0% found this document useful (0 votes)
54 views

SMA 2301 Real Analysis I Notes

This document defines and discusses countable and uncountable sets, as well as open and closed sets in metric spaces. Some key points: - Sets are countable if they are finite or equivalent to the natural numbers, and uncountable otherwise. The real numbers are uncountable while the rational numbers are countable. - A metric space is a non-empty set equipped with a metric or distance function satisfying three properties. - In a metric space, a set is open if it contains an open ball around each of its points, and closed if its complement is open. - The Cartesian product of countable sets is countable, while subsets and countable unions of countable sets

Uploaded by

xtnnjr
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
54 views

SMA 2301 Real Analysis I Notes

This document defines and discusses countable and uncountable sets, as well as open and closed sets in metric spaces. Some key points: - Sets are countable if they are finite or equivalent to the natural numbers, and uncountable otherwise. The real numbers are uncountable while the rational numbers are countable. - A metric space is a non-empty set equipped with a metric or distance function satisfying three properties. - In a metric space, a set is open if it contains an open ball around each of its points, and closed if its complement is open. - The Cartesian product of countable sets is countable, while subsets and countable unions of countable sets

Uploaded by

xtnnjr
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 52

SMA 2301 REAL ANALYSIS I

1 Countable and Uncountable Sets


1.1 Equivalence of Sets
Definition 1.1.1: Let A and B be sets. If there exists a bijection (one-to-one and onto
function) f : A −→ B, then A and B can be put into a one-to-one correspondence. In this
case, we say that A and B are equivalent or equinumerous or have the same cardinality, and
we denote this by A ∼ B.
Proposition 1.1.1: Equivalence of sets is an equivalence relation.
Proof :
(a) Reflexivity: The identity map ι : A −→ A, ι (a) = a for all a ∈ A is a one-to-one and
onto mapping from A to itself. Hence A ∼ A.
(b) Symmetry: Suppose A ∼ B. Then there exists a bijection f : A −→ B. But every
bijection has an inverse which is also a bijection. In this case, f −1 is a bijection and
maps B to A. Hence B ∼ A.
(c) Transitivity: Suppose A ∼ B and B ∼ C. Then there exists bijections f : A −→ B
and g : B −→ C. But composition of two bijections is also a bijection. So g ◦ f is a
bijection and maps A to C. Hence A ∼ C. @
Examples:
1. Let E = 2N, the set of all even natural numbers. Then the function f : N −→ E
defined by f (n) = 2n for all n ∈ N is both one-to-one and onto. Hence E ∼ N.
2. Show that N ∼ Z.
Solution: Define a function f : N −→ Z by
(
n
2
if n is even
f (n) = −(n−1) .
2
if n is odd

Then f is a one-to-one and onto function from N to Z. Hence N ∼ Z.

1.2 Types of Sets


Definition 1.2.1: A set A is finite if A = ∅, i.e., A is empty or A ∼ Nn = {1, · · · , n}, i.e.,
A is equivalent to the set of the first n positive integers. In other words, a set is finite if it
has n elements where n is finite. A set is said to be infinite if it is not finite.
Definition 1.2.2: A set is A countable if A is finite or A ∼ N (i.e., A is equivalent to the
natural number set). If A ∼ N, then A is said to be countably infinite or denumerable. A set
A is said to be uncountable if it is not countable.
Example:
From a previous example N ∼ Z. Hence Z is countable (countably infinite).
Theorem 1.2.1: A set is countable if it is equivalent to a subset of N.

1
Example:
We can also show that Z is countable by showing that there exists a bijection from Z to a
subset of N. For instance, define g : Z −→⊆ N by
(
2x if x ≥ 0
g (x) = .
3−x if x < 0

Theorem 1.2.2: Every subset of a countably infinite set is countable.


Proof : Suppose B is a subset of a countably infinite set A. If B is finite, then B is countable,
from Definition 1.2.2. On the other hand, let B be infinite and let A = {a1 , a2 , · · · }. Suppose
n1 is the smallest subscript for which an1 ∈ B, n2 the next smallest, and so on. Then
B = {an1 , an2 , · · · }. The elements of B are thus labelled with 1, 2, · · · , and so B is countable.
@
Theorem 1.2.3: Any subset of a countable set is countable.
Proof : Suppose A is countable and B ⊂ A. Since A is countable, there is a bijection
′ ′
f : A −→ N. But then f (B) = N is a subset of N, and f is a bijection between B and N .

Hence B is countable (since B is equivalent to a subset N of N). @
Theorem 1.2.4: If B ⊂ A and B is uncountable, then so is A.
Proof : If A is countable, then so is B (by Theorem 1.2.3). So A is uncountable as well. @

Theorem 1.2.5: If A1 , A2 , · · · are countable sets, then ∪ An is countable, i.e., the countable
n=1
union of countable sets is countable.
Proof : Write

A1 = a11 , a12 , a13 , · · ·




A2 = a21 , a22 , a23 , · · ·




..
.
An = {an1 , an2 , an3 , · · · }
..
.

so that ajk is the k th element of Aj (j = 1, 2, · · · ). Define the height of ajk to be j + k.


Then a11 is the only element of height 2; likewise a12 and a21 are the only element of height
3; and so on. Since for any integer m ≥ 2 there are only m − 1 elements of height m

namely a1m−1 , a2m−2 , · · · , am−1

1 , we may arrange (count) the elements of ∪ An according to
n=1
their heights as
a11 , a12 , a21 , a13 , a22 , a31 , a14 , · · ·
removing any ajk that has already been counted. This counting scheme eventually counts

every ajk . Hence ∪ An is countable. @
n=1

Example:
Prove that the set Q of rational numbers is countable.

2

Proof : Clearly, Q = ∪ En where En = {0/n, ±1/n, ±2/n, ±3/n, · · · }. Now, each En is equiv-
n=1
alent to Z and is thus countable (since Z is countable). Since Q is the countable union of
countable sets, it is countable.
Further Examples:
1. The set R of real numbers is uncountable.
Proof : Suppose R is countable, i.e., R = {x1 , x2 , · · · }. Let
 
1 1
I1 = x1 − , x1 +
4 4
 
1 1
I2 = x2 − , x2 +
8 8
..
.
In = xn − 2−(n+1) , xn + 2−(n+1)


..
.

The length of In is 2−n so that the sum of all the lengths of In ’s is 2−1 +2−2 +2−3 +· · · = 1.
∞ ∞
But xn ∈ In so that R = ∪ {xn } ⊂ ∪ In . This means that the whole real line (whose
n=1 n=1
length is infinite) would be covered by (contained in) a union of intervals whose lengths
add up to 1. This is a contradiction. Therefore, R must be uncountable.
2. Let A = (0, ∞). Then A ∼ R since the function f : R −→ A defined by f (x) = ex is
both one-to-one and onto. Since R is uncountable and A ∼ R, then A is uncountable.
3. In general, any interval of real numbers is uncountable.
Theorem 1.2.6: Let A1 , A2 , · · · , An be a finite family of countable sets. Then the Cartesian
product A1 × A2 × · · · × An is countable.
Examples:
1. Z × Z is countable. To prove this
 consider the fuction f : Z × Z −→⊆ N defined by

2a 3b if a ≥ 0, b ≥ 0
5−a 7b

if a < 0, b ≥ 0
f (a, b) = .


11a 13−b if a ≥ 0, b < 0
 −a −b
17 19 if a < 0, b < 0

In this case, f is both one-to-one and onto, and so Z × Z is countable.


2. Q ∼ Z × N ⊂ Z × Z. From Example 1 above, Z × Z is countable. Since Z × N is a
subset of a countable set, it is countable, by Theorem 1.2.3. Since Q is equivalent to a
countable set, it is countable.
3. The set N × N is countable. This is so because the function f : N × N −→ N defined
by f (m, n) = 2m 3n is both one-to-one and onto.
4. If A and B are countable, then A × B is countable.
Proof : Since A and B are countable, there exist bijections f : N −→ A and g : N −→ B.
Define h : N × N −→ A × B by h (m, n) = (f (m) , g (n)). The function h in this case

3
is both one-to-one and onto. Since N × N is countably infinite, there exists a bijection
k : N −→ N × N. Then l : N −→ A × B defined by l = h ◦ k is one-to-one and onto
(since composition of bijections is also a bijection). So the set A × B is countable.

2 Open and Closed Sets


2.1 Metric Spaces
Definition 2.1.1: Let X be a non-empty set. A metric or distance function or simply
distance on X is a function d : X × X −→ R (i.e., d is real-valued) with properties that:
M1 : d (x, y) ≥ 0 for all x, y ∈ X, with d (x, y) = 0 if and only if x = y (i.e., d is positive
definite).
M2 : d (x, y) = d (y, x) for all x, y ∈ X (i.e., d is symmetric).
M3 : d (x, y) ≤ d (x, z) + d (z, y) for all x, y, z ∈ X (triangular property).
Remark 2.1.1 :
(i ) Condition M1 follows from the other two since

2d (x, y) = d (x, y) + d (y, x) ≥ d (x, x) = 0.

(ii ) The name triangular property comes from the property of triangles where the sum of
the lengths of any two sides of a triangle is always greater than or equal to the length
of the remaining side.
(iii ) The distance function d on X is neither one-to-one (since d (x, y) = d (y, x) for all
x, y ∈ X) nor onto (since negatives in R do not have pre-images, by M1 ).
Definition 2.1.2: Let d be a metric on X. The pair (X, d) is called a metric space. Elements
in a metric space are called points.
Remark 2.1.2 : If it is clear what the distance function d on X is, we write the metric space
(X, d) as simply X.
Examples:
1. The most important examples of metric spaces are the Euclidean spaces Rn especially
R (the real line) and R2 (the complex plane). The distance d : Rn × Rn −→ R on Rn is
defined as follows:
Let x = (x1 , x2 , · · · , xn ) and y = (y1 , y2 , · · · , yn ) be points in Rn . Then

d (x, y) = |x − y|
= |(x1 − y1 , x2 − y2 , · · · , xn − yn )|
q
= (x1 − y1 )2 + (x2 − y2 )2 + · · · (xn − yn )2 (positive square root).

For instance,
(a) In R, d −1, 32 = −1 − 3
= | − 52 | = 25 .

2

(b) Let x = (x1 , x2 ) , y = (y1 , y2 ) ∈ R2 . Diagramatically, we have

4
y

b x = (x1 , x2 )

b
y = (y1 , y2 )

Then

d (x, y) = |x − y|
= |(x1 − y1 , x2 − y2 )|
q
= (x1 − y1 )2 + (x2 − y2 )2 (positive square root).

2. Let x = (x1 , x2 , · · · , xn ) , y = (y1 , y2 , · · · , yn ) ∈ Rn and let d : Rn × Rn −→ R be defined


by
d (x, y) = max {|xi − yi |} .
1≤i≤n

Then d is a metric on Rn .
Note 2.1.1 : Every subset Y of a metric space X is a metric space with the same distance
function since if conditions M1 , M2 , and M3 hold for all x, y, z ∈ X, they also hold if we
restrict x, y, z to lie in Y .
Definition 2.1.3: Let X 6= ∅. A topology on X is a collection T of subsets of X such that
(i ) ∅, X ∈ T .
(ii ) The union of any number of members of T is in T .
(iii ) Finite intersection of members of T is in T .
The elements of T in this case are called open (T -open) subsets of X.
Examples:
1. The set of all open intervals of all real numbers is a topology in R.
2. Let X = {a, b, c}. Then P = {∅, {a} , {b} , {c} , {a, b} , {a, c} , {b, c} , X}, the power set
of X, is a topology on X.
Definition 2.1.4: Let x ∈ Rn and let r > 0. Then
(a) the open ball with centre at x and radius r is the set of all y ∈ Rn such that |y −x| < r.
(b) the closed ball with centre at x and radius r is the set of all y ∈ Rn such that |y−x| ≤ r.
Examples:
1. In R, balls are intervals.

5
r1 r1 r2 r2

b b

x1 x2
Open Interval Closed Interval

Suppose x ∈ R and δ > 0. Then X = (x − δ, x + δ) is an open ball centered at x and


having radius δ since if y ∈ X, then x − δ < y < x + δ implying that −δ < y − x < δ
so that |y − x| < δ.
2. In R2 , balls are discs.

b
b b
b

b b

b b
b
b b
b b b
b
b
b b
b
b b
b
b b
b b
b
b b
b
b b
b b
b b b
b b
b b
b
b
b
b b
b b
b
b b b
b b
b b
b
b b
b
b
b
b

b b
b b
b b
b
b
b
b b

b b
b b b b
b
b
b
b
b b
b
b b
b
b
b
b
b
b
b
b

Open Disc Closed Disc

3. In R3 , balls are spheres.

2.2 Neighbourhoods
Definition 2.2.1: Let N ⊆ R and let x ∈ R. Then N is a neighbourhood (written nbhd) of
x if there exist δ > 0 such that (x − δ, x + δ) ⊆ N.
Remark 2.2.1 : Clearly, if N is a neighbourhood of x, then x ∈ N. This implies that if
x∈
/ N, then N cannot be a neighbourhood of x.
Examples:
1. Consider the set A = (1, 4). In this case, A is a neighbourhood of 3 since if we take
δ = 0.1, then (3 − 0.1, 3 + 0.1) = (2.9, 3.1) ⊂ A. In general, if 1 < x < 4, then A is a
neighbourhood of x. On the other hand, A is not a neighbourhood of 1 since for every
δ > 0, (1 − δ, 1) ⊂ (1 − δ, 1 + δ) but (1 − δ, 1) ∩ A = ∅ so that (1 − δ, 1 + δ) * A. In a
similar manner, for any δ > 0, (4, 4 + δ) ⊂ (4 − δ, 4 + δ) but (4, 4 + δ) ∩ A = ∅ so that
(4 − δ, 4 + δ) * A. Therefore, A is not a neighbourhood of 4.
2. In general, the set A = (a, b) or (a, b] or [a, b) or [a, b] is a neighbourhood of every x ∈ A
such that a < x < b. However, A is not a neighbourhood of x if x ∈ / (a, b). For instance,
in Example 1 above, A is not a neighbourhood of 0, 1, 4, 5.5, etc.
Theorem 2.2.1: If U is a neighbourhood of x and U ⊆ V , then V is also a neighbourhood
of x.

6
Proof : Suppose U is a neighbourhood of x. Then there exists δ > 0 so that (x − δ, x + δ) ⊆ U.
But U ⊆ V . So (x − δ, x + δ) ⊆ V . Hence V is also a neighbourhood of x.
Exercise:
Consider the set A = (−1, 3) ∪ {8}. Show that:
(a) A is a neighbourhood of 0.
(b) A is not a neighbourhood of either 3 or 8.
(c) A is not a neighbourhood of 5.

2.3 Interior Points


Definition 2.3.1: A point x is an interior point of a set A if there exists δ > 0 such that
(x − δ, x + δ) ⊆ A. In other words, x is an interior point of A if A is a neighbourhood of x.
The interior of A, denoted by int(A) or Å, is the set of all interior points of A.
To prove that x ∈ int (A), we show that there exists δ > 0 such that (x − δ, x + δ) ⊆ A, and
/ int (A), we show that for all δ > 0, (x − δ, x + δ) * A, i.e., (x − δ, x + δ)
to prove that x ∈
contains points not in A.
Proposition 2.3.1: Int (A) ⊆ A.
Proof : Suppose x ∈ int (A). Then there exists δ > 0 such that (x − δ, x + δ) ⊆ A. Since
x ∈ (x − δ, x + δ), then x ∈ A. Therefore, int (A) ⊆ A. @
Remark 2.3.1 : If x ∈ int (A), then x ∈ A.
Examples:
1. Every point of (a, b) where a < b is an interior point of (a, b). To show this, let x ∈ (a, b)
and let
min {|x − a|, |x − b|}
δ= .
2
Then (x − δ, x + δ) ⊂ (a, b). On the other hand, any point outside of (a, b) is not an 
interior point. Therefore, int ((a, b)) = (a, b). For instance, int −1, − 12 = −1, − 12 .
2. The points a and b are not interior points of [a, b] for a < b. However, every other point
of the interval is an interior point. Therefore, int ([a, b]) = (a, b).
3
3. Let A = [1, 2). Then 2
∈ int (A) but 1 ∈
/ int (A).
4. Let B = [1, 2] ∪ {3}. Then int (B) = (1, 2).
5. The sets Q and Qc = R\Q have no interior points, i.e., int (Q) = ∅ and int (Qc ) = ∅.
Proof : Suppose x ∈ int (Q). Then there exists δ > 0 such that (x − δ, x + δ) ⊆ Q. But
this can’t happen since there is at least one irrational number between x − δ and x + δ
(from the fact that every interval of positive length contains both a rational and an
irrational number). So int (Q) = ∅. A similar argument works for Qc .
6. Let x ∈ R and let δ = 1. Then (x − 1, x + 1) ⊂ R. So every real number is an interior
point of R. Therefore, int (R) = R.
Definition 2.3.2: Let A be a set of real numbers. The exterior of A, denoted by ext (A), is
the interior of its complement, i.e., ext (A) = int (Ac ).

7
Example: If A = (a, b), then ext (A) = int ((−∞, a] ∪ [b, ∞)) = (−∞, a) ∪ (b, ∞).
Exercise:
Show that if A, B ⊆ R, then:
(a) int (A ∩ B) = int (A) ∩ int (B).
(b) it is not necessarily that int (A ∪ B) = int (A) ∪ int (B).

2.4 Isolated Points


Definition 2.4.1: A point x is an isolated point of a set A if there exists δ > 0 such that
(x − δ, x + δ) ∩ A = {x}, i.e., x is an isolated point of A if there exists an open interval
containing x which does not contain any point of A different from x.
Remark 2.4.1 : If x is an isolated point of A, then x ∈ A.
Example:
Let A = [1, 3) ∪ {8}. For δ = 1, (8 − δ, 8 + δ) ∩ A = (7, 9) ∩ A = {8}. Hence 8 is an isolated
point of A. Now, suppose x ∈ [1, 3). Then for all δ > 0, (x − δ, x + δ) ∩ (A\ {x}) 6= ∅, i.e.,
(x − δ, x + δ) ∩ A 6= {x} for all δ > 0. Hence x is not an isolated point of A. Therefore, 8 is
the only isolated point of A.
Exercise:
Consider the set A = [1, 2] ∪ {3} ∪ (4, 5). Show that 3 is the only isolated point of A.

2.5 Limit Points


Definition 2.5.1: Let A ⊆ R. A point x ∈ R is a limit point or an accumulation point or
a cluster point of A if for all δ > 0, (x − δ, x + δ) ∩ (A\ {x}) 6= ∅, i.e., x is a limit point of
A if every neighbourhood of x contains at least one point y ∈ A different from x. In other
words, x is a limit point of A if for any δ > 0, there exists at least one point y ∈ A, with
y 6= x, such that |y − x| < δ. The set of all limit points of a set A, denoted by A′ , is called
the derived set of A.
Remark 2.5.1 :
1. A point x is a limit point of A if there are points in A that are arbitrarily close to x.
2. A limit point of A need not be an element of A. Hence A′ need not be a subset of A.
3. From definition, an isolated point of A can never be a limit point of A. In fact, if x ∈ A
/ A′ , then x is an isolated point of A. On the other hand, if x ∈ A′ , then x is
and x ∈
not an isolated point of A.
Proposition 2.5.1: For a set A, if x ∈ A and x ∈ / A′ , then x is an isolated point of A.
/ A′ , then there exists δ > 0 such that (x − δ, x + δ) ∩ (A\ {x}) = ∅.
Proof : Let x ∈ A. If x ∈
Since x ∈ (x − δ, x + δ) and x ∈ A, this means that (x − δ, x + δ) ∩ A = {x} for some δ > 0.
Hence x is an isolated point of A.
Examples:
1. Let A = (a, b) and let x ∈ A. Then x ∈ A′ since for any δ > 0, the interval (x − δ, x + δ)
contains points of A other than x (infinitely many points in fact). Hence A ⊆ A′ .

8
2. Consider the set A = (1, 3) ∪ {8}. Then for any δ > 0, (1, 1 + δ) ∩ (A\ {1}) 6= ∅. This
implies that (1 − δ, 1 + δ) ∩ (A\ {1}) 6= ∅ for any δ > 0. Hence 1 ∈ A′ . In a similar
manner, for any δ > 0, (3 − δ, 3) ∩ (A\ {3}) 6= ∅ so that (3 − δ, 3 + δ) ∩ (A\ {3}) 6= ∅
for any δ > 0, meaning 3 ∈ A′ . On the other hand, 8 ∈ / A′ since if we take δ = 1, then
(8 − δ, 8 + δ) ∩ (A\ {8}) = (7, 9) ∩ (A\ {8}) = ∅ (in fact, 8 is an isolated point of A).
Remark 2.5.2 : In general, if A = (a, b) or (a, b] or [a, b) or [a, b] for a < b, then a, b ∈ A′ .
This together with Example 1 above mean that [a, b] ⊆ A′ .
3. Let A = (a, b) or (a, b] or [a, b) or [a, b]. Then A′ = [a, b]. To show this, from Remark
2.5.2, we know that [a, b] ⊆ A′ . Hence it suffices to show that if x ∈ R and x ∈ / [a, b],

then x ∈ / A . Suppose x > b. Let δ < x − b. Then b < x − δ and b ∈ / (x − δ, x + δ) so
that (x − δ, x + δ) ∩ (A\ {x}) = ∅. Similarly, suppose x < a. Let δ < a − x. Then
a > x + δ and a ∈ / (x − δ, x + δ) so that (x − δ, x + δ) ∩ (A\ {x}) = ∅. In either case,
x∈/ A′ .
4. Z′ = ∅.
Proof : Suppose Z has a limit point x. Then for any δ > 0, (x − δ, x + δ)∩(Z\ {x}) 6= ∅.
But for any real number x, there exists n ∈ Z such that n < x ≤ n + 1. Now, if we let
δ < min {|x − n|, |x − n − 1|}, the interval n < x ≤ n + 1 contains no integers different
from x, i.e., (x − δ, x + δ) ∩ (Z\ {x}) = ∅. Hence no real number is a limit point of Z.
5. Q′ = R.
Proof : Suppose x ∈ R. Then for any δ > 0, the interval (x − δ, x + δ) contains infinitely
many points of Q other than x. Thus (x − δ, x + δ) ∩ (Q\ {x}) 6= ∅. Therefore, every
real number is a limit point of Q.
Exercise:
Show that if A = n1 : n ∈ N , then A′ = {0}.


Further Examples:
1. Prove that if A, B ⊂ R and A ⊆ B, then A′ ⊆ B ′ .
Solution: Let x ∈ A′ . Then for all δ > 0, (x − δ, x + δ) ∩ (A\ {x}) 6= ∅. If A ⊆ B, it
follows that (x − δ, x + δ) ∩ (B\ {x}) 6= ∅ for all δ > 0, and so x ∈ B ′ . Hence A′ ⊆ B ′ .
2. Prove that if A, B ⊂ R, then (A ∪ B)′ = A′ ∪ B ′ .
Solution: A ⊆ A ∪ B and, by Example 1 above, A′ ⊆ (A ∪ B)′ . Similarly, B ⊆ A ∪ B
and B ′ ⊆ (A ∪ B)′ . Hence A′ ∪ B ′ ⊆ (A ∪ B)′ .
Now, suppose x ∈ (A ∪ B)′ . Then
∀δ > 0, (x − δ, x + δ) ∩ [(A ∪ B) \ {x}] 6= ∅
⇒ ∀δ > 0, (x − δ, x + δ) ∩ [(A\ {x}) ∪ (B\ {x})] 6= ∅
⇒ ∀δ > 0, [(x − δ, x + δ) ∩ (A\ {x})] ∪ [(x − δ, x + δ) ∩ (B\ {x})] 6= ∅
⇒ ∀δ > 0, [(x − δ, x + δ) ∩ (A\ {x})] 6= ∅ or [(x − δ, x + δ) ∩ (B\ {x})] 6= ∅
⇒ x ∈ A′ or x ∈ B ′
⇒ x ∈ A′ ∪ B ′ .
Hence (A ∪ B)′ ⊆ A′ ∪ B ′ .

9
Therefore, (A ∪ B)′ = A′ ∪ B ′ .
Theorem 2.5.1: Let x ∈ R and X ⊂ R. If x is a limit point of X, then any neighbourhood
of x contains infinitely many members of X.
Proof : Let A be a neighbourhood of x containing only a finite number of points of X, i.e.,
A ∩ X is finite. Then A ∩ X\ {x} is also finite. Suppose A ∩ X\ {x} = {y1 , y2 , · · · , yn }. We
show that there exists δ > 0 such that (x − δ, x + δ) ∩ A does not contain any member of
X\ {x}. Since both A and (x − δ, x + δ) are neighbourhoods of x, so is their intersection.
This will prove that there is a neighbourhood of x containing no element of X\ {x} hence
proving x is not a limit point of X. Let δ = min {|x − y1 |, |x − y2 |, · · · , |x − yn |}. Since
x 6= yi (i = 1, · · · , n) and δ > 0, then (x − δ, x + δ) ∩ A contains no point of X other than x,
hence the proof. @
Corollary 2.5.1:
(a) If x has a neighbourhood which only contains finitely many members of X, then x
cannot be a limit point of X.
(b) A finite point set has no limit points.

2.6 Open Sets and Closed Sets


Definition 2.6.1: Let A ⊆ R. Then:
(a) A is open if for any x ∈ A, there exists δ > 0 such that (x − δ, x + δ) ⊆ A, i.e., A is
open if it is (or contains) a neighbourhood of each of its points. In other words, A is
open if every point of A is an interior point of A.
(b) A is closed if every limit point of A is a point of A, i.e., if A′ ⊆ A.
Theorem 2.6.1: A set A is open if and only if its complement is closed.
Proof : Suppose A is open and suppose x is a limit point of Ac . Then for every δ > 0,
(x − δ, x + δ) ∩ (Ac \ {x}) 6= ∅, i.e., for every δ > 0, the interval (x − δ, x + δ) contains a
point y ∈ Ac such that y 6= x. So (x − δ, x + δ) * A for every δ > 0. Hence x ∈ / int (A).
Since A is open, then x ∈ / A, meaning that x ∈ Ac . It now follows that Ac is a closed.
Conversely, suppose Ac is a closed. Let x ∈ A. Then x ∈ / Ac and x is not a limit point
of Ac (since a closed set contains all its limit points). So there exists δ > 0 such that
(x − δ, x + δ) ∩ Ac = ∅, i.e., (x − δ, x + δ) ⊆ A. Thus x ∈ int (A) and it follows that A is
open. @
Corollary 2.6.1: A set A is closed if and only if its complement is open.
Theorem 2.6.2: Let A ⊂ R. Then:
(a) Å is open.
(b) A is open if and only if A = Å.
(c) If B is open and B ⊆ A, then B ⊆ Å.
Proof :
(a) We need to show that if x ∈ Å, then there exists δ > 0 such that (x − δ, x + δ) ⊆ Å.
Suppose x ∈ Å. Then there exists δ > 0 such that (x − δ, x + δ) ⊆ A. We show that
(x − δ, x + δ) ⊆ Å by showing that every point of (x − δ, x + δ) is an interior point of

10
A. Let y ∈ (x − δ, x + δ) and let

min {|y − x + δ|, |y − x − δ|}


ε= .
2

Then (y − ε, y + ε) ⊂ (x − δ, x + δ) ⊆ A. So y ∈ Å by definition, and since y was


arbitrary, (x − δ, x + δ) ⊆ Å.
(b) Suppose A = Å. Then A is open since Å is open, from part (a) of the theorem.
Conversely, supose A is open. Then every point of A is an interior point of A. Hence
A ⊆ Å. But by definition, Å ⊆ A. Therefore, A = Å.
(c) Let x ∈ B. Then x is an interior point of of B since B is open. So there exists δ > 0
such that (x − δ, x + δ) ⊆ B. But since B ⊆ A, then (x − δ, x + δ) ⊆ A and it follows
that x ∈ Å. Therefore, B ⊆ Å. @
Remark 2.6.1 : Parts (a) and (c) of Theorem 2.6.2 show that Å is the largest open subset
of A.
Examples:
1. Let a, b ∈ R such that a < b. Then (a, b) is open since int ((a, b)) = (a, b). This proves
that every open interval is an open set. Similarly, the infinite intervals (−∞, a) and
(a, ∞) are open.
2. Let A = [a, b]. Then A is closed since it contains all its limit points (from a previous
example, A′ = A). Alternatively, it is closed since Ac = (−∞, a) ∪ (b, ∞) which is open.
3. Sets that are both open and closed:
From a previous example, int (R) = R. So R is open. But ∅c = R\∅ = R. So the empty
set ∅ is closed.
Now, since there are no elements in ∅, then for every x ∈ ∅, (x − δ, x + δ) ⊂ ∅ for any
δ > 0. Hence ∅ is open. But Rc = R\R = ∅. Hence R is closed.
Therefore, R and ∅ are sets that are both open and closed; in fact, these are the only
sets with this property.
4. Sets that are neither open nor closed :
(a) Let a, b ∈ R such that a < b. Consider the set A = [a, b). Then A is not open since
a ∈ A but a ∈ / Å. Similarly, A is not closed since b ∈ A′ but b ∈/ A. Alternatively,
consider A = (−∞, a) ∪ [b, ∞). In this case, Ac is not closed since a is a limit point
c

of Ac but a ∈ / Ac . So A is not open. Similarly, Ac is not open since b ∈ Ac but


b∈/ int (Ac ). So A is not closed either. Therefore, [a, b) is neither open nor closed.
The same is true for (a, b].
(b) The set Q is neither open nor closed. As seen earlier none of the points of Q is an
interior point. Thus Q is not open and so Qc is not closed. Similarly, none of the
points of Qc is an interior point. Thus Qc is not open and so (Qc )c = Q is not
closed. Therefore, Q and Qc are examples of subsets of R that are neither open
nor closed.
Note 2.6.1 : Being closed is not the opposite of being open, i.e., a set is not either open or
closed, it can be neither.

11
5. Let x ∈ R. Then {x} is closed since {x}c = (−∞, x) ∪ (x, ∞) is open. Alternatively,
{x} is closed since it contains all its limit points, namely none. On the other hand, {x}
is not open since there does not exist δ > 0 such that (x − δ, x + δ) ⊆ {x}.
Theorem 2.6.3: Let {Gi } be an infinite collection of open sets and {Hi } an infinite collection
of closed sets. Then:

(a) ∪ Gi is open (i.e., arbitrary union of open sets is open).
i=1
n
(b) ∩ Gi is open (i.e., finite intersection of open sets is open).
i=1
n
(c) ∪ Hi is closed (i.e., finite union of closed sets is closed).
i=1

(d ) ∩ Hi is closed (i.e., arbitrary intersection of closed sets is closed).
i=1

Proof :
∞ ∞
(a) We prove that if x ∈ ∪ Gi , then x is an interior point of ∪ Gi , i.e., there exists δ > 0
i=1 i=1
∞ ∞
such that (x − δ, x + δ) ⊆ ∪ Gi . Now, suppose x ∈ ∪ Gi . Then x ∈ Gi for at least
i=1 i=1
one i, and since Gi is open, x is an interior point of Gi . So there exists δi > 0 such

that (x − δi , x + δi ) ⊆ Gi . Therefore, (x − δi , x + δi ) ⊆ ∪ Gi .
i=1
n n n
(b) If ∩ Gi = ∅, then ∩ Gi is open since ∅ is open. On the other hand, suppose ∩ Gi 6= ∅
i=1 i=1 i=1
n
and let x ∈ ∩ Gi . Then x ∈ Gi for all i = 1, 2, · · · , n. But each Gi (i = 1, 2, · · · , n)
i=1
is open. So, for each i = 1, 2, · · · , n, there exists δi > 0 such that (x − δi , x + δi ) ⊆ Gi .
Now, let δ = min {δ1 , · · · , δn }. Since δi > 0 for all i = 1, 2, · · · , n, then δ > 0. Hence

(x − δ, x + δ) ⊆ Gi for every i = 1, 2, · · · , n. Hence (x − δ, x + δ) ⊆ ∩ Gi . Therefore,
i=1
n n n
x is an interior point of ∩ Gi , and since x is an arbitrary point of ∩ Gi , then ∩ Gi
i=1 i=1 i=1
is open.
Remark 2.6.2 : The intersection of an infinite number of open sets need not be open.
Counterexample: 
In R, let Gn = − n1 , n1 , n ∈ N, i.e., G1 = (−1, 1), G2 = − 12 , 12 , G3 = − 13 , 31 , · · · . Then
 

Gn is open for each n = 1, 2, · · · . However, ∩ Gn = {0}, which is not open (by Example 5
n=1
above).
(c) Hi is closed means Hic is open, for each i= 1, 2, ·· · , n. So, by part (b) of the theorem,
n n c
∩ Hic is open, and by Theorem 2.6.1, ∩ Hic is closed. But by the De Morgan’s
i=1  c n i=1
n n n
law, ∩ Hic = ∪ (Hic )c = ∪ Hi . Therefore, ∪ Hi is closed.
i=1 i=1 i=1 i=1

Remark 2.6.3 : The union of an infinite collection of closed sets need not be closed.
Counterexample:
 n n
, n ∈ N, i.e., H1 = − 12 , 12 , H2 = − 32 , 23 , H3 = − 34 , 34 , · · · .
      
In R, let Hn = − n+1 , n+1

Then Hn is closed for each n = 1, 2, · · · . However, ∪ Hn = (−1, 1), which is not closed.
n=1

12
(d ) Hi is closed means Hic is open, for eachi = 1, 2,· · · . So by part (a) of the theorem,
∞ ∞ c
∪ Hic is open, and by Theorem 2.6.1, ∪ Hic is closed. But by the De Morgan’s
i=1  c ∞ i=1
∞ ∞ ∞
law, ∪ Hic = ∩ (Hic )c = ∩ Hi . Therefore, ∩ Hi is closed. @
i=1 i=1 i=1 i=1

2.7 Closure of a Set


Definition 2.7.1: Let E ⊂ R and let E ′ be the set of all limit points of E. Then the closure
of E is the set E = E ∪ E ′ . A point of the closure of E is called an adherent point of E.
Thus, if x is an adherent point of E, then x ∈ E or x ∈ E ′ .
Remark 2.7.1 : E is the intersection of all closed supersets of E.
Examples:
1. (a, b) = [a, b].
2. Q = R.
3. Qc = R.
4. ∅ = ∅.
5. {x} = {x}.
6. If E = [−1, 0), then E ′ = [−1, 0] and E = E ∪ E ′ = [−1, 0].
7. If F = (0, 1] ∪ {3}, then F ′ = [0, 1] and F = F ∪ F ′ = [0, 1] ∪ {3}.
Theorem 2.7.1: Let E ⊂ R. Then:
(a) E is closed.
(b) E = E if and only if E is closed.
(c) If F is closed and E ⊂ F , then E ⊂ F .
Proof :
(a) Since E is the intersection of all closed supersets of E, then E is closed, by Theorem
2.6.3.
(b) Suppose E = E. From part (a) of the theorem, E is closed. Therefore, E is closed.
Conversely, suppose E is closed. Then E ′ ⊆ E, and E = E ∪ E ′ = E.
(c) Suppose F is closed and F ⊃ E. Then F ⊇ F ′ and F ′ ⊃ E ′ . Hence F ⊃ E ′ . Therefore,
F ⊃ E ∪ E ′ = E. @
Remark 2.7.2 : By parts (a) and (c) of Theorem 2.7.1, E is the smallest closed subset of
R that contains E.
Example:
Every singleton set {x} is closed since {x} = {x}.
Exercise:
1. Show that [1, 2) ∪ {3} is neither open nor closed.
2. Let A, B ⊆ R.
(a) Prove that:
(i ) if A ⊆ B, then A ⊆ B.

13
(ii ) A ∪ B = A ∪ B.
(iii ) A ∩ B ⊆ A ∩ B.
(iv ) if A is open, then A′ = A.
(b) Give an example such that A ∩ B 6= A ∩ B.

2.8 Boundary
Definition 2.8.1: Let A ⊆ R. The boundary of A, denoted bdry (A), is the set of points
x ∈ R such that every neighbourhood of x contains at least one point of A and at least one
point not in A. Alternatively, bdry (A) is the set A\Å, i.e., the set of points in the closure of
A not belonging to the interior of A. An element in the boundary of A is called a boundary
point of A.
Examples:
1. bdry ((0, 5)) = bdry ([0, 5)) = bdry ((0, 5]) = bdry ([0, 5]) = {0, 5}.
2. bdry (∅) = ∅.
3. bdry (Q) = bdry (Qc ) = R.
4. bdry (Z) = Z.
5. bdry (Q ∩ [0, 1]) = [0, 1].
Remark 2.8.1 :
1. bdry (A) is closed.
2. bdry (A) = bdry (Ac ).
Theorem 2.8.1: A set is closed if and only if it contains all its boundary points.
Examples:
1. bdry (Z) = Z and hence Z is closed.
2. bdry (Q) * Q and hence Q is not closed.
Corollary 2.8.1: A set is open if and only if it contains none of its boundary points.

2.9 Dense Sets


Definition 2.9.1: Let A ⊂ R. Then A is said to be dense (in R) if A ∩ (a, b) 6= ∅ for each
open interval (a, b). More generally, suppose A, B ⊂ R. If every open interval that intersects
B also intersects A, we say that A is dense in B.
Examples:
1. The set Q is dense in R since for any interval (a, b) in R, Q ∩ (a, b) 6= ∅.
2. Q is dense in Qc .
1 3

3. Z is not dense in R since ,
2 4
∩ Z = ∅.
4. N is not dense in R since (−1, 0) ∩ N = ∅.
5. N is not dense in Z since (−1, 1) ∩ Z 6= ∅ but (−1, 1) ∩ N = ∅.

14
Remark 2.9.1 : A set A is dense in a set B if every point of B is a limit point of A, or a
point of A, or both.
Exercise:
Prove that:
(a) if A is dense in B and C ⊂ B, then A is dense in C.
(b) A is dense in B if and only if A ⊃ B.
(c) every set A is dense in its closure A.
(d ) if A ⊂ B and A is dense in B, then A = B.
Definition 2.9.2: Let A ⊂ R. Then A is said to be nowhere dense in R provided that every
open interval I contains an open subinterval J such that A ∩ J = ∅. In other words, A is
nowhere dense if A contains no open interval.
Examples:
1. Any finite set is nowhere dense.
2. N is nowhere dense.
3. The set n1 : n ∈ N is nowhere dense.


4. Any finite union of nowhere dense sets is nowhere dense.


5. A countable intersection of nowhere dense sets need not be nowhere dense.
n
Theorem 2.9.1: Let A1 , A2 , · · · , An be nowhere dense in R. Then ∪ Ai is also nowhere
i=1
dense in R.
Proof : Let I be any open interval in R. We seek an open interval J ⊂ I such that J ∩ Ai = ∅
for each i = 1, 2, · · · , n. Since A1 is nowhere dense, there exists an open interval I1 ⊂ I such
that A1 ∩I1 = ∅. Now, A2 is also nowhere dense in R. So there exists an open interval I2 ⊂ I1
such that A2 ∩ I2 = ∅. Proceeding in this manner we obtain intervals I1 ⊃ I2 ⊃ I3 ⊃ · · · ⊃ In
such that Ai ∩ Ii = ∅ for i = 1, 2, · · · , n. Since In ⊂ Ii for i = 1, 2, · · · , n, then Ai ∩ In = ∅
for i = 1, 2, · · · , n. Thus
n  n n
∪ Ai ∩ In = ∪ (Ai ∩ In ) = ∪ ∅ = ∅,
i=1 i=1 i=1

which is what we sought to prove. @

2.10 Compact Sets


Definition 2.10.1: Let E ⊆ R. A collection T consisting of open sets is an open cover for
the set E if every element in E is in some set G from T , i.e., if E ⊂ ∪ G. The set E has a
G∈T
n
finite subcover from T if there exists G1 , G2 , · · · , Gn in T such that E ⊂ ∪ Gi .
i=1

Examples:
1. The collection T = 1 − n1 , 2 − n1 : n ∈ N is an open cover for the set E = [1/2, 2).
 

2. The collection T = {(−n, n) : n ∈ N} is an open cover for the set R = (−∞, ∞).
3. Let r ∈ R. Then T = {(r − 1, r + 1) : r ∈ R} is an open cover for the set R.

15
Remark 2.10.1 : The collections in Examples 1 and 2 above are countable while the col-
lection in Example 3 is uncountable.
Definition 2.10.2: The set E ⊆ R is a compact set if every open cover for E has a finite
subcover.
Example:
The set E = [a, ∞) is not compact since T = a − n1 , n : n ∈ N is an open cover for E,
 

with the property that no finite subcollection covers E. Observe that the set E in this case
is closed but not bounded.
Theorem 2.10.1: If E is a compact set, then E is bounded.
Proof : The collection T = {(−n, n) : n ∈ N} is an open cover for any set E ⊆ R. The union
of any finite subcollection will be an interval of the form (−n0 , n0 ). If E is compact, there
must be an interval (−n0 , n0 ), so that E ⊂ (−n0 , n0 ), and so E is bounded. @
Example:
The set E = [a, b) is not compact since the collection T = a − n1 , b− n1 : n ∈ N  is an
 

open cover for E, but the union of any finite subcollection has the form a − n11 , b − n12 , and
does not therefore cover E. Observe that the set E in this case is bounded but not closed.
Theorem 2.10.2: If E is a compact set, then E is closed.
Proof : If E is not
∞  closed,
 1 1
 ∞ be a limit point x of E with x ∈
there must / E. The collection
T = {Gn }n=1 = R\ x − n , x + n n=1 is an open cover for E. In order for E to be compact,
k
there must be a finite subcollection {Gni }ki=1 of T that covers E. Since ∪ Gni = Gn′ where
i=1
k
n′ = max {n1 , n2 , · · · , nk }, we have E ⊂ ∪ Gni = Gn′ = R\ x − n1′ , x + n1′ or equivalently,
 
i=1
E ∩ x − n1′ , x + n1′ = ∅. But E ∩ x − n1′ , x + n1′ = ∅ implies that x is an exterior point of
   

E, and therefore not a limit point of E. This contradiction means that E is not compact,
and the proof is complete. @
Remark 2.10.2 : Theorems 2.10.1 and 2.10.2 prove that a compact set need be both closed
and bounded. The next theorem asserts that these two conditions are also sufficient for
compactness.
Theorem 2.10.3 (Heine-Borel Theorem): The set E ⊆ R is compact if and only if E is
closed and bounded.

3 Upper Bounds and Lower Bounds


Definition 3.1.1: A real number m is called an upper bound (u.b) of a non-empty subset A
of R if x ≤ m for all x ∈ A. A real number n is called an lower bound (l.b) of a non-empty
subset A of R if n ≤ x for all x ∈ A.
The set of all upper bounds of a set A is denoted by UA while the set of all lower bounds of
A is denoted by LA .
Note 3.1.1 :
1. A lower bound or an upper bound of A need not be an element of A.

16
2. A lower bound or an upper bound of A is not unique.
Definition 3.1.2: If A has an upper bound, then A is said to be bounded above. If A has a
lower bound, then A is said to be bounded below. If A has both an upper bound and a lower
bound, then we say that A is bounded. If A is neither bounded nor bounded above or below,
we say that A is unbounded. By definition, A is bounded if A ⊆ [n, m] for some interval
[n, m] of finite length.
Examples:
1. Let A = (1, 3]. Then A = {x | 1 < x ≤ 3}. If n ∈ R such that n ≤ 1, then n ∈ LA , and
if m ∈ R such that m ≥ 3, then n ∈ UA . Therefore, A is bounded.
2. The set N = {1, 2, 3, 4, · · · } is bounded below but not above. For instance, −10 is a
lower bound of N (in fact, every real number n ≤ 1 is a lower bound of N).
Remark 3.1.1 : Examples 1 and 2 above show that a bounded set need not be countable
and vice versa.
n
3. The number 1 is an upper bound of the set B = 12 , 43 , 78 , · · · , 2 2−1

n , · · · . On the other
1 1
hand, 2 is a lower bound of B. In fact, if x ∈ B, then x ≥ 2 and x < 1.
4. The set C = n1 : n ∈ N , i.e., C = 1, 21 , 13 , 41 , · · · is bounded since if x ∈ C, then
 

0 < x ≤ 1. Each real number k ≥ 1 is an upper bound of C while each real number
l ≤ 0 is a lower bound of C. Hence C has infinitely many lower bounds and infinitely
many upper bounds.
5. The set Z = {· · · , −3, −2, −1, 0, 1, 2, 3, · · · } is unbounded.
6. The set Z2 = {z 2 | z ∈ Z} = {0, 1, 4, 9, 16, 25 · · · } is bounded below while the set
−Z2 = {−z 2 | z ∈ Z} = {· · · , −25, −16, −9, −4, −1, 0} is bounded above.
7. The set Y = {sin x | 0 ≤ x < 2π} is bounded by −1 and 1. On the other hand, the set
K = {tan x | 0 ≤ x < 2π} is unbounded. However, K ′ = {tan x | 0 ≤ x ≤ π/2} is bounded
below.
Proposition 3.1.1: Let A ⊂ R.
(a) If m, m′ ∈ R with m′ > m and m ∈ UA , then m′ ∈ UA .
(b) If n, n′ ∈ R with n′ < n and n ∈ LA , then n′ ∈ LA .
From Examples 2 and 3 above, infinitely many real numbers greater than −10 are lower
bounds of N, but there is no number less that 1 which is an upper bound of B. This leads
us to the concept of least upper bound and greatest lower bound.
Definition 3.1.3: Let A ⊂ R. An element u ∈ R is called the least upper bound (l.u.b) of
A or supremum of A, written l.u.b A or sup A, if:
(i ) u ∈ UA (i.e., u is an upper bound of A)
(ii ) u ≤ m for all m ∈ UA .
Definition 3.1.4: Let A ⊂ R. An element l ∈ R is called the greatest lower bound (g.l.b) of
A or infimum of A, written g.l.b A or inf A, if:
(i ) l ∈ LA (i.e., l is a lower bound of A)
(ii ) l ≥ n for all n ∈ LA .

17
Some Consequences of the Definitions
1. If α is an upper bound of a set A ⊂ R and α ∈ A, then α is the supremum of A.
Proof : Suppose α, β ∈ A and α, β ∈ UA . Since α ∈ UA and β ∈ A, then α ≥ β. In a
similarly manner, since β ∈ UA and α ∈ A, then β ≥ α. The two inequalities hold
simultaneously if and only if α = β. So we cannot have two upper bounds in a set.
Therefore, if α ∈ A and α ∈ UA , then α is the supremum of A. @
2. If a set has an infimum, then the infimum is unique.
Proof : Let A be a non-empty subset of R. Suppose α = inf A and also β = inf A. Then
α, β ∈ LA . If α = inf A, then α ≥ l for all l ∈ LA . But β ∈ LA . So α ≥ β. Similarly, if
β = inf A, and α ∈ LA , then β ≥ α. The inequalities α ≥ β and β ≥ α hold simultane-
ously if and only if α = β. Therefore, inf A is unique. @
Remark 3.1.2 : If A, B ⊂ R where A ⊆ B, then inf B ≤ inf A and sup A ≤ sup B.
Examples:
1. Let A = (3, 4). If n ∈ LA , then n ≤ 3, and if m ∈ UA , then m ≥ 4. Hence inf A = 3
and sup A = 4.
n
2. Let B = 21 , 34 , 87 , · · · , 2 2−1 1

n , · · · . Then inf B = 2 and sup B = 1. In this case, inf B ∈ B

and sup B ∈
/ B.
3. Inf N = 1 but and sup N = ∞ since N has no upper bound.
4. Let C = {x} for x ∈ R. Then inf C = sup C = x.
5. If D = 0, 12 , 32 , 43 , 65 , 67 , · · · , then inf D = 0 and sup D = 1.


6. The empty set ∅ is bounded since ∅ ⊂ [n, m] for any interval [n, m] of finite length.
Thus every number m ∈ R is an upper bound of ∅ and so ∅ does not have a supremum.
Similarly, ∅ does not have an infimum.

Completeness Axiom of Real Numbers (Least Upper Bound Axiom)


Statement: Every non-empty set of real numbers that is bounded above has a least upper
bound in R.

4 Sequences and Series of Real Numbers


4.1 Sequences and Subsequences
Definition 4.1.1: A sequence of real numbers is a function from positive integers (natural
numbers) into the set of real numbers, i.e., f : N −→ R, f (n) = an . The sequence may be
denoted as {an } or {an }∞n=1 . We may also write the sequence as {a1 , a2 , · · · , an , · · · } where
a1 is the first term of the sequence, a2 the second, and so on.
Remark 4.1.1 : The sequence illustration should not be confused with set notation. In sets,
the order of elements does not matter, but in sequences, it does. Thus a sequence {1, 2, 3, 4}
is different from {1, 3, 4, 2}. In set notation, these would be considered equal.

18
Examples:
1. {1} = {1, 1, 1, 1, · · · }.
2. {2n}∞
n=1 = {2, 4, 6, 8, · · · }.

3. (−1)n+1 n=1 = {1, −1, 1, −1, · · · }.


Definition 4.1.2: Let {an } be a sequence of real numbers and consider the sequence {ni }
of positive integers such that n1 < n2 < n3 < · · · . Then the sequence {ani } is called a sub-
sequence of {an }. If {an } = {a1 , a2 , a3 , a4 , · · · } is a sequence, then {ani } = {a21 , a42 , a63 , · · · }
is a subsequence.
Remark 4.1.2 : A subsequence {ani } of a sequence {an } is itself a sequence and the numbers
n1 , n2 , · · · themselves form a subsequence of the sequence of positive integers 1, 2, 3, · · · .
Examples:

1. Consider the sequence {(−1)n }n=1 = {−1, 1, −1, 1, −1, 1, · · · }. Then −1, −1, −1, · · ·

and 1, 1, 1, · · · are subsequences of {(−1)n }n=1 .
2. The sequence 1, 2, 1, 3, 1, 4, 1, 5, · · · has subsequences 1, 1, 1, 1, · · · and 1, 2, 3, 4, 5, · · · .

4.2 Convergent and Divergent Sequences


Definition 4.2.1: Let {an }∞ ∞
n=1 be a sequence of real numbers. We say that {an }n=1 has a
limit L ∈ R if for every ε > 0, there exists a positive integer N = N (ε) such that

|an − L| < ε for all n ≥ N.


In this case, we write lim an = L or an −→ L (as n −→ ∞).
n→∞

Remark 4.2.1 : In Definition 4.2.1,


(i ) the limit L must be a real number.
(ii ) the value of N will, in general, depend of the value of ε.
(iii ) the inequality |an − L| < ε must hold for all values of n except at most a finite
number - namely, n = 1, 2, · · · , N − 1.
(iv ) the proof that lim an = L consists, upon given ε > 0, of finding a value of N such
n→∞
that |an − L| < ε for all n ≥ N.
Definition 4.2.2: If a sequence {an }∞ ∞
n=1 of real numbers has a limit L, we say that {an }n=1
converges (or {an }∞
n=1 is convergent) to L.

Definition 4.2.3: If a sequence {an }∞ n=1 of real numbers does not have a limit, we say that
{an }n=1 diverges or {an }n=1 is divergent. A sequence {an }∞
∞ ∞
n=1 of real numbers diverges to
∞, written lim an = ∞ or an −→ ∞, if for any real number M > 0, there exist N ∈ N such
n→∞
that an ≥ M for all n ≥ N. A sequence {an }∞
n=1 of real numbers diverges to −∞, written
lim an = −∞ or an −→ −∞, if for any real number M > 0, there exist N ∈ N such that
n→∞
an < −M for all n ≥ N. If the sequence {an }∞
n=1 of real numbers diverges but lim an 6= ±∞, n→∞
we say that {an }∞
n=1 oscillates.

19
Theorem 4.2.1: A point x is a limit point of a set A if and only if there exists a sequence
{xn }∞
n=1 in A such that lim xn = x.
n→∞

Examples:
1. The sequence n1 = 1, 21 , 13 , 41 , · · · converges to 0.
 

Proof : Given ε > 0 we find N = N (ε) ∈ N such that


1
− 0 < ε for all n ≥ N (1)
n
or
1
< ε for all n ≥ N. (2)
n
If we choose N such that N1 < ε, then (1) and (2) will hold since n1 ≤ N1 if n ≥ N. Now,
1
N
< ε if and only if n > 1ε . So, we can choose N ∈ N such that N > 1ε proving that
1
n
−→ 0.
Note 4.2.1 : The limit 0 if the sequence in Example 1 above is not equal to any term of the
sequence.
2. The sequence {an }∞ ∞
n=1 = {n}n=1 = {1, 2, 3, · · · } diverges.
Proof : Suppose the contrary, i.e., an −→ L for some L ∈ R. Then for every ε > 0 there
exists a positive integer N = N (ε) for which
|n − L| < ε (n ≥ N) .
In particular, there is an N for which
|n − L| < 1 (n ≥ N)
or
−1 < n − L < 1 (n ≥ N)
or
L−1< n<L+1 (n ≥ N) .
This means that all values of n greater than or equal to N lie between L − 1 and L + 1,
a contradiction. Hence the sequence diverges.
Remark 4.2.2 : It is obvious that {an }∞ ∞
n=1 = {n}n=1 diverges to ∞, by Definition 4.2.3. For
given M > 0, choosing N ∈ N such that N ≥ M, then an = n ≥ M for all n ≥ N.

3. The sequence {(−1)n }n=1 diverges.
Proof : The terms of the sequence are −1, 1, −1, 1, · · · . Suppose the sequence converges
1 1

to L for some L ∈ R. Then for ε = 2 there is an N = N 2 ∈ N such that
1
|(−1)n − L| < (n ≥ N) . (1)
2
For n even (1) becomes
1
|1 − L| < (n ≥ N) , (2)
2

20
and for n odd (1) becomes
1
| − 1 − L| < (n ≥ N) . (3)
2
In this case, (2) implies that L is less than 21 unit away from 1 while (3) implies that L
is less than 21 unit away from −1, a contradiction. [Alternatively, the inequality (3) is
equivalent to |1 + L| < 12 . But then

1 1
2 = | (1 + L) + (1 − L) | ≤ |1 + L| + |1 − L| < + = 1,
2 2
which is a contradiction.] Hence the sequence does not have a limit.
n o∞
2n
4. The sequence n+4n 1/2 converges to 2.
n=1
Proof : Given ε > 0 we find (calculate) N ∈ N such that

2n
−2 < ε (n ≥ N) . (1)
n + 4n1/2

The ineqiality (1) is equivalent to

2n − 2n − 8n1/2
<ε (n ≥ N)
n + 4n1/2
or
8n1/2
<ε (n ≥ N) . (2)
n + 4n1/2
1
8n /2 8
Now the left hand side of (2) is less than n
= n1/2
. Hence (2) will be true if

8
<ε (n ≥ N) . (3)
n1/2
8 64
If we choose N such that N 1/2
< ε, that is, choose N > ε2
, then (3) will certainly be
8 8
true. (For n1/2
≤ N 1/2
if n ≥ N). We have thus shown that if N is any positive integer
64
such that N > ε2
, then (3) and hence (2) and finally (1) will be true. This proves that
2n
lim n+4n 1/2 = 2.
n→∞

5. The sequence {log (1/n)}∞


n=1 diverges to −∞.
Proof : Given M > 0 we find N ∈ N such that
1
log < −M (n ≥ N) . (1)
n
But this is equivalent to
log n > M (n ≥ N) ,
or
n > eM (n ≥ N) . (2)

21
Thus if we choose N ≥ eM , the (2) and hence (1) will hold.
Exercise:

Use definition to show that the sequence {an }∞ (−1)n+1

n=1 = n=1
has no limit.
Theorem 4.2.2: Let {an }∞
n=1 be a sequence of non-negative numbers. If lim an = L, then
n→∞
L ≥ 0.
Proof : Supose the contrary, i.e., L < 0. Then for ε = − L2 there exists N ∈ N such that

L
|an − L| < − (n ≥ N) .
2
In particular,
L
|aN − L| < − ,
2
which implies
L
aN − L < −
2
or
L
aN < < 0 for L < 0,
2
a contradiction since, by hypothesis, aN ≥ 0. Hence L ≥ 0. @
Theorem 4.2.3: If {an }∞n=1 converges, then its limit is unique.
Proof : Suppose an −→ L and an −→ M. We show that L = M. Suppose the contrary, i.e.,
L 6= M so that |M − L| > 0. Let ε = 12 |M − L|. By hypothesis an −→ L. So there exists
N1 = N1 (ε) ∈ N such that
|an − L| < ε (n ≥ N1 ) .
Similarly, since an −→ M there exists N2 = N2 (ε) ∈ N such that

|an − M| < ε (n ≥ N2 ) .

Let N = max {N1 , N2 }. Then N ≥ N1 and N ≥ N2 so that

|an − L| < ε (n ≥ N)

and
|an − M| < ε (n ≥ N) .
Thus

|M − L| = |(an − L) − (an − M)| ≤ |an − L| + |an − M| < 2ε = |M − L|.

This contradiction shows that M = L, which is what we wished to show. @


Definition 4.2.4: Let {ani } be a subsequence of a sequence {an } of real numbers. If {ani }
converges, its limit is called a subsequential limit of {an }.
Theorem 4.2.4: A sequence {an }∞ n=1 of real numbers is convergent to L if and only if every

subsequence of {an }n=1 is also convergent to L.

22
Corollary 4.2.1: All subsequences of a convergent sequence {an }∞ n=1 of real numbers con-
verge to the same limit.
Proof : If the sequence {an }∞ ∞
n=1 converges to L, then, by Theorem 4.2.3, {an }n=1 converges
to no other limit. By Theorem 4.2.4, then, all subsequences of {an }∞
n=1 converge to L (and
to no other limit). @
Examples:
1. The sequence 1, − 21 , 13 , − 41 , 15 , − 61 , · · · converges to 0. By Corollary 4.2.1, all its subse-
quences converge to 0. For instace, the subsequences 1, 31 , 51 , · · · and − 12 , − 41 , − 16 , · · ·
both converge to 0.

2. Consider the sequence {(−1)n }n=1 . Then −1, −1, −1, −1, · · · and 1, 1, 1, 1, · · · are sub-

sequences of {(−1)n }n=1 which converge respectively to −1 and 1. Since the two sub-

sequences converge to different limits, then, by Theorem 4.2.4, the sequence {(−1)n }n=1
diverges. This example shows that a divergent sequence may have a convergent subse-
quence.
3. The sequence {n}∞ n=1 shows that a divergent sequence need have no convergent subse-
quence.
4. Consider the sequence 1, 2, 1, 3, 1, 4, 1, 5, · · · . By Corollary 4.2.1, the sequence diverges
since it has the divergent subsequence 1, 2, 3, 4, 5, · · · . Moreover, the sequence does not
diverge to ∞ since there is no N ∈ N for which an > 2 for all n ≥ N. The sequence
obviously does not diverge to −∞. Hence it oscillates.
5. The sequence {an }∞ n=1 = {1, −2, 3, −4, 5, −6, · · · } has the subsequence 1, 3, 5, · · · which
diverges to ∞ and also has the subsequence −2, −4, −6, · · · which diverges to −∞. The
sequence {an }∞n=1 diverges, but lim an 6= ±∞; the sequence oscillates.
n→∞

4.3 Bounded Sequences


Since a sequence {an }∞n=1 of real numbers is a function from N into R, the range of the
sequence, namely {a1 , a2 , · · · }, is a subset of R. The range of a sequence may be finite or it
may be infinite.
Definition 4.3.1: A sequence {an }∞ ∞
n=1 is bounded above if the range of {an }n=1 is bounded
above, it is bounded below if its range is bounded below, and it is bounded if its range is
bounded. Thus {an }∞n=1 is bounded if and only if there exists M ∈ R such that

|an | ≤ M for all n ∈ N.

Example:
Consider the sequence {an } defined by
(
1 if n is even
an = ,
0 if n is odd

i.e., {an } = {0, 1, 0, 1, 0, 1, · · · }. Then the range of {an } is the set {0, 1} which is bounded.
Therefore, the sequence is bounded.

23
If a sequence diverges to ∞ or to −∞, then it is not bounded. A sequence that diverges to
∞ must, however, be bounded below (since such a sequence can have only a finite number
of negative terms). Similarly, if a sequence diverges to −∞, then it is bounded above. An
oscillating sequence may or may not be bounded.
Examples:
1. The sequence 1, −2, 3, −4, 5, −6, · · · oscillates and is neither bounded above nor bounded
below.
2. The sequence −1, 1, −1, 1, −1, 1, · · · oscillates between −1 and 1; it is bounded.
3. The sequence 1, 2, 1, 3, 1, 4, 1, 5, · · · oscillates and is bounded below but is not bounded
above.
Theorem 4.3.1: If the sequence {an }∞
n=1 of real numbers is convergent, then it is bounded.
Proof : Suppose an −→ L. Then given ε = 1, there exists N ∈ N such that

|an − L| < 1 for all n ≥ N.

Since |an | = |L + (an − L)| ≤ |L| + |an − L|, then

|an | < |L| + 1 for all n ≥ N. (1)

If we let M = max {|a1 | , |a2 | , · · · , |aN −1 |}, then

|an | ≤ M for all n = 1, · · · , N − 1. (2)

From (1) and (2) we have that

|an | < M + |L| + 1 for all n ∈ N,

which shows that {an }∞


n=1 is bounded. @

Remark 4.3.1 :
1. The converse of Theorem 4.3.1 is not necessarily true, i.e., a bounded sequence need
not be convergent.
2. From Theorem 4.3.1, an unbounded sequence is divergent.
Examples:

1. The sequence {(−1)n }n=1 is bounded but not convergent.
2. The sequence 1, −2, 3, −4, 5, −6, · · · is divergent since it is neither bounded above nor
bounded below.
Theorem 4.3.2 (Bolzano-Weierstrass Theorem): Let {an }∞ n=1 be a bounded sequence
of real numbers. Then {an }∞
n=1 has a convergent subsequence.

4.4 Monotonic Sequences


As seen in Remark 4.3.1, a bounded sequence need not be convergent. In this section we
consider a condition which, together with boundedness, will ensure that a sequence is con-
vergent.

24
Definition 4.4.1: The sequence {an }∞ n=1 is non-decreasing if an ≤ an+1 for all n; in the
special case where an < an+1 for all n, the sequence {an }∞ n=1 is said to be increasing. The
sequence {an }∞ n=1 is non-increasing if an ≥ a n+1 for all n; in the special case where an > an+1

for all n, the sequence {an }n=1 is said to be decreasing. A monotonic sequence is a sequence
which is either non-increasing or non-decreasing.
Remark 4.4.1 :
1. A non-decreasing sequence {an }∞
n=1 is always bounded below (by a1 ).

2. A non-increasing sequence {an }∞


n=1 is always bounded above (by a1 ).

Examples:
1 ∞
= 1, 1 21 , 1 34 , 1 87 , · · ·
 
1. The sequence 2 − 2n−1 n=1
is non-decreasing and bounded.
2. The sequence {n}∞
n=1 = {1, 2, 3, 4, · · · } is non-decreasing and bounded below (but not
bounded above).
Theorem 4.4.1: A non-decreasing sequence which is bounded above is convergent.
Proof : Suppose {an }∞
n=1 is non-decreasing and bounded above. Then the set A = {a1 , a2 , · · · }
is a non-empty subset of R which is bounded above. By the completeness axiom of real
numbers, this set has a least upper bound. Let

M = l.u.b {a1 , a2 , · · · } = l.u.b of A.

We will prove that an −→ M as n −→ ∞. Given ε > 0 the number M − ε is not an upper


bound for A. Hence, for some N ∈ N, aN > M − ε. But, since {an }∞
n=1 is non-decreasing,
this implies
an > M − ε for all n ≥ N. (1)
On the other hand, since M is an upper bound for A,

M ≥ an for all n ∈ N. (2)

Form (1) and (2) we conclude that

|an − M| < ε for all n ≥ N.

This proves that lim an = M. @


n→∞

Examples:  1 ∞
1. As seen in a previous example, the sequence 2 − 2n−1 n=1
is non-decreasing and
bounded. It converges to 2.
1 n
2. Consider the sequence {an }∞

n=1 where a n = 1 + n
. It can be shown that an ≤ an+1

for every n ∈ N, i.e., {an }n=1 is non-decreasing, and that an < 3 for every n ∈ N, i.e.,
{an }∞ ∞
n=1 is bounded above (by 3). Hence, by Theorem 4.4.1, {an }n=1 is convergent. In
fact, lim an = e where the number e has decimal expansion 2.7182818 · · · .
n→∞

Remark 4.4.2 :
1. Theorem 4.4.1 gives a set of criteria that enables us to prove that a sequence converges

25
without first guessing its limit.
2. A non-decreasing sequence does not oscillate, imlying that a non-decreasing sequence
that is not bounded above diverges to ∞.
Theorem 4.4.2: A non-decreasing sequence which is not bounded above diverges to ∞.
Proof : Suppose {an }∞
n=1 is non-decreasing but not bounded above. Given M > 0 we find
N ∈ N such that
an > M for all n ≥ N. (1)
Now, since M is not an upper bound for {a1 , a2 , · · · } there must exist N ∈ N such that
aN > M. Then, for this N, (1) follows from the hypothesis that {an }∞ n=1 is non-decreasing.
This proves the theorem. @
Theorem 4.4.3: Let {an }∞ n=1 be a non-increasing sequence of real numbers.

(a) If {an }n=1 is bounded below, it is convergent.
(b) If {an }∞
n=1 is not bounded below, it diverges to −∞.
Proof : The proof of the theorem follows the proofs of Theorems 4.4.1 and 4.4.2 exactly,
with all upper bounds and least upper bounds replaced by lower bounds and greatest lower
bounds. @
Remark 4.4.3 : From Theorems 4.4.1 and 4.4.3, a bounded non-decreasing or non-increasing
sequence is convergent. Hence convergence and boundedness are equivalent for monotonic
sequences of real numbers.
Theorem 4.4.4: Let {an }∞ ∞
n=1 be a sequence of real numbers. Then {an }n=1 has a monotonic
subsequence.

4.5 Operations on Convergent Sequences


Remark 4.5.1 : Since sequences of real numbers are real-valued functions, the definition of
the sum, difference, product and quotient of sequences follows the properties of real-valued
functions.
If {an }∞ ∞ ∞ ∞ ∞
n=1 and {bn }n=1 are sequences of real numbers, then {an }n=1 +{bn }n=1 = {an + bn }n=1
and {an }∞ ∞ ∞ ∞ ∞
n=1 · {bn }n=1 = {an · bn }n=1 etc. Also, if c ∈ R, then c {an }n=1 = {can }n=1 .

Theorem 4.5.1: Let {an }∞ ∞


n=1 and {bn }n=1 be sequences of real numbers. If lim an = a and
n→∞
lim bn = b, then lim (an + bn ) = a + b.
n→∞ n→∞
Proof : Given ε > 0 we find N = N (ε) ∈ N such that

|(an + bn ) − (a + b)| < ε for all n ≥ N. (1)

Now,

|(an + bn ) − (a + b)| = |(an − a) + (bn − b)| ≤ |an − a| + |bn − b| .

Hence (1) holds if


|an − a| + |bn − b| < ε for all n ≥ N.

26
Thus we try to make both |an − a| and |bn − b| less than ε/2 by taking n sufficiently large.
Since an −→ a there exists N1 ∈ N such that |an − a| < ε/2 for all n ≥ N1 . Also, since
bn −→ b, there exists N2 ∈ N such that |bn − b| < ε/2 for all n ≥ N2 .
Let N = max {N1 , N2 }. Then |an − a| < ε/2 for all n ≥ N and |bn − b| < ε/2 for all n ≥ N so
that
|an − a| + |bn − b| < ε/2 + ε/2 = ε for all n ≥ N,
hence the proof. @
Theorem 4.5.2: Let {an }∞ n=1 be a sequence of real numbers, and let c ∈ R. If n→∞
lim an = a,
then lim can = ca.
n→∞
Proof : If c = 0, the theorem is obvious. Now, suppose c 6= 0. Given ε > 0 we find
N = N (ε) ∈ N such that

|can − ca| < ε for all n ≥ N. (1)

Now, since an −→ a there exists N ∈ N such that

|an − a| < ε/|c| for all n ≥ N.

But then
|c| · |an − a| < ε for all n ≥ N,
which is equivalent to (1). @
Theorem 4.5.3:
(a) If 0 < x < 1, then {xn }∞
n=1 converges to 0.

(b) If 1 < x < ∞, then {xn }∞n=1 diverges to ∞.


Proof :
(a) If 0 < x < 1, then xn+1 = x · xn < xn . Hence {xn }∞ n
n=1 is non-increasing. Since x > 0
for n ∈ N, {xn }∞ n ∞
n=1 is bounded below. By Theorem 4.4.3, {x }n=1 is convergent. Let
L = lim xn . From Theorem 4.5.2 (with c = x) it follows that lim x · xn = xL. That is,
n→∞ n→∞
∞ ∞
{xn+1 }n=1 converges to xL. But {xn+1 }n=1 is a subsequence of {xn }∞
n=1 . By Corollary
4.2.1, L = xL and so L (1 − x) = 0. Since x 6= 1, this shows that L = 0, and part (a)
is proved.
(b) If x > 1, then xn+1 = x · xn > xn so that {xn }∞
n=1 is non-decreasing. We will show that
{xn }∞n=1 is not bounded above. For if {xn ∞
} n=1 were bounded above, then by Theorem
n ∞
4.4.1 {x }n=1 would converge to some L ∈ R. But the same reasoning as in part (a)
would show that L = Lx, so that L = 0 = lim xn . But xn > 1 and so {xn }∞ n=1 obviously
n→∞
cannot converge to 0. This contradiction proves that {xn }∞
n=1 is not bounded above.
The conclusion follows from Theorem 4.4.2. @
Theorem 4.5.4: Let {an }∞ ∞
n=1 and {bn }n=1 be sequences of real numbers. If lim an = a and
n→∞
lim bn = b, then lim (an − bn ) = a − b.
n→∞ n→∞

27
Proof : Since lim bn = b, it follows from Theorem 4.5.2 (with c = −1) that lim (−bn ) = −b.
n→∞ n→∞
But then, using Theorem 4.5.1,

lim (an − bn ) = lim [an + (−bn )] = lim an + lim (−bn ) = a + (−b) = a − b,


n→∞ n→∞ n→∞ n→∞

which is what we wished to prove. @


Corollary 4.5.1: Let {an }∞ ∞
n=1 and {bn }n=1 be sequences of real numbers such that an ≤ bn
for all n ∈ N. If lim an = a and lim bn = b, then a ≤ b.
n→∞ n→∞
Proof : By Theorem 4.5.4, b − a = lim (bn − an ). But bn − an ≥ 0 for all n ∈ N. Hence by
n→∞
Theorem 4.2.2, b − a ≥ 0, which establishes the result. @
Remark 4.5.1 : Corollary 4.5.1 remains true even if an > bn for a finite number of values
of n.
Theorem 4.5.5: Let {an }∞ ∞
n=1 and {bn }n=1 be sequences of real numbers. If n→∞
lim an = a and
lim bn = b, then lim an bn = ab.
n→∞ n→∞
Proof : Given ε > 0 we find N ∈ N such that

|an bn − ab| < ε for all n ≥ N. (1)

Now,

|an bn − ab| = |(an bn − abn ) + (abn − ab)|


≤ |an bn − abn | + |abn − ab|
= |bn | · |an − a| + |a| · |bn − b| .

Hence (1) holds if

|bn | · |an − a| + |a| · |bn − b| < ε for all n ≥ N. (2)

Since {bn }∞
n=1 is convergent, then by Theorem 4.3.1, it is bounded. So there exists M > 0
such that |bn | ≤ M for all n ∈ N. Then (2) will certainly hold if

M |an − a| + |a| · |bn − b| < ε for all n ≥ N.

Now, choose N1 ∈ N so that


ε
M |an − a| < for all n ≥ N1 ,
2
and choose N2 ∈ N such that
ε
|a| · |bn − b| < for all n ≥ N2 .
2
ε ε
If N = max {N1 , N2 }, then M |an − a| < 2
for all n ≥ N and |a| · |bn − b| < 2
for all n ≥ N,
and it follows that
ε ε
M |an − a| + |a| · |bn − b| < + = ε for all n ≥ N. @
2 2

28
Lemma 4.5.1: Let {an }∞
n=1 be a sequences of real numbers such that lim an = a where n→∞
a 6= 0. Then lim (1/an ) = 1/a.
n→∞
Proof : Either a > 0 or a < 0. We prove the case a > 0 (the case a < 0 can be proved by
applying the first case to {−an }∞
n=1 ). Suppose a > 0. Given ε > 0 we find N = N (ε) ∈ N
such that
|1/an − 1/a| < ε for all n ≥ N
or
|an − a|
< ε for all n ≥ N (1)
|an a|
Now, there exists N1 ∈ N such that |an − a| < a2 for all n ≥ N1 . This implies that
a
an > for all n ≥ N1
2
so that
a2
for all n ≥ N1 .
an a >
2
In addition, there exists N2 ∈ N such that
a2 ε
|an − a| < for all n ≥ N2 .
2
Thus if N = max {N1 , N2 }, we have, for n ≥ N,
|an − a| 1 1 a2 ε
= · |an − a| < a2 · = ε.
|an a| |an a| /2 2
Hence (1) holds for this N. This completes the proof. @
Theorem 4.5.6: Let {an }∞ ∞
n=1 and {bn }n=1 be sequences of real numbers. If lim an = a and n→∞
lim bn = b where b 6= 0. Then lim (ab/bn ) = a/b.
n→∞ n→∞
Proof : Using Theorem 4.5.5 and Lemma 4.5.1 we have
1 1
lim an · =a· ,
n→∞ bn b
which is what we wished to show. @
2
Example: Prove that lim 3n −6n
2 = 3/5.
n→∞ 5n +4
Solution:
3n2 − 6n 3 − 6/n lim (3 − 6/n)
n→∞
lim = lim =
n→∞ 5n2 + 4 n→∞ 5 + 4/n2 lim (5 + 4/n2 )
n→∞
lim 3 + lim (−6/n) lim 3 − 6 lim 1/n
n→∞ n→∞ n→∞ n→∞
= = 2
lim 5 + lim 4/n2

n→∞ n→∞ lim 5 + 4 lim /n1
n→∞ n→∞
3 − 6 (0)
= 2 = /5.
3
5 + 4 (0)

29
4.6 Operations on Divergent Sequences
If {an }∞ ∞
n=1 is a divergent sequence, then {−an }n=1 is also divergent, and the sum of these
two sequences is clearly not divergent. Moreover, the product of the divergent sequence

{(−1)n }n=1 with itself is not divergent. For sequences that diverge to ∞ we have the following
results:
Theorem 4.6.1: Let {an }∞ ∞
n=1 and {bn }n=1 be sequences of real numbers that diverge to ∞.
Then {an + bn }∞ ∞
n=1 and {an bn }n=1 diverge to ∞ also.
Proof : Given M > 0, choose N1 ∈ N such that

an > M for all n ≥ N1 ,

and choose N2 ∈ N such that


bn > 1 for all n ≥ N2 .
(The above is possible since both an −→ ∞ and bn −→ ∞ as n −→ ∞.) Then, for
N = max {N1 , N2 }, we have

an + bn > M + 1 > M for all n ≥ N,

and
an bn > M · 1 = M for all n ≥ N,
hence the proof. @
Theorem 4.6.2: Let {an }∞ ∞ ∞
n=1 and {bn }n=1 be sequences of real numbers. If {an }n=1 diverges
to ∞ and {bn }∞ ∞
n=1 is bounded, then {an + bn }n=1 diverges to ∞.
Proof : By hypothesis there exists Q > 0 such that

|bn | ≤ Q for all n ∈ N.

Given M > 0 choose N ∈ N such that

an > M + Q for all n ≥ N.

Then
an + bn ≥ an − |bn | > (M + Q) − Q = M for all n ≥ N.
That is,
an + bn > M for all n ≥ N
which shows that an + bn −→ ∞ as n −→ ∞. @
Corollary 4.6.1: If {an }∞ ∞ ∞
n=1 diverges to ∞ and {bn }n=1 converges, then {an + bn }n=1 di-
verges to ∞.
Proof : The proof follows directly from Theorems 4.3.1 and 4.6.2. @
Remark 4.6.1 : Theorems 4.6.1 and 4.6.2, and Corollary 4.6.1 hold if we replace ∞ with
−∞.

30
4.7 Limit Superior and Limit Inferior
Consider a sequence {an }∞ n=1 that is bounded above, say an ≤ M for all n ∈ N. Then for a
fixed n ∈ N, the set {an , an+1 , an+2 , · · · } is bounded above and hence has a least upper bound
Mn = l.u.b {an , an+1 , an+2 , · · · }. Moreover, Mn ≥ Mn+1 since Mn+1 = l.u.b {an+1 , an+2 , · · · }
is the least upper bound of a subset of {an , an+1 , an+2 , · · · }. Thus the sequence {Mn }∞ n=1 is
non-increasing and thus either converges or diverges to −∞.
Definition 4.7.1: Let {an }∞ n=1 be a sequence of real numbers that is bounded above, and
let Mn = l.u.b {an , an+1 , an+2 , · · · }.
(a) If {Mn }∞
n=1 converges, then lim sup an = lim Mn .
n−→∞ n→∞

(b) If {Mn }∞
n=1 diverges to −∞ then lim sup an = −∞.
n−→∞

Definition 4.7.2: If {an }∞


n=1 is a sequence of real numbers that is not bounded above, we
write lim sup an = ∞.
n−→∞

Examples:
1. Let an = (−1)n , n ∈ N. Then {an }∞
n=1 that is bounded above. In this case, Mn = 1 for
all n ∈ N and hence lim Mn = 1. Thus lim sup (−1)n = 1.
n→∞ n−→∞

2. Suppose an = −n, n ∈ N. Then Mn = l.u.b {−n, −n − 1, −n − 2, · · · } = −n. In this


case, Mn −→ −∞ as n −→ ∞ and so lim sup (−n) = −∞.
n−→∞

3. Consider the sequence an = 1, −1, 1, −2, 1, −3, 1, −4, · · · . Then Mn = 1 for all n ∈ N.
Hence lim sup an = 1.
n−→∞

4. The sequence {n}∞


n=1 is not bounded above. Hence, by Definition 4.7.2, lim sup n = ∞.
n−→∞

Theorem 4.7.1: If {an }∞


n=1 is convergent sequence of real numbers, then
lim sup an = lim an .
n−→∞ n→∞

Proof : Let L = lim an . Then given ε > 0 there exists N ∈ N such that
n→∞

|an − L| < ε for all n ≥ N,


or
L − ε < an < L + ε for all n ≥ N.
Thus if n ≥ N, then L + ε is an upper bound for {an , an+1 , an+2 , · · · } and L − ε is not a lower
bound. Hence
L − ε < Mn = l.u.b {an , an+1 , an+2 , · · · } ≤ L + ε,
and so, by Corollary 4.5.1,
L − ε < lim Mn ≤ L + ε.
n→∞
But lim Mn = lim sup an . Thus
n→∞ n−→∞

L − ε < lim sup an ≤ L + ε.


n−→∞

31
Since ε was arbitrary, this imlies lim sup an = L, which is what we wished to prove. @
n−→∞

Now, suppose the sequence {an }n=1 of real numbers is bounded below. Then for a fixed
n ∈ N, the set {an , an+1 , an+2 , · · · } is bounded below and hence has a greatest lower bound.
Let mn = g.l.b {an , an+1 , an+2 , · · · }. Then {mn }∞ n=1 is non-decreasing sequence and hence
either converges or diverges to ∞.
Definition 4.7.3: Let {an }∞ n=1 be a sequence of real numbers that is bounded below, and
let mn = g.l.b {an , an+1 , an+2 , · · · }.
(a) If {mn }∞
n=1 converges, then lim inf an = lim mn .
n−→∞ n→∞

(b) If {mn }∞
n=1 diverges to ∞ then lim inf an = ∞.
n−→∞

Definition 4.7.4: If {an }∞


n=1 is a sequence of real numbers that is not bounded below, we
write lim inf an = −∞.
n−→∞

Examples:
n ∞
1. The sequence {an }∞
n=1 = {(−1) }n=1 is bounded below. In this case, mn = −1 for every
n ∈ N and hence lim inf (−1)n = lim mn = −1.
n−→∞ n→∞

2. Let an = n, n ∈ N. Then mn = g.l.b {n, n + 1, n + 2, · · · } = n. Hence lim mn = ∞ so


n→∞
that lim inf n = ∞.
n−→∞

3. The sequence an = 1, −1, 1, −2, 1, −3, 1, −4, · · · is not bounded below. Therefore, by
Definition 4.7.4, lim inf an = −∞.
n−→∞

4. The sequence {−n}∞


n=1 is not bounded below. Hence lim inf n = −∞.
n−→∞

Theorem 4.7.2: If {an }∞is convergent sequence of real numbers, then lim inf an = lim an .
n=1 n−→∞ n→∞
Proof : The proof of this theorem is similar to the proof of Theorem 4.7.1. @
Theorem 4.7.3: If {an }∞
n=1 is a sequence of real numbers, then lim inf an ≤ lim sup an .
n−→∞ n−→∞
Proof : If {an }∞
n=1 is bounded, then

mn = g.l.b {an , an+1 , an+2 , · · · } ≤ l.u.b {an , an+1 , an+2 , · · · } = Mn .


Thus mn ≤ Mn and so, by Corollary 4.5.1, lim inf an ≤ lim sup an . If {an }∞
n=1 is not bounded,
n−→∞ n−→∞
then lim sup an = ∞ or lim inf an = −∞ and lim inf an < lim sup an . @
n−→∞ n−→∞ n−→∞ n−→∞

Remark 4.7.1 : From Theorems 4.7.1 and 4.7.2 we see that if lim an = L, then we must
n→∞
have lim sup an = lim inf an = L.
n−→∞ n→∞

We now prove the converse of Remark 4.7.1.


Theorem 4.7.4: Let {an }∞
n=1 is a sequence of real numbers. If lim sup an = lim inf an = L
n−→∞ n−→∞
where L ∈ R, then {an }∞
n=1 is convergent and n→∞
lim an = L.
Proof : By hypothesis we have
L = lim sup an = lim l.u.b {an , an+1 , an+2 , · · · } .
n−→∞ n→∞

32
Thus given ε > 0 there exists N1 ∈ N such that

|l.u.b {an , an+1 , an+2 , · · · } − L| < ε for all n ≥ N1

implying that
an < L + ε for all n ≥ N1 (1)
Similarly, since lim inf an = L, there exists N2 ∈ N such that
n−→∞

|g.l.b {an , an+1 , an+2 , · · · } − L| < ε for all n ≥ N2

implying that
an > L − ε for all n ≥ N2 (2)
If N = max {N1 , N2 }, then from (1) and (2) we have

|an − L| < ε for all n ≥ N.

This proves that lim an = L. @


n→∞

Theorem 4.7.5: Let {an }∞


n=1 is a sequence of real numbers. If lim sup an = ∞ = lim inf an ,
n−→∞ n−→∞
then {an }∞
n=1 diverges to ∞.
Proof : Since lim inf an = ∞, given M > 0 there exists N ∈ N such that
n−→∞

mn = g.l.b {an , an+1 , an+2 , · · · } > M for all n ≥ N.

This implies that M is a lower bound (but not the g.l.b) for {an , an+1 , · · · } for all n ≥ N, so
that
an > M for all n ≥ N,
which establishes the required conclusion. @
Corollary 4.7.1: If {an }∞
n=1 is a sequence of real numbers diverging to ∞, then

lim sup an = ∞ = lim inf an .


n−→∞ n−→∞

Theorem 4.7.6: Let {an }∞ ∞


n=1 and {bn }n=1 be bounded sequences of real numbers. If an ≤ bn
for all n ∈ N, then lim sup an ≤ lim sup bn and lim inf an ≤ lim inf bn .
n−→∞ n−→∞ n−→∞ n−→∞
Proof : From the hypothesis an ≤ bn for all n ∈ N it is clear that

l.u.b {an , an+1 , an+2 , · · · } ≤ l.u.b {bn , bn+1 , bn+2 , · · · } ,

and
g.l.b {an , an+1 , an+2 , · · · } ≤ g.l.b {bn , bn+1 , bn+2 , · · · } .
Taking the limit as n −→ ∞ on both sides of these inequalities, and using Corollary 4.5.1,
we prove the theorem. @
Remark 4.7.2 :

33
(i ) Theorems 4.7.6 remains true even if an > bn for a finite number of n.
(ii ) It is not always true that lim sup (an + bn ) = lim sup an + lim sup bn , even for bounded
n−→∞ n−→∞ n−→∞
sequences {an }∞ ∞
n=1 and {bn }n=1 .
n ∞ n+1 ∞
Example: Let {an }∞ ∞ 
n=1 = {(−1) }n=1 and {bn }n=1 = (−1) n=1
so that an + bn = 0 for
each n ∈ N. In this case, lim sup an = 1 = lim sup bn so that lim sup an + lim sup bn = 2.
n−→∞ n−→∞ n−→∞ n−→∞
However, lim sup (an + bn ) = 0.
n−→∞

Theorem 4.7.7: Let {an }∞ ∞


n=1 and {bn }n=1 be bounded sequences of real numbers. Then
(a) lim sup (an + bn ) ≤ lim sup an + lim sup bn
n−→∞ n−→∞ n−→∞

(b) lim inf (an + bn ) ≥ lim inf an + lim inf bn


n−→∞ n−→∞ n−→∞
Proof : We prove part (a) of the theorem; the proof of part (b) is much the same. Let

Mn = l.u.b {an , an+1 , an+2 , · · · } ,

and
Pn = l.u.b {bn , bn+1 , bn+2 , · · · } .
Then
ak ≤ Mn for all k ≥ n,
and
bk ≤ Pn for all k ≥ n,
so that
ak + bk ≤ Mn + Pn for all k ≥ n.
Thus Mn + Pn is an upper bound for {an + bn , an+1 + bn+1 , an+2 + bn+2 , · · · } so that

l.u.b {an + bn , an+1 + bn+1 , an+2 + bn+2 , · · · } ≤ Mn + Pn .

By Corollary 4.5.1 and Theorem 4.5.1,

lim l.u.b {an + bn , an+1 + bn+1 , an+2 + bn+2 , · · · } ≤ lim (Mn + Pn ) = lim Mn + lim Pn
n→∞ n→∞ n→∞ n→∞

or
lim sup (an + bn ) ≤ lim sup an + lim sup bn . @
n−→∞ n−→∞ n−→∞

Theorem 4.7.8: Any bounded sequence of real numbers has a converent subsequence.
Exercise:
1. Let {an }∞ ∞
n=1 be a sequence of real numbers and let {ani }n=1 be any subsequence of
{an }∞
n=1 . Prove that if lim sup an = M, then lim sup ani ≤ M.
n−→∞ n−→∞
2. Let {an }∞
n=1 be a bounded sequence of real numbers and let lim inf an = m. Prove that:
n−→∞
(a) there is a subsequence of {an }∞
n=1 which converges to m.

(b) no subsequence of {an }n=1 converges to a limit less than m.

34
4.8 Cauchy Sequences
Definition 4.8.1: Let {an }∞ ∞
n=1 be a sequence of real numbers. Then {an }n=1 called a Cauchy
sequence if for any ε > 0 there exists N = N (ε) ∈ N such that

|am − an | < ε for all m, n ≥ N.

In other words, {an }∞


n=1 is Cauchy if am and an are close together when m and n are very
large.
Remark 4.8.1 : The Cauchy criterion is the most important criterion for proving that a
sequence converges without knowing its limit.
Theorem 4.8.1: Let {an }∞ ∞
n=1 be a convergent sequence of real numbers. Then {an }n=1 is
Cauchy.
Proof : Let L = lim an . Then, given ε > 0 there exists N ∈ N such that
n→∞

ε
|ak − L| < for all k ≥ N.
2
Thus if m, n ≥ N, we have
ε ε
|am − an | = |(am − L) + (L − an )| ≤ |(am − L)| + |(L − an )| < +
2 2
so that
|am − an | < ε for all m, n ≥ N,
which proves that {an }∞
n=1 is Cauchy. @

Lemma 4.8.1: If {an }∞ ∞


n=1 is a Cauchy sequence of real numbers, then {an }n=1 is bounded.
Proof : Given ε = 1, choose N ∈ N such that

|am − an | < 1 for all m, n ≥ N.

Then
|am − aN | < 1 for all m ≥ N. (1)
Hence, if m ≥ N, we have

|am | = |(am − aN ) + aN | ≤ |am − aN | + |aN |


and so, using (1)
|am | ≤ 1 + |aN | for all m ≥ N.
If M = max {|a1 | , · · · , |aN −1 |}, then

|am | < M + 1 + |aN | for all m ∈ N.

Hence {an }∞
n=1 is bounded.

35
Theorem 4.8.2: If {an }∞ ∞
n=1 is a Cauchy sequence of real numbers, then {an }n=1 is conver-
gent.

Proof : By Theorem 4.4.4, {an }∞

n=1 has a monotonic subsequence anj j=1 . By Lemma 4.8.1,
∞ ∞
{an }∞
 
n=1 is bounded. Hence anj j=1 is bounded. Thus anj j=1 converges to some a ∈ R (by

Remark 4.4.3). We will show that {an }∞

n=1 itself converges to a. Fix ε > 0. Since anj j=1
converges to a, there exists J ∈ N such that
ε
anj − a < for all j ≥ J. (1)
2
Since {an }∞
n=1 is Cauchy, there exists N ∈ N such that

ε
|am − an | < for all m, n ≥ N. (2)
2
We may choose N such that N ≥ J.
Now suppose k ∈ N and k ≥ N. Then k ≥ J, so (1) implies
ε
|ank − a| < .
2
Also, nk ≥ k ≥ N, so (2) implies
ε
|ak − ank | < .
2
Therefore,
ε ε
|ak − a| = |(ak − ank ) + (ank − a)| ≤ |ak − ank | + |ank − a| < +
2 2
so that
|ak − a| < ε for all k ≥ N,
and the theorem follows. @

4.9 Infinite Series


4.9.1 Convergence of Infinite Series
Definition 4.9.1: Let {an }∞ n=1 be a sequence of real numbers. An expression of the form
P ∞
P
a1 + a2 + · · · + an + · · · , denoted by an or an , is called an infinite series.
n=1

P
Definition 4.9.2: Let an be an infinite series of real numbers and let sn = a1 +a2 +· · ·+an
n=1
be the sum of the first n terms (nth partial sum) of the sequence {an }∞ ∞
n=1 . Then {sn }n=1 is

P
called the sequence of partial sums of the series an .
n=1

an be an infinite series and let {sn }∞
P
Definition 4.9.3: Let n=1 be the corresponding
n=1
∞ ∞
an is said to converge if {sn }∞
P P
sequence of partial sums. Then n=1 converges, an diverges
n=1 n=1

36

if {sn }∞ an oscillates if {sn }∞ ∞
P
n=1 diverges, and n=1 oscillates. If {sn }n=1 converges to s, then
n=1

P ∞
P
s is called the sum of the series an and we write s = an or s = a1 + a2 + · · · + an + · · · .
n=1 n=1

P n
P
In this case, an = s = lim sn = lim ai . If the series diverges to ∞, we write
n=1 n→∞ n→∞ i=1

P ∞
P
an = ∞, and if it diverges to −∞, we write an = −∞.
n=1 n=1

P ∞
P
Theorem 4.9.1: Let an and bn be convergent series.
n=1 n=1

P ∞
P ∞
P ∞
P
(a) Then (an ± bn ) is also convergent and (an ± bn ) = an ± bn .
n=1 n=1 n=1 n=1
∞ ∞ ∞
(b) If c ∈ R, then
P P P
can is also convergent and can = c an .
n=1 n=1 n=1
Proof :
(a)

X n
X
(an ± bn ) = lim (ai ± bi )
n→∞
n=1 i=1
n n
!
X X
= lim ai ± bi
n→∞
i=1 i=1
n
X n
X
= lim ai ± lim bi
n→∞ n→∞
i=1 i=1

X ∞
X
= an ± bn .
n=1 n=1

∞ n n ∞
(b) If c ∈ R, then
P P P P
can = lim cai = c lim ai = c an . @
n=1 n→∞ i=1 n→∞ i=1 n=1

P
Theorem 4.9.2: If an is a convergent series, then lim an = 0.
n=1 n→∞

P
Proof : Suppose an = s. Then lim sn = s where sn = a1 + a2 + · · · + an . But then
n=1 n→∞
lim sn−1 = s. Since an = sn − sn−1 , then, by Theorem 4.3.4,
n→∞

lim an = lim (sn − sn−1 ) = lim sn − lim sn−1 = s − s = 0. @


n→∞ n→∞ n→∞ n→∞

Remark 4.9.1 : Theorem 4.9.2 gives a necessary but not sufficient condition that a series
be convergent.

P
Corollary 4.9.1: If lim an 6= 0, then an diverges.
n→∞ n=1
Proof : The corollary is just the contrapositive of Theorem 4.9.2. @

37
Examples:
1−n
1. Suppose an = 1+2n . Then lim an = − 21 6= 0. Hence, by Corollary 4.9.1, the series
n→∞

1−n
P
1+2n
diverges.
n=1

(−1)n diverges since lim (−1)n does not exist.
P
2. The series
n=1 n→∞

n2 n2 1
P
3. The series 3n2 +n
diverges since lim 2 = 6= 0.
n=1 n→∞ 3n +n 3

Exercise:
Show that the following series are divergent.
(i ) 21 + 32 + 43 + · · ·
(ii ) − 12 + 23 − 43 + · · ·
P∞
Theorem 4.9.3: Let an be a series of non-negative numbers. Then
n=1

an converges if the sequence {sn }∞
P
(a) n=1 is bounded.
n=1

an diverges if the sequence {sn }∞
P
(b) n=1 is not bounded.
n=1
Proof :
(a) Since an+1 ≥ 0 we have sn+1 = a1 + · · · + an + an+1 = sn + an+1 ≥ sn . Thus {sn }∞
n=1 is
non-decreasing and (by hypothesis) bounded. By Theorem 4.4.1, {sn }∞ n=1 converges,

P
and thus an converges.
n=1

{sn }∞ is not bounded, then by Theorem 4.3.1, {sn }∞
P
(b) If n=1 n=1 diverges. Hence an
n=1
diverges. @
Example:

1
P
The harmonic series n
is divergent.
n=1
Proof : In this case sn = 1 + 12 + · · · + n1 . Consider the subsequence s1 , s2 , s4 , s8 , · · · , s2n−1 , · · ·
of {sn }∞
n=1 . We have
s1 = 1,
1
s2 = 1 + 2
= 32 ,
1 1 3 1 1
s4 = s2 + 3
+ 4
> 2
+ 4
+ 4
= 2,
1 1 1 1
s8 = s4 + 5
+ + +
6 7 8
> 2 + 81 + 81 + 1
8
+ 1
8
= 52 ,
..
.
In general, it may be shown by induction that s2n ≥ n+2
2
. Thus {sn }∞
n=1 contains a divergent

subsequence and hence, by Corollary 4.2.1, {sn }∞ 1
P
n=1 diverges. Therefore, the series n
n=1
diverges. @

38

P
Remark 4.9.2 : The divergence of the harmonic series shows that a series an may diverge
n=1
even if lim an = 0.
n→∞

4.9.2 Alternating Series


Definition 4.9.4: A series of the form a1 − a2 + a3 − a4 + · · · + (−1)n+1 an + · · · where an > 0
(or < 0) for all n is called an alternating series.
Examples:
(a) −1 + 2 − 3 + 4 − 5 + · · ·
(b) 1 − 12 + 13 − 14 + · · ·
(c) 1 − 21 + 1
4
− 18 + · · ·
are all alternating series.
Theorem 4.9.4 (Leibniz Test): If {an }∞ n=1 is a non-increasing sequence of positive numbers

(−1)n+1 an is convergent.
P
such that lim an = 0, then the alternating series
n→∞ n=1

Examples:

P 1
1 ∞
1. From a previous example n
is diverges. However, since n n=1
is non-increasing
n=1

(−1)n+1
= 1 − 12 + 31 − 14 + · · ·
P
and lim an = 0, it follows from Theorem 4.9.4 that n
n→∞ n=1
converges.
 1 ∞
2. The sequence {an }∞
 1 1 1
n=1 = 2n−1 n=1 = 1, 2 , 4 , 8 , · · · is non-increasing and n→∞ lim an = 0.

(−1)n+1 2n−1 1
= 1 − 21 + 41 − 18 + · · · converges.
P
Hence, by Theorem 4.9.4, the series
n=1

4.9.3 Absolute Convergence



P
Definition 4.9.5: Let an be a series of real numbers.
n=1

P ∞
P
(a) If |an | converges, we say that an converges absolutely.
n=1 n=1

P ∞
P ∞
P
(b) If an converges but |an | diverges, we say that an converges conditionally.
n=1 n=1 n=1

Examples:

an = 1 − 21 + 14 − 18 + · · · converges. Also, it is
P
1. From Example 2 above, the series
n=1
∞ ∞
|an | = 1 + 21 + 1
+ 81 + · · · converges. Hence
P P
easy to show that the series 4
an
n=1 n=1
converges absolutely.

an = 1 − 21 + 13 − 14 + · · · converges. However,
P
2. From Example 1 above, the series
n=1

39
∞ ∞
|an | = 1 + 12 + 31 + 41 + · · · diverges. Hence
P P
from a previous example, the series an
n=1 n=1
converges conditionally.

P ∞
P
Theorem 4.9.5: If an converges absolutely, then an converges.
n=1 n=1
Proof : Let sn = a1 + · · · + an . We wish to prove that {sn }∞
n=1 converges. By Theorem

∞ P
4.8.2, it is enough to show that {sn }n=1 is Cauchy. By hypothesis |an | < ∞. Thus if
n=1
tn = |a1 | + · · · + |an |, then {tn }∞ ∞
n=1 converges. By Theorem 4.8.1, {tn }n=1 is Cauchy. Thus
given ε > 0 there exists N ∈ N such that

|tm − tn | < ε for all m, n ≥ N.

But (if m > n, say), |sm − sn | = |an+1 + · · · + am | ≤ |an+1 | + · · · + |am | = tm − tn . Thus

|sm − sn | < ε for all m, n ≥ N.

Hence {sn }∞
n=1 is Cauchy, which is what we wished to show. @

4.9.4 Tests for Convergence


In Subsubsection 4.9.1, we defined the convergence of a series in terms of the convergence of
its associated sequence of partial sums {sn }∞
n=1 . However, it is not always easy to find a simple
formula for the nth partial sums. Consequently, the technique of establishing convergence by
directly finding the limit of {sn }∞
n=1 does not always work. However, the situation is better
for a series whose terms are positive.

P
Remark 4.9.3 : If an is a series of positive terms, then its sequence of partial sums
n=1
{sn }∞
n=1 is monotonic increasing, for sn+1 − sn = an+1 > 0 for all n ∈ N.

P P∞
Theorem 4.9.6: If an is a series of positive terms, then an either converges to a
n=1 n=1
positive number or diverges to ∞.

an and its sequence of partial sums {sn }∞
P
Proof : If the series n=1 is bounded, there exists
n=1
k > 0 such that k > sn+1 > sn ≥ a1 for all n ∈ N. Thus {sn }∞
n=1 being monotonic increasing
and bounded is convergent, by Theorem 4.4.1, and k ≥ lim sn ≥ a1 > 0.
n→∞
If {sn }∞
n=1 is unbounded, then {sn }∞
n=1 being monotonic increasing, by Remark 4.9.3, diverges
P∞
to ∞, by Theorem 4.4.2, i.e., an diverges to ∞. @
n=1

P ∞
P
Theorem 4.9.7 (Comparison Test): Let an and bn be series of positive terms
n=1 n=1
and suppose that an ≤ bn for all n ∈ N.

P ∞
P
(a) If bn converges, then an converges.
n=1 n=1
P∞ ∞
P
(b) If an diverges, then bn diverges.
n=1 n=1

40
Proof : Let sn = a1 + · · · + an and tn = b1 + · · · + bn be the corresponding nth partial sums.
Since an ≤ bn for all n ∈ N, then sn ≤ tn for all n ∈ N.

bn converges, then {tn }∞ ∞
P
(a) If n=1 is bounded. Thus {sn }n=1 is also bounded. Since
n=1

{sn }∞
P
n=1 is increasing and bounded, it is convergent (by Theorem 4.4.1). Hence an
n=1
converges.

an diverges, then {sn }∞ ∞
P
(b) If n=1 is necessarily unbounded and thus {tn }n=1 is also
n=1

unbounded and therefore {tn }∞
P
n=1 is divergent. Hence bn diverges. @
n=1
Examples:
∞ ∞
1 1
P P
1. Consider the series np
where p ≥ 2. It can be shown that n2
is convergent.
n=1 n=1

1 1 1
P
Now, if p ≥ 2, then np
≤ n2
. Therefore, by comparison test, np
is also convergent.
n=1

1
P
The series np
is called a p-series and actually converges for any p > 1.
n=1

n+7
P
2. Determine whether the series n3 +1
converges or diverges.
n=1
Solution: We have that
n+7 n + 7n
3
≤ 3
n +1 n +1
8n
< 3
n
8
= 2.
n
∞ ∞
8 n+7
P P
The series n2
converges and thus, by comparison test, the series n3 +1
also
n=1 n=1
converges.

P ∞
P
Theorem 4.9.8 (Limit Comparison Test): Let an and bn be series of positive
n=1 n=1
an
terms and let lim = L.
n→∞ bn

P ∞
P ∞
P
(a) If L > 0, then an converges if and only if an converges (in other words, an
n=1 n=1 n=1

P
and an converge or diverge together).
n=1

P ∞
P
(b) If L = 0 and bn converges, then an converges also.
n=1 n=1
P∞ ∞
P
(c) If L = ∞ and bn diverges, then an diverges also.
n=1 n=1
Examples:
∞ ∞ ∞ ∞ 1
1 1 an n
P P P P n+1
1. Let an = n+1
and bn = n
. Then L = lim = lim 1 = lim = 1.
n=1 n=1 n=1 n=1 n→∞ bn n→∞ n n→∞ n+1
Since 0 < L < ∞, the two series converge or diverge together. But the harmonic series

41
∞ ∞
1 1
P P
n
diverges. Hence the series n+1
diverges also.
n=1 n=1
∞ ∞ ∞ ∞ 1
1 1 an n2 +1 n2
P P P P
2. If an = n2 +1
and bn = n2
, then L = lim = lim 1 = lim 2 +1 = 1.
n=1 n=1 n=1 n=1 n→∞ bn n→∞ n2 n→∞ n
∞ ∞
1 1
P P
Since 0 < L < ∞, and n2
is a convergent p-series, then n2 +1
converges also.
n=1 n=1

P an+1
Theorem 4.9.9 (Ratio Test): Let an be a series of positive terms and let lim = l,
n=1 n→∞ an
finite or infinite. ∞
P
(a) If l < 1, then an converges.
n=1
P∞
(b) If l > 1, then an diverges.
n=1
(c) If l = 1, the series may converge or diverge and so, the test fails.

an
P
Corollary 4.9.2: Let an be series of positive terms and let lim an+1 = l, finite or infinite.
n=1 n→∞

P
(a) If l > 1, then an converges.
n=1
P∞
(b) If l < 1, then an diverges.
n=1
(c) If l = 1, the series may converge or diverge and so, the test fails.
Examples:

2n
P
1. The series n!
converges since by the ratio test
n=1

an+1 2n+1 n! 2
lim = lim n
= lim = 0 < 1.
n→∞ an n→∞ (n + 1)! 2 n→∞ n + 1

12 ·22 22 ·32 32 ·42


2. The series 1!
+ 2!
+ 3!
+ · · · is convergent.
2 2
Proof : In this case, an = n ·(n+1)
n!
so that

an+1 (n + 1)2 · (n + 2)2 n!


lim = lim ·
n→∞ an n→∞ (n + 1)! n2 · (n + 1)2
(n + 2)2
= lim 2
n→∞ n (n + 1)
 
2 1
= lim 1 + ·
n→∞ n (n + 1)
= 0 < 1.

Hence by the ratio test, the series converges. @



n!xn
P
3. Prove that for x > 0, the series nn
converges if x < e and diverges if x > e.
n=1

42
n!xn
Proof : Since x > 0, the series is of positive terms and an = nn
for all n ∈ N. Thus

an+1 (n + 1)!xn+1 nn
= ·
an (n + 1)n+1 n!xn
(n + 1) n!xn x nn
= ·
(n + 1)n (n + 1) n!xn
 n
n
= x
n+1
x
= n
1 + n1
an+1 x x x
and lim = lim 1 n = e . Hence, by the ratio test, the series converges if <1
n→∞ an n→∞ (1+ n ) e

or x < e, and diverges if xe > 1 or x > e.


Exercise:
3 2 4 3 5 4 n+1 n
Use the ratio test to show that the series 21 +
   
3
+ 5
+ 7
+· · ·+ 2n−1
+· · · converges.

P √
Theorem 4.9.10 (Root Test): Let an be series of positive terms and let lim n an = l,
n=1 n→∞
finite or infinite. ∞
P
(a) If l < 1, then an converges.
n=1
P∞
(b) If l > 1, then an diverges.
n=1
(c) If l = 1, the series may converge or diverge and so, the test fails.
Examples:
∞ n
2−n+(−1) is convergent.
P
1. Show that the series
n=1
√ (−1)n
1
Solution: In this case, lim n an = lim 2−1+ n = 2
< 1. Therefore, by the root test,
n→∞ n→∞
the series is convergent.
∞ 2
nn
P
2. Show that the series n2
is convergent.
n=1 (n+1)
n2
Solution: In this case, an = n n2 > 0 for all n ∈ N. By the root test, the series
(n+1)
√ nn 1 1
converges since lim n an = lim (n+1) n = lim 1 n = e < 1.
n→∞ n→∞ n→∞ (1+ n )

1
P
3. The series n pn
converges if p > 0 and diverges if p ≤ 0.
n=1

Proof : In this case, an = np1n > 0 for all n ∈ N and for all p ∈ R, lim n an = lim n1p .
√ n→∞ n→∞
Now, if p > 0, lim n an = 0 < 1 and the series converges. Since an = 1 for all n ∈ N
n→∞
∞ √
1
P
if p = 0, n pn
diverges. If p < 0, lim n an = ∞ > 1, and by the root test, the series
n=1 n→∞
diverges. @
2 3 2 4 3 5 4
  
4. Test for convergence of the series 1
+ 3
+ 5
+ 7
+ ···.

43
n+1
n √ n+1 1
Solution: In this case, an = 2n−1 so that lim n an = lim = < 1. Hence, by
n→∞ n→∞ 2n−1 2
the root test, the series converges.
Exercise: ∞ 2
3n
P
1. Test for convergence of the series 43n
.
n=1
∞ ∞
e2n en
P P
2. Show that 3n
diverges while 3n
converges.
n=1 n=1
∞  2
1 −n
xn converges if x < e and diverges if x > e.
P
3. Show that for x > 0, 1+ n
n=1

Theorem 4.9.11 (Integral Test): Let f be a continuous, decreasing positive-valued func-



P
tion defined on the interval [1, ∞) and consider the series an with an = f (n).
n=1
R∞ ∞
P
(a) If the improper integral 1
f (x) dx < ∞, then the series an converges.
n=1
R∞ P∞
(b) If the improper integral 1
f (x) dx = ∞, then the series an diverges.
n=1
Examples:

1
P
1. The series np
converges if p > 1 and diverges if p ≤ 1.
n=1
1
Proof : Let f (x) = xp
.Suppose p > 1. Then
Z ∞ Z t
dx
p
= lim x−p dx
1 x t→∞ 1
 1−p t
x
= lim
t→∞ 1 − p
 1−p 1 
t 1
= lim −
t→∞ 1 − p 1−p
 
1 1
= lim p−1
+
t→∞ (1 − p) t p−1
1
= < ∞.
p−1

1
P
Hence, np
converges if p > 1.
n=1
Now, suppose p < 1. Then
Z ∞ t
dx
Z
= lim x−p dx
1 xp t→∞ 1
 1−p t
x
= lim
t→∞ 1 − p
 1−p 1 
t 1
= lim −
t→∞ 1 − p 1−p
= ∞.

44

1
P
Hence, np
diverges if p < 1.
n=1
Finally, if p = 1,
∞ Z t
dx dx
Z
= lim
1 x t→∞ 1 x

= lim [ln x]t1


t→∞
= lim [ln t − ln 1]
t→∞
= ∞.

1
P
Hence, np
diverges if p = 1. @
n=1

1
P
2. Determine the convergence of the series n2 +1
.
n=1
1
Solution: Let f (x) = x+1
. Then f satisfies the hypothesis of the integral theorem.
We now have that
∞ t
dx dx
Z Z
= lim dx
1 x + 1 t→∞ 1 x+1
t
= lim tan−1 x 1

t→∞
= lim tan−1 t − tan−1 1
 
t→∞
= lim tan−1 t − tan−1 1

t→∞
π π π
= − = .
2 4 4
R ∞ dx
Since 1 x+1 exists, the series converges.
Exercise:
Use the integral test to determine whether the following series converge or diverge.

1
P
(a) n3
n=1

n
P
(b) n2 +1
n=1

1
P
(c) en
n=1

ln n
P
(d ) n
n=1

45
5 Limits and Continuity of Functions
5.1 Limits of Functions
Definition 5.1.1: Consider a function f : A −→ R and let c be an accumulation point of
A. We say that f has a limit at c or f converges at c, if there exists a number L ∈ R such
that for any given ε > 0 there exists δ > 0 such that if x ∈ A and 0 < |x − c| < δ then
|f (x) − L| < ε. In this case, we write limf (x) = L or f (x) −→ L as x −→ c and say that
x→c
f converges to L at c or that f has a limit L at c. If f does not converge at c, we say that f
diverges at c.
Remark 5.1.1 : By definition, if limf (x) = L, then for any ε > 0 there exists δ > 0 such
x→c
that for all x ∈ (c − δ, c + δ) ∩ A not equal to c, it holds that f (x) ∈ (L − ε, L + ε).
Theorem 5.1.1: A function f : A −→ R can have at most one limit at c.
Proof : Suppose that f (x) −→ L and f (x) −→ M as x −→ c, and let ε > 0. Then
there exists δ > 0 such that |f (x) − L| < 2ε and |f (x) − M| < 2ε for all x ∈ A satisfying
0 < |x − c| < δ. If 0 < |x − c| < δ, then
|L − M| ≤ |f (x) − L| + |f (x) − M|
ε ε
< +
2 2
= ε.
Since ε > 0 is arbitrary, then |L − M| < ε implies that L = M. @
Examples:
1. Define the function f : R −→ R by f (x) = 5x + 3. Prove that lim f (x) = 13.
x→2
Proof : We have that
|f (x) − 13| = |(5x + 3) − 13|
= |5x − 10|
= 5 |x − 2| .

Now, if 0 < |x − 2| < 5ε , then |f (x) − 13| < 5 5ε = ε. Thus, given ε > 0 we let δ = 5ε


so that if 0 < |x − 2| < δ, then |f (x) − 13| < ε. Hence, by definition, lim f (x) = 13. @
x→2
x+1
2. Let f : R −→ R where f (x) = x2 +3
. Prove that lim f (x) = 21 .
x→1
Proof : We have that
1 x+1 1
f (x) − = 2 −
2 x +3 2
x2 − 2x + 1
=
2 (x2 + 3)
|x − 1|2
=
2 (x2 + 3)
< |x − 1|2 .

46

Let ε > 0 be arbitrary and let δ = ε. Now, if 0 < |x − 1| < δ, then |x − 1|2 < δ 2 = ε.
Hence if 0 < |x − 1| < δ, then f (x) − 21 < ε. Therefore, lim f (x) = 12 . @
x→1

Theorem 5.1.2: Let f : A −→ R be a function and let c be an accumulation point of A.


Then limf (x) = L if and only if for every sequence {xn } in A converging to c (with xn 6= c
x→c
for all n ∈ N) the sequence {f (xn )} converges to L.
Corollary 5.1.1: Let f : A −→ R be a function. Let c be an accumulation 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 to c with xn 6= c for all n ∈ N, and such that {f (xn )} does not converge to L.
Corollary 5.1.2: Let f : A −→ R be a function and let c be an accumulation 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 lim f (xn ) 6= lim f (yn ), then f does not have a
x→∞ x→∞
limit.
Example:
Prove that lim x1 does not exist.
x→0 1 ∞
Proof : Consider {xn }∞ n=1 = n n=1
, which clearly converges to c = 0 and xn 6= 0 for all
n ∈ N. Then {f (xn )} = {n} which is unbounded and thus does not converge. Thus, by
Corollary 5.1.1, lim x1 does not exist. @
x→0

Exercise:
Use the ε-δ definition of the limit of a function to prove that:
(a) lim (2x − 5) = −7
x→−1
(b) lim 2x+3 = 3
x→3 4x−9
2 −3x
(c) lim xx+3 =2
x→6

Theorem 5.1.3: Let f, g : A −→ R be functions and let c be an accumulation point of A.


Suppose that limf (x) = L and limg (x) = M. Then:
x→c x→c
(a) lim (f ± g) (x) = L ± M
x→c
(b) lim (f g) (x) = LM
x→c
 
(b) lim fg (x) = M L
, if M 6= 0.
x→c

Corollary 5.1.3: Let f1 , · · · , fk : A −→ R be functions and let c be an accumulation point


of A. If lim fi (x) exists for each i = 1, 2, · · · , k, then
x→c
Pk k
P
(a) lim fi (x) = limfi (x)
x→c i=1 i=1 x→c
Qk Qk
(b) lim fi (x) = limfi (x)
x→c i=1 i=1 x→c

Examples:
1. If f (x) = a0 + a1 x + a2 x2 + · · · + an xn is a polynomial function, then lim f (x) = f (c)
x→c
for every c ∈ R. If g (x) = b0 + b1 x + b2 x2 + · · · + bm xm is another polynomial function

47
and g (x) 6= 0 in a neighbourhood of x = c and limg (x) = g (c) 6= 0, then lim fg(x)
(x)
= f (c)
g(c)
.
x→c x→c
2
2. Prove that lim xx−2
−4
=4
x→2
x2 −4
Proof : Observe that lim (x − 2) = 0. If x 6= 2, then x−2
= x + 2. Hence the functions
x→2
x2 −4
f (x) = x−2
and g (x) = x + 2 are equal at every point in R\ {2}. Since lim g (x) = 4,
x→2
it follows that lim f (x) = 4. @
x→2

Definition 5.1.2: Let f : A −→ R be a function. Then lim f (x) = L (i.e., f (x) −→ L as


x→∞
x −→ ∞) if given ε > 0, there exists M ∈ R such that |f (x) − L| < ε for all x > M.
Remark 5.1.2 : Definition 5.1.2 requires that the set A contains some interval of the form
(c, ∞).
Example:
Prove that lim x12 = 0.
x→∞
Proof : Given ε > 0, we need to find M ∈ R such that
1
− 0 < ε for all x > M. (1)
x2
Since (1) is equivalent to
1 √
< ε for all x > M,
x
it is clear that (1) will hold if we take M = √1ε . @
Definition 5.1.3: Let f : A −→ R be a function and let c be an accumulation point of A.
We say that f approaches L as x approaches c from the right, if given ε > 0, there exists
δ > 0 such that for all x ∈ A for which c < x < c + δ, then |f (x) − L| < ε. In this case, we
write lim+f (x) = L and the number L is called the right-hand limit of f at c. Similarly, we
x→c
say that f approaches M as x approaches c from the left, if given ε > 0, there exists δ > 0
such that for all x ∈ A for which c − δ < x < c, then |f (x) − M| < ε. In this case, we write
lim− f (x) = M and the number M is called the left-hand limit of f at c.
x→c

Remark 5.1.3 :
1. The statement lim+f (x) = L involves only values of f (x) for x to the right of c, while
x→c
lim− f (x) = M involves only values of f (x) for x to the left of c.
x→c

2. lim+f (x) and lim− f (x) may exist without being equal to each other.
x→c x→c

3. limf (x) = L if and only if lim+f (x) = lim− f (x) = L.


x→c x→c x→c

Example:(
x if 0 ≤ x < 1
If f (x) = , then lim− f (x) = 1 while lim+ f (x) = 2.
3 − x if 1 ≤ x ≤ 2 x→1 x→1

5.2 Continuous and Discontinuous Functions


Definition 5.2.1: Let f : A −→ R be a function. Then f is continuous at c ∈ A if:

48
(i ) limf (x) exists,
x→c

(ii ) f (c) is defined, and


(iii ) limf (x) = f (c).
x→c

In other words, f is continuous at c ∈ A if for each ε > 0, there is some δ > 0 such that if
x ∈ A and |x − c| < δ, then |f (x) − f (c)| < ε. If f is not continuous at c, then we say that
f is discontinuous at c. The function f is continuous on A if f is continuous at each x ∈ A.
Remark 5.2.1 : In Definition 5.2.1, the number δ, in general, depends on ε and the point c.
Definition 5.2.2: Consider a function f : A −→ R and let c ∈ A. Then f is right-continuous
at c if lim+f (x) = f (c), f is left-continuous at c if lim− f (x) = f (c).
x→c x→c

Remark 5.2.2 : A function f : A −→ R is continuous at c if and only if f is both right-


continuous and left-continuous at c.
Theorem 5.2.1: The function f : A −→ R is continuous at c ∈ A if and only if for every
sequence {xn } in A such that lim xn = c, we have lim f (xn ) = f lim xn = f (c).
n→∞ n→∞ n→∞

Theorem 5.2.2: The function f : A −→ R is discontinuous at c ∈ A if and only if there


exists a sequence {xn } in A converging to c but f (xn ) does not converges to f (c).
Theorem 5.2.3: Let f, g : A −→ R be continuous functions at c ∈ A and let b ∈ R. Then
f + g, f − g, f g, and bf are also continuous at c. Furthermore, if g (c) 6= 0, then fg is also
continuous at c.
Proof : Since f and g are continuous at c we have lim f (x) = f (c) and limg (x) = g (c). By
x→c x→c
Theorem 5.1.3, lim [f (x) + g (x)] = f (c)+g (c) or, equivalently, lim (f + g) (x) = (f + g) (c).
x→c x→c
Hence f + g is continuous at c. The remainder of the theorem is proved similarly. @
Corollary 5.2.1: Let f, g : A −→ R be continuous functions on A and let b ∈ R. Then
f + g, f − g, f g, and bf are also continuous on A. Furthermore, if g (c) 6= 0, then fg is also
continuous on A.
Examples:
1. Consider the function f : R −→ R defined by f (x) = 5x + 3. In a previous example,
it was proved that lim f (x) = 13. Furthermore, f (2) = 13. Therefore, f is continuous
x→2
at x = 2.
2. Let f (x) = sinx x where x ∈ R\ {0}. Then f is not continuous at x = 0 since it is not
defined at x = 0, even though lim sinx x exists (and is equal to 1). However, the function
( x→0
sin x
x
if x 6= 0
g defined by g (x) = is continuous at x = 0 since lim g (x) = g (0).
1 if x = 0 x→0

3. Prove that f (x) = x is continuous on R.


Proof : Suppose c ∈ R. Let ε > 0 be arbitrary. Let δ = ε. If 0 < |x − c| < δ, then
|f (x) − f (c)| = |x − c| < δ = ε. @

4. The function f (x) = x is discontinuous at all c < 0, right-continuous at c = 0, and

49
continuous at all c > 0. To prove that f is continuous at all c > 0, we have that
√ √
|f (x) − f (c)| = x − c
√ √
√ √ x+ c
= x− c · √ √
x+ c
|x − c|
=√ √
x+ c
1
≤ √ |x − c| .
c

Hence, given ε > 0, suppose 0 < |x − c| < cε. Then
1 1 √
|f (x) − f (c)| < √ |x − c| = √ cε = ε.
c c

5. If f (x) = a0 + a1 x + a2 x2 + · · · + an xn is a polynomial function, then limf (x) = f (c)


x→c
for every c ∈ R. Thus f is continuous on R. If g (x) = b0 + b1 x + b2 x2 + · · · + bm xm
is another polynomial function and g (c) 6= 0, then lim fg(x)
(x)
= fg(c)
(c)
. Hence, h (x) = fg(x)
(x)
x→c
is continuous at every c where g is non-zero.
(
1
if x 6= 0
6. Determine the points of continuity of the function f (x) = x
0 if x = 0.
1
Solution: Suppose that c 6= 0. Then limf (x) = c = f (c). Hence, f is continuous at
x→c
c ∈ R\ {0}. Now, let c = 0. The sequence xn = n1 converges to c = 0 but f (xn ) = n
does not converge. Hence lim f (x) does not exist. Thus, even though f (0) = 0 is
x→0
well-defined, f is discontinuous at c = 0.
Exercise: 
1
 1−x
 if x < 0
2
1. Show that the function f (x) = x + 1 if 0 ≤ x ≤ 2 is continuous at x = 0 but

x + 4 if x > 2

discontinuous at x = 2. (
1 if x ∈ Q
2. Show that the function f (x) = is discontinuous everywhere.
0 if x ∈ R\Q

5.3 Uniform Continuity of Functions


Definition 5.3.1: A function f : A −→ R is said to be uniformly continuous on A if
for each ε > 0, there exists δ > 0 such that for all x, y ∈ A such that |x − y| < δ, then
|f (x) − f (y)| < ε.
Remark 5.3.1 : In the case of uniform continuity of a function f : A −→ R, the number
δ depends only on ε and not on the points of A. This is unlike in the case of pointwise
continuity, where δ, in general, depends on ε and the points of A. Thus uniform continuity
is a global property whereas the pointwise continuity is a local property.

50
Examples:
1. Let k be any non-zero constant. Show that f (x) = kx is uniformly continuous on R.
Solution: We have that |f (x) − f (y)| = |kx − ky| = |k| |x − y|. Hence, for any ε > 0,
ε
we let δ = |k| , and thus if x, y ∈ R such that |x − y| < δ, then |f (x) − f (y)| < ε.
2. The function f (x) = x1 is uniformly continuous on the set [1, 3].
Proof : Let x, y ∈ [1, 3]. Then 13 ≤ x1 ≤ 1 and 31 ≤ y1 ≤ 1. Now, we have that

1 1
|f (x) − f (y)| = −
x y
y−x
=
xy
1 1
= |x − y|
|x| |y|
≤ |x − y| .

Given ε > 0, we can set δ = ε so that if |x − y| < δ, then |f (x) − f (y)| < ε. @
Exercise:
Show that each of the following functions is uniformly continuous on R:
(a) f (x) = sin x
(b) f (x) = sin 2x
1
(c) f (x) = 1+x 2

Theorem 5.3.1: Let f : A −→ R be a function. The following are equivalent:


(a) The function is not uniformly continuous on A.
(b) There exists ε > 0 such that for every δ > 0, there exists x, y ∈ A such that |x − y| < δ
but |f (x) − f (y)| ≥ ε.
(c) There exists ε > 0 and a pair of sequences {xn } and {yn } in A where lim (xn − yn ) = 0
x→∞
and |f (xn ) − f (yn )| ≥ ε.
Example:
1. The function f (x) = x1 is not uniformly continuous on A = (0, 1]. 
Proof : Let ε = 1. Consider the sequences {xn } = n1 and {yn } = 2n 1
in A. Then
 
1 1 1
lim (xn − yn ) = lim − = lim =0
x→∞ x→∞ n 2n x→∞ 2n

but
|f (xn ) − f (yn )| = |n − 2n| = n ≥ 1 = ε.
Hence, by Theorem 5.3.1, the conclusion follows. @
1
Remark 5.3.2 : In general, the function f (x) = x
is not uniformly continuous on the set
A = (0, ∞).
2. The function f (x) = x2 is not uniformly continuous on [1, ∞), since the function is
1

unbounded as x −→ ∞. We accordingly can set {xn } = {n} and {yn } = n + n so

51
that  2
1 1
|f (xn ) − f (yn )| = n+ − n2 = 2 + > 2.
n n2
Exercise:
2
Show that the function f (x) = x
is not uniformly continuous on A = (0, 2].
Theorem 5.3.2: If the function f : A −→ R is uniformly continuous on A, then it is
continuous on A.
Proof : Suppose f is uniformly continuous on A. Then for ε > 0, there exists δ > 0 such
that for all x1 , x2 ∈ A such that |x1 − x2 | < δ implies |f (x1 ) − f (x2 )| < ε. Now, let x ∈ A.
Then on taking x1 = x, we find for ε > 0, there exists δ > 0 such that for fixed x2 ∈ A such
that |x − x2 | < δ implies |f (x) − f (x2 )| < ε. Hence f is continuous at every x2 ∈ A, i.e., f
is continuous on A. @
Theorem 5.3.3: If the function f : A −→ R is continuous on the closed interval A = [a, b],
then it is uniformly continuous on A.
Proof : Suppose that f is not uniformly continuous on A. Then by Theorem 5.3.1, there
exists ε > 0 and two sequences {xn } and {yn } in A such that lim (xn − yn ) = 0 and
x→∞
|f (xn ) − f (yn )| ≥ ε. Since {xn } is a bounded sequence, by Theorem 4.3.2, there is a sub-
sequence {xnk } of {xn } that converges to a point x0 ∈ [a, b]. But then the corresponding
subsequence {ynk } of {yn } also converges to x0 since
ε ε
|ynk − x0 | ≤ |ynk − xnk | + |xnk − x0 | < + = ε.
2 2
But since the sequences {xnk } and {ynk } both converge to x0 , and f is continuous at
x0 , it follows that the sequences {f (xnk )} and {f (ynk )} must both converge to f (x0 ) or
lim |f (xnk ) − f (ynk )| = 0 which contradicts the assumption that |f (xnk ) − f (ynk )| ≥ ε for
x→∞
all k. @
Example:
Consider the function f (x) = x1 . By Theorem 5.3.3, f is uniformly continuous on any closed
interval that does not contain 0. By Definition 5.3.1, f is uniformly continuous on any set of
the form [c, ∞) for c > 0, because if x, y ∈ [c, ∞) and |x − y| < δ = c2 ε, then
1 1 y−x
− =
x y xy
1 1
= |x − y|
|x| |y|
11 2
< cε
cc
= ε.
Exercise:
1. Let f : A −→ R and g : A −→ R be uniformly continuous functions on A. Prove that
f + g is uniformly continuous on A.
2. Prove that f (x) = x2 is not uniformly continuous on R.
3. Show that f (x) = sin x2 is not uniformly continuous on R.

52

You might also like