Book
Book
to
Real Analysis
Cesar O. Aguilar
May 4, 2022
Contents
1 Preliminaries 1
1.1 Sets, Numbers, and Proofs . . . . . . . . . . . . . . . . . 1
1.2 Mathematical Induction . . . . . . . . . . . . . . . . . . 8
1.3 Functions . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.4 Countability . . . . . . . . . . . . . . . . . . . . . . . . . 19
3 Sequences 69
3.1 Limits of Sequences . . . . . . . . . . . . . . . . . . . . . 69
3.2 Limit Theorems . . . . . . . . . . . . . . . . . . . . . . . 82
3.3 Monotone Sequences . . . . . . . . . . . . . . . . . . . . 91
3.4 Bolzano-Weierstrass Theorem . . . . . . . . . . . . . . . 99
3.5 limsup and liminf . . . . . . . . . . . . . . . . . . . . . . 106
3.6 Cauchy Sequences . . . . . . . . . . . . . . . . . . . . . . 112
3
3.7 Infinite Series . . . . . . . . . . . . . . . . . . . . . . . . 118
5 Continuity 157
5.1 Continuous Functions . . . . . . . . . . . . . . . . . . . . 157
5.2 Combinations of Continuous Functions . . . . . . . . . . 163
5.3 Continuity on Closed Intervals . . . . . . . . . . . . . . . 167
5.4 Uniform Continuity . . . . . . . . . . . . . . . . . . . . . 174
6 Differentiation 181
6.1 The Derivative . . . . . . . . . . . . . . . . . . . . . . . 181
6.2 The Mean Value Theorem . . . . . . . . . . . . . . . . . 191
6.3 Taylor’s Theorem . . . . . . . . . . . . . . . . . . . . . . 196
The material in these notes constitute my personal notes that are used
in the course lectures for MATH 324 and 325 (Real Analysis I, II).
You will find that the lectures and these notes are very closely aligned.
The notes highlight the important ideas and examples that you should
master as a student. You may find these notes useful if:
If you find any typos or errors in these notes, no matter how small,
please email me a short description (with a page number) of the typo/error.
Suggestions and comments on how to improve the notes are also wel-
comed.
Cesar O. Aguilar
SUNY Geneseo
1
Preliminaries
In this short chapter, we will briefly review some basic set notation,
proof methods, functions, and countability. The presentation of these
topics is intentionally brief for two reasons: (1) the reader is likely
familar with these topics, and (2) we include only the necessary material
needed to start doing real analysis.
A ∩ B = {x ∈ S | x ∈ A and x ∈ B}.
1
1.1. SETS, NUMBERS, AND PROOFS
S S
A B A B
A∩B
A∪B
S\A = {x ∈ S | x ∈
/ A},
A × B = {(a, b) | a ∈ A, b ∈ B}.
2
1.1. SETS, NUMBERS, AND PROOFS
(A ∪ B)c = Ac ∩ B c .
N ⊂ Z ⊂ Q.
3
1.1. SETS, NUMBERS, AND PROOFS
4
1.1. SETS, NUMBERS, AND PROOFS
5
1.1. SETS, NUMBERS, AND PROOFS
6
1.1. SETS, NUMBERS, AND PROOFS
Exercises
Exercise 1.1.5. Prove that if x and y are even natural numbers then
xy is even. Do not use proof by contradiction.
7
1.2. MATHEMATICAL INDUCTION
(i) 1 ∈ S
Then S = N.
Proof. Suppose that S is a subset of N with properties (i) and (ii) and
let T = N\S. Proving that S = N is equivalent to proving that T is
the empty set. Suppose then by contradiction that T is non-empty.
By the well-ordering principle of N, T has a smallest element, say it
is a ∈ T . Because S satisfies property (i) we know that 1 ∈ / T and
therefore a > 1. Now since a is the least element of T , then a − 1 ∈ S
(we know that a − 1 > 0 because a > 1). But since S satisfies property
(ii) then (a − 1) + 1 ∈ S, that is, a ∈ S. This is a contradiction because
we cannot have both a ∈ T and a ∈ S. Thus, T is the empty set, and
8
1.2. MATHEMATICAL INDUCTION
therefore S = N.
1 + 2 + 3 + · · · + n = 12 n(n + 1) (1.1)
where n ∈ N. In words, the identity (1.1) says that the sum of all the
integers from 1 to n equals 21 n(n + 1). We use induction to show that
this formula is true for all n ∈ N. Let S be the subset of N consisting
of the natural numbers that satisfy (1.1), that is,
S = {n ∈ N | 1 + 2 + 3 + · · · + n = 21 n(n + 1)}.
If n = 1 then
1
2 n(n + 1) = 12 (1 + 1) = 1.
Thus, 12 n(n + 1) is equal to the sum of all the integers from 1 to n = 1.
Hence, (1.1) is true when n = 1 and thus 1 ∈ S. Now suppose that
some k ∈ N satisfies (1.1), that is, suppose that k ∈ S. Then we may
write that
1 + 2 + · · · + k = 21 k(k + 1). (1.2)
We will prove that the integer k + 1 also satisfies (1.1). To that end,
adding k + 1 to both sides of (1.2) we obtain
1 + 2 + · · · + k + (k + 1) = 12 k(k + 1) + (k + 1).
Now notice that we can factor (k + 1) from the right-hand side and
through some algebraic steps we obtain that
1 + 2 + · · · + k + (k + 1) = 21 k(k + 1) + (k + 1)
= (k + 1)[ 12 k + 1]
= 21 (k + 1)(k + 2).
9
1.2. MATHEMATICAL INDUCTION
(1 + x)k (1 + x) ≥ (1 + kx)(1 + x)
= 1 + (k + 1)x + kx2
≥ 1 + (k + 1)x.
10
1.2. MATHEMATICAL INDUCTION
(i) 1 ∈ S
Then S = N.
11
1.2. MATHEMATICAL INDUCTION
Exercises
12
1.3. FUNCTIONS
1.3 Functions
Let A and B be sets. A function from A to B is a rule that assigns
to each element x ∈ A one element y ∈ B. The set A is called the
domain of f and B is called the co-domain of f . We usually denote
a function with the notation f : A → B, and the assignment of x to y
is written as y = f (x). We also say that f is a mapping from A to B,
or that f maps A into B. The element y assigned to x is called the
image of x under f . The range of f , denoted by f (A), is the set
f (A) = {y ∈ B | ∃ x ∈ A, y = f (x)}.
13
1.3. FUNCTIONS
f (S) = min(S).
Notice that in Example 1.3.5, given any y ∈ N there are many sets
S ∈ P(N) such that f (S) = y. This leads us to the following definition.
14
1.3. FUNCTIONS
f (S) = min(S)
Is f an injection?
f (n) = (2n2, n + 1)
15
1.3. FUNCTIONS
Theorem 1.3.14
If f : A → B and g : B → C are injections (surjections) then the
composition (g ◦ f ) : A → C is an injection (surjection).
16
1.3. FUNCTIONS
Corollary 1.3.15
The composition of two bijections is a bijection.
17
1.3. FUNCTIONS
Exercises
is a bijection.
18
1.4. COUNTABILITY
1.4 Countability
A non-empty set S is said to be finite if there is a bijection from
{1, 2, . . . , n} onto S for some n ∈ N. In this case, we say that S
contains n elements and we write |S| = n. If S is not finite then we say
that S is infinite.
19
1.4. COUNTABILITY
Lk = {(x, y) ∈ N × N | y = −x + k + 1}
20
1.4. COUNTABILITY
ρ(x, y) = 1 + 2 + . . . + (k − 1) + x
1
= (k − 1)k + x
2
1
= (x + y − 2)(x + y − 1) + x.
2
The function ρ is called the Cantor pairing function. Alternatively,
we use a modified version of ρ which we call τ : N × N → N and defined
as (
ρ(x, y), if (x + y − 1) is odd,
τ (x, y) =
ρ(y, x), if (x + y − 1) is even.
We now find a formula for the inverse of τ which we call the Cantor
snake. To write down the formula for τ −1 , we first let for each n ∈ N
p !
−1 + 1 + 8(n − 1)
m = floor
2
and we note that m ≥ 0 is the smallest integer such that 12 m(m+1) < n.
We then set p(n) = n − 21 m(m + 1) a then
(
(p(n), −p(n) + m + 2), if m is even
τ −1 (n) =
(−p(n) + m + 2, p(n)), if m is odd.
We note that the point (x, y) = τ −1(n) is on the line Lk with k = m+1.
The range of τ −1 is
τ −1(N) = {(1, 1), (2, 1), (1, 2), (1, 3), (2, 2), (3, 1), . . .}.
21
1.4. COUNTABILITY
55
10
37 54
9
36 38 53
8
22 35 39 52
7
21 23 34 40 51
6
11 20 24 33 41 50
5
10 12 19 25 32 42 49
4
4 9 13 18 26 31 43 48
3
3 5 8 14 17 27 30 44 47
2
1 2 6 7 15 16 28 29 45 46
1
1 2 3 4 5 6 7 8 9 10
22
1.4. COUNTABILITY
23
1.4. COUNTABILITY
24
1.4. COUNTABILITY
25
1.4. COUNTABILITY
Theorem 1.4.12
The set of rational numbers Q is countable.
Proof. Let Q>0 be the set of all positive rational numbers and let Q<0 be
the set of all negative rational numbers. Clearly, Q = Q<0 ∪ {0} ∪ Q>0,
and thus it is enough to show that Q<0 and Q>0 are countable. In fact,
we have the bijection h : Q>0 → Q<0 given by h(x) = −x, and thus if
we can show that Q>0 is countable then this implies that Q<0 is also
countable. In summary, to show that Q is countable it is enough to
show that Q>0 is countable. To show that Q>0 is countable, consider
the function f : N × N → Q>0 defined as
p
f (p, q) = .
q
26
1.4. COUNTABILITY
C = {a ∈ S | a ∈
/ f (a)}.
Corollary 1.4.14
The set P(N) is uncountable.
27
1.4. COUNTABILITY
Exercises
Exercise 1.4.2. In this exercise you will provide another proof that
N × N is countable.
Q = {{ak }∞
k=1 | ak = 0 or ak = 1} .
a = (0, 0, 0, 0, 0, 0, . . .) ∈ Q
b = (1, 0, 1, 0, 1, 0, . . .) ∈ Q
c = (0, 0, 1, 1, 0, 0, . . .) ∈ Q
d = (1, 1, 1, 1, 1, 1, . . .) ∈ Q
28
1.4. COUNTABILITY
f (a) = {k ∈ N | ak = 1}
f (0, 0, 0, 0, 0, 0, 0, 0, . . .) = ∅
f (1, 0, 1, 0, 1, 0, 1, 0, . . .) = {1, 3, 5, 7, . . .}
f (0, 0, 1, 1, 0, 0, 0, 0, . . .) = {3, 4}
f (1, 1, 1, 1, 1, 1, 1, 1, . . .) = {1, 2, 3, 4, 5, 6, . . .} = N
(c) Combine part (b), Exercise 1.4.1, and Cantor’s theorem to thor-
oughly explain whether Q is countable or uncountable.
29
1.4. COUNTABILITY
30
2
2.1 Introduction
Recall that a rational number is a number x that can be written in
the form x = pq where p, q are integers with q 6= 0. The rational
number system is all you need to accomplish most everyday tasks. For
instance, to measure distances when building a house it suffices to use
1
a tape measure with an accuracy of about 16 of an inch. However, to
do mathematical analysis the rational numbers have some very serious
shortcomings; here is a an example.
Theorem 2.1.1
31
2.1. INTRODUCTION
Thus, both p and q are even, which contradicts the fact that p and q
are relatively prime.
The previous theorem highlights that the set of rational numbers are in
some sense incomplete, or that there are gaps in Q, and that a larger
number system is needed to enlarge the set of math problems that
can be analyzed and solved. Although mathematicians in the 1700s
were using the real number system and resorting to limiting processes
to analyze problems in physics, it was not until the late 1800s that
mathematicians gave a rigorous construction of the real number system.
Motivated by Theorem 2.1.1, we might be tempted to define the real
numbers as the set of solutions of all polynomial equations with integer
coefficients. As it turns out, however, this definition of the reals would
actually leave out almost all the real numbers, including some of our
favorites like π and e. In fact, the set of all numbers that are solutions
to polynomial equations with rational coefficients is countable!
There are two standard ways to construct the set of real numbers.
One standard method to construct R uses the notion of Cauchy se-
quences of rational numbers and is attributed to Georg Cantor [2].
We will cover Cauchy sequences in Section 3.6 and therefore postpone
describing some of the details of the construction until then. The sec-
ond standard method to construct the reals relies on the notion of a
Dedekind cut and is attributed to Richard Dedekind (1831-1916). A
Dedekind cut is a partition {A, B} of Q such that both A and B are
non-empty and
(i) if b ∈ A and a < b then a ∈ A, and
32
2.1. INTRODUCTION
ℝ
⋯ -4 -3 -2 -1 0 1 2 3 4 ⋯
The essential feature that we want to capture with this view of R is that
there are no “holes” in the real number system. This view of R allows us
to quickly start learning the properties of R instead of focusing on the
details of constructing a model for R. Naturally, the rational numbers
Q are a subset of R and we say that a real number x ∈ R is irrational
if it is not rational. As we saw in Thereom 2.1.1, the positive number
x ∈ R such that x2 = 2 is irrational.
33
2.1. INTRODUCTION
Exercises
Exercise 2.1.2. Prove that if 0 < x < 1 then xn < x for all natural
numbers n ≥ 2. Do not assume that x is rational.
34
2.2. ALGEBRAIC AND ORDER PROPERTIES
Definition 2.2.1
A field is a set F with two binary operations + : F × F → F
and × : F × F → F, the former called addition and the latter
multiplication, satisfying the following properties:
(iii) a × (b + c) = a × b + a × c
Example 2.2.2. It is not hard to see that N and Z are not fields. In
each case, what property of a field fails to hold?
35
2.2. ALGEBRAIC AND ORDER PROPERTIES
Theorem 2.2.4
Let a, b, c ∈ R.
a ≤ b ≤ c.
36
2.2. ALGEBRAIC AND ORDER PROPERTIES
a + b ≤ x + y ≤ ζ + ξ.
Theorem 2.2.7
Let x ∈ R be non-negative, that is, x ≥ 0. If for every ε > 0 it
holds that x < ε then x = 0.
Proof. We prove the contrapositive, that is, we prove that if x > 0 then
there exists ε > 0 such that ε < x. Assume then that x > 0 and let
ε = x2 . Then ε > 0 and clearly ε < x.
The next few examples will give us practice with working with in-
equalities.
37
2.2. ALGEBRAIC AND ORDER PROPERTIES
Exercises
38
2.3. THE ABSOLUTE VALUE
Theorem 2.3.1
Let a, b ∈ R and c ≥ 0.
(ii) |a|2 = a2
Proof. Statements (i) and (ii) are trivial. To prove (iii), first suppose
that |a| ≤ c. If a > 0 then a ≤ c. Hence, −a ≥ −c and since a > 0
then a > −a ≥ −c. Hence, −c ≤ a ≤ c. If a < 0 then −a ≤ c, and
thus a ≥ −c. Since a < 0 then a < −a ≤ c. Thus, −c ≤ a ≤ c. Now
suppose that −c ≤ a ≤ c. If 0 < a ≤ c then |a| = a ≤ c. If a < 0
then from multiplying the inequality by (−1) we have c ≥ −a ≥ −c
and thus |a| = −a ≤ c.
39
2.3. THE ABSOLUTE VALUE
To prove part (iv), notice that |a| ≤ |a| and thus applying (iii) we
get −|a| ≤ a ≤ |a|.
a |a|
Example 2.3.2. If b 6= 0 prove that b = |b| .
Example 2.3.3. From Theorem 2.3.1 part (i), we have that |a2 | =
|a · a| = |a| · |a| = |a|2 . Therefore, |a2 | = |a|2 . Similarly, one can show
that |a3 | = |a|3 . By induction, for each n ∈ N it holds that |an | = |a|n .
|x + y| ≤ |x| + |y|.
−|x| ≤ x ≤ |x|
−|y| ≤ y ≤ |y|
and thus
|x + y| ≤ |x| + |y|.
40
2.3. THE ABSOLUTE VALUE
Corollary 2.3.5
For any x1, x2, . . . , xn ∈ R it holds that
Corollary 2.3.6
For x, y ∈ R it holds that
|x − y| = |x + (−y)|
≤ |x| + | − y|
= |x| + |y|.
|x| = |x − y + y| ≤ |x − y| + |y|
41
2.3. THE ABSOLUTE VALUE
|f (x)| = |2x2 − 3x + 7|
≤ |2x2| + |3x| + |7|
= 2|x|2 + 3|x| + 7
≤ 2(2)2 + 3(2) + 7
= 21.
2
Example 2.3.9. Let f (x) = 2x 1−2x
+3x+1
. Find a number M > 0 such that
|f (x)| ≤ M for all 2 ≤ x ≤ 3.
42
2.3. THE ABSOLUTE VALUE
numerator:
2x2 + 3x + 1
|f (x)| =
1 − 2x
|2x2 + 3x + 1|
=
|1 − 2x|
2|x|2 + 3|x| + 1|
=
|1 − 2x|
2 · 32 + 3 · 3 + 1
≤
|1 − 2x|
28
= .
|2x − 1|
28 28
|f (x)| ≤ ≤ .
|1 − 2x| 3
28
Hence, we can take M = 3.
43
2.3. THE ABSOLUTE VALUE
|x − y| = |y − x|
ℝ
y x
Bε (a) = {x ∈ R | |x − a| < ε} = (a − ε, a + ε)
44
2.3. THE ABSOLUTE VALUE
3n + 1 3
|f (n) − 23 | = −
2n + 3 2
6n + 2 − 6n − 9
=
2(2n + 3)
7
=
2(2n + 3)
7
<ε
2(2n + 3)
which after re-arranging can be written as
7 3
n> − .
4ε 2
With ε = 0.0001 we obtain
Example 2.3.13. Let ε1 > 0 and ε2 > 0, and let a ∈ R. Show that
Bε1 (a) ∩ Bε2 (a) and Bε1 (a) ∪ Bε2 (a) are ε-neighborhoods of a for some
appropriate value of ε.
45
2.3. THE ABSOLUTE VALUE
Exercises
Exercise 2.3.1. Prove that if a < x < b and a < y < b then |x − y| <
b−a. Draw a number line with points a, b, x, y satisfying the inequalities
and graphically interpret the inequality |x − y| < b − a.
f (x) = a0 + a1 x + a2 x2 + · · · + an xn .
Prove that
|f (x)| ≤ f (|x|)
for all x ∈ R. Hint: For example, if say f (x) = 2 + 3x2 + 7x3 then you are asked
to prove that
|2 + 3x2 + 7x3 | ≤ 2 + 3|x|2 + 7|x|3 .
| {z } | {z }
|f (x)| f (|x|)
46
2.4. THE COMPLETENESS AXIOM
47
2.4. THE COMPLETENESS AXIOM
(i) S = [0, 1]
(ii) S = (−∞, 3)
(iii) N = {1, 2, 3, . . . , }
(v) S = {x ∈ R | x2 < 2}
48
2.4. THE COMPLETENESS AXIOM
As you will see in the pages that follow, The Completeness Axiom is
the key notion upon which the theory of real analysis depends on.
49
2.4. THE COMPLETENESS AXIOM
(b) S = [0, ∞)
(c) S = (−∞, 3)
The Completeness Axiom is sometimes called the supremum prop-
erty of R or the least upper bound property of R. The Complete-
ness property makes R into a complete ordered field. The following
example shows that Q does not have the completeness property.
Example 2.4.8. The set of rational numbers is an ordered field but it
is not complete. Consider the set S = {x ∈ Q | x2 < 2}. By definition,
S ⊂ Q. Clearly, S is bounded above, for example u = 10 is an upper
√
bound of S, but the least upper bound of S is u = 2 which is not
a rational number. Therefore S ⊂ Q does not have a supremum in Q
and therefore Q does not have the Completeness property. From the
point of view of analysis, this is the main distinction between Q and R
(note that both are ordered fields).
In some cases, it is obvious what sup and inf are, however, to do
analysis rigorously, we need systematic ways to determine sup(S) and
inf(S). To start with, we first need to be a bit more rigorous about
what it means to be the least upper bound, or at least have a more
concrete description that we can work with, i.e., using inequalities.
The following lemma does that and, as you will observe, it is simply a
direct consequence of the definition of the supremum.
Lemma 2.4.9
Let S ⊂ R be non-empty and suppose that u ∈ R is an upper bound
of S. Then u is the least upper bound of S if and only if for any
ε > 0 there exists x ∈ S such that
u − ε < x ≤ u.
50
2.4. THE COMPLETENESS AXIOM
cA = {y ∈ R | ∃ x ∈ A s.t. y = cx}.
51
2.4. THE COMPLETENESS AXIOM
sup(cA) = c sup(A).
52
2.4. THE COMPLETENESS AXIOM
−A = {y ∈ R | ∃ x ∈ A s.t. y = −x}.
53
2.4. THE COMPLETENESS AXIOM
c + A = {y ∈ R | ∃ x ∈ A s.t. y = c + x}.
(b) Prove that inf(AB) = inf(A) · inf(B). Hint: Consider two cases,
when say inf(A) inf(B) = 0 and when inf(A) inf(B) 6= 0.
54
2.4. THE COMPLETENESS AXIOM
A
sup(A)
sup B
= .
inf(B)
Proof. Since A is bounded above, sup(A) exists and x ≤ sup(A) for all
x ∈ A. Since B is bounded below, inf(B) exists and inf(B) ≤ y for all
A
y ∈ B. Let z = x/y be an arbitrary point in B . Then since inf(B) ≤ y
and y > 0 and inf(B) > 0 we obtain that
1 1
≤ .
y inf(B)
Since x ≤ sup(A) and x > 0 (and then clearly sup(A) > 0) we obtain
x x sup(A)
≤ ≤ .
y inf(B) inf(B)
sup(A) A
This proves that z ≤ inf(B) and since z ∈ B was arbitrary we have
sup(A) A
proved that inf(B) is an upper bound of the set B. This proves that
55
2.4. THE COMPLETENESS AXIOM
A
sup( B ) exists. Moreover, by the definition of the supremum, we also
have that
A
sup(A)
sup B ≤ . (2.5)
inf(B)
A A
Now, z ≤ sup( B ) for all z ∈ B and thus xy ≤ sup( BA
) for all x ∈ A and
A
all y ∈ B. If y is held fixed then x ≤ y · sup( B ) for all x ∈ A and thus
A A
sup(A) ≤ y · sup( B ). Therefore, sup(A)/ sup( B ) ≤ y, which holds for
A
all y ∈ B. Therefore, sup(A)/ sup( B ) ≤ inf(B) and consequently
sup(A) A
≤ sup B
. (2.6)
inf(B)
56
2.4. THE COMPLETENESS AXIOM
Exercises
x + A = {y ∈ R | ∃ a ∈ A s.t. y = x + a}.
A + B = {z ∈ R | ∃ x ∈ A, ∃ y ∈ B s.t. z = x + y}.
Prove that A+B is bounded above and that sup(A+B) = sup A+sup B.
Exercise 2.4.6. Let R>0 denote the set of all positive real numbers
and let A, B ⊂ R>0 be bounded. Assume that inf(B) > 0. Define the
set
A n x
o
= z ∈ R | ∃ x ∈ A, ∃ y ∈ B s.t. z = y .
B
Prove that
A
sup(A)
sup B = .
inf(B)
57
2.5. APPLICATIONS OF THE SUPREMUM
58
2.5. APPLICATIONS OF THE SUPREMUM
Corollary 2.5.5
If S = { n1 | n ∈ N} then inf(S) = 0.
Proof. Since 0 < n1 for all n then 0 is a lower bound of S. Now suppose
that 0 < y. By the Archimedean Property, there exists n ∈ N such that
1
y
< n and thus n1 < y. Hence, y is not a lower bound of S. Therefore
0 is the greatest lower bound of S, that is, inf(S) = 0.
59
2.5. APPLICATIONS OF THE SUPREMUM
Corollary 2.5.6
1
For any y > 0 there exists n ∈ N such that n < y.
Corollary 2.5.7
Given y > 0 there exists n ∈ N such that n − 1 ≤ y < n.
Proof. We first prove the claim for the case that 0 < x < y. Suppose
that y − x > 1 and thus x + 1 < y. There exists m ∈ N such that
m − 1 ≤ x < m and thus m ≤ x + 1. Therefore,
x<m≤x+1<y
60
2.5. APPLICATIONS OF THE SUPREMUM
then take r = 0 and if x < y < 0 then apply the previous arguments to
0 < −y < −x.
Hence, between any two distinct real numbers there is a rational num-
ber. This implies that any irrational number can be approximated by
a rational number to within any degree of accuracy.
√ √ √
Proof. We have that 2x < 2y. By the Density Theorem, 2x <
√ √
r < 2y for some r ∈ Q. Then ξ = r/ 2.
61
2.5. APPLICATIONS OF THE SUPREMUM
Exercises
for any x ∈ D. If f (D) and g(D) are bounded above prove that (f +
g)(D) is also bounded above and that
62
2.6. NESTED INTERVAL THEOREM
[a, b] := {x ∈ R | a ≤ x ≤ b}
which we call the closed interval from a to b. The following are called
half-open (or half-closed) intervals:
[a, b) := {x ∈ R | a ≤ x < b}
(a, ∞) = {x ∈ R | x > a}
(−∞, b) = {x ∈ R | x < b}
[a, ∞) = {x ∈ R | x ≥ a}
(−∞, b] = {x ∈ R | x ≤ b}.
Theorem 2.6.1
Let S ⊂ R contain at least two points. Suppose that if x, y ∈ S and
x < y then [x, y] ⊂ S. Then S is an interval.
I1 ⊇ I2 ⊇ I3 ⊇ I4 ⊇ · · ·
63
2.6. NESTED INTERVAL THEOREM
Proof. Since In is a closed interval, we can write In = [an , bn] for some
an , bn ∈ R and an ≤ bn for all n ∈ N. The nested property can be
written as
[a1 , b1] ⊇ [a2, b2] ⊇ [a3 , b3] ⊇ [a4, b4] ⊇ · · ·
Since [an, bn ] ⊆ [a1 , b1] for all n ∈ N then an ≤ b1 for all n ∈ N.
Therefore, the set S = {an | n ∈ N} is bounded above. Let ξ = sup(S)
64
2.6. NESTED INTERVAL THEOREM
Theorem 2.6.3
Let In = [an , bn] be a sequence of nested closed bounded intervals.
If
inf{bn − an | n ∈ N} = 0
T
then nn=1 In is a singleton set.
Example 2.6.4. Let In = 1 − n1 , 1 + n1 for n ∈ N.
(a) Prove that I1 , I2, I3, . . . is a sequence of nested intervals.
T
(b) Find ∞ n=1 In .
65
2.6. NESTED INTERVAL THEOREM
S = {0.a1a2 a3 a4 · · · ∈ R | ak = 0 or ak = 1, k ∈ N}.
x = 0.000000 . . .
x = 0.101010 . . .
x = 0.100000 . . .
x = 0.010100 . . .
66
2.6. NESTED INTERVAL THEOREM
f (0.a1a2 a3 · · · ) = {k ∈ N | ak = 1}.
In other words, f (x) consists of the decimal places in the decimal ex-
pansion of x that have a value of 1. For example,
f (0.000000 . . .) = ∅
f (0.101010 . . .) = {1, 3, 5, 7, . . .}
f (0.100100 . . .) = {1, 4}
f (0.011000 . . .) = {2, 3}
67
2.6. NESTED INTERVAL THEOREM
Exercises
T
Exercise 2.6.1. Let In = 0, n1 for n ∈ N. Prove that ∞
n=1 In = {0}.
68
3
Sequences
69
3.1. LIMITS OF SEQUENCES
xn+1 = xn−1 + xn , n ≥ 2.
Then
x3 = 2, x4 = 3, x5 = 5, . . .
The range of a sequence (xn) is the set
{xn | n ∈ N},
that is, the usual range of a function. However, the range of a sequence
is not the actual sequence (the range is a set and a sequence is a func-
tion). For example, if X = (1, 2, 3, 1, 2, 3, . . .) then the range of X is
{1, 2, 3}. If xn = sin( nπ
2 ) then the range of (xn ) is {1, 0, −1}.
Many concepts in analysis can be described using the long-term or
limiting behavior of sequences. In calculus, you undoubtedly developed
techniques to compute the limit of basic sequences (and hence show
convergence) but you might have omitted the rigorous definition of the
convergence of a sequence. Perhaps you were told that a given sequence
(xn) converges to L if as n → ∞ the values xn get closer to L. Although
this is intuitively sound, we need a more precise way to describe the
meaning of the convergence of a sequence. Before we give the precise
definition, we will consider an example.
70
3.1. LIMITS OF SEQUENCES
Example 3.1.1. Consider the sequence (xn) whose nth term is given
by xn = 3n+2
n+1 . The values of (xn ) for several values of n are displayed
in Table 3.1.
n xn
1 2.50000000
2 2.66666667
3 2.75000000
4 2.80000000
5 2.83333333
50 2.98039216
101 2.99019608
10,000 2.99990001
1,000,000 2.99999900
2,000,000 2.99999950
3n+2
Table 3.1: Values of the sequence xn = n+1
The above data suggests that the values of the sequence (xn) become
closer and closer to the number L = 3. For example, suppose that
ε = 0.005 and consider the ε-neighborhood of L = 3, that is, the
interval (3 − ε, 3 + ε) = (2.995, 3.005). Not all the terms of the sequence
(xn) are in the ε-neighborhood, however, it seems that all the terms of
the sequence from x101 and onward are inside the ε-neighborhood. In
other words, IF n ≥ 101 then 3 − ε < xn < 3 + ε, or equivalently
|xn − 3| < ε. Suppose now that ε = 0.00001 and thus the new ε-
neighborhood is (3−ε, 3+ε) = (2.99999, 3.000001). Then it is no longer
true that |xn − 3| < ε for all n ≥ 101. However, it seems that all the
terms of the sequence from x1000000 and onward are inside the smaller
ε-neighborhood, in other words, |xn − 3| < ε for all n ≥ 1, 000, 000.
We can extrapolate these findings and make the following hypothesis:
For any given ε > 0 there exists a natural number K ∈ N such that if
n ≥ K then |xn − L| < ε.
71
3.1. LIMITS OF SEQUENCES
The above example and our analysis motivates the following defini-
tion.
Hence, xn converges to L if for any given ε > 0 (no matter how small),
there exists a point in the sequence xK such that |xK − L| < ε, |xK+1 −
L| < ε, |xK+2 − L| < ε, . . ., that is, |xn − L| < ε for all n ≥ K. We will
sometimes write lim xn = L simply as lim xn = L or (xn) → L.
n→∞
1
|xn − 0| = n
−0
1
=
n
1
≤
K
< ε.
72
3.1. LIMITS OF SEQUENCES
73
3.1. LIMITS OF SEQUENCES
3
Proof. Let xn = 4nn3+3n
+6 . We want to show that for any given ε > 0,
there exists K ∈ N such that if n ≥ K then
4n3 + 3n
|xn − 4| = − 4 < ε.
n3 + 6
4n3 + 3n
|xn − 4| = −4
n3 + 6
3n − 24
=
n3 + 6
3n − 24 3n + 24
≤
n3 + 6 n3 + 6
27n
≤
n3 + 6
27n
<
n3
27
= 2.
n
27
Hence, if n2< ε then also |xn − 4| < ε by the transitivity property
q of
inequalities. The inequality n272 < ε holds true if and only if 27
ε < n.
Now that we have done a detailed preliminary analysis, we can proceed
with the proof.
74
3.1. LIMITS OF SEQUENCES
q
Suppose that ε > 0 is given and let K ∈ N be such that 27
ε
< K.
Then 27 2 27 27 27
ε < K , and thus K 2 < ε. Then, if n ≥ K then n2 ≤ K 2 and
therefore
3n − 24
|xn − 4| =
n3 + 6
3n + 24
≤
n3 + 6
27n
≤ 3
n +6
27n
< 3
n
27
= 2
n
27
≤ 2
K
< ε.
75
3.1. LIMITS OF SEQUENCES
cos(n)
Example 3.1.7. Let xn = n2 −1 where n ≥ 2. Prove that limn→∞ xn =
0.
Proof. We want to prove given any ε > 0 there exists K ∈ N such that
cos(n)
< ε, n ≥ K.
n2 − 1
Now, since | cos(x)| ≤ 1 for all x ∈ R we have that
cos(n) | cos(n)|
=
n2 − 1 n2 − 1
1
≤ 2
n −1
1
≤ 2 1 2
n − 2n
2
= 2.
n
2 cos(n)
Thus, if n2
< ε then n2 −1
.
q
Let ε > 0 be arbitrary. Let K ∈ N be such that 2ε < K. Then
2
K2
< ε. Therefore, if n ≥ K then n22 ≤ K22 and therefore
cos(n) | cos(n)|
=
n2 − 1 n2 − 1
1
≤ 2
n −1
1
≤ 2 1 2
n − 2n
2 2
= 2≤ 2
n K
<ε
76
3.1. LIMITS OF SEQUENCES
Theorem 3.1.9
Let (xn) be a sequence and let L ∈ R. Let (an ) be a sequence of
positive numbers such that lim an = 0. Suppose that there exists
n→∞
M ∈ N such that
|xn − L| ≤ an , ∀ n ≥ M.
Then lim xn = L.
n→∞
|xn − L| ≤ an < ε.
77
3.1. LIMITS OF SEQUENCES
1
Now since limn→∞ nx = 0 then it follows by Theorem 3.1.9 that
lim rn = 0.
n→∞
n2 −1
Example 3.1.11. Consider the sequence xn = 2n2 +3 . Prove that
limn→∞ xn = 12 .
n2 − 1 1 5 1
− =
2n2 + 3 2 2 (2n2 + 3)
5/2
< 2.
2n
Using the definition of the limit of a sequence, one can show that
limn→∞ 4n5 2 = 0 and therefore limn→∞ xn = 12 .
Proof. Suppose that (xn) → L1 and that (xn) → L2. Let ε > 0 be
arbitrary. Then there exists K1 such that |xn − L1 | < ε/2 for all
n ≥ K1 and there exists K2 such that |xn − L2| < ε/2 for all n ≥ K2.
Let K = max{K1, K2}. Then for n ≥ K it holds that |xn − L1| < ε/2
78
3.1. LIMITS OF SEQUENCES
Hence, |L1 − L2| < ε for all ε > 0, and therefore by Theoreom 2.2.7 we
conclude that |L1 − L2| = 0, that is, L1 − L2 = 0.
79
3.1. LIMITS OF SEQUENCES
Exercises
Exercise 3.1.1. Write the first three terms of the recursively defined
sequence x1 = 1, xn+1 = 21 (xn + x2n ) for n ≥ 1.
Exercise 3.1.3.
(a) Prove that limn→∞ |xn | = 0 if and only if limn→∞ xn = 0.
(b) Combining the previous result and Example 3.1.10, prove that if
1 < r < 0 then limn→∞ rn = 0.
(c) Conclude that for any real number r ∈ R, if |r| < 1 then limn→∞ rn =
0.
80
3.1. LIMITS OF SEQUENCES
81
3.2. LIMIT THEOREMS
Example 3.2.2. Prove that (xn) is bounded if and only if there exists
numbers R1 and R2 such that R1 ≤ xn ≤ R2 for all n ∈ N.
82
3.2. LIMIT THEOREMS
Let ε > 0. There exists K1 such that |xn − L| < ε/2 for n ≥ K1
and there exists K2 such that |yn − M| < ε/2 for n ≥ K2. Let K =
max{K1, K2}. Then for n ≥ K
83
3.2. LIMIT THEOREMS
= ε.
(iii) It is enough to prove that y1n → M1 and then use (ii). Now,
since M 6= 0 and yn 6= 0 then |yn | is bounded below by some positive
number, say R > 0. Indeed, (|yn|) → |M| and |yn | > 0. Thus, |y1n | < R1
for all n ∈ N. Now,
1 1 1
− = |yn − M|
yn M |yn ||M|
1
< |yn − M|.
R|M|
For ε > 0, there exists K ∈ N such that |yn − M| < R|M|ε for all
84
3.2. LIMIT THEOREMS
1
< R|M|ε
R|M|
= ε.
Corollary 3.2.6
Theorem 3.2.7
Suppose that (xn) → L. If xn ≥ 0 for all n ∈ N then L ≥ 0.
Proof. We prove the contrapositive, that is, we prove that if L < 0 then
there exists K ∈ N such that xK < 0. Suppose then that L < 0. Let
ε > 0 be such that L + ε < 0. Since (xn) → L, there exists K ∈ N such
that xK < L + ε, and thus by transitivity we have xK < 0.
85
3.2. LIMIT THEOREMS
Corollary 3.2.9
Suppose that a ≤ xn ≤ b and lim xn = L. Then a ≤ L ≤ b.
n→∞
86
3.2. LIMIT THEOREMS
L − ε < xn < L + ε
Remark 3.2.11. Some people call the Squeeze Theorem the Sandwich
Theorem; we are not those people.
Example 3.2.12. Let 0 < a < b and let xn = (an + bn )1/n. Prove that
lim xn = b.
n→∞
Thus, xK+1 < rxK , and therefore xK+2 < rxK+1 < r2 xK , and induc-
tively for m ≥ 1 it holds that
xK+m < rm xK .
87
3.2. LIMIT THEOREMS
Hence, the tail of the sequence (xn) given by (ym ) = (xK+1, xK+2, . . . , )
satisfies
0 < ym < rm xK .
Since 0 < r < 1 it follows that limm→∞ rm = 0 and therefore limm→∞ ym =
0 by the Squeeze theorem. This implies that (xn) converges to 0
also.
88
3.2. LIMIT THEOREMS
Exercises
Exercise 3.2.1. Use the Limit Theorems to prove that if (xn) converges
and (xn + yn ) converges then (yn ) converges. Give an example of two
sequences (xn) and (yn ) such that both (xn) and (yn) diverge but (xn +
yn ) converges.
Exercise 3.2.4. Show that if (xn) and (yn ) are sequences such that
(xn) and (xn + yn ) are convergent, then (yn) is convergent.
(a) Divergent sequences (xn) and (yn ) such that zn = xnyn converges.
(b) Divergent sequences (xn) and (yn ) such that zn = xnyn diverges.
Exercise 3.2.6. Let (xn) and (yn ) be sequences and suppose that (xn)
converges to L. Assume that for every ε > 0 there exists M ∈ N such
that |xn − yn | < ε for all n ≥ M. Prove that (yn) also converges to L.
89
3.2. LIMIT THEOREMS
Exercise 3.2.9. Apply the Limit Theorems to find the limits of the
following sequences:
r
2n2 + 3
(a) xn =
n2 + 1
(b) xn = (2 + 1/n)2
n+1
(c) xn = √
n n
(d) xn = 2n/n!
90
3.3. MONOTONE SEQUENCES
91
3.3. MONOTONE SEQUENCES
92
3.3. MONOTONE SEQUENCES
L = f (L).
But xn+1 = f (xn) and therefore limn→∞ xn+1 = f (L), which is equiva-
lent to limn→∞ xn = f (L) since (xn+1) is just the 1-tail of the sequence
(xn). Therefore, L = f (L) as claimed. From the equation L = f (L) we
can solve for L if possible.
93
3.3. MONOTONE SEQUENCES
2 n−1 1 − rn
1 + r + r + ··· + r =
1−r
and if 0 < r < 1 then 0 < rn < r < 1 and therefore
1 − rn 1
< .
1−r 1−r
94
3.3. MONOTONE SEQUENCES
95
3.3. MONOTONE SEQUENCES
√ √
Then xk+1 = 2 + xk < 2 + 2 = 2. This proves that (xn) is bounded
above. By the MCT, (xn) converges, say to L. Moreover, since xn ≥ 0
√
(as can be proved by induction), then L ó 0. Therefore, L = 2 + L
and then L2 − L − 2 = 0. Hence, L = 1± 21+8 = 1±3
2 . Since L ≥ 0 then
L = 2.
n
Example 3.3.8. Consider the sequence xn = 1 + n1 . We will show
that (xn) is bounded and increasing, and therefore by the MCT (xn)
convergent. The limit of this sequence is the number e. From the
binomial theorem
n
X n 1
xn =
k nk
k=0
n 1 n(n − 1) 1 n(n − 1)(n − 2)
=1+ + + + ···
n 2! n2 3! n3
1 n(n − 1)(n − 2) · · · (n − (n − 1))
+
n! nk
1 1
= 1+1+ (1 − n1 ) + (1 − n1 )(1 − n1 ) + · · ·
2! 3!
1
+ (1 − n1 )(1 − n1 ) · · · (1 − (n − 1)/n)
n!
1 1 1
< 1+1+ + + ··· +
2! 3! n!
1 1 1
< 1+1+ + 2 + · · · + n−1
2 2 2
1 − (1/2)n
=1+
1 − 1/2
<3
where we used that 2n−1 < n! for all n ≥ 3. This shows that (xn ) is
96
3.3. MONOTONE SEQUENCES
And similarly,
n+1 1 1 2 k−1
= (1 − n+1 )(1 − n+1 ) · · · (1 − n+1 ).
k (n + 1)k
j
It is clear that (1 − nj ) < (1 − n+1 ) for all 1 ≤ j ≤ n. Hence, nk n1k <
n+1
1
k (n+1)k
. Therefore, xn < xn+1, that is, (xn) is increasing. By the
MCT, (xn) converges to sup{xn : n ∈ N}.
97
3.3. MONOTONE SEQUENCES
Exercises
98
3.4. BOLZANO-WEIERSTRASS THEOREM
k ≤ nk
for all k ≥ 1.
99
3.4. BOLZANO-WEIERSTRASS THEOREM
1 1 1
( 11 , 303 , 2000 , . . .)
100
3.4. BOLZANO-WEIERSTRASS THEOREM
Theorem 3.4.5
If (xn) → L then every subsequence of (xn) converges to L.
Proof. Let ε > 0. Then there exists K ∈ N such that |xn − L| < ε for
all n ≥ K. Since nK ≥ K and nK < nK+1 < . . ., then |xnk − L| < ε for
all k ≥ K.
Theorem 3.4.6
Let (xn) be a sequence.
The following is a very neat result that will supply us with a very
short proof of the main result of this section, namely, the Bolzano-
Weierstrass Theorem.
Theorem 3.4.7
Every sequence has a monotone subsequence.
Proof. Let (xn) be an arbitrary sequence. We will say that the term xm
is a peak if xm ≥ xn for all n ≥ m. In other words, xm is a peak if it is
an upper bound of all the terms coming after it. There are two possible
cases for (xn), either it has an infinite number of peaks or it has a finite
number of peaks. Suppose that it has an infinite number of peaks, say
xm1 , xm2 , . . ., and we may assume that m1 < m2 < m3 < · · · . Then,
101
3.4. BOLZANO-WEIERSTRASS THEOREM
102
3.4. BOLZANO-WEIERSTRASS THEOREM
It turns out that if we had taken as our starting axiom the Bolzano-
Weierstrass theorem then we could prove the Completeness property
and then of course the MCT. In other words, all three statements are
equivalent:
Completeness ⇐⇒ MCT ⇐⇒ Bol-Wei
103
3.4. BOLZANO-WEIERSTRASS THEOREM
Exercises
(b) xn = sin(nπ/4)
(a) Suppose that (xn) is increasing. Prove that if (xn) has a sub-
sequence (xnk ) that is bounded above then (xn) is also bounded
above.
(b) Suppose that (xn) is decreasing. Prove that if (xn) has a sub-
sequence (xnk ) that is bounded below then (xn) is also bounded
below.
104
3.4. BOLZANO-WEIERSTRASS THEOREM
Exercise 3.4.6. Suppose that x1 and y1 satisfy 0 < x1 < y1 and define
√ xn + yn
xn+1 = xnyn , yn+1 =
2
for n ≥ 1. Prove that (xn) and (yn) are convergent and that lim xn =
√
lim yn . (Hint: First show that xy ≤ (x + y)/2 for any x, y > 0.)
Exercise 3.4.7. Let (xn) be a sequence and define the sequence (yn)
as
x1 + x2 + · · · + xn
yn =
n
Prove that if (xn) → L then (yn) → L.
105
3.5. LIMSUP AND LIMINF
S = {L ∈ R | (xnk ) → L} .
Example 3.5.2. Here are six examples of sequences and the corre-
sponding subsequences limit set. Notice that in the cases where (xn)
(xn) S
1 1
(1, 2 , 3 , . . .) {0}
(1, −1, 1, −1, . . .) {1, −1}
(1, 2, 3, 4, . . .) ∅
3 1 5 1 7
(1, 2 , 3 , 4 , 5 , 6 , . . .) {0, 1}
1 1 3 1 7
(0, 1, 2 , 4 , 4 , 8 , . . . , 8 , . . .) [0, 1]
(rn) enumeration of Q R
Table 3.2: Limits of subsequences
106
3.5. LIMSUP AND LIMINF
In general, we have seen that for a general set S, sup(S) and inf(S)
are not necessarily in S. This, however, is not the case for the subse-
quences limit set.
Lemma 3.5.3
Let (xn) be a bounded sequence and let S be its subsequences limit
set. Then sup(S) ∈ S and inf(S) ∈ S. In other words, there exists a
subsequence (xnk ) of (xn) such that lim xnk = sup(S) and similarly
k→∞
there exists a subsequence (ynk ) of (xn) such that lim ynk = inf(S).
k→∞
107
3.5. LIMSUP AND LIMINF
Definition 3.5.4
Let (xn) be a bounded sequence and let S be its subsequences limit
set. We define the limit superior of (xn) as
By Lemma 3.5.3, lim sup xn is simply the largest limit of all convergent
subsequences of (xn) while lim inf xn is the smallest limit of all con-
vergent subsequences of (xn). Notice that by definition it is clear that
lim inf xn ≤ lim sup xn . The next theorem gives an alternative charac-
terization of lim sup xn and lim inf xn. The idea is that lim sup xn is a
sort of limiting supremum and lim inf xn is a sort of limiting infimum
of a bounded sequence (xn).
Theorem 3.5.5
Let (xn) be a bounded sequence and let L∗ ∈ R. The following are
equivalent:
(ii) If ε > 0 then there are at most finitely many xn such that
L∗ + ε < xn and infinitely many xn such that L∗ − ε < xn .
108
3.5. LIMSUP AND LIMINF
(ii)→(iii) Let ε > 0. Since L∗ + ε/2 < xm holds for finitely many
m, there exists M such that xm ≤ L∗ + ε/2 for all m ≥ M. Hence,
L∗ + ε/2 is an upper bound of {xn | n ≥ m} and thus um < L∗ + ε.
Since (um) is decreasing, we have that um < L∗ + ε for all m ≥ M.
Now, L∗ − ε/2 < xn holds infinitely often and thus L∗ − ε < um for all
m ∈ N. Hence, L∗ − ε < um < L∗ + ε for all m ≥ M. This proves the
claim.
109
3.5. LIMSUP AND LIMINF
lim inf(pn+1 − pn ) = 2
In other words, the Twin Prime Conjecture is that there are infinitely
many pairs of twin primes.
Theorem 3.5.7
Let (xn) be a bounded sequence and let L ∈ R. If every convergent
subsequence of (xn) converges to L then (xn) converges to L.
Proof. Suppose that (xn) does not converge to L. Then, there exists
ε > 0 such that for every K ∈ N there exists n ≥ K such that |xn −L| ≥
ε. Let K1 ∈ N. Then there exists n1 ≥ K1 such that |xn1 − L| ≥ ε.
Then there exists n2 > n1 + 1 such that |xn2 − L| ≥ ε. By induction,
there exists a subsequence (xnk ) of (xn) such that |xnk − L| ≥ ε for all
k ∈ N. Now (xnk ) is bounded and therefore by Bolzano-Weierstrass
has a convergent subsequence, say (zk ), which is also a subsequence
of (xn). By assumption, (zk ) converges to L, which contradicts that
|xnk − L| ≥ ε for all k ∈ N.
110
3.5. LIMSUP AND LIMINF
Exercises
Exercise 3.5.1. Determine the lim sup xn and lim inf xn for each case:
n→∞ n→∞
(b) xn = 1 + sin(nπ/2)
(c) xn = (2 − 1/n)(−1)n
Exercise 3.5.2. Let (xn) and (yn) be bounded sequences. Let (zn ) be
the sequence zn = xn + yn . Show that
Exercise 3.5.3. Let (xn) and (yn ) be bounded sequences. Show that
if xn ≤ yn for all n then
111
3.6. CAUCHY SEQUENCES
Lemma 3.6.3
If (xn) is convergent then it is a Cauchy sequence.
112
3.6. CAUCHY SEQUENCES
113
3.6. CAUCHY SEQUENCES
Example 3.6.6. Let 0 < r < 1 and suppose that |xn − xn+1| ≤ rn
1
for all n ∈ N. Using the fact that 1 + r + r2 + · · · + rk < 1−r for all
n
r
k ∈ N prove that if m > n then |xn − xm | ≤ 1−r . Deduce that (xn) is a
Cauchy sequence.
114
3.6. CAUCHY SEQUENCES
Notice that the main result used in the Cauchy Criterion is the
Bolzano–Weierstrass (B–W) theorem. We therefore have the following
chain of implications:
Theorem 3.6.8
Suppose that every Cauchy sequence in R converges to a number in
R and the Archimedean Property holds in R. Then R satisfies the
Completeness property, that is, every non-empty bounded above
subset of R has a least upper bound in R.
115
3.6. CAUCHY SEQUENCES
the least upper bound of S. First of all, for fixed x ∈ S we have that
x < bn for all n ∈ N and therefore x ≤ lim bn, that is, u is an upper
bound of S. Since an is not an upper bound of S, there exists xn ∈ S
such that an < xn < bn and therefore by the Squeeze theorem we have
u = lim xn . Given an arbitrary ε > 0 then there exists K ∈ N such
that u − ε < xK and thus u − ε is not an upper bound of S. This proves
that u is the least upper bound of S.
116
3.6. CAUCHY SEQUENCES
Exercises
Exercise 3.6.5. Suppose that (xn) and (yn ) are sequences such that
|xm − xn| ≤ |ym − yn | for all n, m ∈ N. Show that if the sequence (yn)
is convergent then so is the sequence (xn).
Exercise 3.6.6. Suppose that 0 < r < 1. Show that if the sequence
(xn) satisfies |xn − xn−1| < rn−1 for all n ≥ 2 then (xn ) is a Cauchy
sequence and therefore convergent. Hint: If m > n then
117
3.7. INFINITE SERIES
x1 + x2 + x3 + · · ·
s1 = 1
s2 = 1 + ( 32 ) = 1.6666
s3 = 1 + ( 23 ) + ( 23 )2 = 2.1111
s4 = 1 + ( 32 ) + ( 23 )2 + ( 23 )3 = 2.4074
s5 = 1 + ( 23 ) + ( 23 )2 + ( 23 )3 + ( 32 )4 = 2.6049
s6 = 1 + ( 32 ) + ( 23 )2 + ( 23 )3 + ( 32 )4 + ( 23 )5 = 2.7366
s7 = 1 + ( 32 ) + ( 23 )2 + ( 23 )3 + ( 32 )4 + ( 23 )5 + ( 32 )6 = 2.8244
s8 = 1 + ( 32 ) + ( 23 )2 + ( 23 )3 + ( 32 )4 + ( 23 )5 + ( 32 )6 + ( 23 )7 = 2.8829
s9 = 1 + ( 32 ) + ( 23 )2 + ( 23 )3 + ( 32 )4 + ( 23 )5 + ( 32 )6 + ( 23 )7 + ( 32 )8 = 2.9219
s10 = 1 + ( 32 ) + ( 23 )2 + ( 23 )3 + ( 32 )4 + ( 23 )5 + ( 32 )6 + ( 23 )7 + ( 32 )8 + ( 23 )9 = 2.9479
118
3.7. INFINITE SERIES
20
X
s20 = ( 23 )k−1 = 2.999097813 . . .
k=1
50
X
s50 = ( 23 )k−1 = 2.999999994 . . .
k=1
100
X
s100 = ( 32 )k−1 = 2.999999998 . . . .
k=1
∞
X
( 32 )n−1 = 3 = lim sn .
n→∞
n=1
119
3.7. INFINITE SERIES
Definition 3.7.1
Let (xn) be a sequence. The infinite series generated by (xn) is
the sequence (sn) defined by
sn = x1 + x2 + · · · + xn
or recursively,
s1 = x1
sn+1 = sn + xn+1, n ≥ 1.
Example 3.7.2. Let (xn) be the sequence xn = n3 . The first few terms
of the sequence of partials (sn) is
s1 = x1 = 3
3 9
s2 = x1 + x2 = 3 + 2 = 2
9 11
s3 = x1 + x2 + x3 = 2 +1= 2
11 3 25
s4 = s3 + x4 = 2 + 4 = 4
sn+1 = sn + xn+1 ≥ sn
120
3.7. INFINITE SERIES
Solution. We compute:
s1 = x1 = −1
s2 = x1 + x2 = −1 + 1 = 0
s3 = s2 + x3 = 0 − 1 = −1
to converge.
P∞
The notation n=1 xn is therefore a compact way of writing
n
X
lim xk ,
n→∞
k=1
121
3.7. INFINITE SERIES
P
and the question of whether the series ∞ n=1 xn converges is really about
whether the limit limn→∞ sn exists. Often we will write a series such as
P∞ P
n=1 x n simply as xn when either the initial value of n is understood
or is unimportant. Sometimes, the initial n value may be n = 0, n = 2,
or some other n = n0 .
and is called the geometric series. The nth term of the sequence of
partial sums is
sn = 1 + r + r2 + · · · + rn .
If r = 1 then limn→∞ sn does not exist (why?), so suppose that r 6= 1.
n+1
Using the fact that (1 − r)(1 + r + r2 + · · · + rn ) = 1−r
1−r we can write
1 − rn+1
sn = .
1−r
Now if |r| < 1 then limn→∞ rn = 0, while if |r| > 1 then limn→∞ rn does
not exist. Therefore, if |r| < 1 then
1 − rn+1 1
lim sn = lim = .
n→∞ n→∞ 1 − r 1−r
Therefore,
∞
X 1
rn = lim sn = .
n=0
n→∞ 1−r
P
In summary: The series ∞ n
n=0 r is called the geometric series and
1
converges if and only if |r| < 1 and in this case converges to 1−r .
122
3.7. INFINITE SERIES
P (−1)n 2n
Example 3.7.6. Consider the series ∞ n=0 3n . The series can be
P∞ −2 n
written as n=0 3 and thus it is a geometric series with r = − 23 .
Since |r| = | − 23 | < 1, the series converges and it converges to
∞
X (−1)n2n 1 3
= =
n=0
3n 1 − (−2/3) 5
0.999999 . . . = 1.
= 1.
123
3.7. INFINITE SERIES
P∞ 1 1
Example 3.7.8 (Telescoping Series). Consider n=1 n(n+1) = 1·2 +
1 1
2·3 + 3·4 + · · · . Using partial fraction decomposition
1 1 1
= − .
n(n + 1) n n + 1
Therefore, the nth term of the sequence of partial sums (sn) is
n n
X 1 X 1 1
sn = = −
k(k + 1) k k+1
k=1 k=1
This is a telescoping sum because all terms in the middle cancel and
only the first and last remain. For example:
1
s1 = 1 − 2
s2 = 1 − + 12 −
1
2
1
3 =1− 1
3
1
s3 = 1 − + 12 − + 13 −
2
1
3
1
4
= 1 − 41 .
1
By induction one can show that sn = 1− n+1 and therefore limn→∞ sn =
P
1. Therefore, the given series converges and it converges to ∞ 1
n=1 n(n+1) =
limn→∞ sn = 1.
P
Example 3.7.9 (Harmonic Series). Consider the series ∞ 1
n=1 n = 1 +
1 1 1
2 + 3 + 4 + · · · . We are going to analyze a subsequence (snk ) of the
sequence of partial sums (sn ). We will show that (snk ) is unbounded
and thus (sn ) is unbounded and therefore divergent. Consequently, the
P
series ∞ 1
n=1 n is divergent. Consider
s4 = 1 + 21 + 1
3 + 1
4 > 1 + 12 + 1
2 = 1 + 2 21 .
Now consider
s8 = s4 + 51 + 16 + 17 + 1
8 > 1 + 2 21 + 4 18 = 1 + 3 12 .
124
3.7. INFINITE SERIES
Lastly consider,
s16 = s8 + 91 + 1
10 + ··· + 1
16 > 1 + 3 12 + 8 16
1
= 1 + 4 21 .
P
Proof. By definition, if xn converges then the sequence of partial
P
sums (sn ) is convergent. Suppose then that L = limn→∞ sn = xn .
Recall that (sn ) has the recursive definition sn+1 = sn + xn+1. The
sequences (sn+1) and (sn ) both converge to L and thus
125
3.7. INFINITE SERIES
P∞ 3n+1
Example 3.7.11. The series n=1 2n+5 is divergent because limn→∞ 3n+1
2n+5 =
3
2 > 1.
P
Example 3.7.13. A certain series ∞ k=1 xk has sequence of partial sums
2n
(sn ) whose general nth term is sn = n+1 .
P
(a) What is 10 k=1 xk ?
P
(b) Does the series ∞ k=1 xk converge? If yes, what does it converge
to? Explain.
(c) Does the sequence (xn) converge? If yes, what does it converge
to? Explain.
126
3.7. INFINITE SERIES
Theorem 3.7.15
Suppose that (xn) is a sequence of non-negative terms, that is, xn ≥
P
0 for all n ∈ N. Then xn converges if and only if (sn ) is bounded.
In this case,
X∞
xn = sup{sn | n ≥ 1}.
n=1
P
Proof. Clearly, if xn = limn→∞ sn exists then (sn ) is bounded. Now
suppose that (sn ) is bounded. Since xn ≥ 0 for all n ∈ N then sn+1 =
sn + xn+1 ≥ sn and thus sn+1 ≥ sn shows that (sn ) is an increasing
sequence. By the Monotone convergence theorem, (sn) converges and
limn→∞ sn = sup{sn | n ≥ 1}.
P
Example 3.7.16. Consider the series ∞ 1 1
n=1 n2 . Since n2 > 0, to prove
that the series converges it is enough to show that the sequence of
partial sums (sn ) is bounded. We will consider a subsequence of (sn ),
namely, (snk ) where nk = 2k − 1 for k ≥ 2. We have
1 1
s3 = 1 ++
22 32
1 1
<1+ 2 + 2
2 2
1
=1+
2
and
1 1 1 1
s7 = s3 + + + +
42 52 62 72
1 4
<1+ + 2
2 4
1 1
= 1 + + 2.
2 2
127
3.7. INFINITE SERIES
1 1 1
snk < 1 + + 2 + · · · + k−1
2 2 2
1 1 1
snk < 1 + + 2 + · · · + k−1
2 2 2
∞
X 1
< n
= 2.
n=0
2
Theorem 3.7.17
P P
Suppose that xn and yn are convergent series.
P P
(i) Then (xn + yn ) and (xn − yn ) are also convergent and
X X X
(xn ± yn ) = xn ± yn .
P P
(ii) For any constant c ∈ R, cxn is also convergent and cxn =
P
c xn .
128
3.7. INFINITE SERIES
n n n
!
X X X
lim (xk ± yk ) = lim xk ± yk
n→∞ n→∞
k=1 k=1 k=1
n
X n
X
= lim xk ± lim yk
n→∞ n→∞
k=1 k=1
∞
X ∞
X
= xn ± yn .
n=1 n=1
Hence,
∞
X ∞
X ∞
X
(xn ± yn ) = xn ± yn .
n=1 n=1 n=1
n
X n
X ∞
X
lim cxk = c lim xk = c xn .
n→∞ n→∞
k=1 k=1 n=1
Therefore,
∞
X ∞
X
cxn = c xn .
n=1 n=1
129
3.7. INFINITE SERIES
P P
Proof. Let tn = nk=1 xn and sn = nk=1 yk be the sequences of partial
P
sums. Since xn ≤ yn then tn ≤ sn . To prove (a), if yn converges then
(sn ) is bounded. Thus, (tn ) is also bounded. Since (tn ) is increasing and
P
bounded it is convergent by the MCT. To prove (b), if xn diverges
then (tn ) is necessarily unbounded and thus (sn ) is also unbounded and
therefore (sn ) is divergent.
P∞ n2 2
(a) First compute limn→∞ 3nn2 +n = 1/3. Therefore, by
n=1 3n2 +n :
the divergence test, the series diverges.
P∞ P
(b) 1
n=1 np where p ≥ 2: We know that ∞ 1
n=1 n2 is convergent. Now,
P
if p ≥ 2 then n1p ≤ n12 . Therefore by the Comparison test, ∞ 1
n=1 np
is also convergent. This is called a p-series and actually converges
for any p > 1.
130
3.7. INFINITE SERIES
P∞ n+7
(c) n=1 n3 +3 : We will use the comparison test. We have
n+7 n + 7n
≤
n3 + 3 n3 + 3
8n
< 3
n
8
= 2.
n
P P∞ n+7
The series ∞ 8
n=1 n2 converges and thus the original series n=1 n3 +3
also converges.
P
Example 3.7.21. Suppose that 0 ≤ xn ≤ 1 for all n ∈ N. If xn
P
converges prove that (xn)2 also converges. Does the claim hold if we
only assume that xn > 0?
P
Solution. Since 0 ≤ xn ≤ 1 then 0 ≤ x2n ≤ xn. Since xn converges
P
then by the Comparison test then (xn)2 also converges. More gener-
P
ally, suppose that xn > 0 and xn converges. Then (xn) converges to
zero and thus there exists K ∈ N such that 0 < xn < 1 for all n ≥ K.
P P∞ 2
Then since the series ∞ x
n=1 n+K converges it follows that n=1 xn+K
P∞ 2
converges and consequently n=1 xn converges.
The following two tests can be used for series whose terms are not
necessarily non-negative.
0 ≤ xn + |xn | ≤ 2|xn |.
131
3.7. INFINITE SERIES
P P
If |xn| converges then so does 2|xn|. By the Comparison test
P
3.7.18, the series (xn + |xn |) converges also. Then the following series
is a difference of two converging series and therefore converges:
X X X
(xn + |xn |) − |xn | = xn
132
3.7. INFINITE SERIES
Exercises
P
Exercise 3.7.1. Suppose that xn is a convergent series. Is it true
P P
that if yn is divergent then (xn + yn ) is divergent? If it is true,
prove it, otherwise give an example to show that it is not true.
Exercise 3.7.3. Using only the tests derived in this section, determine
whether the given series converge or diverge:
∞
X 2n2 + 3
(a) √
n=1
n2 + 3n + 2
∞
X cos(nπ)3n
(b)
n=1
2n
∞
X n−3
(c)
n=1
n3 + 1
Exercise 3.7.4. Suppose that (xn) and (yn ) are non-negative sequences.
P P P
Prove that if xn and yn are convergent then xnyn is convergent.
P
(Hint: Recall that if zn ≥ 0 then zn converges iff (sn ) is bounded,
Pn
where sn = k=1 zk is the sequence of partial sums. Alternatively, use
the identity (x + y)2 = x2 + 2xy + y 2 .)
P∞
Exercise 3.7.5. Suppose that xn > 0 for all n ∈ N and n=1 xn
P √
converges. You will prove that ∞n=1 xn xn+1 converges.
P∞
(a) Let yn = xn + xn+1. Prove that n=1 yn converges.
133
3.7. INFINITE SERIES
√
(b) Using the fact that (a+b)2 = a2 +2ab+b2 , prove that ab ≤ a+b
if a, b > 0. Deduce that
√
xnxn+1 ≤ xn + xn+1 = yn .
P∞ √
(c) Deduce that n=1 xnxn+1 converges.
∞
X 1
(c)
3k
k=−2
Exercise 3.7.8. Using the fact that 2k−1 < k! for all k ≥ 3, prove that
P
the series ∞ 1
n=0 n! converges.
134
3.7. INFINITE SERIES
for all k ∈ N.
P∞ n
P∞
(b) Conclude that if n=0 2 X(2n ) converges then n=1 X(n).
135
3.7. INFINITE SERIES
136
4
Limits of Functions
137
4.1. LIMITS OF FUNCTIONS
Lemma 4.1.2
A point c is a cluster point of A if and only if there exists a sequence
(xn) in A such that xn 6= c and lim xn = c.
n→∞
Example 4.1.3. Below are some examples of cluster points for a given
set:
138
4.1. LIMITS OF FUNCTIONS
|L − L′ | ≤ |f (x) − L| + |f (x) − L′ |
< ε/2 + ε/2
= ε.
139
4.1. LIMITS OF FUNCTIONS
Thus, given ε > 0 we let δ = ε/5 and thus if 0 < |x − c| < δ then
|f (x) − 13| < ε. Thus, by definition, limx→2 f (x) = 13.
x+1
Example 4.1.7. Consider the function f (x) = x2 +3 with domain A =
R. Prove that
1
lim f (x) = .
x→1 2
140
4.1. LIMITS OF FUNCTIONS
x2 − 2x + 1
=
2(x2 + 3)
|x − 1|2
=
2(x2 + 3)
< |x − 1|2.
√
Let ε > 0 be arbitrary and let δ = ε. Then if 0 < |x − 1| < δ then
|x − 1|2 < δ 2 = ε. Hence, if 0 < |x − 1| < δ then
x+1 1
|f (x) − 21 | = 2
−
x +3 2
< |x − 1|2
< ε.
Thus, by definition, limx→1 f (x) = 21 .
141
4.1. LIMITS OF FUNCTIONS
We now need to analyze how large |x| can become when x is say within
δ > 0 of c. To be concrete, suppose that δ = 1/2. Hence, if 0 < |x−c| <
δ then
|x| = |x − c + c|
≤ |x − c| + |c|
< δ + |c|
< 1 + |c|.
|f (x) − c2 | = |x2 − c2 |
= |x + c||x − c|
≤ (|x| + |c|) · δ
ε
< (1 + |c|) ·
1 + |c|
= ε.
142
4.1. LIMITS OF FUNCTIONS
x2 − 3x
|f (x) − 2| = −2
x+3
x2 − 5x − 6
=
x+3
(x + 1)(x − 6)
=
(x + 3)
|x + 1|
= |x − 6|.
|x + 3|
x2 − 3x
|f (x) − 2| = −2
x+3
|x + 1|
= |x − 6|
|x + 3|
<1·δ
≤ ε.
143
4.1. LIMITS OF FUNCTIONS
Proof. If x ∈ Q then
and if x ∈
/ Q then
3(x − 1)
|f (x)| =
1 + x2
3|x − 1|
=
1 + x2
≤ 3|x − 1|.
Therefore, for all x ∈ R it holds that |f (x)| ≤ 3|x − 1| since π/2 < 3.
Thus, given ε > 0 let δ = ε/3 and thus if 0 < |x − 1| < δ then
|f (x)| ≤ 3|x − 1|
<3·δ
= 3 · ε/3
= ε.
144
4.1. LIMITS OF FUNCTIONS
145
4.1. LIMITS OF FUNCTIONS
Corollary 4.1.12
Let f : A → R be a function and let c be a cluster point of A and
let L ∈ R. Then f does not converge to L at c if and only if there
exists a sequence (xn) in A converging c, with xn 6= c for all n ∈ N,
and such that (f (xn)) does not converge to L.
Corollary 4.1.13
Let f : A → R be a function and let c be a cluster point of A.
Suppose that (xn) and (yn ) are sequences in A converging to c, with
xn 6= c and yn 6= c for all n ∈ N. If f (xn) and f (yn) converge but
146
4.1. LIMITS OF FUNCTIONS
147
4.1. LIMITS OF FUNCTIONS
Exercises
x2 − 3x
(b) lim =2
x→6 x + 3
(c) lim |x − 3| = 1
x→4
148
4.1. LIMITS OF FUNCTIONS
Hint: The Density Theorem will be helpful for (b). In particular, the
Density Theorem implies that for any point c ∈ R, there exists a se-
quence (xn) of rational numbers such that (xn) → c, and that there
exists a sequence (yn) of irrational numbers such that (yn ) → c.
Exercise 4.1.5. Use any applicable theorem to explain why the fol-
lowing limits do not exist.
149
4.2. LIMIT THEOREMS
Theorem 4.2.2
Consider a function f : A → R and let c be a cluster point of A. If
lim f (x) exists then f is bounded locally at c.
x→c
Proof. Let L = limx→c f (x) and let ε > 0 be arbitrary. Then there
exists δ > 0 such that |f (x) − L| < ε for all x ∈ A such that 0 <
|x − c| < δ. Therefore, for all x ∈ A and 0 < |x − c| < δ we have that
|f (x)| = |f (x) − L + L|
≤ |f (x) − L| + |L|
< ε + |L|.
150
4.2. LIMIT THEOREMS
We now state and prove some limit laws for functions. Let f, g :
A → R be functions and define the functions (f + g), (f − g), f g, and
f /g on A as follows:
(f g)(x) = f (x)g(x)
f f (x)
(x) =
g g(x)
where for f /g we require that g(x) 6= 0 for all x ∈ A.
151
4.2. LIMIT THEOREMS
The proofs are left as an exercises. (To prove the results, use the
sequential criterion for limits and the limits laws for sequences).
Corollary 4.2.5
Let f1, . . . , fk : A → R be functions and let c be a cluster point of
A. If limx→c fi (x) exists for each i = 1, 2, . . . , k then
k
X k
X
(i) lim fi(x) = lim fi(x)
x→c x→c
i=1 i=1
k
Y k
Y
(ii) lim fi(x) = lim fi (x)
x→c x→c
i=1 i=1
f (x) f (c)
lim = .
x→c g(x) g(c)
x2 − 4
Example 4.2.7. Prove that lim = 4.
x→2 x − 2
152
4.2. LIMIT THEOREMS
Theorem 4.2.8
Let f : A → R be a function and let c be a cluster point of A.
Suppose that f has limit L at c. If f (x) ≥ 0 for all x ∈ A then
L ≥ 0.
Theorem 4.2.9
Let f : A → R be a function and let c be a cluster point of A.
Suppose that M1 ≤ f (x) ≤ M2 for all x ∈ A and suppose that
limx→c f (x) = L. Then M1 ≤ L ≤ M2 .
153
4.2. LIMIT THEOREMS
Theorem 4.2.12
Let f : A → R be a function and let c be a cluster point of A.
Suppose that limx→c f (x) = L. If L > 0 then there exists δ > 0
such that f (x) > 0 for all x ∈ (c − δ, c + δ), x 6= c.
154
4.2. LIMIT THEOREMS
155
4.2. LIMIT THEOREMS
Exercises
156
5
Continuity
157
5.1. CONTINUOUS FUNCTIONS
Notice that here (xn) is allowed to take on the value c. The following
is immediate.
Theorem 5.1.3
The function f is discontinuous at c if and only if there exists a
sequence (xn) in A converging to c but f (xn) does not converge to
f (c).
158
5.1. CONTINUOUS FUNCTIONS
Proof. Let c be irrational and let ε = 1/2. Then for all δ > 0, there
exists x ∈ Q ∩ (c − δ, c + δ) (by the Density theorem) and therefore
|f (x) − f (c)| = 1 > ε. Hence, f is discontinuous at c. A similar
argument shows that f is discontinuous c ∈ Q. Alternatively, if c ∈ Q
then there exists a sequence of irrational numbers (xn) converging to c.
Now f (xn) = 0 and f (c) = 1, and this proves that f is discontinuous
at c. A similar arguments holds for c irrational.
159
5.1. CONTINUOUS FUNCTIONS
0.5
0.4
0.3
0.2
160
5.1. CONTINUOUS FUNCTIONS
161
5.1. CONTINUOUS FUNCTIONS
Exercises
162
5.2. COMBINATIONS OF CONTINUOUS FUNCTIONS
Theorem 5.2.1
Let f, g : A → R be continuous functions at c ∈ A and let b ∈ R.
Then
163
5.2. COMBINATIONS OF CONTINUOUS FUNCTIONS
have that
Corollary 5.2.2
Let f, g : A → R be continuous functions on A and let b ∈ R. Then
164
5.2. COMBINATIONS OF CONTINUOUS FUNCTIONS
Proof. Let ε > 0 be arbitrary. Then there exists δ > 0 such that
|f (y) − f (d)| < ε for all 0 < |y − d| < δ. Therefore, if 0 < |x − c| =
|(x + α) − (c + α)| < δ then |f (x + α) − f (c + α)| < ε and therefore
Example 5.2.7. Prove that sin(x) and cos(x) are continuous every-
where.
1
≤ 2 |x − c|
2
= |x − c|.
165
5.2. COMBINATIONS OF CONTINUOUS FUNCTIONS
√
Example 5.2.9. Prove that f (x) = x is continuous on A = {x ∈
R | x ≥ 0}.
√ √ √
Proof. For c = 0, we must consider | x − 0| = x. Given ε > 0 let
√
δ = ε2. Then if x ∈ A and x < δ = ε2 then x < ε. This shows that f
is continuous at c = 0. Now suppose that c 6= 0. Then
√ √
√ √ √ √ x+ c
| x − c| = | x − c| · √ √
x+ c
|x − c|
=√ √
x+ c
1
≤ √ |x − c|.
c
√ √ √
Hence, given ε > 0, suppose that 0 < |x − c| < cε. Then | x − c| <
ε.
Proof. Let ε > 0 be given. Let c ∈ A and let d = f (c) ∈ B. Then there
exists δ1 > 0 such that if 0 < |y − d| < δ1 then |g(y) − g(d)| < ε. Now
since f is continuous at c, there exists δ2 > 0 such that if 0 < |x−c| < δ2
166
5.3. CONTINUITY ON CLOSED INTERVALS
then |f (x)−f (c)| < δ1. Therefore, if 0 < |x−c| < δ2 then |f (x)−d| < δ1
and therefore |g(f (x))−g(d)| < ε. This proves that (g ◦f ) is continuous
at c ∈ A. Since c is arbitrary, (g ◦ f ) is continuous on A.
Corollary 5.2.12
If f : A → R is continuous then g(x) = |f (x)| is continuous. If
p
f (x) ≥ 0 for all x ∈ A then h(x) = f (x) is continuous.
Definition 5.3.1
We say that f : A → R is bounded on A if there exists M > 0
such that |f (x)| ≤ M for all x ∈ A.
Theorem 5.3.3
Let f : A → R be a continuous function. If A = [a, b] is a closed
and bounded interval then f is bounded on A.
167
5.3. CONTINUITY ON CLOSED INTERVALS
1
Example 5.3.5. The function f (x) = x is continuous on A = (0, 1].
168
5.3. CONTINUITY ON CLOSED INTERVALS
169
5.3. CONTINUITY ON CLOSED INTERVALS
does not achieve a maximum value on the closed interval [−1, 1].
Corollary 5.3.10
Let f : [a, b] → R be a function and suppose that f (a)f (b) < 0. If
f is continuous then there exists c ∈ (a, b) such that f (c) = 0.
170
5.3. CONTINUITY ON CLOSED INTERVALS
begins his trek down the mountain at 7:00 am, takes the same trail
down as he did going up, and arrives at the base at 7:00 pm. Use the
Intermediate Value Theorem to show that there is a point on the path
that the hiker crossed at exactly the same time of day on both days.
Proof. Let f (t) be the distance traveled along the trail on the way up
the mountain after t units of time, and let g(t) be the distance remaining
to travel along the trail on the way down the mountain after t units of
time. Both f and g are defined on the same time interval, say [0, 12] if
t is measured in hours. If M is the length of the trail, then f (0) = 0,
f (12) = M, g(0) = M and g(12) = 0. Let h(t) = g(t) − f (t). Then
h(0) = M and h(12) = −M. Hence, there exists t∗ ∈ (0, 12) such that
h(t∗ ) = 0. In other words, f (t∗) = g(t∗), and therefore t∗ is the time of
day when the hiker is at exactly the same point on the trail.
Example 5.3.12. Prove by the Intermediate Value Theorem that f (x) =
xex − 2 has a root in the interval [0, 1].
Proof. The function f is continuous on [0, 1]. We have that f (0) =
−2 < 0 and f (1) = e − 2 > 0. Therefore, there exists x∗ ∈ (0, 1) such
that f (x∗) = 0, i.e., f has a zero in the interval (0, 1).
Theorem 5.3.13
If f : [a, b] → R is continuous then f ([a, b]) = [min(f ), max(f )].
Proof. Since f achieves its maximum and minimum value on [a, b], there
exists x∗, x∗ ∈ [a, b] such that f (x∗) ≤ f (x) ≤ f (x∗) for all x ∈ [a, b].
171
5.3. CONTINUITY ON CLOSED INTERVALS
Hence, f ([a, b]) ⊂ [f (x∗), f (x∗)]. Assume for simplicity that x∗ < x∗.
Then [x∗, x∗] ⊂ [a, b]. Let ξ ∈ R be such that f (x∗) < ξ < f (x∗). Then
by the Intermediate Value Theorem, there exists c ∈ (x∗, x∗) such that
ξ = f (c). Hence, ξ ∈ f ([a, b]), and this shows that [f (x∗), f (x∗)] ⊂
f ([a, b]). Therefore, f ([a, b]) = [f (x∗), f (x∗)] = [min(f ), max(f )].
It is worth noting that the previous theorem does not say that
f ([a, b]) = [f (a), f (b)].
172
5.3. CONTINUITY ON CLOSED INTERVALS
Exercises
173
5.4. UNIFORM CONTINUITY
Hence, given any ε > 0 we can set δ = ε/2 and then |x − c| < δ implies
that |f (x) − f (c)| < ε. The punchline is that δ = ε/2 will work for any
c. These motivating examples lead to the following definition.
174
5.4. UNIFORM CONTINUITY
Proof. We have that |f (x)−f (c)| = |kx−kc| = |k||x−c|. Hence, for any
ε > 0 we let δ = ε/|k|, and thus if |x−c| < δ then |f (x)−f (c)| < ε.
Hence, for ε > 0 let δ = ε and if |x−c| < δ then | sin(x)−sin(c)| < ε.
1
Example 5.4.4. Show that f (x) = is uniformly continuous on
1 + x2
R.
1 + c2 − 1 − x2
=
(1 + x2 )(1 + c2 )
x+c
= |x − c|
(1 + x2 )(1 + c2 )
x c
= + |x − c|.
(1 + x2 )(1 + c2 ) (1 + x2)(1 + c2 )
|x|
Now, |x| ≤ 1 + x2 implies that 1+x2
≤ 1 and therefore
|x| 1
≤ ≤ 1.
(1 + x2)(1 + c2 ) 1 + c2
175
5.4. UNIFORM CONTINUITY
It follows that |f (x) − f (c)| ≤ 2|x − c|. Hence, given ε > 0 we let
δ = ε/2, and if |x − c| < δ then |f (x) − f (c)| < ε.
Theorem 5.4.5
Let f : A → R be a function. The following are equivalent:
(ii) There exists ε0 > 0 such that for every δ > 0 there exists
x, u ∈ A such that |x − u| < δ but |f (x) − f (c)| ≥ ε0.
(iii) There exists ε0 > 0 and two sequences (xn) and (un) such that
lim (xn − un) = 0 and |f (xn) − f (un)| ≥ ε0.
n→∞
Example 5.4.6. Let f (x) = x1 and let A = (0, ∞). Let xn = n1 and let
1
un = n+1 . Then lim(xn − un ) = 0. Now |f (xn) − f (un)| = | − 1| = 1.
Hence, if ε = 1/2 then |f (xn) − f (un)| > ε. This proves that f is not
uniformly continuous on A = (0, ∞).
176
5.4. UNIFORM CONTINUITY
exists ε > 0 such that for δn = 1/n, there exists xn, un ∈ [a, b] such
that |xn − un| < δn but |f (xn) − f (un)| ≥ ε. Clearly, lim(xn − un ) = 0.
Now, since a ≤ xn ≤ b, by the Bolzano-Weierstrass theorem there is a
subsequence (xnk ) of (xn) that converges to a point z ∈ [a, b]. Now,
and therefore also limk→∞ unk = z. Hence, both (xnk ) and (unk ) are
sequences in [a, b] converging to z but |f (xnk ) − f (unk )| ≥ ε. Hence
f (xnk ) and f (unk ) do not converge to the same limit and thus f is not
continuous at z. This completes the proof.
π π
= sin πn + 2
+ 16n
π
= cos πn + 16n
π π
= cos(nπ) cos( 16n ) − sin(nπ) sin( 16n )
π
= cos( 16n )
π
≥ cos( 16 )
177
5.4. UNIFORM CONTINUITY
Definition 5.4.9
A function f : A → R is called a Lipschitz function on A if there
exists a constant K > 0 such that |f (x) − f (u)| ≤ K|x − u| for all
x, u ∈ A.
178
5.4. UNIFORM CONTINUITY
179
5.4. UNIFORM CONTINUITY
Exercises
180
6
Differentiation
f (x) − f (c)
− L < ε.
x−c
181
6.1. THE DERIVATIVE
f (x) − f (c)
If f is differentiable at c, we will denote lim by f ′ (c), that
x→c x−c
is,
f (x) − f (c)
f ′ (c) = lim .
x→c x−c
The rule that sends c to the number f ′ (c) defines a function on a pos-
sibly smaller subset J ⊂ I. The function f ′ : J → R is called the
derivative of f .
Example 6.1.2. Let f (x) = 1/x for x ∈ (0, ∞). Prove that f ′ (x) =
− x12 .
x−c
x+c
sin(x) − sin(c) = 2 sin 2 cos 2
= 1 · cos(c) = cos(c).
x
Example 6.1.4. Prove by definition that f (x) = 1+x2
is differentiable
on R.
182
6.1. THE DERIVATIVE
x(1 + c2 ) − c(1 + x2 )
=
(1 + x2)(1 + c2 )(x − c)
1 − cx
= .
(1 + c2 )(1 + x2)
Now
f (x) − f (c) 1 − c2
lim = .
x→c x−c (1 + c2 )2
Hence, f ′ (c) exists for all c ∈ R and the derivative function of f is
1 − x2
f ′(x) = .
(1 + x2 )2
Solution. If x > 0 then f (x) = x and thus f ′ (x) = 1 for x > 0. If x < 0
then f (x) = −x and therefore f ′(x) = −1 for x < 0. Now consider
c = 0. We have that
f (x) − f (c) |x|
= .
x−c x
We claim that the limit limx→0 |x|
x
does not exist and thus f ′ (0) does not
exist. To see this, consider xn = 1/n. Then (xn) → 0 and f (xn) = 1
183
6.1. THE DERIVATIVE
for all n. On the other hand, consider yn = −1/n. Then (yn ) → 0 and
f (yn) = −1. Hence, limn→∞ f (xn) 6= limn→∞ f (yn), and thus the claim
holds by the Sequential criterion for limits. The derivative function f ′
of f is therefore defined on A = R\{0} and is given by
(
1, x>0
f ′(x) =
−1, x > 0.
184
6.1. THE DERIVATIVE
Proof. Parts (i) and (ii) are straightforward. We will prove only (iii)
and (iv). For (iii), we have that
185
6.1. THE DERIVATIVE
186
6.1. THE DERIVATIVE
Now
g(y) − g(f (c))
lim h(y) = lim
y→f (c) y→f (c) y−c
= g ′ (f (c))′
= h(f (c)).
187
6.1. THE DERIVATIVE
From the above formula obtained for f ′ (x), we observe that when x 6= 0
f ′ is continuous since it is the product/difference/composition of con-
tinuous functions. To determine continuity of f ′ at x = 0 consider
1
limx→0 f ′(x). Consider the sequence xn = nπ , which clearly converges
2
to c = 0. Now, f ′ (xn) = nπ sin(nπ) − cos(nπ). Now, sin(nπ) = 0 for all
n and therefore f ′(xn) = − cos(nπ) = (−1)n+1. The sequence f ′(xn)
does not converge and therefore limx→0 f ′(x) does not exist. Thus, f ′
is not continuous at x = 0.
188
6.1. THE DERIVATIVE
189
6.1. THE DERIVATIVE
Exercises
190
6.2. THE MEAN VALUE THEOREM
191
6.2. THE MEAN VALUE THEOREM
Corollary 6.2.3
If f : I → R has a relative maximum (or minimum) at c then either
f ′(c) = 0 or f ′(c) does not exist.
Proof. If f (a) = f (b) then the result follows from Rolle’s Theorem
(f ′(c) = 0 for some c ∈ (a, b)). Let h : [a, b] → R be the line from
(a, f (a)) to (b, f (b), that is,
f (b) − f (a)
h(x) = f (a) + (x − a)
(b − a)
192
6.2. THE MEAN VALUE THEOREM
for x ∈ [a, b]. Then g(a) = f (a) − f (a) = 0 and g(b) = f (b) − f (b) = 0,
and thus g(a) = g(b). Clearly, g is continuous on [a, b] and differentiable
on (a, b), and it is straightforward to verify that g ′ (x) = f ′(x)− f (b)−f
b−a
(a)
.
By Rolle’s Theorem, there exists c ∈ (a, b) such that g ′ (c) = 0, and
therefore f ′(c) = f (b)−f (a)
b−a .
Theorem 6.2.7
Let f : [a, b] → R be continuous on [a, b] and differentiable on (a, b).
If f ′(x) = 0 for all x ∈ (a, b) then f is constant on [a, b].
Proof. Let y ∈ (a, b]. Now f restricted to [a, y] satisfies all the as-
sumptions needed in the Mean Value Theorem. Therefore, there ex-
ists c ∈ (a, y) such that f ′ (c) = f (y)−f (a) ′
y−a . But f (c) = 0 and thus
f (y) = f (a). This holds for all y ∈ (a, b] and thus f is constant on
[a, b].
Example 6.2.8. Show by example that Theorem 6.2.7 is not true for
a function f : A → R if A is not a closed and bounded interval.
Corollary 6.2.9
If f, g : [a, b] → R are continuous and differentiable on (a, b) and
f ′(x) = g ′ (x) for all x ∈ (a, b) then f (x) = g(x) + C for some
constant C.
193
6.2. THE MEAN VALUE THEOREM
Theorem 6.2.12
Suppose that f : I → R is differentiable.
194
6.2. THE MEAN VALUE THEOREM
Exercises
195
6.3. TAYLOR’S THEOREM
Pn (x0) = f (x0)
Pn(1) (x0) = f (1) (x0)
.. .
. = ..
P (n) (x0) = f (n) (x0).
196
6.3. TAYLOR’S THEOREM
and Pn (x) is
Rn (x) = f (x) − Pn (x)
and we call Rn (x) the nth order remainder based at x0. Hence, for
each x∗ ∈ [a, b], the remainder Rn (x∗) is the error in approximating
f (x∗) with Pn (x∗). You may be asking yourself why we would need
to approximate f (x) if the function f is known and given. For exam-
ple, if say f (x) = sin(x) then why would we need to approximate say
f (1) = sin(1) since any basic calculator could easily compute sin(1)?
Well, what your calculator is actually computing is an approximation to
sin(1) using a (rational) number such as Pn (1) and using a large value
of n for accuracy (although modern numerical algorithms for comput-
ing trigonometric functions have superseded Taylor approximations but
Taylor approximations are a good start). Taylor’s theorem provides an
expression for the remainder term Rn (x) using the derivative f (n+1).
where
f (n+1)(c)
Rn (x) = (x − x0)n+1.
(n + 1)!
197
6.3. TAYLOR’S THEOREM
function g : [a, b] → R by
Since f (n+1) exists on (a, b) then g (n+1) exists on (a, b). Moreover,
since P (k) (x0) = f (k) (x0) for k = 0, 1, . . . , n then g (k) (x0) = 0 for
k = 0, 1, . . . , n. Now g(x) = 0 and therefore since g(x0) = 0 by Rolle’s
theorem there exists c1 in between x and x0 such that g ′ (c1 ) = 0. Now
we can apply Rolle’s theorem to g ′ since g ′ (c1) = 0 and g ′ (x0) = 0, and
therefore there exists c2 in between c1 and x0 such that g ′′ (c2 ) = 0. By
applying this same argument repeatedly, there exists c in between x0
and cn−1 such that g (n+1) (c) = 0. Now,
Solution. Note that f (1) = ln(2) and so the estimate of ln(2) using P4
is ln(2) ≈ P4 (1). To determine P4 we need f (0), f (1)(0), . . . , f (4)(0). We
198
6.3. TAYLOR’S THEOREM
compute
1
f (1) (x) = f (1) (0) = 1
1+x
−1
f (2) (x) = f (2) (0) = −1
(1 + x)2
2
f (3) (x) = f (3) (0) = 2
(1 + x)3
−6
f (4) (x) = f (4) (0) = −6.
(1 + x)4
Therefore,
P4 (x) = x − 21 x2 + 13 x3 − 41 x4.
Now P4 (1) = 1 − 21 + 13 − 1
4
= 7
12
and therefore
7
ln(2) ≈ P4 (1) = .
12
The error is R4 (1) = f (1) − P4 (1) which is unknown but we can ap-
proximate it using Taylor’s theorem. To that end, by Taylor’s theorem,
for any x ∈ [0, 2] there exists c in between x0 = 0 and x such that
f (5) (c) 5
R4(x) = x
5!
1 24
= 5
x4
5! (1 + c)
1
= .
5(1 + c)5
Therefore, for x = 1, there exists 0 < c < 1 such that
1
R4(1) = .
5(1 + c)5
Therefore, a bound for the error is
1 1
|R4 (1)| = ≤
5(1 + c)5 5
199
6.3. TAYLOR’S THEOREM
since 1 + c > 1.
(b) Restrict f to a closed and bounded interval of the form [−R, R].
Show that for any ε > 0 there exists K ∈ N such that if n ≥ K
then |f (x) − Pn (x)| < ε for all x ∈ [−a, a].
sin(3) ≈ P8 (3)
1 1 1
= 3 − 33 + 35 − 37
3! 5! 7!
51
=
560
= 0.0910714286
(b) Since f (x) = sin(x) has derivatives of all orders, for any n ∈ N we
have by Taylor’s theorem that
xn+1 an+1 n!
lim = lim
n→∞ xn n→∞ (n + 1)!an
a
= lim = 0.
n→∞ n + 1
Therefore, by the Ratio test limn→∞ xn = 0. Hence, for any ε > 0 there
exists K ∈ N such that |xn − 0| = xn < ε for all n ≥ K. Therefore, for
201
6.3. TAYLOR’S THEOREM
202
6.3. TAYLOR’S THEOREM
Exercises
Exercise 6.3.2. Let f : R → R be such that f (k) (x) exists for all x ∈ R
and for all k ∈ N (such a function is called infinitely differentiable on
R). Suppose further that there exists M > 0 such that |f (k) (x)| ≤ M for
all x ∈ R and all k ∈ N. Let Pn (x) be the nth order Taylor polynomial
of f centered at x0 = 0. Let I = [−R, R], where R > 0. Prove that for
any fixed ε > 0 there exists K ∈ N such that for n ≥ K it holds that
for all x ∈ [−R, R]. Hint: f (n) is continuous on [−R, R] for every
n ∈ N.
203
6.3. TAYLOR’S THEOREM
204
7
Riemann Integration
205
7.1. THE RIEMANN INTEGRAL
Ṗ = {([xk−1, xk ], tk )}nk=1
When f (x) > 0 on the interval [a, b], the Riemann sum S(f ; Ṗ) is
the sum of the areas of the rectangles with height f (tk ) and width
(xk − xk−1).
We now define the notion of Riemann integrability.
The set of all Riemann integrable functions on the interval [a, b] will be
denoted by R[a, b].
Theorem 7.1.4
If f ∈ R[a, b] then the number L in the definition of Riemann
integrability is unique.
206
7.1. THE RIEMANN INTEGRAL
If f ∈ R[a, b], we call the number L the integral of f over [a, b] and
we denote it by Z b
L= f
a
Proof. Let f : [a, b] :→ R be such that f (x) = C for all x ∈ [a, b] and
let Ṗ = {([xk−1, xk ], tk )} be a sampled partition of [a, b]. Then
n
X
S(f ; Ṗ) = f (tk )(xk − xk−1)
k=1
n
X
=C (xk − xk−1)
k=1
= C(xn − x0)
= C(b − a).
Hence, with L = C(b − a), we obtain that |S(f ; Ṗ) − L| = 0 < ε for any
Rb
ε > 0 and therefore a f = C(b − a). This proves that f is Riemann
integrable.
207
7.1. THE RIEMANN INTEGRAL
Proof. We consider the special case that [a, b] = [0, 1], the general case
is similar. Let Q̇ = {([xk−1, xk ], qk )} be a sampled partition of [0, 1]
chosen so that qk = 21 (xk + xk−1), i.e., qk is the midpoint of the interval
[xk−1, xk ]. Then
n
X
S(f ; Q̇) = f (qk )(xk − xk−1)
k=1
n
X
1
= 2 (xk + xk−1)(xk − xk−1)
k=1
n
1X 2
= (xk − x2k−1)
2
k=1
1
= (x2n − x20)
2
1
= (12 − 02)
2
1
= .
2
208
7.1. THE RIEMANN INTEGRAL
n
X
≤ |tk − qk |(xk − xk−1)
k=1
< δ(1 − 0)
= δ.
Hence, we have proved that for arbitrary Ṗ that satisfies kṖk < δ it
holds that |S(f ; Ṗ) − 1/2| < δ. Hence, given ε > 0 we let δ = ε and
R1
then if kṖk < δ then |S(f ; Ṗ) − 1/2| < ε. Therefore, 0 f = 12 .
Theorem 7.1.7
Let f ∈ R[a, b] and let g : [a, b] → R be a function such that
g(x) = f (x) for all x ∈ [a, b] except possibly at a finite number of
Rb Rb
points in [a, b]. Then g ∈ R[a, b] and in fact a g = a f .
Rb
Proof. Let L = a f . Suppose that g(x) = f (x) except at one point
x = c. Let Ṗ = {([xk−1, xk ], tk )} be a sampled partition. We consider
mutually exclusive cases. First, if c 6= tk and c 6= xk for all k then
S(f ; Ṗ) = S(g; Ṗ). If c = tk ∈
/ {x0, x1, . . . , xn} for some k then
209
7.1. THE RIEMANN INTEGRAL
where M = 2(|f (c)| + |g(c)|). Let ε > 0 be arbitrary. Then there exists
δ1 > 0 such that |S(f ; Ṗ) − L| < ε/2 for all partitions Ṗ such that
kṖk < δ1 . Let δ = min{δ1, ε/(2M)}. Then if kṖk < δ then
= ε.
Rb Rb
This proves that g ∈ R[a, b] and a g = L = a f . Now suppose by
induction that if g(x) = f (x) for all x ∈ [a, b] except at a j ≥ 1 number
Rb Rb
of points in [a, b] then g ∈ R[a, b] and a g = a f . Now suppose that
h : [a, b] → R is such that h(x) = f (x) for all x ∈ [a, b] except at the
points c1 , c2, . . . , cj , cj+1. Define the function g by g(x) = h(x) for all
x ∈ [a, b] except at x = cj+1 and define g(cj+1 ) = f (cj+1). Then g
and f differ at the points c1 , . . . , cj . Then by the induction hypothesis,
Rb Rb
g ∈ R[a, b] and a g = a f . Now g and h differ at the point cj+1 and
Rb Rb Rb
therefore h ∈ R[a, b] and a h = a g = a f . This ends the proof.
210
7.1. THE RIEMANN INTEGRAL
Rb
Proof. If k = 0 then (kf )(x) = 0 for all x and then clearly a kf = 0,
so assume that k 6= 0. Let ε > 0 be given. Then there exists δ > 0 such
Rb
that if kṖk < δ then |S(f ; Ṗ) − a f k < ε/|k|. Now for any partition
Ṗ, it holds that S(kf ; Ṗ) = kS(f ; Ṗ). Therefore, if kṖk < δ then
Z b Z b
S(kf ; Ṗ) − k f = |k| S(f ; Ṗ) − f
a a
< |k|(ε/|k|)
= ε.
To prove (b), it is easy to see that S(f + g; Ṗ) = S(f ; Ṗ) + S(g; Ṗ).
Rb
Given ε > 0 there exists δ > 0 such that |S(f ; Ṗ) − a f | < ε/2 and
Rb
|S(g; Ṗ) − a g| < ε/2, whenever kṖk < δ. Therefore, if kṖk < δ we
have that
Z b Z b Z b Z b
S(f + g; Ṗ) − f+ g ≤ S(f ; Ṗ) − f + S(g; Ṗ) − g
a a a a
< ε.
To prove (c), let ε > 0 be arbitrary and let δ > 0 be such that if
211
7.1. THE RIEMANN INTEGRAL
Therefore,
Z b Z b
f< g + ε.
a a
Properties (i), (ii), and (iii) in Theorem 7.1.8 are known as homogeneity,
additivity, and monotonicity, respectively.
We now give a necessary condition for Riemann integrability.
Rb
Proof. Let f ∈ R[a, b] and put L = a f . There exists δ > 0 such that
if kṖk < δ then |S(f ; Ṗ) − L| < 1 and therefore |S(f ; Ṗ)| < |L| + 1.
Suppose by contradiction that f is unbounded on [a, b]. Let P be a
partition of [a, b], with sets I1, . . . , In, and with kPk < δ. Then f is
unbounded on some Ij , i.e., for any M > 0 there exists x ∈ Ij =
[xj−1, xj ] such that f (x) > M. Choose samples in P by asking that
212
7.1. THE RIEMANN INTEGRAL
X
|f (tj )(xj − xj−1)| > |L| + 1 + f (tk )(xk − xk−1) .
k6=j
X
|S(f ; Ṗ)| = f (tj )(xj − xj−1) + f (tk )(xk − xk−1)
k6=j
X
≥ |f (tj )(xj − xj−1)| − f (tk )(xk − xk−1)
k6=j
> |L| + 1.
213
7.1. THE RIEMANN INTEGRAL
214
7.1. THE RIEMANN INTEGRAL
Exercises
Hint: For each n ∈ N we have the real number sn = S(f ; Ṗn), and we
Rb
therefore have a sequence (sn). Let L = a f . We therefore want to
prove that lim sn = L.
n→∞
215
7.2. RIEMANN INTEGRABLE FUNCTIONS
Using the Cauchy Criterion, we show next that the Dirichlet func-
tion is not Riemann integrable.
Proof. To show that f is not in R[0, 1], we must show that there exists
ε0 > 0 such that for all δ > 0 there exists sampled partitions Ṗ and Q̇
with norm less than δ but |S(f ; Ṗ) − S(f ; Q̇)| ≥ ε0. To that end, let
ε0 = 1/2, and let δ > 0 be arbitrary. Let n be sufficiently large so that
1/n < δ. Let Ṗ be a sampled partition of [0, 1] with intervals all of equal
length 1/n < δ and let the samples of Ṗ be rational numbers. Similarly,
let Q̇ be a partition of [0, 1] with intervals all of equal length 1/n and
216
7.2. RIEMANN INTEGRABLE FUNCTIONS
with samples irrational numbers. Then S(f ; Ṗ) = 1 and S(f ; Q̇) = 0,
and therefore |S(f ; Q̇) − S(f ; Q̇)| ≥ ε0.
Rb
Proof. If f ∈ R[a, b] then let α(x) = β(x) = f (x). Then clearly a (β −
α) = 0 < ε for all ε > 0. Now suppose the converse and let ε > 0
be arbitrary. Let α and β satisfy the conditions of the theorem, with
Rb ε
a (β − α) < 3 . Now, there exists δ > 0 such that if kṖk < δ then
Z b Z b
ε ε
α − < S(α; Ṗ) < α+
a 3 a 3
and Z Z
b b
ε ε
β − < S(β; Ṗ) < β+ .
a 3 a 3
For any sampled partition Ṗ it holds that S(α; Ṗ) ≤ S(f ; Ṗ) ≤ S(β; Ṗ),
and therefore Z b Z b
ε ε
α − < S(f ; Ṗ) < β+ . (7.1)
a 3 a 3
If Q̇ is another sampled partition with kQ̇k < δ then also
Z b Z b
ε ε
α − < S(f ; Q̇) < β+ . (7.2)
a 3 a 3
Subtracting the two inequalities (7.1)-(7.2), we deduce that
Z b Z b
ε ε
− (β − α) − 2 < S(f ; Ṗ) − S(f ; Q̇) < (β − α) + 2 .
a 3 a 3
217
7.2. RIEMANN INTEGRABLE FUNCTIONS
Rb ε
Therefore, since a (β − α) < 3 it follows that
Definition 7.2.4
A function s : [a, b] → R is called a step-function on [a, b] if
there is a finite number of disjoint intervals I1, I2, . . . , In contained
S
in [a, b] such that [a, b] = nk=1 Ik and such that s is constant on
each interval.
Lemma 7.2.5
Let J be a subinterval of [a, b] and define ϕJ on [a, b] as ϕJ (x) = 1
Rb
if x ∈ J and ϕJ (x) = 0 otherwise. Then ϕJ ∈ R[a, b] and a ϕJ =
µ(J).
Theorem 7.2.6
If ϕ : [a, b] → R is a step function then ϕ ∈ R[a, b].
218
7.2. RIEMANN INTEGRABLE FUNCTIONS
219
7.2. RIEMANN INTEGRABLE FUNCTIONS
220
7.2. RIEMANN INTEGRABLE FUNCTIONS
Our las theorem is the additivity property of the integral, the proof
is omitted.
221
7.2. RIEMANN INTEGRABLE FUNCTIONS
Exercises
222
7.3. THE FUNDAMENTAL THEOREM OF CALCULUS
Proof. Assume for simplicity that E := {a, b}. Let ε > 0 be arbitrary.
Rb
Then there exists δ > 0 such that if kṖk < ε then |S(f ; Ṗ) − a f | < ε.
For any Ṗ, with intervals Ik = [xk−1, xk ] for k = 1, 2, . . . , n, there exists,
by the Mean Value Theorem applied to F on Ik , a point uk ∈ (xk−1, xk )
such that F (xk ) − F (xk−1) = F ′ (uk )(xk − xk−1). Therefore,
n
X
F (b) − F (a) = F (xk ) − F (xk−1)
k=1
n
X
= f (uk )(xk − xk−1)
k=1
= S(f ; Ṗu)
where Ṗu has the same intervals as Ṗ but with samples uk . Therefore,
if kṖk < δ then
Z b Z b
F (b) − F (a) − f = S(f ; Ṗu) − f
a a
< ε.
Rb
Hence, for any ε we have that F (b) − F (a) − a f < ε and this shows
Rb
that a f = F (b) − F (a).
223
7.3. THE FUNDAMENTAL THEOREM OF CALCULUS
Theorem 7.3.3
Let f ∈ R[a, b]. The indefinite integral F : [a, b] → R of f : [a, b] →
R is a Lipschitz function on [a, b], and thus continuous on [a, b].
224
7.4. RIEMANN-LEBESGUE THEOREM
Definition 7.4.1
A set E ⊂ R is said to be of measure zero if for every ε > 0 there
exists a countable collection of open intervals Ik such that
∞
[ ∞
X
E⊂ Ik and µ(Ik ) < ε.
k=1 k=1
Example 7.4.2. Show that a subset of a set of measure zero also has
measure zero. Show that the union of two sets of measure zero is a set
of measure zero.
225
7.4. RIEMANN-LEBESGUE THEOREM
226
7.4. RIEMANN-LEBESGUE THEOREM
227
7.4. RIEMANN-LEBESGUE THEOREM
228
8
Sequences of Functions
229
8.1. POINTWISE CONVERGENCE
230
8.1. POINTWISE CONVERGENCE
2xn + (−1)nx2
lim fn (x) = lim = 2x.
n→∞ n→∞ n
Hence, (fn ) converges pointwise to f (x) = 2x on R. In Figure 8.1, we
graph fn(x) for the values n = 1, 2, 3, 4 and the function f (x) = 2x.
Notice that fn′ (x) = (2n + 2(−1)nx)/n and therefore limn→∞ fn′ (x) = 2,
and for the limit function f (x) = 2x we have f ′(x) = 2. Hence, the
sequence of derivatives (fn′ ) converges pointwise to f ′ . Also, after some
basic computations,
Z 1 Z 1
2xn + (−1)nx2
fn (x) dx = dx
−1 −1 n
2(−1)n
=
3n
and therefore
Z 1
2(−1)n
lim fn(x) dx = lim
n→∞ −1 n→∞ 3n
= 0.
R1
On the other hand it is clear that −1 f (x) dx = 0.
231
8.1. POINTWISE CONVERGENCE
f1
f2
20
f3
f4
10
0
−4 −2 0 2 4
−10
−20
−30
2xn+(−1)n x2
Figure 8.1: Graph of fn(x) = n for n = 1, 2, 3, 4 and f (x) = 2x
Lemma 8.1.4
Let (fn) be a sequence of functions on A. Then (fn) converges
pointwise to f : A → R if and only if for each x ∈ A and each ε > 0
there exists K ∈ N such that |fn (x) − f (x)| < ε for all n ≥ K.
232
8.1. POINTWISE CONVERGENCE
f8
f12
f25
0.8
0.6
0.4
0.2
0.0
0.0 0.2 0.4 0.6 0.8 1.0
233
8.1. POINTWISE CONVERGENCE
1
= lim x2 +
n→∞
√ n
= x2
= |x|.
and fn′ (x) exists for each x ∈ [−1, 1], in other words, fn is differentiable
on A. However, f (x) = |x| is not differentiable on A since f does not
have a derivative at x = 0. In Figure 8.3, we graph fn for various values
of n.
234
8.1. POINTWISE CONVERGENCE
1.4
1.2
1.0
0.8
0.6
0.4
f1
f3
0.2
f8
f12
f25
0.0 f
−1.00 −0.75 −0.50 −0.25 0.00 0.25 0.50 0.75 1.00
q
nx2 +1
Figure 8.3: Graph of fn(x) = n
for n = 1, 3, 8, 12, 25 and f (x) =
|x|
that
2nx
lim fn (x) = lim 2 = 0.
n→∞ n→∞ enx
and therefore
Z 1
lim fn(x) = lim (1 − e−n ) = 1.
n→∞ 0 n→∞
235
8.1. POINTWISE CONVERGENCE
R1
On the other hand, 0 f (x) dx = 0. Therefore,
Z 1 Z 1
f (x) dx 6= lim fn (x) dx
0 n→∞ 0
236
8.1. POINTWISE CONVERGENCE
Exercises
237
8.2. UNIFORM CONVERGENCE
238
8.2. UNIFORM CONVERGENCE
for all x ∈ A and can therefore be interpreted as saying that the graph
of fn lies in the tube of radius ε > 0 and centered along the graph of
f , see Figure 8.4.
4
fn
3
f(x) + ε
2
f(x) − ε
A
1
−1
Proposition 8.2.2
If (fn) converges uniformly to f then (fn) converges pointwise to f .
Example 8.2.3. Let A = [−5, 5] and let (fn) be the sequence of func-
tions on A defined by fn (x) = (2xn + (−1)nx2)/n. Prove that (fn)
converges uniformly to f (x) = 2x.
239
8.2. UNIFORM CONVERGENCE
2xn + (−1)nx2
lim fn(x) = lim = 2x
n→∞ n→∞ n
2xn + (−1)nx2
|fn (x) − f (x)| = − 2x
n
(−1)nx2
=
n
2
|x|
=
n
52
≤ .
n
52
For any given ε > 0 if K ∈ N is such that K < ε then if n ≥ K then
for any x ∈ A we have
52
|fn (x) − f (x)| ≤
n
52
≤
K
< ε.
√
Example 8.2.4. Show that the sequence of functions fn (x) = sin(nx)/ n
converges uniformly to f (x) = 0 on R.
240
8.2. UNIFORM CONVERGENCE
Solution. We compute
sin(nx)
|fn (x) − f (x)| = √
n
| sin(nx)|
= √
n
1
≤√
n
|fn(x) − f (x)| ≤ Mn , ∀x ∈ A
Theorem 8.2.5
Let fn : A → R be a sequence of functions. Then (fn) converges
uniformly to f on A if and only if there exists a sequence (Mn) of
non-negative numbers converging to zero such that supx∈A |fn (x) −
f (x)| ≤ Mn for n sufficiently large.
241
8.2. UNIFORM CONVERGENCE
242
8.2. UNIFORM CONVERGENCE
To prove the converse, suppose that for every ε > 0 there exists K ∈ N
such that if n, m ≥ K then |fm (x)−fn(x)| < ε for all x ∈ A. Therefore,
for each x ∈ A the sequence (fn(x)) is a Cauchy sequence and therefore
converges. Let f : A → R be defined by f (x) = limn→∞ fn(x). If
ε > 0 let K ∈ N be such that |fm(x) − fn (x)| < ε for all x ∈ A and
n, m ≥ K. Fix m ≥ K and consider the sequence zn = |fm (x) − fn(x)|
and thus zn < ε. Now since limn→∞ fn (x) = f (x) then lim zn exists and
lim zn ≤ ε, that is,
243
8.2. UNIFORM CONVERGENCE
Exercises
244
8.3. PROPERTIES OF UNIFORM CONVERGENCE
245
8.3. PROPERTIES OF UNIFORM CONVERGENCE
|f (x) − f (c)| ≤ |f (x) − fK (x)| + |fK (x) − fK (c)| + |fK (c) − f (c)|
246
8.3. PROPERTIES OF UNIFORM CONVERGENCE
247
8.3. PROPERTIES OF UNIFORM CONVERGENCE
248
8.3. PROPERTIES OF UNIFORM CONVERGENCE
(c) Use Theorem 8.3.5 to show that the convergence is not uniform.
249
8.3. PROPERTIES OF UNIFORM CONVERGENCE
Proof. Let x ∈ [a, b] be arbitrary but with x 6= x0. By the Mean Value
theorem applied to the differentiable function fm − fn , there exists y in
between x and x0 such that
(fm(x) − fn (x)) − (fm(x0) − fn(x0)) ′
= fm (y) − fn′ (y)
x − x0
or equivalently
′
fm(x) − fn (x) = fm (x0) − fn (x0) + (x − x0)(fm (y) − fn′ (y))
Therefore,
′
|fm (x) − fn(x)| ≤ |fm (x0) − fn (x0)| + (b − a)|fm (y) − fn′ (y)|.
250
8.3. PROPERTIES OF UNIFORM CONVERGENCE
and this holds for all x ∈ [a, b]. By the Cauchy criterion for uniform
convergence, (fn) converges uniformly. Let f be the uniform limit of
(fn). We now prove that f is differentiable and f ′ = g. By the Funda-
mental theorem of Calculus (FTC), we have that
Z x
fn (x) = fn(a) + fn′ (t) dt
a
for each x ∈ [a, b]. Since (fn ) converges to f and (fn′ ) converges uni-
formly to g we have
251
8.3. PROPERTIES OF UNIFORM CONVERGENCE
252
8.3. PROPERTIES OF UNIFORM CONVERGENCE
Exercises
253
8.3. PROPERTIES OF UNIFORM CONVERGENCE
254
8.4. INFINITE SERIES OF FUNCTIONS
P
Recall that a series of numbers xn converges if the sequence of partial
sums (tn), defined as tn = x1 + x2 + · · · + xn, converges. Hence, conver-
P
gence of an infinite series of functions fn is treated by considering
the convergence of the sequence of partial sums (sn ) (which are func-
P
tions). For example, to say that the series fn converges uniformly to
a function f we mean that the sequence of partial sums (sn ) converges
uniformly to f , etc. It is now clear that our previous work in Sections
8.1-8.3 translate essentially directly to infinite series of functions. As
an example:
255
8.4. INFINITE SERIES OF FUNCTIONS
Theorem 8.4.2
P
Let (fn) be a sequence of functions on A and suppose that fn
converges uniformly to f . If each fn is continuous on A then f is
continuous on A.
P
Proof. By assumption, the sequence of functions sn (x) = nk=1 fk (x)
for x ∈ A converges uniformly to f . Since each function fn is continu-
ous, and the sum of continuous functions is continuous, it follows that
sn is continuous. The result now follows by Theorem 8.3.3.
256
8.4. INFINITE SERIES OF FUNCTIONS
or
∞
Z bX n Z
X b
fn = lim fk
a n=1 n→∞ a
k=1
or
∞
Z bX ∞ Z
X b
fn = fn
a n=1 n=1 a
P
Proof. Let ε > 0 be arbitrary. Let tn = nk=1 Mk be the sequence of
P
partial sums of the series Mn . By assumption, (tn ) converges and
thus (tn ) is a Cauchy sequence. Hence, there exists K ∈ N such that
257
8.4. INFINITE SERIES OF FUNCTIONS
|tm − tn | < ε for all m > n ≥ K. Let (sn ) be the sequence of partial
P
sums of fn . Then if m > n ≥ K then for all x ∈ A we have
≤ Mm + Mm−1 + · · · + Mn+1
= |tm − tn |
< ε.
Hence, the sequence (sn ) satisfies the Cauchy Criterion for uniform
convergence (Theorem 8.2.7) and the proof is complete.
Z ∞
!
π X n sin(nx) 2e
=
0 n=1
en e2 − 1
n sin(nx) n
≤ .
en en
P
A straightforward application of the Ratio test shows that ∞ n
n=1 en is
a convergent series. Hence, by the M-Test, the given series converges
258
8.4. INFINITE SERIES OF FUNCTIONS
∞
X cos(nx) π
= −
n=1
en 0
∞
X 1 n
−1 n
= e − e
n=1
1 1
= 1−1/e −1 − 1+1/e −1
2e
=
e2 − 1
Since
r(nx) 1/2
≤
n2 n2
P
and ∞ 1
n=1 2n2 converges, then by the M-test the above series converges
uniformly on any interval [a, b]. Let f be the function defined by the
series on [a, b]. Now, on [a, b], the function fn (x) = r(nx)
n2 has only a finite
number of discontinuities and thus fn is Riemann integrable. Therefore,
by Theorem 8.3.5, the function f is Riemann integrable. The graph of
f is shown in Figure 8.6. One can show that f has discontinuities at
the rational points x = 2qp where gcd(p, q) = 1.
259
8.4. INFINITE SERIES OF FUNCTIONS
0.6
0.4
0.2
0.0
−3 −2 −1 0 1 2 3
−0.2
−0.4
−0.6
0.6
0.4
0.2
0.0
0.0 0.5 1.0 1.5 2.0 2.5
−0.2
−0.4
−0.6
P∞ r(nx)
Figure 8.6: The function f (x) = n=1 n2
260
8.4. INFINITE SERIES OF FUNCTIONS
261
8.4. INFINITE SERIES OF FUNCTIONS
Applying the Ratio test again we conclude that the series of the deriva-
tives converges for each x ∈ (a − R, a + R) and a similar argument
as before shows that the series of derivatives converges uniformly on
any interval [a − ρ, a + ρ] where ρ < R. It follows from the Term-
by-Term Differentiation theorem that f is differentiable and f ′ (x) =
P
cn n(x − a)n−1. By the Term-by-Term Integration theorem, we can
also integrate the series and
Z X XZ
fn(x) dx = fn (x) dx
I I
(b) Let f denote the function defined by the series on the left and let
g denote the function defined by the series on the right. Justifying
each step, show that f ′ exists and that f ′ = g.
where an , bn ∈ R.
(a) Suppose that for a given (an ) and (bn), the associated Fourier
series converges pointwise on [−π, π] and let f be the pointwise
limit. Prove that in fact the Fourier series converges on R. Hint:
For any y ∈ R there exists x ∈ [−π, π] such that y = x + 2π.
262
8.4. INFINITE SERIES OF FUNCTIONS
P P
(b) Prove that if |an | and |bn | are convergent series then the
associated Fourier series converges uniformly on R.
(c) Suppose that for a given (an ) and (bn), the associated Fourier
series converges uniformly on [−π, π] and let f be the uniform
limit. Prove the following:
Z π
a0 = f (x)dx
−π
Z π
1
an = f (x) cos(nx)dx
π −π
Z π
1
bn = f (x) sin(nx)dx
π −π
Z Z (
π π
π, m = n
sin(nx) sin(mx)dx = cos(nx) cos(mx)dx =
−π −π 6 n
0, m =
263
8.4. INFINITE SERIES OF FUNCTIONS
(b) Let fn be continuous and suppose that fn+1(x) ≤ fn(x) for all
x ∈ [a, b] and all n ∈ N. Suppose that (fn) converges pointwise to
a continuous function f . Prove that the convergence is actually
uniform. Give an example to show that if f is not continuous
then we only have pointwise convergence.
Proof. We first prove (a). For convenience, write Iµ = (aµ , bµ ) for each
µ ∈ X and assume without loss of generality that {bµ | µ ∈ X} is
bounded above. Let b0 = a and let Iµ1 be such that b0 ∈ Iµ1 = (a1 , b1)
and
b1 = sup{bµ | b0 ∈ Iµ }.
264
8.4. INFINITE SERIES OF FUNCTIONS
265
8.4. INFINITE SERIES OF FUNCTIONS
Exercises
P
Exercise 8.4.1. Show that ∞ n
n=0 x converges uniformly on [−a, a] for
every a such that 0 < a < 1. Then show that the given series does not
converge uniformly on (−1, 1). Hint: This is an important series and
you should know what function the series converges uniformly to.
P∞ P∞
Exercise 8.4.2. If n=1 |an | < ∞ prove that n=1 an sin(nx) con-
verges uniformly on R.
266
9
Metric Spaces
267
9.1. METRIC SPACES
for all x, y, z ∈ R.
Example 9.1.3. Let B([a, b]) denote the set of bounded functions on
the interval [a, b], that is, f ∈ B([a, b]) if there exists M > 0 such that
|f (x)| ≤ M for all x ∈ [a, b]. For f, g ∈ B([a, b]) let
268
9.1. METRIC SPACES
then
Example 9.1.6. Let C([a, b]) denote the set of continuous functions
on the inteval [a, b]. Then C([a, b]) ⊂ B([a, b]) and thus (C([a, b]), d∞)
is a metric subspace of (B([a, b]), d∞).
Rb
Example 9.1.7. For f, g ∈ C([a, b]) let d(f, g) = a |f (t) − g(t)| dt.
Prove that d defines a metric on C([a, b]).
269
9.1. METRIC SPACES
Example 9.1.8. Let Rn×n denote the set of n × n matrices with real
entries. For A, B ∈ Rn×n define
270
9.1. METRIC SPACES
(ii) kαxk = |α| kxk for any scalar α ∈ R and any x ∈ V , and
271
9.1. METRIC SPACES
Notice that when n = 1, k·k2 is the absolute value function since kxk2 =
√
x2 = |x| for x ∈ R. When not specified otherwise, whenever we refer
to Rn as a normed vector space we implicitly assume that the norm is
k·k2 and simply use the notation k·k.
kf k∞ = sup |f (x)|.
a≤x≤b
272
9.1. METRIC SPACES
and Z 1/2
b
2
kf k2 = |f (x)| dx .
a
These norms are important in the analysis of Fourier series.
Let (M, d) be a metric space. For x ∈ M and r > 0, the open ball
centered at x of radius r is by definition the set
273
9.1. METRIC SPACES
Exercises
Exercise 9.1.1. Let H be the set of all real sequences x = (x1, x2, x3, . . .)
such that |xn| ≤ 1 for all n ∈ N. For x, y ∈ H let
∞
X
d(x, y) = 2−n|xn − yn |.
n=1
274
9.2. SEQUENCES AND LIMITS
One can indeed show, just as in Theorem 3.1.12 for sequences of real
numbers, that the point p in Definition 9.2.1 is indeed unique.
275
9.2. SEQUENCES AND LIMITS
Lemma 9.2.4
In a metric space, a convergent sequence is bounded.
Proof. Suppose that p is the limit of (zn). There exists K ∈ N such that
d(zn , p) < 1 for all n ≥ K. Let r = 1+max{d(z1, p), . . . , d(zK−1, p)} and
we note that r ≥ 1. Then {zn | n ∈ N} ⊂ Br (p). To see this, if n ≥ K
then d(zn , p) < 1 ≤ r and thus zn ∈ Br (p). On the other hand, for zj ∈
{z1 , . . . , zK−1} we have that d(zj , p) ≤ max{d(z1, p), . . . , d(zK−1, p)} <
r, and thus zj ∈ Br (p). This proves that (zn ) is bounded.
Lemma 9.2.5
Let M be a metric space and let (zn) be a sequence in M. If (zn) → p
then (znk ) → p for any subsequence (znk ) of (zn).
276
9.2. SEQUENCES AND LIMITS
Lemma 9.2.6
Let M be a metric space and let (zn ) be a sequence in M. The
following hold:
Proof. Proofs for (i) and (ii) are left as exercises (see Lemma 3.6.3-
3.6.4). To prove (iii), let (znk ) be a convergent subsequence of (zn), say
converging to p. Let ε > 0 be arbitrary. There exists K ∈ N such
that d(zn , zm ) < ε/2 for all n, m ≥ K. By convergence of (znk ) to p,
by increasing K if necessary we also have that d(znk , p) < ε/2 for all
k ≥ K. Therefore, if n ≥ K, then since nK ≥ K then
Hence, (zn ) → p.
277
9.2. SEQUENCES AND LIMITS
believe that a general metric space comes equipped with a similar com-
pleteness property. Besides, the completeness axiom of R (Axiom 2.4.6)
relies on the order property of R (i.e., ≤) and there is no reason to be-
lieve that a general metric space comes equipped with an order. We
will have more to say about this in Section 9.4. For now, however, we
will consider an important metric space where almost all the results for
sequences in R carry over (an example of a result not carrying over is
the Monotone convergence theorem), namely, the normed vector space
(Rn , k·k).
Denoting a sequence in Rn is notationally cumbersome. Formally,
a sequence in Rn is a function z : N → Rn . How then should we
denote z(k) as a vector in Rn ? One way is to simply write z(k) =
(z1 (k), z2(k), . . . , zn(k)) for each k ∈ N and this is the notation we will
adopt. It is clear that a sequence (z(k)) in Rn induces n sequences
in R, namely, (zi (k)) for each i ∈ {1, 2, . . . , n} (i.e., the component
sequences). The following theorem explains why Rn inherits almost all
the results for sequences in R.
Proof. Suppose first that (z(k)) converges, say to p = (p1, p2, . . . , pn).
For any i ∈ {1, 2, . . . , n} it holds that
p
|zi (k) − pi | ≤ (z1(k) − p1 )2 + (z2 (k) − p2)2 + · · · + (zn (k) − pn )2
278
9.2. SEQUENCES AND LIMITS
in other words, |zi (k) − pi| ≤ kz(k) − pk. Since (z(k)) → p then
limk→∞ kz(k) − pk = 0 and consequently limk→∞ |zi (k) − pi | = 0, that
is, limk→∞ zi (k) = pi .
Conversely, now suppose that (zi (k)) converges for each i ∈ {1, 2, . . . , n}.
Let pi = limk→∞ zi (k) for each i ∈ {1, 2, . . . , n} and let p = (p1, p2, . . . , pn).
By the basic limit laws of sequences in R, the sequence
xk = kz(k) − pk
p
= (z1(k) − p1 )2 + (z2 (k) − p2)2 + · · · + (zn (k) − pn )2
converges to zero since limk→∞ (zi (k) − pi)2 = 0 and the square root
√
function x 7→ x is continuous. Thus, limk→∞ z(k) = p as desired.
Corollary 9.2.8
Every Cauchy sequence in Rn is convergent.
279
9.2. SEQUENCES AND LIMITS
Definition 9.2.10
Let M be a metric space.
(a) A subset U of M is said to be open if for any x ∈ U there exists
ε > 0 such that Bε(x) ⊂ U .
(b) A subset E of M is closed if E c = M\E is open.
280
9.2. SEQUENCES AND LIMITS
Example 9.2.12. Below are some facts that are easily proved; once
(a) and (b) are proved use DeMorgan’s Laws to prove (c) and (d).
Tn
(a) If U1, . . . , Un is a finite collection of open sets then k=1 Uk is
open.
S
(b) If {Uk } is collection of open sets indexed by a set I then k∈I Uk
is open.
Sn
(c) If E1, . . . , En is a finite collection of closed sets then k=1 Ek is
closed.
T
(d) If {Ek } is collection of closed sets indexed by a set I then k∈I Ek
is closed.
281
9.2. SEQUENCES AND LIMITS
Example 9.2.14. Show that C([a, b]) is a closed subset of B([a, b]).
282
9.2. SEQUENCES AND LIMITS
Exercises
Exercise 9.2.1. Let M be a metric space and suppose that (zn ) con-
verges in M. Prove that the limit of (zn ) is unique. In other words,
prove that if p and q satisfy the convergence definition for (zn ) then
p = q.
(b) Let E be the set of step functions on [a, b]. Then clearly E ⊂
B([a, b]). Prove that the set of continuous function C([a, b]) is
contained in the closure of E. (Hint: See Example 8.2.6)
283
9.2. SEQUENCES AND LIMITS
(c) Perform an internet search and find dense subsets of (C([a, b]), k·k∞ )
(you do not need to supply proofs).
284
9.2. SEQUENCES AND LIMITS
Hint: Recall that for any x = (x1, x2, . . . , xn) ∈ Rn it holds that |xi| ≤
kxk2.
285
9.3. CONTINUITY
9.3 Continuity
Using the definition of continuity for a function f : R → R as a guide,
it is a straightforward task to formulate a definition of continuity for a
function f : M1 → M2 where (M1, d1) and (M2, d2) are metric spaces.
286
9.3. CONTINUITY
Lemma 9.3.4
Let f : M1 → M2 and let g : M2 → M3 , where M1 , M2 , and M3 are
metric spaces. If f is continuous at x ∈ M1 and g is continuous at
f (x) then the composite mapping (g ◦ f ) : M1 → M3 is continuous
at x ∈ M1 .
287
9.3. CONTINUITY
Proof. If limn→∞ xn = p then by Theorem 9.3.2, and using the fact that
f is continuous at p, and g is continuous at g(p):
= g( lim f (xn))
n→∞
= lim (g ◦ f )(xn)
n→∞
Proposition 9.3.5
Let (M, d) be a metric space and let f, g : M → R be continuous
functions, where R is equipped with the usual metric. If f and g
are continuous at x ∈ M then f + g, f − g, and f g are continuous
at x ∈ M.
Proof. In all cases, the most economical proof is to use the sequential
criterion. The details are left as an exercise.
288
9.3. CONTINUITY
is defined as
f −1(S) = {x ∈ A | f (x) ∈ S}.
(i) f is continuous on M1 .
289
9.3. CONTINUITY
Example 9.3.8. Use Proposition 9.3.7 to prove that the level sets of
a function f : M → R on a metric space M are closed sets.
Let ℓ∞ denote the set of all real sequences (xn) that are bounded,
that is, {|xn | : n ∈ N} is a bounded set. If x = (xn) ∈ ℓ∞ , it is
straightforward to verify that kxk∞ = supn∈N |xn | defines a norm on ℓ∞
with addition and scalar multiplication defined in the obvious way. Let
ℓ1 be the set of absolutely summable sequences (xn), that is, (xn) ∈ ℓ1
P
if and only if ∞ n=1 |xn | converges. It is not too hard to verify that ℓ1
P
is a normed vector space with norm defined as kxk1 = ∞ n=1 |xn |. If
P∞
n=1 |xn | converges then (|xn |) converges to zero and thus (xn ) ∈ ℓ∞ ,
thus ℓ1 ⊂ ℓ∞ .
290
9.3. CONTINUITY
Exercises
Exercise 9.3.2. Let (V, k·k) be a normed vector space. Prove that f :
V → R defined by f (x) = kxk is continuous. Hint: kak ≤ ka − bk+kbk
for all a, b, c ∈ V .
Exercise 9.3.3. Consider C([a, b]) with norm k·k∞ . Define the functionΨ :
Rb
C([a, b]) → R by Ψ(f ) = a f (x) dx. Prove that Ψ is continuous in two
ways, using the definition and the sequential criterion for continuity.
(b) Given matrices X, Y ∈ Rn×n , recall that the entries of the prod-
P
uct matrix XY are (XY)i,j = nℓ=1 xi,ℓyℓ,j . Let (A(k))∞ k=1 be a
sequence in Rn×n converging to B and let (C(k))∞ k=1 be the se-
quence whose kth term is C(k) = A(k)A(k) = [A(k)]2. Prove
that (C(k))∞ 2
k=1 converges to B . Hint: By part (a), it is enough
to prove that the (i, j) component of C(k) converges to the (i, j)
component of B2 .
291
9.3. CONTINUITY
f (A) = cm Am + cm−1Am−1 + · · · + c1 A + c0 I
292
9.4. COMPLETENESS
9.4 Completeness
Consider the space P[a, b] of polynomial functions on the interval [a, b].
Clearly, P[a, b] ⊂ C([a, b]), and thus (P[a, b], k·k∞ ) is a metric space.
P
The sequence of functions fn (x) = nk=0 k!1 xk is a sequence in P[a, b]
and it can be easily verified that (fn ) converges in the metric k·k∞ , that
is, (fn) converges uniformly in [a, b] (see Example 8.4.8). However, the
limiting function f is not an element of P[a, b] because it can be verified
that f ′(x) = f (x) and the only polynomial equal to its derivative is the
zero polynomial, however, it is clear that f (0) = limn→∞ fn (0) = 1, i.e.,
f is not the zero function (you may recognize, of course, that f (x) = ex ).
We do know, however, that f is in C([a, b]) because the uniform limit of
a sequence of continuous functions is continuous. The set P[a, b] then
suffers from the same “weakness” as do the rationals Q relative to R,
namely, there are sequences in P[a, b] that converge to elements not
in P[a, b]. On the other hand, because (fn) converges it is a Cauchy
sequence in C([a, b]) and thus also in P[a, b] (the Cauchy condition
only depends on the metric) and thus (fn) is a Cauchy sequence in
P[a, b] that does not converge to an element of P[a, b]. The following
discussion motivates the following definition.
293
9.4. COMPLETENESS
Theorem 9.4.2
Let (M, d) be a complete metric space and let P ⊂ M. Then (P, d)
is a complete metric space if and only if P is closed.
Example 9.4.3. Consider P[0, 1] with induced metric k·k∞ and let
E = {f ∈ P[0, 1] : kf k∞ < 3}, in other words, E is the open ball of
radius r = 3 centered at the zero function. Clearly, E is bounded and
294
9.4. COMPLETENESS
P
thus any sequence in E is bounded. The sequence fn (x) = nk=0 k!1 xk
is in E, that is, kfn k∞ < 3 for all n (see Example 3.3.6). However,
as already discussed, (fn) converges in C[0, 1] but not to a point in
P[0, 1]. On the other hand, (fn) is a Cauchy sequence in P[0, 1] and
thus (fn ) cannot have a converging subsequence in P[0, 1] by part (iii)
of Lemma 9.2.6. Thus, (fn) is a bounded sequence in P [0, 1] with no
converging subsequence in P [0, 1].
The correct notion of boundedness that is needed is the following.
Definition 9.4.4
Let (M, d) be a metric space. A subset E ⊂ M is called totally
bounded if for any given ε > 0 there exists z1 , . . . , zN ∈ E such
S
that E ⊂ N i=1 Bε (zi ).
295
9.4. COMPLETENESS
296
9.4. COMPLETENESS
Lemma 9.4.10
If E is an infinite totally bounded subset of (M, d) then E contains
a Cauchy sequence (xn) such that xn 6= xm for all n 6= m.
297
9.4. COMPLETENESS
Theorem 9.4.11
Let X be a non-empty set. The normed space (B(X), k·k∞ ) is a
Banach space.
Proof. First of all, it is clear that B(X) is a real vector space and thus
we need only show it is complete. The proof is essentially contained
in the Cauchy criterion for uniform convergence for functions on R
(Theorem 8.2.7). Let fn : X → R be a Cauchy sequence of bounded
functions. Then for any given ε > 0 there exists K ∈ N such that if
n, m ≥ K then kfn − fmk∞ < ε. In particular, for any fixed x ∈ X it
holds that
|fn (x) − fm (x)| ≤ kfn − fm k∞ < ε.
Therefore, the sequence of real numbers (fn(x)) is a Cauchy sequence
and thus f (x) = limn→∞ fn (x) exists for each x ∈ X. Now, since (fn)
is a Cauchy sequence in B(X) then (fn) is bounded in B(X). Thus,
there exists M > 0 such that kfnk∞ ≤ M for all n ≥ 1. Thus, for all
x ∈ X and n ≥ 1 it holds that
Thus, f is a bounded function, that is, f ∈ B(X). Now, for any fixed
ε > 0, let K ∈ N be such that |fn (x) − fm (x)| < ε/2 for all x ∈ X and
n, m ≥ K. Therefore, for any x ∈ X we have that
298
9.4. COMPLETENESS
Therefore, kfn − f k∞ < ε for all n ≥ K. This proves that (fn ) con-
verges to f in (B(X), k·k∞ ).
Corollary 9.4.12
The space of continuous functions C([a, b]) on the interval [a, b] with
sup-norm is a Banach space.
Example 9.4.13. Prove that ℓ∞ and ℓ1 are complete and hence Banach
spaces.
P∞
Proof. Suppose that n=1 kzn k converges, that is, suppose that the
P
sequence of partial sums tn = nk=1 kzn k converges (note that (tn ) is
increasing). Then (tn ) is a Cauchy sequence. Consider the sequence of
299
9.4. COMPLETENESS
Pn
partial sums sn = k=1 zk . For n > m we have
n
X
ksn − sm k = zn
k=m+1
Xn
≤ kzn k
k=m+1
= tn − tm
= |tn − tm |
and since (tn ) is Cauchy then |tn − tm | can be made arbitrarily small
provided n, m are sufficiently large. This proves that (sn ) is a Cauchy
sequence in V and therefore converges since V is complete.
Example 9.4.16. Consider the set of matrices Rn×n equipped with the
2-norm !1/2
Xn
kAk2 = a2i,j
i,j=1
300
9.4. COMPLETENESS
P
(c) Let f (x) = ∞ k
k=1 ck x be a power series converging on R. Define
the function f : Rn×n → Rn×n as
∞
X
f (A) = ck Ak .
k=1
∞
X ∞
X
ck A k
≤ ck kAkk2
k=1 2 k=1
Proof. (a) The norm kAk2 is simply the standard Euclidean norm on
(RN , k·k2) with N = n2 and identifying matrices as elements of
RN . Hence, (Rn×n, k·k2 ) is complete.
n
!2
X
(AB)2i,j = ai,ℓbℓ,j
ℓ=1
n
! n
!
X X
≤ a2i,ℓ b2ℓ,j
ℓ=1 ℓ=1
301
9.4. COMPLETENESS
and therefore
!1/2
X
kABk2 = (AB)2i,j
1≤i,j≤n
n
! n
!!1/2
X X X
≤ a2i,ℓ b2ℓ,j
1≤i,j≤n ℓ=1 ℓ=1
1/2 1/2
n
X n
X
= a2i,ℓ b2ℓ,j
i,ℓ=1 j,ℓ=1
= kAk2 kBk2
P
(c) We first note that for any power series ∞ k
k=1 ak x that converges
P
in (−R, R), the power series ∞ k
k=1 |ak |x also converges in (−R, R).
P
The normed space (Rn×n, k·k2 ) is complete and thus ∞ k=1 ck A
k
P
converges whenever ∞ k
k=1 ck A 2 converges. Now by part (b),
P
ck Ak 2 = |ck | Ak 2 ≤ |ck | kAkk2 and since ∞ k
k=1 |ck | kAk2 con-
verges then by the comparison test for series in R, the series
P∞ k
k=1 ck A 2 converges. Therefore, f (A) is well-defined by The-
orem 9.4.14. To prove the last inequality, we note that the norm
function on a vector space is continuous and thus if ck ≥ 0 then
m
X m
X
ck A k
≤ |ck | kAkk2
k=1 2 k=1
Xm
≤ ck kAkk2
k=1
302
9.4. COMPLETENESS
and therefore
∞
X m
X
k
ck A = lim ck Ak
m→∞
k=1 2 k=1 2
m
X
≤ lim ck kAkk2
m→∞
k=1
∞
X
= ck kAkk2 ,
k=1
303
9.4. COMPLETENESS
Exercises
Exercise 9.4.1. Let (M1 , d1) and (M2 , d2) be metric spaces. There
are several ways to define a metric on the Cartesian product M1 × M2.
One way is to imitate what was done in R2 = R × R. We can define
d : M1 × M2 → [0, ∞) as
p
d((x, u), (y, v)) = d1 (x, y)2 + d2 (u, v)2
Note: The △-inequality can only be used on a finite sum. (See Exer-
cise 9.3.2.)
304
9.5. COMPACTNESS
9.5 Compactness
305
9.5. COMPACTNESS
306
9.5. COMPACTNESS
Theorem 9.5.7
Let f : M1 → M2 be a continuous mapping. If E ⊂ M1 is compact
then f (E) ⊂ M2 is compact.
307
9.5. COMPACTNESS
308
9.5. COMPACTNESS
309
9.5. COMPACTNESS
Theorem 9.5.13
If f : (M1 , d1) → (M2, d2) is continuous and M1 is compact then f
is uniformly continuous.
Proof. Let ε > 0. For each x ∈ M1 , there exists rx > 0 such that if y ∈
Brx (x) then f (y) ∈ Bε/2(f (x)). Now {Brx /2(x)}x∈M1 is an open cover
of M1 and since M1 is compact there exists finite x1 , x2, . . . , xN such
that {Bδi (xi)}Ni=1 is an open cover of M1 , where we have set δi = rxi /2.
Let δ = min{δ1, . . . , δN }. If d1 (x, y) < δ, and say x ∈ Bδi (xi), then
d1(y, xi) ≤ d1 (y, x) + d1 (x, xi) < δ + δi < rxi and thus
310
9.6. FOURIER SERIES
Exercises
311
9.6. FOURIER SERIES
for x ∈ [−π, π]. In other words, the series on the right of (9.1) converges
pointwise to f on [−π, π]. The first question we need to answer is what
are the coefficients an and bn in terms of f ? To that end, we use the
following facts. Let n, m ∈ N:
312
9.6. FOURIER SERIES
(ii) If n 6= m then
Z π Z π
sin(nx) sin(mx)dx = cos(nx) cos(mx)dx = 0.
−π −π
Z π ∞ Z π
a0 X
= cos(nx)dx + ak cos(kx) cos(nx)dx
2 −π −π
k=1
∞
X Z π
+ bk sin(kx) cos(nx)dx
k=1 −π
= an π.
Therefore, Z
1 π
an = f (x) cos(nx)dx.
π −π
A similar calculation shows that
Z
1 π
bn = f (x) sin(nx)dx.
π −π
313
9.6. FOURIER SERIES
Finally,
Z π Z ∞ Z π ∞ Z π
a0 π X X
f (x)dx = dx + ak cos(kx)dx + bk sin(kx)dx
−π 2 −π −π −π
k=1 k=1
a0
= 2π
2
= a0 π
and therefore
Z π
1
a0 = f (x)dx.
π −π
Of course, the above calculations are valid provided that the Fourier se-
ries converges uniformly to f on [−π, π] since if all we have is pointwise
convergence then in general we cannot interchange the integral sign and
the infinite sum. Since the functions fn (x) = an cos(nx) + bn sin(nx)
in the Fourier series are clearly continuous, and if we insist that the
convergence is uniform, then we have restricted our investigation of
Fourier series to continuous functions! Relaxing this restriction led to
the development of what we now call the Lebesgue integral; Lebesgue
was interested in extending the notion of integration beyond Riemann’s
development so that a wider class of functions could be integrated and,
more importantly, this new integral would be more robust when it came
to exchanging limits with integration, i.e., uniform convergence would
not be needed. A full development of Lebesgue’s theory of integra-
tion is beyond the scope of this book, however, we can still say some
interesting things about Fourier series.
Motivated by our calculations above, suppose that f ∈ C[−π, π]
314
9.6. FOURIER SERIES
and define
Z
1 π
a0 = f (x)dx
π −π
Z
1 π
an = f (x) cos(nx)dx
π −π
Z
1 π
bn = f (x) sin(nx)dx.
π −π
Assume that the Fourier series of f converges uniformly on C[−π, π]
and let ∞
a0 X
g(x) = + (an cos(nx) + bn sin(nx)).
2 n=1
Then g is continuous on [−π, π]. Does f = g? To answer this question,
our computations above show that
Z Z
1 π 1 π
g(x) cos(nx)dx = f (x) cos(nx)dx
π −π π −π
and therefore Z π
[f (x) − g(x)] cos(nx)dx = 0
−π
for all n ∈ N ∪ {0}. Similarly, for all n ∈ N ∪ {0} we have
Z π
[f (x) − g(x)] sin(nx)dx = 0.
−π
Pn
Let sn (x) = a20 + k=1 ak cos(kx) + bk sin(kx) and recall that (sn) con-
verges uniformly to g. Consider for the moment
Z π Z π Z π Z π
[f (x) − sn(x)]2dx = f (x)2dx − 2 f (x)sn(x)dx + s2n (x)dx.
−π −π −π −π
315
9.6. FOURIER SERIES
and " #
Z π 2 n
a X
s2n (x)dx = π 0 + (a2k + b2k ) .
−π 2
k=1
Therefore,
Z Z π Z π
1 π 1 1
[f (x) − sn (x)]2dx = f (x)2dx − s2n (x)dx
π −π π −π π −π
Rπ
Now since −π [f (x) − sn (x)]2dx ≥ 0 it follows that
Z Z
1 π 2 1 π
sn (x)dx ≤ f (x)2dx,
π −π π −π
or equivalently that
n Z π
a20 X 2 1
+ (ak + b2k ) ≤ f (x)2dx.
2 π −π
k=1
a20 P∞ 2
This proves that the series 2 + k=1 (ak + b2k ) converges and
∞ Z π
a20 X 2 1
+ (ak + b2k ) ≤ f (x)2dx.
2 π −π
k=1
316
10
10.1 Differentiation
Let U ⊂ Rn and let F : U → Rn be a function. How should we define
differentiability of F at some point a ∈ U ? Recall that for a function
317
10.1. DIFFERENTIATION
318
10.1. DIFFERENTIATION
319
10.1. DIFFERENTIATION
linear algebra and the definition of the partial derivative. For the lat-
ter, recall that a function f : U ⊂ Rn → R, has partial derivative at
a ∈ U with respect to xi, if the following limit exists
f (a + ei t) − f (a) − mi t
0 = lim
t→0 t
L(x) = Ax
where A ∈ Rm×n and the the (j, i) entry of the matrix A is the jth
component of the vector Aei ∈ Rm . We can now prove the following.
320
10.1. DIFFERENTIATION
Proof. Let mj,i denote the (j, i) entry of the matrix representation of
DF (a) in the standard bases in Rn and Rm , that is, mj,i is the jth
component of DF (a)ei . By definition of differentiability, it holds that
kF (x) − F (a) − DF (a)(x − a)k
0 = lim .
x→a kx − ak
Let x = a + tei where ei ∈ Rn is the ith standard basis vector. Since U
is open, x ∈ U provided t is sufficiently small. Then since kx − ak =
ktei k = |t| → 0 iff kx − ak → 0 we have
kF (a + tei ) − F (a) − DF (a)ei tk
0 = lim
t→0 |t|
1
= lim F (a + tei ) − F (a) − DF (a)eit .
t→0 t
It follows that each component of the vector 1t F (a + tei ) − F (a) −
DF (a)ei t tends to 0 as t → 0. Hence, for each j ∈ {1, 2, . . . , m} we
have
1
0 = lim (fj (a + tei ) − fj (a) − mj,i t).
t→0 t
321
10.1. DIFFERENTIATION
∂fj ∂fj
Hence, ∂xi
(a) exists and mj,i = ∂xi
(a) as claimed.
It is customary to write
∂f1 ∂f1 ∂f1
∂x1 ∂x2 ··· ∂xn
∂f2 ∂f2
··· ∂f2
DF (a) = ∂x. 1 ∂x2 ∂xn
.. .. ... ..
. .
∂fm ∂fm ∂fm
∂x1 ∂x2 ··· ∂xn
322
10.1. DIFFERENTIATION
≤ kx − ak + kDF (a)k2 kx − ak
Hence, F is continuous at a.
Notice that Theorem 10.1.2 says that if DF (a) exists then all the
relevant partials exist. However, it does not generally hold that if all
the relevant partials exist then DF (a) exists. The reason is that partial
derivatives are derivatives along the coordinate axes whereas, as seen
from the definition, the limit used to define DF (a) is along any direction
that x → a.
∂f ∂f
We determine whether ∂x
(0, 0) and ∂y
(0, 0) exist. To that end, we
compute
f (x + t, 0) − f (0, 0) 0
lim = lim = 0
t→0 t t→0 t
f (0, y + t) − f (0, 0) 0
lim = lim = 0
t→0 t t→0 t
323
10.1. DIFFERENTIATION
Therefore, ∂f ∂f
∂x (0, 0) and ∂y (0, 0) exist and are both equal to zero. It is
straightforward to show that f is not continuous at (0, 0) and therefore
not differentiable at (0, 0).
324
10.1. DIFFERENTIATION
and thus ∂f
∂x (0, 0) = 0. A similar computations shows that
∂f
∂y (0, 0) = 0.
On the other hand, if (x, y) 6= (0, 0) then
∂f xy 2 (x2 + 2y 2)
(x, y) =
∂x (x2 + y 2 )3/2
∂f x2y(2x2 + y 2 )
(x, y) = .
∂y (x2 + y 2 )3/2
To prove that Df (x, y) exists for any (x, y) ∈ R2 , it is enough to show
that ∂f ∂f 2
∂x and ∂y are continuous on R (Theorem 10.1.5). It is clear that
∂f ∂f 2
∂x and ∂y are continuous on the open set U = R \{(0, 0)} and thus Df
exists on U . Now consider the continuity of ∂f∂x at (0, 0). Using polar
coordinates x = r cos(θ) and y = r sin(θ), we can write
∂f xy 2(x2 + 2y 2 )
(x, y) =
∂x (x2 + y 2 )3/2
=0
325
10.1. DIFFERENTIATION
326
10.1. DIFFERENTIATION
Exercises
327
10.2. DIFFERENTIATION RULES AND THE MVT
Example 10.2.2. Verify the chain rule for the composite function H =
G ◦ F where F : R3 → R2 and G : R2 → R2 are
x1 − 3x2
F (x1, x2, x3) =
x1 x2 x3
2y1 + y2
G(y1, y2 ) = .
sin(y2)
328
10.2. DIFFERENTIATION RULES AND THE MVT
330
10.2. DIFFERENTIATION RULES AND THE MVT
y
x
U
Hence,
f (y) − f (x) = Df (γ(t∗))(y − x)
and the proof is complete.
331
10.2. DIFFERENTIATION RULES AND THE MVT
Corollary 10.2.9
Let U ⊂ Rn be open and assume that F = (f1, f2, . . . , fm) : U → Rm
is differentiable on U . Let x, y ∈ U and suppose that the line
segment joining x, y ∈ U is contained entirely in U . Then there
exists c1 , c2, . . . , cm ∈ U on the line segment joining x and y such
that fi(y) − fi(x) = Dfi (ci )(y − x) for i = 1, 2, . . . , m.
332
10.2. DIFFERENTIATION RULES AND THE MVT
Exercises
333
10.3. THE SPACE OF LINEAR MAPS
γ ′(t) = F (γ(t))
Theorem 10.4.1
Let U ⊂ Rn be an open set and let F : U → Rn be a differentiable
function with a continuous derivative
334
10.5. HIGH-ORDER DERIVATIVES
is a linear mapping.
and
335
10.5. HIGH-ORDER DERIVATIVES
These facts are proved by expanding the determinant along the first
column. The same is true if we perform the same computation with a
different column of A. In the case of a 2 × 2 matrix A = [ xx21 yy21 ] we have
det(A) = x1 y2 − y1 x2
and if A is a 3 × 3 matrix with columns x = (x1, x2, x3), y = (y1 , y2, y3),
and z = (z1, z2 , z3) then
336
10.5. HIGH-ORDER DERIVATIVES
n n n
!
X X X
T (x, y, z) = T xiei , yi ei , . . . , zi ei
i=1 i=1 i=1
n X
X n X
n
= xiyj zk · T (ei, ej , ek ).
i=1 j=1 k=1
where the superscripts are not exponents but indices, then from our
337
10.5. HIGH-ORDER DERIVATIVES
computation above
n
X
A1i,j,k · xiyj zk
i,j,k=1
X n
2
Ai,j,k · xiyj zk
T (x, y, z) =
i,j,k=1
.
.
..
n
X m
Ai,j,k · xiyj zk
i,j,k=1
338
10.5. HIGH-ORDER DERIVATIVES
Then
339
10.5. HIGH-ORDER DERIVATIVES
(u, v) 7→ L(u)(v)
340
10.5. HIGH-ORDER DERIVATIVES
Lemma 10.5.8
Let V and W be vector spaces. The vector space L(V, L(V, W )) is
isomorphic to the vector space L2 (V, W ) of multilinear maps from
V × V to W .
m
X
2
D F (a)(u, v) = D2 fj (a)(u, v)ej .
j=1
341
10.5. HIGH-ORDER DERIVATIVES
Therefore,
2 ∂ 2f
D f (a)(ei, ej ) = (a).
∂xj xi
Therefore, for any u = (u1, u2, . . . , un) and v = (v1, v2, . . . , vn), by mul-
tilinearity we have
n X n
2
X ∂ 2f
D f (a)(u, v) = (a)uivj .
i=1 j=1
∂x i x j
342
10.5. HIGH-ORDER DERIVATIVES
k
X ∂kf
D f (a)(u1, u2, . . . , uk ) = (a)u1,i1 u2,i2 · · · uk,ik
1≤i1 ,i2 ,...,ik ≤n
∂x i 1 ∂x i 2 · · · ∂x i k
k
X ∂kf
D f (a)(x, x, . . . , x) = (a)xi1 xi2 · · · xik
1≤i1 ,i2 ,...,ik ≤n
∂x i 1
∂x i 2
· · · ∂x i k
343
10.5. HIGH-ORDER DERIVATIVES
344
10.6. TAYLOR’S THEOREM
where
1
Rr (x) = Dr+1 f (c)(x − a, x − a, . . . , x − a).
(r + 1)!
Furthermore,
Rr (x)
lim r = 0
x→a kx − ak
345
10.6. TAYLOR’S THEOREM
f (x) ≈ Tr (x).
D2 f (a)(u, u) = 0
Therefore,
346
10.6. TAYLOR’S THEOREM
Exercises
Exercise 10.6.1. Find the 2nd order Taylor polynomial of the function
f (x, y, z) = cos(x + 2y)ez centered at a = (0, 0, 0).
L(x) = Dk f (a)(x, x, . . . , x)
347
10.7. THE INVERSE FUNCTION THEOREM
or in vector form
Ax = y,
where the unknown is x = (x1, x2, . . . , xn) ∈ Rn , has a unique solution
if and only if A−1 exists if and only if det(A) 6= 0. In this case, the
solution is y = A−1x. Another way to say this is that the mapping
F (x) = Ax has a global inverse given by F −1(x) = A−1x. Hence,
invertibility of DF = A completely determines whether F is invertible.
Consider now a system of equations
F (x) = y
348
10.7. THE INVERSE FUNCTION THEOREM
1
(f −1)′(f (x)) = .
f ′(x)
349
10.7. THE INVERSE FUNCTION THEOREM
Example 10.7.2. Prove that F (x, y) = (f1(x, y), f2(x, y)) = (x2 −
y 2 , 2xy) is locally invertible at all points a 6= (0, 0).
Proof. Clearly, DF (x, y) exists for all (x, y) since all partials of the
components of F are continuous on R2 . A direct computation gives
2x −2y
DF (x, y) =
2y 2x
and thus det(DF (x, y)) = 2x2 + 2y 2 . Clearly, det(DF (x, y)) = 0 if and
only if (x, y) = (0, 0). Therefore, by the Inverse Function theorem, for
each non-zero a ∈ R2 there exists an open set U ⊂ R2 containing a such
that F : U → F (U ) is invertible. In this very special case, we can find
the local inverse of F about some a ∈ R2 . Let (u, v) = F (x, y), that is,
x2 − y 2 = u
2xy = v
2
v v
If x 6= 0 then y = 2x and therefore x2 − 4x 4 2
2 = u and therefore 4x − v =
4ux2 or
4x4 − 4ux2 − v 2 = 0.
350
10.7. THE INVERSE FUNCTION THEOREM
and therefore √
v 2v
y= = p √
2x 2 u + u2 + v 2
Hence, provided u 6= 0 and v 6= 0 then
q √
u+ u2 +v 2
F −1
(u, v) = √ 2 .
√ √2v
2 u+ u2 +v 2
351
10.7. THE INVERSE FUNCTION THEOREM
Exercises
for (x, y) ∈ R2 .
(a) Prove that the range of F is R2 \{0}. Hint: Think polar coordi-
nates.
x + xyz = u
y + xy = v
z + 2x + 3z 2 = w
352
Bibliography
353