You are on page 1of 116

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/305335787

Elements of Real Analysis

Book · July 2014

CITATIONS READS

0 6,489

1 author:

Akhilesh Chandra Yadav


Mahatma Gandhi Kashi Vidyapith, Cant, Varanasi, India
25 PUBLICATIONS 45 CITATIONS

SEE PROFILE

All content following this page was uploaded by Akhilesh Chandra Yadav on 15 July 2016.

The user has requested enhancement of the downloaded file.


Chapter 1

Sets

‘Set’, ‘belongs to’ and ‘equal to’ are primitive terms of which
the reader has intuitive understanding. Their use is governed by
some postulates in axiomatic set theory. Here in this text, it is not
our purpose to give an account of axiomatic set theory, rather we
develop it in semi axiomatic manner . For the formal development
the reader is referred to S et Theory and Continium hypothesis by
P.J. Cohen.
To take the help of intuition in ascertaining the use of the
primitive terms we regard set as a collection of well defined
objects. Thus ‘A class of students’ , ‘a flock of sheep’ and ‘a
packet of biscuits’ are all examples of sets of things. The notation
‘a ∈ A’ stands for the statement ‘a belongs to A’ (‘a is an element
of A’ , or also ‘a is a member of A’). The negation of the statement
‘a ∈ A’ is denoted by ‘a 6∈ A’. The notation ‘A = B’ stands for the
statement ‘A is equal to B. The negation of ‘A = B’ is denoted
by ‘A 6= B’. The following axiom relates ‘∈’ and ‘=’.

Axiom 1.0.1. (Axiom of extension) A = B ⇐⇒ ∀x(x ∈


A ⇐⇒ x ∈ B).

Thus two sets A and B are equal if they have same members.

1
2

Two equal sets are treated as same. If A = B, then we may


substitute A for B and B for A in any discussion.
Remark 1.0.2. To be logically sound in the use of primitive
terms, axiom of extension is a necessity.
Let A and B be any two sets . We say that A is a subset of
B (A is contained in B or B contains A) if every member of A
is a member of B. Thus, the statement ‘A is a subset of B’ is
same as ‘∀x(x ∈ A =⇒ x ∈ B)’. The notation ‘A ⊆ B’ (or also
’B ⊇ A’) stands for the statement ’A is a subset of B’. Thus
‘A = B’ (axiom of extension) if and only if A ⊆ B and B ⊆ A.
The negation of ‘A ⊆ B’ denoted by ‘A 6⊆ B’ stands for the
negation of ‘∀x(x ∈ A =⇒ x ∈ B)’ which is logically same as
‘∃x(x ∈ A and x 6∈ B)’. Thus to say that A is not a subset of B
is to say that there is an element of A which is not in B.
Observe that every set is a subset of itself for ‘∀x(x ∈ A =⇒
x ∈ A)’ is a tautology. If A ⊆ B and A 6= B, we say that A is
a proper subset of B. The notation for the statement ‘A is a
proper subset of B’ is ‘A ⊂ B’. Thus, A is a proper subset of B if
every member of A is a member of B and there is a member of B
which is not a member of A. More precisely ‘A ⊂ B’ represents ‘
(∀x(x ∈ A =⇒ x ∈ B)) and (∃x(x ∈ B and x 6∈ A))’.
Proposition 1.0.3. ”(A ⊆ B and B ⊆ C) =⇒ (A ⊆ C)” is
always a true statement (a tautology).
Proof: Result follows by Law of Syllogism.✷

Axiom 1.0.4. (Axiom of specification) Let A be a set and


P (x) a valid statement involving the free symbol x. Then there is
a set B such that ‘x ∈ B ⇐⇒ (x ∈ A and P (x))’.
Thus to every set A and to every statement P (x) there is a
unique set B whose members are exactly those members of A for
which P (x) is true. This set B is denoted by B = {x ∈ A | P (x)}.
Clearly B is a subset of A.
3

Proposition 1.0.5. Let A be a set. Then there is a set B such


that B 6∈ A.

Proof: . Consider the statement ‘x is a set and x 6∈ x’.


There is a unique set B = {x ∈ A such that x is a set and x 6∈ x}
(axiom of specification). We show that B 6∈ A. Suppose that
B ∈ A. If B ∈ B, then B 6∈ B. Next if B 6∈ B then since B ∈ A
(supposition) and B is a set, B ∈ B. Thus B 6∈ B ⇐⇒ B ∈ B.
This is a contradiction (P ⇐⇒ −P is a contradiction) to the
supposition that B ∈ A. Hence B 6∈ A. ✷

Corollary 1.0.6. There is no set containing all sets.

Let A and B be sets. Consider the statement ‘x ∈ B’. The


set {x ∈ A | x ∈ B} is denoted by ‘A ∩ B’ and is called the
intersection of A and B. Thus

x ∈ A ∩ B ⇐⇒ (x ∈ A and x ∈ B).

Since [x ∈ A and x ∈ B] ⇐⇒ [x ∈ B and x ∈ A], we have:

Proposition 1.0.7. A ∩ B = B ∩ A.

Proposition 1.0.8. A ∩ B ⊆ A and A ∩ B ⊆ B.


T
Proof: By definition, x ∈ A B ⇐⇒ [x ∈ A and x ∈ B].
Further [x ∈ A and x ∈ B] =⇒ x ∈ A ((P and Q) =⇒ P is
a tautology). Thus x ∈ A ∩ B =⇒ x ∈ A. This shows that
A ∩ B ⊆ A. Similarly A ∩ B ⊆ B. ✷

Proposition 1.0.9. [C ⊆ A and C ⊆ B] =⇒ [C ⊆ A ∩ B].

Proof: Suppose that C ⊆ A and C ⊆ B. Let x ∈ C. Since


C ⊆ A and C ⊆ B, x ∈ A and x ∈ B. Thus x ∈ A ∩ B. Hence
x ∈ C =⇒ A ∩ B. This shows that C ⊆ A ∩ B. ✷
4

Proposition 1.0.10. [A ∩ B = A] ⇐⇒ [A ⊆ B].


Proof: Suppose that A ∩ B = A. Since A ∩ B ⊆ B, therefore
A ⊆ B. Conversely suppose that A ⊆ B. Since A ⊆ A, therefore
A ⊆ A ∩ B. Also A ∩ B ⊆ A. Hence A ∩ B = A. ✷

Proposition 1.0.11. (A ∩ B) ∩ C = A ∩ (B ∩ C).


Definition 1.0.12. (Complement of a set) Let A and B be
sets. Consider the statement x 6∈ B. By the axiom of specification
there is a unique set defined by {x ∈ A and x 6∈ B} = {x ∈ A |
x 6∈ B}. This set is denoted by A−B and is called the complement
of B in A (or A dif f erence B). Clearly A − B is a subset of A.
T
Proposition 1.0.13. A − B = A − (A B).
Axiom 1.0.14. (Axiom of existence of a set) There exists a
set.
Let A be a set. Consider A − A. If B is any set then (x ∈
A and x 6∈ A) ⇐⇒ (x ∈ B and x 6∈ B) (note that (P and −P )
⇐⇒ (Q and −Q) is a tautology). Thus x ∈ (A − A) ⇐⇒ x ∈
(B − B) and so A − A = B − B. The set A − A, therefore is
independent of A. This set is called the empty set or void set or
null set and is denoted by ∅. Thus ∅ = {x ∈ A | x 6∈ A}. Clearly
’x ∈ ∅’ is a contradiction.
Thus ‘x ∈ ∅ =⇒ Q’ is a tautology whatever the statement
Q may be.
Let P (x) be any contradiction involving the symbol x. Then
clearly ∅ = {x ∈ A | P (x)}. Intuitively one may think ∅ as a set
containing no element.
Proposition 1.0.15. The emptyset ∅ is a subset of every set.
Proof: Let B be a set. We have to show that x ∈ ∅ =⇒ x ∈
B. Since x ∈ ∅ is a contradiction, x ∈ ∅ =⇒ x ∈ B is a true
statement. Hence ∅ ⊆ B. ✷
5

Proposition 1.0.16. A − B = ∅ ⇐⇒ A ⊆ B.
Proof: Suppose that A−B = ∅. Let x ∈ A. Since A−B = ∅,
x 6∈ A − B (for x 6∈ ∅ is a tautology). Hence x ∈ B and so A ⊆ B.
Suppose that A ⊆ B. We have to show that A − B = ∅. As we
know that ∅ ⊆ A − B. Now, if x ∈ A − B then x ∈ A and x 6∈ B.
But A ⊆ B, hence x ∈ A and x 6∈ B =⇒ x ∈ B and x 6∈ B. This
in turn implies that x ∈ ∅, a contradiction. Hence A − B ⊆ ∅ and
so A − B = ∅. ✷
The following axiom permits us to generate more sets.
Axiom 1.0.17. (Pairing axiom) Let A and B be sets. There is
a set C such that A ∈ C and B ∈ C.
Consider the statement ‘x = A or x = B’. By the axiom
of specification, we have a unique set {x ∈ C | x = A or x = B}.
This set is also independent of the set C. It contains A and B
as elements and nothing else. We denote this set by {A, B}. The
set {A, A} is denoted by {A} and is called a singleton.
We have the empty-set ∅. Consider {∅}. Since ∅ ∈ {∅} but ∅ 6∈
∅, ∅ 6= {∅}. If {∅} = {{∅}}, then ∅ = {∅} a contradiction. Hence
{∅} 6= {{∅}}. Similarly {{{∅}}} 6= {{∅}}. Axiom of pairing gives
us other new sets such as {∅, {∅}} , {{∅, {∅}}, {{∅}}} and so on.
This way we produce several sets.
Axiom 1.0.18. (Union Axiom) Let A be set of sets. Then
there is a set U such that X ∈ A and x ∈ X implies that x ∈ U .
By the axiom of specification, we have a unique set {x ∈ U | x ∈
X f or some X ∈ A}. This set is denoted by S ∪X∈A X and is called
the U nion of the family A of sets. Thus x ∈ X∈A X ⇐⇒ x ∈ X
for some X ∈ A. What is ∪X∈∅ X? If x ∈ ∪X∈∅ , then there exists
X ∈ ∅ such that x ∈ X. But X ∈ ∅ is a contradiction. Hence
∪X∈∅ X = ∅. Clearly ∪X∈{A} X = A. The set ∪X∈{A,B} X is denoted
by A ∪ B. Thus x ∈ A ∪ B ⇐⇒ x ∈ A or x ∈ B. A ∪ B is called
the union of A and B.
6

Proposition 1.0.19. A ⊆ A ∪ B.
Proof: Suppose thatWx ∈ A. Then the statement ’x ∈ A or
x ∈ B’Sis true (P =⇒ P S Q is a tautology). Hence x ∈ A =⇒
x ∈ A B. Thus A ⊆ A B. ✷
S
Proposition 1.0.20. A ∅ = A.
Proof: Since x ∈ ∅ is always false, x ∈ A ⇐⇒ x ∈ A or x ∈ ∅.
Hence A ∪ ∅ = A. ✷

Proposition 1.0.21. A ∪ B = B ∪ A.
Proof: We have the following equivalences
x ∈ A ∪ B ⇐⇒ x ∈ A or x ∈ B
⇐⇒ x ∈ B or x ∈ A
[
⇐⇒ x ∈ B A
Thus A ∪ B = B ∪ A.✷
S
Proposition 1.0.22. A B = A ⇐⇒ B ⊆ A.
Proof: Suppose that A∪B = A. Let x ∈ B. Then x ∈ A∪B
and hence x ∈ A. Thus B ⊆ A. Conversely suppose that B ⊆ A.
Then x ∈ A or x ∈ B =⇒ x ∈ A. But already x ∈ A =⇒ x ∈
A or x ∈ B. Hence A = A ∪ B. ✷
S
Corollary 1.0.23. A A = A. ♯
S S S S
Proposition 1.0.24. (A B) C = A (B C).
Proof: We have the following equivalences
x ∈ (A ∪ B) ∪ C ⇐⇒ (x ∈ A or x ∈ B) or x ∈ C
⇐⇒ x ∈ A or (x ∈ B or x ∈ C)
⇐⇒ x ∈ A ∪ (B ∪ C)
Hence (A ∪ B) ∪ C = A ∪ (B ∪ C).✷
7

Proposition 1.0.25. Union distributes over intersection and in-


tersection distributes over union in the following sense:
S T S T S
1. A (B C) = (A B) (A C).
T S T S T
2. (A (B C) = (A B) (A C).
Proof: We have the following equivalences:

x ∈ A ∪ (B ∩ C) ⇐⇒ x ∈ A or x ∈ B ∩ C
⇐⇒ x ∈ A or (x ∈ B and x ∈ C)
⇐⇒ (x ∈ A or x ∈ B) and (x ∈ A or x ∈ C)
[ \ [
⇐⇒ x ∈ (A B) (A C)

This proves 1. Similarly one can prove 2. ✷

Theorem 1.0.26. (De Morgan Law) Let A, B and C be sets.


Then
S T
1. A − (B C) = (A − B) (A − C).
T S
2. A − (B C) = (A − B) (A − C).
Proof: We have the following equivalences:

[ [
A − (B C) ⇔ x ∈ A and x 6∈ B C
⇔ x ∈ A and (x 6∈ B and x 6∈ C)
⇔ (x ∈ A and x 6∈ B) and (x ∈ A and x 6∈ C)
\
⇔ x (A − B) (A − C)

This proves 1. Similarly we can prove 2. ✷

Axiom 1.0.27. (Power Set Axiom) Given a set A, there is a


set ℘(A) such that B ∈ ℘(A) ⇐⇒ B ⊆ A.
8

The set ℘(A) is called the Power set of A. Since the empty-set
∅ is a subset of every set, ℘(A) can never be empty set. What is
℘(∅) ? Since ∅ ⊆ ∅ , ∅ ∈ ℘(∅).
Suppose that A ∈ ℘(∅). Then A ⊆ ∅ . But then x ∈ A =⇒
x ∈ ∅. Since x ∈ ∅ is a contradiction, x ∈ A is also a contradiction.
Hence A = ∅. Thus ℘(∅) = {∅}. Further A ∈ ℘({∅}) ⇐⇒ A ⊆
{∅}. This shows that A = ∅ or A = {∅}. Thus ℘({∅}) = {∅, {∅}}.
℘({∅, {∅}}) = {∅, {∅}, {{∅}}, {∅, {∅}}} and so on.
The sets about which we usually think are not members of
them selves . Existence of a set A with the property that A ∈ A
will lead us to a very uncomfortable situation. The situations
(A ∈ B and B ∈ A), (A ∈ B and B ∈ C andC ∈ A) are also
annoying. To avoid these, we have the following axiom :
Axiom 1.0.28. (Axiom Of Regularity) A = 6 ∅ ⇒ ∃X[X ∈
A and∀x(x ∈ X ⇒ x 6∈ A)]. Equivalently: A =6 ∅ ⇒ ∃X[X ∈
A and X ∩ A = ∅].
Thus given a non empty-set A, there is a set X in A such that
no member of X is in A.
Theorem 1.0.29. Let A be a set. Then A 6∈ A.
Proof: Let A be a set. {A} 6= ∅. By the axiom of regularity,
∃X ∈ {A} such that x ∈ X ⇒ x 6∈ {A}. Now X ∈ {A} ⇐⇒ X =
A. Hence x ∈ A ⇒ x 6∈ {A}. Since A ∈ {A} , A 6∈ A. ✷

Theorem 1.0.30. Given sets A and B, A 6∈ B or B 6∈ A.


Proof: Suppose that A ∈ B and B ∈ A. By the axiom of
regularity ∃X ∈ {A, B} such that x ∈ X ⇒ x 6∈ {A, B}. Since
B ∈ A and B ∈ {A, B} and also A ∈ B and A ∈ {A, B}, there is
no X ∈ {A, B} such that x ∈ X ⇒ x 6∈ {A, B}. ✷
Let X be a set. The set X + = X {X} is called the bf
S
successor of X.
9

Proposition 1.0.31. Let X and Y be sets. Then X + = Y + ⇐⇒


X =Y.

Proof: If X = Y , then X + = Y + . Suppose that X 6= Y and


X = Y + . Then X ∪ {X} = Y ∪ {Y }. Since X ∈ X ∪ {X} ,
+

X ∈ Y ∪ {Y } and since X 6= Y , X ∈ Y . Similarly Y ∈ X. This


is a contradiction.✷
A set S is called a successor set if (i) {∅} ∈ S (ii) X ∈ S =⇒
+
X ∈ S.
The following axiom asserts that there is an infinite set.

Axiom 1.0.32. (Axiom of infinity).There exists a successor


set.

Proposition
T 1.0.33. Let X be a set of successor sets. Then
S∈X X is also a successor set.

Proof: Since
T each S is a successor
T set, {∅} ∈ S ∀S ∈ X.
Hence {∅} ∈ S∈X S. Let x ∈ S∈X S. Then x ∈ S ∀S ∈ X.
Since
T each S ∈ X is a successor set, x+ ∈ S ∀S ∈ X. Hence,x+ ∈
S∈X S. ✷

Corollary 1.0.34. Let X be a successor set. Then X contains


the smallest successor set contained in X.

Proof: The intersection of all successor sets contained in X


is the smallest successor set contained in X. ✷

Corollary 1.0.35. Let X and Y be successor sets. Let A be the


smallest successor set contained in X and B the smallest successor
set contained in Y . Then A = B.
T
Proof: X Y is also a successor set.T Thus A and B are both
smallest successor sets contained in X Y . ✷
10

Let X be a successor set . The smallest successor set contained


X which is the smallest successor set contained in any other suc-
cessor set is called the set of N atural N umbers. and is denoted
by N . {∅} is denoted by 1 and is called one. {∅}+ = {∅, {∅}} is
denoted by 2 and is called two and so on.

1.1 Ordered Pair, Cartesian Product


Let X be a set. Let a, b ∈ X. Then {{a}, {a, b}} ⊆ ℘(X). We
denote the set {{a}, {a, b}} by (a, b) and call it an ordered pair.
Thus (a, b) ∈ ℘(℘(X)).

Proposition 1.1.1. (a, b) = (b, a) ⇔ a = b.

Proof: Suppose that (a, b) = (b, a). Then {{a}, {a, b}} =
{{b}, {b, a}}. Since {a, b} = {b, a}, {a} = {b}. Hence a = b.
Clearly a = b implies (a, b) = (a, a) = (b, a). ✷
Observe that (a, a) = {{a}, {a, a}} = {{a}, {a}} = {{a}}.
Let X and Y be sets. Then the set

X ×Y = {(a, b) | a ∈ X and b ∈ Y }

which is a subset of ℘(℘(X ∪Y )) is called the cartesian product


of X and Y .

Proposition 1.1.2. Let A , B and C be sets. Then

1. (A ∪ B) × C = (A × C) ∪ (B × C).

2. (A ∩ B) × C = (A × C) ∩ (B × C).

3. (A − B) × C = (A × C) − (B × C).
11

Proof: We have the following equivalences:

(x, y) ∈ (A ∪ B) × C ⇔ x ∈ A ∪ B and y ∈ C
⇔ (x ∈ A and y ∈ C) or (x ∈ B and y ∈ C)
⇔ (x, y) ∈ (A × C) or (x, y) ∈ (B × C)
⇔ (x, y) ∈ (A × C) ∪ (B × C)

This proves (i). Similarly the rest can be proved.

Proposition 1.1.3. A × B = ∅ ⇔ A = ∅ or B = ∅.

Proof: If A = ∅, then (x, y) ∈ A × B =⇒ x ∈ ∅ and y ∈ B.


Since x ∈ ∅ is a contradiction (x, y) ∈ ∅×B is also a contradiction.
Hence ∅ × B = ∅. Similarly A × ∅ = ∅. Conversely suppose that
A 6= ∅ and B 6= ∅. Then ∃x ∈ A and ∃y ∈ B. Thus (x, y) ∈ A×B
and hence A × B 6= ∅. ✷

Exercise 1.1.1. Show that


T
1. A ∅ = ∅.
S
2. A ∅ = A.

3. A − ∅ = A.

4. ∅ − A = ∅.
T
5. A − (A − B) = A B.
T
6. A − (A B) = A − B.
S
7. A B = A ⇐⇒ B ⊆ A.
T S S T S
8. (A B) C = (A C) (B C).
T S T S
9. (A B) C = A (B C) ⇐⇒ C ⊆ A.
12
S S
10. A ⊆ B =⇒ C A⊆C B.

11. (A − B) − C = (A − C) − B.
T S
12. A (B A) = A.
S T
13. A = A (B A).
L S
14. A B = (A − B) (B − A).
L L L L
15. (A B) C = A (B C).
L L
16. A ∅ = A = ∅ A.
L L
17. A B = B A.
L
18. A B = ∅ ⇐⇒ A = B.
T L T L T
19. A (B C) = (A B) (A C).
L L
20. A C = B C ⇐⇒ A = B.

21. A ⊂ B ⇐⇒ ℘(A) ⊆ ℘(B).


T T
22. ℘(A B) = ℘(A) ℘(B).
S S
23. ℘(A) ℘(B) ⊆ ℘(A B). Show by means of an example
that equality need not hold.

24. Show that A × C ⊆ B × C, C 6= ∅ implies A ⊆ B.

25. Show that (A × B = B × A) ⇐⇒ (A = ∅ or B = ∅ or A =


B).

26. Suppose that A , B and C are nonempty-sets. Is (A × B) ×


C = A × (B × C)? Support.
T T T
27. Show that (A B) × (C D) = (A × C) (B × D).
13

28. A ⊆ A × A =⇒ A = ∅. Hint: Use the axiom of regularity.

29. A = A × B =⇒ A = ∅.

30. Show that if A contains n elements and B contains m ele-


ments then A × B contains n · m elements.

31. Suppose that A contains n elements. Show that ℘(A) con-


tains 2n elements.

32. Can ℘(A) be ∅? Support.

33. Show that a union of successor sets is a successor set.

34. Let A be a successor set. Can ℘(A) be a successor set?


support.

35. Let A and B be successor sets. Can A − B be a successor


set? Support.

36. Show that X + 6= X for every set X.

37. (X + )+ 6= X for every set X.


14
Chapter 2

Relations and Functions

2.1 Relations
Definition 2.1.1. Let A, B be any two sets. Then a subset R of
A × B is called a relation from A to B. Let (a, b) ∈ R then we
write aRb and we read a is related to b. The set Dom R =
{a |(a, b) ∈ R} is called the domain of R and the set Range R =
{b |(a, b) ∈ R} is called the range of R respectively.

Example 2.1.2. 1. ∅ and A × B are subsets of A × B. Thus


these are relations from A to B.

2. If either A or B is empty then there is only one relation


from A to B. This relation is called an empty relation.

3. If |A| = m and |B| = n then there are 2mn relations from A


to B.

4. If A = {a, b, c} and B = {1, 2} then R = {(a, 1), (b, 1)} is


a relation from A to B. Observe that Dom R = {a, b} and
Range R = {1}.

15
16

If R is a relation from A to B then R−1 = {(b, a) |(a, b) ∈ R}


is a relation from B to A. This relation is known as inverse
relation of R. S T
Let R and S be relations from A to B. Then R S, R S
and R − S are subsets of A × B and hence they are relations from
A to B.

2.2 Composition of Relations


Let R and S be relations from X to Y and Y to Z respectively.
We define SoR as follows:

SoR = {(x, z) ∈ X × Z | ∃y ∈ Y with (x, y) ∈ R, (y, z) ∈ S}

SoR is called the composition of R and S. Clearly it is a relation


from X to Z.

Example 2.2.1. Consider A = {1, 2, 3}, B = {a, b, c} and C =


{x, y, z}. Let R = {(1, a), (2, b), (3, c)} and S = {(a, x), (a, y), (b, z)}.
Then SoR = {(1, x), (1, y), (2, z)}.

Definition 2.2.2. Let X be a set. A subset R of X × X is called


a relation on X.

Example 2.2.3. ∅ and X × X are relations on X. If X is an


empty set then there is only one relation which is ∅.

Example 2.2.4. △ = {(x, x) | x ∈ X} is a relation on X called


the diagonal relation on X.

Example 2.2.5. Let X = {a, b, c}. R = {(a, b), (b, a), (a, c)} is a
relation on X.

Example 2.2.6. Let X be a set. Let R = {(A, B) | A, B ∈


℘(X) and A ⊆ B} is a relation on ℘(X).
17

Proposition 2.2.7. Let R, S and T be relations on X. Then


(RoS)oT = Ro(SoT ).

Proof: We have the following equivalences: (x, y) ∈ (RoS)oT


⇔ ∃z ∈ X such that (x, z) ∈ T and (z, y) ∈ RoS
⇔ ∃z, u ∈ X such that (x, z) ∈ T, (z, u) ∈ S and (u, y) ∈ R
⇔ ∃u ∈ X such that (x, u) ∈ SoT and (u, y) ∈ R
⇔ (x, y) ∈ Ro(SoT )
This proves the result.✷

Proposition 2.2.8. If R is any relation on X, then Ro△ = R =


△oR.

Proof: Since (x, x) ∈ △∀ x ∈ X, therefore (x, y) ∈ Ro△ ⇔


(x, y) ∈ R. Thus Ro△ = R. Similarly R = △oR.✷

Proposition 2.2.9. Let R, S and T be relations on X. Then


S S
1. Ro(S T ) = (RoS) (RoT ).
T T
2. Ro(S T ) ⊆ (RoS) (RoT ).
S S
3. (R S)oT = (RoT ) (SoT ).
T T
4. (R S)oT ⊆ (RoT ) (SoT ).
S
Proof: We have the followingS equivalences: (x, y) ∈ Ro(S T )
⇔ ∃z ∈ X such that (x, z) ∈ S T and (z, y) ∈ R
⇔ ∃z ∈ X such that ((x, z) ∈ S and (z, y) ∈ R)or ((x, z) ∈
T, (z, y) ∈ R)
⇔ (x, y) ∈ (RoS) S
or (x, y) ∈ (RoT )
⇔ (x, y) ∈ (RoS) (RoT )

This proves 1. Similarly one can prove the other results ✷


18

Example 2.2.10. Let X = {a, b, c}. Let R = {(a, b), (a, c)} and S =
{(b, c), (b, b)}. Then RoS = ∅ and SoR = {(a, c), (a, b)} = R.
Thus, RoS need not be equal to SoR. Observe that R =
SoR = △oR but S 6= △. Next, if T = {(a, a), (b, c), (b, b)} 6= △,
then RoT = {(a, c), (a, b)} = R = T oR but T 6= ∆. Thus
RoT = R = T oR need not imply that T = △.

Definition 2.2.11. Let R be a relation on X. Then the relation

R−1 = {(x, y) ∈ X × X | (y, x) ∈ R}

is called the inverse of R.

Example 2.2.12. Let R = {(a, b), (a, c)} be a relation on X =


{a, b, c}. Then R−1 = {(b, a), (c, a)}. Observe that RoR−1 =
{(b, b), (c, c)} and R−1 oR = {(a, a)}. Thus, RoR−1 6= R−1 oR.

Proposition 2.2.13. Let R and S be relations on X. Then


(R−1 )−1 = R and (RoS)−1 = S −1 oR−1 .

Proof: Clearly (x, y) ∈ R ⇔ (y, x) ∈ R−1 and (y, x) ∈


R−1 ⇔ (x, y) ∈ (R−1 )−1 . Thus R = (R−1 )−1 . Next, let (x, y) ∈
(RoS)−1 . Then (y, x) ∈ RoS. Hence ∃z ∈ X such that (y, z) ∈
S and (z, x) ∈ R. Thus ∃z ∈ X such that (x, z) ∈ R−1 and (z, y) ∈
S −1 and so (x, y) ∈ S −1 oR−1 . Similarly (x, y) ∈ S −1 oR−1 implies
that (x, y) ∈ (RoS)−1 . This proves (RoS)−1 = S −1 oR−1 .✷

Definition 2.2.14. A relation R on X is called

1. reflexive if (x, x) ∈ R ∀x ∈ X or equivalently △ ⊆ R.

2. symmetric if (x, y) ∈ R ⇒ (y, x) ∈ R or equivalently


R−1 = R.

3. antisymmetric T if−1(x, y) ∈ R and (y, x) ∈ R =⇒ x = y or


equivalently R R ⊆ △.
19

4. transitive if (x, y) ∈ R and (y, z) ∈ R =⇒ (x, z) ∈ R or


equivalently RoR ⊆ R.
Example 2.2.15. Let X = {a, b, c} and
R = {(a, a), (b, b), (c, c), (a, b), (b, c), (c, b)}.
Then R is reflexive but none of the rest of the three.
Example 2.2.16. Let X = {a, b, c} and R = {(a, b), (b, a)}.
Then R is symmetric but none of the rest of the three.
Example 2.2.17. Let X = {a, b, c} and R = {(c, b), (a, c)}.
Then R is anti-symmetric but none of the rest of the three.
Example 2.2.18. Let X = {a, b, c} and
R = {(a, b), (b, a), (a, a), (b, b), (a, c), (b.c)}.
Then R is transitive but none of the rest of the three.
Example 2.2.19. Let X = {a, b, c} and
R = {(a, a), (b, b), (c, c), (a, b), (b, a), (b, c), (c, b)}.
Then R is reflexive, symmetric but neither anti-symmetric nor
transitive.
Example 2.2.20. Let X = {a, b, c} and
R = {(b, c), (c, b), (b, b), (c, c)}.
Then R is symmetric and transitive but neither reflexive nor anti-
symmetric.
Remark 2.2.21. a. If a relation R on X is symmetric and
transitive then it is reflexive.
b. The relation which is reflexive, symmetric and anti-symmetric
is the diagonal relation. Thus, a reflexive, symmetric and
anti-symmetric relation is also transitive.
20

2.3 Equivalence Relations


Definition 2.3.1. A relation R on X which is reflexive, sym-
metric and transitive is called an equivalence relation. Let R
be an equivalence relation on X and x ∈ X. Then the subset
[x] = {y ∈ X | (x, y) ∈ R} is called the equivalence class
of X modulo R determined by element x. Since R is reflexive,
(x, x) ∈ R∀x ∈ X and hence x ∈ [x]∀x ∈ X. In other words
[x] 6= ∅.
Proposition 2.3.2. Let R be an equivalence relation on X. Then
(i) [x] = [y] ⇔ (x, y) ∈ R.

(ii) [x] 6= [y] ⇐⇒ [x] ∩ [y] = ∅.


Proof: (i). Suppose that [x] = [y]. Since R is an equivalence
relation, y ∈ [y] = [x]. Hence (x, y) ∈ R. Conversely, suppose
that (x, y) ∈ R. Let z ∈ [x]. Then (x, z) ∈ R. Since R is
symmetric and transitive therefore (x, y), (x, z) ∈ R gives (y, z) ∈
R; that is, z ∈ [y]. This proves that [x] ⊆ [y]. Interchanging x
and y, we get [y] ⊆ [x]. Thus [x] = [y].
(ii). If [x] ∩ [y] 6= ∅, then there exists z ∈ [x] ∩ [y] and so
(x, z), (z, y) ∈ R. Since R is transitive, (x, y) ∈ R. It follows from
(i) that [x] = [y]. Thus, [x] ∩ [y] = ∅. ✷

Definition 2.3.3. Let X be a set. A subset ℘ of the power set of


X is called a partition of X if (i) union of members of ℘ is X
i.e ∪A∈℘ = X and (ii) A, B ∈ ℘, A 6= B ⇒ A ∩ B = ∅.
Corollary 2.3.4. If R is an equivalence relation on X then ℘R =
{[x] | x ∈ X} is a partition of X.
Proof: Since x ∈ [x]∀x ∈ X (for R is reflexive), union of
members of ℘R is X. Also [x] 6= [y] ⇒ [x] ∩ [y] = ∅. Thus ℘R is
a partition. ✷
21

Thus, the partition ℘R is the partition determined by the


equivalence relation R. The set ℘R is also denoted by X/R and
is called the quotient set of X modulo R.

Proposition 2.3.5. Let ℘ be a partition of X. Define a relation


R℘ = {(x, y) | ∃A ∈ ℘ such that x, y ∈ A} = ∪A∈℘ A × A. Then
R℘ is an equivalences relation such that ℘R℘ = ℘.

Proof: Since union of members of ℘ is X, given x ∈ X, x ∈ A


for some A ∈ ℘. Hence (x, x) ∈ R℘ ∀x ∈ X. Thus R℘ is reflexive.
Suppose that (x, y) ∈ R℘ . Then ∃A ∈ ℘ such that x, y ∈ A and
so y, x ∈ A. Hence (y, x) ∈ R℘ . Thus R℘ is symmetric. Suppose
that (x, y) ∈ R℘ and (y, z) ∈ R℘ . Then ∃A ∈ ℘ and B ∈ ℘
such that x, y ∈ A and y, z ∈ B. Since y ∈ A ∩ B, A ∩ B 6= ∅.
Further since ℘ is a partition, A = B. Hence x, z ∈ A ∈ ℘. Thus
(x, z) ∈ R℘ . This shows that R℘ is transitive. Next Rx℘ is the
member A of ℘ such that x ∈ A. Hence ℘R℘ = ℘. ✷

Proposition 2.3.6. R℘R = R for every equivalence relation R.

Proof: Suppose that (x, y) ∈ R. Then x, y ∈ [x] ∈ ℘R . Hence


(x, y) ∈ R℘R . Suppose that (x, y) ∈ R℘R . Then ∃Rz ∈ ℘R such
that x, y ∈ Rz . Hence ∃z ∈ R such that (x, z) ∈ R and (y, z) ∈ R.
But then since R is symmetric and transitive, (x, y) ∈ R. Thus
R = R ℘R . ✷

Remark 2.3.7. Every partition ℘ of X determines an equivalence


relation R℘ such that X/R℘ = ℘ and conversely. Thus, every
partition can be realized faithfully as an equivalence relation and
every equivalence relation can be realized as a partition.

Example 2.3.8. Let R be a relation (not necessarily equivalence)


on X. Define
Rx = {y ∈ X | (x, y) ∈ R}.
22

Suppose that ℘ = {Rx | x ∈ X} is a partition of X. Can we


infer that R is an equivalence relation? No; For example, take
X = {a, b, c}, R = {(a, b), (b, c), (c, a)}. Then

Ra = {b}, Rb = {c}, Rc = {a}.

Thus, {Ra , Rb , Rc } is a partition of X where as R is not an equiv-


alence relation (it is neither reflexive nor symmetric nor transi-
tive).

Example 2.3.9. Let ℘ ⊆ ℘(X) (not necessarily a partition).


Define a relation R℘ on X by

(x, y) ∈ R℘ ⇔ ∃A ∈ ℘ such that x, y ∈ A.

Suppose that R℘ is an equivalence relation. Can we infer that


℘ is a partition? No; For example, take

℘ = {{a, b}, {b, c}, {c, a}} ⊆ ℘(X)

where X = {a, b, c}. Then clearly R℘ = X × X is an equivalence


relation.
23

EXERCISES

1. Find the number of relations on a set containing n elements.


2
Ans.2n .

2. Let X = {a, b, c} and R, S be relations on X given by

R = {(a, b), (b, c), (c, a)}, S = {(a, a), (a, c), (b, b)}.

S,(iii) RoS, (iv) R−1 .


S T
Find out (i)R S, (ii)R

3. Find out the number of reflexive relations on a set containing


n elements. S
Hint. A reflexive relation on X ×X can be written as ∆ S
where S ⊆ X × X − ∆.

4. Find out the number of symmetric relation on a set contain-


ing n elements.

5. Let R and S be equivalence relations on X. Show that RoS


is an equivalence relation if and only if RoS = SoR.

6. Let pn denote the number of equivalence relations on a set


containing n elements. Show that

pn+1 = Σnr=0 (n Cr )pr

7. Let X = {a, b, c, d} and

R=
{(a, a), (b, b), (c, c), (d, d), (a, b), (b, c), (a, c), (b, a), (c, b), (c, a)}.

Show that R is an equivalence relation. Find ℘R . Can we


find an other relation S such that ℘R = ℘S? Support.
24

2.4 Functions
Definition 2.4.1. Let X and Y be sets. A subset f of X × Y is
called a function or a mapping (or a map) from X to Y if
(i) ∀x ∈ X, ∃y ∈ Y such that (x, y) ∈ f
(ii) (x, y1 ) ∈ f and (x, y2 ) ∈ f ⇛ y1 = y2 .
The set X is called the domain of f and Y is called the
codomain of f . If (x, y) ∈ f , we write y = f (x) and call it
the image of the element x ∈ X under the map f . Thus, under
this notation f = {(x, f (x)) | x ∈ X}.

Intuitively a function f from X to Y is a rule which assigns


each x ∈ X to a unique element y ∈ Y . Any two functions f
and g from X to Y are equal iff f (x) = g(x)∀x ∈ X. We read
f : X −→ Y as ’f is a map from X to Y ’.
Let f : X → Y and g : Y → Z be maps. Then

gof = {(x, z) | (x, y) ∈ f and (y, z) ∈ g}

is also a map from X to Z and is called the composite of f and


g. Thus, the map gof : X → Z is given by

(gof )(x) = g(f (x))∀x ∈ X.

The subset ∆ of X × X is also a map from X to X. This map


is called the identity map on X and is denoted by IX . Thus,
IX (x) = x∀x ∈ X. Clearly f oIX = f = IY of for every map
f :X →Y.
Let Y be a subset of X. Then iY = {(y, y) | y ∈ Y } is a map
from Y to X called the inclusion map from Y to X. This map
is some times denoted by the symbol Y ֒→ X.
Let f : X → A be a map and Y a subset of X. We have the
inclusion map iY : Y → X. The composition f oiY is a map from
Y to A and is called the restriction of f to Y . The map f oiY is
also denoted by f |Y .
25

Let X and Y be sets and y ∈ Y . Then X × {y} is a map f


from X to Y such that f (x) = y ∀x ∈ X. This map is called a
constant map.
Let X and Y be sets. Consider the Cartesian product X × Y .
The map p1 : X × Y −→ X defined by p1 ((x, y)) = x is called the
first projection and p2 : X × Y → Y defined by p2 ((x, y)) = y is
called the second projection map.

Proposition 2.4.2. Let f : X → Y, g : Y → Z and h : Z → U


be maps. Then (hog)of = ho(gof ).

Proof: Since ((hog)of )(x) = (hog)(f (x)) = h(g(f (x))) =


h((gof )(x)) = (ho(gof ))(x) for all x ∈ X. Hence ho(gof ) =
(hog)of .✷
Let f : X −→ Y be a map. Then f ⊆ X × Y . Consider

f −1 = {(y, x) | (x, y) ∈ f }.

Then f −1 ⊆ Y × X need not be a map from Y to X for two


reasons,
(i) for y ∈ Y there may not be any x ∈ X such that (x, y) ∈ f
and so there may not be any x ∈ X such that (y, x) ∈ f −1 .

(ii)(y, x1 ) ∈ f −1 and (y, x2 ) ∈ f −1 need not imply that x1 =


x2 .
Thus, f −1 will be a map if and only if

1. for all y ∈ Y ∃x ∈ X | (x, y) ∈ f and

2. (x1 , y) ∈ f and (x2 , y) ∈ f =⇒ x1 = x2 .

Definition 2.4.3. A map f : X −→ Y is called a surjective map


or an onto map if ∀y ∈ Y there exists x ∈ X such that

(x, y) ∈ f.
26

Thus, f is surjective if ∀y ∈ Y there exists x ∈ X such that

f (x) = y.

A map f ; X −→ Y is called injective or one-one if

f (x1 ) = f (x2 ) ⇒ x1 = x2

or equivalently
x1 6= x2 ⇒ f (x1 ) 6= f (x2 ).
A map f which is injective as well as surjective is called a
bijective map or one- one and onto map.

Thus, f −1 is a map if and only if f is bijective.

Definition 2.4.4. The map f −1 is called the inverse of f and


this is also bijective.

Example 2.4.5. An injective map need not be surjective. Take


X = {a, b}, Y = {x, y, z}. Define f : X −→ Y as f (a) = x,
f (b) = y. Then, f is injective but not surje-ctive because there is
no element in X whose image is z.

Example 2.4.6. A surjective map need not be injective. Take


X = {a, b, c}, Y = {x, y}. Define f : X → Y as

f (a) = x = f (b), f (c) = y.

Then, f is surjective, but not injective because f (a) = f (b)


where as a 6= b.

Proposition 2.4.7. Let f : X −→ Y be a bijective map. Then


f −1 : Y −→ X is also a bijective map. Also (i) (f −1 )−1 =
f, (ii)f −1 of = IX and f of −1 = IY .
27

Proof: Let f be a bijective map. Then we have already


observed that f −1 is a map from Y to X. Suppose that (y1 , x) ∈
f −1 and (y2 , x) ∈ f −1 . Then (x, y1 ) ∈ f and (x, y2 ) ∈ f . Since
f is a map, y1 = y2 . Thus f −1 is injective. Let x ∈ X, then
(x, f (x)) ∈ f and hence (f (x), x) ∈ f −1 . This shows that f −1 is
surjective. We also observe that

(f −1 of )(x) = f −1 (f (x)) = x ∀x ∈ X

and
(f of −1 )(y) = f (f −1 (y)) = y ∀y ∈ Y.
Thus, f −1 of = IX and f of −1 = IY . The fact that (f −1 )−1 = f
follows from the definition of f −1 . ✷

Proposition 2.4.8. (i). Composite of any two injective maps


is injective,

(ii) Composite of two surjective maps is surjective, and

(iii) composite of two bijective map is bijective.

Proof: (i) Let f : X −→ Y and g : Y −→ Z be injective


maps. Suppose (gof )(x1 ) = (gof )(x2 ). Since g is injective, we
have
f (x1 ) = f (x2 ).
Further, since f is injective, we have x1 = x2 . Hence gof is
injective.

(ii) Suppose f and g are surjective. Let z ∈ Z. Since g is


surje- ctive, therefore there exists y in Y such that g(y) = z.
Again, since f is surjective therefore there exists x ∈ X such that

f (x) = y.
28

But, then (gof )(x) = g(f (x)) = g(y) = z. Hence gof is surjec-
tive.
(iii) follows from (i) and (ii). ✷

Proposition 2.4.9. Let f : X −→ Y and g : Y −→ Z be maps.


(i) If gof is surjective then g is surjective.
(ii) If gof is injective, then f is injective.
Proof: (i) Suppose that of gof is surjective. Let z ∈ Z.
Since gof is surjective therefore there exists x in X such that
(gof )(x) = z, i.e; g(f (x)) = z. Hence g is surjective.
(ii) Suppose that gof is injective and f (x1 ) = f (x2 ). Then
g(f (x1 )) = g(f (x2 )) and so (gof )(x1 ) = (gof )(x2 ). Since gof is
injective, therefore x1 = x2 . Hence f is injective. ✷

Corollary 2.4.10. If gof is bijective, then g is surjective and f


is injective. ♯.
Proposition 2.4.11. A map f : X −→ Y is injective if f f can
be left cancelled in the sense that

f og = f oh =⇒ g = h.

Proof: Suppose that f is injective and f og = f oh. Then


(f og)(z) = (f oh)(z)∀z ∈ Z. Since f is injective, g(z) =
h(z)∀z ∈ Z. This shows that g = h. Suppose, now that
f is not injective. Then there exists x1 , x2 in X such that
x1 6= x2 and f (x1 = f (x2 ). Take Z={x1 , x2 }. Define a map
g : Z −→ X as g(x1 ) = x1 = g(x2 ) and a map h : Z −→ X as
h(x1 ) = x2 = h(x2 ). Then g 6= h and f og = f oh. ✷

Proposition 2.4.12. A map f : X −→ Y is surjective if f it can


be right cancelled in the sense that

gof = hof =⇒ g = h.
29

Proof: Suppose that f is surjective and g, h are maps from


Y to Z such that gof = hof . Then g(f (x)) = h(f (x))∀x ∈ X.
Since f is surjective g(y) = h(y) ∀y ∈ Y , i.e; g = h. Suppose
that f is not surjective. Then there exists y0 6= f (x) for all
x ∈ X. Take Z = {a, b}. Define g : Y −→ Z as g(y0 ) = a and
g(y) = b∀y 6= y0 and h : Y −→ Z as h(y) = b∀y ∈ Y . Clearly,
then g 6= h and gof = hof . ✷

Corollary 2.4.13. A map f from X to Y is bijective iff it can


be left as well as right cancelled.♯

Proposition 2.4.14. Let f : X −→ Y be a map. Then f is


bijective iff ∃ a map g : Y −→ X such that gof = IX and f og =
IY and then g = f −1 .

Proof: If f is bijective, then f −1 of = IX and f of −1 = IY .


Let g : Y −→ X be such that gof = IX and f og = IY . Since
gof = IX is injective,f is injective. Since f og = IY is surjective,
f is surjective. Further then f −1 of = IX and f of −1 = IY . Thus
the result follows. ✷

Corollary 2.4.15. Let f : X −→ Y and g : Y −→ Z be bijective


maps. Then (gof )−1 = f −1 og −1 .

Proof: Follows from the following identity:

(f −1 og −1 )o(gof ) = (f −1 o(g −1 og)of = f −1 of = IX

Similarly (gof )o(f −1 og −1 ) = IY . ✷

Proposition 2.4.16. There is no surjective map from any set X


to it’s power set ℘(X).

Proof: Let f : X −→ ℘(X) be a map. Consider the set


A = {x ∈ X | x 6∈ f (x)}. Then A ∈ ℘(X). Suppose that
30

f (y) = A for some y ∈ X. If y 6∈ A = f (y), then y ∈ A. If


y ∈ A = f (y) then y 6∈ f (y) = A. Hence the supposition that
f (y) = A for some y ∈ X is false. This shows that f can not be
surjective. ✷
Let X and Y be sets. Then Y X denotes the set of all maps
from X to Y .
What is X ∅ and ∅X ?
Example 2.4.17. Let X be a set and 2 ={0,1}. Define a map
φ : ℘(X) −→ 2X by φ(A)(x) = 0 if x 6∈ A and φ(A)(x) = 1 if
x ∈ A. Check that the map φ is bijective.
Let f : X −→ Y be a map. Let A ⊆ X and B ⊆ Y . The
subset f (A) = {f (a) | a ∈ A} of Y is called the image of A under
the map f . The subset f −1 (B) = {x ∈ X | f (x) ∈ B} of X is
called the inverseimage of B under f .
What are f −1 (Y ) and f −1 (∅)? To say that f is surjective is to
say that f (X) = Y .
Proposition 2.4.18. Let f be a map from X to Y and A ⊆ X.
Then A ⊆ f −1 (f (A)). Also A = f −1 (f (A))∀A ⊆ X if and only if
f is injective.
Proof: Let a ∈ A. Then f (a) ∈ f (A) and hence by defini-
tion a ∈ f −1 (f (A)). Thus A ⊆ f −1 (f (A)). Suppose that f is
injective. Let x ∈ f −1 (f (A)). Then f (x) ∈ f (A) (by def). Hence
∃a ∈ A such that f (x) = f (a). Since f is injective x = a ∈ A.
Thus f −1 (f (A)) ⊆ A and therefore A = f −1 (f (A)). Suppose
that f is not injective. Then ∃x1 , x2 ∈ X, x1 6= x2 such that
f (x1 ) = f (x2 ) = y (say). Take A = {x1 }. Then f (A) = {y} and
{x1 , x2 } ⊆ f −1 (f (A)). Hence A 6= f −1 (f (A)). ✷

Proposition 2.4.19. Let f : X −→ Y be a map and B ⊆ Y .


Then f (f −1 (B)) ⊆ B. Also B = f (f −1 (B))∀B ⊆ Y if and only
if f is surjective.
31

Proof: Let y ∈ f (f −1 (B)). Then y = f (x) for some x ∈


f −1 (B). But then y = f (x) ∈ B. Hence f (f −1 (B)) ⊆ B. Sup-
pose that f is surjective and y ∈ B. Then ∃x ∈ X such that
f (x) = y. Clearly x ∈ f −1 (B) and hence y = f (x) ∈ f (f −1 (B)).
There fore B = f (f −1 (B)). Suppose now that f is not surjective.
Then ∃b ∈ Y such that b 6∈ f (X). Now f −1 ({b}) = ∅ and hence
f (f −1 ({b})) = ∅ 6= {b}. ✷

Proposition 2.4.20. Let f : X −→ Y be a map. Let A1 and A2


be subsets
S of X. Then S
(i)f (A1 TA2 ) = f (A1 ) Tf (A2 ).
(ii)f (A1 A2 ) ⊆ f (A1 ) f (A2 ).
Further in (ii) equality holds for every pair of subsets A1 and A2
of X if and only if f is injective.

Proof: The proof of (i) and (ii) areT left as exercises. Suppose
now that f is injective. Let y ∈ f (A1 ) f (A2 ). Then ∃a ∈ A1 and
b ∈ A2 such that yT= f (a) = f (b). Since
T f is injective aT= b and so
y = f (a) ∈ f (A1 A2 ). Thus
T T 2 ) ⊆ f (A1 A2 ). But
f (A1 ) f (A
already (from (ii)) f (A1 A2 ) ⊆ f (A1 ) f (A2 ). Thus equality
holds in (ii) if f is injective. Conversely if f is not injective.
Then ∃x1 6= x2 in X such that f (x1T ) = f (x2 ) = b (say). Take
A1 = T {x1 }, A2 = {x2 }. Then f (A1 A2 ) = f (∅) = ∅ whereas
f (A1 ) f (A2 ) = {b} 6= ∅. ✷

Proposition 2.4.21. Let f : X −→ Y be a map. Let B1 and B2


be subsets Tof Y . Then
(i)f (B1 SB2 ) = f −1 (B1 ) Sf −1 (B2 )
−1
T
(ii)f −1 (B1 B2 ) = f −1 (B1 ) f −1 (B2 ) and
(iii)f −1 (B1 − B2 ) = f −1 (B1 ) − f −1 (B2 ).

x ∈ f −1 (B1 B2 )
T
Proof: (i) T
⇐⇒ f (x) ∈ B1 B2
⇐⇒ f (x) ∈ B1 and f (x) ∈ B2
32

⇐⇒ x ∈ f −1 (B1 ) T
and x ∈ f −1 (B2 )
⇐⇒ x ∈ f −1 (B1 ) f −1 (B2 )
This proves (i). Similarly we can prove the rest of the two.✷
Let I be a set and X a set of sets. A surjective mapA : I −→ X
is called a family of sets. We denote the image A(α) of α by
Aα . This family of sets is denoted by {Aα | α ∈ I}. The set I is
called the indexing set of the family.
Let {Aα | α ∈ I} be a family of sets. Then the set
S
α∈I Aα = {x | x ∈ Aα f or some α ∈ I}

is called the union of the family and


T
α∈I Aα = {x | x ∈ Aα ∀α ∈ I}

is called the intersection of the family.


Proposition 2.4.22. (De Morgan’s Law)Let X be a set and {Aα |
α ∈ I} be a family of sets. Then {X − Aα | α ∈ I} is another
family ofSsets and T
(i)X − ( α∈I Aα ) = α∈I (X − Aα )

The proof of the above proposition is left as an exercise.


Let {Xα | α ∈ I} be a family of sets. Then the set
Q S
α∈I Xα = {x : I −→ α∈I Xα such that x(α) ∈ Xα ∀α ∈ I}

is called the cartesian product ofQ the family.


For each α0 ∈ I,the map pα0 : α∈I Xα −→ Xα0 defined by
pα0 (x) = x(α0 ) is called the α0th projection map.
Following is an other axiom of set theory.
Axiom 2.4.23. (Axiom of Choice). Let {Xα | α ∈ I} be a
non empty familyQof non empty sets (i.e.I is non empty and Xα 6=
∅∀α ∈ I).SThen α∈I Xα is non empty set.i.e.there exists a map
c : I −→ α∈I Xα such that c(α) ∈ Xα ∀α ∈ I.
33

Let f : X −→ Y and g : Z −→ U be maps. The map


f × g : X × Z −→ Y × U defined by (f × g)((x, z)) = (f (x), g(z))
is called product of the map f with the map g. Clearly products
of injective maps are injective maps and those of surjective maps
are surjective.
Let f : X −→ Y be a map and S an equivalence relation on
Y . Then (f × f )−1 (S) is an equivalence relation (verify). Let R
be an equivalence relation on X. Then (f × f )(R) need not be an
equivalence relation on Y even if f is surjective (give an example
to support this).
The equivalence relation (f × f )−1 (△) on X is called the
kernel of f and is denoted by kerf . It follows from the defi-
nitions that f is injective if and only if Ker f = △, the diagonal
relation on X.
Proposition 2.4.24. Let f : X −→ Y be a surjective map. Let
R be an equivalence relation on X containing the kernel of f .
Then (f × f )(R) is an equivalence relation on Y such that (f ×
f )−1 ((f × f )(R)) = R.
Proof: Since f is surjective (f × f )(R) is reflexive as well as
symmetric. We prove that it is transitive also. Let (u, v), (v, w) ∈
(f × f )(R). Then ∃(x, y), (z, t) ∈ R such that (f (x), f (y)) =
(u, v) and (f (z), f (t)) = (v, w) This shows that f (y) = f (z) = v.
Hence (y, z) ∈ (f × f )−1 (△) = Kerf ⊆ R. Since R is transitive
(x, t) ∈ R. But, then (u, w) = (f (x), f (t)) ∈ (f × f )(R). Thus
(f × f )(R) is an equivalence relation. Finally we show that (f ×
f )−1 ((f × f )(R)) = R. Clearly R ⊆ (f × f )−1 ((f × f )(R)). Let
(x, y) ∈ (f × f )−1 ((f × f )(R)). Then (f (x), f (y)) ∈ (f × f )(R).
Hence ∃(z, t) ∈ R such that (f (x), f (y)) = (f (z), f (t)). But then
f (x) = f (z) and f (y) = f (t) This shows that (x, z) and (y, t)
belong to (f × f )−1 (△) Since (f × f )−1 (△) is supposed to be
contained in R, (x, z), (y, t) and (z, t) are all in R. Since R is an
equivalence relation (x, y) ∈ R. This completes the proof. ✷
34

Corollary 2.4.25. (Correspondence Theorem).Let f : X −→ Y


be a surjective map. Let R(X) denote the set of all equivalence
relations on X containing Kerf and R(Y ) the set of all equiv-
alence relations on Y . Then f induces a bijective map from
φ : R(X) −→ R(Y ) defined by φ(R) = (f × f )(R).

Proof: From the above proposition it follows that (f ×f )(R) ∈


R(Y )∀R ∈ R(X). Thus φ is indeed a map from R(X) to R(Y ).
Since f is surjective f × f is also surjective. Hence (f × f )((f ×
f )−1 (S)) = S for all S ∈ R(Y ). this shows that φ is surjec-
tive.(note that (f × f )−1 (S) ∈ R(X)). Further suppose that
φ(R1 ) = φ(R2 ). Then (f × f )(R1 ) = (f × f )(R2 ). Since R1
and R2 are equivalence relations containing Kerf it follows from
the above proposition that R1 = (f × f )−1 ((f × f )(R1 )) = (f ×
f )−1 ((f × f )(R2 )) = R2 . This proves that φ is injective. ✷

Let X be a set and R an equivalence relation on X. Consider


the quotient set X/R = {Rx | x ∈ X} The map ν : X −→ X/R
defined by ν(x) = Rx is called the quotient map. Clearly ν is
surjective and (ν × ν)−1 (△) = R. Thus every equivalence relation
is kernel of a map. We shall show that if f ; X −→ Y is a surjective
map then Y can be realized as a quotient set through a bijective
map.

Theorem 2.4.26. Let f : X −→ Y be a surjective map. Let


R be an equivalence relation on X containing Kerf . Let S =
(f × f )(R). Then there is a bijective map f : X/R −→ Y /S such
that the diagram
35

X f ✲ Y

ν ν

❄ f ❄
X/R ✲ Y/S
is commutative.

Proof: Rx1 = Rx2


⇐⇒ (x1 , x2 ) ∈ R
⇐⇒ (f (x1 ), f (x2 )) ∈ (f × f )(R) = S
Sf (x1 ) = Sf (x2 )
Hence f is indeed a map which is injective. Further since f is
surjective, every member of Y /R is of the form Sf (x) = f (Rx ).
This shows that f is surjective. The rest is also clear by the
definition. ✷

Corollary 2.4.27. (Fundamental Theorem Of Maps) Let f :


X −→ Y be a surjective map and R = (f × f )−1 (△) = Kerf .
Then there is a bijective map φ : X/R −→ Y such that φoν = f .

Proof: Clearly (f ×f )((f ×f )−1 (△)) = △. One also observes


that the quotient map ν : Y −→ Y /△ is bijective map given by
y −→ {y}. The result follows from the above theorem. ✷

EXERCISES
36

1. Let X be a finite set containing n elements and Y a set


containing m elements. Suppose that n ≤ m. Find the
number of injective maps from X to Y . What happens if
m ≤ n?

2. Find the number of surjective maps from a set containing n


elements to a set containing m elements.

3. Let X be a set. Show that there is no injective map from


P (X) to X.

4. Let X Y denote the set of all maps Y to X. Suppose that


Y 6= ∅. Show that there is a surjective map from X to Y X
if and only if Y is a singleton set.

5. Let X, Y and Z be sets. Show that there is a bijective map


from X Y ×Z to (X Y )Z .

6. Let R and S be two equivalences relations on a set X such


that R ⊆ S. Show that there is a bijective map

φ : X/S → (X/R)/(ν × ν)(S)

such that the diagram formed by quotient maps is commu-


tative.

7. Let {Xα | α ∈ I} be a family of nonempty sets. Show that


each projection map is a surjective map.
Hint.Use axiom of choice.

8. Letf : X −→ Y be a surjective map. Show that there is an


injective map t from Y to X such that f ot = IY .
Hint.Use axiom of choice.

9. Let f : X −→ Y be an injective map. Show that there is a


surjective map s : Y −→ X such that sof = IX .
37

10. Show that f (A) − f (B) ⊆ f (A − B). Show further that the
equality holds provided that f is injective.

11. Show that there is a bijective map from (Y X )Z to Y Z×X .


38
Chapter 3

Real Number System

Here we assume familiarity with rational numbers. The rational


number system is inadequate for many purposes, both as a field
and as an ordered set. This leads to the introduction of ‘irrational
numbers’ which are written as infinite decimal expansions and are
considered to be
√ ‘approximated’ by corresponding finite decimals.
For example, 2 is approximated by 1.414. To introduce the
construction of real number system, we consider the following
example:

Example 3.0.1. Consider the equation x2 = 2, where x is un-


known in the equation. We now prove that there is no rational
number p such that p2 = 2.

Suppose, if possible, that p is a rational such that p2 = 2.


Then p = m/n, where m and n are non-zero integers and both
are not even. Then

m2 = 2n2 (3.0.1)

This shows that m2 is even and so m will be even (if m is odd then
m2 is odd). Suppose that m = 2k, for some integer k. Putting
this value in equation Eq.3.0.1, we have n2 = 2k 2 . Thus, n will

39
40

also be even. This is a contradiction because both m and n are


not even. Thus, the equation x2 = 2 has no solution in the ratio-
nal number system. Next, consider the following example:

Example 3.0.2.

A = {x ∈ Q |x2 < 2} (3.0.2)

and

B = {x ∈ Q |x2 > 2}, (3.0.3)

where Q denote the set of rational numbers. We, now, claim that
A has no largest rational number and B has no smallest number.
To prove this assume that p is any rational number. Take

p2 − 2 2(p + 1)
q = p− = (3.0.4)
p+2 p+2
Then
 2
2 2(p + 1)
q −2 = −2 (3.0.5)
p+2
4p2 + 8p + 4 − 2p2 − 8p − 8
= (3.0.6)
(p + 2)2
2(p2 − 2)
= (3.0.7)
(p + 2)2

If p ∈ A then p2 < 2. Using Eq. 3.0.4 and Eq. 3.0.5, we get


q ∈ A such that q > p. This shows that A has no largest number.
Similarly, if p ∈ B then we get a rational number q ∈ B such that
q < p. This shows that A has no largest number.

The above discussion shows that neither A has a largest number


nor B has a smallest number. Thus, the rational number system has
41

certain gaps in spite of the fact that between any two rationals p and
q there is a rational r = p+q
2
such that p < r < q. This prompts
us to introduce a number system which fills these gaps.
To begin axiomatic construction of real numbers, we shall in-
troduce the general concepts of ordered sets.

3.1 Ordered sets


Definition 3.1.1. Let S be a non-empty set. Then a relation <
on S is said to be an order on S if it satisfies the following two
properties:
(i) If x, y ∈ S then either statement x < y or x = y or y < x
holds.

(ii) If x, y, z ∈ S and if x < y, y < z then x < z.


We read the statement x < y as ”x is less than y” or ”x is
smaller than y” or ”x precedes y”. We also write y > x in place
of x < y. We use the notation ”x ≤ y” for the statement ”x < y
or x = y”. A set S equipped with order relation < is termed as
the ordered set. In other words, if < is an order on S, then the
ordered pair (S, <) is called an ordered set. The set of rationals is
an ordered set with respect to order < defined by x < y if (y − x)
is a positive rational number.

3.1.1 Lower Bounds, Upper Bounds, Supre-


mum and Infimum
Definition 3.1.2. Let (S, <) be an ordered set and A be a subset
of S. An element u ∈ S is called an upper bound of A if
a ≤ u, ∀a ∈ A. If A has an upper bound then it is said to be
bounded above. If it does not have an upper bound then it is
called unbounded above. Similarly, an element l ∈ S is called
42

a lower bound of A if l ≤ a, ∀a ∈ A. If A has a lower bound


then it is said to be bounded below. If it does not have a lower
bound then it is called unbounded below.

Definition 3.1.3. Let (S, <) be an ordered set and A be a subset


of S. An element u0 ∈ S is called a supremum or least upper
bound of A if it satisfies the following two properties:

(i). u0 is an upper bound of A.

(ii). if x < u0 then x is not an upper bound of A;i.e, no x ∈ S


less than u0 is an upper bound of A.

Similarly, An element l0 ∈ S is called a infimum or greatest


lower bound of A if it satisfies the following two properties:

(i). l0 is a lower bound of A.

(ii). if x > l0 then x is not a lower bound of A;i.e, no x ∈ S


greater than l0 is a lower bound of A.

The supremum and infimum of A is denoted by sup A (l.u.b of A)


and inf A (g.l.b of A) respectively. It is noted that the notions
defined as above may or may not exist.

Example 3.1.4. From Example 3.0.2, we see that B is the set


of all upper bounds of A and A is the set of all lower bounds of
B. Since A has no largest element and B has no least elements
therefore inf B and sup A do not exist respectively.

Example 3.1.5. Consider A = { n1 n ∈ N} ⊂ Q. Then sup A


and inf A both exist. But sup A = 1 ∈ A while inf A = 0 ∈
/ A.

Exercise 3.1.1. Prove that the supremum of the set A = {x ∈


Q |x < 0} exists but not belongs to A. Similarly, construct a set
B whose infimum exists but not belongs to B.
43

Exercise 3.1.2. Prove that the supremum of the set A = {x ∈


Q |x ≤ 0} exists but it belongs to A. Similarly, construct a set B
whose infimum exists but it belongs to B.

Proposition 3.1.6. Supremum (infimum) of the subset A is unique


provided it exists.

Proof: Suppose sup A exists. Let u1 and u2 be any two


supremums of A. Then, u1 and u2 are upper bonds of A. Since
u2 is least upper bounds therefore u2 ≤ u1 . Similarly, u1 ≤ u2 .
Combining these two, we get u2 = u1 . Similarly one can prove
that infimum of the subset A is unique provided it exists. ✷

Definition 3.1.7. (Least upper bound property or Com-


pleteness Axiom). An ordered set S is said to have least upper
bound property if every non-empty subset A of S which is bounded
above has a supremum in S.

Definition 3.1.8. (Greatest lower bound property). An or-


dered set S is said to have greatest lower bound property if every
non-empty subset A of S which is bounded below has an infimum
in S.

Example 3.1.9. The ordered set of rational numbers does not


have least upper bound property.

We now see that these two properties are equivalent.

Theorem 3.1.10. An ordered set has the least upper bound prop-
erty if and only if it has the greatest lower bound property.

Proof: Let S be an ordered set with the least upper bound


property. Let B be a non empty subset which is bounded below.
Consider the set L of all lower bounds of B; i.e,

L = {x ∈ S | x ≤ b, ∀b ∈ B}.
44

Clearly L is non-empty and every element of B is an upper bound


of L. Thus, L is bounded above. By least upper bound property
of S, sup L exists in S, say u0 = sup L. If x < u0 , then x is
not an upper bound of L and so x ∈ / B. Since u0 ≤ b for every
b ∈ B.Thus u0 ∈ L; that is, u0 is a lower bound of B. If u0 < x
then x is not a lower bound of B (because u0 is an upper bound
of L). Thus u0 is a lower bound of B such that no x > u0 is a
lower bound of B; i.e, u0 = inf B. ✷

3.2 Fields
Definition 3.2.1. A field is a non empty set F together with two
binary operations + and ·, called ”addition” and ”multiplication”
respectively, which satisfy the following properties termed as ”field
axioms”:

(i) (Commutative law) x + y = y + x and x · y = y · x for all


x, y ∈ F .

(ii) (Associative law) x + (y + z) = (x + y) + z and (x · y) · z =


x · (y · z) for all x, y, z ∈ F

(iii) F contains an element 0 such that 0 + x = x + 0 = x for


every x ∈ F .

(iv) F contains an element 1 6= 0 such that 1 · x = x · 1 = x for


every x ∈ F .

(v) To every x ∈ F there exists unique element y ∈ F such that


x + y = 0 = y + x. The element y is denoted by −x.

(vi) To every x ∈ F \ 0 there exists unique element y ∈ F such


that x · y = 1 = y · x. The element y is denoted by x1 .
45

(vii) (Distributive law) x · (y + z) = x · y + x · z for all x, y, z ∈ F .

We usually write xy, x − y, xy , x + y + z, xyz, x2 , 3x etc. in


place of x · y, x + (−y), x · y1 , (x + y) + z, (xy)z, xx, x + x + x
respectively.

Proposition 3.2.2. From field axioms, we have the following


statements.

(a) If x + y = x + z then y = z.

(b) If x + y = x then y = 0.

(c) If x + y = 0 then y = −x.

(d) −(−x) = x.

(e) If x 6= 0 and xy = xz then y = z.

(f ) x 6= 0 and xy = x then y = 1.

(g) x 6= 0 and xy = 1 then y = x1 .


1
(h) x 6= 0 then ( x1 )
= x.

(i) 0x = 0 for every x ∈ F .

(j) If x 6= 0 and y 6= 0 then xy 6= 0.

(k) (−x)y = −xy = x(−y) and so (−x)(−y) = xy.

Definition 3.2.3. An ordered field is a field F which is also an


ordered set, such that

(i) If x, y, z ∈ F with y < z then x + y < x + z.

(ii) If x, y ∈ F with x > 0 and y > 0 then xy > 0.


46

If x > 0, we call x positive and if x < 0 then we call x is


negative. Observe that the set Q of all rational numbers is an
orderd field.

Proposition 3.2.4. In an ordered field F , we have:

(i) If x > 0 then −x < 0.

(ii) If x > 0 and y < z then xy < xz.

(iii) If x < 0 and y < z then xy > xz.

(iv) If x 6= 0 then x2 > 0.


1
(v) If 0 < x < y then 0 < y
< x1 .

Proof: If x > 0 then x + (−x) > 0 + (−x). Thus 0 > −x,


that is; −x < 0. This proves (i).
Next let y < z. Then z − y > 0 and so x(z − y) > 0. Thus
xy < xz. This proves (ii).
Next, if x < 0 and y < z. Then −x > 0 and so (−x)(z−y) > 0.
This proves xy > xz, which is (iii).
Now, suppose x 6= 0. If x > 0 then x2 > 0. If x < 0 then
−x > 0 and so (−x)2 > 0. Thus x2 > 0 for all x 6= 0. This proves
(iv). In particular 1 = 12 > 0.
Next, let 0 < x < y. Observe that if v ≤ 0 then vx ≤ 0. Thus
x(1/x) = 1 > 0 gives 1/x > 0. Similarly 1/y > 0 for y > 0. Since
0 < x < y therefore (1/x)(1/y) > 0 and so 0 < x(1/x)(1/y) <
y(1/x)(1/y). This gives 0 < 1/y < 1/x. This proves the last
result. ✷

3.3 Axiom of Real number system


Axiom: There exists an ordered field R which satisfies complete-
ness axiom. The elements of R are called real numbers. The
47

construction of Real number system is given in Appendix on page


17 [1].

Theorem 3.3.1. Let a and b be any two real numbers. If a ≤ b+ǫ


for every ǫ ≥ 0 then a ≤ b.

Proof: Suppose that the result is not true. Then b < a.


Choose ǫ = a−b
2
. Then

a−b
b+ǫ = b+
2
b+a
=
2
a+a
< =a
2
which is a contradiction. Thus, a ≤ b. ✷

3.4 Geometric representation of real


numbers
The real numbers are represented geometrically as points on a
line called the real line or real axis. A point is selected to represent
0 and another to represent 1. This choice determines the scale.
Under an appropriate set of axioms for Euclidean Geometry, each
point on the real line corresponds to one and only one real number
and conversely each real number is represented by one and only
one point on the line. The order relation has a simple geometric
interpretation. If x < y then the point x lies left to the point y.
The positive real numbers lie right to 0 and negative real numbers
lie left to 0. If a < b, a point x satisfies a < x < b if and only if x
is between a and b.
48

3.5 Intervals
The set of all real numbers between a and b is called an Interval
Let a, b be any two real numbers such that a < b. Then the open
interval (a, b) is defined by the set {x ∈ R|a < x < b}. The closed
interval [a, b] is defined by the set {x ∈ R|a ≤ x ≤ b}. The left
open interval (a, b] is defined by the set {x ∈ R|a < x ≤ b} and the
right open interval [a, b) is defined by the set {x ∈ R|a ≤ x < b}
respectively. These intervals are also termed as half-open intervals.
The real numbers a and b of intervals are called end points. All
such intervals defined as above are also called finite intervals. The
subsets of real numbers of the form {x ∈ R|a < x}, {x ∈ R|x < b},
{x ∈ R|a ≤ x} and {x ∈ R|x ≤ b}are called infinite intervals of
real numbers. These infinite intervals are denoted by (a, +∞),
(−∞, b), [a, +∞) and (−∞, b] respectively. The real line R is
sometimes denoted by (−∞, +∞). Note that +∞ and −∞
are just symbols. They are not to be considered as real
numbers. The singleton set {a} is considered as a degenerate
closed interval [a, a].

3.6 Integers, Rational numbers and Ir-


rational numbers
In this section we describe the set of integers as a special subset
of R. To define it, first we define inductive set.
Definition 3.6.1. A subset I ⊂ R is called an inductive set if it
satisfies the following two properties:
(i). 1 ∈ I.
(ii). if x ∈ I then x + 1 ∈ I
Observe that R and R+ (the set of all positive real numbers)
are inductive sets.
49

Definition 3.6.2. A real number is called a positive integer (nat-


ural numbers) if it belongs to every inductive set. The set of all
positive integers is denoted by Z+ . Observe that this is the smallest
inductive set. The negative of positive integers are called negative
integers. The positive integers together with negative integers and
zero form the set Z of all integers.
Definition 3.6.3. The quotient a/b of integers a and b with b 6=
0, are called rational numbers. The set of rational numbers is
denoted by Q. Clearly Z ⊂ Q. Observe that a, b ∈ Q implies
that a + b, ab ∈ Q. The real numbers which are not rational are
called irrational numbers. For example, e, ex , π, sqrt2 and sqrtn,
where n is not a perfect square. Note that the composition of two
irrationals may or may not be an irrational number.
Proposition 3.6.4. There are infinitely many rationals between
any two given rationals.
Proof: Let a, b ∈ Q. Then a < a+b 2
< b, where a+b
2
∈ Q.
Repeating this fact, we have the result. ✷

Proposition 3.6.5. (Approximation Property) Let S be a


non empty set of real numbers with a supremum (infimum), say
u0 . Then for every a < u0 (u0 < a) there exists s ∈ S such that
a < s ≤ u0 (u0 ≤ s < a).
Proof: Since u0 = sup S. Therefore s ≤ u0 for all s. Suppose,
if possible that s ≤ a for all s ∈ S. Then u0 ≤ a. But a < u0 ,
therefore we get a contradiction. Thus a < s for some s ∈ S and
so a < s ≤ u0 for some s ∈ S. ✷

Proposition 3.6.6. (Additive property) Let A and B be any


two subsets of R. Define C = {a + b |a ∈ A, b ∈ B}. If both A and
B have supremums then sup C exists and is given by sup C =
sup A + sup B. Similar result holds for infimum.
50

Proof: Let a = supA and b = sup B. Clearly x+y ≤ a+b for


all x ∈ A and y ∈ B. Thus C is bounded above. By completeness
axiom C has a supremum in R, say c. Clearly c ≤ a + b as a + b is
an upper bound of C. Next, let ǫ be a positive real number. Then
a − ǫ and b − ǫ will not be upper bounds of A and B respectively.
Thus there exists x ∈ A and y in B such that a − ǫ < x and
b − ǫ < y respectively. By property of real numbers, we get
a + b − 2ǫ < x + y ≤ c as c is an upper bound of A + B. That is,
a + b ≤ c + 2ǫ, where ǫ is arbitrary. Thus a + b ≤ c. But c ≤ a + b
and hence c = a + b. This proves the result. ✷

Exercise 3.6.1. Let A, B be any two subsets of R such that


a ≤ b for every a ∈ A and b ∈ B. Then sup A ≤ sup B and
inf A ≤ inf B provided A has infimum and B has a supremum.
If C = {xy |x ∈ A, y ∈ B} and both A, B has supremums then
prove that C has a supremum. Indeed sup C = sup A · sup B.

Exercise 3.6.2. Let A, B be any two subset of real numbers such


that A ⊂ B. If B is bounded then prove that A is bounded. Also
prove that supA ≤ sup B and inf B ≤ inf A.

Proposition 3.6.7. The set of positive integers Z+ (natural num-


bers) is unbounded above.

Proof: Suppose that the result is not true. Then it is bounded


above. By completeness property it has a supremum, say a =
sup Z+ . Then a − 1 will not be an upper bound and so a − 1 < n
for some n ∈ Z+ . This gives a < n + 1 where n + 1 ∈ Z+ . This
is a contradiction as a is an upper bound of Z+ . Thus Z+ is
unbounded above. ✷

Proposition 3.6.8. For every real number x there exists a posi-


tive integer n such that n > x.
51

Proof: Let x be any real number. Suppose n ≤ x for every


natural number n, then Z+ will be bounded above. But this is a
contradiction because Z+ is unbounded above. Thus there exists
n ∈ Z+ such that n > x. ✷

Theorem 3.6.9. (Achimedean Property) If x, y ∈ R with


x > 0 then there exists a natural number(positive integer) n such
that nx > y.
Proof: Using Proposition3.6.8, we have a natural number n
such that n > xy . This gives nx > y. ✷

Theorem 3.6.10. (Dense property of Q) Between any two


given real numbers there is a rational one.
Proof: Let x, y be any two real numbers such that x < y.
Then y − x > 0. By archimedean property (Theorem 3.6.9), we
get a natural number n such that n(y − x) > 1, that is;

ny > nx + 1 (3.6.3)

Next, by archimedean property of R, we get natural numbers m1


and m2 such that m1 > nx and m2 > −nx, that is; −m2 < nx.
Combining these two, we have −m2 < nx < m1 . Choose an
integer m with −m2 ≤ m ≤ m1 such that

(m − 1) ≤ nx < m or nx < m < nx + 1 (3.6.4)

By Eqns. 3.6.3, 3.6.4, we have nx < m < ny. This gives


m
x< <y
n
m
where n
is a rational number. ✷
Repeating this result, we find that there are infinitely many ratio-
nals between any two given real numbers.
52

3.7 Decimal representations of real num-


bers
A real number r of the form
a1 a2 an
r = a0 + + 2 + . . . + n = a0 .a1 a2 . . . an
10 10 10
is called a finite decimal representation of r, where a0 is non-
negative integer and a1 , a2 , . . . , an are integers such that 0 ≤ ai ≤
9 for all i = 1, 2, . . . , n. Observe that
a0 a1 a2 . . . an
r= .
10n
Thus such real numbers are rational numbers.

Theorem 3.7.1. (Finite decimal approximations to real


numbers) If x ∈ R with x ≥ 0. Then for every natural number
n there is a finite decimal rn = a0 .a1 a2 . . . an such that
1
rn ≤ x ≤ rn + .
10n
Proof: Let S = {n ∈ Z|0 ≤ n ≤ x}. Clearly it is non empty
and bounded above. Thus it has a supremum, say a0 = supS.
Observe that a0 ∈ S and a non-negative integer. Such number
is called the greatest integer in x. We denote it by [x]. Thus
a0 = [x]. Clearly a0 ≤ x < a0 + 1. Next, let a1 = [10x − 10a0 ].
Since 0 ≤ 10x − 10a0 = 10(x − a0 ) < 10, therefore 0 ≤ a1 ≤ 9 and

a1 ≤ 10x − 10a0 < a1 + 1.

Thus we have a largest integer a1 with 0 ≤ a1 ≤ 9 satisfying


a1 a1 + 1
a0 + ≤ x < a0 + .
10 10
53

Now, having chosen a0 , a1 , . . . , an−1 with 0 ≤ ai ≤ 9, let an be


the greatest integer satisfying
a1 an a1 an + 1
a0 + + . . . + n ≤ x < a0 + + ... + (. 3.7.1)
10 10 10 10n
Then 0 ≤ an ≤ 9 and we have
1
rn ≤ x ≤ rn + .
10n
where rn = a0 .a1 a2 . . . an . This completes the proof. ✷
The decimal representation x = a0 .a1 a2 . . . an . . . where 0 ≤ ai ≤
9 and satisfying Equation 3.7.1 is called the infinite decimal rep-
resentation of real number x.

3.8 Absolute values and Triangle In-


equality
Definition 3.8.1. The absolute value |x| of real number x is de-
fined by 
 x if x > 0
|x| = 0 if x = 0
−x if x < 0

Observe that −|x| ≤ x ≤ |x| for all real x.


Proposition 3.8.2. Let a ≥ 0. Then |x| ≤ a if and only if
−a ≤ x ≤ a.
Proof: Assume that |x| ≤ a. If x > 0 then |x| ≤ a gives
x ≤ a and so −a ≤ −x < 0 < x ≤ a. Next, if x ≤ 0 then |x| ≤ a
gives −x ≤ a and so −a ≤ x ≤ 0 ≤ a, i.e; −a ≤ x ≤ a. Thus
−a ≤ x ≤ a for all x.
Conversely assume that −a ≤ x ≤ a. If x ≥ 0 then |x| = x ≤
a. If x < 0 then |x| = −x ≤ a. Thus proved. ✷
54

Corollary 3.8.3. Let ǫ > 0 and x, a ∈ R. Then |x − a| < ǫ if


and only if a − ǫ < x < a + ǫ.
Theorem 3.8.4. (Triangle inequality) Let x, y ∈ R. Then
|x + y| ≤ |x| + |y|.
Proof: Since −|x| ≤ x ≤ |x| and −|y| ≤ y ≤ |y|, therefore
−(|x| + |y|) ≤ x + y ≤ |x| + |y| and so |x + y| ≤ |x| + |y|. This
proves the theorem. ✷
Putting x = a − c and y = c − b in this inequality, we get
|a − b| ≤ |a − c| + |c − b|.
If we put x = a + b and y = −b in the triangle inequality, we get
|a| ≤ |a + b| + |b|. This implies |a| − |b| ≤ |a + b|. Interchanging
a and b we get |b| − |a| = −(|a| − |b|) ≤ |a + b|. Combining these
two we get
||a| − |b|| ≤ |a + b|
where a and b are arbitrary real numbers.

3.9 Extended Real Number System


A set R∗ = R∪{−∞, +∞} obtained from R by adjoining two sym-
bols denoted by +∞ and −∞, is called an extended real number
system if it satisfies following properties
1. if x ∈ R then x + ∞ = ∞ = x − (−∞), x + (−∞) = −∞ =
x x
x − ∞ and +∞ = 0 = −∞ .
2. If x > 0 then x(+∞) = +∞ and x(−∞) = −∞.
3. If x < 0 then x(+∞) = −∞ and x(−∞) = +∞.
4. (+∞)+(+∞) = +∞ = (+∞)(+∞) = (−∞)(−∞), (−∞)+
(−∞) = −∞ = (+∞)(−∞).
5. If x ∈ R then −∞ < x < +∞. We denote it by [−∞, +∞].
55

3.10 Similar (Equinumerous) Sets,and


Countability
Two sets A and B are said to be equinumerous or similar sets if
there is bijective function from A to B. If A and B are similar
then we write A ∼ B and read as A is similar to B. Observe that
if A ∼ B then B ∼ A. Also if A ∼ B and B ∼ C then A ∼ C.
A set S is said to be finite set containing n elements
if S ∼ {1, 2, . . . , n}. The number n is called cardinal number
or cardinality of S. The empty set is also considered as finite
set and its cardinal number is defined to be 0. Observe that
{1, 2, . . . , n} ∼ {1, 2, . . . , m} if and only if m = n. A set which is
not finite is called an infinite set. A set S is called a countable set
if it is either finite or S ∼ Z+ (countably infinite). A set which
is not countable is called an uncountable set. The cardinality of a
countably infinite set is said to have cardinal number N0 (read as
aleph nought)

Proposition 3.10.1. Every subset of a countable set is countable.

Proof: Let S be a countable set and A be any subset of S.


If A is finite then nothing to do. Suppose that A is countably
infinite then so is S, that is; S ∼ Z+ under bijective map f . Then
S = {f1 , f2 , . . .}. Define a function k : N → N as follows:
k(1) is the smallest natural number m such that fm ∈ A. Suppose
k(1), k(2), . . . k(n − 1) have been defined, then we define k(n) is
the smallest natural number m such that fm ∈ A where k(n) >
k(n − 1). This process determines a bijective function k with
k(1) < k(2) < . . . < k(n − 1) < k(n) . . . such that f ok ∈ A. Since
domain of f ok is N, range is A and f ok is bijective, therefore A
is countable.✷

Theorem 3.10.2. The set R of all real numbers is uncountable.


56

Proof: To prove this it is sufficient to prove that the open


interval (0, 1) is uncountable. Suppose that (0, 1) is countable.
Then (0, 1) = {f1 , f2 , . . .}, where f : N → (0, 1) is a bijective
map. For each n, consider the decimal representation of fn as :

0.gn,1 gn,2 gn,3 . . .

where 0 ≤ gn,i ≤ 9 for all i ∈ N. Construct a real number x as

0.v1 v2 . . . vn . . .

where

6 1
1 if gi,i =
vi =
2 if gi,i = 1

Then x 6= fn for all n because x differs from f1 in first place,


differs from f2 in second place, .., differs from fn in nth place.
But x ∈ (0, 1). This is a contradiction. Thus (0, 1) is uncountable
and so is R. ✷

Corollary 3.10.3. Each interval is uncountable.


57

Facts sheet
• Z, Q, R and N are unbounded sets.

• Let k ∈ R. Any subset of the form A = {x ∈ R|x ≥ k}


is bounded below but unbounded above. In this case k =
inf A. Similarly any subset of the form B = {x ∈ R|x ≤ k}
is bounded above but unbounded below. In this case k =
sup B.

• Supremum and infimum of a subset of an ordered set may


or may not exists. But if they exist then they are unique.

• Supremum and infimum of a subset A of an ordered set S


may or may not be an element of A.

• If l = inf A then we have (i) l ≤ a, ∀a ∈ A, (ii) for every


ǫ > 0, l + ǫ is not a lower bound, that is; a < l + ǫ for some
a ∈ A, and (iii) l is unique, i.e; if l1 = inf A and l2 = inf A
then l1 = l2 .

• If u = sup A then we have (i). u ≥ a, ∀a ∈ A, (ii). for every


ǫ > 0, u − ǫ is not an upper bound, that is; u − ǫ < a for
some a ∈ A, and (iii). u is unique, i.e; if u1 = sup A and
u2 = sup A then u1 = u2 .

• If l is a lower bound of A and l ∈ A then l = inf A.

• If u is an upper bound of A and u ∈ A then u = sup A.

• Every finite A set is bounded. Its smallest element is infi-


mum of A and its largest element is supremum of A.

• sup A = inf A if and only if A is singleton.

• Every infinite subset of R need not be unbounded. For


example, { n1 |n ∈ N}.
58

• There are infinitely many rationals between any two given


rationals.

• (Approximation Property). Let S be a non empty set


of real numbers with a supremum (infimum), say u0 . Then
for every a < u0 (u0 < a) there exists s ∈ S such that
a < s ≤ u0 (u0 ≤ s < a).

• For every real number x there exists a positive integer n


such that n > x.

• (Archimedean Property). If x, y ∈ R with x > 0 then


there exists a natural number(positive integer) n such that
nx > y. Indeed such n are infinitely many. In particular
there exists a natural number n such that nx > 1, that is;
1/n < x for some n. Also such n are infinitely many.

• Between any two given real numbers there is a rational one.

• For each x ∈ R there is a unique integer n such that x − 1 <


n ≤ x, here n = [x] and so n ≤ x < n + 1.

• Every superset of an unbounded subset is unbounded.

• Every subset of a bounded set is bounded.

• (Least upper bound property or Completeness Ax-


iom). An ordered set S is said to have least upper bound
property if every non-empty subset A of S which is bounded
above has a supremum in S.

• (Greatest lower bound property). An ordered set S


is said to have greatest lower bound property if every non-
empty subset A of S which is bounded below has an infimum
in S.
59

• The ordered set of rational numbers does not have least


upper bound property. Thus it is an ordered field but not
complete.

• An ordered set has the least upper bound property if and


only if it has the greatest lower bound property.

• R is a complete ordered field (by axiomatic construction)


but Q is an ordered field which is not complete. Observe
that A = {x ∈ Q|x2 < 3} is a non
√ empty subset of Q which
is bounded above but sup A = 3 ∈ / Q.

Exercise 3.10.1. Find lower bounds, upper bounds, supremum


and infimum of each of the following subsets of R:
n 1
1. { n+1 |n ∈ N} = {1 − n+1
|n ∈ N}.
n 1
2. {− n+1 |n ∈ N} = {−1 + n+1
|n ∈ N}.
(−1)n
3. {1 + n
|n ∈ N}.
1
4. {l + n
|n ∈ N} where l ∈ R.

5. {(−1)n |n ∈ N}.

6. {(−2)n |n ∈ N}.

7. {xn |n ∈ N}, where 0 < x < 1. Also find for |x| > 1.

Exercise 3.10.2. Find lower bounds, upper bounds, supremum


and infimum of each of the following subsets of real numbers:

1. All numbers of the form 2−p + 3−q + 5−r , where p, q and r


take all positive integral values.

2. {x ∈ R |3x2 − 10x + 3 < 0}.


60

3. {x ∈ R |(x − a)(x − b)(x − c)(x − d) < 0}, where a < b <


c < d.

Exercise 3.10.3. Given x > 0 and an integer k ≥ 2. Let a0 = [x]


and assuming a0 , a1 , . . . an−1 have been defined, let an = [x] such
that
a1 an
a0 + + . . . + n ≤ x.
k k
(a). Prove that 0 ≤ ai ≤ k − 1 for all 1 ≤ i ≤ n.
(b). Let rn = a0 + a1 k −1 + a2 k −2 + . . . + an k −n and show that x is
the supremum of all rational numbers r1 , r2 , . . . .. Indeed it provide
a representation in scale k. For k = 2, the representation is
binary system and for k = 10 the representation is the
decimal representation system.
Chapter 4

Basic Topology of R

This chapter is devoted to study subsets of R and its points.

4.1 Definitions
Definition 4.1.1. Let A ⊂ R and x ∈ R.

1. A subset A of R is called a neighbourhood (nbd) of a point


x if there exists a positive real number ǫ such that x ∈ (x −
ǫ, x + ǫ) ⊂ A, i.e; there is an infinitesimal line segment
with mid point x is contained in A. Open interval
(x − ǫ, x + ǫ) is called ǫ-nbd of the point x. For example,
every open interval is a nbd of each of its points. But it is
not a nbd of its end points.

2. A point x ∈ R is called a limit point of A if every nbd N of x


contains a point of A other than x, i.e; (N ∩A)−{x} 6= ∅ for
every nbd N of x. The set of all limit points of A is denoted
by D(A) and is called a derived set. For example, every point
of an open interval is a limit point of that interval.

3. If x ∈ A and x is not a limit point of A then x is called an

61
62

isolated point of A. For example, each natural number is an


isolated point.

4. A is called a closed set if every limit point of A belongs to


A, i.e; D(A) ⊂ A. For example, N is a closed set because
D(N) = ∅ ⊂ N.

5. A point x ∈ R is called an interior point of A if there exists a


nbd N of x such that x ∈ N ⊂ A. Observe that each interior
point is a limit poin. The set of all interior points of A is
denoted by i(A) or A0 or Int A. For example, every point
of an open interval is an interior point of that interval.

6. A is called an open set if i(A) = A, i.e; every point of A is


an interior point of A . In other words, A is a nbd of each
of its points. For example, every open interval is an open
set.

7. A is called perfect if A is closed and every point of A is a


limit point of A. For example, every closed interval of R
and R are perfect but Q is not perfect.

8. A is called dense in R if every point of R is a limit point of


A or a point of A. For example, Q is dense in R.
Example 4.1.2. Since every ǫ -nbd of a point x ∈ Q contains
infinitely many rationals as well as irrationals except x, therefore
no ǫ -nbd of a point x ∈ Q is a subset of Q. Thus Q is not a nbd of
any of its points. In other words no point of Q is an interior point.
Thus Q is not an open set. Since every ǫ-nbd of a real number
x contains infinitely many points of Q other than x. Thus every
real number is a limit point of Q and so D(Q) = R 6⊂ Q. This
shows that Q is not closed.
Example 4.1.3. No finite set is a nbd of its points. Indeed their
derived sets are empty (i.e; no point of a finite set is a limit point).
63

Thus a finite set is a closed set. It also follows that every point
of a finite set is an isolated point.
Example 4.1.4. Every interval is a nbd of each of its points
except end points. Thus each point are its interior points except
end points. Also observe that each point of an interval are its
limit points. Thus no point of an interval are isolated points.

4.2 Some results on neighbourhoods


Since ∅ contains no points therefore it is a nbd of each of its points,
i.e; ∅0 = ∅ and so ∅ is an open set. Indeed D(∅) = ∅ ⊂ ∅ and so
it is open as well as closed set.
Proposition 4.2.1. If N is a nbd of x, then x ∈ N .
Proof: By definition of nbd of x, x ∈ (x − ǫ, x + ǫ) ⊂ N for
some ǫ-nbd. Thus x ∈ N . ✷ Thus if x ∈ / N then N is not a
nbd of x.
Proposition 4.2.2. Every superset M of a nbd N of x is a nbd
of x.
Proof: If N ⊂ M and N is a nbd of x then x ∈ (x−ǫ, x+ǫ) ⊂
N for some ǫ-nbd. But N ⊂ M . Thus we have x ∈ (x−ǫ, x+ǫ) ⊂
M which shows that M is a nbd of x. ✷
This motivates us to have the following:
Proposition 4.2.3. Union of a family of nbds of x is a nbd of x.
But subset L of a nbd N of x need not be a nbd of x. For,
N = (0, 1) is a nbd of 0.5 but M = (0.56, 1) ⊂ N is not a nbd of
0.5 because 0.5 ∈/ M (cf. Proposition 4.2.1).
Proposition 4.2.4. Finite intersection of nbds of x is a nbd of
x.
64

Proof: Let {Ni |1 ≤ i ≤ n} be a finite family of nbds of x.


Let N = ∩ni=1 Ni . Then x ∈ N = ∩ni=1 Ni . Since each Ni is a nbd
of x, therefore we have an ǫ-nbd (x − ǫi , x + ǫi ) ⊂ Ni . Choose
ǫ = min{ǫi |1 ≤ i ≤ n}. Then (x − ǫ, x + ǫ) ⊂ N . Thus proved. ✷

Remark 4.2.5. The result is not true for infinite family. For,
consider the infinite family {(−1/n, 1/n)|n ∈ N} of nbds of 0,
then ∩∞
n=1 (−1/n, 1/n) = {0}, which is not a nbd of 0.

4.3 Interior points and open sets


Proposition 4.3.1. If x ∈ R is an interior point of A ⊂ R. Then
x ∈ A. In particular A0 ⊂ A.
Proof: Proof follows from the definiton of interior point and
Proposition 4.2.1. ✷

Theorem 4.3.2. Every open interval I is an open set. Thus


I 0 = I.
Proof: Let I = (a, b) be an open interval of R. Let x ∈ I.
Choose ǫ < min{x − a, b − x}. Then ǫ < x − a and b − x and so
a < x − ǫ < x + ǫ < b. This gives x ∈ (x − ǫ, x + ǫ) ⊂ I. Thus x
is an interior point for all x and so I is open. ✷

Theorem 4.3.3. Union of an arbitrary collection of open sets


(nbds) is an open set (nbd).
Proof: Let {Oα |α ∈ δ} be any collection of open set. Let
O = ∪α∈δ Oα and x ∈ O. Then x ∈ Oα for some α. But Oα is
open and so there is an open interval Nα = (x − ǫα , x + ǫα ) such
that x ∈ Nα ⊂ Oα ⊂ O. Thus every point of O is an interior point
(O is a nbd of each of its points). Consequently O is open.✷
65

Example 4.3.4. Union of two open sets is open. Finite union of


open sets is open.

Theorem 4.3.5. Intersection of two open sets is open. Indeed


intersection of a finite collection of open sets is open.

Proof: Let O1 and O2 be any two open sets. Let O = O1 ∩O2 .


We claim that every point of O is an interior points. If O is empty
then nothing to prove. Assume that O is non-empty. Let x ∈ O.
Then x ∈ O1 and x ∈ O2 . But O1 , O2 are open sets. Therefore we
have open intervals (x − ǫ1 , x + ǫ1 ) and (x − ǫ2 , x + ǫ2 ) such that
(x − ǫ1 , x + ǫ1 ) ⊂ O1 and (x − ǫ2 , x + ǫ2 ) ⊂ O2 respectively, where
ǫ1 , ǫ2 > 0. Chose ǫ = min{ǫ1 , ǫ2 }. Then ǫ ≤ ǫi and −ǫi ≤ −ǫ for
all i = 1, 2. Thus
−ǫi ≤ −ǫ < ǫ ≤ ǫi
which gives (x − ǫ, x + ǫ) ⊆ (x − ǫi , x + ǫi ) ⊂ Oi for i = 1, 2. Thus
(x − ǫ, x + ǫ) ⊂ O. This proves the result. ✷

Remark 4.3.6. Intersection of an arbitrary collection of sets need


not be open, e.g;

∩∞
n=1 (−1/n, 1/n) = {0} (4.3.1)

is a closed set where as each open interval (− n1 , n1 ) is an open


set.

Exercise 4.3.2. 1. ∅0 = ∅,

2. A0 ⊆ A,

3. A ⊆ B then A0 ⊆ B 0

4. Q0 = ∅, in other words Q is not a nbd of any real number.


But each real number is a limit point of Q,
66

5. If A is finite then A0 = ∅, but converse is not true. For


example, A = {1/n|n ∈ N}, N, Z and Q are infinite sets but
A0 = ∅.

6. A0 ∪ B 0 ⊆ (A ∪ B)0 .
Hint: A0 ⊂ A, B 0 ⊂ B gives A0 ∪ B 0 ⊂ A ∪ B. But A0 ∪
B 0 is open so A0 ∪ B 0 = (A0 ∪ B 0 )0 ⊂ (A ∪ B)0 .

7. A0 ∩ B 0 = (A ∩ B)0 .
Hint: A0 ⊂ A, B 0 ⊂ B gives A0 ∩ B 0 ⊂ A ∩ B. But A0 ∩
B 0 is open so A0 ∩ B 0 = (A0 ∩ B 0 )0 ⊂ (A ∩ B)0 . Now
A ∩ B ⊂ A, B gives (A ∩ B)0 ⊂ A0 , B 0 and so (A ∩ B)0 ⊂
A0 ∩ B 0 .

Proposition 4.3.7. A0 is an open subset of R. Indeed it is the


largest open subset of A.

Proof: Let x ∈ A0 . Then A is a nbd of x so x ∈ (x − ǫx , x +


ǫx ) ⊂ A, for some ǫx . But each point of (x − ǫx , x + ǫx ) is an
interior point. Thus (x − ǫx , x + ǫx ) ⊂ A0 for each x. Next, let B
be an open subset of A with A0 ⊆ B. Then every point of B will
be an interior point of A and so B ⊆ A0 . Thus A0 = B. Thus
proved. ✷

4.3.1 Structures of open sets in R


Let S be an open subset of R. Then an open interval I is called a
component interval of S if I ⊂ S and there is no open interval
J ⊂ S which properly contains I. Thus a component interval S
is not a proper subset of any other open interval contained in S.

Theorem 4.3.8. Every point of a non-empty open subset S ⊂ R


belongs to one and only one component interval of S.
67

Proof: Let x ∈ S. If S is an open interval then nothing


to do. Suppose S is not an interval. Since S is open therefore
there is an open interval I such that x ∈ I ⊂ S(by definition
I = ǫ-nbd for some ǫ). Let ax = inf {a ∈ S|(a, x) ⊂ S} and
bx = sup{b ∈ S|(x, b) ⊂ S} and take Ix = (ax , bx ). Then Ix is an
open interval contained in S with x ∈ Ix . By definition of Ix it is
clear that there is no open interval J containing Ix contained in
S. Thus Ix is a component interval. If Jx is another component
interval then Ix ∪ Jx will be an open interval containing Ix and Jx .
Thus Ix = Ix ∪ Jx = Jx . This proves the uniqueness of component
interval. ✷

Theorem 4.3.9. (Representation theorem for open sets on


real line) Every non-empty open subset S of R is the union of
a countable collection of disjoint component intervals of S. (This
collection is unique)
Proof: If S is an interval then nothing to do. Suppose S is
not an open interval. By theorem 4.3.8, each x ∈ S is contained
in a unique component interval Ix of S. Thus S = ∪x∈S Ix . First
we claim that these component intervals are disjoint. For if Ix , Iy
be two component intervals with z ∈ Ix ∩ Iy . Then Ix ∪ Iy will
be an open interval containing Ix , Iy and so Ix = Iy . Let Λ be
an index set obtained by selecting one only one element from
each component intervals of S. Then T = {Ix |x ∈ Λ} will be
disjoint collection of component intervals. Let {x1 , x2 , . . . , xn . . .}
be a countable collection of rational numbers in S. Since each
component intervals Ix , x ∈ Λ contains infinitely many xn , but
among these xn there will be a unique xn with smallest index
n. This process defines a function f : T → N as f (Ix ) = n,
where xn ∈ Ix is a rational number with smallest index n. Now
f (Ix ) = f (Iy ) = xn gives xn ∈ Ix ∩ Iy and so Ix = Iy , i.e; f is
one-one. Thus T is countable.

68

4.4 Limit points, Closed sets and Per-


fect sets
From the defintion we observe the following:

Exercise 4.4.1. 1. The sets Z, N and finite sets have no limit


points, that is, there derived sets are empty. Thus, all of
these sets are closed sets but none of them is a perfect set
(no point is a limit point).

2. The set A = {1/n |n ∈ N} has limit point 0 ∈


/ A.

3. Every point of Q, R and R − Q (the set of all irrationals)


are limit points. Indeed D(Q) = D(R) = D(R − Q) = R.
Thus Q, R − Q and R are dense in R.

4. Every closed interval is a closed set. Since every point is a


limit point, therefore it is also a perfect set.

Proposition 4.4.1. A set X is closed if and only if its comple-


ment R − X is open. In particular, each closed interval [a, b] is a
closed set

Proof: Let X be closed and x ∈ R − X. Then x ∈ / X. Since


X is closed and hence x is not a limit point of X. Thus we have
a nbd N of x such that N ∩ X − {x} = ∅. But x ∈ / X and so
N ∩ X = ∅ which gives N ⊂ R − X. Thus N is an nbd of x.
But x ∈ R − X is arbitrary, hence R − X is open. Conversely
suppose that R − X is open and x ∈ / X. Then x ∈ (R − X).
But R − X is open and so there exists a nbd N of x such that
x ∈ N ⊆ (R − X). Thus N ∩ (R − X) \ {x} = ∅ and so x is not
a limit point of X. This gives that x ∈ D(X) ⇒ x ∈ X. Hence
X is a closed set. Next R − [a, b] = (−∞, a) ∪ (b, ∞) = open set,
thus [a, b] is closed.✷
69

By De Morgan’s law

R \ (∪ni=1 Fi ) = ∩ni=1 (R \ Fi ) and R \ (∩α∈∆ Fα ) = ∪α∈∆ (R \ Fα )

Using Theorems 4.3.3 and 4.3.5, we have

Proposition 4.4.2. Finite Union of closed sets is a closed set.


Arbitrary intersection of closed sets is a closed set.

Remark 4.4.3. Infinite union of closed sets need not be closed.


For, ∩∞n=1 [1/n, 1] = (0, 1] is not a closed set although each [1/n, 1]
is a closed set.

Exercise 4.4.2. Prove that

1. If A ⊂ B, then D(A) ⊂ D(B).

2. D(A) ⊂ A if and only if A is closed.

3. D(A ∩ B) ⊂ D(A) ∩ D(B) for all sets A and B. Equality


may not holds, for example, A = (1, 2) and B = (2, 3) then
D(A ∩ B) = ∅ but D(A) ∩ D(B) = {2}.

4. If A and B be any two subsets of R, then D(A ∪ B) =


D(A) ∪ D(B).
Hint: A, B ⊂ A ∪ B implies D(A) ∪ D(B) ⊂ D(A ∪ B).
Next, if x ∈
/ D(A) ∪ D(B) then x ∈/ D(A) and x ∈ / D(B).
Thus there is common nbd N of x such that N ∩ A = {x} =
N ∩ B and so N ∩ (A ∪ B) = {x}. Thus x ∈ / D(A ∪ B).
This proves that D(A ∪ B) ⊂ D(A) ∪ D(B).

Proposition 4.4.4. If x is a limit point of A then every nbd


contains infinitely many distinct points of A. Thus no finite set
has a limit point.
70

Proof: Suppose not. Then there is a ǫ-nbd (x − ǫ, x + ǫ)


containing only finitely many points a1 , a2 , . . . , an of A. Take
ǫ0 = (1/2)min{|x−ai | |i = 1, 2, . . . , n}, then ǫ0 -nbd (x−ǫ0 , x+ǫ0 )
contains no points of A and so x is not a limit point. This is a
contradiction. Thus proved. ✷

Proposition 4.4.5. Derived set D(A) of a subset A ⊂ R is a


closed set. In particular A ∪ D(A) is a closed set containing A.
Proof: To prove this we prove that D(A) has all its limit
points. Let x ∈ D(D(A)). Then x is a limit point of D(A). Then
every ǫ-nbd (x − ǫ, x + ǫ) of x contains a point y ∈ D(A) with
y 6= x. But y ∈ D(A), so every ǫ-nbd (y − ǫ, y + ǫ) of y contains
infinitely many points of A. This enforces us to say that ǫ-nbd
(x − ǫ, x + ǫ) of x contains infinitely many points of A and so x is
a limit point of A, i.e; x ∈ D(A). Thus D(D(A)) ⊂ D(A) and so
D(A) is a closed set. Next,
D(A ∪ D(A)) = D(A) ∪ D(D(A))
= D(A) ⊂ A ∪ D(A) as D(D(A)) ⊂ D(A).
Thus A ∪ D(A) is a closed set containing A. ✷
Indeed if B is a closed set containing A and B ⊂ A ∪ D(A),
then A ⊂ B gives A ∪ D(A) ⊂ B ∪ D(B) = B as D(B) ⊂ B.
Thus A ∪ D(A) = B.

Note: A ∪ D(A) is a smallest closed set containing A.


This set is denoted by A.
Theorem 4.4.6. (Bolzano-Weierstrass theorem) Every bounded
infinite subset of R has a limit point.
Proof: Let A be an infinite bounded subset of R. Let M =
sup A and m = inf A. Consider the set P defined by
P = {x ∈ R| x exceeds at most f inite members of A}.
71

Since m ≤ a ∀a ∈ A, therefore m ∈ P and so P is non-empty.


Since any real number x ≥ M exceeds infinitely many elements
of A, therefore x can not belong to P . Thus p < M for all p ∈ P
and hence P is bounded above. By completeness property of R,
P has a supremum say, l = sup P . Clearly l ≤ M . We claim that
l is a limit point of P . Let ǫ > 0 and take the ǫ-nbd (l − ǫ, l + ǫ).
Since l is supremum of P , therefore there exists p ∈ P such that
l − ǫ < p ≤ l < l + ǫ. Then p ∈ (l − ǫ, l + ǫ). Since l + ǫ ∈
/ P thus
it exceeds infinitely many points of A. Also p exceeds at most
finite members of A therefore ǫ-nbd (l − ǫ, l + ǫ) contains infinitely
many points of A. But ǫ is arbitrary, therefore l is a limit point
of A. Thus proved. ✷
72

QUESTIONS
1. An open set which is not an interval is
(a) (1, 2) ∪ [0, 1) (b) (1, 2) ∪ {3} (c) (1, 5) ∩ (3, 6) (d)
None

2. An interval which is an open set is ........ and which is not


an open set is ........

3. An interval which is a closed set is ......... and which is not


closed set is ........

4. An interval which is neither open nor closed is ..............

5. An interval, whose derived set is the same as that interval,


is .........

6. The derived set of finite set is....................(empty set/ non


empty set) and the derived set of bounded infinite set is..............
(empty set/ non empty set).

7. The derived set of infinite subset is ..........(necessarily/ not


necessarily) empty set.

8. A subset of real numbers is closed if and only if it contains


.......(all/ some) its .......(limit points/interior points).

9. Interior set and Derived set of sets Z, N are ..........(empty/


non empty) where as Q0 = ∅ and D(Q) = R.

10. Interior set and Derived set of finite sets are ..........(empty/
non empty).

11. Derived set of open interval (a, b) is [a, b].


Chapter 5

Sequences of real numbers

This chapter is devoted to the study of sequences of real numbers.


First we gave the notion of sequences and then describe their na-
tures. We discuss the relationship between convergent sequences
and bounded sequences and discuss also the nature of limits of
sequences. We give its application through fruitful discussions.

Definition 5.0.1. Let A be a set. Then a function f : N → A


is called a sequence in A. If A ⊂ R then the sequence f is called
a sequence of real numbers. The sequence f is denoted by hf (n)i
or hfn i or {fn }. Thus f = hfn i. The image f (n) of n under
function f is also denoted by fn and is called the nth -term of
the sequence. If hfn i is a sequence then we also write f =
{f1 , f2 , . . . , fn , . . .}. If fn = c for all n ∈ N, then the sequence f is
called a constant sequence and we frequently write f = c instead
of hci. The set {fn |n ∈ N} is called the range of sequence
f = hfn i.

Example 5.0.2. 1. hni = {1, 2, 3, . . . , n, . . .} is a sequence.


Observe that the range of hni is N.

2. h(−1)n i is a sequence with range set {−1, 1}.

73
74

1 1
3. hfn i is a sequence with fn = 1 + 1!
+ ... + n!
.

4. hci is a sequence with range set {c}, where c ∈ R.

Definition 5.0.3. Let σ = hnk i be a sequence of natural numbers


such that n1 < n2 < . . . nk < . . . and f = hfn i be a sequence of real
numbers. Then the sequence f oσ = hfnk i is called a subsequence
of the sequence hfn i.

Example 5.0.4. Consider the sequence hfn i. Take σ1 = h2k i


and σ2 = h3k i. Clearly 2k < 2k+1 and 3k < 3k+1 for all k. Then
f oσ1 = hf2k i will be hf1 , f2 , f4 , f8 , . . . f2k , . . .i the subsequence of
hfn i. Similarly hf3k i will be subsequence of hfn i.

Exercise 5.0.1. Write down the nth -term a sequence 1, 3, 6, 10, 15, . . ..
Also write its two subsequences.

Definition 5.0.5. A sequence hfn i is said to be a bounded se-


quence if there exists real numbers m and M such that m < fn <
M for all n ∈ N (i.e;its range set is bounded). In this case we find
a positive real number M such that |fn | < M for all n. The se-
quence fn is said to be unbounded if it is not bounded. Thus
if hfn i is unbounded then for each positive real number M either
fn < −M for some n (unbounded below) or fn > M for some
n (unbounded above). A sequence hfn i is called a monotonic
increasing (decreasing) sequence if fn ≤ fn+1 (fn ≥ fn+1 ) for all n.
It is said to be strictly monotonic increasing (decreasing) sequence
if fn < fn+1 (fn > fn+1 ) for all n.

5.1 Limit of a sequence


Definition 5.1.1. A real number l is said to be a limit of a
sequence hfn i as n approaches infinity ( n tends to ∞) if for
75

each positive real number ǫ there exists a natural number n0 such


that
n ≥ n0 ⇒ |fn − l| < ǫ or |fn − l| < ǫ ∀n ≥ n0
that is; fn ∈ (l − ǫ, l + ǫ) for all n ≥ n0 .
We write l = limn→∞ fn . For convenience we write lim fn = l.
A sequence hfn i whose limit is l is called a convergent sequence
and converges to l. A sequence hfn i which is not convergent is
called a divergent sequence.
Example 5.1.2. c is the limit of a constant sequence hci because
for each ǫ > 0 we have |c − c| = 0 < ǫ for all n. Thus every
constant sequence is a convergent sequence. Observe that the
range {c} of sequence hci is finite and hence no point is a
limit point of the set {c}. Thus limit point of a set and
limit of a sequence are two distinct concepts.
Example 5.1.3. The sequence h1/ni is a convergent sequence
converges to 0 because for each ǫ > 0 we get a natural number
n0 > (1/ǫ) such that |1/n − 0| = 1/n < ǫ for all n ≥ n0 .
Example 5.1.4. 1 is not a limit of sequence h(−1)n i because for
ǫ = 1/2, |(−1)2n+1 − 1| = 2 > ǫ for all n. Similarly no l ∈ R is
a limit of h(−1)n i. Thus it is a divergent sequence. Such type of
sequence is callled an oscillatory sequence.
Proposition 5.1.5. (Uniqueness of limit). If limit of a se-
quence exists then it is unique. Thus, a convergent sequence con-
verges to a unique limit.
Proof: Let l1 and l2 be any two distinct limits of hfn i. Choose
ǫ = |l1 −l
2
2|
. Then by definition of limit of sequence, we get natural
numbers n1 and n2 such that

n ≥ n1 ⇒ |fn − l1 | < ǫ and (5.1.1)


n ≥ n2 ⇒ |fn − l2 | < ǫ (5.1.2)
76

Choose n0 = max {n1 , n2 }. Then we have

n ≥ n0 ⇒ |fn − l1 | < ǫ and |fn − l2 | < ǫ (5.1.3)

Thus, for n ≥ n0 we have

|l1 − l2 | = |(fn − l1 ) − (fn − l2 )|


≤ |fn − l1 | + |fn − l2 |
< 2ǫ = |l1 − l2 |

which is a contradiction. Thus l1 = l2 . ✷

Proposition 5.1.6. Any two subsequences of a convergent se-


quence converges to the same limit.

Proof: Follows from the definition of subsequence and con-


vergent sequence.✷

Example 5.1.7. The sequence h(−1)n i is divergent because the


subsequences h(−1)2n i =< 1 > and h(−1)2n+1 i =< −1 > con-
verges to different limits 1 and −1 respectively.

Proposition 5.1.8. (Bounded property) Every convergent se-


quence is bounded.

Proof: Let < fn > be a convergent sequence and lim fn = l.


By definition of limits, for ǫ = 1 we get a natural number N such
that

n≥N ⇒ |fn − l| < 1 i.e; |fn | − |l| < 1 (as |a| − |b| < |a(5.1.4)
− b|)
i.e; |fn | < 1 + |l| (5.1.5)

Chose M = max {f1 , f2 , . . . , fN −1 , 1 + |l|} then |fn | ≤ M for all


n. ✷
77

Remark 5.1.9. A bounded sequence < fn > may or may not be


convergent sequence. For example, < (−1)n > is bounded but not
convergent.

Proposition 5.1.10. (Preservence of sign) If lim fn = l and


fn ≥ 0 then l ≥ 0.

Proof: Suppose if possible that l < 0. Then for ǫ = −l/2


there is a natural number n0 such that |fn − l| < −l/2 for all
n ≥ n0 , that is; l + l/2 < fn < l − l/2 = l/2 < 0 for n ≥ n0 . But
fn ≥ 0, thus our assumption is wrong. Hence l ≥ 0. ✷

Corollary 5.1.11. If lim fn = l and lim gn = m with fn ≤ gn


then l ≤ m.

Hint: |fn − l| < ǫ/2 and |gn − m| < ǫ/2 gives |(gn − fn ) −
(m − l)| < ǫ. Hence m − l = lim(gn − fn ). But (gn − fn ) ≥ 0
therefore m − l ≥ 0, i.e; l ≤ m.

Proposition 5.1.12. (Sandwich theorem or Squeeze Lemma)


Let hfn i, hgn i and hhn i be sequences such that fn ≤ gn ≤ hn for
all n. If lim fn = l = lim hn then lim gn = l.

Proof: Since lim fn = l = lim hn , therefore for ǫ > 0 there


are natural numbers n1 and n2 such that

n ≥ n1 ⇒ |fn − l| < ǫ and


n ≥ n2 ⇒ |hn − l| < ǫ
that is; n ≥ n1 ⇒ l − ǫ < fn < l + ǫ and
n ≥ n2 ⇒ l − ǫ < hn < l + ǫ

Choose n0 = max {n1 , n2 }. Then we have

n ≥ n0 ⇒ l − ǫ < fn < l + ǫ and l − ǫ < hn < l + ǫ


78

But fn ≤ gn ≤ hn for all n. Thus, for n ≥ n0 , we get

l − ǫ < fn ≤ g n ≤ h n < l + ǫ

, i.e;
|gn − l| < ǫ

Definition 5.1.13. Let < an >, < bn > be any two sequences and
c ∈ R. Then

1. sum < an > + < bn > of sequences < an > and < bn > is
defined by < an + bn > whose nth -term is an + bn .

2. difference < an > + < bn > of sequences < an > and


< bn > is defined by < an − bn > whose nth -term is an − bn .

3. product c. < an > of constant c and sequence < an > is


defined by < c.an > whose nth -term is c.an .

4. Product < an > . < bn > of sequences < an > and < bn >
is defined by < an .bn > whose nth -term is an .bn .

5. quotient <a n>


<bn >
of sequences < an > and < bn > is defined
by h bn i whose n -term is abnn provided bn 6= 0 for all n.
an th

5.2 (Algebra of limits or Algebra of


convergent sequences)
Proposition 5.2.1. Let hfn i, hgn i be any two sequences and c ∈
R. If lim fn = l and lim gn = m. Then

(a) lim(fn + gn ) = l + m = lim fn + lim gn .


79

(b) lim c.fn = c.l = c. lim fn and so lim(fn − gn ) = l − m =


lim fn − lim gn .

(c) lim(fn .gn ) = l.m = lim fn . lim gn .

(d) lim g1n = 1


m
= 1
lim gn
provided gn 6= 0 for all n and m 6= 0.
Proof: (a) Given ǫ > 0 we have natural numbers n1 and n2
such that

n ≥ n1 ⇒ |fn − l| < ǫ/2 and (5.2.1)


n ≥ n2 ⇒ |gn − l| < ǫ/2 (5.2.2)

Choose n0 = max {n1 , n2 }. Then

n ≥ n0 Rrightarrow |fn + gn − (l + m)| ≤ |fn − l| + |gn − l|


< ǫ/2 + ǫ/2 = ǫ

(b). If c = 0 then nothing to do. If c 6= 0 then for given ǫ > 0


there is a natural number N such that n ≥ N ⇒ |sn − l| < ǫ/|c|
and so |cfn − cl| < ǫ for all n ≥ N .
(c) Since

|fn gn − f g| = |fn (gn − g) + g(fn − f )| (5.2.3)


≤ |fn | |gn − g| + |fn − f | |g| (5.2.4)

and < fn > is convergent therefore there exists a positive real


number M such that

|fn | ≤ M ∀n (5.2.5)

Let ǫ > 0. By definition of limit of sequences we get natural


numbers N1 and N2 such that
ǫ
n ≥ N1 ⇒ |fn − f | < and (5.2.6)
2(|g| + 1)
ǫ
n ≥ N2 ⇒ |gn − g| < (5.2.7)
2M
80

Choose N = max{N1 , N2 }. Then we have


ǫ
n ≥ N ⇒ |fn − f | < and (5.2.8)
2(|g| + 1)
ǫ
|gn − g| < (5.2.9)
2M
Using Eqs. 5.2.3, 5.2.5, and 5.2.8 we have
ǫ ǫ
n ≥ N ⇒ |fn gn − f g| < M 2M + |g| 2(|g|+1) < ǫ (5.2.10)

This proves the required result.


(d) Choosing n1 such that

|gn − m| < (1/2)|m| f or n ≥ n1 .

Then
1 2
|gn | > (1/2)|m|, that is; < f or n ≥ n1 .(5.2.11)
|gn | |m|
Let ǫ be a given positive real number. By definition of limit
of < gn >, we have a natural number N such that

|m|2
|gn − m| < f or all n ≥ N. (5.2.12)

Choose k = max{n1 , N }. Hence, for n ≥ k, Equations 5.2.11
and 5.2.12 hold and so
1 1 |gn − m|
− =
gn m |m||gn |
2|gn − m|
<
|m|2
< ǫ.


81

Theorem 5.2.2. (Monotone convergence theorem) Every mono-


tonic increasing sequence, which is bounded above, is convergent.

Proof: Let < xn > be a monotonic increasing sequence which


bounded above. Then its range set X = {xn |n ∈ N} is non-
empty and bounded above. Thus by completeness axiom, it has
supremum sup xn = x in R. Let ǫ be any positive real number.
Then x − ǫ is not an upper bound of X and so there is a natural
number N such that x − ǫ < xN . But xN ≤ xn for all n ≥ N and
xn ≤ x < x + ǫ for all n. Thus for n ≥ N we have

x − ǫ < xn < x + ǫ, that is; |xn − x| < ǫ

Thus lim xn = x. ✷

Corollary 5.2.3. Every monotonic decreasing sequence, which is


bounded below, is convergent.

Proof: Left as an exercise. ✷


Combining above two results, we have

Corollary 5.2.4. Every monotonic sequence, which is bounded,


is convergent.

Remark 5.2.5. A convergent sequence need not be monotonic,.


For example, consider the sequence < xn >, where

−1 if n = 1
xn = 1
n−1
if n ≥ 2

Then lim xn = 0 but it is neither a monotonic increasing nor


monotonic decreasing sequence. Indeed C ⊂ B and B ∩ M ⊂
C, where B, C and M denote the set of all bounded sequences,
convergent sequences and monotonic sequences of real numbers
respectively. Equality need not hold.
82

Theorem 5.2.6. (Cauchy’s first theorem on limits) If lim xn = x


then
x1 + x2 + . . . + xn
lim = x. (5.2.13)
n
Proof: Define a sequence < yn > such that yn = xn − x for
all n ∈ N. By properties of limit of sequence we have lim yn =
lim xn − x = 0.
x1 + x2 + . . . + xn y1 + 22 + . . . + yn
lim = x + lim .
n n
Thus to prove the result it is sufficient to prove that
y1 + 22 + . . . + yn
lim = 0 (5.2.14)
n
Let ǫ > 0. Since lim yn = 0 therefore we have natural number n1
such that
ǫ
|yn | < f or all n ≥ n1 (5.2.15)
2
Next, every convergent sequence is bounded and so we have a
positive real number k such that

|yn | < k ∀ n. (5.2.16)

Choose n2 such that n2 > 2nǫ1 k and let N = max{n1 , n2 }. Then,


for n ≥ N equations 5.2.15 and 5.2.16 hold. Thus
y1 + 22 + . . . + yn
n≥N ⇒
n
|y1 | + |y2 | + . . . + |yn1 |yn1 +1 | + |yn2 +2 | + . . . + |yn |
≤ +
n n
n1 k (n − n1 )ǫ
< +
n 2n
ǫ ǫ
< + =ǫ
2 2
83

because (n−n
n
1)
< 1 and n ≥ N > 2n1 k
ǫ
, that is; nn1 k < ǫ
2
. This
proves the required result. ✷

Theorem 5.2.7. (Cauchy’s second theorem on limit). Let hsn i be


a convergent sequence of non-negative real numbers converges to
1
s then lim(s1 s2 . . . sn ) n = s.

Proof: Define tn by tn = logsn for all n ∈ N. Then htn i


is a convergent sequence with lim tn = logs. By Cauchy’s first
theorem on limit, we have
t1 + t2 + . . . + tn
lim = logs.
n
1 1
This gives log(s1 s2 . . . sn ) n = logs and so lim(s1 s2 . . . sn ) n = s.

Definition 5.2.8. . A real number p is called a limit point of


sequence hsn i if every ǫ − nhd (p − ǫ, p + ǫ) of p contains infinitely
many terms of the sequence hsn i. Thus −1 and +1 both are limit
point of sequence h(−1)n i but no point is a limit of its range set.
Also this sequence has no limits.

Remark 5.2.9. The limit point of sequence, limit of sequence and


limit point of range set of sequence are distinct concepts.

Theorem 5.2.10. (Bolzano-Weierstrass theorem for sequence). Ev-


ery bounded sequence has a limit point.

Proof: Let hxn i be a bounded sequence. Then its range set


X = {xn | n ∈ N} will be non-empty bounded subset of R.
Case I:
If X is finite then there exists xn0 ∈ X such that xn = xn0 for
infinitely many n ∈ N. Thus xn0 is a limit point of hxn i.
Case II: If X is infinite then it will be bounded infinite subset of
84

R. Thus, by Bolzano-Weierstrass theorem, X has a limit point p,


say. Then every ǫ-nhd (p − ǫ, p + ǫ) of p contains infinitely many
points of X and so infinitely many terms of hxn i. Thus p is a limit
point of the sequence.

Note: In this case for each k we select a term xnk such that
xnk ∈ (p − k1 , p + k1 ) so that n1 < n2 . . . nk . . .. This gives us a
subsequence hxnk i of hxn i such that lim xn = p. Thus we have the
following:
Corollary 5.2.11. Every bounded sequence has a convergent sub-
sequence.
Proposition 5.2.12. If lim xn = x then x is a limit point of the
sequence hxn i.
Proof: The proof follows immediately from the definition of
limit of the sequence. ✷

Remark 5.2.13. The converse of above result is not true. For


example, h(−1)n i has two limit points but it has no limit, i.e;
lim(−1)n does not exist.

5.3 Cauchy Sequence


Definition 5.3.1. A sequence hxn i is called a Cauchy sequence
if for every positive real number ǫ there is a natural number N
(depending upon ǫ only) such that

n, m ≥ N ⇒ |xn − xm | < ǫ. (5.3.1)

Example 5.3.2. h n1 i is a Cauchy sequence.


Proposition 5.3.3. Every convergent sequence is a Cauchy se-
quence.
85

Proof: Let hxn i be a convergent sequence with lim xn = x.


Let ǫ be any positive real number. Then, for 2ǫ > 0 we have a
natural number N such that
ǫ
n ≥ N ⇒ |xn − x| <
2
Thus, for n, m ≥ N , we have
|xn − xm | = |(xn − x) + (x − xm )|
≤ |(xn − x)| + |(x − xm )|
ǫ ǫ
< + =ǫ
2 2
Thus hxn i is a Cauchy sequence. ✷

Proposition 5.3.4. Every Cauchy sequence is a bounded sequence.


Proof: Let hxn i be a Cauchy sequence. Then, for ǫ = 1 we
have a natural number N such that
n, m ≥ N ⇒ |xn − xm | < 1
In particular |xn − xN | < 1 ∀n ≥ N , that is; |xn | < |xN | + 1 for
all n ≥ N . Choose M = max{|x1 |, |x2 |, . . . , |xN −1 |, |xN | + 1}, the
|Xn | ≤ M . Thus hxn i is bounded. ✷

Theorem 5.3.5. (Cauchy Convergence Criterion ) A sequence in


R is convergent if and only if it is a Cauchy sequence.
Proof: Let hxn i be a sequence of real numbers. If it is con-
vergent then it is a Cauchy sequence (by Proposition 5.3.3). Con-
versely suppose that it is a Cauchy sequence. Then it will be
bounded. By Bolzano-Weierstass theorem, hxn i has a limit point
x, say(i.e; it has a convergent subsequence). Let ǫ > 0 be any real
number. Then, for 3ǫ > 0 we have a natural number N such that
ǫ
n, m ≥ N ⇒ |xn − xm | <
3
86

and 3ǫ -nhd (x − 3ǫ , x + 3ǫ ) contains infinitely many points of the


sequence. Choose xn0 ∈ (x − 3ǫ , x + 3ǫ ) with n0 > N . Then
ǫ
|xn0 − x| < .
3
Thus, for n ≥ n0 we have

|xn − x| = |(xn − xm ) + (xm − xn0 ) + (xn0 − x)| (where m ≥ n0 )


≤ |(xn − xm )| + |(xm − xn0 )| + |(xn0 − x)|
ǫ ǫ ǫ
< + + = ǫ.
3 3 3
Thus hxn i is a convergent sequence. ✷
Chapter 6

Sequence of functions

This chapter is devoted to study the sequences of real valued func-


tions defined on certain domain I ⊆ R. Sometimes the sequence
of functions converges for each x ∈ J ⊆ I. This determines a
function on J which we call limit function of the sequence. It is
not always possible to carry the common property of each term
to its limit function. But if we put certain conditions known as
uniform convergence, then we can carry the common property of
sequence to its limit function.
Definition 6.0.1. (Sequence of functions) Let I be non-empty
subset of R. A sequence hfn i, where each fn is a function from I
into R, is called a sequence of real valued functions defined on I.
Definition 6.0.2. (Pointwise convergence) Let hfn i be a se-
quence of real valued functions fn defined on I. If the sequence
hfn (x)i converges to a real number, say f (x), for each x ∈ J ⊆ I,
that is; if f (x) = limn→∞ fn (x) exists for each x ∈ J, then unique-
ness of limit determines a function f : I → R given by
f (x) = lim fn (x), ∀x ∈ J.
n→∞

In this case we say that the sequence hfn i converges pointwise


to f on J ⊆ I and we write fn (x) → f (x) pointwise on J. The

87
88

function f , where f (x) = limn→∞ fn (x), ∀x ∈ J ⊆ I, is called the


limit function of the sequence hfn i of functions. Thus, a sequence
hfn i converges pointwise to a function f on J ⊆ I if for each
x ∈ J, the sequence hfn (x)i of real numbers converges to real
number f (x), i.e; if for each x ∈ J and for each positive real
number ǫ, we have a natural number N (ǫ, x) (depending on ǫ and
x) such that

n ≥ N (ǫ, x) ⇒ |fn (x) − f (x)| < ǫ

Thus f (x) = limn→∞ fn (x) ∀x.


P∞
Definition 6.0.3. Let n=0 un (x) be an infinite series,Pwhere

hun (x)i be a sequence of real valued functions. If the series
P∞ n=0 un (x)
converges for each x ∈ I ⊆ R. Then, we define P∞f (x) = n=0 un (x),
for all x ∈ I and we say that the series n=0 un (x) converges
pointwise on I to a function f (x).P∞ The function f is called the
sum or sum function
P∞ of the series n=0 un . Thus, the sum function
f of the series n=0 un (x) is given by

X
f (x) = un (x), f or all x ∈ I.
n=0

6.1 Why Uniform Convergence?


In this section we gave some guiding examples to illustrate the
importance of uniform convergence of sequence. Indeed we shall
now show by means of several examples that the limit process can
not in general be interchanged without affecting the result.

Example 6.1.1. A sequence of continuous functions whose limit


function is discontinuous.
89

Consider the sequence hfn i of real valued functions fn defined


by
x2n
fn (x) = , x ∈ R.
1 + x2n
Then each fn is continuous on R. But its limit function

 0 if |x| < 1
1
f (x) = if |x| = 1
 2
1 if |x| > 1
is discontinuous at x = 1 and x = −1. Observe that fn (x) → f (x)
for each x ∈ R, i.e; fn → f pointwise on R but the property of
continuity of each fn is not carried on its limit function f .
Example 6.1.2. A sequence of continuous functions whose limit
function is continuous.

Consider the sequence hfn i of real valued continuous functions


fn defined by
xn
fn (x) = x ∈ (0, 1).
1 + xn
Then its limit function f given by f (x) = 0 is continuous every-
where on (0, 1). Thus continuity property of each term of sequence
of real valued functions is carried on its limit function.
Example 6.1.3. A sequence of functions for which limit of integrals
is not equal to integral of limit, i.e;
Z 1 Z 1
R1
lim fn (x)dx 6= 0 limn→∞ fn (x)dx = f (x) dx.
n→∞ 0 0

Consider the sequence hfn i of real valued functions fn defined


by
fn (x) = n2 x(1 − x)n , x ∈ [0, 1].
90

Then its limit function f is given by



 0 if x=0
f (x) = limn→∞ n2 x(1 − x)n = 0 if x=1
0 if 0 < x < 1

Thus f (x) = 0 for all x and so


Z 1 Z 1
f (x)dx = lim fn (x) = 0
0 0 n→∞

But
Z 1 Z 1 Z 1
2 n 2
fn (x)dx = n x(1 − x) dx = n x(1 − x)n dx
0 0 0
Z 1
= n2 tn (1 − t)dt (put x = 1 − t, dx = −dt)
0
n2

2 1 1
= n − =
n+1 n+2 (n + 1)(n + 2)
R1 n2
and so limn→∞ 0 fn (x)dx = limn→∞ (n+1)(n+2) = 1 which is not
R1
equal to 0 f (x)dx = 0.
Thus the operations ‘limit’ and ‘integration’ can not
always be interchanged.

Example 6.1.4. A sequence of functions for which limit of integrals


is equal to integral of limit, i.e;
Z 1 Z 1
R1
lim fn (x)dx = 0 limn→∞ fn (x)dx = f (x) dx.
n→∞ 0 0

Consider the sequence hfn i of real valued functions fn defined


by

fn (x) = xn , x ∈ (0, 1).


91

Then its limit function f is given by


f (x) = limn→∞ xn = 0.
and so Z 1 Z 1
f (x)dx = lim fn (x) = 0
0 0 n→∞
Next,
1 1
xn+1

1
Z
R1 n
fn (x)dx = 0
x dx = =
0 n+1 0 n+1
R1 1
and so limn→∞ 0 fn (x)dx = limn→∞ n+1 = 0 which is equal to
R1
0
f (x)dx = 0.
Thus the operations ‘limit’ and ‘integration’ can be
interchanged.
Example 6.1.5. A sequence of differentiable functions hfn i with
limit 0 but hfn′ i diverges.
Consider the sequence hfn i of real valued functions fn defined
by
sin nx
fn (x) = √ , x ∈ R.
n
Then its limit function f is given by f (x) = 0, ∀x ∈ R. Hence

f ′ (x) = 0 for all x. Observe that fn′ (x) = n cos nx → ∞ as
n → ∞ for all x. Thus hfn′ i does not converge to f ′ .
Example 6.1.6. Consider the series ∞
P
n=0 un (x), where

X x2
un (x) = , x ∈ R.
n=1
(1 + x2 )n−1

Then
n
(
X 0 if x = 0
fn (x) = ur (x) = x2 (1−(1+x2 )−n )
1−(1+x2 )−1
6 0
if x =
r=1
92

Then

0 if x = 0
f (x) = lim fn (x) = 2
n→∞ 1 + x if x 6= 0

This shows that series ∞


P
n=0 un (x) is a convergent series of con-
tinuous function whose sum function f (x) is discontinuous.
Exercise 6.1.1. Observe that

lim lim f (m, n) = 1 6= 0 = lim lim f (m, n),


n→∞ m→∞ m→∞ n→∞

where f (m, n) = m/(m + n) for all m, n ∈ N. Thus limits can


not be interchanged in general.

6.2 Uniform Convergence


Definition 6.2.1. A sequence hfn i of real valued functions defined
on interval I is said to converge uniformly on a subset J ⊆ I to
a function f if for every positive real number ǫ we have a natural
number N (depending on ǫ only) such that

n ≥ N ⇒ |fn (x) − f (x)| < ǫ, ∀x ∈ J.

We write fn → f uniformly on J.
From this it follows that if fn → f uniformly on J then
fn (x) → f (x) for each x ∈ J, i.e; fn → f pointwise on J.
Definition 6.2.2. A series ∞
P
n=0 un (x) converges uniformly
Pn on J
if the sequence hsn i of partial sums defined by sn (x) = r=0 ur (x)
converges uniformly on J.
Definition 6.2.3. A sequence hfn i of real valued functions is said
to be uniformly bounded on a subset I ⊆ R if we have a positive
real number M such that |fn (x)| < M for all n ∈ N and x ∈ I.
93
P
Similarly, a series un (x) is uniformly bounded on [a, b] if the
the sequence hsn (x)i of its nth -partial sum is uniformly bounded on
[a, b]; that is, we have positive real numberPkn such that |sn (x)| < k
for all x ∈ [a, b] and n ∈ N; that is, | r=1 ur (x)| < k for all
x ∈ [a, b] and n ∈ N.

Example 6.2.4. A sequence hxn i where x ∈ R is uniformly


bounded on [−k, k], for each k ∈ R.

Proposition 6.2.5. Every sequence of functions, which converges


uniformly on a subset I ⊆ R, is uniformly bounded on I.

Proof: Let hfn i be sequence of real valued functions converg-


ing uniformly to a function f on J ⊂ R. Then, for ǫ = 1 we have
a natural number N (depending on ǫ only) such that

n ≥ N ⇒ |fn (x) − f (x)| < 1, ∀x ∈ J.

that is;

|fn (x)| < |f (x)| + 1, f or all x ∈ J and n ≥ N.

Chose M = supx∈J {|f1 (x)|, |f2 (x)|, . . . , |fN −1 (x)|, |f (x)|+1}. Then

|fn (x)| ≤ M, ∀n ∈ N and x ∈ J.

Thus hfn i is uniformly bounded on J. ✷

Proposition 6.2.6. Assume that fn → f uniformly on J. If each


fn is continuous at c ∈ J then the limit function f is continuous
at c. Thus
lim lim fn (x) = lim lim fn (x).
x→c n→∞ n→∞ x→c

Proof: Let c ∈ J is an isolated point then f is automatically


continuous. Assume that c ∈ J is a limit point of J. Let ǫ > 0.
94

By our assumption, for every 3ǫ > 0 we have a natural number N


such that
ǫ
n ≥ N ⇒ |fn (x) − f (x)| <
3
ǫ
But fN is continuous at c and so for 3 there exists δ > 0 such that
ǫ
|x − c| < δ ⇒ |fN (x) − fN (c)| < , where x ∈ J.
3
Thus, for x ∈ J with |x − c| < δ we have
|f (x) − f (c)| = |{f (x) − fN (x)} + {fN (x) − fN (c)} + {fn (c) − f (c)}|
≤ |f (x) − fN (x)| + |fN (x) − fN (c)| + |fn (c) − f (c)|
ǫ ǫ ǫ
< + + =ǫ (using above two equations).
3 3 3
Thus f is continuous at c, that is; limx→c f (x) = f (c) = limn→∞ fn (c),
that is;
lim lim fn (x) = lim lim fn (x).
x→c n→∞ n→∞ x→c

Remark 6.2.7. The converse is not true; that is, a sequence of


continuous functions hfn i may converge to a continuous function,
although the convergence is not uniform. For example, fn (x) =
nx(1 − x)n for x ∈ [0, 1] and n ∈ N. Then each fn is contin-
uous with continuous limit function f (x) = 0, x ∈ [0, 1] but the
convergence is not uniform (we discuss it in latter). Thus, the
uniform convergence of a sequence (series) of continuous
functions is only a sufficient condition but not a necessary
condition for the continuity of limit (sum) function.
Theorem 6.2.8. (Cauchy criterion for uniform convergence). The
sequence of functions hfn i , defined on J, converges uniformly on
J if and only if for every ǫ > 0 there is a natural number N such
that
n, m ≥ N ⇒ |fn (x) − fm (x)| < ǫ ∀x ∈ J. (6.2.1)
95

Proof: Suppose that the sequence of functions hfn i, defined


on J, converges uniformly on J and f is the limit function. Let
ǫ > 0. By definition of uniform convergence, we have a natural
number N such that
ǫ
n ≥ N ⇒ |fn (x) − f (x)| < , ∀x ∈ J. (6.2.2)
2
Then, for n, m ≥ N and ∀x ∈ J, we have
|fn (x) − fm (x)| = |{fn (x) − f (x)} + {f (x) − fm (x)}|
≤ |fn (x) − f (x)| + |f (x) − fm (x)|
ǫ ǫ
< + =ǫ (by Eqn.6.2.2)
2 2
Conversely assume the hypothesis 6.2.1 holds. Then it follows
fn (x) → f (x) pointwise on J. Keeping n fixed and making m →
∞ in Eqn. 6.2.1, we have
n≥N ⇒ |fn (x) − lim fm (x)| < ǫ ∀x ∈ J.
m→∞

that is;
n≥N ⇒ |fn (x) − f (x)| ≤ ǫ ∀x ∈ J.
Thus fn → f uniformly on J. ✷
Taking m = n + p in Eqn. 6.2.1, we have
n ≥ N, p ∈ N ⇒ |fn+p (x) − fn (x)| < ǫ ∀x ∈ J.(6.2.3)
Corollary 6.2.9. A series ∞
P
n=0 un (x) of functions un (x), defined
on J, converges uniformly on J if and only if for every ǫ > 0 there
is a natural number N such that
n>m≥N ⇒ |um (x) + um+1 (x) + . . . + un (x)| < ǫ, ∀x ∈ J.
Putting m = n + p, p ∈ N, we have
n ≥ N, p ∈ N ⇒ |un (x) + um+1 (x) + . . . + un+p (x)| < ǫ, ∀x ∈ J.
96

Theorem 6.2.10. (Mn -Test) Suppose that limn→∞ fn (x) = f (x), x ∈


J. Put Mn = supx∈J |fn (x) − f (x)|. Then fn → f uniformly on
J if Mn → 0 as n → ∞.
Proof: Let ǫ > 0. By convergence of hMn i with lim Mn = 0,
we have a natural number N such that |Mn | < ǫ for all n ≥ N .
But Mn = supx∈J |fn (x) − f (x)|. Thus, for all x ∈ J and n ≥ N
we have |fn (x) − f (x)| < ǫ. This proves that fn → f uniformly
on J. ✷

Theorem 6.2.11. Let hfn i be a sequence of functions defined


P∞ on
J. Let |fn (x)| ≤ Mn for all n ∈PN and x ∈ J. Then n=0 fn
converges uniformly if the series Mn converges.
P
Proof: Let ǫ > 0. Assume that the series Mn converges.
By defintion of convergence, we have a natural number N such
that
n ≥ N, p ∈ N ⇒ |Mn + Mn+1 + . . . + Mn+p | < ǫ.
But |fn (x)| ≤ Mn for all n ∈ N and x ∈ J. Thus, for n ≥ N, p ∈ N
and x ∈ J we have
|fn (x) + fn+1 (x) + . . . + fn+p (x)| ≤ |Mn + Mn+1 + . . . + Mn+p | < ǫ.
This proves the result. ✷

Definition 6.2.12. A point x = a ∈ J is called a point of non-


uniform convergence of sequence hfn i if hfn i does not converge
uniformly to a function f in the nhd of a.
Example 6.2.13. The sequence hfn i, where fn (x) = xn , 0 < x <
1, converges pointwise on (0, 1) to a function f (x) = 0 for all
x ∈ (0, 1). Since
Mn = sup xn ≥ (1 − 1/n)n f or x = 1 − 1/n
→ 1/e 6= 0 as n → ∞
97

Thus, by Mn -test, the sequence hxn i does not converge uniformly.


Observe that x = 1 − 1/n → 1 as n → ∞. Thus x = 1 is a point
of non-uniform convergence.
P cos nx
Example 6.2.14. The infinite series np
, where x ∈ R and
p > 1, is uniformly convergent. For,
cos nx 1
p
≤ p f or all x ∈ R and n ∈ N
n n
P 1 P cos nx
and np
is convergent. Thus, by Weierstrass M -test, np
is
uniformly convergent. P sin nx
Similarly,
P xn by Weierstrass M -test, n2
, where x ∈ R and
np
with x ∈ [0, 1], p > 1 are uniformly convergent.
Exercise 6.2.4. Prove that the following sequences hfn (x)i do
not converge uniformly, where
nx
1. fn (x) = 1+n2 x2
, x ∈ R.
n2 x
2. fn (x) = 1+n4 x2
, x ∈ [0, 1].
2
3. fn (x) = nx exp−nx , x ∈ R.
Hint: (i) f (x) = limn→∞ fn (x) = 0, ∀x ∈ R and Mn =
n.1/n
sup |fn (x) − f (x)| ≥ 1+n 2 .1/n2 = 1/2 for x = 1/n. Thus Mn 6→ 0

as n → ∞. The point x = lim 1/n = 0 is a point of non-uniform


convergence.
(ii). f (x) = limn→∞ fn (x) = 0, ∀x ∈ [0, 1] and Mn = sup |fn (x) − f (x)| ≥
n2 .1/n2
1+n4 .1/n4
= 1/2 for x = 1/n2 . Thus Mn 6→ 0 as n → ∞. The point
x = lim 1/n2 = 0 is a point of non-uniform convergence.
(iii). f (x) = limn→∞ fn (x) = 0, ∀x ∈ R and

Mn = sup |fn (x) − f (x)| ≥ n.1/ n. exp(−n.1/n) → ∞

as n → ∞, for x = 1/ n. Thus Mn 6→ 0 as n → ∞. Thus
x = lim 1/n = 0 is a point of non-uniform convergence.
98

Alternate Method: To obtain Mn , we proceed as follows:


Take yn = |fn (x) − f (x)| for each n ∈ N. Consider the case (iii).
Then yn = nx exp(−nx2 ) and so dyn /dx = √ 0 gives n exp(−nx2 )+
2
nx exp(−nx )(−2nx) = 0; that is, x = 1/ 2n. Now

d2 yn /dx2 = [−2nx.(n − 2n2 x2 ) − 4n2 x] exp(−nx2 )


√ √
is negative for x = 1/ 2n. Thus Mn = sup |yn | = n/2 exp(−1/2) →
∞ as √ n → ∞. Thus, it is not uniformly convegent and x =
lim 1/ 2n = 0 is a point of non-uniform convergence.

Example 6.2.15. The sequence hfn i converges uniformly on [0, 1],


where fn (x) = xn−1 (1 − x).
For, the limit function f (x) = 0 for all x. Hence yn =
|fn (x) − f (x)| = xn−1 (1 − x). To compute Mn , we obtain a point
where yn is maximum. Now dyn /dx = 0 gives [(n − 1)(1 − x) −
x.(−1)]xn−2 = 0; that is, x = 0, 1 − n1 . Since d2 yn /dx2 is negative
at x = 1 − 1/n, therefore Mn = (1 − 1/n)n−1 .1/n → 1/e .0 = 0
as n → ∞. Thus, by Mn - test we have the result.
P
Theorem 6.2.16. (Abel’s Test) The series un (x)vn (x) will
converge uniformly on [a, b] if
P
(i). the series un (x) is uniformly convergent on [a, b], and

(ii). the sequence hvn (x)i is monotonic for every x ∈ [a, b] and is
uniformly bounded on [a, b].

Proof: Assume the hypothesis. Let ǫ > 0. Since hvn i is


uniformly bounded therefore there is a positive real number k
such that
|vn (x)| < k f or all x ∈ [a, b].
Next, since the sequence hvn (x)i is monotonic for every x ∈
P therefore {vn (x)−vn+1 (x)} has same sign for all n ∈ N. Now,
[a, b],
un (x) is uniformly convergent, therefore by Cauchy criterion for
99

uniform convergence of series, we have a natural number N such


that
ǫ
n ≥ N, p ∈ N ⇒ |un+1 (x) + . . . + un+p (x)| < , ∀x ∈ [a, b].
3k
Let rn,p (x) = un+1 (x) + . . . + un+p (x) for all n, p ∈ N. Then we
have
ǫ
n ≥ N, p ∈ N ⇒ |rn,p (x)| < , ∀x ∈ [a, b] (6.2.5)
3k
Now, for n ≥ N, p ∈ N and ∀x ∈ [a, b], we have

|un+1 (x)vn+1 (x) + un+2 (x)vn+2 (x) . . . + un+p (x)vn+p (x)|

= |rn,1 (x)vn+1 (x) + {rn,2 (x) − rn,1 (x)}vn+2 (x)+

. . . + {rn,p (x) − rn,p−1 (x)}vn+p (x)|

= |rn,1 (x){vn+1 (x) − vn+2 (x)} + rn,2 (x){vn+2 (x) − vn+3 (x)} + . . . +

rn,p−1 (x){vn+p− (x) − vn+p (x)} + rn,p (x)vn+p (x)|

≤ |rn,1 (x)| |{vn+1 (x) − vn+2 (x)}|+|rn,2 (x)| |{vn+2 (x) − vn+3 (x)}|

+ . . .+|rn,p−1 (x)| |{vn+p− (x) − vn+p (x)}|+|rn,p (x)| |vn+p (x)|

ǫ ǫ
< 3k
{vn+1 (x) − vn+2 (x)} + 3k
{vn+2 (x) − vn+3 (x)}+

ǫ ǫ
... + 3k
{vn+p− (x) − vn+p (x)} + 3k
|vn+p (x)|

ǫ
= 3k
[{vn+1 (x) − vn+p (x)} + |vn+p (x)|] (using Eqn.6.2.5)

ǫ
≤ 3k
[|vn+1 (x)| + |vn+p (x)| + |vn+p (x)|]
100

ǫ
< 3k
(k + k + k) = ǫ. ✷

P (−1)n−1
Example 6.2.17. Since the series n
is convergent and
the sequence hxn i is uniformly bounded and monotonic decreasing
P (−1)n−1 n
on [0, 1], therefore by Abel’s-test the series n
x is uni-
formly convergent on [0, 1]. Observe that the test is not applicable
on [0, k], where k > 1 because the monotonicity of hxn i is dis-
turbed.
P
Theorem 6.2.18. (Dirichlet’s Test) If a series un (x) is uni-
formly bounded
P on [a, b]; that is, there is a positive real number k
such that | nr=1 ur | < k, ∀n ∈ N, x ∈ [a, b] and a sequence hvn (x)i
is monotonic
P decreasing converging uniformly to 0 on [a, b] then
the series un (x)vn (x) converges uniformly on [a, b].

P Proof: Assume the hypothesis. Let ǫ > 0 be arbitrary. Since


un (x) is uniformly bounded on [a, b]. Therefore we have a pos-
itive real number k such that
n
X
ur (x) < k f or all x ∈ [a, b] and n ∈ N.
r=1

Thus,
n+p
X
ur (x) = |sn+p (x) − sn (x)| ≤ |sn (x)| + |sn+p (x)| < 2k
r=n+1

for all x ∈ [a, b] and n, p ∈ N. Next, the sequence hvn (x)i is


monotonic decreasing and converging uniformly to 0 on [a, b], then
for ǫ/2k > 0 we have a natural number N such that

n ≥ N ⇒ |vn (x) − 0| < ǫ/2k f or all x; thatis,


|vn (x)| < ǫ/2k f or all x
101

Thus, for x ∈ [a, b] and n ≥ N , p ∈ N, we have


n+p n+p
X X
ur (x)vr (x) ≤ ur (x)vp+1 (x) (because vp+1 ≤ vr ∀r ≥ n + 1)
r=n+1 r=n+1
n+p
X ǫ
< ur (x) .
r=n+1
2k
ǫ
< 2k. =ǫ
2k
P
Thus, by Cauchy convergence criterion, un (x)vn (x) is uniformly
convergent. ✷

Example 6.2.19. The series (−1)n is uniformly bounded be-


P
cause |sn (x)| ≤ 1 for all n and the sequence hvn (x)i, where vn (x) =
1
np +x2
, p ≥ 1 is monotonic decreasing sequence for each x ∈ R and
converging uniformly to 0 on R. Therefore, by Dirichlet’s test, the
P (−1) n
series np +x2
is uniformly convergent on R.
x
Exercise 6.2.6. 1. Using Mn -test, prove that the sequence 1+nx2
,
converges uniformly on R.

2. Using Mn -test, prove that the series


X  nx (n − 1)x


1 + n2 x2 1 + (n − 1)2 x2
does not converge uniformly on [0, 1] and x = 0 is a point
of non-uniform convergence.
P −nx
3. Using Mn -test, prove that the series xe does not con-
verge uniformly on [0, 1] and x = 0 is a point of non-uniform
convergence.
P x
4. Using Weierstrass M -test, prove that the series (n+x2 )2
x
P
and n(1+nx2 )2
are uniformly convergent on R.
102
P
Proposition 6.2.20. Assume that un (x) converges uniformly
on [a, b]. If each un is continuous at c ∈ [a, b] then the sum
function f is continuous at c.

Proof: The proof follows from the Proposition 6.2.6. ✷

6.2.1 Term By Term Integration and Differen-


tiation
Theorem 6.2.21. Let hfn i be a sequence of Riemann integrable
functions fn : [a, b] → R. If it converges uniformly to f on [a, b]
then f is Riemann integrable on [a, b] such that
Z b Z
f (x) dx = lim fn (x) dx.
a n→∞

Proof: We will prove it after studying Riemann integration.



P
Corollary 6.2.22. Let un be a series of Riemann integrable
functions un : [a, b] → R. If it converges uniformly to f on [a, b]
then f is Riemann integrable on [a, b] such that
Z b Z b X  ∞ Z
X
f (x) dx = un dx = un (x) dx.
a a n=1

Proof: The proof follows immediately by applying above the-


orem to the sequence of its n-th partial sum.✷ This result is known
as term by term integration
P
Remark 6.2.23. Uniform convergence of un is only a suffi-
cient condition but not a necessary condition for term by term
integration. For,
103

Consider the series


X  nx (n − 1)x

− , x ∈ [0, 1]
1 + n2 x2 1 + (n − 1)2 x2

whose each term is continuous and so integrable on [0, 1]. Its


sum function f (x) = limn→∞ sn (x) = 0 is continuous and so
integrable. Observe that the series does not converges uniformly
to f (x = 0 is a point of non-uniform convergence) but

XZ 1 n Z
X 1
un (x) dx = lim um (x)dx
0 n→∞ 0
m=1
Z 1Xn
= lim um (x)dx
n→∞ 0 m=1
Z 1
nx
= lim dx
n→∞ 0 1 + n2 x 2
1
= lim log(1 + n2 ) = 0
n→∞ 2n

PR1 R1P
Thus, u (x) dx = 0
0 n
un (x)dx = 0 whereas x = 0 is a
point of non-uniform convergence of the series.

Exercise 6.2.7. Using Weierstrass M - test, prove that


1 ∞
xn
X 
1
Z X
dx =
0 n2 n=1
n2 (n + 1)

Exercise 6.2.8. Examine term by term integration of the series


whose n-th partial sum is n2 x(1 − x)n , where 0 ≤ x ≤ 1.
104
Chapter 7

Numeric functions

Definition 7.0.1. A function f : N∪{0} → R is called a numeric


function. The image of n under f is denoted by fn . We frequently
say that fn is a numeric function instead of f . fn = 2n and 2n − 1
are examples of numeric function where n ∈ N ∪ {0}. Let a and
b be any two numeric functions then the sum a + b of a and b is
defined by (a + b)n = an + bn for all n = 0, 1, 2, ...... The product
Pann bn . The convolution a ∗ b of a and b
a.b is defined by (a.b)n =
is defined by (a ∗ b)n = i=1 ai bn−i . Thus the value of a ∗ b at n
is given by (a ∗ b)n = a0 bn + a1 bn−1 + . . . an b0 .

Exercise 7.0.1. Find the sum, product and convolution of nu-


meric functions a and b, where

0 if 0 ≤ n ≤ 7
an =
2n if n>7

and

3n

if 0 ≤ n ≤ 3
bn =
2n − 3n + 4 if n>3

Solution: The sum of two numeric functions an and bn is

105
106

given by
 n
 3 if 0 ≤ n ≤ 3
an + b n = 2n − 3n + 4 if 3 < n ≤ 7
 n
2 + 3n − 3n + 4 if n>7

The product of two numeric functions an and bn is given by



0 if 0 ≤ n ≤ 7
an .bn =
3n (2n − 3n + 4) if n>7

The convolution of two numeric functions an and bn is given


by

an ∗ bn = a0 bn + a1 bn−1 + a2 bn−2 + . . . + an b0
Xn
= ai bn−i
i=8
n−4
X
= [2i (2n−i − 3(n − i) + 4)] + 2n−3 33 + 2n−2 32 + 3.2n−1 + 2n
i=8

Definition 7.0.2. The modulus of a numeric function an is a


numeric function denoted and defined by |an |. Thus the value
of |a| at n is |an | for all n. If a is a P
numeric function then its
n
accumulated sum at n is given by i=0 ai . Let k be a fixed
natural number then the numeric functions S k a and S −k a are
defined by

k 0 if 0 ≤ n < k
(S a)n =
an−k if n≥k

and

(S −k a)n = an+k

respectively.
107

Exercise 7.0.2. Find the numeric function b if a ∗ b = c, where


a and c are given numeric functions.

Exercise 7.0.3. Rahul deposits Rs. 1 lakh in his GPF account


at an interest rate 8 percent per annum compounded annually.
Evaluate his amount after n years.

Solution: Let an be the amount after n years. Then an =


Rs. (1.08)n lakh.

7.1 Big-oh or Asymptotic dominance


Definition 7.1.1. Let a be a numeric function. Then the be-
haviour of limn→∞ an is known as the asymptotic behaviour of
a. Let b be another numeric function. If there exists n0 ∈ N such
that |an | ≤ kbn for n ≥ n0 then we say b asymptotically domi-
nates a or a is asymptotically dominated by b. We also say
that |an |/bn is bounded on [n, ∞]. Observe that if a is dominated
by b then ka is also dominated by b for any fixed constant k.

Exercise 7.1.1. The numeric function an = 2−n is asymptotically


dominated by bn = 1.

Definition 7.1.2. Let a be a numeric function. Then Order a is


the set of all numeric functions that are asymptotically dominated
by a. It is denoted by O(a) or ‘big -oh of a’. Thus, b ∈ O(a) if
b is asymptotically dominated by a. Observe that the numeric
function n is dominated by n2 . Thus n ∈ O(n2 ). Similarly 15n2 ,
n2 + 340, n, n − 5 ∈ O(n2 ).

Exercise 7.1.2. Prove that an = 12 n2 log n − n2 is not dominated


by bn = n2 . In other words an ∈
/ O(bn ).

Solution: For if, let k > 0 and n0 ∈ N such that |an | ≤ kbn
for all n ≥ n0 . Then 21 n2 log n − n2 ≤ kn2 for all n ≥ n0 . This
108

gives log n ≤ 2(k + 1) for all n ≥ n0 . This is a contradiction.


Thus proved.

Definition 7.1.3. Let A and B be any two sets and k is any


constant. Then the sets A + B, kA and A.B are given by
A + B = {a + b | a ∈ A, b ∈ B}
k.A = {k.a | a ∈ A}
A.B = {a.b | a ∈ A, b ∈ B}.

7.1.1 Properties
Theorem 7.1.4. For numeric functions a, b, c, we have follow-
ing:

(i). a is O(|a|).
(ii). if a ∈ O(b) then ka ∈ O(b), where k is constant.
(iii). if b ∈ O(a) then S i (b) ∈ O(S i (a)) for all i.
(iv). if b, c ∈ O(a) then b + c ∈ O(a).
(v). if a ∈ O(b) and b ∈ O(c) then a ∈ O(c).
(vi). if a ∈ O(b) then O(a) ⊆ O(b). Thus, if a ∈ O(b) and
b ∈ O(a) then O(a) = O(b).
(vii) O(a) + O(a) = O(a) for every numeric function a. If a ∈
O(b) then O(a) + O(b) = O(b).
(viii). kO(a) = O(ka) = O(a) for every positive constant k and
numeric function a.
(ix). O(a).O(b) = O(ab) for any nemeric function a and b.
109

7.2 Generating functions


Definition 7.2.1. Let an be a numeric function
P∞ thenn the sum
function A(x) of the infinite series A(x) = n=0 an x is called
a generating function in x corresponding to numeric function an .
Observe that eachPnumeric function an corresponds a generating
function A(x) = ∞ n
n=0 an x and conversely.

To obtain the generating function, first we note the following


formulae:


X
−1
(1 − x) = xr
r=0
(x + a)n = xn + nC 1 a.xn−1 + . . . + nC r ar xn−r + . . . + nC n an
ax (ax)2 (ax)r
exp(ax) = 1 + + + ... + + ...
1! 2! r!
n(n + 1) 2 n(n + 1)(n + 2) 3
(1 + x)−n = 1 − nx + x − x + ...
2! 3!
n(n + 1) 2 n(n + 1)(n + 2) 3
(1 − x)−n = 1 + nx + x + x + ....
2! 3!
In particular, for n = 2, we have

X
−2 2 3
(1 + x) = 1 − 2x + 3x − 4x + . . . = (−1)r (r + 1)xr .
r=0
X∞
(1 − x)−2 = 1 + 2x + 3x2 + 4x3 + . . . = (r + 1)xr .
r=0

Example 7.2.2. The generating function corresponding to nu-


meric function an = an is
∞ ∞
X
n
X 1
A(x) = an x = an x n = .
n=0 n=0
1 − ax
110

Example 7.2.3. The generating function corresponding to nu-


1
meric function 2n is 1−2x .

The following table gives generating functions corresponding


to some simple numeric functions.

numeric functions generating functions


1
ar = 1 1−x
ar = r!1 exp(x)
x
ar = r (1−x)2
2x
ar = r(r + 1) (1−x)3

By a little effort, one may easily observe the following:

Theorem 7.2.4. Let A(x), B(x) and C(x) denote the generating
functions corresponding to numeric functions an , bn and cn . Then

1. B(x) = kA(x) if bn = kan .

2. C(x) = A(x) + B(x) if cn = an + bn .

3. C(x) = A(x)B(x) if cn is the convolution of an and bn .

4. B(x) = A(kx) if bn = k n an

Exercise 7.2.1. Find the generating functions of following func-


tions:

1. 2, 4, 8, 16, . . .

2. 0, 1, −2, 4, −8, . . .
 
r+4
3. ar = for all r = 0, 1, 2, 3 . . . .
r
1
4. ar = (r+1)!
for all r = 0, 1, 2, 3, . . ..

5. ar = 2r + 4r for all r = 0, 1, 2, 3, . . ..
111

Example 7.2.5. Prove that the numeric functions corresponding


to generating functions ar = 7.2r and ar = r+1
3r
7
are 1−2x 9
and (3−x) 2

respectively.
Solution: The
P numeric function
P∞ corresponding to generating
function 7.2r is ∞
r=0 7.2 r r
x = 7 r=0 (2x) r
= 7
1−2x
.
The numeric function corresponding to generating function
r+1
3r
is
Xr+1
f (x) = xr
r
3r
x  x 2  x 3  x r
= 1+2 +3 +4 + . . . + (r + 1) + ...
3 3 3 3
1
=
(1 − x3 )2

Exercise 7.2.2. Find the numeric functions corresponding to fol-


lowing generating functions:
1. (1 + x)n + (1 − x)n .
(1+x)2
2. (1−x)3
.
f (x) P∞
3. 1+x
, where f (x) =ar x r .r=0

4. (1 + x)f (x), where f (x) = ∞ r


P
r=0 ar x .

Solution: Since numeric functions of (1 + x)n and (1 − x)n


are
Crn if 0 ≤ r ≤ n

ar =
0 otherwise
and
(−1)r Crn if 0 ≤ r ≤ n

br =
0 otherwise
112

respectively.
Thus, the numeric function of function (1 + x)n + (1 − x)n is
given by

 2Crn if 0 ≤ r ≤ n and ris even.


cr = 0 if 1 ≤ r ≤ n and r is odd.
0 if r > n

2
Similarly, the numeric function of (1+x) (1−x)3
is given by the coefficeint
of x in the expansion (1 + 2x + x )(1 + 3x + (3.4/2!)x2 + . . . +
r 2

(3.4 . . . (r + 2)/r!)xr + . . .) for r ≥ 0.


The numeric function of (1 + x)−1 f (x)is given by

f (x) X
= ar xr .(1 + x)−1
1+x r=0
= (a0 + a1 x + a2 x2 + . . . + ar xr + . . .).
(1 − x + x2 + . . . + (−1)r xr + . . .)

Then, the coefficient of xr is equal to (−1)r a0 + (−1)r−1 a1 + . . . −


ar−1 + ar . Therefore the numeric function of f1+x (x)
is ar − ar−1 +
r
. . . + (−1) a0 .
2
Similarly, the numeric function br of (1+x)
(1−x) 3 is given by the

coefficeint of x in the expansion (1 + x) ar xr and is


r
P


a0 if r = 0
br =
ar + ar+1 if r > 0
Bibliography

[1] W. Rudin, Principles of Mathematical Analy-


sis, Third Edition, Mc Graw Hill, 1976

[2] T M Apostol, Mathematical Analysis, Second


edition, Narosa Publishing House, 1985

113
Index

Axiom of extension , 1 Dirichlet’s Test, 100

limit function, 88 finite set, 55

Abel’s Test, 98 Greatest lower bound property,


Achimedean Property, 51 43
Approximation Property, 49
Axiom of specification, 2 half-open interval, 48

Bolzano-Weierstrass theorem, 70 inductive set, 48


bounded above, 41 infimum, 42
bounded below, 42 infinite intervals, 48
infinite set, 55
Cauchy Convergence Criterion, interior point, 62
85 interval, 48
Cauchy criterion for uniform con-
vergence of sequence, 94 Least upper bound property, 43
Cauchy’s first theorem on lim- left open interval, 48
its, 82 limit point, 61
Cauchy’s second theorem on limit,limit point of a sequence, 83
83 lower bound, 42
closed interval, 48 Monotone convergence theorem,
closed set, 62 81
countable set, 55
neighbourhood, 61
De-Morgan Law, 7
dense, 62 open interval, 48
difference, 5 open set, 62

114
115

Order a, 107
ordered set, 41

perfect, 62
Pointwise convergence, 87

rational numbers, 49
right open interval, 48

Sequence of functions, 87
supremum, 42

triangle inequality, 54

uncountable set, 55
Uniform Convergence, 92
upper bound, 41

View publication stats

You might also like