)| : z ∈ Ω} and |f (z2 )| = inf{|f (z)| : z ∈ Ω}.

Example: There is no continuous function from D onto C or from ∂D onto D.

Uniform continuity: A function f : Ω ⊂ C → C is said to be uniformly continuous if for every


ε > 0, there exists δ > 0 such that |f (z) − f (w)| < ε for all z, w ∈ Ω satisfying |z − w| < δ.
If f : Ω → C is uniformly continuous, then f is continuous. However, the converse is not true in
general, as an example below shows.

Sequential criterion of uniform continuity: A function f : Ω ⊂ C → C is uniformly contin-


uous iff for every sequences (zn ) and (wn ) in Ω with |zn − wn | → 0, |f (zn ) − f (wn )| → 0.
1 1
Example: If f (z) = z−2i and g(z) = 1−z for all z ∈ D, then f : D → C is uniformly continuous
and g : D → C is continuous but not uniformly continuous.

Proposition: If Ω is a nonempty compact subset of C and f : Ω → C is continuous, then


f : Ω → C is uniformly continuous.

Connectedness: A subset E of C is said to be disconnected if there exist nonempty disjoint


open sets G and H in E such that E = G ∪ H.
E ⊂ C is called connected if it is not disconnected.
Note that a nonempty subset E of C is connected iff ∅ and E are the only subsets of E which are
both open and closed in E.
If E ⊂ C, then a maximal connected subset of E is called a component of E.
Proposition: Let Ω be a nonempty connected set in C. If f : Ω → C is continuous, then f (Ω) is
connected.

Example: Since D is connected (as seen below) and Z is disconnected, there is no continuous
function from D onto Z.

Path: Let E be a nonempty subset of C. A path in E is a continuous map γ : [a, b] → E for


some a, b ∈ R with a < b.
If γ : [a, b] → E is a path in E, then the trace (or, range) of γ is the set {γ} = {γ(t) : t ∈ [a, b]},
which is a compact and connected subset of E.
If z, w ∈ E and γ : [a, b] → E is a path such that γ(a) = z and γ(b) = w, then we say that γ joins
z to w, or γ is a path in E from z to w, or γ is a path in E with initial point z and final point w.
If γ1 : [a, b] → E and γ2 : [a, b] → E are paths in E such that γ1 (b) = γ2 (a), then the sum (or,
join/union) of  γ1 and γ2 is the path γ1 + γ2 : [a, b] → E, defined by
γ1 (2t − a) if t ∈ a, a+b
2 
,
(γ1 + γ2 )(t) =
γ2 (2t − b) if t ∈ a+b
2
, b .
If z, w ∈ C and γ(t) = (1 − t)z + tw for all t ∈ [0, 1], then γ is called the line segment in C
joining z to w. In this case we denote {γ} by [z, w]. If γ1 , . . . , γn : [0, 1] → C are finitely many
line segments in C such that γj (1) = γj+1 (0) for j = 1, . . . , n − 1, then γ1 + · · · + γn is called a
polygonal path (or, a polygon) in C.

Path connectedness: A subset E of C is said to be path connected if for each z, w ∈ E, there


exists a path in E from z to w.
If Ω is a nonempty path connected set in C and if f : Ω → C is continuous, then f (Ω) is path
connected.

Proposition: Every path connected set in C is connected.


However, a connected set in C need not be path connected. For example,
x + i sin x1 : 0 < x ≤ 1 ∪ {ix : −1 ≤ x ≤ 1} is connected but not path connected.


Examples of connected and disconnected sets:


(a) Every convex set in C is (path) connected. In particular, C and {z ∈ C : Re(z) > 1} are
(path) connected. Also, every open ball and every closed ball in C are (path) connected.
Consequently D and D are (path) connected.
(b) C \ {0} and ∂D are (path) connected.
(c) D ∪ {z ∈ C : |z + 2| ≤ 1} is (path) connected.
(d) D ∪ {z ∈ C : Im(z) = 1} is (path) connected.
(e) D ∪ {z ∈ C : |z − 3| ≤ 1} is disconnected.

Proposition: Let E be an open and connected set in C. Then E is path connected.


In fact, if z, w ∈ E, then there exists a polygonal path in E joining z to w. Moreover, each of the
line segments forming the polygonal path can be taken to be parallel to either the real axis or the
imaginary axis.

Domain: A (nonempty) open and connected set in C is called a domain in C.

Differentiability: A function f : Ω ⊂ C → C is said to be differentiable at z0 ∈ Ω0 if


lim f (z)−f
z−z0
(z0 )
(or, equivalently, lim f (z0 +h)−f
h
(z0 )
) exists (in C).
z→z0 h→0
If f is differentiable at z0 , then the derivative of f at z0 is defined as
f 0 (z0 ) = lim f (z)−f
z−z0
(z0 )
= lim f (z0 +h)−f
h
(z0 ) 
.
z→z0 h→0
If Ω is open in C, then f : Ω → C is called differentiable if f is differentiable at each point of Ω.

Example:
(a) If f (z) = z 2 for all z ∈ C, then f : C → C is differentiable and f 0 (z) = 2z for all z ∈ C.
(b) If f (z) = z for all z ∈ C, then f : C → C is not differentiable at any point of C.
(c) If f (z) = |z|2 for all z ∈ C, then f : C → C is differentiable only at 0 and f 0 (0) = 0.

Proposition: If f : Ω ⊂ C → C is differentiable at z0 ∈ Ω0 , then f is continuous at z0 .


The converse of this proposition is not true, in general. For example, if f (z) = |z| for all z ∈ C,
then f : C → C is continuous but not differentiable at 0.

Combinations of differentiable functions: Let f : Ω ⊂ C → C and g : Ω → C be differentiable


at z0 ∈ Ω0 . Then the functions f + g : Ω → C and f g : Ω → C are differentiable at z0 . Moreover,
if g(z0 ) 6= 0, then there exists δ > 0 such that Bδ (z0 ) ⊂ Ω and the function fg : Bδ (z0 ) → C
differentiable at z0 .
It follows that every polynomial function on C is differentiable and every rational function
p
q
: {z ∈ C : q(z) 6= 0} → C is differentiable.
 z+3
z 2 −3z+2
if z ∈ C \ {1, 2},
Example: Let f (z) =
1 if z ∈ {1, 2}.
Then f : C → C is differentiable on C \ {1, 2} and not differentiable at 1 and 2.

Chain rule: Let f : Ω1 ⊂ C → C and g : Ω2 ⊂ C → C be such that f (Ω1 ) ⊂ Ω2 . If f is


differentiable at z0 ∈ Ω01 and g is differentiable at f (z0 ) ∈ Ω02 , then g ◦ f : Ω1 → C is differentiable
at z0 and (g ◦ f )0 (z0 ) = g 0 (f (z0 ))f 0 (z0 ).

Remark: The following criterion of differentiability is useful in proving the chain rule.
A function f : Ω ⊂ C → C is differentiable at z0 ∈ Ω0 iff there exist a function ϕ : Ω → C such
that ϕ is continuous at z0 and f (z) − f (z0 ) = (z − z0 )ϕ(z) for all z ∈ Ω.
Also, in such case f 0 (z0 ) = ϕ(z0 ).

Complex differentiability and real differentiability: Let u and v be respectively the real
part and the imaginary part of f : Ω ⊂ C → C and let z0 = x0 + iy0 ∈ Ω0 .
(a) If f is differentiable at z0 , then all the partial derivatives ux , uy , vx and vy exist at (x0 , y0 )
and ux (x0 , y0 ) = vy (x0 , y0 ) and uy (x0 , y0 ) = −vx (x0 , y0 ).
Also, in this case f 0 (z0 ) = ux (x0 , y0 ) + ivx (x0 , y0 ).
(b) If all the partial derivatives ux , uy , vx and vy exist at (x0 , y0 ) and ux (x0 , y0 ) = vy (x0 , y0 ) and
uy (x0 , y0 ) = −vx (x0 , y0 ), then fneed not be differentiable at z0 . For example, for the function
z2
if z 6= 0,
f : C → C, defined by f (z) = z
0 if z = 0,
ux (0, 0) = 1 = vy (0, 0) and uy (0, 0) = 0 = −vx (0, 0) but f is not differentiable at 0.
(c) f is differentiable at z0 iff u and v are differentiable at (x0 , y0 ) and ux (x0 , y0 ) = vy (x0 , y0 ) and
uy (x0 , y0 ) = −vx (x0 , y0 ).
In particular, therefore, if there exists δ > 0 such that ux , uy , vx , vy exist on Bδ ((x0 , y0 )),
are continuous at (x0 , y0 ) and ux (x0 , y0 ) = vy (x0 , y0 ) and uy (x0 , y0 ) = −vx (x0 , y0 ), then f is
differentiable at z0 .
(d) Let Ω0 = {(x, y) ∈ R2 : x + iy ∈ Ω} and let f˜(x, y) = (u(x, y), v(x, y)) for all (x, y) ∈ Ω0 . Then
f is differentiable (as a function of a complex variable) at z0 iff f˜ : Ω0 → R2 is differentiable
(as a function of two real variables) at (x0 , y0 ) and the R-linear map f˜0 (x0 , y0 ) : R2 → R2 is a
C-linear map (i.e. a linear map over the field R2 ).
Put in an equivalent way, f is differentiable (as a function of a complex variable) at z0 iff f˜
is differentiable (as a function of two real variables) at (x0 , y0 ) and ux (x0 , y0 ) = vy (x0 , y0 ) and
uy (x0 , y0 ) = −vx (x0 , y0 ).
The equations ux = vy and uy = −vx are known as the Cauchy-Riemann equations.

Example:
(a) If f (z) = y + ix for all z = x + iy ∈ C, then f : C → C is not differentiable at any point of C.
(b) If f (z) = ex (cos y + i sin y) for all z = x + iy ∈ C, then f : C → C is differentiable and
f 0 (z) = f (z) for all z ∈ C.
(c) If f (z) = zRe(z) for all z ∈ C, then f : C → C is differentiable only at 0 and f 0 (0) = 0.

Looman-Menchoff theorem: Let u and v be the real part and the imaginary part of a continu-
ous function f : Ω → C, where Ω is an open set in C. If ux (x, y) = vy (x, y) and uy (x, y) = −vx (x, y)
for all (x, y) ∈ R2 with x + iy ∈ Ω, then f is differentiable (on Ω).
Montel stated (and Tolstov proved) that the continuity of f in the above result can be replaced
by boundedness of f .

Polar forms of Cauchy-Riemann equations: Let Ω be an open set in C and let f : Ω → C be


such that f (z) = u(x, y)+iv(x, y) = u(r, θ)+iv(r, θ) for all z = x+iy = r(cos θ +i sin θ) ∈ Ω\{0}.
The Cauchy-Riemann equations ux (x, y) = vy (x, y), uy (x, y) = −vx (x, y), expressed in polar forms,
become ur (r, θ) = 1r vθ (r, θ), 1r uθ (r, θ) = −vr (r, θ) for all z = x + iy = r(cos θ + i sin θ) ∈ Ω \ {0}.

Holomorphic function: If Ω is an open set in C, then a function f : Ω → C is said to be


holomorphic (or, analytic) if f is differentiable (on Ω).
If Ω is an open set in C, then the set of all holomorphic functions from Ω to C is denoted by H(Ω).
A holomorphic function f : C → C is also called an entire function.
We say that a function f : Ω ⊂ C → C is holomorphic at z0 ∈ Ω if there exists δ > 0 such that
Bδ (z0 ) ⊂ Ω and f is differentiable on Bδ (z0 ).

Example: If f (z) = e−x (cos y − i sin y) and g(z) = x3 + iy 3 for all z = x + iy ∈ C, then f : C → C
is holomorphic (hence entire) and g : C → C is not holomorphic at any point of C.

Proposition: Let Ω be a domain in C. If f : Ω → C is holomorphic and f 0 (z) = 0 for all z ∈ Ω,


then f is a constant function.

Remark: The connectedness of Ω in  the above proposition is, in general, necessary. For example,
1 if |z| < 1,
f : C \ ∂D → C, defined by f (z) =
2 if |z| > 1,
0
is not a constant function although f (z) = 0 for all z ∈ C \ ∂D.
In the above proposition, if Ω is not assumed to be connected, then we can conclude that f is
constant on each component of Ω.

Proposition: Let f : Ω → C be holomorphic, where Ω is a domain in C. If either of the functions


Re(f ) or Im(f ) or |f | on Ω is a constant, then f is a constant function.

Failure of mean value theorem: A direct generalization of the mean value theorem (similarly,
Rolle’s theorem) in R to C is not valid, in general.
For example, if f (z) = z 3 for all z ∈ C, then f : C → C is holomorphic, but there is no c ∈ [1, i]
(the line segment from 1 to i) such that f (i) − f (1) = (i − 1)f 0 (c).

Harmonic function: If Ω is an open set in R2 , then ϕ : Ω → R is called a harmonic function if


all the second order partial derivatives ϕxx , ϕyy , ϕxy , ϕyx of ϕ are continuous on Ω and
ϕxx (x, y) + ϕyy (x, y) = 0 ( Laplace’s equation) for all (x, y) ∈ Ω.
For example, if ϕ(x, y) = x2 − y 2 for all (x, y) ∈ R2 , then ϕ : R2 → R is a harmonic function.

Proposition: Let Ω be an open set in C. If f : Ω → C is holomorphic, then Re(f ) : Ω0 → R and


Im(f ) : Ω0 → R are harmonic functions, where Ω0 = {(x, y) ∈ R2 : x + iy ∈ Ω}.

Example: There is no holomorphic function f : C → C such that Re(f (x + iy)) = x2 + y 2 for all
x + iy ∈ C.

Harmonic conjugate: Let Ω be a domain in R2 and let Ω0 = (x + iy : (x, y) ∈ Ω}. If u : Ω → R


and v : Ω → R are harmonic functions such that u + iv : Ω0 → C is holomorphic, then v is said to
be a harmonic conjugate of u.
If v1 and v2 are harmonic conjugates of a harmonic function u : Ω → R, where Ω is a domain in
R2 , then v1 − v2 : Ω → R is a constant function.

Existence of harmonic conjugate:


(a) If Ω is a domain in R2 and u : Ω → R is harmonic, then u need not have a harmonic conjugate
v : Ω → R. For example, take u(x, y) = 21 log(x2 + y 2 ) for all (x, y) ∈ R2 \ {(0, 0)}.
x 2 2
However, if u(x, y) = x2 +y 2 for all (x, y) ∈ R \ {(0, 0)}, then u : R \ {(0, 0)} → R has a
y
harmonic conjugate v : R2 \ {(0, 0)} → R, where v(x, y) = − x2 +y 2
2 for all (x, y) ∈ R \ {(0, 0)}.

(b) If Ω is either R2 or a disk in R2 or a rectangle in R2 with sides parallel to the axes and if
u : Ω → R is harmonic, then u has a harmonic conjugate v : Ω → R.
If u(x, y) = 2x(1 − y) for all (x, y) ∈ R2 , then all the harmonic conjugates v : R2 → R of the
harmonic function u : R2 → R are given by v(x, y) = x2 − y 2 + 2y + c for all (x, y) ∈ R2 , where
c ∈ R. The corresponding holomorphic functions f : C → C are given by f (z) = iz 2 + 2z + ic
for all z ∈ C, where c ∈ R.
Similar method can be used to find the real part of a holomorphic function when its imaginary
part is given.

an (z − z0 )n , where an ∈ C for all
P
Power series: A power series in C is a series of the form
n=0
n ∈ N ∪ {0}, z0 ∈ C and z is a complex variable. Such a series is said to be centred at z0 .

an (z − z0 )n converges for z = z0 .
P
Every power series
n=0

an z n (i.e.
P
By replacing z − z0 by z, it is sufficient to consider only power series of the form
n=0
series centred at 0).

Example:

z n converges absolutely for all z ∈ D and diverges for all z ∈ C \ D.
P
(a) The power series
n=0

zn
P
(b) The power series n!
converges absolutely for all z ∈ C.
n=0

n!z n converges only for z = 0.
P
(c) The power series
n=0

zn
P
(d) The power series n2
converges absolutely for all z ∈ D and diverges for all z ∈ C \ D.
n=1

zn
P
(e) The power series n
converges absolutely for all z ∈ D, converges (but not absolutely) for
n=1
all z ∈ ∂D \ {1} and diverges for all z ∈ (C \ D) ∪ {1}.

an z n converges for z = z1 ∈ C\{0}, then it converges absolutely
P
Proposition: If a power series
n=0

an z n diverges for z = z2 ∈ C, then it diverges
P
for all z ∈ C with |z| < |z1 |. If the power series
n=0
for all z ∈ C with |z| > |z2 |.

an z n , there exists a unique R with
P
Radius of convergence: For every power series
n=0

n
P
0 ≤ R ≤ ∞ such that the series an z converges absolutely if |z| < R and diverges if |z| > R.
n=0

an z n .
P
We define R to be the radius of convergence of the power series
n=0
∞ ∞ ∞ ∞ ∞
zn zn zn
zn, n!z n ,
P P P P P
Example: The radii of convergence of the power series n!
, n2
and n
n=0 n=0 n=0 n=1 n=1
are 1, ∞, 0, 1 and 1 respectively.

an z n , then the series can converge
P
Remark: If R is the radius of convergence of a power series
n=0
for all or some or none of z satisfying |z| = R, as the examples given above show.

an z n is given
P
Cauchy-Hadamard formula: The radius of convergence R of a power series
n=0
1 1
by R
= lim sup |an | n .
n→∞

Examples:
z2 z4
(a) For a ∈ C \ {0}, the radius of convergence of the power series 1 + az + a2
+ a3 z 3 + a4
+ · · · is
1
|a| if |a| ≤ 1 and |a| if |a| > 1.
P∞
(−1)n 2n

(b) The radius of convergence of the power series 2n z is 2.
n=0

z n! is 1.
P
(c) The radius of convergence of the power series
n=0

an z n , where an 6= 0 for
P
Ratio formula: Let R be the radius of convergence of a power series
n=0
1 an+1
all n ∈ N ∪ {0}. Then R = lim an , provided that the limit exists (as a real number or ∞).
n→∞

n! n
P
Example: The radius of convergence of the power series nn
z is e.
n=1

an z n , where
P
Remark: If R is the radius of convergence of a power series an ∈ C \ {0} for all
n=0 
2n if n is even,

1 an+1
n ∈ N∪{0}, then R need not be equal to lim sup an . For example, if an =
n→∞ 2n−1 if n is odd,

an+1
an z n is 21 .
P
then lim sup an = 4 whereas the radius of convergence of the power series
n→∞ n=0

Pointwise convergence and uniform convergence: Let X be a metric space. Let (fn ) be a
n
P
sequence of complex-valued functions defined on X and let sn (x) = fj (x) for all n ∈ N and for
j=1
all x ∈ X.
(a) The sequence (fn ) is said to be pointwise convergent on X if for each x ∈ X, the sequence
(fn (x)) converges in C.
Thus (fn ) is pointwise convergent on X iff there exists a function f : X → C such that
fn (x) → f (x) for each x ∈ X, i.e. for each x ∈ X and for each ε > 0, there exists n0 ∈ N such
that |fn (x) − f (x)| < ε for all n ≥ n0 .
In this case, f is called the pointwise limit (function) of (fn ) on X and we write fn → f
pointwise on X.
(b) The sequence (fn ) is said to be uniformly convergent on X if there exists a function f : X → C
such that for every ε > 0, there exists n0 ∈ N satisfying |fn (x) − f (x)| < ε for all n ≥ n0 and
for all x ∈ X.
In this case, f is called the uniform limit (function) of (fn ) on X and we write fn → f
uniformly on X.

P
(c) The series fn is said to be pointwise (respectively, uniformly) convergent on X if the
n=1
sequence (sn ) is pointwise (respectively, uniformly) convergent on X and we define
P∞
fn = lim sn pointwise (respectively, uniformly) on X.
n=1 n→∞

Weierstrass M -test: Let X be a metric space. Let (fn ) be a sequence of complex-valued


functions defined on X and let (Mn ) be a sequence of non-negative real numbers such that

P
|fn (x)| ≤ Mn for all n ∈ N and for all x ∈ X. If the series Mn converges (in R), then the series
n=1

P
fn is uniformly convergent on X.
n=1

Uniform convergence and continuity: Let X be a metric space. For each n ∈ N, let
fn : X → C be continuous at x0 ∈ X.. If f : X → C such that fn → f uniformly on X, then f is
continuous at x0 .

Uniform convergence of power series: If R is the radius of convergence of a power series


∞ ∞
an z n and if 0 < r < R, then the power series an z n converges uniformly on {z ∈ C : |z| ≤ r}.
P P
n=0 n=0

n
P
However, the power series an z need not converge uniformly on {z ∈ C : |z| < R}. For
n=0

z n (with radius of convergence 1) does not converge uniformly on
P
example, the power series
n=0
{z ∈ C : |z| < 1}.

an (z − z0 )n . Then the
P
Proposition: Let R > 0 be the radius of convergence of a power series
n=0

n−1
P
radius of convergence of the power series nan (z − z0 ) is also R.
n=1

an (z − z0 )n for all z ∈ Ω, where Ω = C if R = ∞ and Ω = BR (z0 ) if R < ∞. Then
P
Let f (z) =
n=0

f : Ω → C is holomorphic and f 0 (z) = nan (z − z0 )n−1 for all z ∈ Ω.
P
n=1
f (n) (z0 )
Thus f is infinitely differentiable and an = n!
for all n ∈ N ∪ {0}.
∞ ∞
an (z −z0 )n and bn (z −z0 )n be power series with radii of
P P
Uniqueness of power series: Let
n=0 n=0
∞ ∞
an (z − z0 )n = bn (z − z0 )n
P P
convergence R1 and R2 respectively. If 0 < r ≤ min{R1 , R2 } and
n=0 n=0
for all z ∈ C with |z − z0 | < r, then an = bn for all n ∈ N ∪ {0}.

zn
P
Exponential function: The power series n!
converges for all z ∈ C. We define
n=0

zn
for all z ∈ C. We shall also write ez for exp(z), where z ∈ C.
P
exp(z) = n!
n=0
If f (z) = exp(z) for all z ∈ C, then f : C → C is holomorphic and f 0 (z) = exp(z) for all z ∈ C.
We note the following properties of the exponential function.
(a) ez = ez for all z ∈ C.
(b) ez ew = ez+w for all z, w ∈ C.
(c) ez 6= 0 and e−z = e1z for all z ∈ C.
∞ 2n+1
∞ 2n
z z
(−1)n (2n+1)! (−1)n (2n)!
P P
Sine and cosine functions: The power series and converge (ab-
n=0 n=0
∞ ∞
z 2n+1 z2n
(−1)n (2n+1)! (−1)n (2n)!
P P
solutely) for all z ∈ C. We define sin z = and cos z = for all z ∈ C.
n=0 n=0
If f (z) = sin z and g(z) = cos z for all z ∈ C, then f : C → C and g : C → C are holomorphic and
f 0 (z) = cos z and g 0 (z) = − sin z for all z ∈ C.
We note the following properties of sine, cosine and exponential functions.
(a) cos z = 21 (eiz + e−iz ) and sin z = 2i1 (eiz − e−iz ) for all z ∈ C.
(b) eiz = cos z+i sin z for all z ∈ C and in particular, we get the Euler’s formula: eiθ = cos θ+i sin θ
for all θ ∈ R and Euler’s equation: eiπ + 1 = 0.
(c) |eiθ | = 1 for all θ ∈ R and |ez | = eRe(z) for all z ∈ C.
(d) If z ∈ C \ {0}, then z = |z|eiθ , where θ ∈ arg(z).
(e) {ez : z ∈ C} = C \ {0}.
(f) sin2 z + cos2 z = 1 for all z ∈ C.
(g) The functions z 7→ sin z and z 7→ cos z from C to C are unbounded.
(h) {z ∈ C : sin z = 0} = {nπ : n ∈ Z} and {z ∈ C : cos z = 0} = (2n + 1) π2 : n ∈ Z .


Other trigonometric functions: The other trigonometric functions are defined as follows.
sin z
for all z ∈ C \ (2n + 1) π2 : n ∈ Z .

(a) tan z = cos z
(b) cot z = cos z
sin z
for all z ∈ C \ {nπ : n ∈ Z}.
(c) sec z = cos z for all z ∈ C \ (2n + 1) π2 : n ∈ Z .
1


(d) cosec z = sin1 z for all z ∈ C \ {nπ : n ∈ Z}.

Hyperbolic functions: We define cosh z = 12 (ez + e−z ) and sinh z = 12 (ez − e−z ) for all z ∈ C.
Note that cosh2 z − sinh2 z = 1, cos(iz) = cosh z and sin(iz) = i sinh z for all z ∈ C.
The other hyperbolic functions tanh, coth, sech and cosech are defined in terms of sinh and cosh
in the usual way.

Periodicity of ez : A function f : C → C is called periodic if there exists c ∈ C \ {0} such that


f (z + c) = f (z) for all z ∈ C. Also, any such c is called a period of f .
The function z 7→ ez from C to C is periodic and the set of all its periods is {2nπi : n ∈ Z \ {0}}.
In particular, the function z 7→ ez from C to C is not one-one.

Logarithm of a complex number: If z ∈ C \ {0}, then we define the logarithm of z as



log z = {w ∈ C : ew = z} = log |z| + i(Arg(z) + 2nπ) : n ∈ Z = log |z| + i arg(z), where log |z|
is the usual log of the positive real number |z|.
Also, the principal value of the logarithm of z ∈ C \ {0} is defined as Log z = log |z| + i Arg(z).
Thus log z = {Log z + 2nπi : n ∈ Z}.

Example: Log(1 + i) = 21 log 2 + πi


1 πi

4
and log(1 + i) = 2
log 2 + (8n + 1) 4
: n ∈ Z .

Remark:
(a) If z ∈ R and z > 0, then Log z = log z, where log on the right hand side is the usual log of
the positive real number z.
(b) If z, w ∈ C \ {0}, then log(zw) = log z + log w but Log(zw) need not be equal to Logz + Logw.
Similarly, if z ∈ C \ {0}, then Log(z 3 ) need not be equal to 3 Log z.

Continuity of Arg: The function Arg : C \ {0} → (−π, π] is discontinuous at each point of
(−∞, 0) and continuous at all other points of C \ {0}.

Logarithm as a function: The function z 7→ log z with domain C \ {0} is set-valued (or,
multi-valued). The function z 7→ Log z from C \ {0} to H0 = {x + iy : x ∈ R, y ∈ (−π, π]}
is not continuous at any point of (−∞, 0) although it is the inverse of the function z 7→ ez
from H0 to C \ {0}. In fact, for each n ∈ Z, the function z 7→ Log z + 2nπi from C \ {0} to
Hn = {x + iy : x ∈ R, y ∈ (2nπ − π, 2nπ + π]} is not continuous at any point of (−∞, 0) although
it is the inverse of the function z 7→ ez from Hn to C \ {0}.

Branch of logarithm: Let Ω be a domain in C. A branch of the logarithm on Ω is a continuous


function f : Ω → C such that f (z) ∈ log z for all z ∈ Ω, i.e. ef (z) = z for all z ∈ Ω.
If there exists a branch of the logarithm on a domain Ω in C, then 0 ∈ / Ω.
Let f : Ω → C be a branch of the logarithm on a domain Ω in C. Then g : Ω → C is a branch of
the logarithm on Ω iff there exists n ∈ Z such that g(z) = f (z) + 2nπi for all z ∈ Ω.

Proposition: If f : Ω → C is a branch of the logarithm on a domain Ω in C, then f is holomorphic


and f 0 (z) = z1 for all z ∈ Ω.

Example:
(a) If f (z) = Log z for all z ∈ C \ (−∞, 0], then f : C \ (−∞, 0] → C is a branch of the logarithm

(−1)n−1
(z − 1)n for all z ∈ B1 (1). We say that f is the principal
P
on C \ (−∞, 0] and f (z) = n
n=1
branch of the logarithm on C \ (−∞,  0].
Log z if Im(z) ≥ 0,
(b) For all z ∈ C \ [0, ∞), let f (z) =
Log z + 2πi if Im(z) < 0.
Then f : C \ [0, ∞) → C is a branch of the logarithm on C \ [0, ∞).
(c) Let P = {t + it(t − 1) : t ≥ 0},
Ω1 = {z ∈ C : Re(z) < 0, Im(z) ≥ 0} ∪ {z ∈ C : Re(z) ≥ 0, Im(z) > Re(z)(Re(z) − 1)} and
Ω2 = {z∈ C : Re(z) < 0, Im(z) < 0} ∪ {z ∈ C : Re(z) ≥ 0, Im(z) < Re(z)(Re(z) − 1)}. If
Log z if z ∈ Ω1 ,
f (z) =
Log z + 2πi if z ∈ Ω2 ,
then f : C \ P → C is a branch of the logarithm on C \ P .
(d) There does not exist any branch of the logarithm on C \ {0}.
(e) There does not exist any branch of the logarithm on {z ∈ C : 1 < |z| < 2}.

Complex power: If z, w ∈ C with z 6= 0, then we define z w = ew(Log z+2nπi) : n ∈ Z = ew log z .
Also, the principal value of z w is defined as ewLog z .

Remark:
(a) Let z ∈ C \ {0}. Then the present definition of z n for n ∈ N ∪ {0} is consistent with the usual
definition of z n . Similarly, the present definition of z −n for n ∈ N is consistent with the usual
1
definition of z −n . Further, if n ∈ N, then z n (according to the present definition) represent
the nth roots of z.
(b) If z ∈ C, then exp(z) (which we also write as ez ) is equal to ez (as per the present definition),
if we consider only the principal value.
1 1 1
(c) If z, w ∈ C \ {0}, then (zw) 2 need not be equal to z 2 w 2 , if we consider principal values only.
 π π
Example: ii = e− 2 −2nπ : n ∈ Z and the principal value of ii is e− 2 .

Branch of z w : Let w ∈ C. If f : Ω → C is a branch of the logarithm on a domain Ω in C, then


a branch of z w on Ω is given by the function g : Ω → C, where g(z) = ewf (z) for all z ∈ Ω.
Since f is holomorphic, g is also holomorphic and g 0 (z) = we(w−1)f (z) for all z ∈ Ω. (This can be
d
interpreted as dz (z w ) = wz w−1 for all z ∈ Ω, where the same branch of the logarithm on Ω has
been used for both z w and z w−1 .)

Rectifiable path: A path γ : [a, b] → C in C is called rectifiable (or, of bounded variation) if


n
P
there exists M > 0 such that γ(tj ) − γ(tj−1 ) ≤ M for every partition {t0 , t1 , . . . , tn } of [a, b].
j=1
It can be seen that a path γ : [a, b] → C in C is rectifiable iff both the functions Re(γ) : [a, b] → R
and Im(γ) : [a, b] → R are of bounded variation.

Example:
t + it2 cos 1t if 0 < t ≤ 1,

(a) Let γ(t) =
0 if t = 0.
Then γ : [0,1] → C is a rectifiable path in C.
π
t + it cos 2t if 0 < t ≤ 1,
(b) Let γ(t) =
0 if t = 0.
Then γ : [0, 1] → C is a path in C which is not rectifiable.

Differentiation of f : [a, b] → C: A function f : [a, b] → C is said to be differentiable at


t0 ∈ [a, b] if both the functions Re(f ) : [a, b] → R and Im(f ) : [a, b] → R are differentiable at t0 .
Also, if f : [a, b] → C is differentiable at t0 ∈ [a, b], then we define the derivative
f 0 (t0 ) = Re(f )0 (t0 ) + i Im(f )0 (t0 ).
The function f : [a, b] → C is said to be differentiable if f is differentiable at each point of [a, b].
If f, g : (a, b) → C are differentiable at t0 ∈ [a, b] and α ∈ C, then f + g and αf are differentiable
at t0 and (f + g)0 (t0 ) = f 0 (t0 ) + g 0 (t0 ), (αf )0 (t0 ) = αf 0 (t0 ).

Smooth path: A path γ : [a, b] → C in C is said to be


(a) continuously differentiable if γ is differentiable and γ 0 : [a, b] → C is continuous.
(b) smooth if γ is continuously differentiable and γ 0 (t) 6= 0 for all t ∈ (a, b).
(c) piecewise smooth or a contour if there exists a partition {t0 , t1 , . . . , tn } of [a, b] such that
γ|[tj−1 ,tj ] is smooth for each j ∈ {1, . . . , n}.
Similarly, we can define piecewise continuously differentiable path in C.

Integration of f : [a, b] → C: A function f : [a, b] → C is said to be integrable on [a, b] if both


the functions Re(f ) : [a, b] → R and Im(f ) : [a, b] → R are (Riemann) integrable on [a, b]. Also, if
f : [a, b] → C is integrable on [a, b], then we define the integral
Rb Rb Rb
f (t) dt = Re(f (t)) dt + i Im(f (t)) dt.
a a a
If f : [a, b] → C is bounded and f is continuous at all except possibly finitely many points of [a, b],
then f is integrable on [a, b].

Proposition: Let f : [a, b] → C and g : [a, b] → C be integrable on [a, b], α ∈ C and c ∈ (a, b).
Then
Rb Rb Rb
(a) f + g : [a, b] → C is integrable on [a, b] and (f (t) + g(t)) dt = f (t) dt + g(t) dt.
a a a
Rb Rb
(b) αf : [a, b] → C is integrable on [a, b] and αf (t) dt = α f (t) dt.
a a
Rb Rc Rb
(c) f (t) dt = f (t) dt + f (t) dt.
a a c b
R Rb
(d) |f | : [a, b] → C is integrable on [a, b] and f (t) dt ≤ |f (t)| dt.

a a
Rb
(e) f (t) dt = F (b) − F (a), provided that there exists a differentiable function F : [a, b] → C
a
such that F 0 (t) = f (t) for all t ∈ [a, b].
R1 R1 R1 2
Example: By definition, (1 + it)2 dt = (1 − t2 ) dt + i 2t dt = 3
+ i and by using (e) above,
0 0 0
R1 2 (1+it)3
(1 + it)2 dt = F (1) − F (0) = 3
+ i, where F (t) = 3i
for all t ∈ [0, 1].
0

Length of path: If γ : [a, b] → C is a rectifiable path in C, then the length of γ is defined as


n
nP o
`(γ) = sup γ(tj ) − γ(tj−1 ) : {t0 , t1 , . . . , tn } is a partition of [a, b] .
n=1
If γ : [a, b] → C is a piecewise continuously differentiable path in C, then γ is rectifiable and
Rb
`(γ) = |γ 0 (t)| dt.
a

Example:
(a) Let z0 ∈ C and r > 0. If γ(t) = z0 + reit for all t ∈ [0, 2π], then γ : [0, 2π] → C is a smooth
path in C and `(γ) = 2πr.
(b) Let z, w ∈ C. If γ(t) = (1 − t)z + tw for all t ∈ [0, 1], then γ : [0, 1] → C is a continuously
differentiable path in C and `(γ) = |z − w|.

Complex line integral: Let γ : [a, b] → C be a rectifiable path in C and let Ω ⊂ C such that
{γ} ⊂ Ω. A function f : Ω → C is said to be integrable along γ if there exists I ∈ C such that
for every ε > 0, there exists δ > 0 such that for every partition P = {t0 , t1 , . . . , tn } of [a, b] with
n
P
kP k < δ and for every ξj ∈ [tj−1 , tj ] (j = 1, . . . , n), f (γ(ξj ))[γ(tj ) − γ(tj−1 )] − I < ε.

j=1
Such an I, when it exists, is unique and is called the (complex line) integral of f along γ. We
R R R R
denote this by I = f or by I = f dγ or by I = f (z) dz. Thus f can be viewed as the
γ γ γ γ
Riemann-Stieltjes integral of f ◦ γ with respect to γ, where we consider the Riemann-Stieltjes
integral for complex-valued functions instead of real-valued functions.

Existence of integral: Let γ be a rectifiable path in C and let Ω ⊂ C such that {γ} ⊂ Ω. If
f : Ω → C is continuous on {γ}, then f is integrable along γ.

Example:
(a) Let α ∈ C and let f (z) = α for all z ∈ C. If γ is any rectifiable path in C joining z1 ∈ C to
R
z2 ∈ C, then f is integrable along γ and f dγ = α(z2 − z1 ).
γ
(b) Let f (z) = z for all z ∈ C. If γ is any rectifiable path in C joining z1 ∈ C to z2 ∈ C, then f
is integrable along γ and f dγ = 21 (z22 − z12 ).
R
γ

Existence and evaluation of integral: Let γ : [a, b] → C be a piecewise smooth path in C and
let Ω ⊂ C such that {γ} ⊂ Ω. If f : Ω → C is continuous on {γ}, then f is integrable along γ and
Rb
f (γ(t))γ 0 (t) dt.
R
f dγ =
γ a

Example:
(a) For z0 ∈ C, r > 0 and n ∈ Z, let f (z) = (z − z0 )n for all z∈ C \ {z0 } and γ(t) = z0 + reit for
R 0 if n 6= −1,
all t ∈ [0, 2π]. Then f is integrable along γ and f dγ =
γ
2πi if n = −1.
it
(b) Let γ1 (t) = e for all t ∈ [0, π] and γ2 (t) = 1 − 2t for all t ∈ [0, 1]. If f (z) = z for all z ∈ C,
R R
then f (z) dz = πi and f (z) dz = 0.
γ1 γ2

Change of parameter: Two paths γ : [a, b] → C and σ : [c, d] → C in C are said to be


equivalent if there exists a strictly increasing continuous function ϕ : [c, d] → [a, b] such that
ϕ(c) = a, ϕ(d) = b and γ ◦ ϕ = σ. In this case, the function ϕ is called a change of parameter.
Let γ and σ be equivalent rectifiable paths in C and let Ω ⊂ C such that {γ} ⊂ Ω. If f : Ω → C
R R
is integrable along γ, then f is integrable along σ and f = f .
σ γ

Example: If γ(t) = 1 + t(i − 1) and σ(t) = 1 + t2 (i − 1) for all t ∈ [0, 1], then γ and σ are
equivalent rectifiable paths in C and Re(z) dz = Re(z) dz = 21 (i − 1).
R R
γ σ

Opposite path: If γ : [a, b] → C is a path in C, then the opposite (or, reverse) path of γ is the
path −γ : [−b, −a] → C, defined by (−γ)(t) = γ(−t) for all t ∈ [−b, −a].
By a change of parameter, the opposite path of γ is equivalently given by the map −γ : [a, b] → C,
where (−γ)(t) = γ(a + b − t) for all t ∈ [a, b].
Note that −γ is rectifiable if γ is rectifiable and in such case `(−γ) = `(γ).

Proposition: Let γ : [a, b] → C be a rectifiable path in C and let Ω ⊂ C such that {γ} ⊂ Ω.
R R
(a) If f : Ω → C is integrable along γ, then f is integrable along −γ and f = − f .
−γ γ
(b) If f : Ω → C and g : Ω → C are integrable along γ, then f + g : Ω → C is integrable along γ
R R R
and (f + g) = f + g.
γ γ γ
(c) If f : Ω → C is integrable along γ and α ∈ C, then αf : Ω → C is integrable along γ and
R R
(αf ) = α f .
γ γ
(d) If σ : [a, b] → C is also a rectifiable path in C with {σ} ⊂ Ω and γ(b) = σ(a) and if f : Ω → C
R R R
is integrable along both γ and σ, then f is integrable along γ + σ and f = f + f.
γ+σ γ σ

Another integral: Let γ : [a, b] → C be a rectifiable path in C and let Ω ⊂ C such that {γ} ⊂ Ω.
R
For f : Ω → C, we say that f (z) |dz| exists (in C) if there exists I ∈ C such that for every ε > 0,
γ
there exists δ > 0 such that for every partition {t0 , t1 , . . . , tn } of [a, b] and for every ξj ∈ [tj−1 , tj ]
n
P
(j = 1, . . . , n), f (γ(ξj ))|γ(tj ) − γ(tj−1 )| − I < ε.

j=1 R
Such an I, when it exists, is unique and we write I = f (z) |dz|.
R γ
If f is continuous on {γ}, then f (z) |dz| exists (in C).
γ R
If f is continuous on {γ} and γ is piecewise smooth, then f (z) |dz| exists (in C) and
γ
Rb
f (γ(t))|γ 0 (t)| dt.
R
f (z) |dz| =
γ a
M L-inequality: Let γ : [a, b] → C be a rectifiable path in C and let Ω ⊂ C such that {γ} ⊂ Ω.
If f : Ω → C is integrable along γ and if there exists M > 0 such that |f (z)| ≤ M for all z ∈ {γ},
R R
then f (z) dz ≤ |f (z)| |dz| ≤ M `(γ).
γ γ
Z dz π
Example: ≤ , where γ(t) = 2eit for all t ∈ [0, π2 ].

γ z2 + 1 3
Proposition: Let f : Ω → C be continuous, where Ω is an open set in C, and let γ : [a, b] → Ω
R everyR ε > 0, there exists a polygonal path Γ : [a, b] → Ω with
be a rectifiable path. Then for
Γ(a) = γ(a), Γ(b) = γ(b) and f − f < ε.

γ Γ

Antiderivative: Let Ω be an open set in C. A differentiable function F : Ω → C is called an


antiderivative (or, a primitive) of a function f : Ω → C (on Ω) if F 0 (z) = f (z) for all z ∈ Ω.
If Ω is a domain in C and both F1 : Ω → C and F2 : Ω → C are primitives of f : Ω → C, then
there exists c ∈ C such that F2 (z) = F1 (z) + c for all z ∈ Ω.

Fundamental theorem of calculus for complex line integral: Let f : Ω → C be continuous,


where Ω is an open set in C, and let γ be a rectifiable path in Ω from z1 ∈ Ω to z2 ∈ Ω. If there
R
exists a primitive F : Ω → C of f , then f = F (z2 ) − F (z1 ).
γ

Example:
n+1
(a) For n ∈ N, let f (z) = z n and F (z) = zn+1 for all z ∈ C. Then F : C → C is a primitive
of f : C → C and hence if γ is any rectifiable path in C from z1 ∈ C to z2 ∈ C, then
1
z2n+1 − z1n+1 .
R n 
z dz = n+1
γ
(b) Let f (z) = sin z and F (z) = − cos z for all z ∈ C. Then F : C → C is a primitive of f : C → C
R
and hence if γ is any rectifiable path in C from z1 ∈ C to z2 ∈ C, then sin z dz = cos z1 −cos z2 .
γ

Closed path: A path γ : [a, b] → C in C is called closed if γ(a) = γ(b). A closed polygon with
three sides is called a triangle.

Proposition: Let f : Ω → C be continuous, where Ω is a domain in C. Then the following


statements are equivalent.
(a) There exists a primitive of f on Ω.
R
(b) f = 0 for every closed rectifiable path in Ω.
γ R R
(c) For every z1 , z2 ∈ Ω, if γ1 and γ2 are rectifiable paths in Ω joining z1 to z2 , then f = f
R γ1 γ2
(i.e f depends only on the endpoints of the rectifiable path γ in Ω).
γ

(The connectedness of Ω is required only in proving (a) from either (b) or (c).)

Example:
(a) Let f (z) = a0 + a1 z + · · · + an z n and F (z) = a0 z + a21 z 2 + · · · + n+1
an n+1
z for all z ∈ C, where
R
a0 , . . . , an ∈ C. Then F : C → C is a primitive of f : C → C and hence f = 0 for every
γ
closed rectifiable path γ in C.

an (z − z0 )n . Then R is also
P
(b) Let R > 0 be the radius of convergence of a power series
n=0
∞ ∞
an
− z0 )n+1 . If f (z) = an (z − z0 )n
P P
the radius of convergence of the power series n+1
(z
n=0 n=0

an
− z0 )n+1 for all z ∈ BR (z0 ), then F : BR (z0 ) → C is a primitive of
P
and F (z) = n+1
(z
n=0 R
f : BR (z0 ) → C and hence f = 0 for every closed rectifiable path γ in BR (z0 ).
γ
(c) If Ω is an open set in C, then a continuous function f : Ω → C need not have a primitive. For
example, consider Ω = C and f (z) = z for all z ∈ C.
(d) If Ω is a domain in C containing a circle centred at 0, then there cannot exist any branch of
the logarithm on Ω.

Cauchy-Goursat theorem: Let f : Ω → C be holomorphic, where Ω is an open set in C. If γ


R
is a triangle in Ω and conv(γ) ⊂ Ω, then f = 0.
γ

Star-shaped set: A subset Ω of C is called star-shaped if there exists z0 ∈ Ω such that [z0 , z] ⊂ Ω
for all z ∈ Ω.
Note that every convex set in C is star-shaped although the converse is not true. Also, every
star-shaped set in C is path connected although the converse is not true.

Cauchy’s theorem for star-shaped domain: Let f : Ω → C be holomorphic, where Ω is a


R
star-shaped open set in C. If γ is a closed rectifiable path in Ω, then f = 0.
γ

ez
Z
Example: 2
dz = 0, where γ(t) = eit for all t ∈ [0, 2π].
γ z +4
Extension of Cauchy-Goursat theorem: Let Ω be an open set in C and let z0 ∈ Ω. Let
f : Ω → C be continuous and f ∈ H(Ω \ {z0 }). If γ is a triangle in Ω and conv(γ) ⊂ Ω, then
R
f = 0.
γ

Extension of Cauchy’s theorem for star-shaped domain: Let Ω be a star-shaped open set
in C and let z0 ∈ Ω. Let f : Ω → C be continuous and f ∈ H(Ω \ {z0 }). If γ is a closed rectifiable
R
path in Ω, then f = 0.
γ

Uniform convergence and integration: Let γ : [a, b] → Ω be a rectifiable path, where Ω ⊂ C.


For each n ∈ N, let fn : Ω → C be continuous on {γ} and let f : Ω → C.
R R
(a) If fn → f uniformly on {γ}, then fn → f .
γ γ

P ∞
R P ∞ R
P
(b) If fn converges uniformly on {γ}, then fn = fn .
n=1 γ n=1 n=1 γ

Example:
(a) If f (z) = z1 for all z ∈ C \ {0}, then f : C \ {0} → C is continuous but there does not exist
any sequence (pn ) of polynomials (over C) such that pn → f uniformly on {z ∈ C : |z| = 1}.
it
Given z0 ∈ C
(b) Z  and r > 0, let γ(t) = z0 + re for all t ∈ [0, 2π]. Then for a ∈ C,
dz 0 if |a − z0 | > r,
=
z−a 2πi if |a − z0 | < r.
γ

Z or winding number: Let γ be a closed rectifiable path in C and z0 ∈ C \ {γ}. Then


Index
1 dz
is an integer, which is denoted by indγ (z0 ) and is called the index of γ with respect
2πi γ z − z0
to z0 (or, the winding number of γ around z0 ).

Proposition: Let γ be a closed rectifiable path in C. Then


(a) ind−γ (z) = −indγ (z) for all z ∈ C \ {γ}.
(b) indγ+σ (z) = indγ (z) + indσ (z) for all z ∈ C \ ({γ} ∪ {σ}), where σ is also a closed rectifiable
path in C with same initial point as that of γ.

Example: For z0 ∈ C, n ∈ Z \ {0} and r > 0, let γ(t) = z0 + reint for all t ∈ [0, 2π]. If z ∈ C,
n if |z − z0 | < r,
then indγ (z) =
0 if |z − z0 | > r.
Cauchy’s integral formula for star-shaped domain: Let f : Ω → C be holomorphic, where
Ω is Za star-shaped open set in C. If γ is a closed rectifiable path in Ω, then
1 f (z)
dz = indγ (z0 ) · f (z0 ) for all z0 ∈ Ω \ {γ}.
2πi γ z − z0
Cauchy’s integral formula for circle: Let f : Ω → C be holomorphic, where Ω is an open
it
set in C. LetZ a ∈ Ω and r > 0 such that Br [a] ⊂ Ω. If γ(t) = a + re for all t ∈ [0, 2π], then
1 f (z)
f (z0 ) = dz for all z0 ∈ Br (a).
2πi γ z − z0
Gauss’s mean value theorem: Let f : Ω → C be holomorphic, where Ω is an open set in C. If
1 R2π
z0 ∈ G and r > 0 such that Br [z0 ] ⊂ Ω, then f (z0 ) = f (z0 + reit ) dt.
2π 0

You might also like