H1BP English
H1BP English
1
Functions of one complex variable
3 Integration 47
1. Integration of a complex function . . . . . . . . . . . . . . . . . . . . . . . . 47
1.1. Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
1.2. Reminder: Line integral of a vector field . . . . . . . . . . . . . . . . 48
1.3. Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
1.4. Primitive. The Newton - Leibniz formula . . . . . . . . . . . . . . . . 52
1.5. The index of a closed curve (see references) . . . . . . . . . . . . . . . 54
2. Cauchy theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.1. Cauchy’s integral formula . . . . . . . . . . . . . . . . . . . . . . . . 54
2.2. Cauchy theorem for multiply domain . . . . . . . . . . . . . . . . . . 56
2.3. The existence of antiderivative . . . . . . . . . . . . . . . . . . . . . . 56
3. Cauchy’s integral formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.1. Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.2. The Cauchy Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.3. Cauchy integral formula for derivatives . . . . . . . . . . . . . . . . . 60
4. 4. Some corollaries of Cauchy integral formula . . . . . . . . . . . . . . . . . 61
4.1. Theorem Morera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.2. Cauchy inequality for derivatives . . . . . . . . . . . . . . . . . . . . 61
4.3. Liouville’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.4. Fundamental Theorem of Algebra . . . . . . . . . . . . . . . . . . . . 62
4.5. Mean Value Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
1. Complex numers
1.1. Definition
On R × R = R2 we define these operations:
4
Functions of one complex variable
−b
Å ã
−1 a
+ The inverse element of (a; b) is (a, b) = ; .
a2 + b 2 a2 + b 2
We call (R2 , +, •) is the complex field. Every elements z with z = (a; b) ∈ (R2 , +, •) is called
a complex number
We denote it as C.
We identify (a; 0) with a. One has:
(a; 0) + (b; 0) = (a + b; 0) = a + b
(a; 0).(b; 0) = (ab; 0) = ab
Example 1.1.
z = (a, b) = M
b
arg z
O a x
−−→
We call r = OM is a modulus of the complex number z.
−−→
arg z: is a trigonometric angle between Ox and OM .
arg z: is the value of arg z such that 0 ≤ z ≤ 2π.
We have: Arg z = {arg z + k2π, k ∈ Z} .
√
Let φ = arg z, it can be seen that r = a2 + b2 .
z = a + ib = r cos φ + i.r sin φ = r(cos φ + i sin φ).
The above representation is called the trigonometric form of the complex number, one has:
z = r(cos φ + i sin φ) := r.eiφ với eiφ = cos φ + i sin φ . Indeed, the Maclaurin expansion gives
us:
x 2 x4 x6
cos x = 1 − + − + · · · ⇒ thay φ.
2! 4! 6!
x3 x5 x7
sin x = x − + − + · · · ⇒ thay φ nhận thêm i
3! 5! 7!
2 3 4
x x x
ex = 1 + x + + + + · · · ⇒, substitute i.φ
2! 3! 4!
Then,
(iφ)2 (iφ)4 (iφ)6
cos φ = 1 + + − + . . . (Do i2 = −1, i4 = 1)
2! 4! 6!
(iφ)3 (iφ)5 (iφ)7
i sin φ = iφ + + + + . . . (Do i3 = −i, i5 = i)
3! 5! 7!
2 3
(iφ) (iφ) (iφ)4
eiφ = 1 + iφ + + + + . . . nên với eiφ = cos φ + i sin φ
2! 3! 4!
Indeed, we have:
2. z 7→ z̄ is called a standard on C.
1. z = z̄¯.
2. z1 + z2 = z1 + z2 .
3. z1 .z2 = z1 .z2 .
4. z + z = 2. Re z, z − z = 2. Im z, z.z = |z|2 = a2 + b2 .
z1 |z1 |
5. |z1 .z2 | = |z1 | |z2 | , = .
z2 |z2 |
6. |z1 + z2 | ≤ |z1 | + |z2 |.
z1 z1 z2 z1 z2
Note: = = .
z2 z2 z2 |z2 |2
Given M (xM ; yM ) = z1 , N (xN ; yN ) = z2 , then
»
δ (z1 , z2 ) = (xM − xN )2 + (yM − yN )2
r
√
n
r w1
O x
w2
So the conplex number z has two square roots, located symmetrically about the origin.
When representing the cube root of a complex number z: They lie on a circle with radius
√ φ
3
r with the first point making an angle with Ox-axis is , the following two points are
3
2π
separated by an angel of and form an equilateral triangle.
3
Then, we generalize to the n-th root case n
√
Example: Find the 4-th root of z = 1 + i. 3
π π
Rewrite z: z = 2 cos + i. sin .
3 3
Set w = ρ. (cos ψ + i. sin ψ) , we have:
√
4
ρ= 2
π
w4 = z ⇒ + k2π
ψ= 3
, k = 0, 3
4
So,the 4-th root of z is
√
4
π π
w1 = 2 cos + i. sin
12 12 ã
√
4
Å
7π 7π
w2 = 2 cos + i. sin
12 12
√
4
Å
13π 13π
ã
w3 = 2 cos + i. sin
12 12
√
4
Å
19π 19π
ã
w4 = 2 cos + i. sin .
12 12
1. wn ± zn → w ± z.
2. wn .zn → w.z.
zn z
3. → .
wn w
Proposition 2.2. Cauchy’s criterion. The sequence zn is convergent if and only if:
n+p
X
∀ε > 0, ∃n0 : n > n0 : p ∈ N ⇒ zk < ε
k=n+1
n+p
P
Remark zk = |sn+p − sn | .
k=n+1
Note:
∞
P
1. zn converges ⇒ zn → 0.
n=0
∞
P ∞
P ∞
P
2. zn → z, wn → w then (zn + wn ) → z + w.
n=0 n=0 n=0
∞
P ∞
P
Definition 2.2. The complex series zn is absolute convergence |zn | hội tụ.
n=0 n=0
∞
P ∞
P ∞
P
If zn converges but |zn | is not convegent zn is conditionally convergent.
n=0 n=0 n=0
∞
Å ã
1P 1
Example 2.2. 2
+ √ i is absolute convergence
n=0 n … n3
1 1 1 1 1 P∞ 1
Indeed, 2 + √ = + ∼ √ for n large enough. Besides, √ is conver-
n n3 i n4 n3 n3 Ç å n=1 n 3
P∞ 1 1 P∞ 1 1
gent, then 2
+ √ i is convergent. ⇒ + √ 3 i is absolute convergence.
n=1 n n3 n=1 n2 n
Next, here is an example of conditionally sequence:
1 1
Ö è
n
P (−1) . cos n i. sin
n
Example 2.3. + is a conditionally sequence. Rediscovering
n n
Leibnitz’s Theorem (of convergence of alternating series):
∞
X
(−1)n un , un > 0
n=0
X∞
If un is strictly decreasing to 0, then (−1)n un is convergent.
n=0
∞
Å ã
P 1 1 1 1
we have cos + i. sin is not convergent since cos + i. sin → 1 ̸= 0 .
n=1 n n n n
On the other hand, we can consider the following alternating series:
Ã
1 1 1 1
(−1)n cos sin cos2 + sin2
n + i. n = n n = 1
n n n2 n
1 1
∞ cos cos
P
(−1)n n . Consider u = n.
n
n=1 n n
1
cos
Set f (x) = x , x ≥ 1.
x
1 1 1 1 1
2
.x sin − cos sin − x. cos
f ′ (x) = x x x = x x ⇒ f ′ (x) < 0, with x ≥ x .
0
x2 x3
1
∞ sin
P n có
n=1 n
1 1
sin
n ≤ n = 1.
n n n2
∞
P
Proposition: Suppose that zn is absolute convergence and its sum is S.
n=0
σ:N→N
is a bijection (a permutation).
∞
P
Then zσ(n) is also absolute convergence and its sum is S. S.
σ(n)=1
∞
P
Proof: For all ε > 0, since zn is absolute convergence, then ∃n0 , ∀n > n0 :
n=0
n
P ε n
P ε
|zk | − S < and |zk | <
k=0 2 k=n0 +1 2
Let’s consider N0 = max {σ(0), σ(1), . . . , σ(n0 )} thì N0 ≥ n0 .
If σ is taken from 1 → n0 then N0 = n0 .
we have:
N0
X n0
X N0
X
zσ(k) − S = zσ(k) − S + zσ(k)
σ(k)=0 σ(k)=0 σ(k)=n0
k>n0
n0
X N0
X
≤ zσ(k) − S + zσ(k)
σ(k)=0 σ(k)=n0
k>n0
n0 ∞
X X ε ε
≤ zσ(k) − S + |zk | < + = ε.
k=n+1
2 2
σ(k)=0
∞
P ∞
P ∞
P
Definition 2.3. Cho zn , tn be 2 complex sequences. We call Wn with
n=0 n=0 n=0
n
X
wn = zi .tn−i
i=0
P P
is the product chain of zn , tn .
∞
P ∞
P
Proposition 2.3. Giả sử zn , tn are 2 absolutely convergent series corresponding to
n=0 n=0
S and T
∞
P P P
Then, the product chain Wn of zn , tn converge absolutely to the same sum S.T .
n=0
If one series converges absolutely and the other converges (not necessarily absolutely) to the
same sum.0
∞
P ∞
P
Proposition: Proof: we have zn absolutely converges ⇒ |zn | is convergent .
n=0 n=0
Consider
Now, we proof:
z0 + z1 + z2 + · · · + zn → z
z0 (Tn − T ) + z1 (Tn−1 − T ) + · · · + +zn (T0 − T ) → 0.
< ε. (|T | + S + M + |T |) → 0.
∞ 1
z n with |z| ≤ q < 1 converges absolutely to
P
Example 2.4. .
n=0 1−z
1 − z n+1 1
1 + z + z2 + · · · + zn = converges to when n → ∞.
1−z 1−z
∞
|z|n ≤ q n , q n converges with q < 1.
P
n=0
The extended complex plane C = C ∪ {±∞} is the set of extended complex numbers
1
zn → ∞ ⇔ |zn | → +∞ ⇔ → θC
zn
z
z + ∞ = ∞, =0
z ∞
z ̸= 0, = ∞, z.∞ = ∞.
0 ÅÅ ã ã
1 1
In the space Oxyz consider the sphere B 0; 0; , .
Å 2 ã22
1 1
Then the equation of this sphere is x2 + y 2 + z − = .
2 4
z
P (0; 0; 1)
|z|2
Å ã
x0 y
π(M ) π(M ) = ; ;
1 + |z|2 1 + |z|2 1 + |z|2
O x
M (x; y; z)
y
ConsiderÅÅthe following
ã ã map:
1 1
π: B 0; 0; ; → Oxy ∪ {∞} = C
2 2
M 7→ π(M ) = P M ∩ Oxy
P 7→ ∞.
Clearly, π is a bijection.
x = 0 + x0 t
Suppose M (x0 ; y0 ; 0) ̸= ∞ we have (P M ) : y = y0 t , t ∈ R.
z =1−t
ÅÅ ã ã
1 1
We haveπ(M ) ∈ B 0; 0; , nên
2 2
1 2 1
Å ã
2 2
(x0 t) + (y0 t) + 1 − t − =
2 4
Å ã2
1 1
⇔ t2 x20 + y02 + t −
=
2 4
2 2 2 2
⇔ t x0 + y 0 + t − t = 0
⇔ t x20 + y02 + 1 − 1 = 0 (Dot = 0 ⇒ M = ∞)
1 1
⇔t= 2 2
=
x0 + y 0 + 1 1 + |z|2
Ç å
x0 y0 |z|2
⇒ π(M ) = ; ; .
1 + |z|2 1 + |z|2 1 + |z|2
x3
P (0, 0, 1)
π(M )
x2
O
M (x, y, 0)
x1
C = Ox1 x2 → S
z = M (x, y) 7→ π(M ) = P M ∩ S
• Find π(M ):
−−→ x1 = xt,
P M = (x, y, −1) ⇒ P M : x2 = yt, (t ∈ R)
x3 = 1 − t,
1 2 1
Å ã
2 2
(xt) + (yt) + 1 − t − =
2 4
⇔ t2 (x2 + y 2 + 1) − t = 0
t = 0 (absurd)
⇔ 1 1
t= 2 2
=
x +y +1 1 + |z|2
|z|2
Å ã
x y
Thus, π(M ) = , ,
1 + |z|2 1 + |z|2 1 + |z|2
• Conversely, given M ′ (x1 , x2 , x3 ) ∈ S, find M ∈ Ox1 x2 = C such that π(M ) = M ′ :
Assume z = M (x, y) = M (x, y, 0). We have:
|z|2 2 x3
x3 = ⇒ |z| =
1 + |z|2 1 − x3 Å ã
x 2 x3 x1
Thus, x1 = 2
⇒ x = x1 (1 + |z| ) = x1 1 + =
1 + |z| Å 1 − x3
ã 1 − x3
y x 3 x 2
x2 = ⇒ y = x2 (1 + |z|2 ) = x2 1 + =
1 + |z|2 Å ã 1 − x3 1 − x3
x 1 x 2
Therefore, π −1 (M ′ ) = ,
1 − x3 1 − x3
The projection π isÅa bijection. If zã= ∞, then π(z) = P .
x1 x2
Suy ra W = N , là tạo ảnh của N ′ .
1 − x3 1 − x3
It is obvious that M → ∞ then π(M ) → P.
Let’s consider z = M (x, y), w = N (x′ , y ′ ), we have: π(z) = (x1 , x2 , x3 ), π(∞) = (x′1 , x′2 , x′3 ).
p
We define λ(z, w) = (x1 − x′1 )2 + (x2 − x′2 )2 + (x3 − x′3 )2 spherical distance between z
and w.
With z, w ̸= ∞:
|z − w|
⇒ d(z, w) = 1 1
(1 + |z|2 ) 2 (1 + |w|2 ) 2
1
If w = ∞, then d(z, ∞) = 1
(1 + |z|2 ) 2
Proposition 4.1.
Proof: iii)
Property 4.1.
i) ∅, C is open (closed).
• clA =
S
α Fα , Fα is closed, A ⊂ Fα .‘
Proof:
Proposition 4.3. | · | và d(·, ·) are equivalent in terms of sequences in C. That is, the
sequence {zn } converges to z ∈ C according to | · | if {zn } is convergent to z according to
d(·, ·).
Proof:
(⇐) Suppose that {zn } converges z according to d(·, ·), that is, d (zn , z) → 0.
First, we prove the existence, M > 0 such that |zn | ≥ M, ∀n.
Suppose that, {zn } is not bounded, in this case, there exists a subsequence {znk } such
that {znk } → ∞.
In this case,
d(z, ∞) ≥ d (znk , z) + d (znk ) , ∞)
1 d (zn , z)
= d (znk , z) + Ä 1/2
·
|znk |
ä
1 + |znk |2
Whereas, d (zn , z) → 0, |znk | → ∞.
We deduce that, z = ∞, which contradicts z ∈ C. Do đó |zn | ≥ M for all n.
In this case,
Ä ä1/2 1/2
|zn − z| = 1 + |zn |2 · 1 + |z|2 · d (zn , z)
1/2 1/2
≥ 1 + M2 · 1 + |z|2 · d (zn , z) → 0
γ : [a, b] → C
Giving the curve γ is equivalent to giving two real functions x and y on the interval [a, b].
We always assume the curve is continuous, which means the functions x(t) and y(t) are
continuous.
We often identify γ (as a mapping) with the set L = γ ([a, b]), which is the image of the
mapping γ. This is done to simplify notation when there is no risk of confusion. Note that
each set L is the image of numerous different curves.
Two curves γ1 : [a, b] → C and γ2 : [c, d] → C are called equivalent if there exists a
bijective, continuous, strictly increasing function g : [a, b] → [c, d] such that γ1 (t) = γ2 (g(t))
for every t ∈ [a, b].
We do not distinguish between equivalent curves. It is easy to see that every curve
γ : [a, b] → C is equivalent to some curve σ : [0, 1] → C.
Example 1
1. Given γ1 (t) = t, γ2 (t) = sin πt, t ∈ [0, 1]. Both curves γ1 and γ2 have the image as the
interval [0, 1] ⊂ C but are not equivalent. Indeed, as t varies from 0 to 1, γ1 (t) varies
from 0 to 1 whereas γ2 (t) varies from 0 to 1 and then back to 0.
2. For every n ∈ N consider the curve εn (t) = eint , t ∈ [0, 2π]. These curves all have
the image as the unit circle {z ∈ C : |z| = 1}. However, as t varies from 0 to 2π, εn (t)
travels around the unit circle n times. Therefore, εm is not equivalent to εn if m ̸= n.
• A curve γ is called to be simple if γ (t1 ) ̸= γ (t2 ) for t1 ̸= t2 and {t1 , t2 } ̸= {a, b}. A
simple, closed curve is called a Jordan curve or contour.
exists and it is continuous and nonzero for all t ∈ [a, b] (considering one-sided deriva-
tives at a and b).
• A curve γ is called to be piecewise smooth if there are points a = t0 < t1 < . . . < tm = b
such that for every i = 1, . . . , m, the curve γi (t) = γ(t), t ∈ [ti−1 , ti ] is smooth.
F1 ∩ F2 = ∅, F1 ∪ F2 = A (1.1)
Theorem 5.1. For a set A ⊂ C to be connected, it is necessary and sufficient that for every
set B ̸= ∅ which is both open and closed in A, we have B = A.
Proof:
γ : [a, b] → A
such that γ(a) = z and γ(b) = ω, then we say that z and ω can be connected by a curve
within A.
A set A ⊂ C is called path-connected if any two points in A can be connected by a path
lying within A.
Proof:
Assume A is path-connected but not connected. Then there exist two closed subsets
F1 , F2 satisfying (1.1). Choose z ∈ F1 and ω ∈ F2 and let γ : [a, b] → A be a path connecting
a+b
z and ω. Bisect the interval [a, b] at the midpoint m0 = . If γ (m0 ) ∈ F1 , then choose
2
[a1 , b1 ] = [m0 , b]. If γ (m0 ) ∈ F2 , then choose [a1 , b1 ] = [a, m0 ].
a1 + b 1
Continuing this process of bisecting the interval [a1 , b1 ] at the midpoint m1 = and
2
selecting subintervals [a2 , b2 ] such that γ (a2 ) ∈ F1 and γ (b2 ) ∈ F2 , and so on. This process
results in a sequence of narrowing intervals {[an , bn ]} where γ (an ) ∈ F1 and γ (bn ) ∈ F2 .
Let c be the unique common limit point of all intervals [an , bn ]. Since an → c, γ (an ) →
γ(c). Since the sequence {γ (an )} ⊂ F1 and F1 is closed, it follows that γ(c) ∈ F1 . Similarly, as
bn → c, γ(c) ∈ F2 . Hence, γ(c) ∈ F1 ∩ F2 , which contradicts the assumption that F1 ∩ F2 = ∅
0
Theorem 5.3. For A being an open subset in C, the following statements are equivalent:
i) A is connected;
ii) A is path-connected;
iii) Any two points in A can be connected by a piecewise smooth curve within A;
Proof:
5.3. Domain
A set D ⊂ C is called a domain if D is open and connected.
If D is a domain then D = D ∪ ∂D is called the closure of the domain. Note that the
closure of a domain is not a domain, but the interior of the closure is a domain.
If the domain D is bounded, then D is called a bounded domain. If D is a bounded
domain, then D is a compact set; therefore, a bounded domain is also referred to as a
compact domain. To denote D as a compact domain, we write D ⋐ C. If A is a bounded set
in D and A ⊂ D, then we also write A ⋐ D.
Example 4:
a) Open disks or more generally any open convex sets are domains.
b) A closed disk minus a point is a closed domain. Two curves γ0 , γ1 : [a, b] → D are
called to be homotopic in D, denoted γ0 ∼ γ1 , if there exists a continuous mapping
φ : [a, b] × [0, 1] → D
such that φ(t, 0) = γ0 (t), φ(t, 1) = γ1 (t) and φ (a, s) = φ(b, s) for every t ∈ [a, b], s ∈ [0, 1].
The homotopy relation among curves defined over the interval [a, b] is an equivalence
relation.
Remark 2 If we define γs (t) = φ(t, s), t ∈ [a, b] then for each s ∈ [0, 1] there exists a
closed curve γs : [a, b] → D. We see that γ0 ∼ γ1 , which means that as s varies from 0 to 1,
γs continuously transitions from γ0 to γ1 .
If γ1 (t) =∗ ∈ D, (where γ1 (t) is a constant point) then instead of γ0 ∼ γ1 we write γ0 ∼ ∗
a) A set D is called star-shaped if there exists z0 ∈ D such that the segment [z0 , z] ⊂ D
for every z ∈ D. Clearly, all convex sets are star-shaped, and all star-shaped sets are
connected. Thus, any open, star-shaped set is a domain. We will demonstrate that a
star-shaped domain is simply connected.
Then φ is continuous, and φ(t, 0) = γ(t), φ(t, 1) = z0 for all t ∈ [a, b]. Therefore, γ ∼∗ .
b) The domain D = C\ {0} is multiply connected since the unit circle γ(t) = eit , t ∈ [0, π]
cannot be continuously contracted to a point.
It has been shown that a domain D ⊂ C is simply connected if and only if the boundary
of D in C is connected (note that the boundary is in C, not in C). According to Remark
1, D is simply connected if ∂C D has only one connected component.
Example 6
l2
l1
H3 H4
A multiply connected domain can be transformed into a simply connected domain by adding
certain line segments to its boundary. For example, the domain D in figure 3 can be made
simply connected by adding two segments, l1 and l2 , to its boundary.
The positive direction of the boundary of a domain is defined such that when traversing
along the boundary, the domain being considered lies to the left.
In figure 3, the direction indicated by the arrows is the positive direction of the boundary.
The positively oriented boundary of the domain D is denoted as ∂D+ , while the boundary
oriented in the opposite direction is denoted as ∂D− .
w = z2
y y
w = f (z)
O x O x
z
Example 1.2. Consider the function f (z) = w =
z+1
28
Functions of one complex variable
Proposition 1.2. Suppose f (z) = u(x, y) + iv(x, y), where z = (x, y) ∈ D, z0 = x0 + iy0 ,
and L = L1 + iL2 . Then,
lim u(x, y) = L1
(x,y)→(x0 ,y0 )
lim f (z) = L ⇔
z→z0 lim v(x, y) = L2
(x,y)→(x0 ,y0 )
Proof:
Property 1.1. a) The properties of complex variable functions correspond to the proper-
ties of continuous two-variable functions.
Proof:
Suppose f is not uniformly continuous on K. Then,
1 1
∃ε > 0, ∀δn = , ∃zn , wn ∈ D such that |zn − wn | < and |f (zn ) − f (wn )| ≥ ε0 .
n n
As K is compact, there exist sub-sequences (assume these are (znk ) and (wnk )) such that
Thus,
Resulting in w∗ ≡ z ∗ .
Due to the continuity of f , we have:
Then
ε0 ≤ |f (znk ) − f (z ∗ )|
≤ |f (znk ) − f (z ∗ )| + |f (z ∗ ) − f (wnk ) | (Contradiction)
Proposition 1.4. If f is continuous on a compact set K, then |f | attains its maximum and
minimum values on K, meaning there exist z1 , z2 such that
So, lim fn = f ⇔ (∀z ∈ D, ∀ε > 0, ∃n0 : n > n0 ⇒ |fn (z) − f (z)| < ε).
n→∞
Indeed, for all z such that |z| < 1, for all ε > 0, we choose n0 > log|z| ε, with n > n0 , we
have:
∀ε > 0, ∃n0 = logq ε, ∀n > n0 ⇒ |fn (z) − θ(z)| = |z|n ≤ q n < q0n < ε.
Proof:
Take an arbitraryz0 ∈ D.
ε
|fn (z) − f (z)| < , ∀z ∈ D.
3
ε
In specific, |fn (z0 ) − f (z0 )| < .
3
Supervisor: Bùi Thế Quân - Email: quanbt@hcmue.edu.vn Trang 31
Functions of one complex variable
|f (z) − f (z0 )| ≤ |f (z) − fn (z)| + |fn (z) − fn (z0 )| + |fn (z0 ) − f (z0 )|
3ε
≤
3
=ε
Proposition 2.2. The sequence of function {fn } is uniformly convergent on D if and only
if
Proof:
(⇒): It is obvious.
(⇐): Let’s consider f (n) satisfies (*). So, for all z ∈ D then {fn (z)} is a Cauchy sequence
in C, so it is convergence.
Set f (z) = lim fn (z), ∀z ∈ D.
n→∞
From (*) fixing n, given m ⇒ ∞, we get
∞
X 1 − z n+1 1
Example 2.3. z k = sn = → .
k=0
1−z 1−z
∞
X
Proposition 2.3. The series of functions fn is uniformly convergent on D if and only
n=0
if
Proof:
we have
n+p n+p
X X
| fk (z)| ≤ |fk (z)|
k=n+1 k=n+1
n+p
X
≤ cn
k=n+1
⇒0
∞ √
X nz + 1
Example 2.4. , |z| ≤ R,
n=1
n2
we have:
√ √
nz + 1 nz 1
≤ +
n2 n2 n2
R 1
≤ 3 + 2
n2 n
R+1
≤ 3 .
n2
∞ ∞ √
X R+1 X nz + 1
moreover 3 is convergent so is uniformly and absolute convergent |z| ≤
n=1 n2 n=1
n2
R.
∞
X (2z)n
Example 2.5.
n=1
3n + 1
we have
n (2z)n 2 2 3
lim n
= z < · = 1.
n→∞ 3 +1 3 3 2
3
• Nếu z < thì chuỗi đã cho hội tụ đều và tuyệt đối.
2
3
• Nếu z > thì chuỗi đã cho phân kỳ.
2
Proof:
∞
X
cn z0n is convergent ⇒ cn z0n = 0.
n=0
a) If the sequence {cn z0n } is convergent then, it is bounded, which means there exists
M > 0 such that |cn z0n | ≤ M, ∀n ∈ N. Then, with |z| ≤ q < |z0 |, we have
Å ãn
n n z
|cn z | = cn z0 ·
z0
z n
= |cn z0n | ·
z0
q n
Å ã
≤M· .
|z0 |
∞ Å
q n
ã
q X
Vì < 1 nên hội tụ.
|z0 | n=0
|z0
∞
X
Vậy cn z n in uniformly converges and absolute in |z| ≤ q < |z0 |.
n=0
b) Similarly prove.
∞
X
Proposition 2.7. There exists a real number R such that cn z n is uniformly continuous
n=0
and absolute in |z| ≤ q < R and it is divergent |z| > R with R is convergent radius.
cn+1 »
Proposition 2.8. (Cauchy - H’Adamard) If lim = l or lim n |cn | = l then
n→∞ cn n→∞
0, l = ∞
R = ∞, l = 0 .
1
, l ∈ (0, ∞)
l
The series identifying those transcendental functions have the same radius of convergence,
which is +∞, so those functions are identified and continuous on the plane. According to
Maclaurin expansions of real functions, narrowing those functions on R to get the corre-
sponding real functions that we have known.
From these equations (1) and (2) we have (i) and (ii).
By the definition of the transcendental functions and the commutative property of the
absolute convergent series we get:
ez = ch z + sh z,
and then
e−z = ch z − sh z.
Remark 3.2. From the formula (3) with x = 0, y = ϕ we have the Euler formula
Theorem 3.5. The function ez is periodic with period being multiples of 2πi.
On the other hands, if ω = a+ib moreover ez+ω = ez then eω = 1. So, ea (cos b+i sin b) = 1
or ea = 1, cos b = 1, sin b = 0. From that a = 0, b = 2kπ, which means ω = 2kπi, k ∈
Z.Therefore, the function ez is periodic with period being multiples of 2πi. Using this and
the Theorem 3, we have the circulation and the period of cos z và sin z.
4. Elementary transformations
4.1. Linear function
Definition 4.1. (Linear function) A linear function is a function having the form w =
w(z) = az + b, where a, b ∈ C and a ̸= 0.
dw
Remark 4.1. (i) Since = a ̸= 0, w is (analytic and) conformal on C.
dz
w−b
(ii) We have w = az + b ⇐⇒ z = , hence w is a bijection from C to C. If we consider
a
w(∞) = ∞, then w is a bijection from C to C.
(iii) If |a| = 1, b = 0 then a = eiα then w = az = eiα |z|ei arg z = |z|ei(arg z+α) is the rotation
R0,α .
̸ 0, b = 0, then w = az = |a||z|ei(arg z+α) is the composition of the rotation R0,α
If |a| =
and the dilation D0,|a| .
In general, w = az + b = |a||z|ei(arg z+α) + b is the composition of the rotation R0,α , the
dilation D0,|a| , and the translation Tb .
Proposition 4.1. A linear transformation turns circles into circles, turns lines into lines.
Remark 4.2. A linear function is uniquely determined by either two distinct points and
dw
their images, or a point along with its image and = a.
dz
w − w1 z − z1
Indeed, if w1 = w(z1 ) and w2 = w(z2 ), then = is the desired linear function.
w2 − w1 z2 − z1
dw
If w(z0 ) = w0 and = a, then w − w0 = a(z − z0 ) is the desired linear function.
dz
Example 4.1. Find all linear transformations that turn B(z0 , r1 ) into B(w0 , r2 ).
Solution.
After the translation T−z0 : B(z0 , r1 ) 7→ B(0, r1 ).
After the dilation D0, rr2 : B(0, r1 ) 7→ B(0, r2 ).
1
Proof.
Let us consider a circle (S) having equation
A(x2 + y 2 ) + Bx + Cy + D = 0.
−d −d
Remark 4.4. (i) If belongs to the circle (S), the image of (S) is a line. If doesn’t
c c
belong to the circle (S), the image of (S) is a circle.
(ii) If z1 , z2 are on opposite sides of (S), then w(z1 ), w(z2 ) are on opposite sides of w(S).
Consequently, if z1 , z2 are on the same side of (S), then w(z1 ), w(z2 ) are on the same
side of w(S).
Indeed, suppose by contradiction that z1 , z2 are on opposite sides of (S) and w(z1 ), w(z2 )
are on the same side of w(S). Observe that there exists a circle ∆ passing through
w(z1 ), w(z2 ) without intersecting w(S). It follows that w−1 (∆) is a circle passing through
z1 , z2 without intersecting (S). This is impossible as z1 , z2 are on opposite sides of (S).
The proof is complete.
2z + 1
Example 4.2. Let w = . In C, find:
z−1
a) The image of x2 + y 2 ≤ 1. c) The preimage of u2 + v 2 < 1.
b) The image of Imz > 2. d) The preimage of Rew = 2.
Solution.
2z + 1 w+1
a) For z ̸= 1, we have w ̸= ∞ and: w = ⇔z= . Setting w = u + vi gives:
z−1 w−2
ß ß
|z| ≤ 1 u + vi + 1 |u + vi + 1| ≤ |u + vi − 2|
⇔ ≤1⇔
z ̸= 1 u + vi − 2 (u, v) ̸= (2, 0)
1
⇔ |u + 1 + iv| ≤ |u − 2 + iv| ⇔ (u + 1)2 + v 2 ≤ (u − 2)2 + v 2 ⇔ u ≤ .
2
ß ™
1
Therefore, the image of x2 + y 2 ≤ 1 is w Rew ≤ ∪ {∞}.
2
b) For z ̸= 1, we have:
w+1 u + vi + 1 (u + 1 + vi)(u − 2 − vi) u2 − u − 2 + v 2 −3v
z= = = 2 = 2 + i.
w−2 u + vi − 2 (u − 2) + v 2 (u − 2) + v 2 (u − 2)2 + v 2
Note that 1 ∈
/ {z|Imz > 2}, so:
u2 − u − 2 + v 2
ß 2
u − u − 2 + v 2 > 2(u − 2)2 + 2v 2
Imz > 2 ⇔ > 2 ⇔
(u − 2)2 + v 2 (u, v) ̸= (2, 0)
7 2
Å ã
9
⇔ u− + v2 < .
2 4
ÅÅ ã ã
7 3
Therefore, the image of Imz > 2 is B ,0 , .
2 2
c) B((−1, 0), 1).
d) {∞} ∪ {z|z ̸= 1, Rez = 1}. □
Definition 4.3. (Cross ratio) Given 4 distinct points z1 , z2 , z3 , z4 in C. The cross ratio
of these points is denoted by (z1 , z2 , z3 , z4 ).
If the points are not ∞, then
z1 − z3 z1 − z4
(z1 , z2 , z3 , z4 ) := : . (4.1)
z2 − z3 z2 − z4
If zi = ∞ for some unique i, say i = 1, we define (∞, z2 , z3 , z4 ) = lim (z1 , z2 , z3 , z4 ).
z1 →∞
Particularly, this limit is obtained by replacing the two sums containing z1 in the equation
(4.1), which are z1 − z3 and z1 − z4 , with 1. Similar definitions apply to i = 2, 3, 4.
1−i 1 − (1 + i)
Example 4.3. a) (1, −1 + 2i, i, 1 + i) = : = 1 + 2i.
(−1 + 2i) − i (−1 + 2i) − (1 + i)
1 − i 1 − (1 + i)
b) (1, ∞, i, 1 + i) = : = 1 + i.
1 1
Proposition 4.3. A linear fractional function preserves the cross ratio. In other words, if
wi = w(zi ), i = 1, 4, then (z1 , z2 , z3 , z4 ) = (w1 , w2 , w3 , w4 ).
Proof.
We have:
azi + b azj + b ad(zi − zj ) − bc(zi − zj ) (zi − zj )(ad − bc)
wi − wj = − = = .
czi + d czj + d (czi + d)(czj + d) (czi + d)(czj + d)
Thus:
w1 − w3 w1 − w4
:
w2 − w3 w2 − w4
(z1 − z3 )(ad − bc) (cz2 + d)(cz3 + d) (z1 − z4 )(ad − bc) (cz2 + d)(cz4 + d)
= · : ·
(cz1 + d)(cz3 + d) (z2 − z3 )(ad − bc) (cz1 + d)(cz4 + d) (z2 − z4 )(ad − bc)
z1 − z3 z1 − z4
= : .
z2 − z3 z2 − z4
□
Example 4.4. Find the linear fractional function that maps 1, 0, i to ∞, 1, 1 + i respectively.
Solution.
Using (4.2) gives:
Example 4.5. Find the linear fractional function mapping 1, i, 1+2i to i, 1, 3−i respectively.
Solution.
We have:
w−i z−i −1 3i
Å ã
= · −
w − (3 − i) z − (1 + 2i) 5 5
5w − 5i z−i
= · (−1 − 3i)
w−3+i z − 1 − 2i
15 − 10i z−i
5+ = · (−1 − 3i)
w−3+i z − 1 − 2i
(15 − 10i)(z − 1 − 2i)
w−3+i =
(z − 1)(−1 − 3i) − 5(z − 1 − 2i)
z(15 − 10i) + (15 − 10i)(−1 − 2i)
w = +3−i
z(−1 − 3i) + 1 + 3i − 5z − 5(−1 − 2i)
z(−6 − 13i) − 4 + 13i
Hence, w = . □
z(−6 − 3i) + 6 + 13i
Definition 4.4. (Points symmetric with respect to a circle) Given a circle S(z0 , r).
We say the points z, z ∗ are symmetric with respect to S(z0 , r) if arg(z − z0 ) = arg(z ∗ − z0 )
and |z − z0 ||z ∗ − z0 | = r2 .
Example 4.6. Find the point z ∗ symmetric to z = 1 + i with respect to B((0, 0), 2).
Solution.
Let z0 = 0. Since z ̸= z0 , arg(z − z0 ) = arg(z ∗ − z0 ) means there exists k > 0 such that
z ∗ − z0 = k(z − z0 ), or z ∗ = k + ki. Besides:
√ √
|z − z0 ||z ∗ − z0 | = 22 ⇔ 2( 2k) = 4 ⇔ k = 2.
Thus z ∗ = 2 + 2i. □
Proposition 4.5. z, z ∗ are symmetric with respect to S(z0 , r) iff any circle (S ∗ ) passing
through z, z ∗ is orthogonal to (S).
Proof.
(⇒) Assume that z, z ∗ are symmetric with respect to S(z0 , r) and S ∗ (w0 , R) is a circle passing
through z, z ∗ . Let P be an intersection of (S) and (S ∗ ).
We have: (P z0 )2 = r2 = |z0 − z| · |z0 − z ∗ | = (z0 w0 )2 − (P w0 )2 .
Thus, ∆z0 P w0 is right angled at P , and hence, (S) and (S ∗ ) are orthogonal.
(⇐) Suppose that every circle (S ∗ ) passing through z, z ∗ is orthogonal to (S). Particularly,
the line zz ∗ is orthogonal to (S), so z, z ∗ , z0 are collinear.
If z0 is between z and z ∗ , the circle with diameter zz ∗ is not orthogonal to (S), a contradiction.
Thus, arg(z − z0 ) = arg(z ∗ − z0 ).
Let (S ∗ ) be a circle passing through z, z ∗ , and let P be an intersection of (S) and (S ∗ ). Since
(S) and (S ∗ ) are orthogonal, ∆z0 P w0 is right angled at P . Hence:
Proof.
Take arbitrarily a circle (C ∗ ) passing through w and w∗ . Then, f −1 (C ∗ ) = C is also a circle.
Since (C) goes through z and z ∗ , then according to Proposition 4.5, (C) is orthogonal to
(S). Since f preserves angles, C ∗ = f (C) is orthogonal to S ∗ = f (S). By Proposition 4.5
again, w and w∗ are symmetric with respect to (S ∗ ). □
Example 4.7. Find all linear fractional transformations that map Imz > 0 to |w| < 1.
Solution.
Assume that w is a satisfactory transformation. Let z0 ∈ {z|Imz > 0} be such that w(z0 ) = 0.
Since z0 and z0 are symmetric with respect to Imz = 0, so w(z0 ) = ∞ by the symmetry
z − z0
preservation of w. It follows that w has the form w = A .
z − z0
Then |w(0)| = 1 (since Im0 = 0), which is equivalent to |A| = 1, so A = eiα with α ∈ R.
z − z0
Hence, w = eiα (Imz0 > 0, α ∈ R) is the general solution. □
z − z0
Now, for each z ∈ C∗ , we denote by Logz the set {t ∈ C|et = z}. Suppose that t = α + iβ.
We have: ß
t α iβ i arg z α = ln |z|
e = z ⇔ e e = |z| e ⇔ .
β = arg z + k2π, k ∈ Z
Integration
Suppose that an I ∈ C exists with the property: For every ϵ > 0, there exists δ > 0 such
that for every partition τ of [a, b] with |τ | < δ, one has |στ (f ) − I| < ϵ for every way to
mark τ . Then, we write lim στ (f ) = I and call I the (line) integral of f along γ, denoted
|τ |→0
Z
by f (z) dz:
γ Z
I = f (z) dz.
γ
47
Functions of one complex variable
Hence:
(ii) The integration of a complex function has the same properties as a line integral of a
vector field.
Solution.
a) We have u(x, y) = x2 + y 2 + xy, v(x, y) = x − y. In addition, x = t2 and y = t + 1, so
dx = 2tdt, dy = dt.
Hence:
Z Z Z
f (z)dz = udx − vdy + i vdx + udy
γ γ γ
Z 1
= {[t4 + (t + 1)2 + t2 (t + 1)]2t − [t2 − (t + 1)]}dt
0
Z 1
+i {[t2 − (t + 1)]2t + [t4 + (t + 1)2 + t2 (t + 1)]}dt
0
157 39
= +i .
30 20
b) A parametrization of γ is γ(t) = cos t + i sin t, t ∈ [0, π]. We have u(x, y) = y 2 , v(x, y) =
−x2 . In addition, x = cos t and y = sin t, so dx = − sin tdt, dy = cos tdt.
Hence:
Z Z Z
f (z)dz = udx − vdy + i vdx + udy
γ γ γ
Z 0
− sin3 t + cos3 t dt
=
Zπ 0
cos2 t sin t + sin2 t cos t dt
+i
π
4 2
= −i .
3 3
□
Example
Z 1.2. Let γ be a piecewise smooth curve from A = i + 1 to B = 2i − 3. Calculate
z 2 dz.
γ
Solution.
2 2 2 2
Z z = (x + iy) = x − y + i2xy.
Note that Z Z
Hence, z dz = I1 + iI2 , where I1 = (x − y )dx − 2xydy, I2 = 2xydx + (x2 − y 2 )dy.
2 2 2
γ γ γ
∂Q ∂P
Let P (x, y) = x2 − y 2 , Q(x, y) = −2xy. Since = = −2y on R2 , I1 is independent of
∂x ∂y
γ, and P dx + Qdy is the total differential of some U (x, y) on R2 . Moreover, U is given by:
Z x Z y
U (x, y) = P (x, y)dx + Q(0, y)dy + C
Z0 x 0
Z y
= (x2 − y 2 )dx + 0dy + C
0 0
Å 3
x3
ã
x 2
= − xy + 0 + C = − xy 2 + C.
3 3
òB
x3
Z Å ã ï 3
2 x 2
Therefore, I1 = d − xy = − xy .
γ 3 3 A
òB
y3 y3
Z Å ã ï
2 2
Similarly, I2 = d x y − = x y− .
γ 3 3 A
So:
ï 3 ãò2i−3
y3
Z Å
2 x 2 2
z dz = − xy + i x y −
γ 3 3 i+1
1 3 2i−3
= x + 3x(iy)2 + 3x2 (iy) + (iy)3 i+1
3
1 2i−3
= (x + iy)3 i+1
3
ï 3 ò2i−3
z
=
3 i+1
11 44
= +i .
3 3
□
1.3. Properties
Let γ be a piecewise smooth curve and f, g be continuous functions on γ.
Z Z Z
(i) (αf (z) + βg(z))dz = α f (z)dz + β g(z)dz, where α, β ∈ C.
γ γ γ
(iii) Suppose that γ1 : [a, c] → C and γ2 : [c, b] → C are piecewise smooth curves with
γ1 (c) = γ2 (c) (a < c < b). Let γ0 : [a, b] → C be such that γ0 |[a,c] = γ1 , γ0 |[c,b] = γ2 .
Then
Z Z Z
f (z)dz = f (z)dz + f (z)dz.
γ0 γ1 γ2
(v) We have
Z Z Z
f (z)dz ≤ |f (z)||dz| = |f (z)|ds,
γ γ γ
Indeed, suppose that γ(t) = x(t) + iy(t). Then dz = (x′ (t) + iy ′ (t))dt, so |dz| =
p
x′2 (t) + y ′2 (t)dt = ds. The conclusion follows from
Z n
X n
X n
X
|f (z)||dz| = |f (ck )||∆zk | and f (ck )∆zk ≤ |f (ck )||∆zk |.
γ k=1 k=1 k=1
(vi) It is immediate from (v) that if |f (z)| ≤ M on γ and lγ is the length of γ, then:
Z Z Z Z
f (z)dz ≤ |f (z)||dz| = M |dz| = M ds = M lγ .
γ γ γ γ
∞
X
(viii) Suppose that fn is a series of continuous functions on the region D and converges
n=0
uniformly to the function f , γ ⊂ D. Then:
∞ Z
X ∞
Z X Z
fn (z)dz = fn (z)dz = f (z)dz.
n=1 γ γ n=1 γ
Indeed, take any ϵ > 0. Let lγ > 0 be the length of γ. By the uniform convergence of
X∞
fn :
n=0
n
X ϵ
∃n0 : n > n0 =⇒ fi (z) − f (z) < , ∀z ∈ D.
i=1
lγ
Then, for all n > n0 , we have:
n Z Z Z n
!
X X ϵ
fi (z)dz − f (z)dz = fi (z) − f (z) dz ≤ · lγ (by (vi)) = ϵ.
i=1 γ γ γ i=1
lγ
Z
Example 1.3. Calculate I = zdz, where
γ
a) γ(t) = t2 + it, t ∈ [0, 1].
b) γ is the upper half of the unit circle x2 + y 2 = 1 from (−1, 0) to (1, 0).
Solution.
a) Using (iv) gives:
Z 1
I = t2 + it · (t2 + it)′ dt
Z0 1
= (t2 − it)(2t + i)dt
Z0 1
= [(2t3 + t) − it2 ]dt
Z0 1 Z 1
3 1
= (2t + t)dt − i t2 dt = 1 − i.
0 0 3
b) A parametrization of γ is γ(t) = eit , t ∈ [0, π]. Using (iv) gives:
Z 0 Z 0
I= it it
e ie dt = e−it ieit dt = −iπ.
π π
Proof.
Suppose that F is a primitive of f on D. It is easy to see that for each c ∈ C, F + c is also
a primitive of f .
Let G be a primitive of f on D. Set G(z) − F (z) = u(x, y) + iv(x, y). We have (G − F )′ =
G′ − F ′ = f − f = 0. On the other hand:
Theorem 1.1. (The Newton - Leibniz formula) Suppose f has a primitive F on the
region D and z1 , z2 ∈ D. Then, for any piecewise smooth curve γ ⊂ D from z1 to z2 , we
have: Z z2
f (z) dz = F (z2 ) − F (z1 ) = F (z) .
γ z1
Proof.
Let γ = γ(t), t ∈ [a, b], γ(a) = z1 , γ(b) = z2 . We have:
Z Z b Z b
′
f (z) dz = f (γ(t)) · γ (t) dt = [F (γ(t))]′ dt.
γ a a
□
Z
Example 1.4. Calculate I = ez dz, where γ is a piecewise smooth curve from i to 1 + i.
γ
Solution.
Applying the Newton - Leibniz formula yields
1+i
I = ez = e1+i − ei = ei (e − 1).
i
f (z)dz.
γ
is called the index of γ with respect to the point z0 . It is also sometimes called the winding
number of γ around z0 .
Example 1.5. Find the index of the curve γ with respect to the point z0 , where
Solution. Z
2π Z 2π Z 2π
1 (z0 + reint )′ 1 inreint 1
j(γ, z0 ) = dt = dt = in dt = n. □
2πi 0 (z0 + reint ) − z0 2πi 0 reint 2πi 0
2. Cauchy theorem
2.1. Cauchy’s integral formula
Theorem 2.1. Assume that f is holomorphoic on G. The, every simply domain D such
that D ⊂ G, we have Z
f (z)dz.
∂D
Proof:
Step 1: Assume that D is a triangular domain, consider l equals perimeter D.
Z
Let’s consider f (z)dz = M .
∂D
Dividing D into 4 triangles by centers of their edges
Let’s consider γi , i = 1, 4 is a boundary of 4 small triangles we have:
Z 4 Z
X 4
X Z
M= f (z)dz = f (z)dz ≤ f (z)dz .
∂D i=1 γi i=1 γi
Then, there exists a small triangle D1 with the boundary γi0 such that
Z
M
f (z)dz ≥
∂D1 4
l
in which the perimeter D1 = .
2
Dividing D1 into 4 triangles by centers of their edges, similarly, there exists triangle D2
such that Z
M
f (z)dz ≥
∂D2 42
l
in which the perimeter D2 = .
22 Z
M
By continuing that process Dn such that Dn+1 ⊂ Dn , f (z)dz ≥ , the perimeter
∂Dn 4n
l
Dn = n .
2
The sequence {Dn } thắt dần so there exists z0 ∈ Dn for all n.
As f is holomorphic z0 , then
As ϵ(z) is small z → z0 for all ϵ0 , there exists δ such that |ϵ(z)| < ϵ0 , ∀z ∈ B(z0 , δ).
With n is large enough, we have Dn ⊂ B(z0 , δ). Then,
ϵl2
Z
M
≤ ϵ(z)(z − z0 )dz ≤ |ϵ(z)| max |z−z0 |. chu vi Dn < ϵ. chu vi Dn . chu vi Dn = .
4n ∂Dn z∈∂Dn 4n
Step 3: With an arbitrary D , for all ϵ > 0, choose a polygon Dϵ with the vertices ∂D
such that Z Z
f (z)dz − f (z)dz < ϵ. (∗)
∂D ∂Dϵ
Z Z
As f (z)dz = 0 so f (z)dz < ϵ. Given ϵ → 0 we get the proof.
∂Dϵ ∂Dϵ
We need to show that with ϵ > 0, there exists Dϵ such that (∗). As f is continuous on D
so it is uniformly continuous D. Then, with ϵ > 0, there exists δ > 0 such that |z − z ′ | < δ,
so |f (z) − f (z ′ )| < ϵ.
>
Z zi ∈ ∂D, i = 1, n such that l(zi , zi+1 ) < δ.
Choose Dϵ with the vertices
Note: zn+1 = z1 , set l = ds is a perimeter of domain D.
∂D
Z ∞
X
We have known: f (z)dz = lim f (ck )∆zk .
n→+∞
γ k=1
Z Z Z n
X Z
f (z)dz − f (z)dz = f (z)dz − f (z)dz
∂D ∂Dε ∂D i=1
[zi ;zi+1 ]
Z n
X Z n
X Z
= f (z)dz − f (z1 ) dz + [f (z1 ) − f (z)] dz
∂D i=1 i=1
[zi ;zi+1 ] [zi ;zi+1 ]
Z n
X Z n
X Z
≤ f (z)dz − f (zi ) dz + [f (z1 ) − f (z)] dz
∂D i=1 > i=1
zi zi+1 [zi ;zi+1 ]
n
Z X
P
≤ [f (z) − f (zi )] dz +
i=1
>
zi zi+1
[zi ;zi+1 ](f (zi )−f (z))dz
Xn Z Z
≤ |f (z) − f (zi )| |dz| + |f (zi ) − f (z)| |dz|
i=1 >
zi zi+1 [zi ;zi+1 ]
n
ε > ε
l (zzi+1 ) + l (>
X
≤ zi zi+1 )
i=1
2l 2l
ε
= 2l = ε
2l
with l is the length ∂D.
Proof: Firstly, we prove: f is a function. Assume that γ1 , γ2 are the two curves D con-
necting z0 và z.
Let’s consider γ = γ1 ∪ γ2 . Then, γ is a closed curve on D.
By applying the Cauchy theorem Dγ (a domain limited by γ) we have:
Z Z Z Z Z
0 = f (z)dz = f (z)dz + f (z)dz = f (z)dz − f (z)dz
γ γ1 γ2 γ1 γ2
f (η)
connected domain D \ B(z, r) : . We get:
η−z
Z Z Z
f (η) f (η) f (η)
0= dη = dη − dη
η−z η−z η−z
∂D∪ ∂D ∂Br
Therefore, Z Z
f (η) f (η)
dη = dη
η−z η−z
∂D ∂Br
Notice Z
f (η)
dη = 2πi
η−z
∂Br
Therefore,
1
= . Maxη∈Br |f (η) − f (z)| .2πr −→ 0 khi r ⇒ 0
2πr
Example 3.1. Compute the integral
ez
Z
I= dz
√
z2 + 2
|z−i 2|=1
we have
ez eη
Z √ Z √
z + i√2 η+i 2
dz = √ dη
√
z−i 2 √
η−i 2
|z−i 2|=1 |η−i 2|=1
Applying Cauchy’s integral domain, we have
√
√ ei 2
I = f (i 2).2πi = √ .2πi
2i 2
Example 3.2. Comupute Z
z+2
I= dz
z2 + 1
|z|=2
Let r be small enough B(i; r) ⊂ B(O; 2), B(−i; r) ⊂ B(O; 2), B(i; r) ∩ B(−i; r) = ∅
z+2
f (z) = 2 which is holomorphic on the multi-domain Dand B(O; 2)\(B(i; r)∪B(−i; r)) =
z +1
D
Applying the Cauchy’s theorem on the domain D, we have:
Z Z Z Z
0 = f d(z) = f d(z) − f d(z) − f d(z)
∂D ∂ ∂1 ∂2
So, Z Z Z
f d(z) = f d(z) − f d(z)
∂ ∂1 ∂2
With
Z Z z+2
z+2 z + i d(z) = 2πi.g(i) = 2πi. i + 2
dz =
z2 + 1 z−i 2i
∂1 |z−i|=r
z+2
z − i d(z) = 2πi.h(−i) = 2πi. −i + 2
Z Z
z+2
dz =
z2 + 1 z+i −2i
∂2 |z−i|=r
F : C \ γ −→ C
Z
1 f (η)
z 7→ . dη
2πi η−z
γ
Proposition 3.1. The function F defined above has derivatives of all orders, and:
Z
(n) n! f (η)
F (z) = . dη
2πi (η − z)n+1
γ
b b
f (γ(t)).γ ′ (t)
Z Z Z
1 f (η) 1
F (z) = dη = dt = g(t, z)dt
2πi η−z 2πi a γ(t) − z a
γ
Let g(t, z) = u(t, x, y) + iv(t, x, y). As gz′ exists, then u, v have partial derivatives with respec
to x, y and satisfying C − R.
Suppose that F (z) = U(x, y) + iV(x, y) then
Z b Z b
U(x, y) = u(t, x, y)dt, V(x, y) = v(t, x, y)dt
a a
and Z b Z b
U′x (x, y) = u′x (t, x, y)dt, U′y (x, y) = u′y (t, x, y)dt
a a
Z b Z b
Vx′ (x, y) = vx′ (t, x, y)dt, Vy′ (x, y) = vy′ (t, x, y)dt
a a
Then U, V satisfy the C − R condition so F is differentiable
Z b
′ ′ ′
F (z) = Ux + Vy = (u′x (t, x, y) + ivx′ (t, x, y)) dt
a
b b
f (γ(t)).γ ′ (t)
Z Z Z
1 1 f (η)
= gz′ (t, z)dt = 2
dt = dη
a 2πi a (γ(t) − z) 2πi (η − z)2
γ
1. f is infinitely differentiable on D.
Proof: 2)
1 R f (η)
Đặt F (z) = dz. According to the Cauchy integral formula, we have F = f
2πi ηD η − z
n! R f (η)
According to Cauchy integral classification formula, we have F (n) (z) = f (n) (z) = dη.
2πi ηD (η − z)n+1
R ez
Example 3.4. Compute the integral I = 3
dz.
1 z(z − 1)
|z−1|=
2
ez 1
Consider f (z) = holomorphic on|z − 1| ≤ . So,
z 2
ez
Z z Z
e 2πi 2! z 2πi ′′
dz = . dz = f (z).
z(z − 1)3 2! 2πi (z − 1)3 2!
1 1
|z−1|= |z−1|=
2 2
Z M ax |f (η)|
(n) n! f (η) n! |η−z0 |=R M.n!
f (z0 ) = . dη ≤ . − 2πR =
2πi (η − z0 )n+1 2π Rn+1 Rn
|η−z0 |=R
Z2π
1
f z0 + Reit dt.
f (z0 ) =
2π
0
Z2π Z2π
f (z0 + Reit )
Z
1 f (z) 1
f z0 + Reit dt.
f (z0 ) = dz = dt =
2πi z − z0 Reit 2π
|z−z0 |=R 0 0
Proof : Suppose that there exist z0 ∈ D such that |f (z)| = max{|f (z)| : z ∈ D}. Indeed,
then f is a constants function o D.
Choose R > 0 such that B(z0 , R) ⊂ D. According to the Mean Value Theorem:
Z 2π Z 2π
1 it 1
|f (z0 )| = f (z0 + Re )dt ≤ f (z0 + Reit ) dt.
2π 0 2π 0
1 2π
Z
We also have |f (z0 )| = |f (z0 )|dt.
Z 2π π 0
1
We deduce 0 ≤ ( f (z0 + Reit ) − |f (z0 )|)dt.
2π 0
Hence, |f (z0 + Reit )| = |f (z0 )| = M for all t ∈ [0; 2π].
Let R → 0, we get |f (z0 )| = M for all z ∈ B ∗ (z ′ , r). Thus A is open.
Take (zn ) ⊂ A, zn → z. We have |f (zn )| = M and f is continuous, so |f (z)| = M .
In conclusion, A = D.
5. Harmonic function
5.1. Harmonic function
The function of 2 real variables u(x, y) in domain D is called harmonic function (or
potential functions) if its second partial derivatives are continuous and satisfy the Laplace
equation
∂ 2u ∂ 2u
∆u = + =0 (3.1)
∂x2 ∂y 2
for all (x, y) ∈ D.
Harmonic functions have many applications in mechanics and physics. Here we will es-
tablish some relationships between differential functions and harmonic functions, which show
the applicability of differentiable functions and of complex analysis in general.
Theorem 5.1. The real and imaginary parts of differentiable functions are harmonic func-
tions.
Proof. Let’s consider f (z) = u(x, y) + iv(x, y). As f is holomorphic then, by Cauchy -
Riemann
∂u ∂v ∂u ∂v
= , =− (3.2)
∂x ∂y ∂y ∂x
Differentiating both sides of the first equality w.r.t x, and the second one w.r.t y and taking
the sum give we have:
∂ 2u ∂ 2u
∆u = + =0
∂x2 ∂y 2
∂ 2v ∂ 2v
Similarly, we have ∆v = + = 0.
∂x2 ∂y 2
The harmonic function v is called harmonic conjugate with u if it satisfies equations.
The following theorem is obvious.
In the case (x0 , y0 ) can be joined with (x, y) by the parallel lines and the Cartesian coordinate
system, then the derivative of (3) or (4) can be formed as a defined derivative:
Zx Zy
∂u (t, y0 ) ∂u (x0 , t)
v(x, y) = − dt + dt, (3’)
∂y ∂x
x0 y0
Zx Zy
∂v (t, y0 ) ∂v (x0 , t)
u(x, y) = dt − dt. (4’)
∂y ∂x
x0 y0
Example 5.1. Find the holomorphic function f on a plane, knowing the real part of f is
u = xy.
y 2 x2
Å ã
Therefore, the holomorphic function is f (z) = xy + i − + ic, c ∈ R
2 2
Reit + z
The function is called Schwartz’s kernel
Reit − z
Proof. For all z ∈ B(0, R) by the Cauchy’s integral formula, one has:
Z2π
Reit
Z
1 f (η) 1
f Reit
f (z) = dη = dt, (3.6)
2πi η−z 2π Reit − z
SR 0
Especially
Z Z2π
1 f (η) 1
f Reit dt.
f (0) = dη = (3.7)
2πi η 2π
SR 0
Since |z| = |z| < R, R|z| < R2 , therefore, from Cauchy’s theorem
Z2π
z̄eit
Z
1 f (η) 1
f Reit
0= 2 dη = dt (3.8)
2πi R 2π z̄eit − R
SR η− 0
z̄
From (6) and (7), it implies
Z2π
1 1 1 Reit + z
f (z) − f (0) = f Reit . . it dt. (3.9)
2 2π 2 Re − z
0
and then
Z2π
1 1 1 Reit + z
f (0) = f (Reit ). . it dt. (3.10)
2 2π 2 Re − z
0
Note . The Schwartz formula can be used to define the holomorphic f in a circle B(0, R) wehn
we know the value of the real parts on the circumference SR and the value of the imaginary
part at 0.
Knowing the function f , then u = Ref . So, the harmonic function u in B(0, R) is defined
when we know the value of it on the circumference SR . In specific, we have:
Z2π
1 R2 − |z|2
u(z) = u Reit dt. (3.11)
2π |Reit − z|2
0
R2 − |z|2
The function Poissonend’s kernel
|Reit − z|2
Proof Since
Reit + z (Reit + z) (Re−it − z̄)
=
Reit − z |Reit − z|2
R2 − |z|2 2R (Imz.e−it )
= + i (3.12)
|Reit − z|2 |Reit − z|2
From the Schwartz’s Theorem, we have:
Z2π
1 R2 − |z|2
u(z) = Ref (z) = u Reit dt
2π |Reit − z|2
0
Note From (12) we see that the Poisson’s kernel is the real part of Schwartz’s kernerl.
1. Taylor series
1.1. Taylor series of holomorphic
Theorem 1.1. (Taylor). Assume f is a holomorphic function on domain D. Then at every
point z0 ∈ D, the function f can be expanded into a Taylor series.
∞
X
f (z) = cn (z − z0 )n . (4.1)
n=0
The series (1) converges and its sum is equal to f (z) at all z ∈ B (z0 , R), here R = d (z0 , ∂D)
and the coefficients cn of the series are uniquely determined by the formula:
f (n) (z0 )
cn = .
n!
Proof: Let r be an arbitrary number such that 0 < r < R, By the Cauchy integral
formula, for any z ∈ B (z0 , r)
Z
1 f (η)
f (z) = dη. (4.2)
2πi η−z
Sz ,x0
1
Since |z − z0 | < |η − z0 | so , one has:
η−z
1 1 1 1
= = .
η−z η − z0 − (z − z0 ) η − z0 1 − − z0
z
η − z0
∞
X (z − z0 )n
= (4.3)
n=0
(η − z0 )n+1
68
Functions of one complex variable
f (n) (z0 )
Z
1 f (η)
cn = dη =
2πi (η − z0 )n+1 n!
Sr
do đó ta được
∞ ∞
X n
X f (n) (z0 )
f (z) = cn (z − z0 ) = (z − z0 )n (4.4)
n=0 n=0
n!
Since the series (4) converges on every B(z0 , r) with r < R so the series (4) is convergent on
B(z0 , R).We have to check the uniqueness of the coefficients cn . Suppose
∞
X
f (z) = an (z − z0 )n
n=0
is another expansion. Taking the nth derivative of both sides of this equation and substituting
z = z0 we have:
f (n) (z0 ) = n!an
f (n) (z0 )
Then, an = .
n!
Note Theorem 1 is not valid in the case of real differentiability. For instance, a real
function 1
e− x2
®
nếu x ̸= 0
φ(x) =
0 nếu x = 0
may be infinitely differentiable at 0, φ(n) (0) = 0 for all n. This means that
∞
X
φ(n) (0)xn = 0
n=0
we have:
∞ Å
1X z−1 k
ã
1 1 1 1
=− =− . =−
z−3 2 − (z − 1) 2 z−1 2 k=0 2
1−
2
So,
∞ Å ã
1 X 1
= − k+1 (z − 1)k
z − 3 k=0 2
The convergence disk of the series is B(1, 2).
Lemma 1.1. Let {zn } be as in theorem 2, and φ a holomorphic function on D such that
φ(zn ) = 0 for every n ∈ N. Then φ ≡ 0 on D.
Proof: Since zn → z0 and φ are continuous φ(z0 ) = lim φ (zn ) = 0. Consider the Taylor
n→∞
expansion of φ in the neighborhood z0
∞
X
φ(z) = cn (z − z0 )n ,
n=0
Note: Theorem 2 implies that an identity holding between holomorphic functions for real
values also holds for complex values. For instance, the Å
identity
ã sin2 z +cos2 z = 1 is equivalent
1 1
to f (z) = sin2 z + cos2 z − 1 = 0. As ⇒ 0 and f = 0 for all n so f (z) ≡ 0 on C,
n n
which means sin2 z + cos2 z = 1 for all z ∈ C