MA271 Lecture Notes
MA271 Lecture Notes
Contents i
1 Introduction 2
1.1 Review of limits of sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Review of continuity and differentiability . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Review of integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
i
CONTENTS
6 The Derivative 39
6.1 Directional derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
6.1.1 Directional derivative and continuity . . . . . . . . . . . . . . . . . . . . . . . . . 40
6.2 The (Fréchet) Derivative as an affine linear approximation . . . . . . . . . . . . . . . . . 40
6.2.1 Affine linear approximation in the 1-variable case . . . . . . . . . . . . . . . . . . 40
6.2.2 The (Fréchet) Derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
6.2.3 Differentiability of components of vector-valued functions . . . . . . . . . . . . . . 41
6.2.4 Relation between the derivative and directional derivative . . . . . . . . . . . . . . 42
6.3 Partial derivatives, gradient and Jacobian matrix . . . . . . . . . . . . . . . . . . . . . . . 42
6.3.1 Algebraic rules for partial derivatives. . . . . . . . . . . . . . . . . . . . . . . . . 43
6.3.2 Gradient and Jacobian matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
6.3.3 Why so many different notations for the same thing?! . . . . . . . . . . . . . . . . 44
6.4 Geometric approximation and approximation of functions (NOT COVERED IN CLASS AND
NOT EXAMINABLE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.4.1 Tangent to a curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.4.2 Tangent plane of a surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
6.4.3 Graph of a scalar function of 2 variables . . . . . . . . . . . . . . . . . . . . . . . 46
6.4.4 Orders of approximation of a function . . . . . . . . . . . . . . . . . . . . . . . . 46
6.5 Examples of direct calculation of the derivative from its definition (NOT COVERED IN
CLASS AND NOT EXAMINABLE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
6.5.1 Differentiation of matrix-valued functions . . . . . . . . . . . . . . . . . . . . . . 47
6.6 The Chain Rule (NOT COVERED IN CLASS AND NOT EXAMINABLE) . . . . . . . . . 48
6.6.1 Jacobian form of chain rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
6.6.2 Calculating with the chain rule and gradient . . . . . . . . . . . . . . . . . . . . . 50
6.6.3 Another proof of Proposition 6.9 . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.6.4 Application of chain rule to the verification of a PDE satisfied by a function . . . . 51
6.7 Continuity of partial derivatives implies differentiability (NOT COVERED IN CLASS AND
NOT EXAMINABLE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.7.1 The space of continuously differentiable functions . . . . . . . . . . . . . . . . . . 53
7 Complex Analysis 54
7.1 Review of basic facts about C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
7.2 Power Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
7.2.1 The exponential and the circular functions . . . . . . . . . . . . . . . . . . . . . . 62
7.2.2 Argument and Log . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
7.3 Complex integration, contour integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
7.3.1 Links with Green’s and Gauss’ Theorems . . . . . . . . . . . . . . . . . . . . . . . 70
7.4 Additional material (NOT COVERED AND NON-EXAMINABLE) . . . . . . . . . . . . . 75
7.4.1 Consequences of Cauchy’s Theorem (NOT COVERED AND NOT EXAMINABLE) 77
7.4.2 Applications of Cauchy’s formula to evaluate integrals in R (NOT COVERED AND
NOT EXAMINABLE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
Bibliography 87
ANALYSIS III 1
Chapter 1
Introduction
MA271, Mathematical Analysis 3 is a self-contained module. Only material lectured in class will be
examined. Non-examinable will be clearly marked.
The course covers the following topics:
• Differentiation.
The module does not follow any specific source. There references at the end of the notes cover most
of the topics in the module. I would be happy to supply a list of references that can be used to expand
any of the Chapters in the notes.
The course relies heavily on material covered in first-year analysis. We recall several of the main notions
that we will use.
• Cauchy sequences
• Summation
• The derivative
• Uniform continuity.
2
CHAPTER 1. INTRODUCTION
Definition 1.1. A sequence (an ) converges to a limit l ∈ R if for every ε > 0 there exists N ∈ N such
that
|an − l| < ε for every n > N.
In this case we write an → l as n → ∞.
• The shift rule: For any fixed k ∈ N, an → l as n → ∞ if and only if an+k → l as n → ∞. (This
effectively means that for the question of convergence, we can disregard the first k terms of the
sequence (for any fixed k we like).
(i) an + bn → a + b;
(ii) an bn → ab;
(iii) if b 6= 0 then an /bn → a/b.
Definition 1.2. Given f : Ω → R, we say that f is continuous at x ∈ Ω if for every ε > 0 there exists
δ = δ(x, ε) > 0 such that
The key point to note from the definition above is that given a function f , ε > 0 and a point x there
exists δ, but δ can depend on ε and x (and of course f ).
Definition 1.3. Given f : Ω → R, we say that f is uniformly continuous if for every ε > 0 there exists
δ = δ(ε) > 0 such that
x, y ∈ Ω and |x − y| < δ =⇒ |f (y) − f (x)| < ε. (1.2)
ANALYSIS III 3
CHAPTER 1. INTRODUCTION
Before we prove the result let’s consider a couple of examples in which the closed, bounded interval
[a, b] is replaced by an unbounded or an open domain.
Consider f (x) = ex , defined in R. Clearly this is a continuous function, but not uniformly continuous.
Indeed, since f grows faster and faster for larger x it is possible to find arbitrarily small intervals in which
f changes by at least ε. This example shows that the result in Theorem 1.4 is not necessarily true for
unbounded domains.
We can also consider g(x) = x1 on (0, 1). Just as in the previous example, near zero, the function g
grows to infinity faster and faster as we approach the origin, making it impossible to find δ independent of
x that satisfies (1.2).
Proof of Theorem 1.4. We will argue by contradiction. That would mean that there exist ε > 0 and xn , yn
such that |xn − yn | ≤ n1 but |f (xn ) − f (yn )| > ε.
The sequences {xn } and {yn } are bounded, as they are in [a, b], and therefore we can apply the Bolzano–
Weierstrass theorem to obtain convergent subsequences {xnk }∞ k=1 to x and {ynk }k=1 to y. Notice that
∞
1
|x − ynk | ≤ |x − xnk | + |xnk − ynk | ≤ |x − xnk | + −−−−−−−→ 0,
nk k→∞
which implies that x = y. However we know that |f (xnk ) − f (ynk )| > ε for all k. Since f is continuous,
taking limits as k goes to infinity we obtain 0 = |f (x) − f (x)| > ε, which is a contradiction.
Lemma 1.6. Suppose I is an open interval, f : I → R and c ∈ I. Then f is differentiable at c if and only
if there exists a number A and a function ε with the properties that for all x
Definition 1.7. Given f : [a, b] → R and a partition P = {I1 , . . . , In } of [a, b] we define the upper
Riemann sum of f with respect to P as
n
X
U (f, P ) := Mk |Ik |,
k=1
ANALYSIS III 4
CHAPTER 1. INTRODUCTION
Figure 1.1 shows the intuitive idea for calculating an integral, displaying the Lower and Upper Riemann
sums, for a uniform partition with 10 intervals (for f (x) = x2 ).
x2
x2
1.
0.8
0.6
0.4
0.2
0. x 0 x
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
U (f ) := inf U (f, P ).
P ∈P
Definition 1.9. Given f : [a, b] → R bounded we say that it is Riemann integrable if and only if L(f ) =
´b ´b
U (f ), and define its Riemann integral, denoted by a f (x)dx or a f , by
ˆ b
f (x)dx := L(f ) = U (f ).
a
The following result will also prove useful in showing that a function is integrable.
Theorem 1.10. Let f : [a, b] → R be a bounded function. Then f is integrable if and only if for every
ε > 0 there exists a partition P of [a, b] such that
Theorem 1.13. Let f, g : [a, b] → R be Riemann integrable functions such that f ≤ g. Then
ˆ b ˆ b
f≤ g.
a a
ANALYSIS III 5
CHAPTER 1. INTRODUCTION
Theorem 1.14. Let f : [a, b] → R be an integrable function. Then |f | is integrable and we have
ˆ b ˆ b
f ≤ |f |.
a a
The Fundamental Theorem of Calculus explores the relationship between integration and differentiation,
and how under sufficient conditions they can be understood as inverse operations. The first result we
consider is when the integral of a derivative is the original function.
Theorem 1.15. Let F : [a, b] → R be a continuous function that is differentiable on (a, b) with F 0 = f .
Assume that f : [a, b] → R is an integrable function. Then
ˆ b
f (x)dx = F (b) − F (a).
a
Theorem 1.16. Let f : [a, b] → R be an integrable function and define the function F : [a, b] → R by
ˆ x
F (x) := f (t)dt.
a
Then F is continuous on [a, b]. Additionally if f is continuous at c ∈ [a, b] then F 0 (c) = f (c), with the
derivatives at a and b understood as one-sided derivatives.
ANALYSIS III 6
Chapter 2
In this Chapter we will consider sequences and series of functions and aspects relating to pointwise and
uniform convergence and its interactions with continuity, integrability and differentiability questions.
f4 f3 f2 f1
0.6
0.4
0.2
Notice that fn (0) = 0 for every n, but that for every x ∈ (0, 1] we have limn→∞ x1/n = 1. As a result
the limit of the sequence (fn ) is (
0 x = 0,
f (x) =
1 x ∈ (0, 1].
Remark 2.3. Notice that the above example shows that the pointwise limit of a sequence of continuous
functions need not be continuous. It also produces a counterexample for the commutativity of the limits.
We have
lim lim fn (x) 6= lim lim fn (x),
n→∞ x→0+ x→0+ n→∞
as the left-hand side equals zero, while the right-hand side equals one.
7
CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Pointwise convergence clearly does not preserve continuity. It can also be very non-uniform, in the sense
that while fn (x) → 0 for every x we may have supx |fn (x)−f (x)| → C > 0 or even supx |fn (x)−f (x)| →
∞ as n goes to infinity, as shown in the next examples.
It is easy to see that gn and hn are continuous and converge to the function f = 0. However, for every n
we have gn (1/(2n)) = 1 (with that being the maximum of gn ) and therefore
The situation is worse for the sequence (hn ), known as the Witch’s hat. Indeed hn (1/(2n)) = n, which
shows that while hn → 0 we have
sup |hn (x) − 0| → ∞.
x∈[0,1]
Pointwise convergence and integrability do not interact´as one ´would hope. Indeed, even if we assume
that the pointwise limit is integrable we may not have lim fn = lim fn .
Example 2.5. Consider fn (x) = χ[n,n+1) (x), where χI is the indicator of the set I, i.e., takes value 1 if
x ∈ I and zero otherwise. Clearly fn converges pointwise to f = 0. However,
ˆ ˆ
1 = fn 6= f = 0.
We can think of this, as “the mass escaping to´ infinity” (along the x axis). In the latter calculation, we
+∞
are considering the improper Riemann integral −∞ on all of R.
Another example of this phenomena,´ can be found by considering gn (x) = nχ(0,1/n) (x) we also have
that gn converges to 0, while having gn = 1 for every n. We can think of this as “pointwise convergence
allowing the mass to go to infinity” (along the y axis this time). The Witch’s hat above also provides a
similar example, in this case with continuous functions.
Example 2.6. Another sequence that will play a role in several modules this year is fn (x) = sin(n x). This
sequence is connected to Fourier series and will be heavily studied in MA250 PDE for example. Notice
that for x = kπ with k ∈ Z the limit exists and equals 0. If x = p/q π with p/q ∈
/ Z then there is no limit.
Indeed sin(nqx) = 0 while sin((2nq + 1)x) = sin(x) 6= 0. If x is an irrational multiple of π, then the rest
of the division of nx by 2π is dense in [0, 2π] and there is no limit.
Despite the fact that sin(nx) does not have a limit for most x, one can show that for every integrable
function f ˆ π
f (x) sin(nx)dx → 0 as n → ∞.
−π
This result, known as the Riemann–Lebesgue Lemma, suggests that sin(nx) goes to zero in some sense.
We can also consider the sequence gn (x) = cos(nx)
n . As cosine is a bounded function it is easy to see that
gn converges pointwise to 0. Since gn are smooth we can also consider gn0 (x) = − sin(nx). This tells us
that even for smooth functions, having gn converge pointwise to g does not imply that gn0 converges to g 0
even if g is smooth.
´
The final example we consider is one of a sequence (fn ) such that (fn − f )dx converges to zero, but
where fn does not converge pointwise to f .
ANALYSIS III 8
CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Definition 2.8. Let fn : Ω → R be a sequence of functions. We say that (fn ) converges uniformly to
f : Ω → R if and only if for every ε > 0 there exists N (ε) such that |fn (x) − f (x)| < ε for every x ∈ Ω
and for all n > N (ε).
The key different with pointwise convergence is that N depends only on ε and not on x. For pointwise
convergence we first froze x and consider the convergence of fn (x) to f (x). We will denote uniform
convergence by fn ⇒ f .
As before we are not making any assumption on Ω. In order to simplify the presentation we introduce
the notation
kf k∞ = sup |f (x)|.
x∈Ω
Remark 2.9. Clearly uniform convergence implies pointwise convergence. The converse is of course false,
as can be seen by considering the sequence from Remark 2.3. Namely, we note that fn (1/2n ) = 1/2 and
so kfn − f k∞ ≥ 1/2. Alternatively, one can argue by contradiction and apply Theorem 2.13 below.
Definition 2.10. A sequence (fn ) of functions in Ω is called uniformly Cauchy if and only if for every ε > 0
there exists N (ε) such that kfn −fm k∞ < ε for all n, m > N (ε) (or alternatively supx∈Ω |fn (x)−fm (x)| <
ε for all n, m > N (ε)).
Theorem 2.11. A sequence (fn ) is uniformly convergent if and only if it is uniformly Cauchy.
ANALYSIS III 9
CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Proof. Assume that (fn ) is uniformly convergent to f , i.e. for every ε there exists N such that kfn −f k∞ <
ε/2 for all n > N . Then, for m, n > N
For the converse, assume (fn ) is uniformly Cauchy. That means that for every x, fn (x) is a Cauchy
sequence in R and therefore convergent. That means there exists f (x) such that fn (x) converges to f (x)
at least pointwise. Now, we know that given ε > 0 there exists N (ε) > 0 such that |fn (x) − fm (x)| < ε/2
for every x and all n, m > N (ε). That is
fm (x) − ε/2 < fn (x) < fm (x) + ε/2 for all x, and all n, m > N (ε).
As the left-hand side holds for all m > N (ε) we can take limits as m goes to infinity. We find
f (x) − ε/2 ≤ fn (x) ≤ f (x) + ε/2 for all x, and all n > N (ε).
Remark 2.12. It is worth noting that k · k∞ is a norm in the space of bounded functions in Ω ( we make
no assumptions about it being open, closed, bounded or unbounded). k · k∞ is referred to as the supremum
norm or uniform norm. Recall that by norm we mean that it satisfies
3. kf + gk∞ ≤ kf k∞ + kgk∞ .
Theorem 2.13. Let (fn ) be a sequence of continuous functions in Ω that converges uniformly to f : Ω → R.
Then f is continuous.
Proof. First notice that the uniform convergence implies that given any ε > 0 there exists N > 0 such
that kfn − f k∞ < ε/3 for all n > N . In order to show that f is continuous at x0 ∈ Ω we need to show
that given ε there exists δ = δ(ε) such that for all x ∈ (x0 − δ, x0 + δ) ∩ Ω we have |f (x) − f (x0 )| < ε.
With N as above, we choose n > N , fixed from now on. Since fn is continuous at x0 we know that there
exists δ = δ(ε) such that for all x ∈ (x0 − δ, x0 + δ) ∩ Ω we have |fn (x) − fn (x0 )| < ε/3.
We estimate |f (x) − f (x0 )| using the triangle inequality
We will denote the space of bounded, continuous functions with the uniform norm by (Cb ; k · k∞ ).
Theorem 2.14. (Cb ; k · k∞ ) is a complete space, i.e. every Cauchy sequence converges to a continuous
bounded function.
ANALYSIS III 10
CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Proof. We need to show that if (fn ) is Cauchy in the space (Cb ; k · k∞ ), then there is a limit, and that
the limit is bounded and continuous. First notice that a Cauchy sequence in (Cb ; k · k∞ ) is, by definition,
a uniformly Cauchy sequence. Theorem 2.11 implies that the sequence is convergent and since all the
functions are continuous Theorem 2.13 implies the limit is continuous.
To see that it is bounded, notice that for every x ∈ Ω
|f (x)| ≤ |f (x) − fn (x)| + |fn (x)|
for every n. Since fn converges uniformly to f there exists n large enough |fn (x) − f (x)| < 1. For that
n, since fn is bounded we have |fn | ≤ M . These two inequalities lead to |f (x)| ≤ M + 1 for every x ∈ Ω,
proving the boundedness of f .
Remark 2.15. We could consider the interaction of uniform convergence and differentiation or integration.
2 x)
Consider for example fn (x) = sin(n
n . The sequence (fn ) converges to f = 0 uniformly. Indeed
sin(n2 x) 1
−0 ≤ ∀x.
n n
Clearly all the functions fn are smooth. The derivatives are given by fn0 (x) = n cos(n2 x). It is easy to see
that the sequence (fn0 ) does not converge uniformly (or pointwise). This example shows that while fn ⇒ f
we may not have fn0 ⇒ f 0 or even fn0 → f 0 .
1
To explore integrability, we consider gn (x) = 2n χ[−n,n] . Recall that strictly speaking we have not defined
Riemann
´ integration in R, but rather improper integration, via a limiting procedure. It is clear however
that gn =´ 1 for every n. The ´ sequence gn converges uniformly to g = 0 as we have |gn − 0| ≤ 1/(2n),
and so lim gn = 1 6= 0 = g. We reiterate that strictly speaking´ gn are´not Riemann integrable and we
will prove that in fact, on a bounded interval fn ⇒ f does imply fn → f .
Theorem 2.16. Lef (fn ), fn : [a, b] → R be a sequence of Riemann
´ integrable
´ functions that converges
uniformly to f : [a, b] → R. Then f is Riemann integrable and fn → f .
Proof. First we need to show that f is Riemann integrable, that is show that for every ε > 0 there exists
a partition P of [a, b] such that
U (f, P ) − L(f, P ) < ε.
Now, since fn ⇒ f we know that for any ε > 0 there exists N such that kfn − f k∞ < ε/(4(b − a)) for
n > N . For a fixed n > N since fn is integrable we know that given ε > 0 there exists a partition P of
[a, b] such that
ε
U (fn , P ) − L(fn , P ) < .
2
Now, for that P
X X
U (f, P ) − L(f, P ) = [sup f − inf f ]|Ik | = [sup(f − fn + fn ) − inf (f − fn + fn )]|Ik |
Ik Ik Ik Ik
" #
X
≤ kf − fn k∞ + sup fn + kf − fn k∞ − inf fn |Ik |
Ik Ik
X X
=2 kf − fn k∞ |Ik | + [sup fn − inf fn ]|Ik |
Ik Ik
Clearly the right hand side goes to zero as n goes to infinity by the uniform convergence of (fn ) to f .
ANALYSIS III 11
CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
In many circumstances it is necessary to consider functions of two variables (or more) from which we
construct new functions by integrating out some of the variables. We want to study several results in this
direction; we start by reviewing the notions of continuity and uniform continuity in two dimensions. The
definitions are analogous to Definitions 1.2 and 1.3.
Definition 2.17. Given f : Ω ⊂ R2 → R, we say that f is continuous at x if for every ε > 0 there exists
δ = δ(x, ε) > 0 such that
Note that | · | has been used both to denote Euclidean distance in the plane, as in |x − y|, as well as for
absolute value of a real number, in |f (y) − f (x)|.
Definition 2.18. Given f : Ω ⊂ R2 → R, we say that it is uniformly continuous if for every ε > 0 there
exists δ = δ(ε) > 0 such that
The key point here is that δ can be chosen independently of x. Similarly to Theorem 1.4 we have the
following result.
Theorem 2.19. Let f : Ω ⊂ R2 → R be a continuous function. Assume that Ω is closed and bounded.
Then it is uniformly continuous.
Proof. We will argue by contradiction. That would mean that there exists ε > 0 and xn , yn such that
|xn − yn | ≤ n1 but |f (xn ) − f (yn )| > ε.
The sequences {xn } and {yn } are bounded, as they are in Ω, which is closed and bounded, and therefore
we can apply Bolzano–Weierstrass to each component to obtain convergent subsequences {xnk }∞ k=1 to x
and {ynk }∞k=1 to y. Notice that
1
|x − ynk | ≤ |x − xnk | + |xnk − ynk | ≤ |x − xnk | + −−−−−−−→ 0,
nk k→∞
which implies that x = y. However we know that |f (xnk ) − f (ynk )| > ε for all k. Since f is continuous,
taking limits as k goes to infinity we obtain 0 = |f (x) − f (x)| ≥ ε, which is a contradiction.
Since f is continuous on [a, b]×[c, d] it is p uniformly continuous, and therefore given ε > 0 there exists δ such
that (x1 , t1 ), (x2 , t2 ) ∈ [a, b]×[c, d] with (x1 − x2 )2 + (t1 − t2 )2 < δ implies that |f (x1 , t1 )−f (x2 , t2 )| <
ε/(b − a). Therefore if |t − t0 | < δ we have |f (x, t) − f (x, t0 )| < ε/(b − a). As a result (2.3) becomes
ˆ b ˆ b
ε
|I(t) − I(t0 )| ≤ |f (x, t) − f (x, t0 )|dx < dx = ε,
a a b−a
ANALYSIS III 12
CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
We can also consider differentiating I with respect to t under sufficient regularity results for f .
Theorem 2.21. Lef f, ∂f ∂t be continuous functions on [a, b] × [c, d]. Then, for t ∈ (c, d)
ˆ ˆ b
d b ∂f
f (x, t)dx = (x, t)dx.
dt a a ∂t
´b ´b
Proof. Set F (t) := a f (x, t)dx, and G(t) := a ∂f ∂t (x, t)dx. We want to show that F is differentiable on
(c, d) and F = G. We consider the difference between the incremental quotient that is used to define a
0
derivative of F and the function we expect to be derivative, namely G. Let t0 ∈ (c, d) be given. Consider
h ∈ R \ {0} such that t0 + h ∈ [c, d]. We write
ˆ b
F (t0 + h) − F (t0 ) f (x, t0 + h) − f (x, t0 ) ∂f
− G(t) = − (x, t0 )dx ,
h a h ∂t
which by the Mean Value Theorem, becomes, for some τ between t0 and t0 + h
ˆ b ˆ b
∂f ∂f ∂f ∂f
= (x, τ ) − (x, t0 )dx ≤ (x, τ ) − (x, t0 ) dx.
a ∂t ∂t a ∂t ∂t
Now, since ∂f
∂t is continuous on [a, b] × [c, d] it is uniformly continuous, and therefore for every ε > 0 there
exists δ such that for |h| < δ and τ as above we have
∂f ∂f ε
(x, τ ) − (x, t0 ) <
∂t ∂t b−a
This is implies that for |h| < δ
ˆ b
F (t0 + h) − F (t0 ) ε
− G(t0 ) ≤ dx = ε.
h a b−a
Such δ > 0 can be found for all ε > 0. Hence F is differentiable at t0 and F 0 (t0 ) = G(t0 ).
By the FTC (Theorem 1.16), we know that F is continuous, with F (a) = 0. Also the first integral is
differentiable with ˆ ˆ d ! ˆ d
d t
f (x, y)dy dx = f (t, y)dy.
dt a c c
We know that ˆ t
d
f (x, y) dx = f (t, y).
dt a
We would now like to differentiate the second integral in F , namely
ˆ d ˆ t !
− f (x, y) dx dy
c a
ANALYSIS III 13
CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
by differentiating inside the first integral. For that Theorem 2.21 requires that
ˆ t !
f (x, y) dx
a
be continuous in [a, b] × [c, d] as a function of t and y. Theorem 2.20 proves that it is continuous as a
function of y and we actually know that it is differentiable as a function of t. However, continuity in each
of the variables separately does not ensure that the function is continuous on [a, b] × [c, d]. However, one
can modify the proof of Theorem 2.20 to show continuity in [a, b] × [c, d]. (This is left as an exercise.)
Then we are allowed to differentiate inside the integral and we obtain
ˆ ˆ t ! ˆ d ˆ t ! ˆ d
d d ∂
f (x, y) dx dy = f (x, y) dx dy = f (t, y) dy.
dt c a c ∂t a c
Therefore ˆ ˆ
d d
0
F (t) = f (t, y) dy − f (t, y) dy = 0,
c c
Since F is continuous on [a, b], F (a) = 0 and F 0 (t) = 0 we find F (b) = 0. This implies the result.
Remark 2.23. The continuity requirement is necessary in the previous Theorem. The following is a
counterexample to Fubini’s theorem when continuity fails at just a point. Let
x2 − y 2
f (x, y) = .
(x2 + y 2 )2
Notice that f is not continuous at the origin. We have
ˆ 1 2 y=1
x − y2 y 1
2 2 2
dy = 2 = ,
0 (x + y ) x + y2 1 + x2
y=0
and ˆ ˆ ! ˆ
1 1 1
x2 − y 2 1 π
dy dx = dx = .
0 0 (x2 + y 2 )2 0 1+x 2 4
−π
In the opposite direction we get by symmetry. The key here is that the function is not in L1 , i.e. |f |
4
is not integrable.
ˆ 1 ˆ 1 ! ˆ 1 ˆ x 2 ! ˆ 1 y=x ˆ 1
x2 − y 2 x − y2 y 1
2 + y 2 )2
dy dx ≥ 2 + y 2 )2
dy dx = 2 + y2
dx = dx = ∞.
0 0 (x 0 0 (x 0 x y=0 0 2x
Differentiation revisited.
We will use the notation C k (a, b) to denote functions that are k times continuously differentiable on
(a, b), and C ∞ (a, b) for functions that are infinitely differentiable on (a, b).
We have seen examples of sequences (fn ) that are differentiable, with (fn ) converging uniformly to f
but for which fn0 does not converge to f 0 . In fact it is easy to construct examples of C 1 functions that
converge uniformly for which f 0 does not exist. Consider
1/2
fn (x) = x2 + 1/n .
They are clearly C 1 as the x2 + 1/n never vanishes for fixed n. (fn ) converges uniformly to f (x) = |x|,
which is not smooth at the origin. To see this notice that if
1/2
A := x2 + 1/n − |x|
then √ 1/2 1
A ≤ (x + 1/ n)2 − |x| ≤ √ ,
n
and the uniform convergence follows.
ANALYSIS III 14
CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
Notice that since g is continuous this means that f is continuous. While the Theorem does not assume
that g is continuous, that
´ x is a consequence of the uniform convergence of fn to g, since fn are C . Now,
0 1
with fk : Ω → R. We begin by establishing the notion of pointwise convergence and uniform convergence
for a series.
Definition 2.25. Let (fk ) be a sequence of functions fk : Ω → R. Let (Sn ) be the sequence of partial
sums, with Sn : Ω → R defined by
n
X
Sn (x) = fk (x).
k=1
Then the series ∞
X
fk (x)
k=1
converges pointwise to S : Ω → R in Ω if Sn → S pointwise on Ω and it converges uniformly to S in Ω if
Sn ⇒ S uniformly in Ω.
Theorem 2.26. Let (fk ), with fk : [a, b] → R, be a sequence of Riemann integrable functions on [a, b].
Assume that ∞
P P∞
k=1 fk converges uniformly. Then k=1 fk is Riemann integrable on [a, b] and
ˆ bX∞ X∞ ˆ b
fk = fk .
a k=1 k=1 a
Proof. Sn is a finite sum of integrable functions and therefore integrable (by additivity). Since Sn converges
uniformly Theorem 2.16 implies that S = limn→∞ Sn is integrable and moreover
ˆ b ˆ b
lim Sn = S.
n→∞ a a
´b ´b P∞
Since fk and S = limn→∞ Sn = we obtain the result.
Pn
a Sn = k=1 a k=1 fk
ANALYSIS III 15
CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
P∞
Theorem 2.27. Let (fk ), with fk : [a, b] → R, be a sequence of C 1 functions such that k=1 fk converges
pointwise. Assume that ∞ 0
P
k=1 fk converges uniformly. Then
∞
!0 ∞
fk0 (x),
X X
fk (x) =
k=1 k=1
0
but since Sn0 = ( we obtain the result, namely
Pn Pn 0
k=1 fk ) = k=1 fk
∞ ∞
!0
fk0 (x) =
X X
fk (x) .
k=1 k=1
Theorem 2.28 (The Weierstrass M-test). Let (fk ) be a sequence of functions fk : Ω → R, and assume
that for every k there exists Mk > 0 such that |fk (x)| ≤ Mk for every x ∈ Ω and ∞
k=1 Mk < ∞. Then
P
∞
X
fk
k=1
converges uniformly on Ω.
Proof. Notice that it suffices to show that Sn := nk=1 fk is uniformly Cauchy (recall Theorem 2.11). Now
P
n
for all m, n > N.
X
Mk < ε
k=m+1
Now
n
X m
X n
X n
X n
X
|Sn (x) − Sm (x)| = fk (x) − fk (x) = fk (x) ≤ |fk | ≤ Mk ≤ ε,
k=1 k=1 k=m+1 k=m+1 k=m+1
ANALYSIS III 16
CHAPTER 2. SEQUENCES AND SERIES OF FUNCTIONS
1 1 1
0 ≤ fn ≤ n
φ≤ ,
4 2 4n
and that by the Weierstrass M-test we have the uniform convergence of the series. Since each fn is
continuous, and the convergence is uniform we have that f is C 0 .
Given x ∈ R we will choose the sign of hn = ± 4n+1 1
in such a way that the points 4n x and 4n (x + hn )
2 ] for some k ∈ Z. We make this choice of sign for
both belong to the same interval of length 1/2, [ k2 , k+1
hn because on each of these intervals [ 2 , 2 ], the function φ has constant slope +1 or −1.
k k+1
`
4m x < < 4m x + 4m hn .
2
Multiplying the above inequalities by 4n−m , we get
` 1
4n x < 4n−m < 4n x + .
2 4
Since 4n−m 2` , this is a contradiction to our choice of hn . Therefore,
Therefore An is an even integer if n is odd and an odd integer if n is even. Hence there is no limit as n
goes to infinity. Since hn goes to zero that proves that f is not differentiable.
ANALYSIS III 17
Chapter 3
3.1 Notation in Rn
The main vector spaces that we shall consider in this module are Rn , n ∈ N. Thus, by a vector x ∈ Rn
we mean the n-tuple (x1 , . . . , xn ), xi ∈ R, 1 6 i 6 n.
For ease of writing, a vector x ∈ Rn will be written as a row vector x = (x1 , . . . , xn ), xi ∈ R, 1 6 i 6 n.
However, in calculations vectors will be written as column vectors
x1
..
x = . .
xn
a11 . . . a1n
.. ..
A= . .
ak1 . . . akn
with respect to the standard bases of Rn and Rk , then the vector y := Ax is obtained by multiplying the
column vector x by the matrix A on the left. In index notation, if y = (y1 , . . . , yk ), then
n
X
yi = aij xj , i ∈ {1, . . . , k}.
j=1
fk (x1 , . . . , xn )
We normally use a, b and c at the start of the latin alphabet to denote real numbers (i.e. scalars) and
we shall use letters like x, y, p, q, u, v and w in the second half of the latin alphabet to denote vectors.
Thus we shall write ax for the vector (ax1 , . . . , axn ) without always spelling out that a ∈ R and x ∈ Rn .
In two and three dimensions, we will often be written f (x, y) and f (x, y, z). Finally, 0 will denote both
the zero vector (though we shall occasionally write it as (0, . . . , 0)) and the zero scalar!
1
A real valued function f : U → R will be referred to as a scalar function.
18
CHAPTER 3. BASIC RESULTS ABOUT RN
|x · y| 6 kxkkyk.
It follows from
2
0 6 kyk2 x − (x · y)y = kyk4 kxk2 − (x · y)2 kyk2 .
Definition 3.1 (Angle between two nonzero vectors). The Cauchy-Schwarz inequality implies that, if x
and y are both nonzero, then there exists unique θ ∈ [0, π] such that
Proof.
kx − yk 6 kx − zk + kz − yk, (3.3)
which corresponds to the familiar fact that the distance from x to y is less than or equal to the sum of the
distances from x to z and z to y. Regarding x, y and z as the vertices of a triangle, (3.3) says that the
length of the edge joining x and y is less than or equal to the sum of the lengths of the edges joining x to
z and z to y; this is the usual triangle inequality.
2
Note that the zero vector does not have a direction.
ANALYSIS III 19
CHAPTER 3. BASIC RESULTS ABOUT RN
Proposition 3.3.
Warning 3.4. It is true that y = x if, and only if, kx − yk = 0. However, kxk = kyk does not imply that
y = ±x. (Think of points on the unit circle.)
Remark 3.5 (k · k satisfies the definition of a norm). A norm is a non-negative valued function
k · k : X → R+ on a real vector space X which satisfies
(iii) The triangle inequality (3.2): for every x, y ∈ X we have kx + yk 6 kxk + kyk.
3.3 Convergence in Rn
Definition 3.6. A sequence (xj ) of vectors in Rn converges to x ∈ Rn if
Proposition 3.7 (Uniqueness of limits). Let (xj ) be a sequence in Rn . If it converges to both x and x̃,
then x = x̃.
Proof. If we assume, by contradiction that x 6= x̃, then ε := 21 kx − x̃k > 0. Since xj converges to x,
∃ N1 ∈ N such that
j > N1 ⇒ kxj − xk < ε. (3.5)
The notation when we consider each of the coordinates of one of the elements xj in the sequence can
get a bit awkward. We will donte the i-th coordinate of xj ∈ Rn by xj,i .
ANALYSIS III 20
CHAPTER 3. BASIC RESULTS ABOUT RN
Proof that componentwise convergence implies convergence. Given ε > 0 and i ∈ {1, . . . , n}, ∃ Ni ∈ N
√
such that j > Ni ⇒ |x0,i − xj,i | < ε/ n. Set N := max{N1 , . . . , Nn }. Then
n
!1/2
X
2
j > N ⇒ kx0 − xj k = (x0,k − xj,k ) < ε,
k=1
i.e., limj→∞ xj = x0 .
Remark 3.9. Proposition 3.8 allows us to reduce questions of convergence of a vector-valued sequence to
the corresponding (more familiar) questions of convergence of a sequence of real numbers.
The norm k · k that we defined in (3.1) corresponds to the standard notion of distance we are used to.
However we could have defined other alternative norms.
Definition 3.10 (Max-norm k · k∞ ). The max-norm, which is denoted by k · k∞ , is defined by
n
!1/p
X
kxkp := |xi |p .
i=1
The Euclidean norm (3.1) corresponds to p = 2, and k · k∞ corresponds to taking the limit as p → ∞.
Exercise 3.2 (Comparison of the Euclidean norm with k · k∞ and k · k1 ). Prove that
√
kxk∞ 6 kxk 6 n kxk∞ (3.9)
and that
√
kxk 6 kxk1 6 n kxk. (3.10)
Furthermore, verify that k · k∞ and k · k1 satisfy the triangle inequality and the relations stated in
Remark 3.5 for a norm; indeed, k · k1 and k · k∞ are actually norms.
This exercise shows that in Definition 3.6 we could have equivalently used k · k1 or k · k∞ instead of
k · k to define the limit of a sequence.
Since Proposition 3.8 reduces convergence to componentwise convergence, we have the following result.
ANALYSIS III 21
CHAPTER 3. BASIC RESULTS ABOUT RN
Exercise 3.3 (Sequence product rules). Let (aj ) be a sequence of real numbers that converges to a and
let (xj ) and (yj ) be sequences of vectors in Rn that converge to x and y respectively. Prove that
Definition 3.13 (Boundedness of a sequence). A sequence (xj ) is bounded if there ∃ M > 0 such that
kxj || 6 M for every j ∈ N.
Proposition 3.14 (Boundedness of a convergent sequence). If (xj ) converges to x, then (xj ) is bounded.
It is possible to prove this proposition by using componentwise convergence and the boundedness of
real sequences. A more direct proof is based on the following lemma.
Lemma 3.15. If (xj ) converges to x then the sequence of real numbers kxj k converges to kxk.
Proof. Given ε > 0, ∃ N ∈ N such that j > N ⇒ kxj − xk < ε. It follows from the reverse triangle
inequality that
for j > N, | kxj k − kxk | 6 kxj − xk < ε.
Proof of boundedness of a convergent sequence. By the lemma, the convergence of (xj ) implies the con-
vergence of kxj k. The boundedness of a convergent sequence of real numbers then implies that kxj k is
bounded and therefore, by definition, (xj ) is bounded.
Remark 3.16. Note that the converse of Lemma (3.15) does not hold, not even when n = 1. (Use a
mathematical software package to plot the sequence ak = (cos k, sin k) in R2 for a demonstration of how
badly the converse of Lemma (3.15) can fail.)
Proposition 3.17 (Completeness of Rn ). Let (xj ) be a Cauchy sequence in Rn , that is, ∀ ε > 0, ∃ N ∈ N
such that j, k > N ⇒ kxj − xk k < ε. Then (xj ) converges to some x ∈ Rn .
Sketch proof. Show that each component xj,i , 1 6 i 6 n, is a Cauchy sequence of real numbers. Then
use the completeness of R and componentwise convergence of xj .
Theorem 3.18 (Bolzano-Weierstrass for a bounded sequence of vectors). A bounded sequence (xj ) in Rn
has a convergent subsequence (xj` ).
Sketch of the proof. The proof of the Bolzano-Weierstrass in Chapter 3 in MA141 is done in R. The
argument below is a complete proof in two dimensions, and with Let xj = (xj,1 , . . . , xj,n ) be a bounded
sequence in Rn . Then xj,1 is a bounded sequence in R and therefore, by the Bolzano-Weierstrass Theorem
it has a convergent subsequence xjk ,1 which converges to x∗1 ∈ R. Since we are only interested in finding
a subsequence, we can consider the following sequence, indexed by k, (xjk ,1 , . . . , xjk ,n ). So far we have
constructed a subsequence of the original for which the first coordinate is a convergent sequence.
ANALYSIS III 22
CHAPTER 3. BASIC RESULTS ABOUT RN
Consider now the sequence xjk ,2 . The sequence is of course bounded and therefore, by Bolzano-
Weierstrass, it has a subsequence xjkl ,2 which converges to x∗2 ∈ R. Notice that since xjk ,1 is convergent,
so is xjkl ,1 . Therefore, if we consider the sequence indexed by l, (xjkl ,1 , . . . , xjkl ,n ), we now have convergent
sequences in the first two components.
It is hopefully clear that, aside from running out letters (and having to resort to cleverer notation), we
can repeat this procedure n times, iteratively constructing subsequences to ensure that every component
is convergent.
3.5 Continuity
3.5.1 Definitions of continuity and continuous limit
We define continuity following the results in year 1 (see Definition 1.2). The only changes are in the
dimension of the domain and the target of the function. We consider U ⊂ Rn , p ∈ U and a function
f : U → Rk .
Notice that the two norms k · k in the definition above correspond to norms in different spaces, namely
Rn and Rk , but we dot make a distinction in the notation.
Hint: The argument is the same as that given in First Year Analysis. That the ε-δ definition implies
the sequential definition is straightforward. The converse proceeds by proving the contrapositive, i.e., one
assumes the failure of the ε-δ definition and then one constructs a sequence xj in U which converges to p
but for which f (xj ) does not converge to f (p).
We say that f is continuous, without specifying a particular point, if it is continuous at all points of its
domain. If we wish to emphasize the domain U on which f is continuous, then we say that f is continuous
on U .
Notation 3.21. The space of functions continuous on U with values in Rk is denoted by C(U, Rk ) or
C 0 (U, Rk ).4 When k = 1, we simply write C(U ) or C 0 (U ).
Just as for limits of sequences, continuous limits are unique. It is also clear that f is continuous at p if,
and only if, limx→p f (x) = f (p). Notice that the definition of continuous limit tacitly assumes that there
exist points in U , different from x, which are arbitrarily close to x.
4
The superscript will later denote the number of derivatives.
ANALYSIS III 23
CHAPTER 3. BASIC RESULTS ABOUT RN
As for question (ii), the example below shows that separate continuity does not imply continuity.
f (x, y) := 1 , if xy 6= 0, f (x, y) := 0 , if xy = 0.
g 0 (x) = 0 for every x ∈ R and h0 (y) = 0 for every y ∈ R. In particular, both g 0 and h0 are continuous
at 0 and therefore, f is separately continuous at (0, 0).
However, f is not continuous at (0, 0) because lim(x,y)→(0,0) f (x, y) does not exist. We can estab-
lish this by finding two sequences (aj , bj ) and (αj , βj ) both of which converge to (0, 0) but for which
limj→∞ f (aj , bj ) 6= limj→∞ f (αj , βj ); we also require (aj , bj ) 6= (0, 0) and (αj , βj ) 6= (0, 0) ∀ j ∈ N.
So take, for example, aj = αj = βj = 1/j and bj = 0. Then f (aj , bj ) = 0 and f (αj , βj ) = 1 ∀ j and
therefore limj→∞ f (aj , bj ) = 0 6= 1 = limj→∞ f (αj , βj ). By uniqueness of limits, if lim(x,y)→(0,0) f (x, y)
were to exist, limj→∞ f (aj , bj ) and limj→∞ f (αj , βj ) would have to have the same value. Since they do
not, we conclude that lim(x,y)→(0,0) f (x, y) does not exist.
The following are easy to verify (left as exercises):
• f is not separately continuous at points (x, 0) such that x 6= 0 (because hx is then not continuous
at 0) and similarly,
Proposition 3.25 (The sum of continuous functions is continuous). If f, g : U → Rk are both continuous
at p then, af + bg is continuous at p.
The proof of this is just an application of the sum rule for limits of sequences of vectors.
ANALYSIS III 24
CHAPTER 3. BASIC RESULTS ABOUT RN
Proposition 3.26 (The product of a continuous scalar (real) valued function with a continuous vector-val-
ued function is continuous). If f : U → R and g : U → Rk are both continuous at p then, f g is continuous
at p where (f g)(x) := f (x) g(x) .
The proof of this is just an application of the product rule for a convergent sequence of real numbers
and a convergent sequence of vectors.
Exercise 3.6. Suppose that f : U → Rk is continuous at p. Prove that if f (p) 6= 0 then there exists δ > 0
such that kf (x)k > 21 kf (p)k ∀ x ∈ U for which kx − pk < δ.
Exercise 3.7. Suppose that f : U → R is continuous at p ∈ U and f (x) 6= 0 ∀ x ∈ U . Prove that 1/f is
continuous at p.
Corollary 3.27. Suppose that f : U → R and g : U → Rk are both continuous at p and that f (x) 6=
0 ∀ x ∈ U . Then g/f is continuous at p.
Proposition 3.28 (The composition of continuous functions is continuous). If U ⊂ Rn , V ⊂ Rk , f : U →
Rk is continuous at p ∈ U, f (U ) ⊂ V, g : V → Rm is continuous at f (p) ∈ V , then g ◦ f : U → Rm is
continuous at p.
The proof of this is just an application of the sequential definition of continuity.
Proposition 3.29 (Componentwise continuity). Recall that f : U → Rk can be written as
3.5.4 Constructing continuous functions of several variables from continuous real valued
functions of a single real variable (NOT EXAMINABLE).
xy
A function like f (x, y) = looks continuous on R2 \ {(0, 0)}, but how do we prove it without
+ y2x2
resorting to the ε-δ definition or the sequential definition of continuity? We know g(x) = x and h(x) = x2
are continuous as functions of the single real variable x. However, what we need to know is that the
functions γ : R2 → R and η : R2 → R defined by
γ(x, y) := x, η(x, y) := x2
are continuous as functions of two variables. That is precisely the content of the next proposition, which
follows from the following easy lemma.
Lemma 3.31. Write Rn+` as Rn ⊕ R` , that is
Rn+` = {(x, y) : x ∈ Rn , y ∈ R` }.
π1 (x, y) := x, π2 (x, y) := y, x ∈ Rn , y ∈ R` .
ANALYSIS III 25
CHAPTER 3. BASIC RESULTS ABOUT RN
Proof. By Lemma 3.31, πi is continuous on Rn and therefore, by the continuity of composition of contin-
uous functions, f = g ◦ πi is continuous on πi−1 {a}.
xy
We can use Proposition 3.32 and the results in section 3.5.3 to prove the continuity of f (x, y) =
x2 + y2
on R2 \ {(0, 0)} as follows. Consider the four functions, each defined on R2 by
Proposition 3.32 tells us that the continuity of these four functions follows from the continuity (proved in
First Year Analysis) of g(t) = t and h(t) = t2 as functions of the single real variable t. Now
and therefore, the continuity of f on R2 \ {(0, 0)} follows from the continuity of the product, sum and
quotient of continuous functions at points where the denominator does not vanish.
A similar approach can be followed for most functions given by explicit formulas. However, the continuity
of a function at points where the function is given special values (not by a formula) has to be investigated
by separate arguments.
The following two examples are intended to clarify what is meant by ‘natural domain of definition’ of
a function defined by an expression involving familiar continuous functions. The natural domain of
x2 sin(y)
F (x, y) =
ex − cosh y
log(x + y)
q
f (x, y, z) := , arccos(y) 1 + (cos(xez ))2
sin z
ANALYSIS III 26
CHAPTER 3. BASIC RESULTS ABOUT RN
(0, 0). For example, we can proceed along (x, x2 ), i.e. along the parabola y = x2 . Indeed, we can approach
(0, 0) along the graph of any continuous function ψ(x) for which ψ(0) = 0. This still does not exhaust all
possibilities because, for instance, we may approach (0, 0) along a spiral like (t cos(1/t), t sin(1/t), t > 0.
So, by Proposition 3.28, if f : R2 → R is continuous at (0, 0) then limt→0 f (ϕ(t), ψ(t)) would have to
exist for any pair of functions ϕ, ψ : R → R that are continuous at 0 and equal to 0 there. This should
make it clear that continuity is much more restrictive than separate continuity and indeed, more restrictive
than continuity along lines, which we shall now define.
Definition 3.33 (continuity along lines, also called linear continuity). A function f : Rn → Rk is continuous
along lines (also referred to as linearly continuous) at x0 if the restriction f L of f to the line L passing
through x0 is continuous for every such line L.
that is, limt→0 f (x0 + tv) is independent of v. We have seen above that continuity implies continuity along
lines.
In the next example we will exhibit a function which is separately continuous at all points of R2
but which is not continuous along lines through (0, 0).
Therefore, lim(x,y)→(0,0) f (x, y) depends on the line in R2 along which we approach (0, 0). In particular, it
is not possible to assign any value to f at (0, 0) that would make it continuous along lines through (0, 0).
Remark 3.35. (This is not examinable) One may be tempted to think that a separately continuous
function fails to be continuous only at isolated points. This is not the case. For a fixed (a, b) ∈ R2
define f(a,b) : R2 → R by f(a,b) (x, y) := f (x − a, y − b) where f is as in Example 3.34. Let (an , bn ) be
an enumeration of Q × Q, i.e., of all points in R2 both of whose coordinates are rational. Then define
F : R2 → R by
∞
2−n f(an ,bn ) (x, y).
X
F (x, y) :=
n=0
It can be shown that F is separately continuous, but discontinuous precisely on Q × Q. Conceptually, the
sum in the definition of F disperses the discontinuity of f to all the rational points in R2 .
The next example exhibits a function f : R2 → R which is continuous along lines through (0, 0)
but which is not continuous at (0, 0).
Example 3.36. Define f : R2 → R by f (x, y) = 1 if 0 < y < x2 and f (x, y) = 0 otherwise. Show that
limt→0 f (tv) = 0 = f (0, 0) ∀ v ∈ R2 . However, show also that f is discontinuous at (0, 0).
ANALYSIS III 27
CHAPTER 3. BASIC RESULTS ABOUT RN
In the diagram below, the grey shaded region E is the set on which f = 0, i.e.,
Given v ∈ R2 , ∃ τ > 0 such that, |t| < τ ⇒ tv ∈ E. (If v = (a, b), take τ = b/a2 if ab = 6 0
and τ = +∞ if ab = 0.) In other words, f (tv) = 0 ∀ t ∈ (−τ, τ ). It follows that limt→0 f (tv) = 0 =
f (0, 0) ∀ v ∈ R2 , as claimed.
Finally, limx→0 f (x, 12 x2 ) = 1 6= 0 = f (0, 0) which shows that f is discontinuous at (0, 0).
In the next example, we consider the following question: Suppose we demand that f : R2 → R be
continuous along any line in R2 and not just the ones that pass through a chosen point. Would f then
have to be continuous? Remarkably, this is still not the case, as demonstrated by the following example.
x2 y
f (x, y) = , if (x, y) 6= (0, 0), f (0, 0) := 0.
x4 + y 2
For each k ∈ R, define g k : R → R to be the restriction of f to the line y = kx through the origin in
R2 ,i.e.,
kx
g k (x) := f (x, kx) = 2 ∀ x ∈ R.
x + k2
Also define g ∞ : R → R to be the restriction of f to the y-axis, i.e.,
g ∞ (x) := f (0, y) = 0 ∀ y ∈ R.
x2 x2 1
ϕ(x) := f (x, x2 ) = 4 2 2
= , if x 6= 0, ϕ(0) = f (0, 0) = 0.
x + (x ) 2
Thus ϕ is not continuous. It follows from Propositon 3.28 that, since g(x) := (x, x2 ) is continuous, f
cannot be continuous at (0, 0) because, if it were, then ϕ = f ◦ g would also have to be continuous, which
it is not. Indeed, we have also shown that lim(x,y)→(0,0) f (x, y) does not exist. In particular, the function
f is continuous away from (0, 0) (we will not show this). It is also continuous along lines at (0, 0), but it
is not continuous at (0, 0).
ANALYSIS III 28
Chapter 4
Remark 4.4. Most textbooks first define an open set and then define a closed set to be the complement
of an open set. Of course, these textbooks then have to prove that a closed set satisfies Definition 4.1.
The definition of open set motivates the following definition.
Definition 4.5 (Open (Euclidean) ball). The open ball of radius r > 0 centred at a ∈ Rn is denoted by
B(a, r) or Br (a) and is defined by
29
CHAPTER 4. RUDIMENTS OF TOPOLOGY OF RN AND CONTINUITY
This is the definition of an open set that is given in most textbooks. The following proposition justifies the
use of the adjective ‘open’ in the definition of an open ball.
Proposition 4.6. An open ball is open, i.e., it satisfies the definition of an open set.
Proof. For each y ∈ B(a, r) we need to find ρy > 0 so that the open ball B(y, ρy ) ⊂ B(a, r).
To this end, set ρy = r − ky − ak. (This value of ρy is suggested by a picture which you should draw.)
Then, since ky − ak < r, we have ρy > 0 and, for x ∈ B(y, ρy ),
kx − ak 6 kx − yk + ky − ak < ρy + ky − ak = r,
Definition 4.7 (Closed ball). The closed ball of radius r > 0 centred at a ∈ Rn is denoted by B(a, r) or
Br (a) and is defined by
Br (a) ≡ B(a, r) := {x ∈ Rn : kx − ak 6 r}.
We abbreviate Br (0) to Br and B1 to just B.
The following proposition justifies the use of the adjective ‘closed’ in the definition of a closed ball.
Proposition 4.8. A closed ball is closed, i.e., it satisfies the definition of a closed set.
Sketch proof. This proposition can be proved in at least two ways. One way is to prove that the complement
of a closed ball is open. Another way is to prove that, if xj is a sequence in B(a, r) which converges to x,
then kx − ak 6 r, i.e., x ∈ B(a, r).
Proposition 4.9 (An arbitrary union of open sets is open). If Uλ is open for all λ ∈ Λ, where Λ is an
S
indexing set (which could be uncountable), then λ∈Λ Uλ is open.
Proof. If p ∈ λ∈Λ Uλ then ∃ λ∗ ∈ Λ such that p ∈ Uλ∗ . But Uλ∗ is open and therefore, ∃ ε > 0 such
S
Definition 4.10 (ε-neighbourhood). Let E be any subset of Rn . Given ε > 0, the ε-neighbourhood
N (E, ε) of E is defined by
N (E, ε) :=
[
B(x, ε).
x∈E
j=1
Then O is open. The sum of the lengths of the intervals that make up O is ε ∞ 1−j = 2ε. Therefore,
P
j=1 2
if ε < 21 , O cannot contain all the irrationals between zero and 1. This example shows how complicated
open sets can be.
Proposition 4.12 (The finite intersection of open sets is open).
If U1 , U2 , . . . , Um are all open, then m
T
j=1 Uj is also open.
Proof. If p ∈ then
Tm
j=1 Uj
ANALYSIS III 30
CHAPTER 4. RUDIMENTS OF TOPOLOGY OF RN AND CONTINUITY
Corollary 4.13. An arbitrary intersection of closed sets is closed and the finite union of closed sets is
closed.
Sketch proof. Consider the complements of the relevant closed sets and apply the preceding propositions
together with de Morgan’s laws on complements, unions and intersections.
Remark 4.14. Note that a subset of Rn may be neither open, nor closed. For example [0, 1) in R or
Equivalently,
∀ ε > 0 ∃ δ > 0 such that B(p, δ) ∩ U ⊂ f −1 B(f (p), ε) . (4.2)
Informally and pictorially, f is continuous at p if f (x) can be guaranteed to stay near f (p) (ε-near) by
requiring x to stay sufficiently close (δ-close) to p in U .
Theorem 4.15 (Continuity via open sets and closed sets). The following statements are equivalent.
Proof. Suppose that f is continuous at all points of Rn and let V ⊂ Rk be open. Then, for each
p ∈ f −1 (V ), ∃ ε > 0 such that B(f (p), ε) ⊂ V . By continuity of f at p as stated in (4.2), ∃ δ(p) > 0
such that B(p, δ(p)) ⊂ f −1 B(f (p), ε) ⊂ f −1 (V ), which shows that f −1 (V ) is open. (We have assumed
Remark 4.16. If f : Rn → Rk is continuous, then the image of an open set need not be open. For instance,
consider the constant map f (x) = 0 ∀ x ∈ Rn .
If f : Rn → Rk is continuous, then the image of a closed set need not be closed. For instance, consider
x2
the map f : R → R given by f (x) = 2 ∀ x ∈ R. Then f (R) = [0, 1) which is neither a closed subset
x +1
nor an open subset of R.
Example 4.17 (Open and closed sets via Theorem 4.15). Show that
ANALYSIS III 31
CHAPTER 4. RUDIMENTS OF TOPOLOGY OF RN AND CONTINUITY
(ii) The set E := {(x, y) ∈ R2 : xy sin(1/x) cos(1/y) > −1} is an open subset of R2 .
Solution.
(i) Define f ∈ C(Rn ) by f (x) = kxk. Then S n−1 = f −1 ({1}) which, by part (iii) of Theorem 4.15, is a
closed subset of Rn because {1} is a closed subset of R.
(ii) Define f : R2 → R by
cos(1/y) if xy 6= 0, f (x, y) = 0 if xy = 0.
f (x, y) := xy sin(1/x)
Then f is continuous and E = f −1 (−1, ∞) which, by part (ii) of Theorem 4.15, is an open subset of
Remark 4.18. Close inspection of the proof of Theorem 4.15 will reveal that if U ⊂ Rn is open then
f : U → Rk is continuous at all points of Rn if, and only if, for all open subsets V of Rk , f −1 (V ) is open.
However, it is no longer true that the preimage of a closed set is necessarily closed, that is, statement (iii)
of Theorem 4.15 no longer applies.
Similarly, if U ⊂ Rn is closed then f : U → Rk is continuous at all points of Rn if, and only if, for all
closed subsets F of Rk , f −1 (F ) is closed. However, it is no longer true that the preimage of an open set
is necessarily open, that is, statement (ii) of Theorem 4.15 no longer applies.
The extension of Theorem 4.15 to functions f : U → Rk , where U is an arbitrary subset of Rn , requires
the notion of sets that are open/closed relative to U . We will not explore these notions in this module.
Theorem 4.21. K ⊂ Rn is sequentially compact if, and only if, K is closed and bounded.
Proof. Suppose that K is sequentially compact. To prove that K is closed, we consider a sequence xj in
K which converges to x ∈ Rn and then we have to show that x ∈ K. By the sequential compactness of
K, xj has a subsequence xj` whose limit is in K. But x = limj→∞ xj = lim`→∞ xj` ∈ K. The proof
that K is closed is complete.
To prove that K is bounded, assume, for a contradiction, that it is unbounded. Then there exists a
sequence xj in K such that kxj k > j ∀ j ∈ N. By the sequential compactness of K, xj has a subsequence
xj` whose limit is in K. In particular, xj` is bounded, i.e., ∃ M > 0 such that kxj` k 6 M ∀ ` ∈ N. But
by definition of subsequence, j` > ` and, by the way the sequence xj was chosen, kxj` k > j` . Therefore,
Theorem 4.21 is important because it enables us to determine easily whether a set is sequentially
compact. For instance, the theorem asserts that a closed ball B(a, r) is sequentially compact without
having to check whether all its sequences contain a convergent subsequence! Similarly, we can assert that
the sphere Sn−1 (a, r) := {x ∈ Rn : kx − ak = r} is sequentially compact; it is clearly bounded and we
showed above (for a = 0 and r = 1, but the proof is virtually identical) that it is closed.
ANALYSIS III 32
CHAPTER 4. RUDIMENTS OF TOPOLOGY OF RN AND CONTINUITY
Proof. Let (yj ) be a sequence in f (K). Then, for each j ∈ N, ∃ xj ∈ K such that f (xj ) = yj . By the
sequential compactness of K, there exists a convergent subsequence (xj` ) of (xj ) such that lim`→∞ xj` =
x ∈ K. By continuity of f at x, lim`→∞ yj` = lim`→∞ f (xj` ) = f (x) ∈ f (K), i.e., f (K) is sequentially
compact.
Theorem 4.23 (Extreme Value Theorem). Let K ⊂ Rn be sequentially compact and let f : K → R be
continuous. Then ∃ x∗ , x∗ ∈ K such that
This theorem asserts that a continuous real valued function on a sequentially compact space attains
its extreme values, i.e., max and min. This theorem was proved in First Year Analysis in the case that K is
a closed, bounded interval. It is one of the most important theorems of elementary mathematical analysis
because, for instance, it is used in the proof of Rolle’s Theorem which, in turn, is used in the proof of
Taylor’s theorem.
Proof of Extreme Value Theorem. By the previous theorem and Theorem 4.21, f (K) ⊂ R must be closed
and bounded. Therefore, M := sup f (K) and m := inf f (K) are both finite because f (K) is bounded. By
definition of sup and inf, there exist sequences aj , bj ∈ f (K) such that limj→∞ aj = m and limj→∞ bj =
M . But f (K) is closed and therefore, m, M ∈ f (K), i.e., ∃ x∗ , x∗ ∈ K such that
Remark 4.24. The notion of supremum cannot be extended from R to Rk , k > 2. That is why we had
to restrict ourselves to scalar functions in the preceding theorem. The best we can do for vector-valued
functions is stated in the following corollary.
Proof. Let us note that the function g : K → R given by g(x) := kf (x)k is continuous. In order to see
this, we use the triangle inequality to obtain that for all x, y ∈ K, we have
More precisely, given x ∈ K and ε > 0, by continuity of f at x, it follows that there exists δ > 0 such
that kf (x) − f (y)k < ε for all y ∈ K with kx − yk < δ. Substituting this into (4.3), we obtain that
|g(x) − g(y)| < ε for all such y. The result now follows from Theorem 4.23.
ANALYSIS III 33
Chapter 5
Notation 5.1.
(i) The space of linear maps, i.e., {A : Rn → Rk | A is linear}, shall be denoted by L (Rn , Rk ) and
L (Rn , Rn ) shall be abbreviated to L (Rn ).1
(ii) The space of k × n matrices with real entries shall be denoted by Rk,n .2
With a matrix
a11 . . . a1n
.. .. ∈ Rk,n
(aij ) = . .
ak1 . . . akn
we associate (subconsciously?!) A ∈ L (Rn , Rk ) defined by
x1 a11 . . . a1n x1
.. .. .
.. ...
n
R 3 x = . 7→ Ax := . k
∈R . (5.1)
xn ak1 . . . akn xn
Let {v1 , . . . , vn } and {w1 , . . . , wk } be the standard bases of Rn and Rk respectively, i.e.,
vj = (0, . . . , 0, 1, 0, . . . , 0) ∈ Rn , wi = (0, . . . , 0, 1, 0, . . . , 0) ∈ Rk .
↑ ↑
j th position among n entries ith position among k entries
Then,
a1j
k
.. X
Avj = . = aij wi , j ∈ {1, . . . , n}, (5.2)
i=1
akj
and therefore, (aij ) is the matrix representation of A with respect to the standard bases on Rn and Rk .
On a few occasions it shall be useful to express this association of (aij ) with A as defined above more
formally as a map µ : L (Rn , Rk ) → Rk,n , i.e.,
It is easy to verify that µ is a linear isomorphism. Moreover, since we shall be using standard bases on Rn
and Rk throughout (unless otherwise explicitly stated), we shall switch between the linear map A and the
associated matrix µ(A) = (aij ) without warning.
1
Other notations in use are HomR (Rn , Rk ) and End(Rn ).
2
Other notations in use are Rk×n , M (k × n, R), Mk×n (R) and Mkn (R). M (n, R) is sometimes used as an abbreviation
of M (n × n, R).
34
CHAPTER 5. THE SPACE OF LINEAR MAPS AND MATRICES
This is fine, but we shall make more use of the operator norm (defined below) as it turns out to be more
convenient.3
The operator norm arises from studying how large kAxk can get relative to kxk as x ranges over Rn .
We can do this using (5.1) and the Cauchy-Schwarz inequality:
X n
k X 2 k
X n
X n
X X n
k X
kAxk2 = aij xj 6 a2ij x2j = a2ij kxk2 = k(aij )k2F kxk2 .
i=1 j=1 i=1 j=1 j=1 i=1 j=1
ANALYSIS III 35
CHAPTER 5. THE SPACE OF LINEAR MAPS AND MATRICES
Observe that the supremum in (5.9) is being taken over a sequentially compact set (namely, the unit
sphere S n−1 in Rn ). We shall see in Proposition 5.5 below that this can be an advantage of (5.9) over
(5.7).
k(A + B)xk
|||A + B||| = sup 6 |||A||| + |||B|||.
x∈Rn \{0} kxk
Proposition 5.5. A ∈ L (Rn , Rk ) is injective if, and only if, ∃ α > 0 such that kAxk > αkxk ∀ x ∈ Rn .
(Note that k does not have to be equal to n.)
Proof. If Ax = 0 and kAxk > αkxk for some α > 0 then x = 0, i.e., A is injective.
The converse is proved by establishing the contrapositive, i.e., suppose that there is a sequence xj
in Rn \ {0} such that kAxj k/kxj k → 0 as j → ∞. Set uj := xj /kxj k. Then kuj k = 1 ∀ j ∈ N and
Auj → 0 as j → ∞. Since S n−1 is sequentially compact, there exists a subsequence uj` which converges
to u ∈ S n−1 . Let us note that the map x 7→ kAxk is continuous. Namely, by the triangle inequality and
linearity, we have for all x, y ∈ Rn that
kAxk − kAyk ≤ kAx − Ayk = kA(x − y)k ≤ |||A||| kx − yk ,
from where we deduce the continuity of x 7→ kAxk. Therefore, kAuk = limj→∞ kAuj k = 0, i.e.,
u ∈ ker(A). It follows that A is not injective.
ANALYSIS III 36
CHAPTER 5. THE SPACE OF LINEAR MAPS AND MATRICES
Remark 5.6. Proposition 5.5 can be regarded as a quantitative measure of injectivity. An injective
linear map has to keep a nonzero vector x away from zero. The larger the value of α in the inequality
kAxk > αkxk, the more the linear map A pushes x away from zero.
A better way of saying this is to consider a perturbation of A by a matrix B to get the matrix A + B.
We then have the following
Proof.
k(A + B)xk > kAxk − kBxk > αkxk − |||B|||kxk = δkxk,
where δ := α − |||B||| > 0. Therefore (A + B)x = 0 ⇒ x = 0, which proves that A + B is injective.
This proposition can be interpreted as saying that if (5.13) holds then the open ball B(A, α) ⊂
L (Rn , Rk )4 is contained in the set of injective linear transformations in L (Rn , Rk ).
So, a larger value of α in (5.13) indicates that A is able to withstand perturbations by ‘larger’ (as
measured by the operator norm) linear transformations while maintaining injectivity.
x 7→ ... .. : U → Rk,n
f
.
ak1 (x) . . . akn (x)
is continuous at x if, and only if, ∀ i ∈ {1, . . . , k}, j ∈ {1, . . . , n}, x 7→ aij (x) is continuous at x.
A function F : U → L (Rn , Rk ) is continuous at x ∈ U if ∀ ε > 0, ∃ δ > 0 such that ky − xk < δ ⇒
|||F (y) − F (x)||| < ε.
4
This ball is taken with respect to the operator norm on L (Rn , Rk ); see §5.2.
5
A space X is complete if every Cauchy sequence in X converges to an element of X.
ANALYSIS III 37
CHAPTER 5. THE SPACE OF LINEAR MAPS AND MATRICES
Remark 5.8. Because of (5.12), we see that F : U → L (Rn , Rk ) is continuous at x ∈ U if, and only if,
µ(F ) : U → Rk,n is continuous at x ∈ U .
This remark is very useful because it provides a practical way of checking the continuity of
F : U → L (Rn , Rk ). Namely, we simply have to check whether all the matrix entries of the
matrix representation µ(F ) (with respect to the standard bases on Rn and Rk ) are continuous.
The continuity of f : Rk,n → R` and of F : Rk,n → R`,m is defined similarly by identifying (Rk,n , k · kF )
with (Rnk , k · k), as in the next proposition and the example below it.
Proposition 5.9 (Continuity of the determinant function). The map ∆ : Rn,n → R defined by ∆(aij ) :=
det(aij ) is continuous with respect to the norm k · kF on Rn,n .
6
See Proposition 3.32 and use the algebraic properties of continuous functions. Each term of ∆(aij ) is, in fact, linear
in each of the n2 variables (5.14)and this is what makes ∆ an example of a multilinear map. The special case of the
determinant of a 2 × 2 matrix ac db may help to clarify matters. This determinant is then just the function on R4 defined by
∆(a, b, c, d) := ad − bc, which is easily seen to be continuous by Proposition 3.32 and the algebraic properties of continuous
functions.
ANALYSIS III 38
Chapter 6
The Derivative
From now on, the domain U ⊂ Rn of a function f : U → Rk shall be an open subset of Rn , unless
otherwise stated. In particular, this means that when p ∈ U and a limit like limx→p is considered, x is
allowed to approach p from any direction.
Since gx,v (t) = (g1 (t), . . . , gk (t)) is a function of a single real real variable, we can differentiate it component
by component in the usual way.1
d
∂v f (x) := gx,v (t)
dt t=0
d
= f (x + tv) (6.1)
dt t=0
f (x + tv) − f (x)
= lim . (6.2)
t→0 t
Example 6.2. Calculate ∂v f (x, y) for the function f (x, y) := x2 − y 2 in the direction of v = (a, b).
d
f (x + tv) = 2ax + 2ta2 − 2by − 2tb2 ,
dt
d
∂v f (x, y) = f (x + tv) = 2ax − 2by.
dt t=0
1
In the definition 3.33 of linear continuity, gx,v was denoted by f L .
39
CHAPTER 6. THE DERIVATIVE
ANALYSIS III 40
CHAPTER 6. THE DERIVATIVE
Exercise 6.1. Show that the linear map A in (6.5), if it exists, is unique. This justifies saying that the
Fréchet derivative provides the optimal linear approximation of f .
Remark 6.5. A real number a can also be viewed as the linear map A : R → R defined by Ah = ah.
Indeed a is the 1 × 1 matrix representation of A with respect to the standard basis 1 of R. Similarly, the
real number f 0 (x) in (6.3) is the 1 × 1 matrix representation of the linear map h 7→ f 0 (x) h in (6.4).
Notation 6.6. If a linear map A that satisfies (6.5) exists, it is called the (Fréchet) derivative of f at x
and it is denoted by Df (x).
Exercise 6.1 justifies calling the affine linear map h 7→ f (x)+Df (x)h the best affine linear approximation
of the map h 7→ f (x + h). The ε-δ formulation of (6.6) provides us with a way of quantifying how good
this approximation is. Namely, by definition of continuous limit we have that
We shall refer to (6.7) and (6.6) also as the definition of the derivative Df (x) of f at x. Note that equality
has to be allowed in (6.7) because we have allowed the possibility that h = 0, which can be convenient in
many situations.
Set δ∗ := min{δ, ε/(|||Df (x)||| + ε)}. Then khk < δ∗ ⇒ khk < δ and therefore we can use the above
chain of implications to conclude that
Remark 6.8. Compare Exercise 6.2 with Proposition 3.29 on componentwise continuity.
ANALYSIS III 41
CHAPTER 6. THE DERIVATIVE
Proof. If v = 0 there is nothing to prove. So, we assume that v 6= 0. Then, replacing h in (6.6) by tv and
removing k · k where that is allowed, we get
Multiply both sides of (6.9) by kvk and use Df (x)(tv) = tDf (x)v by the linearity of Df (x) so as to get
f (x + tv) − f (x)
lim = Df (x)v, i.e.,∂v f (x) = Df (x)v.
t→0 t
Finally, by linearity of Df (x),
∂av+bw f (x) = Df (x)(av + bw) = aDf (x)v + bDf (x)w = a ∂v f (x) + b ∂w f (x).
Example 6.3 shows that the converse of Proposition 6.9 does not hold.
Remark 6.10. We give another proof of Proposition 6.9 after we have proved the chain rule; see §6.6.3.
vi = (0, . . . , 0, 1, 0, . . . , 0) ∈ Rn .
↑
ith position among n entries
Definition 6.11. For 1 6 i 6 n, ∂vi f (x) is called the ith -partial derivative of f : U → Rk at x ∈ U . It is
more simply denoted by ∂i f (x).
Since
f (x + tvi ) − f (x)
∂vi f (x) = lim
t→0 t
f (x1 , . . . , xi−1 , xi + t, xi+1 , . . . xn ) − f (x1 , . . . , xi−1 , xi , xi+1 , . . . xn )
= lim
t→0 t
∂i f (x) is calculated by differentiating f (x1 , . . . , xn ) with respect to the ith variable, treating all the other
∂f
variables as constant2 . It is therefore also common to write ∂x i
(x) or ∂x
∂
i
f (x1 , . . . , xn ) instead of ∂i f (x).
Bearing in mind that f (x) = (f1 (x), . . . , fk (x)) we have
If f is a function of a few variables, say two, it is common to write f (x, y) instead of f (x1 , x2 ), and
to write fx instead of ∂f ∂f
∂x or ∂1 f . Similarly, fy is shorthand for ∂y . Similar shorthand applies to functions
of 3 or 4 variables. For 5 variables or more, it is usually more convenient to number the variables, rather
than choose distinct letters!
2
See Example 6.17 further down.
ANALYSIS III 42
CHAPTER 6. THE DERIVATIVE
if f, g : U → Rk , then ∂i (f + g) = ∂i f + ∂i g.
Similarly,
if f : U → R and g : U → Rk then ∂i (f g) = (∂i f )g + f ∂i g.
where ∂1 f (x), . . . , ∂n f (x) are the vector-valued partial derivatives of the vector-valued function f (the
values of f are vectors in Rk ) and ∂f1 , . . . , ∂fk are the Jacobian 1 × n matrices (row vectors) of the
scalar-valued functions f1 , . . . , fk .
Definition 6.13. The gradient at x, ∇f (x), of a scalar valued function f : U → R is defined to be the
column vector
∂1 f (x)
∇f (x) := ... .
∂n f (x)
T
Thus ∇f (x) is the vector in Rn which is the transpose of the row vector ∂f (x), ∇f (x) = ∂f (x) .
Remark 6.14. For a scalar valued function f : U → R, ∂f (x) represents a linear functional on Rn defined
by
Rn 3 h 7→ ∂f (x) (h) := ∂1 f (x) h1 + · · · + ∂n f (x) hn ∈ R,
h = (h1 , . . . , hn ).
Using the Euclidean inner product, this linear functional ∂f (x) is identified with the vector ∇f (x):
∂f (x) (h) = ∇f (x) · h.
However, be warned that the distinction between ∇f and ∂f is often suppressed, even in these notes!
Proposition 6.15. If f : U → Rk is differentiable at x ∈ U and h ∈ Rn then
Remark 6.16. It is important to appreciate the difference between the two sides of (6.10). On the left hand
side we have the linear map Df (x) acting on the vector h whereas on the right hand side we have the matrix
∂f (x) multiplying the vector h. In other words, ∂f (x) is the matrix representation of Df (x) with respect
to the standard bases on Rn and Rk . More formally, ∂f (x) = µ(Df (x)) where µ : L (Rn , Rk ) → Rk,n is
defined by (5.3).
Proof of Proposition 6.15. h = h1 v1 + · · · + hn vn and therefore, by linearity of Df (x),
n
X n
X
Df (x)h = hi Df (x)vi = hi ∂i f (x) = ∂f (x)h,
i=1 i=1
ANALYSIS III 43
CHAPTER 6. THE DERIVATIVE
All the entries of the Jacobian matrix ∂f are continuous functions and we shall see in §6.7 that this implies
the existence of the derivative Df (x, y, z) ∈ L (R3 , R2 ), which is the linear map defined by
3 +2y 3 +2y
3x2 ex 2ex r r
0
Df (x, y, z)(r, s, t) = s = ∂f (x, y, z) · s .
4
yz sin x 2 3
2y z sin x
√ cos x − (1+y 2 z 4 )3/2 − (1+y 2 z 4 )3/2
1+y 2 z 4 t t
So why bother with three different ways of writing the same thing?
The reason is that even when Df (x) does not exist, ∂h f (x) and ∂f (x)h may both still exist but
it may happen that they are not equal! (See Example 6.19 below.) In other words, the existence
of the Jacobian matrix ∂f (x) does not guarantee that the linear map it defines is the derivative
Df (x), unless Df (x) is known to exist.
The reader would be right to wonder at this stage how to go about calculating the derivative
Df if this cannot be simply done by computing the Jacobian matrix ∂f . Fortunately, as has
already been pointed out in Example 6.17, it suffices to verify further that all the entries of ∂f
are continuous for then, by Theorem 6.30, Df exists and its matrix representation with respect
to the standard bases of Rn and Rk is given by ∂f .
However, there are situations where it may be necessary to to calculate Df directly from the definition
(6.6). This would be the case, for example, at points outside the natural domain of definition of f where it
is assigned a special value. There are also situations where it is more convenient, to calculate ∂v f directly
from the definition (6.1), as in §6.5.1.
Remark 6.18. The notions of partial derivative, directional derivative and (Fréchet) derivative are the
differentiable analogues of separate continuity, continuity along lines and continuity.
x3
f (x, y) = if (x, y) 6= (0, 0), f (0, 0) = 0.
x2 + y 2
(i) Show that ∂v f (0, 0) exists for all v ∈ R2 . In particular, calculate ∂f (0, 0).
(ii) Show that ∂v f (0, 0) 6= ∂f (0, 0) v and that ∂v f (0, 0) is not linear in v.
(iii) Calculate fx (x, y) and fy (x, y) for (x, y) 6= (0, 0) and show that fx and fy are not continuous at
(0, 0).
ANALYSIS III 44
CHAPTER 6. THE DERIVATIVE
ta3
(i) As always, ∂(0,0) f (0, 0) = 0. If (a, b) 6= (0, 0) we have f (ta, tb) = ∀ t ∈ R and therefore,
a2 + b2
d d ta3 a3
∂v f (0, 0) = f (ta, tb) = = .
dt t=0 dt a2 + b2 t=0
a2 + b2
the Jacobian matrix ∂f (0, 0) = (∂(1,0) f (0, 0) , ∂(0,1) f (0, 0)) = (1, 0).
a3
(ii) ∂f (0, 0) (a, b) = a 6= ∂(a,b) f (0, 0), unless b = 0. ∂(a,b) f (0, 0) is not linear in (a, b) because a2 +b2
is
not a linear function of a and b.
(iii)
x2 (x2 + 3y 2 ) −2x3 y
fx (x, y) = , fy (x, y) = .
(x2 + y 2 )2 (x2 + y 2 )2
We have seen that fx (0, 0) = ∂(1,0) f (0, 0) = 1 but limy→0 fx (0, y) = 0. Therefore, fx is not
continuous at (0, 0). Similarly fy (0, 0) = ∂(0,1) f (0, 0) = 0 but limx→0 fy (x, x) = − 12 . Therefore, fy
is also not continuous at (0, 0).
(iv) ∂v f (0, 0) is not linear in v and therefore, by Proposition 6.9, Df (0, 0) does not exist. The situation is
similar to that in Example 3.37 of a real valued function of two variables which fails to be continuous
even though its restriction to any line in R2 is continuous. We shall see below that the lack of
differentiability of f at (0, 0) can be understood geometrically as the failure of the graph of f (which
lies in R3 ) to have a tangent plane at (0, 0, 0).
r (t) = ( dt , . . . , dt ) 6= 0 ∀ t ∈ [a, b], i.e., the parameterisation r is regular. Using the rate of change
0 dx1 dxk
definition of derivative given by (6.3), we can then interpret r0 (t) as the vector tangent to C at r(t).3 The
line Lr(t) tangent to C at r(t) is parameterised by
But r0 (t) = ∂r(t) and therefore, the affine linear approximation of h 7→ r(t+h) by h 7→ r(t)+∂r(t)h = `(h)
is a parameterisation of the tangent line Lr(t) . In other words, the affine linear approximation of h 7→ r(t+h)
by h 7→ r(t) + ∂r(t)h for small h corresponds to the geometric approximation of C by Lr(t) near r(t).
In the special case that C is itself a line, then Lr(t) is the same as C. This is the geometric manifestation
of the fact that, as discussed in Example 6.20, the best affine linear approximation of an affine linear map
is itself!
3
We can also view r(t) as the position of a particle at time t and then r0 (t), also denoted ṙ(t), is the velocity of the particle.
ANALYSIS III 45
CHAPTER 6. THE DERIVATIVE
we see that ∂r is of rank 2 if, and only if, ru and rv are linearly independent.4 As in the preceding discussion
for a curve C, the affine linear approximation of (h, k) 7→ r(u + h, v + k) by
is then a parameterisation of the plane Tr(u,v) S tangent to S at r(u, v). Once again, the affine linear
approximation of (h, k) 7→ r(u + h, v + k) for small h and k corresponds to the geometric approximation
of S by Tr(u,v) S near r(u, v).
Thus we see that f is not differentiable at (x0 , y0 ) ∈ pU if, and only if, Gf does not have a tangent plane at
(x0 , y0 , f (x0 , y0 )). For example, (x, y) → |(x, y)| = x2 + y 2 is not differentiable at (0, 0) because none
of its partial derivatives exist at 0. We see this geometrically by noting that the graph of (x, y) → |(x, y)|
on R2 is a circular cone about the z-axis with an apex at the origin where the cone does not have a tangent
plane.
ANALYSIS III 46
CHAPTER 6. THE DERIVATIVE
ANALYSIS III 47
CHAPTER 6. THE DERIVATIVE
Solution. As in Example 6.20, we consider f (A + H) − f (A) and look for the term linear in H:
f (A + H) − f (A) = (A + H)(A + H) − A2 = AH + HA + H 2 .
The term linear in H is AH + HA and so, we define a linear map ΛA : L (Rn ) → L (Rn ) by ΛA (H) :=
AH + HA. Then, f (A + H) − f (A) − ΛA (H) = H 2 and therefore,
Remark 6.22. We can also calculate the directional derivative ∂H f (A) directly from its definition:
d
∂H f (A) = f (A + tH)
dt t=0
d 2
= (A + tAH + tHA + t2 H 2 )
dt t=0
2
= (AH + HA + 2tH )
t=0
= AH + HA.
This is in agreement with Proposition 6.9 according to which (Df (A))(H) = ∂H f (A).
Indeed, since the entries of f (A) are quadratic expressions in the entries of A5 then we know that the
partial derivatives of f with respect to the variables of the entries of A are continuous and therefore, as we
shall see in §4.7, f is differentiable and the calculation of (Df (A))(H) can be reduced to that of ∂H f (A),
which is often much simpler.
6.6 The Chain Rule (NOT COVERED IN CLASS AND NOT EXAM-
INABLE)
Theorem 6.23. Let U and V be open subsets of Rn and Rk respectively. Suppose that f : U → Rk is
differentiable at x ∈ U and that f (x) ∈ V . Suppose further that g : V → Rm is differentiable at f (x).
Then g ◦f : Rn → Rm is differentiable at x and
The following two lemmas will be useful in the proof of the Chain Rule.
Lemma 6.24. Given f : U → Rk , x ∈ U, r > 0 such that B(x, r) ⊂ U and A ∈ L (Rn , Rk ), define
∆x,A f : B(0, r) → Rk by
f (x+h)−f (x)−Ah , if h 6= 0,
∆x,A f (h) = khk
(6.12)
0, if h = 0.
Then f is differentiable at x with Df (x) = A if, and only if, ∆x,A f is continuous at 0.
Proof. If ∆x,A f is continuous at 0 then limh→0 k∆x,A f (h)k = k limh→0 ∆x,A f (h)k = k∆x,A f (0)k = 0,
Therefore, (6.5) holds and f is differentiable at x with Df (x) = A.
Conversely, if f is differentiable at x and we set A = Df (x) in (6.12) then, (6.6) asserts that
limh→0 k∆x,A f (h)k = 0. But then limh→0 ∆x,A f (h) = 0 = ∆x,A f (0), which is precisely the statement
that ∆x,A f is continuous at 0.
Notation 6.25. If f is differentiable at x, then we let ∆x f (h) denote ∆x,Df (x) f (h).
a2 +bc ab+bd
5 a b
For instance, if A = c d then f (A) = ca+dc cb+d2
.
ANALYSIS III 48
CHAPTER 6. THE DERIVATIVE
Lemma 6.26. Let τ > 0 and consider a function δ from the open ball Bτ ⊂ Rn to Rk defined by
δ(h) := ξ(h) η(h), 0 < khk < τ, δ(0) := 0,
where, ξ : Bτ \ {0} → R is bounded and η : Bτ → Rk is continuous at 0 ∈ Bτ and η(0) = 0. Then δ is
continuous at 0 ∈ Bτ .
Proof. By continuity of η at 0, given ε > 0, ∃ σ ∈ (0, τ ) such that khk < σ ⇒ kη(h)k < ε.
By boundedness of ξ, ∃ M > 0 such that kξ(h)k < M ∀ h ∈ Bτ \ {0}.
Therefore, 0 < khk < σ ⇒ kδ(h)k < M ε, i.e., limh→0 δ(h) = 0 = δ(0) and this completes the proof of
the lemma.
Proof of Chain Rule. As in the proof of Proposition 6.7 we have
f (x + h) = f (x) + Df (x)h + ∆x f (h)khk
and
g(f (x) + k) = g(f (x)) + Dg(f (x))k + ∆f (x) g(k)kkk (6.13)
where
f (x+h)−f (x)−Df (x)h , if h 6= 0,
khk
∆x f (h) :=
0, if h = 0.
and
g(f (x)+k)−g(f (x))−Dg(f (x))k , if k 6= 0,
kkk
∆f (x) g(k) :=
0, if k = 0.
Set k(h) := Df (x)h + ∆x f (h)khk in (6.13). Then, by linearity of Dg(f (x)),
g(f (x + h)) = g(f (x)) + Dg(f (x))(Df (x)h)
+ khkDg(f (x))(∆x f (h)) + kk(h)k ∆f (x) g(k(h)).
Therefore,
g(f (x + h)) − g(f (x)) − Dg(f (x)) ◦ Df (x)h = khk δ1 (h) + δ2 (h)
where,
δ1 (h) := Dg(f (x))(∆x f (h)),
kk(h)k
and δ2 (h) := ∆f (x) g(k(h)), h 6= 0, δ2 (0) := 0.
khk
The proof of the Chain Rule will be complete once we prove that
lim kδ1 (h)k = 0 and lim kδ2 (h)k = 0.
h→0 h→0
We start with δ1 (h).
kδ1 (h)k 6 |||Dg(f (x))k||| k∆x f (h)k
and, by differentiability of f at x, we have limh→0 k∆x f (h)k = 0. It follows immediately that limh→0 kδ1 (h)k =
0.
We move on to δ2 (h). For h 6= 0, set
kk(h)k kDf (x)hk
ξ(h) := 6 + k∆x f (h)k 6 |||Df (x)||| + k∆x f (h)k.
khk khk
The continuity of ∆x f at 0 implies that ξ(h) is bounded on Bτ \ {0} for some τ > 0. Next set
η(h) := ∆f (x) g(k(h)).
k(h) is a continuous function of h and k(0) = 0. Therefore, by differentiability of g at f (x) and Proposition
6.7, η(h) is a continuous function of h and η(0) = 0. We may therefore apply Lemma 6.26 to δ2 (h) =
ξ(h) η(h) to conclude that limh→0 kδ2 (h)k = 0.
The proof that g ◦f is differentiable at x and (6.11) holds is complete.
ANALYSIS III 49
CHAPTER 6. THE DERIVATIVE
∂ g ◦f (x) = ∂g(f (x)) · ∂f (x) where · stands for matrix multiplication. (6.14)
More eexampleicitly,
∂
The entry in the j th row and ith column of ∂ g ◦f (x) can be written as gj (f (x) . If we set y = f (x)
∂xi
and we see g as a function of y = (y1 , . . . , yk ) then the entry in the j th row and rth column of ∂g(f (x))
∂gj
can be written as (f (x)). Then, (6.15) can be written as
∂yr
k
∂ ∂gj ∂yr ∂yr ∂fr
(x) where, by (x) we mean (6.16)
X
gj (f (x) = (f (x)) (x).
∂xi r=1
∂yr ∂xi ∂xi ∂xi
(6.16) is perhaps more memorable than (6.15) because we can imagine cancelling ∂yr from the denominator
and numerator in the terms of the sum in (6.16). However, it is important to appreciate the difference
∂ ∂gj
between gj (f (x) and (f (x)). In the first expression, the function gj ◦f is being differentiated
∂xi ∂yi
with respect to its ith variable (1 6 i 6 n) and evaluated at x whereas in the second expression it is the
function gj that is being differentiated with respect to its ith variable (1 6 i 6 k) and then evaluated at
y = f (x).
and therefore,
∇ g ◦f (x) = g 0 (f (x))∇f (x).
In the example that follows, the gradient of a function will be written as a row vector (to make it easier to
write)!
n 2
√ Calculate ∇kxk, x ∈ R \ {0}, by applying the chain rule to g ◦f where f (x) := kxk
Example 6.27.
and g(t) := t, t > 0.
Solution. g 0 (t) = 2√ 1
t
and, since f (x) = x21 + · · · + x2n , we have that ∂i f (x) = 2xi , i.e., ∇f (x) =
(∂1 f (x), . . . , ∂n f (x)) = (2x1 , . . . , 2xn ) = 2x. Therefore,
1 x
∇kxk = ∇ g ◦f (x) = g 0 (f (x))∇f (x) = p 2x = . (6.17)
2 kxk2 kxk
ANALYSIS III 50
CHAPTER 6. THE DERIVATIVE
Another common application of the chain rule occurs in the calculation of the derivative of f ◦ r where
r : R → Rn is a parameterisation of a path in Rn and f : Rn → R is a scalar function. In this case, if
r(t) = (x1 (t), . . . , xn (t)) then
n
0
X ∂f dxi
(f ◦ r) (t) = (r(t))
i=1
∂xi dt
Example 6.28. Fix x ∈ Rn and define r : R → Rn by r(t) := tx, i.e., if x 6= 0 then r is a parameterisation
of the line through 0 and x. Given f : Rn → R calculate (f ◦ r)0 (t) in terms of ∇f .
d ∂f ∂f
f (r(t)) = x1 (tx) + · · · + xn (tx).
dt ∂x1 ∂xn
d
∂v f (x) = Df (x) (x + tv) (chain rule)
dt t=0
= Df (x)v.
Solution.
∂ 2 −y
ux (x, y) = f 0 (x2 e−y ) (x e ) = 2xe−y f 0 (x2 e−y ),
∂x
∂
uy (x, y) = f 0 (x2 e−y ) (x2 e−y ) = −x2 e−y f 0 (x2 e−y ).
∂y
Therefore xux + 2uy = 2x2 e−y f 0 (x2 e−y ) + (−2x2 e−y f 0 (x2 e−y )) = 0.
ANALYSIS III 51
CHAPTER 6. THE DERIVATIVE
Theorem 6.30. Consider f : U → Rk and suppose there exists B(x, r) ⊂ U such that the Jacobian matrix
∂f (y) exists at all points of B(x, r) and that ∂f is continuous at x. Then f is differentiable at x and
Df (x) h = ∂f (x) h ∀ h ∈ Rn .
Remark 6.31. Recall from §5.2.1 that, writing f as (f1 , . . . , fk ), ∂f is continuous at x if, and only if,
∂1 f1 , . . . , ∂n f1 , ∂1 f2 , . . . , ∂n f2 , . . . , ∂1 fk , . . . , ∂n fk are all continuous at x.
Note, too, that we need to make an assumption on the behaviour of the partial derivatives of f at all
points y sufficiently near x in order to conclude the existence of Df at just x.
If we succeed, then we would have proved that Df (x1 , x2 ) is the linear map from R2 to R defined by
Partial derivatives only provide information along lines parallel to the axes. Therefore, we have to break
f (x1 + h1 , x2 + h2 ) − f (x1 , x2 ) into differences along the axes as follows:
f (x1 + h1 , x2 + h2 ) − f (x1 , x2 ) = f (x1 + h1 , x2 + h2 ) − f (x1 + h1 , x2 ) + f (x1 + h1 , x2 ) − f (x1 , x2 )
= II + I.
The second term I can be written in terms of ∂1 f by applying the mean value theorem to f (·, x2 ), x2
fixed. Namely,
By continuity of ∂1 f and ∂2 f at (x1 , x2 ), given ε > 0, ∃ δ > 0 (which we may, and shall, assume to
be less than r) such that
Now k(h1 , θ2 h2 )k < k(h1 , h2 )k and k(θ1 h1 , 0)k < k(h1 , h2 )k and therefore, if k(h1 , h2 )k < δ then, using
(6.24) in (6.23) with (h̃1 , h̃2 ) = (θ1 h1 , 0) in the first term and (h̃1 , h̃2 ) = (h1 , θ2 h2 ) in the second term,
yields √
k∆f (h1 , h2 )k < ε(|h1 | + |h2 |) 6 ε 2 k(h1 , h2 )k.
In other words, we have established (6.20) and completed the proof of Theorem 6.30.
ANALYSIS III 52
CHAPTER 6. THE DERIVATIVE
Proof. Recall that L (Rn , Rk ) and Rk,n are identified via the map µ : L (Rn , Rk ) → Rk,n defined by
(5.3) and that, by Proposition 6.15, if Df (x) exists, then ∂f (x) = µ(Df (x)). Furthermore, by the
discussion in §5.2.1, the continuity at p of x 7→ Df (x) : U → L (Rn , Rk ) implies the continuity at p of
x 7→ ∂f (x) : U → Rk,n .
Conversely, if ∂f is continuous on U then Theorem 6.30 assures us that Df exists at all points of U
and that ∂f = µ ◦ Df . We can then appeal again to the discussion in §5.2.1 to assert that the continuity
at p of ∂f implies the continuity at p of Df .
This proposition is useful because it provides us with a practical way of checking continuous
differentiability, namely, we simply have to compute all the first order partial derivatives ∂i fj
of f = (f1 , . . . , fk ) and verify that they are all continuous. This means that most functions
that one can eexampleicitly write down in terms of polynomials, exponential, logarithm, etc. are
continuously differentiable on their ‘natural’ domain of definition.
Notation 6.34.
C 1 (U, Rk ) := {f : U → Rk | ∂f : U → Rk,n is continuous}.
C 1 (U ) := C 1 (U, R).
ANALYSIS III 53
Chapter 7
Complex Analysis
This part of the course is an introduction to complex analysis. The main topics will be complex differ-
entiability, power series and contour integrals. Basic notions and properties for complex numbers were
introduction in Year 1 and we only provide a quick review here.
C = {z = x + iy, x, y ∈ R},
with i2 = −1. For z = x + iy as above we say that x is the real part of z, denoted by x = Re z and that
ypis the imaginary part of z, denoted by y = Im z. By |z| we denote the modulus (or norm) of z, given by
x2 + y 2 . We denote by z̄ the complex conjugate of z. That is, if z = x + iy then z̄ = x − iy. It is easy
to see that
1. z̄¯ = z,
2. z + w = z̄ + w̄,
3. zw = z̄ w̄,
Notice that we can identify C with R2 , simply by identifying z = x + iy with (x, y). In this way |z|
corresponds to the Euclidean norm in R2 . We will not use k · k, the notation that we used for the norm in
R2 .
The notions of convergence, open and closed for C are identical to those in the plane (see results in
Definition 4.1.
Definition 7.1. We say that (zn )∞n=1 ⊂ C converges to z if and only if |zn − z| tends to zero as n goes to
∞. That is, if for every ε > 0 there exists N ∈ N such that |zn − z| < ε for all n > N .
Definition 7.2. We say that Ω ⊂ C is open if and only for every x ∈ Ω there exits r > 0 such that
Br (x) = {z ∈ C |z − x| < r} ⊂ Ω. We say that Ω is closed if and only if Ωc is open.
Definition 7.3. A set K ⊂ C is sequentially compact if and only if for every sequence (xj )j∈N ⊂ K has a
convergent subsequence (xj(l) )l∈N whose limit is in K.
Now, maps in f : Ω ⊂ C → C, are given by a pair of real-valued functions f (z) = u(z) + iv(z), the
real part u of f and the imaginary part v of f . We can think of those two functions as functions of z or
as functions in R2 of x and y, the real and imaginary part of z. This means we can also think of f as a
function from Ω ⊂ R2 to R2 .
54
CHAPTER 7. COMPLEX ANALYSIS
Definition 7.4. Given f : Ω ⊂ C → C we say that it is continuous at z0 ∈ Ω if and only if for every ε > 0
there exists δ such that |z − z0 | < δ, with z ∈ Ω implies that |f (z) − f (z0 )| < ε.
Notice that the notion of continuity coincides with the one defined for maps from R2 to R2 . We will
now consider the notion of differentiability, where the two notions differ very significantly.
Recall that a function f : Rn → Rk is differentiable at a point p if and only if there exists a linear map
Df (p) ∈ L (Rn ; Rk ) such that
kf (p + h) − f (p) − Df (p)hk
lim = 0. (7.1)
h→0 khk
The reason for introducing that definition arose from the fact that when k > 1 we have no notion of
division for the quantity we would like to study
f (p + h) − f (p)
lim
h→0 h
as division by h ∈ Rn , n > 1 is not well defined. However, in C we do have a notion of multiplication and
therefore we can use that quotient to define differentiability.
Definition 7.5. Let Ω ⊂ C be an open set and z ∈ Ω. We say that f is complex differentiable at z if and
only if the limit
f (z + h) − f (z)
lim (7.2)
h→0 h
exists. We denote the limit by f 0 (z).
In contrast to what happened in the real valued case, where the derivative was a linear map from Rn to
Rk , which in our case would mean from R2 to R2 , corresponding to a 2 by 2 matrix, in the complex case we
obtain a complex number. Before studying how to reconcile this difference, we look at the consequences
of the definition for the real and imaginary part of f . Let’s write h = ∆x + i∆y, and f (z) = u(z) + iv(z),
which we can also think of as f (x, y) = u(x, y) + iv(x, y). Then the quotient in the definition of complex
derivative can be rewritten as
f (z + h) − f (z) u(x + ∆x, y + ∆y) − u(x, y) + i[v(x + ∆x, y + ∆y) − v(x, y)]
= .
h ∆x + i∆y
We could consider multiple ways of sending ∆x + i∆y to zero, obtaining the same answer if the limit exists.
We will consider the two obvious options, sending ∆x first to zero followed by ∆y, and the reverse, ∆y
first followed by ∆x. We find
ANALYSIS III 55
CHAPTER 7. COMPLEX ANALYSIS
ux = vy uy = −vx (7.3)
These equations are known as the Cauchy–Riemann equations. These are clearly necessary conditions, but
at this point in no way guarantee that a complex derivative would exists if (7.3) is satisfied.
By considering two simple examples, it is easy to see that the notion of complex derivative is highly
restrictive as functions that are obviously smooth when considered as a map from R2 to R2 are not actually
complex differentiable. First we consider f (z) = z. Notice that f 0 (z) exists and equals 1. Indeed
z+h−z
f 0 (z) = lim = 1.
h→0 h
However if we consider g(z) = z̄, we obtain a function that is not complex differentiable. We have
g(z + h) − g(z) z̄ + h̄ − z̄ h̄
lim = lim = lim ,
h→0 h h→0 h h→0 h
a limit that does not exist. (Consider for example the limits obtained by taking h along the real or the
imaginary axis.) The function g does not satisfy the Cauchy–Riemann equations. We have g(z) = x − iy,
and therefore
ux = 1, vy = −1, uy = 0, vx = 0.
When considering g as a function from R2 to R2 we have g(x, y) = (x, −y) we clearly have a differentiable
function, as all components are smooth functions. (The existence of continuous partial derivatives suffices
to obtain differentiability, see Theorem 6.30.)
Definition 7.6. We say that f : Ω → C is analytic (or holomorphic) in a neighbourhood U of z if it is
complex differentiable everywhere in U . We say that f is entire if it is analytic in the whole of C.
A function can be differentiable at one point, but not necessarily analytic. Consider as an example
the function f (z) = |z|2 . We will show that the function is complex differentiable at 0, but that it is not
analytic, as it is not complex differentiable outside the origin. Notice that f (z) = x2 + y 2 , and u = x2 + y 2
and v = 0. When computing the Cauchy–Riemann equations we find
ux = 2x uy = 2y, vx = vy = 0.
The Cauchy–Riemann equations mean 2x = 0 and 2y = 0, which is only satisfied at the origin. Now, to
check that f is complex differentiable at the origin
ANALYSIS III 56
CHAPTER 7. COMPLEX ANALYSIS
Therefore
∂f 1 1 1 1
= uz̄ + ivz̄ = ux − uy + i vx − vy ,
∂ z̄ 2 2i 2 2i
which we can simplify to
∂f 1 1
= [ux − vy ] + i [vx + uy ] .
∂ z̄ 2 2
Notice that if the function is complex differentiable, it satisfies the Cauchy–Riemann equations and therefore
the expression above is identically zero. In this sense we say that if a function is complex differentiable,
then
∂f
= 0.
∂ z̄
This illustrates why f (z) = z̄ or g(z) = |z|2 = z z̄ were not complex differentiable.
We conclude this section by proving that f (z) = z n is complex differentiable for every n ∈ N. Using
Theorem 7.7 it suffices to show that it has a derivative at every point and that it satisifies the Cauchy–
Riemann equations. Notice that since it is a polynomial (once expanded in terms of x and y and considered
as map from R2 to R2 we trivially have that it has a derivative). To see that it satisfies the Cauchy–Riemann
equations, notice that (and similarly for v)
ux = (Re f )x = Re(fx ).
Therefore
fx = ux + ivx = n(x + iy)n−1 fy = uy + ivy = n(x + iy)n−1 i.
Without computing what ux , vx , uy , vy are, notice that it follows from the expression above that
which implies that ux = vy and uy = −vx , which are the Cauchy–Riemann equations.
The remainder of Section 7.1 is optional. It was not covered in class, and it is not examinable.
Now that we know that the Cauchy–Riemann equations need to be satisfied for a function to be complex
differentiable we can identify the complex plane with a subspace of 2 × 2 matrices. This identification will
allow us to connect directly complex differentiability with the standard notion of differentiability discussed
in Chapter 6. We have already identified a + ib with the point in R2 given by (a, b). We can also identify
it with the matrix !
a −b
.
b a
Note that which factor of b contains a minus sign is just a convention. Notice that the determinant of that
matrix equals |a + ib|2 , and that therefore the matrix is invertible unless a + ib = 0. This identification
preserves the basic operations we have for complex numbers, for example summation and multiplication.
That is it is possible to perform the operation (a + ib) + (c + id) as complex numbers or as the sum of
the two corresponding matrices, with the results agreeing (modulo the identification). For the product we
have
(a + ib)(c + id) = (ac − bd) + i(bc + ad)
and ! ! !
a −b c −d ac − bd −(bc + ad)
= ,
b a d c bc + ad ac − bd
proving the result. Sometimes it is useful to consider a hybrid of both identification, the one as a matrix,
and the one as a point (or vector) in R2 . For example, for the product of two complex numbers that we
have just considered, we could identify it with
! !
a −b c
.
b a d
ANALYSIS III 57
CHAPTER 7. COMPLEX ANALYSIS
Before we prove this result, we emphasize that some books will replace the right-hand side by asking
that the Cauchy–Riemann equations are satisfied and that all partial derivatives are continuous. Notice
that this last condition implies the existence of a derivative.
f (z + h) − f (z)
lim = f 0 (z),
h→0 h
which we can rewrite as
f (z + h) − f (z) − f 0 (z)h
lim = 0. (7.4)
h→0 h
In order to prove that f is differentiable as a map in R2 we need to find a linear map Df that satisfies
(7.1), which translates in finding a 2 × 2 matrix. Notice that (7.4) suggest that f 0 (z) ∈ C should be the
map. Indeed if we identify f 0 (z) with the corresponding matrix, and think of f 0 (z)h not as a product of
two complex numbers but as a matrix acting on the vector h then we have in fact proven that f has a
derivative. Since we already know that all complex differentiable functions satisfy the Cauchy–Riemann
equations we have completed that implication.
For the reverse, assuming that we have a derivative, that means that we have a 2 × 2 matrix which is
given by !
ux uy
Df ((a, b)) =
vx vy
and that satisfies
|f ((a, b) + h) − f ((a, b)) − Df ((a, b))h|
lim = 0.
h→0 |h|
Since the Cauchy–Riemann equations are satisfied we know that this matrix does in fact have the form
!
ux −vx
,
vx ux
meaning that we could identify it with a complex number as before. We could therefore identify Df h with
the product of the complex numbers f 0 (z) = ux + ivx and h. Identifying (a, b) with z we obtain
|f (z + h) − f (z) − f 0 (z)h|
lim =0
h→0 |h|
ANALYSIS III 58
CHAPTER 7. COMPLEX ANALYSIS
As a consequence of the above result, since we can connect complex derivatives with derivatives of
maps from R2 to R2 , we have the following results:
Theorem 7.8. Lef f, g : Ω ⊂ C → C be complex differentiable functions. Then (asumming g 6= 0 in the
third expression) we have that the familiar expressions
0
f f 0g − f g0
(f + g)0 = f 0 + g 0 (f g)0 = f 0 g + f g 0 = (f (g))0 = f 0 (g)g 0
g g2
apply to the complex-valued case as well. For the final expression one needs to assume that the composition
makes sense, i.e. the range of g is contained in the domain of f .
reviewing (in a very utilitarian way) some basic ideas of series for complex numbers covered in year 1.
P∞ PN
Definition 7.9. The series n=0 an , with an ∈ C is convergent if and only if the sequence SN = n=0 an
is convergent in C.
P∞
Definition 7.10. The series n=0 an , with an ∈ C is absolutely convergent if and only if the series
P∞
n=0 |an | is convergent.
The geometric series ∞ n=0 z is convergent if and only if |z| < 1, and sums up to 1/(1 − z) (with
n
P
partial sums SN = (1 − z N +1 )/(1 − z) ). We review a couple of the convergence tests from year 1.
P∞
Theorem 7.11 (Ratio Test). Consider n=0 an and assume that an 6= 0 for all n. Then
with an , z ∈ C.
Theorem 7.13. Given (an )∞
n=0 there exists R ∈ [0, ∞] such that
∞
X
an z n
n=0
1
converges for all |z| < R and diverges for |z| > R. (As we will see in the proof R = lim sup |an |1/n
.) The
quantity R is called the radius of convergence of the series.
ANALYSIS III 59
CHAPTER 7. COMPLEX ANALYSIS
P∞
Proof. We consider z given, but fixed, and apply the root test to the series n=0 an z
n. This series is
convergent if
lim sup |an z n |1/n
is less than 1 and divergent if it is greater than 1. But that translates to convergence if
1
|z| <
lim sup |an |1/n
Theorem 7.14. Let an 6= 0 for all n ≥ N , and assume that lim |a|an+1 | P∞ n
n|
exists. Then n=0 an z has radius
|an |
of convergence R = lim |an+1 | .
Next we will show that within the radius of convergence a power series is actually differentiable, and
that we can in fact compute the derivative term-by-term. More precisely:
Proof. First we will show that the power series for f 0 (z) does have the same radius of convergence. Notice
that the radius of convergence of ∞ n−1 and P∞ na z n (i.e. where we have multiplied the
P
n=1 nan z n=1 n
expression by z) is the same. To see this notice that if ∞ n−1 is convergent for |z| < R and
P
n=1 na nz
divergent for |z| > R then the same will apply to the second series. Therefore we just need to consider
lim sup |nan |1/n = lim n1/n lim sup |an |1/n = lim sup |an |1/n ,
which shows that the radius of convergence is the same as for ∞ n=0 an z . Notice that the series
n P
P∞ n−2 also has the same radius of convergence (we will need this result in our estimate
n=2 n(n − 1)an z
below, even though we never formally compute second derivatives).
Next, notice that for k ∈ N we have
wk − z k
= wk−1 + wk−2 z + · · · + wz k−2 + z k−1 . (7.5)
w−z
Now, in order to prove that f is complex differentiable and compute its derivative we study
∞
f (z + h) − f (z) X
− nan z n−1 .
h n=1
We denote by w = z + h (and so h = w − z), and substitute the expression for f in terms of a series to
find
∞ n
w − zn
X
an − nz n−1 .
n=0
w−z
We look more carefully at the term in brackets. Using (7.5) we find (taking k = n)
wn − z n
− nz n−1 = wn−1 + wn−2 z + · · · + wz n−2 + z n−1 − nz n−1
w−z
ANALYSIS III 60
CHAPTER 7. COMPLEX ANALYSIS
wk − z k
= |wk−1 + wk−2 z + · · · + wz k−2 + z k−1 | < krk−1
w−z
and therefore
wk − z k n−k−1
z ≤ krk−1 rn−k−1 ≤ krn−2 ,
w−z
which substituted in (7.6) yields
h i
≤ |w − z| |(n − 1)rn−2 + (n − 2)rn−2 + · · · 2rn−2 + rn−2 |
1
≤ |w − z|rn−2 n(n − 1).
2
We have shown that
∞ ∞
f (z + h) − f (z) X 1X
− nan z n−1 ≤ |w − z| n(n − 1)|an |rn−2 ≤ M |h|,
h n=1
2 n=0
which goes to zero as h goes to zero. Notice that in the last inequality we have used that the series
P∞ n−2 is finite, since we observed that the radius of convergence of the corresponding
n=0 n(n − 1)|an |r
power series was also R.
We have the following simple consequence of the Theorem above, which allows us to compute the
coefficients an in terms of derivatives of f .
Proof. The result is trivial for f (0), as it clearly equals a0 . A simple induction argument using the formula
for the derivative of f in the previous Theorem yields the desired result.
Proof. We show the result by proving that (fk ) is uniformly Cauchy in |z| ≤ r. We have (assuming that
j ≤ k)
k
X k
X ∞
X
|fk (z) − fj (z)| = | an z n | ≤ |an |rn ≤ |an |rn .
n=j+1 n=j+1 n=j+1
P∞
Since by assumption n=0 |an |rn is finite, given any ε > 0 we can choose N large enough to make
|fk (z) − fj (z)| < ε for all j, k > N , concluding the proof. (This proof is essentially an application of the
Weierstrass M-test that we covered a few weeks ago.)
ANALYSIS III 61
CHAPTER 7. COMPLEX ANALYSIS
The ratio test shows (Exercise) that the radius of convergence of all of the series above is R = ∞.
Notice that using Theorem 7.15 we can prove well known identities like (ez )0 = ez . Indeed
∞
!0 ∞ ∞
z 0
X 1 n X n n−1 X 1 n
(e ) = z = z = z = ez .
n=0
n! n=1
n! n=0
n!
In fact, we can easily relate all the circular functions to the exponential.
∞ n ∞
" #
1 X i + (−i)n n X (−1)k 2k
= z = z ,
2 n=0 n! k=0
(2k)!
where we have used that
n even
(
n n 2(i)n = 2(−1)n/2
i + (−i) = .
0 n odd
There are additional relationships between sine and cosine and their hyperbolic counterparts. Notice
that we have
which shows that sine and cosine are unbounded functions in the complex plane. Just consider z = iy for
y ∈ R together with the fact that the real valued sinh and cosh grow exponentially at infinity.
ANALYSIS III 62
CHAPTER 7. COMPLEX ANALYSIS
2. ez 6= 0 for all z ∈ C.
3. ez = 1 if and only if z = 2kπi for k ∈ Z, and as a result ez+w = ez if and only if w = 2kπi, k ∈ Z.
Notice that in particular we have shown ez+2kπi = ez for all k ∈ Z, so in this sense the exponential
is periodic in the imaginary variable.
Proof. We present the direct proof of part 1, without using more advanced tools from complex analysis
that would reduce the heavy computational nature.
∞ ∞ ∞ X
! !
z w
X 1 n X 1 k
X z n wk
e e = z w =
n=0
n! k=0
k! l=0 n,k
n! k!
n+k=l
∞ X
l ∞
!
X 1 l j l−j X 1
= z w = (z + w)l = ez+w .
l=0 j=0
l! j l=0
l!
For part 2, notice that ez e−z = 1, proving that ez 6= 0. For part 3, denoting z = x + iy we find
which equals 1 if and only if ex = 1 and cos y + i sin y = 1. These only happen if x = 0 and y = 2πk,
k ∈ Z. Similarly for part 4.
1.5
r
1.0
θ
0.5
x
-1.5 -1.0 -0.5 0.5 1.0 1.5
-0.5
-1.0
-1.5
It is not a function as such, as the image is not uniquely defined, and if θ ∈ arg(z) then so is θ + 2kπ.
The following are easily verified properties of arg.
ANALYSIS III 63
CHAPTER 7. COMPLEX ANALYSIS
Proposition 7.21.
4. arg(1/z) = − arg(z),
The ambiguity of the argument function can be solved by defining the principal value Arg of the arg
function to take values in (−π, π]. That is for any z ∈ C we have Arg(z) ∈ (−π, π].
Notice that it is impossible to define the Arg function continuously in the entire plane. In particular
as we approach any point in the negative real axis, if we do it from above the Arg function will yield π,
while if we do it from below it will −π. Observe that if we had made any other choice for the range of Arg
there would always be a half-line where we have the same issue, the difference between the values of the
argument when approaching from opposite sides is always 2π.
We want to define the logarithm by analogy of what happens in R. In the real valued case we say (here
w, z ∈ R)
w = log(z) if and only if ew = z.
If we could extend this for w, z ∈ C, since we know that ew = ew+2πik for any k ∈ Z we would have that
if w = log(z) the so is w + 2πik. Therefore we would have that log(z) is a multivalued function (just like
it happened before with arg(z), the argument function).
Let’s write z = |z|ei arg(z) and w = log(z) = u + iv. We have
and therefore comparing the two expressions in polar form we must have
That means that u = log |z|, with this logarithm being the real logarithm. We will denote by Log the
logarithm in R to distinguish it from the complex valued we want to define. We define the multivalued
function
log(z) = Log|z| + i arg(z). (7.13)
In terms of the Arg function we have
Notice that the definition above makes sense provided that z 6= 0, where the real logarithm is not defined.
We can now compute logarithms of negative numbers.
The complex logarithm we have just defined obeys many of the properties that we know for the real
logarithm, with the caveat that we have to take care of the multi-valuedness of the function. For example
ANALYSIS III 64
CHAPTER 7. COMPLEX ANALYSIS
To prove this result, notice that since Log|zw| = Log(|z||w|) = Log|z| + Log|w| and that arg(zw) =
arg(z) + arg(w) we have
This equality needs to be understood modulo 2πi, that is there exists k ∈ Z such that
Similarly we have
log(z/w) = log(z) − log(w).
If we want to consider the (complex) differentiability of the log we have to deal with the multi-valuedness
of the arg function. Indeed if we consider the incremental quotient
This function, defined on C\{x ≤ 0} is single valued. If we consider points of the form z = x ± iε, for
x < 0 and ε > 0 small, we find
lim Log(x ± iε) = Log(−x) ± iπ,
ε→0
showing that the function could not be extended continuously along {x < 0}. This half-line is called a
branch cut. It is possible to compute the derivative of Log directly from the definition, or in terms of its
inverse. However, for practical purposes, once we know it is differentiable, from the identity
eLogz = z
we find
eLogz (Logz)0 = 1
from which it follows that (Logz)0 = 1/z.
Once we have defined the notion of logarithm it is possible to consider defining complex powers of
complex numbers. Given α ∈ C, and z 6= 0 we define the α-th power of z by
The multi-valuedness of arg means that the same is true for z α . If we rewrite the above as
for k ∈ Z the multi-valuedness becomes more evident. The number of α powers, whether it is one, finitely
many or infinitely many will depend on α.
Indeed if α is an integer for example then e2παki = 1, which means that in fact there is only one value
of z α . If α is rational, say α = p/q, with p, q coprime, then z α will have finitely many powers. It is easy
to see that for α = p/q (with p, q coprime, and q ∈ N)
e2παki = e2πα(k+q)i
ANALYSIS III 65
CHAPTER 7. COMPLEX ANALYSIS
eαLog(z) e2παki , k = 0, 1, . . . , q − 1.
We also have the following estimate (which we will use repeteadly below)
ˆ b ˆ b
f (t)dt ≤ |f (t)|dt. (7.16)
a a
ANALYSIS III 66
CHAPTER 7. COMPLEX ANALYSIS
´b ´b
To prove this result, assume that a f (t)dt = Reiθ , where R = a f (t)dt . As a result of this
representation R also equals
ˆ b ˆ b
R = e−iθ f (t)dt = e−iθ f (t)dt.
a a
Now, if we write e−iθ f (t) = u + iv, with u and v real valued. Then we must have
ˆ b ˆ b
R= udt vdt = 0.
a a
where Γ is a curve in the complex plane. To define a curve in C, consider a function γ : [a, b] → C,
given by γ(t) = x(t) + iy(t). We will ask that the curve γ be C 1 . The primary reason is that we want to
have a well defined tangent at every point of the curve (which is also integrable). We say that the curve
Γ = γ([a, b]) ⊂ C is parametrised by the map γ.
Definition 7.22. Given a function f : Ω ⊂ C → C along the path Γ ⊂ Ω ⊂ C parametrised by
γ : [a, b] → C the integral of f over Γ is given by
ˆ ˆ b ˆ b ˆ b
0 0
f dz = f (γ(t))γ (t)dt = Re(f (γ(t))γ (t))dt + i Im(f (γ(t))γ 0 (t))dt.
Γ a a a
Notice that we are not making any regularity assumptions on f , just that the integrals are well defined.
Sometimes ´ we will ´consider more than ´one parametrisation of a curve Γ, say γ1 and γ2 and will use the
notation γ1 f and γ2 f in addition to Γ .
On many occasions we want to consider curves that are not C 1 but perhaps just piece-wise C 1 . For
example a square. In this case we can think of Γ as a union of n curves Γj , each one C 1 , and parametrised
in the right direction, so that connected in the right order they describe the entire curve Γ. We can define
ˆ Xn ˆ
f dz := f dz.
Γ j=1 Γj
It is straightforward from the definition (details are left as an Exercise) that given a curve Γ, and two
functions f, g : C → C and α, β ∈ C we have
ˆ ˆ ˆ
(αf (z) + βg(z))dz = α f (z)dz + β g(z)dz.
Γ Γ Γ
If we allow for not to exist at finitely many points, this can be defined as a single integral, with clearly
γ 0 (t)
both formulations being equivalent.
Example 7.23. Let f : C → C be given by f (z) = f (x + iy) = x4 + iy 4 and the curve joining the origin in
a straight line to the point 1 + i, parametrized by γ : [0, 1] → C, γ(t) = (1 + i)t. Notice that γ 0 (t) = 1 + i
and so we have ˆ ˆ 1 ˆ 1
4 4 2
f= (t + it )(1 + i)dt = 2it4 dt = i.
Γ 0 0 5
ANALYSIS III 67
CHAPTER 7. COMPLEX ANALYSIS
´
In the next Lemma we want to show that Γf depends only on the orientation of the parametrisation
of the curve. More precisely
Lemma 7.24. Let Γ be a curve in C, parametrised by γ : [a, b] → C, that is γ([a, b]) = Γ. Given
f : Ω ⊂ C → C and Γ ⊂ Ω we have:
1. if γ − represents the parametrisation of γ in the opposite direction, then
ˆ ˆ
f = − f.
γ− γ
If a curve Γ has attached a sense of direction we will call it a directed curve. In this case we will
denote by −Γ the same curve swept in the opposite direction. Without the need to specify the
parametrisation we can reformulate the above result by
ˆ ˆ
f dz = − f dz.
Γ −Γ
We refer to this fact as reparametrisation invariance. [In practise, with the regularity we are demanding
on the curves, this means that there exists φ : [ã, b̃] → [a, b], bijective and increasing, such that
γ̃ = γ(φ).]
Proof. 1. Notice that if γ : [a, b] → C parametrises the curve in one direction then γ − is given by
γ − : [a, b] → C with γ − (t) = γ(a + b − t). Therefore
ˆ ˆ b ˆ b
− − 0
f= f (γ (t)) γ (t)dt = f (γ(a + b − t))(−γ 0 (a + b − t))dt
γ− a a
ˆ a ˆ b ˆ
0 0
= f (γ(s))(−γ (s))(−1)ds = − f (γ(s))γ (s)ds = − f.
b a γ
where we have made the change of variables φ(t) = s and therefore φ0 (t)dt = ds
Consider the function f (z) = 1 as a complex-valued function and a curve γ : [a, b] → C. Then
ˆ ˆ b
f dz = γ 0 (t)dt.
γ a
where γ : [a, b] → C is given by γ(t) = x(t) + iy(t), and l(γ) stands for the length of the curve γ.
ANALYSIS III 68
CHAPTER 7. COMPLEX ANALYSIS
Therefore we obtain ˆ
f dz ≤ max|f (z)|l(γ).
γ z∈Γ
In fact ˆ 2π
eint dt = 0, for all n 6= 0.
0
ANALYSIS III 69
CHAPTER 7. COMPLEX ANALYSIS
Example 7.27. Integrate f (z) = z along the circle of centred at 1 + i of radius 2 (oriented counterclock-
wise). As before γ(t) = (1 + i) + 2eit for t ∈ [0, 2π]. We have γ 0 (t) = 2ieit . Therefore the integral
becomes
ˆ ˆ 2π ˆ 2π ˆ 2π
it it it
f (z)dz = (1 + i) + 2e 2ie dt = 2(1 + i)i e dt + 4ie2it dt = 0,
γ 0 0 0
´ 2π
using that 0 eint dt = 0, for all n 6= 0.
dF
Theorem 7.28. Assume that F : Ω ⊂ C → C is analytic (Ω open) and set f (z) = dz , with f continuous.
Let γ : [a, b] → Ω be a C 1 curve. Then
ˆ
f dz = F (γ(b)) − F (γ(a)).
γ
Proof. We have
ˆ ˆ b ˆ b ˆ b
dF d
f dz = f (γ(t))γ 0 (t)dt = (γ(t))γ 0 (t)dt = F (γ(t))dt = F (γ(b)) − F (γ(a)).
γ a a dz a dt
We remark that there are no assumptions made about Ω other than it is open. That is, all we need
for the result to be true, is that f is analytic in an open neighborhood of the curve. The notion of simply
connected (for a domain) will be defined later, but we emphasize that there is no such requirement on Ω
above result to be true.
Here n represents the normal, with the following convention. If the curve C is parametrised by r(t) =
(x(t), y(t)), and r0 (t) = (x0 (t), y 0 (t)) has the same direction of the tangent, we choose
When considering the curves determining the boundary of a regular domain we will consider them as
positively oriented. That is, choose the orientation so that the corresponding n as defined above corresponds
(i.e. has the same direction as) to the outward normal.
The following results (considered here only for two dimensions) correspond to Green’s and Gauss’
Theorems. For a positively oriented regular domain Ω we have
¨ ‰
curl Fdxdy = F · dr
Ω ∂Ω
ANALYSIS III 70
CHAPTER 7. COMPLEX ANALYSIS
and ¨ ‰
divFdxdy = F · ndt.
Ω ∂Ω
´
Now let’s consider our contour integral γ f (z)dz for a function f = u + iv and a curve γ(t) =
γ1 (t) + iγ2 (t). We have
ˆ ˆ b
f (z)dz = [u(γ(t)) + iv(γ(t))][γ10 (t) + iγ20 (t)]dt
γ a
ˆ b ˆ b
= u(γ(t))γ10 (t) − v(γ(t))γ20 (t)dt +i u(γ(t))γ20 (t) + v(γ(t))γ10 (t)dt
a a
ˆ b ˆ b
= (u, −v) · (γ10 , γ20 )dt + i (u, −v) · (γ20 , −γ10 )dt
a a
ˆ ˆ
= (u, −v) · dr + i (u, −v) · ndt,
γ γ
and so if we define the vector field f = (u, −v), we have just shown that
ˆ
f dz = circulation(f ) + i flux(f ).
γ
Using the above expression, together with Green’s and Gauss’ Theorem we can prove the following
result.
Theorem 7.29 (Cauchy’s Theorem). Let f : Ω → C be an analytic function, with Ω a simply connected
domain. Let γ be a C 1 closed curve in Ω. Then
ˆ
f dz = 0.
γ
Before we prove the result we define simply connected. Loosely speaking means that the domain
contains no holes. A set of more formal definitions is as follows.
Definition 7.30. A set Ω ⊂ C is connected if it cannot be expressed as the union of non-empty open sets
Ω1 and Ω2 such that Ω1 ∩ Ω2 = ∅. An open, connected set Ω ⊂ C is called simply connected if every
closed curve in Ω can be continuously deformed to a point.
Proof. (THIS PROOF IS NOT EXAMINABLE.) The proof presented here assumes that the curve is
a simple, regular curve and that f 0 is continuous. If the domain is simply connected, the region inside the
curve does not have any holes, and f is analytic in it. We know
ˆ
f dz = circulation(f ) + i flux(f )
γ
¨ ¨
= curl f dxdy + i divf dxdy ,
A A
where A is the region encircled by γ. We claim that both terms are actually 0, because curl f = divf = 0.
Since f = (u, −v) we have
ux = vy vx = −uy
ANALYSIS III 71
CHAPTER 7. COMPLEX ANALYSIS
Notice that Cauchy’s Theorem applies to Example 7.27, where the function is analytic, but obviously
not to Example 7.26, where the function is not analytic.
Cauchy’s Theorem works for more general curves. Consider the shaded region Ω in Figure 7.2. If we
think of its boundary as a one curve Γ, even though it is formed by two separate curves we have
ˆ
f dz = 0,
Γ
provided that Γ is oriented positively. That means that the exterior curve, that we denote by γ1 needs to
be oriented counter-clockwise, while the interior curve, denoted by γ2 has to be oriented clockwise.
An equivalent formulation of this fact, which will be extremely useful is known as the deformation of
contour Theorem.
Theorem 7.31. Let Ω ⊂ C be a region bounded by two closed simple curves γ1 (the exterior curve) and
γ2 (the interior). Assume they are oriented positively (meaning that Ω is to the left of the curves as we
traverse them), and let f be an analytic function in a neighbourhood of Ω ∪ γ1 ∪ γ2 . Then
ˆ ˆ
f dz + f dz = 0.
γ1 γ2
If we denote by γ2− the anti-clockwise parametrization of γ2 , then the result can be rephrased as
ˆ ˆ
f dz = f dz,
γ1 γ2−
that is the integral is the same along both curves when both are parametrised counter-clockwise.
Proof. The proof is based on creating two new contours of integration, the boundaries of two simply
connected regions where f is analytic so that we can apply Cauchy’s Theorem 7.29.
To achieve this we add two new curves to the previous picture, now in yellow in Figure 7.3. They join
the points A (in γ1 ) with D (in γ2 ) and the points B (in γ1 ) with C (in γ2 ). The two curves we want to
consider are denoted by ρ and η. Each one of them is piecewise C 1 and formed by four sections. Each one
of these curves is oriented positively with respect to the region they enclose, that is, they are both oriented
counter-clockwise.
By Cauchy’s Theorem
ˆ ˆ ˆ ˆ ˆ
f dz = f dz + f dz + f dz + f dz = 0, (7.17)
ρ ρ1 ρ2 ρ3 ρ4
ˆ ˆ ˆ ˆ ˆ
f dz = f dz + f dz + f dz + f dz = 0. (7.18)
η η1 η2 η3 η4
ANALYSIS III 72
CHAPTER 7. COMPLEX ANALYSIS
We observe that η1 and ρ4 correspond to the same curve but with parametrisations in opposite directions.
Similarly for η3 and ρ2 . Therefore
ˆ ˆ ˆ ˆ
f dz + f dz = 0 f dz + f dz = 0.
η1 ρ4 η3 ρ2
Adding (7.17) and (7.18) and using the above identities we find
ˆ ˆ ˆ ˆ
f dz + f dz + f dz + f dz = 0
ρ1 ρ3 η2 η4
Also notice that ρ1 and η4 together build γ1 , while ρ3 and η2 build γ2 . Therefore, the above equality can
be rewritten as ˆ ˆ
f dz + f dz = 0.
γ1 γ2
Since ˆ ˆ
f dz = − f dz
γ2 γ2−
we obtain ˆ ˆ
f dz = f dz
γ1 γ2−
as required.
We now compute one of the fundamental contour integrals. We will show that
ˆ (
2πi n = −1,
(z − a) dz =n
(7.19)
∂Br (a) 0 n 6= 1,
where ∂Br (a) denotes the boundary of the ball of radius r, parametrised counter-clockwise (i.e. positively
oriented with respect to Br (a)).
Observe that the result is uniform with respect to r. That is a natural consequence of Theorem 7.31,
given than the functions we are integrating only fail to be analytic at one point (at most, depending on
n). In fact we could have chosen any curve that wraps around a once and obtain the same result.
ANALYSIS III 73
CHAPTER 7. COMPLEX ANALYSIS
Now, to compute the integral above, notice that we can parametrise the curve as γ(t) = a + reit , for
t ∈ [0, 2π). Therefore we have (since γ 0 (t) = ireit )
ˆ ˆ 2π ˆ 2π
n it n it n+1
(z − a) dz = (re ) ire dt = ir ei(n+1)t dt.
∂Br (a) 0 0
Notice that in the case n = −1 that expression equals 2πi. When n 6= −1 notice that we obtain 0, since
for all k 6= 0 we have
ˆ 2π 2π
1 1 1
eikt dt = eikt = − = 0.
0 k 0 k k
We restate, in the notation that will be most convenient for the next few results, the fundamental integral
above in the case n = −1, noting that the result does not depend on r. We have
ˆ
1
dw = 2πi.
∂Br (z) w − z
Definition 7.32. Given a simple closed C 1 curve γ we denote by I(γ) the interior region to γ. We denote
by O(γ) the exterior region to γ.
Notice that by the deformation of contours Theorem we have
ˆ ˆ
1 1
dw = dw = 2πi (7.20)
γ w − z ∂Br (z) w − z
for every z ∈ I(γ) and every r sufficiently small so that Br (z) ⊂ I(γ).
Theorem 7.33. Let γ : [a, b] → C be a positively oriented simple closed C 1 curve. Assume that f is
analytic in γ and on the interior of γ, I(γ). Then
ˆ
1 f (w)
f (z) = dw for all z ∈ I(γ). (7.21)
2πi γ w − z
Proof. Fix z ∈ I(γ), and choose r small enough so that Br (z) ⊂ I(γ). By the deformation of contours
theorem we have ˆ ˆ
1 f (w) 1 f (w)
dw = dw,
2πi γ w − z 2πi ∂Br (z) w − z
reducing the problem to considering γ as a ∂Br (z). Observe that the integral is the same for every r
sufficiently small, and later on we will exploit this fact by talking limits as r tends to zero. For now, we
have
ˆ ˆ ˆ
1 f (w) 1 f (z) 1 f (w) − f (z)
dw = dw + dw =: I + II.
2πi ∂Br (z) w − z 2πi ∂Br (z) w − z 2πi ∂Br (z) w−z
Notice that the first integral I equals f (z). Indeed, using (7.20)
ˆ ˆ
1 f (z) 1 1
dw = f (z) dw = f (z).
2πi ∂Br (z) w − z 2πi ∂Br (z) w − z
All that remains to is to show that II = 0. Notice that since f is analytic in I(γ), given any ε > 0 we can
find r sufficiently small so that
|f (w) − f (z)| ≤ ε for all w ∈ ∂Br (z).
We parametrise ∂Br (z) counterclockwise by γ(t) = z + reit for t ∈ [0, 2π]. We have γ 0 (t) = ireit and
therefore ˆ ˆ 2π
1 f (w) − f (z) 1 f (z + reit ) − f (z) it
|II| = dw ≤ ire dt
2πi ∂Br (z) w−z 2πi 0 reit
ˆ 2π
1
≤ |f (z + reit ) − f (z)|dt ≤ ε.
2π 0
Since ε is arbitrary we obtain the desired result.
ANALYSIS III 74
CHAPTER 7. COMPLEX ANALYSIS
has remarkable consequences for analytic functions. First notice that it claims that we can recover the
value of f at any point by integration along a curve around that point (provided the curve is sufficiently
regular, positively oriented, and contained in I(γ)). This is a very significant difference with respect to
smooth functions in R2 for example.
THIS IS THE END OF THE MATERIAL THAT WE COVERED IN THE MODULE. THE
REST IS NOT EXAMINABLE.
Of course we need to justify moving the derivative inside the integral sign. We assumed that f was analytic,
which means that f 0 (z) exists. The expression above would produce a formula for it, a way to compute it.
The key observation is that without assuming that f has more derivatives it seems that the right hand side
can be differentiated arbitrarily many times, which would suggest that f has infinitely many derivatives.
This is indeed the case as we will show in the next Theorem.
Theorem 7.35. Let γ : [a, b] → C be a positively oriented simple closed C 1 curve. Assume that f is
analytic in γ and on the interior of γ, I(γ). Then f (n) (z) exists for all n ∈ N and the derivative is given by
ˆ
n! f (w)
(n)
f (z) = dw for all z ∈ I(γ). (7.22)
2πi γ (w − z)(n+1)
Proof. Notice that Theorem 7.33 would correspond to the case n = 0 in the current Theorem. In order to
prove the result for n = 1 we consider the incremental quotient, and use (7.21) to obtain
" ˆ ˆ #
f (z + h) − f (z) 1 1 f (w) 1 f (w)
= dw − dw .
h h 2πi γ w−z−h 2πi γ w−z
By the deformation of contours Theorem we can choose γ as ∂B2r (z), with B2r (z) ⊂ I(γ). We have,
operating on the right-hand side
ˆ
f (z + h) − f (z) 1 f (w)
= dw
h 2πi ∂B2r (z) (w − z − h)(w − z)
ˆ ˆ
1 f (w) 1 1 1
= 2
dw + f (w) − dw
2πi ∂B2r (z) (w − z) 2πi ∂B2r (z) (w − z − h)(w − z) (w − z)2
ˆ ˆ
1 f (w) 1 hf (w)
= dw + dw.
2πi ∂B2r (z) (w − z)2 2πi ∂B2r (z) (w − z − h)(w − z)2
To conclude the proof all that we need to do is show that the limit of the last integral as h tends to zero
is zero, that is (ignoring factors of 2πi)
ˆ
hf (w)
lim dw = 0,
h→0 ∂B2r (z) (w − z − h)(w − z)2
ANALYSIS III 75
CHAPTER 7. COMPLEX ANALYSIS
and recall that we are able to choose r arbitrarily small without affecting the value of the integrals above.
First we choose |h| < r so that for all w ∈ ∂B2r (z) we have
|w − z − h| ≥ |w − z| − |h| > r.
Here we have used the reverse triangle inequality in the first case, and the fact that |w − z| = 2r for points
w ∈ ∂B2r (z). Choosing γ(t) = z + 2reit for t ∈ [0, 2π), we have γ 0 (t) = 2rieit , and therefore |γ 0 (t)| ≤ 2r.
Since f is analytic, in particular it is continuous and therefore there exists M > 0 such that |f (w)| ≤ M
for all w ∈ ∂B2r (z). Using these facts we have
ˆ ˆ 2π
hf (w) hM πM
2
dw ≤ 2
2rdt = 2 h,
∂B2r (z) (w − z − h)(w − z) 0 r(2r) r
which goes to zero as h goes to zero, proving the result for n = 1. The general case is proven by induction.
If we assume the result for n = 1, 2, · · · , k − 1 we want to prove it for n = k. That is, in particular we
assume ˆ
(k − 1)! f (w)
f (k−1)
(z) = (k)
dw for all z ∈ I(γ).
2πi γ (w − z)
We write the corresponding incremental quotient, just as before
" ˆ ˆ #
f (k−1) (z + h) − f (k−1) (z) 1 (k − 1)! f (w) (k − 1)! f (w)
= k
dw − k
dw .
h h 2πi γ (w − z − h) 2πi γ (w − z)
By the deformation of contours Theorem we can choose γ as ∂B2r (z), with B2r (z) ⊂ I(γ). We have,
operating on the right-hand side
ˆ
f (k−1) (z + h) − f (k−1) (z) (k − 1)! f (w)[(w − z)k − (w − z − h)k ]
= dw
h 2πih ∂B2r (z) (w − z − h)k (w − z)k
ˆ
k! f (w)
= dw
2πi ∂B2r (z) (w − z)(k+1)
ˆ " #
(k − 1)! [(w − z)k − (w − z − h)k ] k
+ f (w) − dw
2πi ∂B2r (z) h(w − z − h)k (w − z)k (w − z)(k+1)
ˆ ˆ " #
k! f (w) (k − 1)! (w − z)k+1 − (w − z − h)k (w − z) − kh(w − z − h)k
= dw+ f (w) dw.
2πi ∂B2r (z) (w − z)k+1 2πi ∂B2r (z) h(w − z − h)k (w − z)k+1
(7.23)
As before, all that remains is to show that the last integral tends to zero as h tends to zero. We choose h
and the parametrisation as above. The result will follow if we show that
(w − z)k+1 − (w − z − h)k (w − z) − kh(w − z − h)k
≤ C|h|,
h
where the constant might depend on r. This is the case because, as before |f | ≤ M and |w − z − h| ≥
|w − z| − |h| > r implies
1 1
≤ .
(w − z − h)k (w − z)k (2r)k rk
In order to prove (7.23), notice that the binomial formula implies
k
!
k
X k
(w − z − h) = (w − z)k−j (−h)j
j=0
j
and therefore
(w − z)k+1 − (w − z − h)k (w − z) − kh(w − z − h)k
k k
! !
X k X k
=− (w − z)k+1−j (−h)j − kh (w − z)k−j (−h)j
j=2
j j=1
j
which is of order h2 , proving the result.
ANALYSIS III 76
CHAPTER 7. COMPLEX ANALYSIS
where γ is any positively oriented simple closed curve (piece-wise C 1 ) that is contained in BR (a) with
a ∈ I(γ).
Proof. Given some z ∈ BR (a) we will take γ to be ∂Br (a) (positively oriented), for r small enough so
that |z − a| < r < R. We can use the Theorem of deformation of contours to prove the integrals over all
curves γ as above are the same. Cauchy’s formula (7.21) gives
ˆ
1 f (w)
f (z) = dw. (7.24)
2πi ∂Br (a) w − z
Notice that since |w − a| = r and we have chosen r so that |z − a| < r we have |z − a| < |w − a| for all
w ∈ ∂Br (a). As a result
|z − a|
<1
|w − a|
and we can use the geometric series expansion to obtain
∞ n
1 1 1 1 X z−a
= = .
w−z w−a 1− z−a w − a n=0 w − a
w−a
For w ∈ ∂Br (a) the series converges absolutely (Weierstrass M-test), and therefore we can exchange the
order of the summation and integration to obtain
∞ ˆ ∞
X 1 f (w) n
X
f (z) = dw(z − a) = cn (z − a)n ,
n=0
2πi ∂Br (a) (w − a)n+1 n=0
obtaining the desired result. It remains to show that the coefficients are unique. Now, assume that
f (z) = ∞ k=0 bk (z − a) for some bk ∈ C. We have
k
P
ˆ ˆ ∞
f (w) X 1
dw = bk (w − a)k dw
∂Br (a) (w − a)n+1 ∂Br (a) k=0 (w − a)n+1
∞
X
ˆ
= bk (w − a)k−n−1 dw = 2πibn ,
k=0 ∂Br (a)
where we have used the fundamental integrals, together with the fact that we can commute the summation
and integration. This proves that bn = cn , concluding the proof.
ANALYSIS III 77
CHAPTER 7. COMPLEX ANALYSIS
Example 7.37. We consider an example of a Taylor series. We consider the function (1 + z)a for a ∈ C
and |z| < 1.
When we consider logarithms we noticed that z n is well defined for n ∈ N, but not for any a, without
making any specific choice of the argument function. In this case
n
!
X n k
(1 + z)n = z ,
k=0
k
which is a polynomial of order n, and equals the Taylor series expansion centred at the origin. This series
converges for every z ∈ C, not just |z| < 1. However, we defined
having made a choice of the argument function defining the logarithm, which meant creating a branch cut
where the function was not defined. Choosing the argument in (−π, π), and since our function is translated
(not z a ) we obtained a function that is not defined for z ∈ (−∞, −1].
We want to show that in fact a binomial expansion is possible for all a ∈ C. We know by Taylor’s
Theorem 7.36 that we have a Taylor expansion. To compute we need to work out the derivatives of (1+z)a .
Using the definition we have (for (a ∈/ N))
0 a
eaLog(1+z) = eaLog(1+z) a(Log(1 + z))0 = (1 + z)a = a(1 + z)a−1 .
1+z
Notice that since by induction we have
dk aLog(1+z)
e = a(a − 1) · · · (a − k + 1)(1 + z)a−k .
dz k
Therefore we obtain the Taylor series (centred at 0)
∞
X a(a − 1) · · · (a − k + 1) k
z .
k=0
k!
Notice that the radius of convergence of this series is 1, as we know there are issues for z ∈ (−∞, −1].
The binomial coefficient, for integer values n and k is
!
n n! n(n − 1) · · · (n − k + 1)
= =
k (n − k)!k! k!
which would naturally a radius of convergence R equal to the distance from the point z0 to the half line
{x ≤ −1}, where we have made a brach cut for the Log function. You may ignore the issue of the radius
of convergence for this series for the exam.
ANALYSIS III 78
CHAPTER 7. COMPLEX ANALYSIS
Proof. Assume that |f (z)| ≤ M for all z ∈ C. Let a 6= b be two points in C. Choose R large enough so
that 2 max{|a|, |b|} < R. That means that if we consider w ∈ ∂BR (0), that is |w| = R then
R R
|w − a| > |w − b| > .
2 2
Since f is analytic in C we can use Cauchy’s formula to compute f (a) and f (b) using ∂BR (0) as the curve
γ (of course positively oriented!). We have
ˆ ˆ
1 f (w) 1 f (w)
f (a) − f (b) = dw − dw
2πi ∂BR (0) w − a 2πi ∂BR (0) w − b
ˆ ˆ
1 1 1 a−b f (w)
= f (w) − dw = dw.
2πi ∂BR (0) w−a w−b 2πi ∂BR (0) (w − a)(w − b)
Therefore ˆ
|a − b| M |a − b|4M
|f (a) − f (b)| ≤ 1dw = ,
2π R2 /4 ∂Br (0) R
´
as ∂BR (0 1dw is just the length of the curve, which equals 2πR. Notice that since R is arbitrary (provided
that it is big enough, as indicated above) we can send R to infinity, showing that |f (a) − f (b)| = 0 for
any a and b in C, therefore proving that the function is constant.
Theorem 7.39 (Fundamental Theorem of Algebra). Every non-constant polynomial p on C has a root,
that is, there exists a ∈ C such that p(a) = 0.
Proof. We will prove the result by contradiction. Assume that |p(z)| = 6 0 for every z ∈ C. Define
f : C → C by f (z) = p(z) 1
. Now, since p does not vanish, the function f is analytic in all of C, since it is
the composition of two holomorphic functions (1/z is holomophic outside the origin).
Notice that if we assume p(z) = nk=0 ck z k , with cn 6= 0 (n > 0), then at infinity the polynomial
P
behaves like cn z n , as that is the highest power. That means |p(z)| goes to infinity as z goes to infinity,
and satisfies |p(z)| > 1 for all |z| > R for some R > 0. As a result the function f (z) = p(z) 1
is bounded
in C. It is less than 1 for all |z| > R based on our analysis of p, and it is bounded on the compact set
|z| ≤ R since it is continuous.
Liouville’s Theorem implies that f is in fact constant, which would force p to be constant, which is a
contradiction.
Recall that for a function to be analytic at one point we require that the function be differentiable in
a neighbourhood of the point, and therefore the assumption on Ω being open is natural. Being analytic is
a local property, and requiring that the uniform convergence holds only on compact sets would suffice.
Proof. Let z ∈ Ω. Choose r > 0 sufficiently small so that Br (z) ⊂ Ω. Since fn is analytic in Ω we can
apply Cauchy’s formula to obtain
ˆ
1 fn (w)
fn (z) = dw.
2πi ∂Br (z) w − z
Taking limits as n goes to infinity, and assuming that we can move the limit inside the integral we would
obtain ˆ
1 f (w)
f (z) = dw.
2πi ∂Br (z) w − z
ANALYSIS III 79
CHAPTER 7. COMPLEX ANALYSIS
We have seen before that this implies that f is differentiable (in fact infinitely differentiable) and obtained
an expression for its derivative (see Theorem 7.35). So the only thing left is to justify moving the limit
inside the integral. Notice that this is really a one dimensional integral and we can apply the results learnt
earlier in the year. Taking γ(t) = z + reit for t ∈ [0, 2π), we have γ 0 (t) = ireit and so
ˆ ˆ 2π ˆ 2π
fn (w) fn (z + reit ) it
dw = ire dt = i fn (z + reit )dt. (7.25)
∂Br (z) w−z 0 reit 0
For fixed z, as a function of t we have that fn (z + reit ) converges uniformly to f (z + reit ) and applying
Theorem 2.16 we can move the limit inside the integral, obtaining
ˆ ˆ 2π ˆ 2π
fn (w) it
lim dw = lim i fn (z + re )dt = i f (z + reit )dt.
n→∞ ∂B (z)
r
w−z n→∞ 0 0
Notice that we have (reading the expression (7.25) backwards, now for f instead of for fn )
ˆ 2π ˆ
it f (w)
i f (z + re )dt = dw,
0 ∂Br (z) w−z
obtaining the result.
γ1 is formed by the segment joining −R and R, together with the half circle or radius R. The contour γ2
is a circle centred at i and of radius r < 1. To understand the choice of curves, notice that we can rewrite
the integral as ˆ ∞
1
dx.
−∞ (x − i)(x + i)
ANALYSIS III 80
CHAPTER 7. COMPLEX ANALYSIS
Notice that in the region enclosed by γ2 the function (the integrand extended to a function on C)
1
f (z) :=
(z − i)(z + i)
is analytic except for at z = i. By the deformation of contours Theorem we know that
ˆ ˆ
f (z)dz = f (z)dz,
γ1 γ2
Therefore ˆ ˆ π
R R
f (z)dz ≤ dt = π 2 .
arc 0 R2
−1 R −1
´
As R tends to infinity the arc f (z)dz equals zero. Therefore
ˆ ∞ ˆ ˆ
f (z)dz = f (z)dz = f (z)dz.
−∞ γ1 γ2
Now ˆ ˆ
1 1
f (z)dz = dz.
γ2 ∂Br (i) z+iz−i
Recall that by Cauchy’s formula if g(z) is analytic in the interior of a positively oriented curve then
ˆ
1
g(z) dz = 2πig(a).
γ z−a
which yields ˆ ∞
1
dx = π.
−∞ 1 + x2
and so if we choose a contour similar to the one above (an expanding semi-circle) there will be two
singularities in the interior. We obtain the picture 7.5.
ANALYSIS III 81
CHAPTER 7. COMPLEX ANALYSIS
Now γ1 is built out of the line joining −R and R, together with the semi-circle or radius R centred at
0. γ2 and γ3 correspond to circles centred at e3πi/4 and eπi/4 , oriented clock-wise (positively with respect
to both the blue-shaded and yellow-shaded regions). Notice that with those orientations we have
ˆ ˆ ˆ
1 1 1
4
dz + 4
dz + 4
dz = 0.
γ1 1 + z γ2 1 + z γ3 1 + z
and we have used Cauchy’s formula as g is analytic inside the curve γ2− . Now
1 1
g(e3iπ/4 ) = −iπ/4 −3iπ/4
= √ √ √ √ .
(e3iπ/4 − eiπ/4 )(e3iπ/4 −e )(e3iπ/4 −e ) (− 2)(− 2 + 2i)( 2i)
Now we consider the integral over γ3
ˆ ˆ
1 1
dz = − dz
γ3 1 + z
4
γ3− (z − e
iπ/4 )(z − e3iπ/4 )(z − e−iπ/4 )(z − e−3iπ/4 )
ANALYSIS III 82
CHAPTER 7. COMPLEX ANALYSIS
ˆ
h(z)
=− iπ/4
dz = −2πih(eiπ/4 ),
γ3− (z − e )
where the above defines h as
1
h(z) = ,
(z − e3iπ/4 )(z − e−iπ/4 )(z − e−3iπ/4 )
and we have used Cauchy’s formula as h is analytic inside the curve γ3− . Now
1 1
h(eiπ/4 ) = −iπ/4 −3iπ/4
=√ √ √ √ .
(eiπ/4 − e3iπ/4 )(eiπ/4 −e )(eiπ/4 −e ) 2( 2i)( 2 + 2i)
Since we have ˆ ∞ ˆ ˆ
1 1 1
dz = − d − dz
−∞ 1 + z4 γ2 1 + z4 γ3 1 + z4
we obtain
ˆ ∞
√
1 1 1 π 2
dz = 2πi √ √ √ √ + 2πi √ √ √ √ = .
−∞ 1 + z4 (− 2)(− 2 + 2i)( 2i) 2( 2i)( 2 + 2i) 2
In addition to being able to integrate quotients involving polynomials, we can integration some trigono-
metric functions. For example ˆ ∞
cos(3x)
2
dx.
−∞ 4 + x
As before, ˆ ˆ
e3iz e3iz
dz = dzdz.
γ1 (z − 2i)(z + 2i) γ2 (z − 2i)(z + 2i)
Also ˆ ˆ ˆ
R
e3iz e3iz e3iz
dz = dz + |dz|.
γ1 (z − 2i)(z + 2i) −R (z − 2i)(z + 2i) arc (z − 2i)(z + 2i)
ANALYSIS III 83
CHAPTER 7. COMPLEX ANALYSIS
We consider first the integral over the arc (half circle of radius R). We have, for R >> 4
ˆ ˆ ˆ
e3iz e3iz e−3 Im z πR
dz ≤ 2
|dz| ≤ 2
|dz| ≤ 2 −−−−→ 0,
arc (z − 2i)(z + 2i) arc |z + 4| arc R − 4 R − 4 R→∞
where we have used that along the arc, Im z ≥ 0 and so e−3 Im z ≤ 1. Now for γ2 (remember it is oriented
anti-clockwise)
ˆ ˆ
e3iz g(z)
dz = dz = 2πig(2i),
γ2 (z − 2i)(z + 2i) γ2 z − 2i
where
e3iz
g(z) = dz
z + 2i
and we have used Cauchy’s formula since g is analytic inside γ2 . We have
e−6
g(2i) = .
4i
Therefore
ˆ ∞ ˆ ˆ
e3iz e3iz e3iz π
dz = = = 2πig(2i) = e−6 .
−∞ (z − 2i)(z + 2i) γ1 (z − 2i)(z + 2i) γ2 (z − 2i)(z + 2i) 2
Notice the real part of the integral vanishes as the integrand is odd (hence dividing the i). We consider
the following contours of integration We consider the contours (notice they are both oriented counter-
clockwise) By Cauchy’s Theorem since the integrand is analytic in the region between the curves we
have ˆ ˆ
zeiz zeiz
dz = dz.
γ1 1 + z2 γ2 1 + z2
Now for γ2
ˆ ˆ
zeiz h(z)
dz = dz = 2πih(i),
γ2 (z − i)(z + i) γ2 (z − i)
ANALYSIS III 84
CHAPTER 7. COMPLEX ANALYSIS
for
zeiz
h(z) = ,
z+i
and so ˆ
zeiz ie−1 π
dz = 2πi = i.
γ2 (z − i)(z + i) 2i e
We now consider the integral over γ1 . We look at the integral along the arc.
ˆ ˆ π ˆ π
zeiz Reit e−R sin t+Ri cos t it R2 −R sin t
2
dz = Rie dt ≤ e dt
arc 1 + z 0 1 + R2 e2it 2
0 R −1
As the final example we consider an integrand that has a singularity along the natural path of integration
ˆ ∞
sin x
dx.
−∞ x
since the real part of the integrand is odd. Since the denominator vanishes for a point in the real axis we
need to modify the contours we consider.
The contour is formed of 4 curves, and since the integrand is analytic we know that
ˆ ˆ ˆ ˆ
eiz eiz eiz eiz
dz + dz + dz + dz = 0.
γ1 z γ2 z γ3 z γ4 z
Now for γ1
ANALYSIS III 85
CHAPTER 7. COMPLEX ANALYSIS
Therefore ˆ ˆ
∞ ∞
sin x 1eiz
dx = dz
−∞ x −∞ z
i
"ˆ ˆ # " ˆ ˆ #
1 eiz eiz 1 eiz 1 eiz
= lim lim dz + dz = lim lim − dz − dz = π.
ε→0 R→∞ i γ2 z γ4 z ε→0 R→∞ i γ1 z γ3 i z
ANALYSIS III 86
Bibliography
[1] Rudin, W. Principles of mathematical analysis. Third edition. International Series in Pure and Applied
Mathematics. McGraw-Hill Book Co., 1976.
[5] Gelbaum, B.R., & Olmsted, J.M.H. Counterexamples in Analysis, Holden Day Inc, 1964.
87