You are on page 1of 156

Srivatsan: 1st Proof dated 20.09.

2021
nd
2 Proof dated 18.10.2021
All Corrected : As on 23.10.2021
Fit for printing after alignment.
By
Dr. P. Manoharan 9994545666
Dr. J. Jayaraj 9994045676
018E1110
1 - 10

ANNAMALAI UNIVERSITY
DIRECTORATE OF DISTANCE EDUCATION

M.Sc. Mathematics
First Semester

ABSTRACT ALGEBRA
LESSONS : 1 - 10

Copyright Reserved
(For Private Circulation Only)
M.Sc. MATHEMATICS
FIRST SEMESTER
ABSTRACT ALGEBRA

Editorial Board
Members
Dr.Nirmala P. Ratchagar
Dean
Faculty of Science and
Professor of Mathematics
Annamalai University
Annamalainagar

Dr. R. Singaravel The Director


Director Directorate of Academic Affairs
Directorate of Distance Education Annamalai University
Annamalai University Annamalainagar
Annamalainagar

Dr.S. Sriram Dr. M. Seenivasan


Professor and Head Associate Professor and Coordinator
Department of Mathematics Mathematics Wing, DDE
Annamalai University Annamalai University
Annamalainagar Annamalainagar

Internals

Dr. S. Pazhani Bala Murugan Dr. G. Kannadasan


Associate Professor Assistant Professor
Department of Mathematics(FEAT) Department of Mathematics
Annamalai University Annamalai University
Annamalainagar Annamalainagar
Externals

Dr. G. Ayyappan Dr. D. Arivudainambi


Professor Professor
Department of Mathematics Department of Mathematics
Pondicherry Technical University CEG Campus
Pondicherry Anna University

Lesson Writers

Dr. P. Manoharan, Associate Professor


Dr. J. Jayaraj, Assistant Professor
Department of Mathematics
Annamalai University
Annamalainagar
M.Sc. MATHEMATICS
FIRST SEMESTER
ABSTRACT ALGEBRA
SYLLABUS
Unit I –Group theory
Definition of a Group – Some Examples of Groups – Some preliminary Lemmas
– Subgroups – A counting Principle – Normal Subgroups and Quotient Groups
Unit II - Group theory
Homomorphisms – Automorphisms – Cayley’s Theorem - Permutation Groups-
Another counting principle – Sylow’s Theorem
Unit III- Ring Theory
Definitions and examples of Rings- Some Special classes of Rings-
Homomorphisms – Ideals and Quotient Rings
Unit IV - Ring Theory
More ideals and Quotient Rings – The Field of Quotients of an Integral domain–
Euclidean rings- A particular Euclidean Ring- Polynomial Rings
Unit V - Ring Theory and Vector Spaces and inner Product Spaces
Polynomials over the rational fields – Polynomial Rings over commutative
Rings.
Vector Spaces and inner Product Spaces
Elementary basic concepts – Linear independence and bases – Dual spaces –
Inner products spaces
Content and Treatment as in the Book
Topics in algebra by I.N. Herstein, Second edition, John Wiley & Sons, 2002
REFERENCE BOOKS
1. Algebra by Micheal Artin, Prentice Hall of India, New Delhi, 1994.
2. Lectures in Abstract Algebra Volumes I, II and III by N. Jacobson, D. Van
Nostrand Co., New York, 1976.
2

M.Sc. MATHEMATICS
FIRST YEAR
ABSTRACT ALGEBRA
CONTENTS

Lesson Nos Chapter Name Page No.


Lesson – 1 Introduction 1
Lesson – 2 Group Theory 7
Lesson – 3 Homomorphisms 21
Lesson – 4 Cayley’s Theorem 39
Lesson – 5 Permutation groups 49
Lesson – 6 Rings 62
Lesson – 7 Ideals 84
Lesson – 8 Fermat’s Theorem 104
Lesson – 9 Vector Spaces 124
Lesson – 10 Inner Product Spaces 142
LESSON – 1

INTRODUCTION
1.1. INTRODUCTION
In this Lesson basic notations of sets and mapping, cartesion product, Binary
operations, permutation and its degree, equivalence relation, partial order relation,
equivalence class are discussed with suitable examples.
1.2. OBJECTIVES
Sets and mapping, Cartesian product, Binary operations, permutation and its
degree, equivalence relation, partial order relation, equalance class are explained in
detail.
1.3. CONTENTS
1.3.1. Definition : Sets - Mapping
1.3.2 Definition: Cartesian Product
1.3.3 Binary Operation
1.3.4 Definition: Permutation and its degree
1.3.5 Definition: Equivalence relation and Partial order relation
1.3.6 Definition: Equivalence Class
1.3.1 DEFINITION : SETS -MAPPING
The union of the two sets A and B written as AB={ x| xA or xB } is a set
and the intersection of the two sets A and B, written as AB = { x|xA and xB} is
also a set.
For any two sets A, B, the difference A-B = { xA| xB } is a set. For example,
if A = {1, 2, 5, 6, 7} and B = {1, 2, 3, 6, 8} then AB = {1, 2, 3, 5, 6, 7, 8},
AB = {1, 2, 6}, A – B = {5, 7} and B – A = {3, 8}.
Given the universal set X and the set A, the complement of A written as A is
the set : { xX | xA}. i.e. A = X – A. Two sets A and B are said to be equal if and
only if A  B and B  A.
The following fundamental properties of union, intersection and complements
can be easily proved. (Here, A, B, C denote arbitrary sets)
(a) AB = BA and AB = BA (commutativity of  and )
(b) A(BC) = (AB)C and A(BC) = (AB)C (associativity of  and )
(c) A(BC) = (AB)(AC) (Distributivity of  over )
(d) A(BC) = (AB)(AC) (Distributivity of  over )
(e) (A  B) = A   B and (A  B)  = A   B (De Morgan’s law)
1.3.2 DEFINITION: CARTESIAN PRODUCT
Let A, B are two given sets. Then their Cartesian product is defined as the set
of all ordered pairs (a, b) where aA and bB. The Cartesian product is denoted
2
by A  B. We declare the pair (a 1, b 1) to be equal to (a2, b2) if and only if a1= a2 and
b1= b2.
A mapping from one set S into another set T is a rule that associates with
each element s in S a unique element t in T. Let φ be a mapping from S to T. We
often denote this by writing.
φ
φ : S  T or S  T, if tT is the image of sS under φ . We represent this
fact by φ (s) = t. The set of all φ - images of the elements of S will be denoted by
the symbol φ (S).
We may thus write φ (S) = {(x)|xS}. Clearly  (S)  T. S is called the Domain
and T is called the Co-domain and  (S) is called the range of φ .
1.3.2.1 Equality of Mappings
Two mappings φ and f of S into T are said to be equal if and only if φ (s) = f(s)
for every sS.
Definition
The mapping φ :S  T is said to be onto if for a given t  T, there exists an
element s  S such that φ (s) = t.
An onto mapping is also called a Surjective mapping.
Example

Let Z denotes the set of all integers. Define : ZZ by (x)=x+3 for xZ. Then
 is onto.
Definition
The mapping  of S into T is said to be one-to-one (briefly 1-1), if for any
s1, s2  S such that s1  s2 then (s1)   (s2) ( otherwise if  (s1) = (s2) then s1 = s2 )
A one-to-one mapping is also called injective mapping.
Example
Let Z denote the set of all integers. Define f: Z  Z by f(x) = 2x. Then clearly f is
a one-to-one mapping ( but not onto.).
Definition
A one-to-one and onto mapping of a set onto another set is called a bijective
mapping.
It is easy to see that when a mapping  is defined between two finite sets, then
 is surjective if and only if  is injective. The above result is false when the sets
are not finite.
1.3.2.2 Inverse Mapping
Let f be a one-to-one mapping of a set S onto T. Let y be a member of T. Since f
is onto there exists at least one member x of S such that y = f(x). Since f is one-to-
one, the member x of S is unique for the given y.
3
Thus we can define a mapping of T into S which associates to each yS
(uniqueness and existence of x is seen) such that f(x) = y. This mapping is called the
inverse of f and is denoted by f-1. Clearly f-1 is also well defined, one one and onto.
Thus if f: ST is a one one and onto, then f-1: TS is given by f-1(t) = s where
f(s) = t.
1.3.2.3 Definition: Composition of Mappings
If : S  T and : T  U, then the composition (also known as the product) of
 and  is the mapping ( • ) : S  U defined by (•) (s) = ((s)) for every sS.
We list below some of the essential properties satisfied by composition of
mappings and they are easily established from definitions.
Property I
: ST and  : T  U, then
(a) (•) is onto if both  and  are onto.
(b) is necessarily onto if • is onto.
(c) (•) is one-to-one if both  and  are one-to-one.
(d)  is necessarily one to one if (•)is one-to-one.
Property II
The product of mappings is associative. That is, if  : ST,  : TU and
 : UC then (•)•=• (•)
Property III
If : S  T be a one-to-one and onto mapping and -1 is the inverse of  then
-1•= iT and •-1 =is where is : SS and iT : TT are identity mappings (in fact
bijective mappings) defined by is(s) = s, it(t)= t, V sS and tT. In particular taking
T = S the inverse of any bijective map : SS has the property -1 •  = •-1 = is
With the above properties in mind, one can obtain the following facts relating to the
set A(S) of all bijective mapping of any non empty set S:
(a) If ,A(S) then •  A(S).
(b) • (•) =(•) • for any , ,   A (S)
(c) i•i  A (S) and  •i = i •  =  for all   A(S)
(d) For any   A (S), -1  A(S) such that •-1 = -1•=I.
Because of the above properties, it will be mean later that A(S) becomes a group
under composition as the binary operation. This group A(S) turns out to be an
important example of a group, as much as that any group can be identified later with
one such group A(S) for a suitable set S. You may be aware of notions of binary
operation defined on a set S and also of a group. We will deal with these notions in
detail in the second lesson, though we formally give the definition of Binary operation.
1.3.3 BINARY OPERATION
A Binary operation on a set S is a rule which associates for every ordered pair
of elements a, b in S an element c in S. Thus it is a map from SSS.
4
1.3.4 DEFINITION: PERMUTATION AND ITS DEGREE
A bijective mapping on a finite set S is called a permutation on S. The Degree
of the permutation is nothing but the number of elements in the finite set S. The
set of all permutations of the set S of n elements is denoted by Sn. It may be seen
that the number of members in the set S is n!
If S = (a1, a2 … … … an) then a permutation of S is sometimes conveniently
represented by the symbol

 a1 a2 ... ... ... ... an 


 
 ai ai2 ... ... ... ... ain 
 1
where i1, i2 … in are the integers 1, 2, … … n arranged in some order.
The symbol with two rows is such that the upper row consists of n elements
a1, a2 … … an of the set S and each element in the lower row is the image of the one
immediately above it, under the said permutation.
For example let S = {1, 2, 3} . Here n=3. Then Sn is the set of all permutations
on S and will contain 3! = 6 elements and let them be f1, f2, f3, f 4, f5 and f6 where
1 2 3  1 2 3   1 2 3
f1   ; f 2    ; f 3   
1 2 3  1 3 2   3 2 1
 1 2 3  1 2 3 1 2 3
f 4   ; f 5    ; f 6   
 2 1 3  2 3 1 3 1 2
Note :

(1). Here (f5 f4)= (f5 • f4 )= 1 2 3   f 3


1 3 2 

(2). Also (f4 f 5)= (f4 •f5)= 1 2 3   f 2 etc


1 3 2 
1.3.4.1 Definition: Relation
The relation R is defined on a set S, as an ordered pair of elements a, b of S,
such that a has a relation R on b.
We write aRb if a has the relation R on b. It may be seen that every relation R
in a set S gives rise to a subset A of the product set SS defined by A = {(a, b) : aRb}
and conversely any subset A of SS defines a relation R on S by setting aRb
whenever (a, b)  A. Thus a relation on S may also be defined as a subset of SS.
1.3.4.2 Definition: Types of Relations
A relation R defined in a set S is said to be
(a) Reflexive if aRa, aS
(b) Symmetric If aRb  bRa (a, b S)
(c) Anti-Symmetric: If aRb, bRa  a = b (a, bS)
(d) Transitive: If aRb and bRc then aRc (a, b, cS)
5
1.3.5 DEFINITION: EQUIVALENCE RELATION AND PARTIAL ORDER RELATION
A relation is said to be an equivalence relation if it is
(i) reflexive
(ii) symmetric and
(iii) transitive.
A relation is said to be a partial order relation if it is
(i) reflexive
(ii) anti-symmetric and
(iii) transitive.
Example
In the set of triangles, the relation “Smaller” is an equivalence relation.
1.3.6 DEFINITION: EQUIVALENCE CLASS
If S is a set and if R is an equivalence relation on S, then the equivalence class
of aS is the set (a)= {xS|aRx}.
For example in the set of all triangles if similarity is the equivalence relation,
then the equivalence class determined by a given triangle is the set of all triangles
which are similar to the given triangles.
Corresponding to an equivalence relation R defined in a set S there exists a
partition of S into a set of mutually disjoint equivalence classes such that
(i) For any two members a and b of the same equivalence class, a has the
relation R to b.
(ii) For any two members a, b of different equivalence classes, a does not have
the relation R to b. We thus have the following basic theorem.
Theorem
The distinct equivalence classes of an equivalence relation on S provide us
with a decomposition of S as a union of mutually disjoint subsets. Conversely,
given a decomposition of S as a union of mutually disjoint, non-empty subsets, we
can define an equivalence relation on S for which the subsets are the distinct
equivalence classes.
1.4. REVISION POINTS
Sets, Mapping, cartesion product, Permutation and its degree Equivalence
relation, partial order relation, Equivelence class
1.5. INTENT QUESTION
As this lesson contains only basic definitions and examples, there is no intent
question.
1.6. SUMMARY
Sets, mapping, cartesion product, binary operation, permutation and its
degree, Equivalence relations, partial order relation, Equivalence Classes are
Explained.
6
1.7. TERMINAL EXERCISES
As this lesson deals only about basic definitions and examples, there is no
terminal exercise.
1.8. SUPPLEMENTARY MATERIAL
1. Algebra by Micheal Artin, Prentice Hall of India, New Delhi, 1994.
2. Lectures in Abstract Algebra Volumes I, II and III by N. Jacobson, D. Van
Nostrand Co., New York, 1976.
1.9. ASSIGNMENT
As this lesson deals only about basic definitions and examples, there is no
assignment.
1.10. SUGGESTED READINGS / REFERENCE BOOK / SET BOOK
Alegbra – By Michacl Artin – PHI Learning (P) Ltd
1.11. LEARNING ACTIVITIES
Students are requested to attend the P.C.P Classes and get the basic idea
related to this lesson
1.12. KEYWORDS
Set, mapping, cartesion Product, binary operations, permutation, degree of
permutation, equivalence relations, partial order relation and equivalance class.

7
LESSON – 2

GROUP THEORY
2.1. INTRODUCTION
In this lesson, group, abelian group, cyclic group, multiplication modulo m,
subgroup, right coset, left coset are defined and examples are given. Some problems
are solved. Lagranger’s theorem, Euler’s theorem, Fermat’s theorem are stated and
proved. Some problems related to these are also solved.
2.2 OBJECTIVE
Group, abelian group, cyclic group, subgroup are defined and suitable examples
are given. Lagrange’s theorem is stated and proved. Euler’s theorem, Fermat’s
theorem are also proved. Some problems related to these theorem are also solved.
2.3. CONTENTS
2.3.1 Definition : Group
2.3.2 Definition : Abelian Group
2.3.3 Definition : Cyclic Group
2.3.4 Definition : Multiplication modulo m
2.3.5 Definition : Sub-group
2.3.6 Definition : Right coset of H in G
2.3.7 Definition : Left coset of H in G
2.3.8 Legrange’s Theorem
2.3.9 Corollary : (Euler’s Theorem)
2.3.10 Corollary : (Fermat’s Theorem)
2.3.1 DEFINITION - GROUP
A non-empty set G together with a binary operation ‘.’, is said to be a group, if
(G, .) satisfies the following axioms:
(i) For any a, b  G, a.b  G (closure)
(ii). For any a, b, c  G, a. (b.c) = (a.b).c (associative law)
(iii). There exists an element c  G such that a.e. = e.a. = a for all a  G (the
existence of an identity in G).
(iv). For every a  G there exists an element a-1  G such that a.a-1 = a-1a =
e (the existence of inverse in G).
Note:
1. Notation-- < G, . >
2. The closure axiom (1) is redundant in as much as this axiom is implied even
in the definition of the binary operation ‘.’.
3. A set G together with the binary operation, satisfies the conditions (i) and (ii)
stated above is called a semi-group. We already seen that the set A(S)
together with the binary operation “composition of maps” satisfies the above
8
four conditions and hence A(S) becomes a group under composition of maps.
Still we can find elements ,fA(S) such that •f  f•. This enables us to
single out a special class of groups.
2.3.2 DEFINITIONS : ABELIAN GROUP
A group G is said to be an Abelian Group (or commutative group) if a.b = b.a
for every a, b  G
The number of elements in a group G is called the order of the group G and
denoted by O(G). If the group G has only a finite number of elements then we say G
is a finite.
Immediate Consequences from the Definition
(1) The following lemma can be proved from the definition.
Lemma 3.2.1
If G is a group then
(a) The identity elements of G is unique
(b) Every a  G has a unique Inverse in G.
(c) For every a  G, (a-1)-1=a
(d) For all a, b  G, (ab)-1 = b-1 a-1.
(II) General Associative Law
Since in a group G, the associative law for three elements is true, it can
be shown that all the possible products of any finite set of elements of G arrived at
by different ways of parenthesizing are equal. For example if we consider an ordered
set {a1, a2, a3, a4} of four elements of G, then the following four products are
equal.
{ { {a1, a2 } a3 } a4 } , {a1 { { a2, a3 } a4 } } {a1 {a2, {a3 a4 }}}. Hence if we consider the
product of n elements each equal to a. This product is independent of the manner
of parenthesizing and there is a unique value of the product. We shall denote this
product by the symbol an.
n
Thus a i  a n (n being a positive integer and ai = a for all i)
i 1

n 1
We also define a0 = e and a {( a )}n and we note that the inverse of an is(an)-1
The laws of indices also can be verified, namely;
am. an=am+n = an+m = an.am
(am)n = amn. (Here m and n are any integers)
Example of groups
Example 3.2.2
Let G consist of the integers, 0,  1,  2 … and let the binary operation be the
usual addition of integers. Then G becomes a group in which 0 plays the role of e
and – a that of a-1.
9
Example 3.2.3
The set of all non-zero rational numbers is a group under the binary operation
1
“the usual multiplication”, in which ‘1’ plays the role of e and that of a-1.
a
Example 3.2.4
Let G = {1, -1}. Under the multiplication of real numbers, G forms a group of
order 2.
Example 3.2.5
S3 (the set of all permutations of 3 elements) forms a group under the binary
operation ‘composition’.
Example 3.2.6
Let n be any integer. We construct a group of order n as follows. G will
consist of all symbols ai, i = 0, 1, 2,…(n-1) where we to insists that a0 = an = e;
ai aj =ai+j ; if (i + j)  n and ai . aj = ai+j-n ; if (i + j) > n. It can be easily verified that
G is a group. It is a cyclic group of order n, as per the following definition.
2.3.3 DEFINITION: CYCLIC GROUP
In a group, if every element can be expressed as some power of any particular
element of the group, then the group is called cyclic group and that element is
called a generator for the cyclic group. ■
Example 3.3.1
The set of all n, nth roots of unity with multiplication as composition is a finite
cyclic group, n being any positive integer. The following lemma can be easily proved.
Lemma 3.3.2
Given a, b in the group G, then the equations ax = b and ya = b have unique
solutions for x and y in G. In particular, the two cancellation laws au = aw implies
u =w (left cancellation law) and ua = wa implies u = w (right cancellation ) hold in G. ■
Now let us solve some problems.
Problem 1
If G is a group in which (ab)i = ai bi for three consecutive integers i for all
a b  G then G is abelian.
Solution
Let (ab)i = ai bi …(1)
Then
(ab)i+1 = ai+1 b i+1 …(2)
Similarly
(ab)i+2 = ai+2 b i+2 …(3)
From the last we have
(ab)i+1 (ab) = ai+2 bi+2
10
ai+1 bi+1 a = ai+2 b i+1 (By (2) and the right cancellation of b)
aib i ba = ai+1 bi+1 (By the left cancellation of a)
(ab)i ba = (ab)i (ab) From (1) and (2).
By left cancellation law, we get
ba = ab. Hence G is a abelian. ■
Problem 2
Let G be a nonempty set closed under an associative operation and satisfies
the following.
(i) There exists an eG such that ae = a for all a  G.
(ii) Given a  G there exists an element y  G such that ay = e
Then G is a group under this product.
Solution
To prove G is a group, it suffice to show that, (1)e is also left identity i.e. ea = a
for all a in G and (2) all right inverse is also a left inverse, i.e. ay = e  ya = e.
First we will show that if ay = e then ya = e also. By (ii) for a given y we can
find y1 in G such that yy1 = e.
Consider yayy1 = y(ay)y1 = (ye)y1= yy1 = e
But yayy1 = ya(yy1) = y(ae) = ya
Therefore ya = e
Hence ay = ya = e.
Therefore all right inverse is also a left inverse.
Now we have ae = a. Just now we proved that for given a, there exists y such
that ay = ya = e.
Hence, ea = (ay) a
= a (ya)
= ae = a.
Thus, e is also a left identity. ■
Hence we can take the condition stated in the problem 2 as the alternative
definition of a group.
Thus we have Alternative Definition No.1 (for a group)
Alternative Definition No. 1 for a group: If G is a semi-group in which
atleast one right identity exists and for every element in G, there exists atleast one
right inverse with respect to the right identity, then G is a group.
Note: In the above definition we may replace and word “right” by the word
“left” throughout and we obtain an analogous definition for a group.
Problem 3
Prove, by an example, that the conclusion of the problem 2 is false if we
assume that
11
(a) There exists an eG such that ae = a for all aG.
(b) Given a in G, there exists y in G such that ya = e.
Solution

Consider the set G of matrices of the form a 0 where a, b are rational


b 0
numbers, and a  0. Let us have the usual matrix multiplication as a binary

operation. Clearly G is a semi-group. Also we find that 1 0 acts as right identity


1 0
1 
 a 0
and for given a 0 , a 0
 1  acts as left inverse. Here the set of matrices b 0 is not
b 0  
 0
b 
a group under matrix multiplication as can be verified. ■
Definition 3.3.3 ( Alternative Definition No.2 (for a group) as per the above problem.)
If G is a semi-group, in which for every pair of element a, b of G, there exists
elements x, y of G such that ax = b and ya = b then G is a group. ■
In order to prove that G is a group, we have to show that identity exists in G,
and that for every element in G there exists an inverse in G.
Let us first show the existence of identity. Consider an element a of G. By data,
there exists ea, fa in G such that aea = a and faa = a.
Let b be any arbitrary element of G. Then there exist elements x and y in G
such that ax = b and ya = b. Then, we have
be a = (ya) ea = y(ae a) = ya = b
fab = fa (ax) = (faa)x = ax = b
Thus ea and fa are two elements such that for every element b of G bea = b,
fab = b. In particular if we take b as fa or ea we have
fa ea = fa, f ae a = ea
Therefore fa = fa ea = e a
 fa = e a
Let we denote the element by e and so for every element b of G, eb = be = b
We show that inverse exists for every element a of G. By data for a given a in
G, there exist z, t in G such that. az = e and ta = e
Hence t(az) = te = t. But t(az) = (ta). z = ez = z. Hence t = z. Thus for every
element a of G, there exists an element t of G such that at = ta = e. This
implies that there exixts a unique inverse to each element of G. Hence G is a
group.■
Problem 4
If G is a finite semi-group in which both the cancellation laws are true then G
is group.
12
Proof
In order to show that G is a group, we have to prove the existence of the
identity element in G and an inverse for every element of G. Let
a1, a2 … ……ag …(1)
be the g distinct elements of the group G. If a is any one of them, then
aa1, aa2 … ……aag …(2)
are elements of G because G is closed and they are all distinct. Otherwise if
aai = aaj by cancellation law ai = aj which is a contradiction. Thus the g elements
written in (2) are same as that of (1) but for possibly a different order. Therefore if b
is any element of G, there exists an element a such that aan = b, for some
n{1,2,3,…,g} Similarly on post-multiplying a by a , we conclude that there exists
an element a, such that aa = b. Hence given a, b in G, the equations ax = b; ya = b
have solutions namely x = a, y = a. Hence G is a group by alternative definition No.2
for a group. ■
2.3.4 MULTIPLICATION MODULO M
Let m be any fixed positive integer which will be called the modulus. We define a
congruence relation  (relative to m) on the set of integers by setting a  b (mod m ), iff
(a – b) is divisible by m. One can easily check that the relation is an equivalence relation
and the equivalence class determined by the integer a is the set of all integer which are
congruent to a modulo m. One can also show that there are precisely m equivalence
classes and that we can denote them by (0), (1), … ,(m-1) since any integer is congruent
to one and only one of the numbers 0, 1, … (m –1) modulo m. For convenience we
restrict our attention to these numbers 0, 1, 2, …, (m-1) which form a complete set of
residue classes modulo m, We define Multiplication modulo m as follows.
If a, b  { 0, 1, 2, … (m-1) } which is congruent to the usual product a*b. It is
almost immediate that * is associative and commutative.
We now do a problem in which the multiplication modulo m is used and
incidentally we get further examples for finite abelian groups.
Problem 5
(a) Prove that, the non-zero integers modulo p(p, a prime number) form a
group under multiplication mod p.
(b) Do part (a) for non-zero integers relatively time to n under multiplication
mod n.
Solution
Consider the set S = {0, 1, 2, … (p-1)} where p is prime number. Clearly S is a
finite semi-group under multiplication modulo p. By problem 4, it is enough if we
show that the cancellation laws are valid in S. If k, x, y  S then kx  ky (mod p)
implies that p divides k(x-y), which means that p divides k or (x-y), p being a prime.
But p cannot divide k as k < p. Therefore p divides (x – y) which means
x  y (mod p). Moreover the multiplication mod p is commutative. Hence the other
cancellation law is also true. Hence S becomes a group and (a) follows
13
To Prove (b) : Let n be any positive integer. Let S denote the set {a1, a2 … a(n) }
of (n) numbers which are less than n and prime to n. [Here (n) denotes the Euler’s
phi function]. Under the operation ‘*’ namely multiplication mod n we note that S is
closed (i.e.) if a1, a 2  S then a1 * a2  S. For this it is enough to prove a is
relatively prime to n. If a1 *a2 is not relatively prime to n then there is a common
prime divisor say d (i.e.) d divides n as well as a1 * a2 [Because we can show that if
x  y (mod n) and d is a divisor of n then d divides x iff d divides y]. Since d is
prime, d divides either a1 or a2 contradicting the fact that a1 and a2 are relatively
prime to n. Clearly * is associative and by part (a), cancellation laws are also true.
Hence S becomes a group. ■
Now let us study the notion of sub-group.
Note: (a) is a particular case of (b) in problem 5.
2.3.5 DEFINITION: SUB-GROUP
A nonempty subset H of a group G is said to be a sub-group of G, if H itself
forms a group under the same operation on G■
It is clear that if H is a sub group of G, and K is a sub group of H, then K is a
subgroup of G.
The next two lemmas help us to decide whether a given subset of a group is a
sub-group or not.
Lemma 3.5.1
A non-empty subset H of a group G is a subgroup of G if and only if
(i) a, b  H implies that ab  H
(ii) a  H implies that a-1  H.
Proof
If H is subgroup of G, then it is clear that (i) and (ii) must hold.
Conversely let us assume that H is a subset of G for which (i) and (ii) holds. To
show that H is a subgroup we have to show that associative law holds in H, and
that H contains an identity element and inverse exists for every element. Since the
associative law holds for G, it holds for H also which is a subset of G. If a  H by (ii)
a-1  H and therefore by (i) a-1a = e  H. Thus e is the identity element for H. By (ii)
the inverse of a belongs to H. Thus H is a subgroup of G ■
Note: If H is finite then the lemma becomes nicer.
Lemma 3.5.2
If H is a non-empty finite subset of a group G and H is closed under
multiplication, then H is a subgroup of G.
Proof
By data, H is a finite semi group. Also since G is a group both cancellation
laws are true in G and hence in H. Therefore by problem 4, H is a group. Since H is
subset of G, H is a subgroup of G.
14
Example 3.5.3
Any group G having more than one element has atleast two sub-groups
namely G and the trivial sub-group containing the identity only.
Example 3.5.4
Let G be the group of integers under addition and H is the sub-set consisting
of all the multiples of 3. Then H is a sub-group of G.
Example 3.5.5
Let S be any set and let A(S) be the group consisting of all 1-1 mappings of S
on to itself. If x0S, let H(x0) = {A(S): ( x0)= x0}. Then one can check that H(x0) is a
subgroup of A(S).
Definition 3.5.6
Let G be a group and H be a sub-group of G. For a, b  G, we say a is
congruent to b mod H, written as a  b (mod H), if ab -1G. One can easily check
that the relation a  b (mod H) is an equivalence relation.
2.3.6 DEFINITION: RIGHT COSET OF H IN G
Let H be a sub-group of G and let aG. Then Ha = {ha : hH}, is called a right
cosset of H in G.
2.3.7 DEFINITION: LEFT COSET OF H IN G
Let H be a sub-group of G and let a G. Then aH = {ah:hH}, is called a left
coset of H in G.
Proposition: For any a,bG, Ha = Hb, iff ab-1  H and aH = bH, iff b -1 aH.
Proof: Let ab-1H. Then, for any ha  Ha where hH, we can write
ha = h(ab -1b)Hb and hence Ha  Hb. Again, for hb  Hb, we can write
hb = h(ba-1a)Ha, since ba-1 = (ab -1)-1H and H being a subgroup and ab-1H.
Hence Hb  Ha. Thus Ha=Hb. Conversely suppose that Ha = Hb. Then ab-1 = h –1h

 H; H being a subgroup and h, hH. Thus if Ha = Hb, we get that ab-1  H. We


have therefore proved that Ha = Hb if ab-1  H. Similarly it can be shown that
aH = bH iff b -1a  H.
As a consequence of the above proposition we can easily deduce that
Ha = { xG |x  a (mod H)} and this is the content of the following Lemma .
However, we give an (apparently) independent proof for the lemma. ■
Lemma 3.7.1
For all a  G, Ha = {xG| a  x (mod H)}
Proof: Let (a) denotes the equivalence class of a, corresponding to the
congruence relation  (mod H). Then (a) = { x  G : x  a (mod H) }. We want to show
(a) = Ha. If x  Ha, then x = ha for some h  H. Now, xa-1=h and x aH. Thus
(a)  Ha  (a). Hence the lemma. ■
15
2.3.8 LEGRANGE’S THEOREM
If G is a finite group and H is a sub-group of G, then O(H) is a divisor
of O(G).
Proof: We know that the relation congruence  is an equivalence relation on G,
(a) and thus Ha is the equivalence class of a in G. These equivalence classes yields
a decomposition of G into disjoint subsets. Thus any two right cosets of H in G are
either identical or have no elements in common.
Also we know that between any two right cosets Ha and Hb of H in G there
exists a 1-1 correspondence by associating with any element haHa, where hH,
hbHb. Clearly this mapping is onto . This mapping is also 1-1. For if h1b = h2b
with h1, h2 H, then by the cancellation law in G, h1a=h2a. Since G is finite, H is
also finite and therefore any two right cosets of H have the same number of
elements. Since He = H is itself a right coset of H, it will have O(H) elements. Thus
any right coset of H in G has 0(H) elements. Thus if k represents the number of
distinct right cosets of H in G, we must have that k.0(H) = 0(G). ■
Note 1: Lagrange’s Theorem has many applications. If G is a finite group of
order n and m is not a divisor of n then there can be no subgroup of order m. Thus
if G is a group of order 6, then there can be no subgroup of G of order 5 or 4.
Note 2: The converse of Lagrange’s Theorem is not true. If m is divisor of n,
then it is not necessary that G must have a subgroup of order m. The alternating
group A4 of degree 4 is of order 12. But there is no subgroup of A4 of order 6, even
though 6 is a divisor or of 12.
Definition 3.8.1
If G is a group and a  G, then the order(or period) of an element aG, is the least
positive integer m such that am = e. If no such integer m exists,then we say that a is
of order zero or of infinite order.
Corollary 3.8.2
If G is a finite group and a  G, then O(a) divides O(G).
Proof:
Let us consider the cyclic subgroup H of G generated by a. Clearly O(a) is finite
as G is finite. Also H consists of e, a, a2 … we claim that H has O(a) elements. Since
aO(a)= e, we will have atmost o(a) elements. If the sub-group has fewer than O(a)
elements then ai = aj for some integers 0  i < j < o(a). Then ai-j = e which is a
contradiction to the definition of O(a). Thus the sub-group H of G will have O(a)
elements. By Lagrange’s theorem O(a) | O(G). ■
Corollary 3.8.3
If G is a finite group and a  G then aO(G) = e.
Proof:
By corollary 1, O(a) | O(G), thus O(G) = m O(a) for some integer m. Therefore a
aO(G) = aO(a)m = (aO(a))m=e m=e. ■
16
2.3.9 COROLLARY: (EULER’S THEOREM)
If n is a positive integer and a is relatively prime to n then a(n)  1(mod n)
where  (n) is the number of positive integers less than n and relatively prime to n.
Proof
We have already proved in the problem 5 that the set H of numbers less than n
and relatively prime to n form a group under multiplication mod n. Hence H
becomes group under multiplication mod whose order is (n). Hence from corollary
1.2.2, we have for a in H, a(p)  1 (mod n) ■
2.3.10 COROLLARY : (FERMAT’S THEOREM)
If p is a prime number and a is any integer, then ap  a (mod p).
Proof: Since p is a prime numberwe have that (p) = (p – 1). Therefore the
given integer is either relatively prime to p or not relatively prime to p.
Case I: Let a be relatively prime to p.
If a is relatively prime to p, then by corollary 3,
a(p)  1 (mod p) (i.e.) ap-1  1 (mod p). Hence ap  a (mod p).
Case II: Let a be not relatively prime to p. Then since p is a prime p| a, p
divides a(ap-1-1) also. (i.e.) ap –a is divisible by p. Therefore ap  a (mod p).■
Problem
Prove that any group of prime order is cyclic and can be generated by any
element of the group except the identity.
Solution
Let G be a group of prime order p. Let a  G and a be not the identity in G.
(Such an element exists, as p being a prime, greater than 1). Let H = (a) be the
cyclic subgroup of G generated by a. Then, by Lagrange’s theorem, O(H) is a divisor
of the prime p, which is the order of G. This means that O(H) = 1 or p. However, as
a  H and a  e, O(H) > 1. Thus O(H) = p and  H = G and hence any group of
prime order is cyclic whose generator may not be an identity element. ■
In the proof of corollary 1, we mentioned about the cyclic subgroup H
generated by an element a in a group G. This notion of cyclic subgroup H = (a) can
be formally defined in either of two equivalent ways, viz. (i) (a) = the set of all powers
of a [positive, negative and zeroth powers of a] (ii) (a) is the intersection of all
subgroups of G which contain a; in other words, (a) is the smallest subgroup of G
which contains a. It is not difficult to prove the equivalence of these two definitions
for the cyclic subgroup generated by a. [i.e denoted by (a)]. More generally, we may
define the subgroup of G generated by any non empty set M of G, as the
intersection of all subgroups of G which contain M; in other words the subgroup
generated by M [denoted as (M)] is the smallest subgroup of G which contains M.
One can show that this subgroup (M) is the set of all finite products of powers
(positive or negative or zeroth powers) of elements in M.
17
If H, K are two sub-groups of G, let HK = { x  G : x = hk, hH and kK}. One
can see from the following example that HK need not be a sub-group of G. Consider
S3, the permutation group on 3 elements {1, 2, 3}. The permutations in S3 are:

 1 2 3  1 2 3  1 2 3 1
f4  
2 3
 1 2 3, f  1 2 3
, f5  
f1  ; f  ; f  
1 2 3 2 1 3 2 3 3 2 1 2 1 3 2 3 1 6 3 1 2
If H = {f1, f2}, K = {f3, f 5} then H, K are sub-groups. HK = {f1, f2, f 3, f 5}. Since HK
contains four elements and 4 does not divide 6 [6 being the order of S3], we can
immediately conclude from Lagrange’s theorem that HK cannot be a sub-group.
Our aim is to find when this HK will become a subgroup. The following lemma
answer this question.
Lemma 3.10.1
If H and K are subgroups of a group G, then HK is a sub-group of G if and only
if HK = KH.
Proof
First let us assume that HK = KH (i.e.) if hH, kK, then we can find some
h1, h2  H and k 1, k 2  K such that hk = k 1 h1 and kh = k2h2. We will show that HK
is a sub-group. For that we have to show that HK is closed and every element in HK
has its inverse in HK. Suppose x = hk  HK and y = hk  HK. Then xy = h(kh)k.
Since kh  KH = HK, we can find hH and kK such that kh = h1k1. Therefore xy
= h(h1k 1)k = (hh1) (k 1 k)  HK [Since h h1 H, k1 kK, H, K being subgroups).
Hence HK is closed. Also x-1 = (hk)-1 = k-1h -1  KH = HK. So x-1  HK. Thus HK is a
subgroup of G.
Conversely let us assume that HK in a sub-group of G.
Now we will show that HK = KH.
For any hH, kK, (h-1k-1) HK and so kh = (h-1k-1)-1 HK, as HK is sub-group.
Thus KH  HK . Similarly we can prove that HK  KH. Thus HK = KH. ■
Corollary 3.10.2
If H, K are sub-groups of an abelian group G, then HK is sub-group of G.
Theorem 3.10.3
If H and K are finite sub-groups of a group G of orders O(H) and O(K)
respectively then.
O(H)O(K)
O(HK) =
O(H  K)
Proof
In this proof, we have two separate cases as HK = {e} and HK contains more
than one element.
Case(i) : Let us assume that HK = {e}. In that case we will show that O(HK) = O(H)O(K).
Suppose this is not true, then among the elements of HK, atleast two elements, h1k1
18
and hk are equal for different elements h,h1H (i.e.) kh = h1k 1. That is h1-1h = k1k-1. Clearly
h1-1h  H, since k1 k1-1  K also. Therefore h1-1h  H  K. Therefore h1-1h=e. Hence h = h1
which is a contradiction. Therefore O(HK) = O(H) O(K).
Case (ii) : Let us assume that HK contains more than one element. Out aim is to
find how may elements are repeated in HK. We will show that it is only O(HK).
Let h1  H  K and h  H and k  K.
Then hk = (hh1 (h1-1k). Clearly hh1  H. Since h1-1  HK, HK being a sub-
group h1-1  K. Hence h1-1H and h1-1kK. Since h1 is arbitrary in HK, hk may be
written in yet another form (hh2) (h2-1k) by taking some other element h1 HK.
Thus hk is duplicated in the product atleast O(HK) times. However if hk = h1k 1
then h-1h 1 = k(k 1)-1 = u and u  HK. Now h1=hu, h1 – u-1k so that h1k 1=(hu) u-1k
where u HK. Hence hk is repeated at the most O(HK) times. Thus the number
of distinct elements in HK is the total number in the listing of HK that is O(H).
O(K) divided by the number of times a given element appears, namely O(HK).
O(H)O(K)
Thus O(HK) = ■
O(H  K)
Aliter
HK is a subset of G. O(HK) denotes the number of distinct elements of H.
Let L = HK. L is a subgroup of G and L  K.
Therefore L is a subgroup of K.
The number of distinct right cosets of L in K is finite and = m (say) ( K is
finite). Let the cosets be
Lk1, Lk2, … Lkm
By Lagrange’s Theorem
O(K)
m=
O(HK)
K = Lk1  Lk2 … Lkm
HK = HLk 1  HLk2 … HLkm
= Hk1  Hk2 … Hkm ( Since L  H  HL = H)
Claim: Hk1, Hk2 … Hkm are pair wise distinct.
For, if Hki = Hkj  k ik j-1  H
 k ik j-1  H K (Since ki, k j  K  kikj-1  K)
 k i = kj ( Since Lk1, … Lkm are distinct)
Thus Hk1…Hk m distinct right cosets and so they are pair wise disjoint. The number
of elements in each of them = the number of elements in H. ( Since H itself is a
right coset)
Therefore O(HK) = m × the number of elements in each right coset
19
= m O(H)
O(K )
= O(H)
O(L )
O(H) O(K )
= ■
O(H  K )
Corollary 3.10.4

If H and K are sub-groups of G and O(H) > O(G ) , O(K) > O(G ) ,then
H  K  {e}.
Proof: Since HK  G, O(HK) < O(G)
But we have

O( H ) O(K) O (G ) O(G) O(G )


O(G) > O(HK) =  
O( H  K ) O( H  K ) O( H  K )
Thus O(HK) > 1. Therefore H K  {e}. ■
Corollary 3.10.5
If G is a group of order pq, where p and q are primes ,then there is atmost one
(cyclic) subgroup of order p in G.
Proof: Let H and K are two subgroups in G each of order p. Since p > q,
O(H) = O(K) = p > pq  O (G ) . Hence by the last corollary HK {e}. But H and K
are subgroups of prime order p. Using Lagrange’s theorem, we find that
O(HK) = p, as O (HK)>1. This is possible only if HK = H = K. ■
Note: Later, using Cauchy’s Theorem, we can show that the group G of order
pq must have atleast one subgroup or order p (and at least one subgroup of order
q). Hence, using a corollary, we find that the group G has precisely one subgroup of
order p (and p, being a prime, this subgroup is necessarily cyclic).
2.4. REVISION POINTS
Group, Abelian group, Cyclic group, Sub-group, Lagrange’s Theorem, Euler’s
Theorem, Fermat’s Theorem.
2.5. INTEXT QUESTIONS:
1. If G is a group in which (ab)i = aibi for three consecutive integers i for all
a,b in G, then prove that G is abelian
2. It G is a finite semi group in which both concelation laws are ture then
prove that G is Abelian.
3. Prove that any group of prime order is cyclic and can be generated by any
element of the group except the identity
4. If G is a group, then prove the following
(i) The identify element of G is unique
(ii) Every a in G has a unique inverse in G
20
(iii) For every a in G, (a-1) -1 = a
(iv) For any a,b in G, (ab) -1 = b-1a-1
5. If H and K are sub groups of a cyclic group G, then prove that HK itself is
a cyclic group
6. How many generators does a cyclic group of order n have?
2.6 SUMMARY
Group, Abelian group, Cyclic group, Sub group, Multiplication madulo m,
Right coset, Left coset, Lagrange’s theorem, Euler’s theorem and Fermat’s theorem
are discussed in detail with suitable examples and related problems are also solved.
2.7 TERMINAL EXERCISE
1. If H is a sub group of G and a  h and if aHa-1 = { aha-1: h ∈ H }, show
that aHa-1 is a subgroup of G.
2. Show that a group of prime order is Cyclic
2.8 SUPPLEMENTARY MATERIALS
1. Algebra by Micheal Artin, Prentice Hall of India, New Delhi, 1994.
2. Lectures in Abstract Algebra Volumes I, II and III by N. Jacobson, D. Van
Nostrand Co., New York, 1976.
2.9 ASSIGNMENT
1. State and prove Lagrange’s Theorem.
2. State and Prove Euler’s Theorem.
3. If H and K are subgroups of a group G then prove that HK is a subgroup of
G iff HK = KH.
2.10 SUGGESTED READINGS / REFERENCE BOOK/ SET BOOK
Algebra – By Michael Artin – PHI Learning P Ltd
2.11 LEARNING ACTIVITIES
Students are requested to attend P.C.P. Classes and work out the problems
given in the lesson.
2.12 KEYWORDS
Group, Abelian group, Cyclic group, Sub group, Multiplication modulo m, Left
coset, Right coset, Lagrange’s Theorem, Euler’s Theorem, Fermat’s Theorem



21
LESSON – 3

HOMOMORPHISM
3.1 INTRODUCTION
In this Lesson, normal subgroup, coset multiplication, homomorphism, kernel
of homomorphism, Isomorphism, Cauchy’s theorem for abelian group, sylow’s
theorem for abelian group, automorphism, Inner homomorphism, center of a group
are discussed in detail with suitable examples and some problems are solved.
In the previous lesson, we have seen about the right cosets and left cosets of a
subgroup. We can easily see that for some subgroups, the notion of left coset
coincides with that of right coset, whereas for some subgroups these conceptions
may differ. The subgroups for which the right coset and left coset coincide are
distinguished and we will see about them now.
3.2 OBJECTIVES
Normal subgroup, coset multiplication, homomorphism, kernel of
homomorphism, Isomorphism, Cauchy’s theorem for abelian group, Taylor’s
theorem for abelian group, Automorphism, Inner automorphism, Center of a group
are discussed in detail.
3.3 CONTENTS
3.3.1 Normal subgroup
3.3.2 Coset Multiplication
3.3.3 Kernel of a Homomorphism
3.3.4 Theorem : Cauchy’s Theorem
3.3.5 Theorem on Abelian Groups
3.3.6 Automorphism
3.3.7 Inner Automorphism
3.3.8 Center of a Group
3.3.1 DEFINITION : NORMAL SUBGROUP
A subgroup N of G is said to be a normal subgroup of G if for every g in G and
n in N, gng-1  N.
It is clear from the definition that N is a normal subgroup of G if and only if
gNg-1 N for every g in G, where gNg-1 = {gng-1 : n N}.
3.3.1.1 Lemma
N is a normal subgroup of G if and only if gNg-1 = N for every g in G.
Proof
Let gNg-1 = N for every g in G.
Then gNg-1  N and hence N is a normal subgroup of G.
Conversely let N be a Normal subgroup of G.
22
1 1 1 1 1
Then for g  G, gNg-1  N and g Ng  g N g   N  g  G 
   
Consider N = g g-1 N gg-1 = g(g-1Ng)g1  gNg-1  N. Therefore gNg-1 = N. ■
3.3.1.2 Lemma
The subgroup N of G is a normal subgroup of G if and only if every left coset of
N in G is a right coset of N in G.
Proof
Let N be a normal subgroup of G. Then for every g, gNg-1 = N. Hence
(gNg )g = Ng. So gN = Ng. Hence the left coset gN is same as the right coset Ng.
-1

Conversely assume that every left coset of N in G is a right coset of N in G.


Thus for gG, gN, being a left coset, must be a right coset say Ng (gG). We claim
that this right coset must be Ng. Since Ng and Ng contain the element g in common
they are the same. Thus gN = Ng . So (gN)g-1 = N and hence N is a normal subgroup
of G. ■
3.3.1.3 Lemma
A subgroup N of G is a normal subgroup of G if and only if the product of any
two right cosets of N in G is, again a right coset of N in G. The subgroup N of G is a
normal subgroup of G if and only if every left coset of N in G is a right coset of N in
G.
First let us assume that N is a normal subgroup of G. Then for every g in G, we
have gNg-1 = N. Hence (gNg-1)g = Ng. Similarly we have g-1(gNg-1) = gN. So gN = Ng.
Hence the left coset gN is same as the right coset Ng.
Conversely let us assume that every left coset of N in G is a right coset of N in
G. Thus for any gG, gN, being a left coset, must be a right coset say Ng where
gG. We claim that this right coset must be Ng. Since gN and Ng contain the
element g in common they are the same. Thus gN = Ng. So (gN)g-1 = N and hence N
is a normal subgroup of G. ■
3.3.3.4 Lemma
A subgroup N of G is normal subgroup of G if and only if the product of any
two right coset of N in G is a right coset of N in G.
Proof
Let N be a normal subgroup of G. Let Na and Nb are two right cosets of N
where a and b are in G. Since N is a normal subgroup of G, aN = Na.
Hence (Na)(Nb) = N(aN) b = NNab = Nab=Nc where c = abG
Thus the product of any two right cosets Na and Nb is Nc which is another
right coset of N in G.
Conversely let us assume that the product of any two right cosets Na and Nb is
again a right coset Nc, where c G. We can take Nc = Nab, for some a, b G.
Therefore Nc = Nab = (Na)(Nb)
 (Na) (Nb) = Nab, therefore (Na)(Na-1)= Naa-1N=NN=N
23
 N(aNa-1) = N.
 for n, n  N, (ana-1)  N.  ana-1  N, as n N.  aNa-1 = N and hence N is
a normal subgroup. ■
Note :
It is clear that every subgroup of an abelian group is normal as the right coset
and left coset of the subgroup will obviously coincide.
3.3.2 COSET MULTIPLICATION
Definition 3.3.2.1
Let N be a normal subgroup of a group G. Let Na and Nb are two right cosets N
in G, where a,b  G. Then the product (Na) (Nb) = Nab.
The above said product is well defined. To show this, let Na = Na and
Nb = Nb. Then we must show that Nab = Nab. But Na = Na  a(a)-1  N and
Nb = Nb  b(b)-1  N.
Now (ab)(ab)-1 = ab(b)-1(a)-1 = an(a)-1, where n = b(b1)-1 N
= (a n (a-1a)(a)-1]  N
= (a n a-1 )[a(a)-1]  N
 (ab)(ab)-1  N
 Nab = Nab ( N is normal ana-1  N and a(a)-1  N ) ■
3.3.2.2 Theorem
G
If G is a group N is normal subgroup of G, then , the collection of all cosets
N
of N in G, is a group under coset multiplication.
Proof
G
Let = { Na : a  G}. It is closed under operation coset multiplication which is
N
a well-defined operation. Further this coset multiplication is associative.

G
For that, take three elements X, Y, Z  .
N

Then by definition X = Na, Y = Nb and Z = Nc some a, b, c  G.

(XY)Z = (NaNb)Nc =(Nab)Nc = N(ab)c=Na(bc)=Na(Nbc)= Na(NbNc) = X(YZ)

 Coset multiplication is associative.

G
Now N = Ne  and NNa = (Ne)(Na) = N (ea) = Na
N

G
 NNa= Na which shows that N acts as an identity in
N
24
G G
Finally, for any X = Na  , the inverse of X is Na-1  ,
N N
G
Thus becomes a group. ■
N
Note: (1). This group of cosets of N in G is called the Quotient group or factor
G G
group of G and is denoted by . (2). This Quotient group is abelian,
N N
whenever G is abelian
3.3.2.3 Lemma

 G  0(G)
If G is a finite group and N is a normal subgroup of G, then O = .
 N  0(N)
Proof
G 0(G )
If G is a finite group then will have precisely distinct cosets. Hence
N 0(N)
G 0(G )
O  = .■
N 0( N )
We shall now study the notion of Homomorphism, by which we mean a
mapping from one algebraic system to a algebraic system which preserves
structure.
3.3.2.4 Definition

A mapping  from a group (G, .), into a group ( G , * ) is said to be a


homomorphism if for all a, b  G,  (a.b) = (a) *  (b).
3.3.2.5 Example
If C denotes the additive group of complex numbers and R denotes the additive
group of real numbers, then the map  : C  R, given by  (x + iy) = x is a
homomorphism.
3.3.2.6 Example

Let G = Z be the additive group of integers and G = G. For x G, define  such
that  (x) = 2x. Then  is a homomorphism.
3.3.2.7 Example
Let G be the group of non-zero real numbers under multiplication of real
numbers. Let G = {1, -1} be the group under the same multiplication.

Let us define  : G  G by setting  (x) = 1 if x is positive and – 1 if x is


negative. Then  is a homomophism.
3.3.2.8 Lemma
Let G be a group and let N be a normal subgroup of G. Define the mapping 
G G
from G to by  (x) = Nx for all x in G. Then  is a homomorphism from G onto .
N N
25
Proof
Consider (xy) = Nxy = (Nx)(Ny) = (x)(y)   is a homomorphism.
G
For any X , it will be of the form Ny for some y  G.
N
Then by definition (y) = X.   is onto. ■
G
Note: This homomorphism  : G  , where (x) = Nx is called the natural
N
G
homomorphism of G onto .
N
3.3.3. KERNEL OF A HOMOMORPHISM
3.3.3.1 Definition
If  is a homomorphism of G into G , then the kernel of  denoted by K is
defined by K = { xG : (x) = e , where e is the unit element of G }.
Note: In the Example 1, (given earlier to illustrate the concept of group) the
kernel of homomorphism is the set of all complex numbers with real part zero. In
example 3, all positive real numbers form the kernel while, in the Example 2 the
kernel is trivial, consisting of the integer 0 only.
Hereafter, whenever we mention about a homomorphism G  G , let us take
multiplication ‘.’ as the operation for both the groups G and G . This assumption
does not mean that both groups G and G are groups under the same operation.
3.3.3.2 Lemma
If  is homomorphism of G into G , then

(i)  (e) = e where e is the unit element in G and e is the unit element in G .

  1
(ii) φ x 1  φ(x) for all x  G

Proof : (i) We know that (x). e =  (x),


and (xe) = (x)(e) ( is a homomorphism )
 (x) = (x)(e)
 (x) e = (x)(e)

 e = (e) ( by cancellation law )


(ii) Now (e) = e   (xx-1) = e

  (x). (x-1) = e
  (x-1)=[(x)]-1 ■
Lemma 3.3.2 tells us that the kernel of any homomorphism will not be empty,
as it certainly contains the unit element of G. The next lemma tells, more about
this kernel.
26
3.3.3.3 Lemma
Let  be a homomorphism of G into G with kernel K. Then K is a normal
subgroup of G.
Proof
First let us prove that K is a subgroup. To show this it is enough if we prove
that K is closed under multiplication and has inverse in it for every element
belonging to K. If x, y  K, then  (x) = e =  (y), where e is the unit element in G .

Hence  (xy) = (x) (y) = e e  e . Hence xy  K, so K is closed. If x  K, let us


show that x-1 is also  K.
We have by lemma ,  (x-1) = [ (x)]-1 = (e)1  e . Thus x-1 is in K. So K is a
subgroup. Now we will show that K is a normal subgroup also. To prove this one
must establish that for any g in G, and for any k  K, gkg-1K. So we have,
 (gkg-1) = e whenever (k) = e .
Now (gkg-1) = (g) (k) (g-1)
= (g) e (g-1)
= (g) (g-1)
= (g) [(g)]-1 = e .
Hence K is a normal subgroup of G ■.
3.3.3.4 Definition
A homomorphism : G  G is said to be an isomomorphism if  is one-to-one.
3.3.3.5 Lemma
A homomorphism  : G  G , where G and G are two groups, is an
isomorphism if and only if its kernel contains only the unit element of G.
Proof
Let : G  G , be an isomorphism. Then we should show that its Kernel contains
only its unit element. Let us assume that x ( e) is in the Kernel. Then (x) = e , and  (e)
= e is the element in G . Since x  e, is not one-to-one so that  will not be an
isomorphism which is a contradiction. Hence kernel must contain only the identity
element e. Conversely, let us assume that the kernel is trivial. We must show that 
is one-to-one.
Let (x) = (y).
Now (xy-1) = (x)(y-1)
= (x) [(y)]-1
= (y) [(y)]-1 ( (x) =  (y)]
= e
27
 xy-1 = e
 x = y.
  is an isomorphism. ■
3.3.3.6 Theorem
Two groups G and G are said to be isomorphic if there exists an isomorphism
of G . In this homomorphism, the kernel K is a normal subgroup of G such that
G
the quotient group is isomorphic with G . Conversely, if K is any normal
K
G
subgroup of G then there exists a suitable homomorphism of G onto such that
K
the kernel of this homomorphism is precisely K.
Proof
It follows from the lemma that, the kernel K of the homomorphism  of the
group G on to the group G is a normal subgroup of G,. Therefore it is possible to
G
talk about the quotient group , whose element are the cosets Kx on K in G.
K
G
Consider the mapping :  G given by σ(Kx)  φ(x), for any x  G. We claim that
K
G
the mapping is the desired isomorphism of onto G . Since  is surjective ,  is
K
also surjective. Also  is well-defined and one-to-one as seen by the following
implications:
Kx = Ky (for x, y  G)
 xy-1  K
 (xy-1) = e (  K is the kernel of  )

 (x) ((y))-1 = e (  is homomorphism )

 (x) = (y)
 (Kx) = (Ky) (by definition of ).
Finally  is a homomorphism. Since  ((Kx) (Ky)) =  (Kxy) = (xy) = (x) (y) =
G
(Kx) (Ky). Thus  is an isomorphism of onto G and the claim is established.To
K
G
see the converse, consider the natural mapping  : G  given by v(x)  Kx .Clearly
H
v is surjective and since, for x, y  G, v(xy) = Kxy = (Kx) (Ky) = v(x) v(y), v is a
G
homomorphism of G onto . Moreover the kernel of this homomorphism v is K,
K
28

 G 
since x is in the kernel of v iff vx = K  the identity element of    which is possible
 K 
iff Kx = K, which again is possible iff x  K. Thus the converse is also established. ■
Note: For xG, since (•v)(x) = (v(x)) = (Kx) = (x), we find that •v =  yields
the factorization of the given surjective homomorphism  as a product of the
natural homomorphism v and an isomorphism . This fact is exhibited in the
following commutative diagram.


G
G

v
6

G
K

G 
When the mapping  is not surjective the same p of will show that   is
K 
isomorphic with  (G), the image of G under . Before proving the next theorem, viz.
Cauchy’s Theorem for Abelian groups, let us solve some (five) problems involving
the notions of normal sub-groups quotient groups and group homomorphism.
Problem 1
If N is a normal subgroup of a group G and H is any subgroup of G, prove that
NH is a subgroup of G.
Solution
Since N and H are subgroups of G, to prove that NH is a subgroup of G, it is
enough if we show NH = HN, (in view of a theorem proved in Lesson 2). Now N is a
normal subgroup of G and so for any gG, Ng = gN. Hence, in particular, for hH,
Nh = hN which immediately implies that NH = HN, as desired. Hence the problem is
solved. ■
Note: If H is also normal in G, we can show that NH is also a normal subgroup
of G.
Problem 2
Let A and B are normal subgroups of a group G such that AB = {e}. Prove
that, ab = ba for a  A, b  B,
Solution
To prove ab = ba, it is enough if we show that a-1 b-1 ab = e.
29
For this reason the element a-1b-1 ab is called the commutator of a and b.
Since AB = {e}, to prove a-1b-1 ab = e, it is enough if we show that a-1b-1 ab  AB.
Since a  A and A is normal in G, b -1 ab  A and a-1  A, A being a subgroup, so
that a-1b-1 ab = a-1(b-1ab)  A. Again, since bB, b-1  B and since B is a normal
subgroup of G, a-1b-1a  B so that a-1b-1 ab = (a-1b -1a)bB. Thus a-1b-1 ab  AB={e},
and hence a-1b -1 ab=e  ab=ba for all aA and bB. ■
Problem 3
If a group G has no proper non-trivial subgroup, then show that G must be a
finite cyclic group of prime order.
Solution
First we show that G must be a cyclic group and it can be generated by any
element a of G, provided a  e. Then, we will show that G is necessarily finite and is
of prime order. Let a  G and a  e and so H is a non-trivial subgroup of G. Thus
the group G must be cyclic and can be generated by any element a of G, with a  e.
Now we claim that G cannot be an infinite cyclic group. For if a is a generator of G,
then no other power of a other than a and a -1 can generate G. This means, that the
subgroup H = (am), viz. the cyclic subgroup generated by a m where m = 1 or m = – 1
will be a proper, non-trivial subgroup. This is impossible as G has no proper, non-
trivial subgroup. This is impossible as G has no proper, non trivial subgroup. Thus
G must be a finite cyclic group of order, say, n. We claim that this order n must be
a prime, for otherwise, if d is a (positive) proper divisor of n and G = (a) then the
cyclic subgroup generated by ad will be a proper, non-trivial subgroup of G. [This is
 
because, o a d  n
d
 n ]. Hence the problem is solved. ■ (The fact proved in this
problem will be used in the proof of Cauchy’s Theorem for Abelian groups ).
Problem 4.
If a group G has an endomorphism  : G  G defined by  (g) = g for g  G, prove
that G must be abelian (A homomorphism of a group G into itself is called an
endomorphism of G)
Solution
If a, b  G, then  (ab) =  (a)  (b),  being an endomorphism of G. This means
that (ab)2 = a2 b2, and hence we can get ba = ab. Hence G is abelian. ■
Similarly it can be shown that if G has an endomorphism  : G  G defined by
 (g) = g-1ng then G is abelian. Here it may be noted that such an endomorphism of
G, if exists, is actually an automorphism of G, i.e. an isomorphism of G onto G.
PROBLEM 5
NM N
If N and M are normal subgroups of G, prove that  .
M NM
The result of this problem is some times referred to as the Second
Isomorphism theorem for groups, the first isomorphism theorem being the
fundamental Homomorphism theorem for groups
30
Solution
Since N and M are normal subgroups of G, NM = MN and NM is a (normal)
subgroup of G and hence it is a group . Since M is a normal subgroup of G and
NM
since M  NM (obviously) M is a normal subgroup of the group NM,  is a
M
quotient group, whose elements are the cosets (nm)M, where n  N, m  M and
since m  M, these cosets can be taken as nM with n  N.
NM
Now, consider the mapping : N  defined by (n) = nM, for n N. Since
M
NM
any element in is a coset of the form nM, with n  N, we find that this
M
mapping is a well defined surjective mapping. Further, if n1, n 2  N then 
(n1n 2) = (n1n2) M = (n1M) (n2M) =  (n1) (n2) so that this mapping  turn to be a
(surjective) homomorphism. If n is in the kernel of , then n    (n) = M, the
NM
identity element of  nM = M  n NM. Thus K = NM and so, by the
M
N NM
fundamental theorem of homomorphism , we get that  .■
NM M
3.3.4 THEOREM : CAUCHY’S THEOREM
(Cauchy’s Theorem for Abelian Groups). Suppose G is a finite abelian group
and p divides O(G), where p is a prime number, then there is an element a ( e) in G
such that an = e.
Proof
Let us prove the theorem by the method of induction. We assume that the
theorem is true for all abelian groups having fewer elements than G. Then we will
show that the theorem is true for G also, if G has got more than one element.
If G has no proper subgroups (other than identity subgroup), then G must be a
cyclic group of prime order. This prime must be p, since G has got p elements. Now
we can take any one element a  e in G such that ap = a0(G) = e. Hence the theorem.
If G has no proper subgroups, we have proved the theorem. So let us assume
that G has a proper nontrivial subgroup N of G. If P|O(N), by induction hypothesis,
there is an element b  N, b  e such that bn = e. Since b, n  G, b  N, is a normal
G G O (G )
subgroup and is a group. Moreover O   = . Since p does not divide O(N)
N N O( N )
G G
and p|O(G), p| O   . Further   < O(G). So by our induction hypothesis, there
N N
G
is an element Nx  such that (Nx)p = N. Since Nx  N and Nxp = N. we have xp  N
N
and x  N. Now xpO(N) = e. That is xO(N)p = e. Let c = xO(N). Then cp = e. In order to show
that e is the required element we have to show that x  e. If x = e, then xO(N) = e and
31
so (Nx)O(N) = N. We already have (Nx)p=N and p does not divide O(N). Therefore
(Nx) = N and hence x  N, contradicting our assumption that x  N. Hence x  e and
the theorem is proved.■
3.3.5 THEOREM : ABELIAN GROUPS
(Sylow’s theorem for abelian groups). If G is an abelian group of order p and
if p is a prime number such that p  |O(G), p+1 does not divide O(G) then G has a
subgroup of order p .
Proof
If  = 0, the result is trivial. If   0, then definitely p divides 0(G). By the
previous theorem there is an element a (e) in G such that ap=e. Let
S = { x  G: xpn = e, for some integer n}. Clearly S contains not only e but also a. Let
us show that S is a subgroup of G. Since G is finite it is enough if we show that S is
multiplicatively closed. If x, y  S, then there exist integers n and m such that
pn pm p(n+m) p(n+m) p(n+m)
x = e and y = e and hence (xy) =x y

( G is abelian (xy) k  x k .y k for any interger k )

x  . y 
pn
pm
pm
pn m
 ep . ep  e
n

Hence xy  S, So S is a subgroup.
Out next aim is to find O(S). We claim that O(S) = p with  is an integer and
O(M)=  < . If some prime q | O(S), q  p, then by the previous theorem, there is an
pn
element c  S, c  e such that c = e for some integer n. Since pn and q are
relatively prime, we can find integers ,  such that q + p n = 1. Hence c = c1 =

c q  pn  c q   
 
 cp
n 
 e , contradicting c  e. Hence q does not divide O(S). So O(S)
= P with 0 <  < , since by lagrange’s theorem O(S) | O(G) and p+1 does not divide

O(G). Hence  < ,

 G  O(G ) G
Suppose  < , then O   . Since p |O(G) and O(S) = p , p | O   .
 S  O( S ) S
G
Hence by Cauchy’s theorem, there is an element Sx  , for x  G satisfying
S
Sx  S and (Sx)P = S. But S = (Sx)p = Sxp and so xp  S.

O(S)  P +1
Consequently e = (xp) = (xp)p = x . Therefore x  S.
Hence Sx = S contradicting the fact Sx  S, Hence  =  and so S is the
required subgroup of order p  ■
Corollary 3.5.1

If G is an abelian group of order O(G) and p | O(G) and p+1 does not divide
O(G), then there is a unique subgroup of order p.
32
Proof

Let T be another subgroup of order p, other than S which we have got from
the theorem. Since G is abelian ST = TS, so that ST is a subgroup of G.
p α
Hence OST   O(S) O(T)  P , P since S  T, O(S  T)  p α . So O(ST) = pv, v > . Since ST is a
O(S  T) O(S  T)
subgroup of G, O(ST)|O(G). Thus pv divides O(G), contradicting the fact that p is
the largest power of p which divides O(G). Hence S = T. ■
If there is a homomorphism of a group G onto a group G with kernel K, it is
possible to set up a one-to-one correspondence between subgroups of G which
contains K and all subgroup of G. This fact and other connected results are given in
the following lemma.
Lemma 3.5.2
Let  be a homomorphism of G onto G with Kernel K. Let H be a subgroup of
G . Let H = { xG:  (x)  H }. Then H is a subgroup of G and H  K. If H is normal
in G , then H is normal in G. This association H H is a one-to-one
correspondence from the set of all subgroups of G onto the set of all subgroups of
G which contains K.
Proof

First let us show that H is a subgroup. Let x, y  H. Then (x), (y)  H . Hence
(x) (y) = (xy)  H . So xy  H. So H is closed in G. Moreover if x  H, then
 (x)  H . Hence [ (x)]-1  H , so that  (x-1)  H (i.e.) x-1  H. So H is a subgroup.
Clearly K  H. So H is a subgroup of G, containing K.
Now let us show that if H is normal, then H is also normal. Let g  G and
h  H. Let us show that ghg-1  H whenever H is normal.
(ghg-1) = (g) (h) [(g)]-1  H since H is normal, So ghg-1  H. Hence H is
normal too.
Conversely let L be a subgroup of G containing K. Let
L = {x  G : x   (l ), l  L}. Now we have to prove that L is a subgroup. For that let
x, y  L . Then there exist l1, l2  L such that  (l1) = x and (l 2 )  y. Since l1, l2  L,

 (l1l2) = (l1)(l2) = x y  L. Hence L is closed. If now x  L , then there exists l  L

such that  (l) = x . Since l  L, l-1L, therefore (l-1) = [ (l)]-1 = ( x ) 1  L. Hence L is


a subgroup of G.

If we define T = { yG: (y) L }, then our question is whether T=L. Clearly


L  T, for if lL, (l)  L . Hence lT. Further for any element tT, (t) is in L ,
Hence by the definition of L, there exists an l in L such that (t) = (l).
33
Thus (t l -1) = (t) [(l)]-1 = e, Hence(t l -1) K  L. Thus t is in L, l =L [since lL].
Hence T L. So T= L. Hence there is a one-to-one correspondence between the
subgroups of G and the subgroups of G containing K, the one-to-one
correspondence being given by H  H .■
Theorem 3.5.2
Let  be a homomorphism of G into G with kernel K, and let N be a subgroup

G
G G G K
 
of G, where N  x  G :  ( x )  N . Then 
N N
equivalent to 
N N
.

K
Proof

G
Since N is a normal subgroup of G , we can talk of the quotient group . Let
N
G
us define the map  : G  with the help of the given homomorphism : G G .
N
For any g  G, let us define (g) = N (g). Clearly  is onto.
Consider (g1g2) = Nφg1g 2   Nφg1 φg 2 

= Nφg1 Nφg 2 

= φg1  φg 2  for all g1 , g2  G

This implies that  is a homomorphism.


Let we find T which is the kernel T with respect to . We claim T = N,
By definition T = { g  G :  (g) = N}

So, g  T  (g) = N  N (g)  N

  (g)  N

 g N
Therfore T = N.

G G G
Hence for the homomorphism  : G  , the kernel is N. Hence  .
N N N
G
Since K is the kernel of the homomorphism  : G  G , we have  G and
K
similarly restricting our attention to
G
G K
: N  N where  (n)  N where n  N and henceN  G , we have  ■
N N
K
34
3.3.6 AUTOMORPHISM
Definition 3.6.1
An Automorphism of a group G, is an isomorphism of G onto itself.
For any group G, there exists atleast one automorphism namely the identity
mapping.
Lemma 3.6.2
If G is a group, then A(G), the set of all automorphisms of G is also a group,
under the composition of mappings.
Proof
We have already shown in the first lesson that A(S), the set of all one-to-one
mappings of a set S onto S is a group under composition. Hence A(G) is also a
group, so A(G)  A(G). Actually A(G) is going to be the subgroup of A(G). So it is
enough if we show that A(G) is closed under composition and that inverse of any
element in A(G) is in A(G) itself.
Let T1, T2  A(G). We want to show that (T1.T 2 )  A(G). That is we want to show
that (T1.T 2) is an automorphism. Let we prove that (T1.T2) is a homomorphism. So,
let x, y  G. Then T1(xy) = T1(x).T1(y) and T2(xy) = T2(x).T2(y)
Hence (T1.T 2)(xy) = T 1[T2(xy)]
= T 1[T2(x)T 2(y)]
= T 1(T2(x))T1(T 2(y))
= (T 1.T2) (x) (T1.T 2) (y)
Hence (T1.T 2) is a homomorphism.
Now let us show that, if T 1  A(G), (T1)-1  A(G).
Since T1 is one-one and onto, (T1)-1exists.
Now let us show (T1)-1is a homomorphism.
For any x, y  G,
Consider T1[(T1)-1(xy)] = T1[(T1)-1(x)(T1)-1(y)]
= [ T1(T1)-1 (x)] [T1(T1)-1 (y)]
= xy
 [(T1)-1(xy)] = (T1)-1(x))(T 1)-1 (y))
 (T 1)-1is a homomorphism.
Hence (T1)-1 A(G).
A(G) is a group under composition of mappings. ■
3.3.7 INNER AUTOMORPHISM
Definition 3.7.1
Let G be a group, For g  G, let us define
Tg: G  G by T g(x) = g-1xg, for all x  G.
35
Now we will show that Tg is an automorphism of G corresponding to g  G.
First let us show that T g is onto. For a given yG, there exists xG, such that
x = gyg-1G. Then T g(x) = g-1xg = g-1(gyg-1)g = y. Hence T g is onto.
To prove that T g is a homomorphism
For a,bG, consider T g(ab) = g-1(ab)g
= g-1(agg-1b) g
= (g-1ag) (g-1bg)
= Tg(a)T g(b)
 T g is a homomorphism
To prove that T g is one-to-one
Let Tg(a) = Tg(b)  g-1ag = g-1bg  a = b
 Tg is one-to-one
 Tg is a one one and on to homomorphism of G on to G
 Tg is an automorphism on G. This is called Inner automorphism of G
Note: Hence for every g  G, we can have an Inner automorphism T g. Let I(G)
denote the set of all inner automorphisms of G.
Our next aim is to show that this I(G) is a group under usual composition of
maps, and that I(G) is a normal sub-group of A(G).
First we claim that Tg.Tb = Tbg for any Tg, Tb  I(G).
For any xG, consider (T g.Tb)(x) = (TgTb(x)) = Tg(b-1xb) where hG
= g-1(b-1xb) g
= (bg)-1x(bg)
= Tbg(x) is true for all x
Tg.Tb = T bg
I(G) is closed.
Since I(G)  A(G), it is enough, if we show that inverse for every element in I(G)
exists in I(G) itself. Clearly Tg is the identity element in I(G) where e is the identity
element in G. Let Tg  I(G), then the inverse of T g is (Tg)-1 I(G) as Tg (Tg)-1= T g-1g = Te.
Hence I(G) is subgroup of A(G)
To show I(G) is normal in A(G)
Let Tg  I(G) and   A(G) .
For x  G, consider (-1.T g. )(x) = (-1.Tg )((x))
= -1 (Tg((x))
= -1(g-1(x)g)
= (-1(g-1))(-1((x)) -1(g)
= (g)-1x-1(g)
= T-1(g)(x).
Hence (-1.Tg .) =T -1(g)  I(G) (-1(g) G).
36
3.3.8 CENTER OF A GROUP
Definition 3.8.1
The center of G, is the set Z(G) of elements in G, such that it commute with
every element of G. Therefore Z(G)= { xG: xy = yx for every yG}
G
Note: Z(G) is a normal subgroup of G. So, we can discuss , which is a
Z(G )
quotient group
Lemma 3.8.2
G
Let G be a group. Then I(G)  where I(G) is the set of automorphisms of G
Z(G)
and Z(G) is the centre of G.
Proof
Let  : G I(G) defined by (g) = Tg-1 for any gG.
Consider (gh) = T (gh)-1 = T(h-1 g-1)= Tg-1 Th-1 = (g) (h).
Therefore  is a homomorphism.
Let K be the kernel of the homomorphism .
Let gK. Then (g) = Te, the identity of I(G) (i.e.) Tg-1 = Te, This means that for
x  G, (T g )-1(x) = x. But Tg-1(x) = gxg-1. Hence gxg-1 = x which implies that xg = gx,
Hence g commutes with every element x of G. So g  Z(G), whenever g  K. So
k  Z(G). If g Z(G) then for any x  G. Tg(x) = g-1xg = g-1gx = x = Te(x)
for every g  Z(G)
Hence T g = Te and hence g  K. Hence Z(G)  K.
Since we have already shown that K  Z(G) we have Z(G) = K.
G
Hence by theorem, I(G)  . ■
Z(G )
Lemma 3.8.3
Let G be a group and  an automorphism of G. If a  G is of order O(a) > 0
then 0((a)) = 0(a).
Proof
Let O(a) = n
Then [(a)]n = (an) [ is an automorphism] = (e) = e.
Therefore O((a)) < n. If [(a)]m = e for some m and 0 < m < n, then
[(a)]m = (am) = e. Since  is an automorphism and am=e we arrived at a
contradiction to our assumption that O(a) = n and m<n. Therefore O((a)) cannot be
less than n. Therefore 0((a)) = n = O(a). ■
3.4 REVISION POINTS
Normal sub group, coset multiplication, Homorphism, Kernel of a
homomorphism, Isomorphism, Cauchy’s theorem for abelian group, Sylow’s
theorem for abelian group, automorphism, Inner automorphism, Centre of a group.
37
3.5 INTEXT QUESTION
1. Prove that N is a normal subgroup of G it and only it gNg-1  N for every
g in G
2. Prove that a sub group N of G is a normal subgroup of G if and only if the
product of two right cosets of N in G is again a right coset of N in G
3. If G is a group, N is a normalsubgroup of G, then prove that G/N the
collection of all cosets of N in G, is a group under coset multiplication.
4. If N is a normal subgroup of a group G and H is any subgroup of G, prove
that NH is a subgroup of G.
5. If a group G has no proper non- trivial subgroups, show that G must be a
finite cyclic group of prime order.
NM N
6. If N and M are normal subgroup of G, prove that 
M N M
7. If  is a homomorphism of G into G with Kernel K, and if N be a subgroup
G G/K
of G where N ={ x G:  (x)  N }, then prove that 
N N/K
G
8. If G is a group, then prove that I(G)  , where I(G) is the set of
Z (G )
automrphisms of G and Z(G) is the centre of G.
9. If G is a group and  be an automorphism of G and if O(a)> zero, then
Prove that O(  (a)) = O(a)
3.6 SUMMARY
Normal subgroup, Coset multiplication, homomorphism, Kernel of
homomorphism, Isomorphism, Cauchy’s theorem for abelian group, Sylow’s
theorem for ablian group, automorphism, Inner automorphism, Center of a
group.
3.7 TERMINAL EXERCISES
1. If A and B are normal subgroups of a group G such that A  B = C then
prove that for a A, b  B, ab = ba
2. If a group G has no proper nontrivial subgroup show that a must be a
finite cyclic group of prime order
3.8 SUPPLIMENTARY MATERIALS
1. Algebra by Micheal Artin, Prentice Hall of India, New Delhi, 1994.
2. Lectures in Abstract Algebra Volumes I, II and III by N. Jacobson, D. Van
Nostrand Co., New York, 1976.
3.9 ASSIGNMENT
1. State and prove Cauchy’s Theorem for abelian group
38
2. State and prove Sylow’s Theorem for abelian group
3. State and prove first fundamental theorem of homomorphism
4. If G is a group, then prove that A(G), the set of all automorphism of G is
also a group under composition of mappings.
3.10 SUGGESTED READINGS / REFERENCE BOOK / SET BOOK:
Algebra – By Michael Artin – PHI Learning P Ltd.
3.11 LEARNING ACTIVITIES
Students are requested to attend the P.C.P Classes and work out the problems
given in the lesson.
3.12 KEYWORDS
Normal subgroup, Coset multiplication, homomorphism, Kernel of
homomorphism, Isomorphism, Cauchy’s Theorem for abelian group, sylow’s
theorem for ablian group, Automorphism, Inner automorphism, Center of a
group

39
LESSON – 4

CAYLEY’S THEOREM
4.1. INTRODUCTION
In this Lesson, Cayley’s theorem is discussed in detail and related problems
are also solved.
From the following Cayley’s theorem, one can see, how any group can be
realised as a subgroup of A(S), the permutation group for a suitable set S.
4.2. OBJECTIVE
Cayley’s Theorem, related lemmas are explained in detail and some related
problem are solved.
4.3. CONTENTS
4.3.1 Theorem (Cayley)
4.3.1 Theorem (Cayley)
Every group is isomorphic to a subgroup of A(S) for some appropriate S.
Proof
Let G be a group and let us take S as the set of all elements in G.(i.e S = G).
Now let us define Tg : SS by tg(x) = gx for every x  S and gG
Since for any yS, there exists (g-1y)  S such that tg (g-1y)= g(g-1y)= y and
hence tg is onto.
Let tg(x) = tg(y)  gx = gy  x = y  tg is one one and hence tg is one-to-one
Hence tg  A(S). Moreover for g, h  S, (tg.th)(x) = tg(hx) = g(hx) = (gh)x = tgh(x)
for every xS, Hence (tg.th)= tgh.
Now let us define  : G  A (S) by  (g) = tg.
Consider (gh)= tgh = tg th = (g)(h)   is a homomorphism.
Let us find the kernel K of the homomorphism .
If g0 K, then (go) = tg = I (identity map on S).
o

Hence tg (e) = e ( since tg = I )


o o

But tg (e) = goe = go


o

Hence go = e, Hence K = {e}.


So  is an isomorphism of G into A(S).
Therefore G is isomorphic to a subgroup of A(S) ■
If we look into the order of A(S) = A(G), we find it will have O(G)! elements
provided G is a finite group. So by Cayley’s theorem we have an isomorphism from
a group G into the group A(S) of O(G)! elements. The size of A(S) is quite large
compared to that of G. Our next aim is to find an isomorphism of G into a group
smaller than A(S). Our next theorem shows the way.
40
Theorem 3.1.1
If G is a group and H, a subgroup of G and if S is the set of all left cosets of H
in G, then there is a homomorphism  of G into A(S) with the kernel of  as the
largest normal subgroup of G which is contained in H.
Proof

For any g  G, consider the map tg : SS by tg(xH) = gxH for any x  tg.
To prove that is tg on to:
It is very obvious that for any (yH)S in the co-domain, there exists (g-1yH)S
in the domain such that tg (g -1yH) = g (g-1yH) = yH and hence tg is on to.
To prove that tg is one one also.
Assume that tg(xH) = tg(yH), where (xH),(yH)S
 gxH = gyH
 (gy)-1gxH, (Since aH= bH iff b -1a  H)
 (y-1g-1gx) H.
 ( y-1x)  H
 xH = yH.  tg is one one and hence tg is one –to-one
Moreover for any xH S and g1, ge  G and
 
(t g1 .t g 2 )(xH)  t g1 t g 2 (xH)  t g1 (g 2 xH)  g1 (g 2 xH)  t g1g 2 (xH)
Hence tg1g = tg1. tg2
2

Now let us define a map : G  A(S) by setting A(g) = tg for any gG. Since
(gh) = tgh = tgtg = (g)•(h),  is a homomorphism. Let K be the Kernel of .
If go  K then  (go) = tgo is the identity mapping on S. Since S is the set of all left
cosets of H, the above  tgo(aH) = aH for every sG. But by the definition of tgo,
tgo (aH) =goaH. Hence we get aH = goaH for every aH.
Hence K = {bG : bxH = xH for all xH}. Since K is the kernel, it is a normal
subgroup of G. We claim K  H. For if bK, bxH = xH for every x in G and in
particular for the identity e, Hence beH = eH bH = H  b H. Therefore K  H
Now in order to show that K is the largest normal subgroup of G contained in
H, we must show that if N is the normal subgroup contained in H then N  K.
Let N be a normal subgroup of G contained in H. If n  N and x  G then
x-1nx  N  H  (x-1nx)H=H. Thus nxH = xH for all x in G. Hence n  K by the
definition of K. Hence N  K.
So K is the largest normal subgroup contained in H. ■
Lemma 3.1.2
If G is a finite group and H  G, is a subgroup of G such that O(G) does not
divide IG(H)! where IG(H) is the index of H in G (that is the number of left cosets of H
in G), then H must contain a nontrivial normal subgroup of G.
41
Proof
Since O(G) does not divide iG(H) = the number of elements in A(S) (where A(S) is
the same A(S) mentioned in the previous theorem), by Lagrange’s theorem A(S) can
have no subgroup of order O(G). But A(S) contains (G) where (G) = { (g) : gG}.
Hence  cannot be an isomorphism. That is, the kernel K of the previous theorem is
nontrivial. Hence H must contain a non-trivial normal subgroup K of G.
Example 3.1.3
Let G be a group of order 36 and suppose G has a subgroup H of order 9, then
iG(H) = 4, Hence 4! = 24 < 36, Hence A(S) has got 24 elements. But G has got 36
elements. Hence G must have a non-trivial normal subgroup.
Example 3.1.4
Let H be a group of order 99 and suppose that H is a subgroup of G of order H,
then iG(H) = 9 and since 99 does not divide 9!. G contains subgroups other than (e).
Solved Problems
1. If G is a group and H is a subgroup of Index 2 in G, prove that H is a
normal subgroup of G.
Solution
If a  H, then H = aH = Ha
If a  H, ah is a left coset different from H
Hence HaH = 
Since index of H in G is 2, HaH = G
Hence aH = G – H
Similarly Ha = G – H
Ha = aH
 H is a normal subgroup of G. ■
2. Show that the intersection of two normal subgroups of G is a normal
subgroup of G.
Solution
Let H and K be two normal subgroups of G.
Now HK   since cHK
Let a, bHK  a, bH and a, bK
 ab -1H and ab -1K
 ab -1HK
 HK is a subgroup of G.
Let gG and xHK  gG and xH & xK.
Since H and K are normal, gxg-1  H,
Therefore gxg-1H, gxg-1 K
 HK is a normal subgroup of G. ■
42
3. If H is a subgroup of G and N is a normal subgroup of G show that HN is
a normal subgroup of H. Show by an example that HN need not be
normal in G.
Solution
Let xHN, and gH.
xN and g H  gxg-1  N (N is a normal subgroup)
Also gxg-1 H ( gH and x  H)
Hence gxg-1 HN
HN is a normal subgroup of H.
1 2 3  1 2 3 
LET G = S3 TAKE N = G AND H =  ,  
1 2 3  1 3 2 
HN = H which is not normal in G. ■
4. Suppose H is the only subgroup of order O(H) in a finite group G. Prove
that H is a normal subgroup G.
Solution
Let O(H) = m
To prove that gHg-1 is also a subgroup of G, where g  G.
Consider e = geg-1  gHg-1. Therefore gHg-1 
Let x, y  gHg-1. Then x = gh1g-1 and y = gh2g-1
xy-1 = (gh1g-1) (gh2g-1)-1
= (gh1g-1) (gh2-1g-1)
= g(h1h2)g-1
 gHg-1
gHg-1 is a subgroup of G.
To prove that 0(H) = 0(gHg-1) = m
Consider f: HgHg-1 defined by f(h) = ghg-1
Now f(h1) = f(h2)  gh1g -1 = gh2g-1
 h1 = h2
 f is one one.
Let x = ghg-1  gHg-1. Then f(h) = ghg-1= x. Therefore the pre image of
x is h under f and hence f is on to  f is one-to-one.
Hence 0(H) = 0 (gHg-1) = m. But H is the only subgroup of G of order
m. gHg-1 = H  H is normal in G.■
5. Let H be a subgroup of G, let N(H) = { gG/gHg-1 = H }. Prove that
(i) N(H) is a subgroup of G, (ii) H is normal in N(H),(iii) If H is a normal
subgroup KG, then K  N(H).(iv) H is normal in G iff N(H) = G.
43
Solution:
(i). Let a, b  N(H)
 aHa-1 = H and bHb-1 = H
Now bHb-1 = H  b-1 (bHb-1) b=b -1Hb
 H = b-1Hb
Consider (ab-1) H(ab -1)-1 = a(b -1Hb) a-1
= aHa-1
=H
 ab-1  N(H)
Thus a,b  N(H)  ab -1 N(H)
Hence N(H) is a subgroup of G.
(ii). Let h H. Since hHh-1 = H, h  N(H)
Thus H  N(H). Hence H is a subgroup of N(H)
Let x  N(H)  xHx-1 = H
 H is a normal subgroup of N(H)
(iii). Let k  K
Since H is a normal subgroup of K, kHk-1 = H  kN(H)
Thus k  K  k  N(H)  K  N(H)
(iv). Let H be normal in G. Let x  G. Then xHx-1 = H  x  N(H).
Thus x  G  x  N(H) G N(H)
But N(H)  G. Hence G = N(H)
Let N(H) = G. We have to prove that H is normal in G.
x  G  x  N(H) ( N(H) = G).
 xHx-1 = H. x  G.
 H is normal in G. ■
6. If N and M are normal subgroups of G, then prove that NM is also a
normal subgroup of G.
Solution:
Since N is a normal subgroup and a normal subgroup commutes with
every element, we have NM = MN  NM is a subgroup of G.
Let g  G and nm  nM. Then n  N and m M
Consider g(nm)g-1 = (gng-1) (gmg-1)
 NM [ N is normal  gng-1  N & M is normal  gmg-1  M)
Hence NM is a normal subgroup of G. ■
7. Let N and M are normal subgroups of G and MN = {e}. Show that every
element of N commutes with every element of M.
44
Solution :
Let n  N & m  M
To prove that nm = mn
Consider the element of nmn-1 m-1
Since N is normal, mn-1 m-1  N. Also n  N  nmn-1m-1N.
Since M is normal, nmn-1  M. Also m-1  M  nmn-1m-1M.
Thus nmn-1m-1  NM = (e)
 nmn-1m-1 = e
 (nm) (mn)-1 = e
 nm = mn ■
8. If a cyclic group N of G is normal in G, show that every subgroup of N is
normal in G.
Solution:
Let N = <a>. Let H be a subgroup of N. We know that every subgroup of a
cyclic group is cyclic. Hence H is cyclic and am will be a generator of H
where m is the least positive integer such that am  H.
To prove that H is normal in G.
Let x  G & h  H
Then h = (am)n where n is some integer.
Consider xhx-1 = x(am)n x-1 where n is some integer.
= x(an)m x-1
= (xanx-1) (xanx-1) …… to m factors
= (xanx-1)m
Now a is a generator of N an  N
Also N is normal in G  x  G, an  N  xanx-1 N
Hence xanx-1 = an for some integers (a is generator of N)
 xhx-1 = (an)m = (am)n H  xhx-1  H
(am is a generator of H and x  H, h  H  xhx-1  H)
Hence H is a normal subgroup of G. ■
9. Prove that the centre Z of a group G is a normal subgroup of G.
Solution :
We know that Z = { z  G : zx = xz,  x  G}
Let z1, z2  Z. Then z1x = xz1 and z2x = xz 2 for all x  G
Consider z2 x = xz2  x  G
 z 2-1 (z2x) z 2-1 = z2-1 (xz2)z 2-1
 xz2-1 = z 2-1x  x  G
45
 z 2-1  Z.
 (z1z2-1) x = z1 (z2-1x)
= z1 (xz 2-1)
= (z 1x) z1-1
= (xz1) z2-1
= x(z1z2-1)
 z 1z 2-1  Z.
Thus z1, z2  Z  z1z 2-1  Z
Z is a subgroup of G.
To prove that Z is a normal subgroup of G
Let x  G, z  Z
xzx-1 = (xz)x-1 = (zx) x-1 = z  Z
Z is normal subgroup of G. ■
10. Let G be a group and G be the commutator subgroup of G prove that

G
(i) G is normal in G and (ii)   is abelian.(iii) If N is normal subgroup of
 G 
G
G, then is abelian iff G  N.
N
Solution:
Let U = { xyx-1y-1 : x, y  G }. If G is the commutator subgroup of G, then
G is the smallest subgroup of G containing U.
(i) Let x  G & y  G
Then x-1yx = yy-1 (x-1yx) = y (y-1 x-1yx)  G. ( y  G and y-1 x-1yx  G)
Hence G is a normal subgroup of G.

G
(ii) Let x, y  G. Then Gx, Gy are element of  
 G 
xy x-1y-1  U  xyx-1y-1  G (U  G)
 (xy) (yx)-1  G  G (xy) = G(yx).
 (Gx) (Gy) = (Gy) (Gx)

G
   is abelian.
 G 
(iii). Let N be any normal subgroup of G.
G
Let x, y  G. Then Nx, Ny 
N
46
G
Let be abelian.
N
Then (Nx) (Ny) = (Ny) (Nx)
 Nxy = Nyx
 (xy) (yx)-1  N
 xy xy-1  N
 U  N.
Thus N is a subgroup of G containing U.
Since G is the smallest subgroup of G containing U,
G must be contained in N. i.e. G  N.
To prove the converse, let G  N.
Then U  N (U  G)
 xyx-1 y-1  N
 (xy) (yx)-1  N
 Nxy = Nyx
 (Nx) (Ny) = (Ny) Nx)
G
 is abelian ■
N
11. Show that the group of automorphisms of an infinite cyclic group is of
order 2 and that the group of automorphisms of a finite cyclic group of
order n is of order (n).
Solution :
Let G = <a> is a cyclic group generated by a.
An automorphism T of G is defined by T(x) = xm …(1)
For any integer k, we have that T(xk) = [T(x)]k = (xm)k …(2)
Let y by any element of G.
Since T is a mapping of G onto itself, there must exist an element xk  G
such that y = T(xk ) = (xm)k . Thus we conclude that for the mapping T
defined in (1) is an automorphism of G, only if xm must be a generator of
G. If G is infinite, the only generators of G are x and x-1 . So in this case,
the only automorphisms of G are the identity mapping I for which I(x) = x
 I(xk) = xk  integer k and the mapping T defined by T(x) = x-1. Hence the
group of automorphisms of an infinite cyclic group is of order 2. Suppose
G is finite and of order n. Then xm is a generator of G iff m is prime to n
and < n. So, for each such m, the mapping T defined in (2) is an
automorphism of G.
To prove that T is 1-1:
Let xk1 and xk2 be any two elements of G, where 1 < k1<n and 1 < k2 < n.
47
Then T(xk1) = T(xk2)
 [T(x)]k1 = [T(x)]k2
 xm (k1-k2) = c
 n|m (k 1-k2)
But m is prime to n and 0 < h1 h2 < h
n | m (k1-k 2)
 k1 – k 2 = 0
 k1 = k 2
 xk 1 = xk2
 xm(k1-k2) = e
Thus T is 1-1
To prove that T is onto :
Since C is finite and T is 1-1, T must be onto.
Let xk1, xk2 be any two elements of G.
Then T (xk1 xk2) = T(xk1+k2)
= T [(x)]k1+k2
= xm(k1+k2)
= (xm) k1 (xm)k2
= [T(x)]k1 [T(x)]k2
= T(xk1) T(xk2)
Thus T is an automorphism for each positive integer m less than n and prime
to n. Hence the group of automorphism of a finite cyclic group of order n is of order
(n) Where (n) denotes the number of integers less than n and prime to n. ■
4.4 REVISION POINTS
Cayley’s Tehorem, related Lemmas and Problems are solved.
4.5 INTEXT QUESTIONS
1. If G is a group and H is a subgroup of idea 2 in G, prove that H is a normal
subgroup of G
2. If H is a subgroup of G and N is a normal subgroup of G, show that H  N is
a normal subgroup of H. Show by an example that H  N need not be
normal in G.
3. If H is a subgroup of G, and if N(H) = { g  G/gHg-1 = H } the prove the
following.
(i) N(H) is a subgroup of G, (ii) H is normal in N(H), (iii) If H is normal
subgroup of the subgroup K in G then K  N(H), (iv) H is normal in G its
N(H) = G
4. Prove that the centre Z of a group G is a normal subgroup of G
48
4.6 SUMMARY
Cayley’s theorem was stated and proved with relevant problems are solved.
4.7 TERMINAL EXERCISES
1. For G = S3 prove that G  I(G)
4.8 SUPPLEMENTARY MATERIAL
1. Algebra by Micheal Artin, Prentice Hall of India, New Delhi, 1994.
2. Lectures in Abstract Algebra Volumes I, II and III by N. Jacobson, D. Van
Nostrand Co., New York, 1976.
4.9 ASSIGNMENTS
1. State and prove Cayley’s theorem
2. If N is normal in G and a  G is of order O(a), prove that the order m of Na in
G/N is a divisor of O(a)
3. If Z is the centre of G, prove that G is abelian of the quotient group G/N is
cyclic
4.10 SUGGESTED READINGS / REFERENCE BOOK / SET BOOK
Algebra – By Michael, Artin – PHI Learning P Ltd
4.11 LEARNING ACTIVITIES
Students are requested to attend the P.C.P classes and work out the problems
given in the lesson
4.12 KEYWORDS
Cayley’s theorem, Commutator subgroup, Centre of a group


49
LESSON – 5

PERMUTATION GROUPS
5.1 INTRODUCTION
In this lesson, permutation group, cycle, another counting principle,
conjucacy, cauchy’s theorem for non abelian group, Sylow’s theorem, P-Sylow
subgroup, are discussed in detail with suitable examples. Some related problems
are solved.
5.2 OBJECTIVES
Permutation group, another counting principle, cauchy’s theorem for non
abelian group, sylow’s theorem, P-Sylow subgroup are explained.
5.3 CONTENTS
5.3.1 Definition - Permutation Groups
5.3.2 Another Counting Principle
5.3.3 Theorem (Sylow)
5.3.4 Definition : p-Sylow subgroup of G
5.3.1 DEFINITION - PERMUTATION GROUPS
We have shown that every group can be represented isomorphically as a
subgroup of A(S) for some appropriate set S and in particular, a finite group G can
be represented as a subgroup of Sn, for some n.
If   A(S) = Sn then  is a one one mapping of S onto itself and we could write
 by showing what it does to every element. For example if we consider S4 and if
T: x1 x3 , x2 x 4 , x3 x 1 , x4 x2 then we write
x x2 x3 x4  1 2 3 4
 =  1  and we can briefly write  as =  
 x3 x4 x1 x2  3 4 1 2
Let S be a set and A(S). Given two elements a, b  S, we define a = b if and
only if b = ai for some integer i (i can be positive, negative or 0) [Here, contrary to
our earlier notation, we write ai instead of i(a), to denote the image of a under i].
The above relation is an equivalence relation for,
1. a  a since a = ao

2. If a  b then b = ai, so that a = b-i, Hence ba

3. If a  b and b  c so that b = ai, c = b-i then c = (b)j = (ai)j = a(i+j)


Hence a  c.
The above equivalence relation induces a decomposition of S into disjoint
equivalence classes.
We call the equivalence class of an element sS, the orbit of s under .
50
In particular, if S is a finite set and sS, there is a smallest positive integer m
such that sm = s. The orbit of s under  then consists of the elements, s, s, s2,
……s(m-1).
By a cycle of  we mean the ordered set (s, s, … … s(m-1)). If we know all the
cycles of , we clearly know , since we would know the image of any element under
.
1 2 3 4 5 6
For example of A =  
2 1 3 5 6 4
It can be seen that the cycles of  are (1, 2), (3), (4,5,6). Now let us assume that
the cycle (ii, i2 ……ir) means the permutation which sends ii into i2, i2 into i3 ……
ir-1 into ir and ir into i1 and leaves all other elements of S fixed. Thus for example if S
consists of the elements 1, 2, ……9 then the symbol (1,2,4,7,8) means the
permutation
1 2 3 4 5 6 7 8 9
 
2 4 3 7 5 6 8 1 9
Now we multiply cycles by multiplying the permutation they represent. For
example if S has 9 elements, then (1, 2, 3) (5, 6, 4, 1, 8)
1 2 3 4 5 6 7 8 9 1 2 3 4 5 6 7 8 9
=    
2 3 1 4 5 6 7 8 9 8 2 3 1 6 4 7 5 9
1 2 3 4 5 6 7 8 9
=  
2 3 8 1 6 4 7 5 9
Because of a change in our notation regarding the manner in which we exhibit
the image of any element sS under a permutation A(S). i.e. we would write s
instead of (s) as was the case earlier. You will find a change with respect to the
multiplication of two permutations.
5.3.1.1 Lemma
Prove that every permutation is the product of disjoint cycles.
Proof
Let  be a permutation. Then its cycles are of the form (s, s …… s-1). Since
the multiplication of cycles is defined and since the cycles of  are disjoint, the
image of any element sS under  and the image of a under the product of cycles
are the same. Hence  = product of its cycles and the result follows. ■
Note: A cycle of the form (i1, i2,…im) in which there are m distinct symbols is
called a m-cycle or a cycle of length m.
5.3.1.2 Lemma
Prove that every permutation is a product of 2 cycles.
Proof:
In the last lemma we have shown that every permutation can be expressed as a
51
product of disjoint cycles. Now let us take a m-cycles (1, 2, ……m). We see that we
can write (1, 2, …m) = (1, 2) (1,3) … (1, m). Hence each cycle can be expressed as a
product of 2-cycles. So every permutation can be expressed as product of 2-cycles.
Any 2-cycle is also called a transposition.
5.3.1.3 Definition
A permutation S, is said to be an even or an odd permutation if it can be
represented as a product of an even or odd number of transpositions.
Note: It can be shown that when any given permutation is written as a
product of 2-cycles, the number of 2-cycles in the product is always even or odd,
even though the number is not a constant. Because of this reason it is meaningful
to define even and odd permutations as we have done above. The following facts
can be easily verified.
1. The product of two even permutation is an even permutation.
2. The product of an even permutation and an odd permutation is an odd
permutation.
3. The product of two odd permutations is an even permutation.
Let An be the subset of Sn consisting of an even permutation. Since the product
of two even permutations is even, An is closed and hence is a subgroup of Sn. This
subgroup An is called alternating group of degree n.
5.3.1.4 Lemma
Prove that Sn has a normal subgroup of index 2.
Proof
Let W be the group of the real numbers 1 and –1 under multiplication. Let us
define : Sn  W by (s) = 1, if s is an odd permutation = -1, if s is an even
permutation=1.
Clearly  is a homomorphism. The kernel of  is An . An being the kernel, An is
normal in Sn. Moreover
Sn S  O S n 
 W and so 2  O (W)  O  n  
An  An  O  An 
We also note that since O(Sn)= n!; O(An) = ½n!
A group G is said to be a permutation group of degree n if G is a group of Sn
the symmetric group of permutations of degree n. In other words a permutation
group is a group in which the elements are permutations. We have the following
standard and interesting problem, concerning a permutation group. ■
Problem 1
In any permutation group, either all or exactly half of the elements are even
permutations and these even permutations form a normal subgroup.
Solution
Let G be the given permutation group. Clearly G is a finite group. If all the
elements in G are even permutations there is nothing to prove. On the other hand,
52
let G contains both even and odd permutations. Let E1, E2 … Em be the only
distinct even permutations and O1, O2 … Op be the only distinct odd permutations
in G. We show that m = p. If O denotes a typical odd permutation in G, consider the
elements OE1, OE2, … OEm. All these elements are distinct odd permutations in G,
since OEi = OEj implies Ei = Ej (by cancellation law valid in G) and Ei = Ej implies
i = j. Moreover the product of any odd permutation in G and any even permutation
in G is necessarily an odd permutation, in G. Thus G contains atleast m odd
permutations, viz., OE1, OE2, … OEm, and so p>m. Again, the elements,
OO1, … OOp are distinct even permutations in G. Distinct because of the
cancellation law valid in G and even because the product of any two odd
permutations in G is necessarily an even permutation in G. Thus G contains atleast
p even permutations, viz OO1, OO2, … OOp, whence m>p. Earlier we have shown
that p>m. Hence p = m. i.e. exactly half of the permutations in G are even. Clearly
these even permutation form a subgroup of G, as the set of all even permutations in
G is clearly closed under multiplication on G. As G is a finite group, this set is a
subgroup of G. The order of this subgroup of all even permutations in G being half
the order of the group G, this subgroup has index 2 and hence is normal. ■
5.3.2 ANOTHER COUNTING PRINCIPLE
5.3.2.1 Definition
Let G be a group and a, b  G. The element b is said to be a conjugate of a in
G if there exists an element c  G such that b = c-1ac. We shall call this relation of
conjugacy and write a ~ b, whenever b = c-1ac for some element c  G.
5.3.2.2 Lemma
Prove that conjugacy is an equivalence relation in G.
Proof
1. a ~ a since a = e-1ae where e is the identity in G.
2. If a ~ b then b= c-1ac  cbc-1 = a a=(c-1)-1 bc-1b~a
3. If a~b, and b~c then b=g-1ag and c=h-1bh where g,hG.
Hence c=h-1(g-1ag)h= (gh)-1a(gh).  a ~ c.
For a  G, let Cn = { xG| a~x }. Then Gn is the equivalence class determined
by a and this is usually called the conjugate class of a in G. It consists of all
elements of the form y-1ay as y ranges over G. Let us denote by Cn as the number
of distinct elements of G. ■
5.3.2.3 Definition
If a  G, then N(a) the normalizer of a in G, is the set N(a) = { xG|xa=ax }.
5.3.2.4 Lemma
N(a) is a subgroup of G.
Proof
First let us show that N(a) is closed. Let x,y  N(a). Then xa = ax and ya = ay.
Hence (xy)a = x(ya) = x (ay) = (xa) y = (ax)y= a(xy). Hence xyN(a). Now we will show
53
that inverse also exists in N(a) itself. Let xN(a) so that xa = ax. Therefore
x-1 (xa)x-1 = x-1 (ax)x-1  ax-1 = x-1a. So x-1N(a). Hence N(a) is a subgroup of G. ■
5.3.2.5 Theorem
O(G)
If G is a finite group, then Cn =
O(N(a))
(The number of elements conjugate to a in G is the index of the normalizer of a
in G).
Proof
To prove the theorem we will show that two elements in the same right coset of
N(a) (G yield the same conjugate of a, whereas two elements in different right cosets
of N(a)) in G give rise to different conjugates of a. Thereby there will be a 1-1
correspondence between conjugates of a and right cosets of N(a).
Let x, y  G are in the same right coset of N(a) in G. Then y = nx where nN(a)
and hence na = an. So y-1 = (nx)-1 = x-1n-1.
Therefore y-1ay = x-1ax.
Hence x and y give rise to the same conjugate of a. If x and y are in different
right coset of N(a) in G we claim that x-1ax  y-1.
If not, let x-1ax = y-1ay. Then y(x-1ax) x-1 = y (y-1ay) x-1 ( yx-1)a = a(yx-1)
Hence (yx-1) N(a). Hence x and y are in the same right coset of N(a) in G,
contradicting the fact that they are in different cosets.
O(G)
Hence Cn = ■
O(N(a))
Corollary
O(G)
Prove that 0(G) =  where the sum runs through one element a in
O(N(a))
each distinct conjugate class of a.
Proof
Since conjugacy is an equivalence relation in G, G can be written as the union
of disjoint equivalence classes C(a) having c elements. Hence G=UC(a) and
O(G) =  C(a) where the sum run over one element a in each conjugate class.
The above equation is called the class equation of G. ■
5.3.2.6 Lemma
a  Z(G) if and only if N(a) = G. If G is finite, then aZ (G) if and only if
0(N(a)) = 0(G).
Proof
If aZ (G) then xa = ax for all xG. Hence H(a) = G. Conversely if N(a) = G,
then xa = ax for all xG, so that a Z(G). Hence if G is finite, O(N(a)) = O(G) iff
aZ(G). ■
54
Remark
Using the above fact the class equation can be modified and rewritten as
O(G)
0(G) = 0(Z) + 
aG O(N(a))
where Z stands for Z(G).
aZ

5.3.2.7 Theorem
If 0(G) = p n where p is a prime, then that Z(G)  <e>.
Proof
Let aG, since N(a) is a subgroup of G, O(N(a)) is a divisor of O(G)p n. Therefore
0(N(a)) = pn. Therefore 0(N(a)) = pn where n < n provided aZ(G). If z = 0(Z(G)),
then by the class equation we have

pn
pn  z  
n  n p n

pn
Now p is the divisor of the left hand side. Moreover p | and hence p divides
p n
each term in the  and hence

pn
p| p n  i. e. p | z
n  n pn
Since e Z(G), z>1. Thus z is a positive integer divisible by the prime p. Hence
z > 1. So there must be an element besides e in Z(G).
Hence the theorem is proved. ■
Note
Since p| z, z > p, and z is a power of p, as Z(G) is a subgroup of G.
Corollary
If O(G) = p2, where p is a prime, then G is abelian.
Proof
We will show that Z(G) = G itself so that G is abelian. By the previous theorem
Z(G)  <e> is subgroup of G, so that (Z(G)) = p or p2.
If O(Z(G)) = p2 we are done. Suppose that O(Z(G)) = p, let aG, such that
aZ(G). Let us take N(a). Clearly Z(G)  N(a) and aN(a), so that O(N(a))>p. But by
Lagrange’s theorem 0(N(a)) | 0(H). O(N(a)) = p2. Hence N(a) = G so that a  Z(G)
which is a contradiction. Hence o(Z(G))p
Therefore Z(G) = G itself. Hence G is abelian. ■
5.3.2.8 Theorem
(Cauchy’s theorem for any group G, not necessarily abelian).
If G is a finite group and if p is a prime number such that p|O(G), then G has
an element of order p. ■
55
Proof
Let we prove the theorem by the method of induction on O(G). Let we assume
that the theorem is true for all groups of order < O(G).
Let W be a proper subgroup of G such that p|O(W). Then by our induction
hypothesis there would exist an element of order p in W and hence also exist such
an element in G. Thus we may assume that p is not a divisor of the order of any
proper subgroup of G.
In particular if aZ(G), so that N(a)G, then p cannot divide 0(N(a)). Let us
write down the class equation.
O (G )
O(G) = O(Z(G)) =
O ( N ( a))
Since p| O(G) and p cannot divide O(N(a)), we have

O(G )
p for every a  Z(G)
O( N ( a))

O(G )
p  . We also have p|O(G)
N ( a )G O( N (a ))

Hence from the class equation p|O(Z(G)). Thus Z(G) is a subgroup of G whose
order is divisible by p. But by our assumption, p does not divide the order of any
proper subgroup of G. Hence Z(G) = G and G is abelian. Since we have already
proved the fact for abelian groups, the result follows. ■
SYLOW’S FIRST THEOREM
We have seen in Lagrange’s theorem that the order of any sub-group of a finite
group G is a divisor of the order of G. We now consider whether the converse is
true. It can be shown by means of suitable examples that the converse is false.
Especially in the case of non-abelian groups. In this connection we have the
following theorem due to Sylow.
Theorem (Sylow)
 
If p is a prime such that p |O(G), then G has atleast one sub-group of order p .
Proof
We require the following theoretical result of number theory
  p m
If n = p m , pr|m but p(r+1) does not divide m then pr|    , p (r+1) does not
 p 
 p m
divide    , where we write  n  for n.

 p  k

 p m p  m(p  m  1) (p  m  1)...(p  m  p   1)
In fact, consider    =  . We note that
 p  p  (p   1) (p   1)...(p   p   1)
pk /(p m-i) if and only if pk|i and hence if and only if pk / p  - I, where we assume
56

k< and I<i<p -i. Therefore all powers of p canceled out except the power which
 pm
divides m in the expression for    . Thus the result follows.
 p 
Now let us prove the theorem. Let O(G) = n = pm. Let M be the set of all
 p m
subsets of G which have p  elements. Thus M has    elements. Now let us
 p 
define a relation “~” on M as follows.
For M1,M2, we can write M1~M2 iff there exists an element gG such that
M1=M2g. We claim that this relation is an equivalence relation. Since M1 = M1e
where e is the identity in G, M1 ~ M1 and hence the relation is reflexive.
If M1~M2 then there exists g such that M1 = M2g which implies that
M2 = M1g-1 . Hence M2~M1 and hence the relation is symmetric.
If M1~M2 and M2 ~M3 then there exists g and f in G such that M1 = M2g and
M2 = M3f which implies that M1 = M3fg. Hence M1~M3 and the relation is transitive.
So the relation “~” is an equivalence relation on M.
This equivalence relation gives rise to a partition of the set M into equivalence
classes. We claim that there exists atleast one equivalence class such that the
number of elements in this class is not a multiple of p(r+1). For if p(r+1) is a divisor of
the number of elements in each equivalence class then p (r+1) would be a divisor of
 p m 
the number of elements in M which is not true since M has    elements and by
 p 
 p m
the number theoritical result p(r+1) does not divide    .
 p 
Let M1, M2, …. Mn be such an equivalence class where p(r+1) does not divide t.
Let H = { gG|Mtg = M }. Clearly H is a sub-group of G.
For if a,bH, then M1a = M2a, M2b = M1.
Hence M, ab = (M1ab) = M1b = M1, so that ab H; i.e. H is closed under
multiplication and G being a finite group H becomes a subgroup.

Now we show that O(H) = p so that H turns out to be the required subgroup of
G. We claim that there is a 1-1 correspondence between the elements in the
equivalence class { M1, M2 … Mn } and the right cosets of H in G.
For M1g = M1 g  Mg(g)-1=M1
 g(g)-1  H
 Hg = Hg
O(G )
Hence t = the number of right cosets of H in G =
O( H )

 t 0(H) = 0(G) = n = p m.
57
Since p(r+1) does not divide t and p(+r)|t0(H) = n, it follows that p|0(H) and so
0(H) > p.
Moreover, if M1  m1, then for all h  H we have m,h  M1.
Thus M1 has atleast 0(H) distinct elements. However M1 was a subset of G
containing p elements. ■
Corollary

If pm|O(G), p (m+1) does not divide O(G) then G has a subgroup of order p m.
Definition : p-Sylow subgroup of G
A subgroup of G of order p m where pm | o(G) and pm+1 does not divide 0(G) is
called a p-sylow subgroup of G.
Alternate Proof
We shall prove the sylow’s theorem by induction on O(G) by assuming that the
theorem is true for groups of order less than that of G. We shall show that, it is also
true for G.
Since the theorem is true if 0(G) = 1, we have a basis for induction.
Let O(G) = pmn where p is not divisor of n. If m = 0, theorem is obviously true.
If m = 1, the theorem is true by Cauchy’s theorem. So let m > 1. Then G is a group
of composite order and so G must posses a subgroup H such that H  G.
O (G ) O(G )
If p is not a divisor of then pm|O(H). Since 0(G) = pmn = O(H) .
O( H ) O( H )
pm+1 does not divide O(H). For, if p m+1|0(H), then pm+1|0(G) which is not the
case. Further 0(H) < 0(G). Thus by the induction hypothesis the theorem is true for
H.  H has a subgroup of order pm and this will be a subgroup of G.
O(G )
So let us assume that p is a divisor of .
O( H )
Consider the class equation
 O(G )
0(G) = 0(Z) +
a  Z O( N (a ))
Since a  Z,  N(a)  G, by our assumption
 0(G )
P . Also p O(G )
a  Z 0( N (a ))
 O (G )
Hence, p O (G )  p 0( Z )
aZ O ( N ( a ))
Then by Cauchy’s theorem, Z has an element c of order p where Z is the
Centre. Also N = <c> is a cyclic subgroup of Z of order pN is a cyclic subgroup of
G of order P. Since cZ, N is a normal sub-group of G of order p
58
m
G 0(G ) p n
Let G = .  0(G) =   p m1n . Thus 0(G) < 0(G).
N 0( N ) p
Also p m-1 |O(G). But pm does not divide O(G). By induction hypothesis G has
G
a subgroup U of order pm-1. Let f be defined by f: G  such that f(x) = Nx
N
is a natural homomorphism with kernel N . Let U = { xG|f(x)U }. Then U is a
U  U  O (U )
subgroup of G and U   O (U )  O  
N  N  O( N )
Thus U is a subgroup of order pm ■
SYLOW’S SECOND THEOREM 2.5.5
Let G be a finite group and p be a prime number such that p|0(G). Then any
two p-Sylow subgroups of G are conjugate.
Proof
Let 0(G) = p mq, where (p, q) = 1.
Then every p-Sylow subgroup of G is of order pm.
Let H be any p-Sylow subgroup of G.
Let M be the set of all conjugates of H in G.
To prove that every p-Sylow subgroup K of G is conjugate to H i.e. kM.
Assume the contrary. That is assume that there exists a p-sylow subgroup K
not conjugate to H. For any p-sylow subgroup P(K), K  P
For if K  P & O(K) = O(P)  K = P.
Hence there exists x  (K-P). If x-1Px = x  x  N(P)
Since x  K which is of order p m, x P. This is a contradiction.
Hence x-1Px  x and P has more than one conjugate in K.
The number of conjugates of P in K = Index of N(P) in H.
Also Index of Na(P) in K |O(K) = pm
The number of conjugates of P in K = a multiple of p.
Define a relation ~ in M as follows:
If H1, H2 M then H1 ~ H2  there exists x  k such that H2 = x-1 H 1x.
This is an equivalence relation on M and it partitions M into equivalence
classes. As before, for any PM, (PK), the number of conjugates of P in K = a
multiple of p.  the equivalence class of P determined by the elements in K
contains n elements where p|n. Thus O(M) = a multiple of p
But O(M) = Index of N(H) in G
Index of H in G = q i.e. iG(H) = q
Since H is a subgroup of N(H) & O(H) = pm

By Lagrange’s theorem, O(N(H)) = p mr


59
0(H) iG(H) = 0(N(H)) . iG(N(H))
 pm . q = pmr O(M)
 q = rO(M)
Since p|O(M) we have p|q
This is a contradiction to the fact that (p, q) = 1
Hence K is a conjugate of H.
Thus any two p-Sylow subgroups are conjugates. ■
Problem
If a group G has only one p-Sylow subgroup H then H is normal in G.
Solution
Assume that the group G has only one p-Sylow subgroup H. Let x  G.
To prove that x-1Hx is a p-Sylow subgroup of G.
Let O(G) = pmn where p is a prime and p|n since H is a p-Sylow subgroup of
G, O(H) = pm. We try to prove that x-1Hx is a subgroup of G.
For that, let x-1h1x, x-1h2x  x-1Hx, where h1, h2  H.
But (x-1h1x)(x-1h2x)-1 = x-1h1x x-1(h2)-1(x-1)-1 = x-1h1e(h2)-1x= x-1(h1h 2-1)x x-1Hx
 x-1 Hx is a subgroup of G.
Define : H  x-1Hx by  (h) = x-1hx
Let x-1yx x-1 Hx
Then y  H & (y) = x-1yx
Hence  is on to.
Let h1, h2  H
Let (h1) = (h2) then x-1h1x = x-1h2x
 h1 = h2
  is one one
Thus  is a 1-1 correspondence between the elements of H and the
elements of x-1Hx. 0(x-1Hx) = 0(H) = pm . Hence x-1Hx is a p-Sylow
subgroup of G. Since H is the only p-Sylow subgroup of G, H = x-1 Hx
 H is a normal subgroup of G. ■
SYLOW’S THIRD THEOREM:
Let p|O(G) where G is a finite group and p, a prime number. Then the number
of p-Sylow subgroups is (1+rp) where r is a non-negative integer and (1+rp)|O(G).
Proof
Let M be the set of all p-Sylow subgroup of G. Let K M be fixed. Define ~ in M
as follows. If H1, H2  M, then H1 ~ H2  there exists x  K such that H2 = x-1 Hx.
60
Now ~ is equivalence relation. By Sylow’s second theorem, if P is any p-Sylow
subgroup of G different from K, the number of Conjugates of P in K = a multiple of
p. Hence M contains (1+rp) elements where r is some non negative integer. Also
O(M) = index of N(K) in G which divides O(G). Hence (1+rp) |O(G) ■
Problem
Show that group of order 28 has a normal subgroup of order 7.
Solution
Let G be a group of order 28. Since 7 is a prime divisor of 28 by Sylows first
theorem there exists at least one 7-Sylow subgroup H (say) of G. Further by Sylow’s
II & III theorems, any two 7-sylow subgroups are conjugate and if the number of 7
sylow subgroups is n, then n  1 (mod 7) and n is a divisor of 28, the order of G. It
is easy to check that the only divisor of 28 which is  1 (mod 7) is 1 and so H is the
only 7-Sylow subgroup. This means that H is normal in G■
5.4. REVISION POINTS
Permutation group, Cycle, Conjucacy, Cocuchy’s theorem for non abelian
group, Sylow’s theorem, p-Sylow subgroup.
5.5. INTEXT QUESTION
1. In any permutation group, either all or exactly half of the elements are even
permutations and these even permutations form a normal subgroup
2. If group G has only one p-Sylow subgroup H, then prove that H is normal in G
3. Show that the group of order 28 has a normal subgroup of order 7.
5.6. SUMMARY
Permutation group, Another Counting Principle, Sylow’s theorem, p-Sylow
subgroup of G are explained.
5.7. TERMINAL EXERCISES
1. Prove that any group of order 15 is cyclic
2. If G is a group such that  (x) = xn defines an automorphism of G, prove that
for a  G, a-1 is the center of G
3. If O(G)  42, prove that its 7-Sylow Subgroup is a normal subgroup.
5.8. SUPPLEMENTARY MATERIAL
1. Algebra by Micheal Artin, Prentice Hall of India, New Delhi, 1994.
2. Lectures in Abstract Algebra Volumes I, II and III by N. Jacobson, D. Van
Nostrand Co., New York, 1976.
5.9. ASSIGNMENTS
1. Express as the product of disjoint Cycles
(a). (1,2,3) (4,5)(1,6,7,8,9)(1,5)
(b). (1,2)(1,2,3)(1,2)
2. Compute a-1ba where
61
(i) . a = (1,3,5)(1,2) & b = (1,5,7,9)
(ii) a = (5,7,9) & b = (1,2,3)
3. Prove that Sn has a normal subgroup of index 2.
O (G )
4. If G is a finite group, then prove that Cn =
O ( N ( a ))
5. State and prove Cauchy’s theorem for non abelian group.
6. State and Prove Sylow’s theorem.
5.10. SUGGESTED READINGS / REFERENCE BOOK / SET BOOK
Algebra – By Michael Artin – PHI Learning P Ltd
5.11. LEARNING ACTIVITIES
Students are requested to attend the P.C.P Classes and work out the problems
given in the Lesson
5.12. KEYWORDS
Permutation group, Cycle, Conjucacy, Cauchy’s theorem for non abelian group,
Sylow’s theorem, p-Sylow subgroup.

62
LESSON – 6

RINGS
6.1. INTRODUCTION
In the previous lessons we introduced the basic notion of groups and studied
their important properties, which are highly useful for our further studies in
Modern Algebra. In this lesson and in the subsequent two lessons we introduce the
notion of rings and study their essential properties, besides other relevant facts.
Just as groups, rings are algebraic systems which serve as building blocks for
various structures encountered in Modern Algebra. While the abstract notion of a
group is founded on the set of bijections of a set onto itself, the notion of a ring will
be seen to have arisen as generalizations from the set of integers and the additive
and multiplicative properties of integers. We know that groups are one-operational,
(i.e.) structures in which only one operation (generally called multiplication) is
defined. But in the case of rings, there are two operations usually called addition
and multiplication. Despite this difference, the analysis of rings will follow the same
pattern as for groups. In that we will be interested in studying analogs of
homomorphism, normal subgroup, factor or quotient groups etc. We begin with a
formal and complete definition of a ring.
6.2. OBJECTIVES
Rings, types of rings, related therems, Lemmas and problems are discussed in
detail.
6.3. CONTENTS
6.3.1 Definition : Types of Rings
6.3.2 THEORM : Homomorphism of Ring
6.3.1 Definition : Types of Rings
A ring R  < R, +, , > is a system R of double composition, (i.e.) a set R in
which two binary operations + and  (respectively called addition and
multiplication) are defined and which associate with every pair of elements a, b
(distinct or otherwise) of R with unique elements of R, viz. (a + b) and (a.b) (or
simply ab) respectively called their sum and product such that the following axioms
are satisfied:
A0 : (a+b)R for every a, bR. (closure property with reference to addition)
A1 : a+(b+c) = (a+b)+c every a, b, c R (associativity of addition)
A2 : There exists an element 0R called the zero or the null element of R
such that a+0 = 0+a = a for every a  R (existence of the zero element).
A3 : Corresponding to any element a of R, there exists an element – a in R,
called the negative of a such that a + (-a) = (-a) +a = 0. (existence of the
additive inverse)
A4 : (a+b) = (b+a) for every a, b  R. (commutativity of addition)
63
M0 : ab  R for every a,b  R (closure property with reference to
multiplication)
M1 : a(bc) = (ab)c for every a, b, c  R (associativity of multiplication)
D1 : a(b+c) = ab + ac for every a, b, c  R (left distributive law)
D2 : (b+c) a = ba + ca for every, a, b, c  R (Right distributive law)
Remark on the Above
1. Definition
The axioms A0 – A4 given above shows that, for the ring R, < R, + > is an
additive abelian group. This group is called the additive group of the ring R and is
usually denoted by R+. So also the axioms M0 and M1 shows that <R,  > is a semi-
group and this semi-group is called the multiplicative semi-group of the ring R. The
axioms D1 and D2 shows that the multiplication ‘’ in the ring R is both left
distributive and right distributive over addition ‘+’ in R.
In view of the above remark, we may define the notion of a ring more briefly as
follows (one may give this definition for a ring):
2. Definition
A ring R  < R, +, > is an algebraic structure in which two binary (or algebraic)
operation viz. addition ‘+’ and multiplication ‘’ are defined, such that (i) <R, + > is
an abelian group (called the additive group of the ring R), (ii) < R,  > is a semi-
group (called the multiplicative semi-group of the ring R). and (iii) the multiplication
is both left distributive and right distributive over addition.
We give below some additional axioms, some of which may or may not be
satisfied in a ring R.
M2 : ab = ba for every a, b  R (Commutativity of multiplication)
M3 : There exists an element 1R, called the unit element or the identity in
R, such that a.1 = 1.a = a for every a R. (Existence of the multiplicative
identity)
M4 : R has at least two elements and the non-zero elements of the R form a
group under multiplication. (Multiplicative group structure of the non-
zero elements)
M5 : For some non-zero element a of R, there exists element bR such that
ab = 0 or ba =0 where b 0. (Existence of Zero divisors)
M5 : If a, b  R and ab = 0 then either a or b or both are zero (absence of
non-trivial zero divisors) ■
3. Definition
A ring R in which axiom M2 is satisfied is called a commutative ring and a ring
R in which axiom M2 is not valid (i.e. in which there are at least two elements a, b
such that ab  ba) is called a non-commutative ring.
64
Note: Throughout our lessons, we always mean by a ring, unless otherwise
stated, a non-commutative (associative) ring. Some authors (for instance, refer to
“General Algebra” by A.G. Kurosh) exclude the axiom M1 (viz. Associativity of
multiplication) in the definition for a ring, so that according to them a ring need not
be associative. Examples of rings that are not necessarily associative can be given.
4 Definition
A ring R in which axiom M3 holds good is called a ring with identity (or unity or
unit element).
5. Definition
A ring R in which axiom M4 holds good is called a field (not necessarily
commutative) more precisely a division ring or a skew-field or an S-field. If axiom
M2 is also satisfied (besides axiom M4), then the ring R is called a commutative field
or simply a field.
6. Definition
A ring R in which axioms M2 and M3 satisfied is called an integral domain. In
other words a commutative ring (with at least two elements) which has no non-
trivial zero divisors is called an Integral domain.
Note: Some authors like N. Jacobson (vide: his book Lectures in Abstract
Algebra’, Vol. I) do not insist on the axiom M2 (i.e. commutativity) in the definition
of an integral domain.
We will not give a good many standard examples of different types of rings.
Example 6.1
The set Z of all integers (positive, negative and zero) is a (commutative) integral
domain with identity under the usual operations of addition and multiplication of
integers.
Example 6.2
The set nZ of all multiples of a fixed integer n[  +1 & -1] is an (commutative)
integral domain with identity, the operations being the usual addition and
multiplication of integers.
Example 6.3
All rational numbers, so also all real numbers, all complex numbers form
(commutative) fields Q, R, C the operations being the usual operations of addition
and multiplication of the numbers, rational or real or complex, as the case may be.
Example 6.4
The set Z[i] of all Gaussian integers, viz, all complex numbers of form (a + ib)
where a, b  Z and i=  1 , is an (commutative) integral domain with identity under
the usual operations of addition and multiplication of complex numbers. Z[i] is only
an integral domain and not a field, since the inverse of any Gaussian integer (a+ib)
(0) is not a Gaussian integer, unless, the Gaussian integer (a+ib)(0) is not a
Gaussian integer, unless, the Gaussian integer is  1 or  i.
65
Example 6.5

All real numbers of the form a+b c , where c is a fixed positive integer which is
not a perfect square and a, b  Z form a commutative integral domain, which is not
a field.
Example 6.6
Let < R, + > be any non-trivial additive abelian group. Define the trivial
multiplication ‘.’ on R by setting ab = 0, the zero (or null) element of the group
<R, + >. Then one can easily check that <R, +,  > is a commutative ring, every one
of whose elements  0 is a zero divisor.
All the above examples are infinite commutative rings provided in Example 6,
the additive group is an infinite group. We will now give examples of finite rings,
which are either commutative or non commutative.
Example 6.7
Let n be a fixed positive integer and let Zn denote the set of all residue classes
modulo n. For example, if n = 4. = 0, 1, 2, 3 , where i denotes, for i = 0, 1, 2, 3 the
residue class modulo 4 determined by I, viz, the set of all integers which are
congruent to i modulo 4. Now < Zn ,  > is a finite commutative ring with identity
viz 1, if  and  respectively are the usual addition and multiplication modulo n,
(Some authors use the notation +n, ×n instead of  and  ). This ring will turn out to
be a finite field iff n is a prime, while it will contain non-trivial divisors of zero iff n
is not a prime. For example Z5 is a field of 5 elements and Z6 is a commutative ring
with identity containing 2, 3, 4 as the only non-trivial divisors of zero, since
2  3  0, 3  4  0 .
Example 6.8
Let Fn be the set of all n  n matrices (aij), where aij are elements chosen from a
field F. For example, F can be taken as Z m, mentioned in the last example, where m
n2
is a prime, in which case Fn will have m elements. Because, matrix multiplication
is both left distributive and right distributive over matrix addition. These facts are
supposed to be known, otherwise one may refer to any standard book on Matrices –
Fn under the usual operations of matrix addition and matrix multiplication
becomes a finite non-commutative ring with identity, viz, the identity matrix (bij)
where bij = 0 if i  j and b ii = 1 of the field F. This ring F n contains non-trivial
divisors of zero. For instance, the matrix
 1 0 0 ... ... 0 
 0 0 0 ... ... 0 
... ... ... ... ... ...
... ... ... ... ... ...
 0 0 0 ... ... 0 

in Fn is a divisor of zero, since


66
 1 0 0 ... ... 0  0 0 ... ... ... 0 
 0 0 0 ... ... 0  a 21 a 22 ... ... ... a 2n 
... ... ... ... ... ...  ... ... ... ... ... ... 
... ... ... ... ... ...  ... ... ... ... ... ... 
 0 0 0 ... ... 0  a a n2 ... ... ... a nn 
 n1

is a zero matrix
and similarly,

 0 a12 a13 ... ... a1n   1 0 0 ... ... 0 


 0 a 22 a 23 ... ... a 2n   0 0 0 ... ... 0 
... ... ... ... ... ...  ... ... ... ... ... ...
... ... ... ... ... ...  ... ... ... ... ... ...
0 a a n3 ... ... a nn   0 0 0 ... ... 0 
 n2

is a zero matrix
Thus the matrix

 1 0 0 ... ... 0 
 0 0 0 ... ... 0 
... ... ... ... ... ...
... ... ... ... ... ...
 0 0 0 ... ... 0 

in Fn turns out to be a non-trivial two sided zero-divisor. Thus Fn contains


atleast one non-trivial divisor of zero; in fact, it contains many more.
Note: In in the above example, if F is the field of rational (or real or complex)
numbers, Fn turns out to be an example of an infinite, non-commutative ring with
identity and with (many) non-trivial divisors of zero.
It will be subsequently shown that any finite integral domain is a field. This
being so, to give an example of a non-commutative integral domain which is not a
field, we need try to get an example of an infinite non-commutative integral domain.
Actually we give below an important example of an infinite skew-field, viz, the non-
commutative field of Quaternion (in the form of matrices) and this skew field is also
a non-commutative integral domain – one can show that a field and a skew field
cannot contain non-trivial divisors of zero.
Example 6.9
(The division ring or skew-field of quaternions): Let F denote the set of all
u v
complex matrices of the form  where u, v are complex numbers and
 v u
u, v denote their conjugates. One can easily check that F is a skew-field, under the
usual operations of matrix addition and matrix multiplication. In fact, it is not
difficult to verify that the sum and product of any two matrices in F, besides the
zero matrix, the identity matrix and the negative of any matrix in F all belong to F.
67
u v
Also for any non-zero matrix A =  , the multiplicative inverse belongs to F, the
 v u

1 u  v
inverse being  .
A v u 
Note: The above mentioned division ring of Quaternion can also be realized as
a non-commutative algebra of dimension 4 over the field of real numbers.
Having listed certain important examples of rings, let us now mention certain
familiar properties of rings.
Lemma 7
If R is ring and a, bR then, (i) a0 = 0a = 0; (ii) a(-b)=(-a)b=-(a b);
(iii)(-a)(-b) = ab. If further R is a ring with identity 1, then (iv) (-1)a = –a;
(v) (–1)(–1) = 1.
Proof
(i) a0 = a(0+0) = a0 + a0, whence a0 = 0 using the cancellation law (or the
property of idempotent) in the group < R, + >. Similarly, 0a = 0 follows:
(ii) 0 = a0 = a [b +(-b)] = ab + a (-b), which yields a(-b) = - (a b) by definition of
inverse in the group < R + > Similarly (-a) b = - (a b) follows.
(iii) Using (2), we get (-a) (-b) =-[a(-b)] =-[-(ab)] = ab (iv) and (v) are only special
cases of (ii) and (iii). ■
Making use of the additive group < R, + > of the ring R, we define as usual,
multiple of any element a  R where n is an integer by setting na = a + a ….. + ,
(n times, if n is a positive integer).
= (-a) + (-a) + … + (-a) (-n times, if n is a negative integer)
= 0, where n is a integer 0.
With this familiar notion of a multiple of any element a, we can state the following
(easily proved) facts: If a, b  R and m, n be any two integer (positive or negative or
zero) then (i) m (a + b) = ma + mb;(ii) m(-a) = -(m a);(iii) (ma) (nb) = (mn) (ab).
Note: The result (i) above is a consequence of the abelian nature of the
additive group < R, + > of the ring and the result (iii) is a consequence of the two
distributive laws valid in a ring. So also, in order to prove a0 = 0 and 0a = 0 we
need respectively the left and the right distributive laws. The following
generalization of the distributive laws can be easily proved.
Lemma 7.1
If a1, a2 … am, b1, b2, … b n  R, where R is a ring, then a1+a2+… +am) (b1+b2 +…+ bn)
= (a1b 1 + a1b2+ ...+a1b n) + (a2b1 + a1b2+ …+a2bn) + … + (amb 1 + amb 2 + … + ambn).
More briefly the above fact may be stated as

m   n  m n
  a i    bi    a b i i
 i 1   i 1  i 1 j 1
68
Note that since addition is commutative in the ring R it is immaterial in what order
the terms in the sum on the right hand side of the result (stated above) are written.
Lemma 7.2
Perove that any finite integral domain is a field.
Proof
Let D =< D, +, ∙ > be a finite integral domain. To prove that D is a field it is
enough if we prove that the non-zero elements of D is a field. So, it is enough if we
prove that the non-zero elements of D form a group under multiplication. It suffice
to show that D* = D- {0} is a semigroup under multiplication and that the equations
ax = b, ya = b have solutions x, y  D*, given a,bD* (vide alternative Definition
No.2 for a group, mentioned in Lesson 2). From the definition of an integral domain
it follows that if a, b  D, (so that a, b  D and a, b  0) then a, bD*. Hence D* is a
semigroup under multiplication. Suppose now the finitely many elements of D* are
a1, a2 … an (all distinct). If a is any arbitrary element in D*, consider the elements
aa1, aa2, … aan. Since <D*, * > is a semigroup all the above elements are in D* and
further they are all distinct; for otherwise, suppose a.ai = a.aj then a(ai – aj)
= a [ai+(-aj)] =a.ai+(a.-aj)], by Lemma 1.
= a.ai – a.aj = 0
ai – aj = o, since a  0 and D is an Integral domain. Thus a.ai= a.aj  ai = aj.
This shows that all the n elements aa1, aa2, …, aan are distinct and belong to D*.
The above n elements are same as the n elements a1, a2, …, an but for possible
order. That is for given any bD* there exists an element aiD* such that aai = b,
which means that the equation ax = b has a solution x in D*. Similarly we can show
that the equation y.a = b has a solution y in D*. Similarly, we can show that the
equation y.a = b has a solution y in D*. Thus <D*,. > turns out to be a group, as
desired. Hence the Lemma. ■
Note: The above proof must be reminiscent of the proof we have given for
problem 4, in Lesson 2. In fact, if we make use of this problem, namely that any
finite cancellative semi-group this Lemma follows almost automatically in view of
the following:
Remark
A (commutative) ring R with atleast one non-zero element is an integral
domain if and only if the non-zero elements R form a commutative semigroup under
multiplication. In fact, if R is an integral domain and R* = R - { 0 }, then by
definition of an integral domain it follows that <R*, . > is a semi-group (See
condition M5). Further if a  R and ab = ac then a(b-c) = ab –ac = 0 which implies
that (b-c) = 0 as a  0, and hence b = c. Thus ab = ac  b = c whenever a  0.
Similarly ba = ca  b = c whenever a  0. Thus < R*,. > turns out to be a
(commutative ) semigroup. Conversely suppose <R*,> is a (commutative) semigroup.
Then for a, b 0 in R ab  0, which means that R has no non-trivial divisors of zero.
(In other words, for a, b R, ab = 0 always implies that atleast either a or b is equal
69
to zero). Thus R becomes an integral domain. We will now formally define the notion
of divisors of zero or zero divisors in a ring R.
Definition 8
Let R be a ring (not necessarily commutative) and let a  R. Then a is said to
be a left zero divisor iff there exists a0 and b  0 in R such that ab = 0. Similarly a
is said to be a right zero divisor iff there exists an element c  0 in R such that
ca = 0. The element a is said to be a two-sided zero divisor or simply a zero divisor if
a is both a left zero divisor and a right zero divisor. (Note that 0 is always a zero
divisor in R, if R has at least one non zero element). We now show in passing that
no field or skew-field contains a non-trivial zero divisor. This property of a field or a
skew-field follows from the following more general result.
Result
If R is a ring with identity e (not necessarily commutative) and a  R, then a
has no multiplicative[ left/right] inverse in R if a is a [left/right zero divisor] inverse.
Proof
Suppose a has a multiplicative left inverse, say a-1R such that a-1 a=e. Then a
cannot be a left zero divisor, for otherwise there exists an element b  0 in R such
that ab = 0. But this implies that a -1(ab) = a -10=0, and hence (a-1a) b=0, i.e. eb = 0
or b = 0 which is contradiction to the assumption that b0. Thus a cannot have a
left inverse. Similarly, it can be shown that if a has a right inverse, say a 1-1R, such
that aa1-1=e then a cannot be a right zero divisor. Hence the result. ■
Problem I
Prove that in a ring R the product of any two (and hence the product of any
finitely many) left (right) zero divisors is again a left (right) zero divisor.
Problem 2
The units (or invertible elements) in a ring with identity form a group under
multiplication.
(By a unit in ring R with identity e(say) is meant that an element u which has a
multiplicative inverse v in R, such that uv=vu=e). We have the following important
fact concerning the ring Zn of residue classes modulo n ■.
Lemma 7.3
The ring z n is a field if and only if n is a prime.
Proof
First assume that n is a prime. We show that Z n is a field. The elements of Z n
are the class consisting of all integers (of the form n+i, Z ) which leaves the
remainder i on division by n, (i.e.) which are congruent to i modulo n. Suppose for
a, b  Z n , a b = 0 then ab  0 (mod n). Since n is a prime the above implies that
either a or b is a  0 (mod n) or b  0 (mod n) . That is either a or b is equal to o .
70
Zn is a finite (commutative) ring containing no non-trivial zero divisors (i.e.) Z n
is a finite integral domain and hence is a field. Conversely suppose Zn is a (finite)
field. We prove that n must be a prime. Suppose not, then we can write n = pq
where 1 < p, q < n. (i.e) ( p and q are proper divisors of n ) Now pq  (mod n) implies
p q  0 and hence p q  0 . But p q  0 , since, 1 < p, q<n. This means that Zn has
non-trivial zero divisors, namely p, q. But Zn being a field cannot contain non-trival
zero divisors and so we have a contradiction. Hence n must be a prime, as claimed.
Definition 8.2
Let R be a ring. Then R is said to be of characteristic m (>0), if m is the least
positive integer such that ma=0 for all aR. R is said be of characteristic zero if
there exists no positive integer m such that ma = o for all aR.
Note: In the case of an integral domain R, if for some integer m and for some
non-zero element aR, ma=0, then mx=0 for all xR. To see this fact, consider
(mx) a = m (xa) (by right distributive law)
= x(ma) (by left distributive law)
= x0 = 0
Hence mx = 0, since a  0 and R is an integral domain.
Thus in the case of an integral domain R, the characteristic of R is the least
positive integer m (if it exists) such that ma = 0 for some non-zero element a in R. If
no such positive integer m exists then R is of characteristic zero. ■
We will now prove the following problem, which gives a definite information
about the characteristic of an integral domain.
Problem 3
Prove that the characteristic of an integral domain is either zero or a prime.
Proof
Let R be an integral domain. If the characteristic is zero, there is nothing to prove.
(However, we may note that when a ring R, not necessarily an integral domain, is of
characteristic zero then R must be an infinite ring). For otherwise if the ring R is finite
and has only finite number of elements, then by a corollary under Lagrange’s theorem
on groups, we get that na = 0 for all aR, which means that the characteristic of R is
either n or less then n. (In fact a divisor of n is not zero). Now, the integral domain R is
of characteristics n  0. Then we claim that n must be a prime. For otherwise, let
n = pq where 1<p, q<n. Then for any non-zero element a in R, we must have
(pa) (qa) = (pq) a2 (using the second distributive laws)
= na2 = 0 ( since n is the characteristic of R)
The above implies that either pa = 0 or qa = 0. However both the possibilities
are impossible since both p, q are < n and pa = 0 will imply that px = 0 for all xR.
So also qa =0 will imply that qx = 0 for all xR. Thus n must be a prime and the
problem has been established. ■
71
The following problem for which the hint given may be used, proves that the
commutativity of addition in a ring R with identity is a consequence of the
remaining axioms which defines R.
Problem 4
Let R  < R, +,  > be an algebraic system satisfying all the axioms for a ring
with identity with the possible exception of the commutativity of addition +
(i.e.) < R, + > is a group not necessarily abelian, < R ,  > is a semigroup with
identity (known as a monoid) and multiplication ‘’ is both left distributive and right
distributive over addition ‘+’. Show that R is a ring.
The proof is left to the reader.■
Homomorphism:
Just as for groups, we introduce now the important notion of homomorphism
between rings.
Definition 8.3
A mapping  of a ring R into a ring R is said to be a homomorphism of R into R
if
 (a+b) = (a) + (b) and
 (a.b) =  (a). (b), for all a, bR.
In the above definition it must be noted (as in the case of groups) that +
and . that occur in the left hand sides of the above conditions (1) and (2) are those
operations of + and . in the ring R, while the + and . that occur on the right hand
sides are those of the ring R.
If there exist a homomorphism of a ring R onto a ring R, we write R ~ R is a
homomorphic image of R. More briefly, a homomorphism of a ring R into (onto) a
ring R is a mapping of R into (onto) R which preserves the operations. A
homomorphism of ring R onto a ring R is also known as an epimorphism (more
precisely, a ring epimorpism). If the homomorphism  of a ring R into a ring R is
one-to-one, (i.e.)  (a) = (b)  a = b, where a, b  R, then the homomorphism  is
said to be a monomorphism (more precisely, a ring monomorphism). A
homomorphism : RR which is both a monomorphism and an epimorphism is
called an isomorphism and in such case we write R~R and say that the ring R is
isomorphic with ring R. It is easy to see that any two rings which are isomorphic
have the same algebraic properties pertaining to addition and multiplication. For
instance if the rings R and R are isomorphic, then (i) R is commutative iff R is (ii) R
has identity iff R has, (iii) R is an integral domain iff R is, (iv) R is a field iff R is, (v)
R is finite iff R is and so on. Such a statement as above cannot be made when
there exists only a homomorphism and nor an isomorphism of the ring R into (or
onto) the ring R. We will say more about this a little later.
Let : RR be a homomorphism of a ring R into a ring R. Since  preserves
the additive structures of the rings R and R, making use of the known properties
72
of a group homomorphism, we have the following facts: (i)  (0) = 0, where 0 in the
left hand side is the zero element in R and 0 in the right-hand side is the zero
element in R, (ii)  (-a) = -(a) for all aR.
Definition 8.4
For the homomorphism  : R  R, the set K = { aR : (a) = 0, the zero in R }
is (as in the case of groups) called the kernel of the homomorphism .
One can immediately notice (from group theory) that  is an isomorphism iff K
is trivial i.e. K contains only the zero element of R. Further K, under addition, is a
(normal) sub group of the additive group <R, +> of the ring R. Also for any aK and
rR both ra and ar are in K, since (ra) = (r) (a)=(r)∙0 = 0 and similarly (ar) = 0.
These properties of K are used to define the notion of an ideal in the ring R. This
notion being analogous to the notion of normal subgroups in Group Theory.
Definition 8.5
Let R be any ring. A non-empty subset A of R is said to be a two sided ideal or
simply an ideal of R if
(i) x, y A  x – y  A (Module property)
(ii) xA, rR  rxA (Left-multiple property)
(iii) xA, rR  xrA (Right multiple property)
Note: A is said to be a left ideal of R if conditions (i) and (ii) stated above are
satisfied and is said to a right ideal of R if conditions (i) and (iii) are satisfied. Thus
an ideal of R is both a left ideal and a right ideal of R. Also one may prove that
when the ring R is commutative both the notions of left ideal and right ideal
coincide and all ideals in R are only two sided ideals.
Definition 8.6
A non-empty subset A of a ring R is said to be a sub-ring of R, if A is also a
ring under the same operations of R.
It is easy to prove that a non-empty subset A of a ring R is a sub-ring of R iff
(i) x, yA  x - yA and (ii) x, yA  xy  A
For the ring Z of integers, all multiples, of a fixed integer n(say) form the
subset nZ which is clearly a sub-ring of Z. So also all the integers from a sub-ring
of the ring Z[i] of Gaussian integers. The real numbers of the form a+b c where
a, b Z and c, a fixed positive integer which is not a perfect square form a sub-ring
of the field R of real numbers. Many examples of sub-rings can be given. One can
also show that every ideal A in a ring R, (even a one-sided ideal) is a sub-ring of R,
though not conversely. We give below examples of (i) a sub-ring which is not an
ideal (even one-sided) (ii) a left ideal which is not a right ideal, (iii) a right ideal
which is not a left ideal, besides examples of two-sided ideals.
73
Example 8.7
Let F2 denote the (non-commutative) ring of all real 2  2 matrices (vide:
Example 8). Let A be the set of all real upper triangular matrices of the form
a b  then one can easily check that A is only a sub-ring of F and is neither a left
2
 o c 
nor a right ideal of F2. Similarly the set B of all real lower triangular matrices of the
form a o , (which are transposes of the matrices in A) is also, only a sub-ring of F2
b c 
and is neither a left nor a right ideal of F2. On the other hand if C is the set of all
real matrices of the form a b  , one can easily show that C is a right ideal and not
 o c 
a left ideal of F2. Similarly the transposes of the above mentioned matrices, viz. the
set D of all real matrices of the form a o is a left ideal and not a right ideal of F2.
b o
Analogous examples can be given from the (more general) ring Fn of all nn real (or
rational or complex) matrices.
Example 8.8
In the ring Z of integers the subset nZ, where n is a fixed integer, is not only a
sub-ring of Z, but even a two sided ideal in Z. It can be shown that every ideal in Z is of
this form, viz. nZ = { nm: mZ } for some fixed integer n. For this reason Z is a principal
ideal ring about which we will mention in greater details in the next lesson.
Example 8.9
Let R be any ring not necessarily commutative. If a is any fixed element in R,
the set aR = { ax: xR } can be shown to be a right ideal of R. In general aR need not
be a left ideal of R. Similarly the set Ra = { xa: xR } is a left ideal of R and is, in
general, not a right ideal of R. Moreover, in general, a need not belong to either the
right ideal aR or the left ideal Ra. For example, if R is the (commutative) ring of all
even integers, the subset 2R (of all multiples of 4) is an ideal of R not containing 2.
However, if the element a in the ring R belong to the centre, viz. the set
Z(R) = { zR: xz=zx,  xR }, then aR = Ra will be a two-sided ideal of R.
The following standard problem gives an important characterisation of a
division ring in terms of right or left ideals, (one may also note in passing that, any
ring R, not necessarily commutative but however containing atleast one non-zero
element, always has two ideals viz. the null ideal (0) containing the zero element 0
only and the unit ideal which is the whole ring R.)
Problem 5
Let R be a ring with identity element not necessarily commutative. Then R is a
division ring if and only if the only right (left) ideals of R are the null ideal (0) and
the unit ideal R.
Proof :
Suppose the only right ideals in R are the null ideal (0) and the unit ideal R.
Then we have to show that R is a division ring. It suffice to show that the non-zero
74
elements of R form a group under multiplication. In our present case since R has
the (multiplicative) identity element e (say) it is enough if we show that every non-
zero element in R has a (multiplicative) right (or left) inverse and that the product of
any two non-zero elements in R is again a non-zero element. First we will show that
by assuming that a, bR and a, b are  0. Now aR is a right ideal of R and since
a = ae  aR and a0, aR=R as R contains only the 2 ideals viz. (0) and R. Similarly
bR=R. Suppose, if possible, let ab = 0. Then (ab) R = (0). But (ab) R = a (bR) = aR =
R  (0). Hence we obtain a contradiction and so ab  0. Thus the non-zero elements
of R form a semigroup under multiplication with identity e. Also since aR = R,
there exists an element a (a right inverse of a) in R such that aa = e. Hence, by
definition number 1 for a group ( vide: Lesson 2 ), the non-zero elements in R form
a group under multiplication, as required. Conversely suppose R is a division ring,
then we prove that R has only two ideals namely the null ideal (0) and the unit ideal
R. Suppose A is a non- null ideal of R and a  A, a  0. then for any b  R, there
exists an element x  R such that ax = b. (This fact follows from we get that b = ax
 A). But b is any arbitrary element of R, then it follows that R  A  R. Thus A = R,
the unit ideal, as desired. Hence the problem. ■
Bearing in mind that the definition of an ideal (left or right or two-sided), it is
not difficult to show that if A and B are both left (or right two-sided) ideals in a ring
R, then A  B is also a left (or right or two-sided) ideal in R. More generally the
intersection of any non empty family of left (or right or two-sided) ideals in R is also
a left (or right or two-sided) ideal in R. This being so, we can define the left (or right
or two sided) ideal generated by any non-empty subset M of R as the intersection of
all left (or right or two sided) ideals of R which contain M. The above definition is
meaningful in as much as that there is atleast one left (right, two-sided) ideal in R,
viz, the unit ideal R which contains M. When M is the singleton set {a}, the left
(right, two-sided) ideal generated by M is called the principal left (right, two sided)
ideal generated by the element a and is usually denoted by (a)l (or (a)r, or (a)), as the
case may be. ■
Problem 6
If R is a ring with identity, not necessarily commutative then show that, for
elements aR, aR is the left ideal generated by a. Ra is the right ideal generated by
a and the set of all finite sums of the form  ras is the two sided ideal generated
r , sR
by a
However, if R has no identity element then show that
the set { na+ra|nZ, rR } is the left ideal (a)l,

the set { na+ar|nZ, rR} is the right ideal (a)r, and

the set { na+ras : nZ ,  being a finite sum } r, s R is the two sided ideal (a).
75
It may be noted that the left (right, two-sided) ideal generated by the non-
empty subset M of R is the set-theoretically smallest left (right or two sided) ideal in
R which contains M.
Problem 7
If A, B are two left (right, two-sided) ideals in a ring R, not necessarily
commutative, then show that A+B = { a+b : aA, bB } is also a left (right two-sided)
ideal in R, which contains AUB.
Problem 8
Let A, B are two ideals in a ring R. Let AB denote the set of all elements that
can be written as finite sums of elements of the form ab with aA, bB. Then prove
that AB is also an ideal in R and that AB  AB.
Note: The above problems 6, 7, 8 which concern the algebra of ideals are almost
straight forward and follow from definition of ideals. So you may try to do them all.
We saw earlier that the definition of an ideal was motivated by the module and
multiple properties of the kernel of any ring homomorphism. Just as for groups,
where we mentioned the close relationship that exists between group
homomorphism, normal subgroup and factor group, we proceed to explain in the
case of a ring, a similar relationship that exists between ring homomorphism, ideal
and factor-ring or residue-class ring. We now introduce the concept factor ring (or
quotient ring or residue – class ring).
Factor ring or quotient ring
Let R be a ring and A be an ideal. ( i.e. a two-sided ideal ) of R, A being a
(normal) subgroup of the additive group <R, +>, we can talk of cosets A + x, xR.
These cosets, which are now called residue classes of A. We know, from a group
under the well-defined addition given by (A+x)+(A+y) = A+(x+y) for x, yR. Now, we
define multiplication of residue classes by setting (A+x)(A+y) = A+(xy). That the above
definition of multiplication of residue classes is well-defined, can be seen as follows:
Suppose A + x = A + x and A + y = A + y where x, y, x, yR.
Then we show (A+x) (A+y) = (A+x) (A+y), i.e. A+xy=A + xy
Now A+x = A+x  x-x A …(1)
A+y = A+y  y-y A …(2)
Since A is an ideal ( x-x)  A  (x-x) y  A
 xy-x y  A …(3)
Again (y-y)  A  x(y-y)  A
 xy - xy  A …(4)
The above implications (1), (2), (3), (4) shows that (xy-xy) + (xy -xy)  A.
i.e. xy - xy  A, and hence A+xy = A+xy, as desired. ■
R
Now we make the formal definition of factor ring
A
76
Definition 8.9
Let R be any ring, not necessarily commutative and let A be two sided ideal of
R
R. Let = { A+x : xR } be the set of all residue classes of A ( or modulo A ). Define
A
R
addition + and multiplication ∙ on by setting (A+x) + (A+y) = A + (x+y),
A
R
(A+x).(A+y) = A + xy. Then become a ring under the above operation of addition
A
R
and multiplication of residue classes of A and this ring is called the quotient
A
ring or factor ring or residue class ring of R modulo A.
We already know from group theory that, if we restrict ourselves to the additive
R
structures only, then the mapping : R  defined by (x) = A + x for xR is the
A
canonical (group) homomorphism. The same mapping turns out to be a ring
R
epimorphism also if we take the ring structures of R and into consideration.
A
Thus we have the following important example of a ring epimorphism.
R
For any two sided ideal A in R the mapping : R  defined by (x) = A+x is a
A
ring epimorphism known as the canonical (ring) homomorphism. In fact, for x, yR,
(x+y) = A + (x+y) = (A+x) + (A+y) = (x).(y) so that  preserves, both the operations
of addition and multiplication. Moreover  is clearly surjective, as every element in
R
is of the form A+x and (x) = A+x.
A
Applying the same procedure as for groups one can establish the following
(analogous) fundamental ring homomorphism theorem.
6.3.2 THEOREM : HOMOMORPHISM OF RINGS
Let  : R R be a homomorphism of a ring R onto a ring R and let K be the
R
kernel of . Then K is a two sided ideal of R such that the residue clas ring is
A
isomorphic with R (in other words, any homomorphic image of a ring R is isomorphic
R
with residue class ring where K is the kernel of the homomorphism). Conversely if
K
K is any two sided ideal of R then there exists a homomorphism (viz. the canonical
R R
homomorphism : R of the ring R onto the ring such that the kernel of the
K K
homomorphism  is precisely K. ( Write down the proof of this basic theorem ). ■
The following theorem is analogous to the group – theoretical results, namely
Lemma 9 and Theorem 5 of Lesson 3 and its proof is an exact verbatim translation
of the proofs for the above mentioned group theoretical results into the language of
77
rings. So this theorem is also stated without proof and you are invited to write
down the proof.
Theorem 3.2.1
Let : R  R be a homomorphism of a ring R onto a ring R and let K be the
kernel of . If A is any sub-ring of R, let A = { xR: (x)A }. Then A is a sub-ring
of R containing K and A is an ideal in R iff A is an ideal in R iff A is an ideal in R.
Moreover there is a one-to-one correspondence between the ideals A of R and the
ideals A (as defined earlier) of R which contain K. Further if the ideals A of R and A
R
of R correspond under the above mentioned one-to-one correspondence, then is
A
R
isomorphic with .■
A
We conclude this lesson with some standard examples of ring homomorphism
and certain observations on them.
Example 3.2.2
Let R and R be two arbitrary rings and let : R R be a mapping defined by
(x) = 0  xR. It is quite easy to see that  is (trivially) a homomorphism. This
homomorphism is called the zero-homomorphism and is never an epimorphism
unless R is the trivial ring consisting of the zero element only. Similarly  is never a
monomorphism unless R is the trivial ring consisting of the zero element only.
Similarly  is never a monomorphism unless R is the trivial ring consisting of the
zero element only.
Example 3.2.3
Let R be a ring and I : R  R be the identity mapping given by I(x) = x, xR.
Then clearly I is an isomorphism and is really the identity automorphism of the
ring R.
Note: An automorphism of a ring R is an isomorphism of R onto itself.
Example 3.2.4
Let R  Z[i] be the ring (in fact an integral domain) of all Gaussian integers
a + bi where a, b Z. Consider the mapping : RR defined by  (a+bi) = a-bi. Then
one can easily check  is an automorphism of K, which is not the identity
automorphism. So also, if R is the field C of complex numbers, then the mapping
which maps every complex number z into its conjugate z is an automorphism,
which is not the identity automorphism. Similarly, if R is the ring (in fact, an
integral domain) of all real numbers of the form a+b c ( where a,bZ and c is a
fixed positive integer which is not a perfect square) – vide: Example 5-the mapping
: R R defined by  (a+b c ) = a-b c can be easily verified to be an
antomorphism of R, which is not identity automorphism.
78
Example 3.2.5
If, as usual, Z denotes the ring (in fact, an integral domain) of integers and Z n
(where n is a fixed positive integer) denotes the ring of residue classes modulo n-
vide: Example 7, consider the mapping : Z Zn defined by  (m) = m where m
denotes the residue class modulo n determined by the integer m, and is the set of
all integers which are congruent to m modulo n. Since n  m  m for CZ this
mapping  is certainly not one-to-one, though however it is onto. Since one can
easily check that p  q  p  q and pq  pq where p, qZ, the mapping  is a ring
epimorphism, whose kernel is the ideal nZ of Z. Thus Zn is a homomorphic image of
Z, with kernel nZ and hence , using the basic homomorphism Theorem 1 , we are
Z
that  Zn . Earlier we had mentioned that Zn is a field if and only if the integer
nZ
n is a prime. Thus by choosing a nonprime n, we find that the homomorphic image
of an integral domain, viz., Z used not be an integral domain, as Zn contains non-
trivial zero divisors when n is not a prime. So also, one can show that the
homomorphic image of a ring with non-trivial zero divisors can be a ring without
non-trivial zero divisors.
Example 3.2.6
Let R be the set of all continuous real valued functions defined on the closed
unit interval I = [0, 1]. Define addition ‘+’ and multiplication ‘.’ on R, as usual, by
setting, (f+g) (x) = f(x) + g(x) and (fg)(x) = f(x) g(x) xI, where f, gR. Then since sum
and product of two real valued continuous functions, one can easily check that R
becomes a commutative infinite ring with identity, under the above defined
operations of ‘+’ and ‘.’ . Now let F be the field of real numbers. Define a mapping
: RF by setting (f) = f() where fR and , a fixed real number in I. (Hence f() is
the value of the function f  f(x) when x = ). Then it is not difficult to verify that  is
a ring epimorphism, whose kernel consists of all functions in R which vanish at
x = . That is, the mapping  is surjective follows from the fact that the ring R
contains all constant functions.
Solved Problems
1. Prove that any field is an integral domain, but not vice-versa.
Solution
The statement suffices to show that a field F has no zero divisors.
Let a, b  F such that a  o & ab = 0.
ab = 0  a-1(ab) =a-1 0 (since a  0  a-1 exists)
 (a-1a) b = 0
 1.b = 0
 b = 0.
Similarly, let ab = 0 & b  0.
79
Since b  0, b-1 exists & we have
ab = 0  (ab) b-1 = 0.b-1
 a(bb-1) = 0
 a.1 = 0
a=0
Thus F has no zero divisors.
Hence every field is an integral domain.
But the converse is not true. For example the ring of integers which is
an integral domain is not a field since the only inevertible elements of
the ring of integers are 1 & -1. ■
2. A ring R is said to be a Boolean ring if a2 = a for all aR. Prove that any
Boolean ring is of characteristic 2 and is commutative.
Solution
Let R be a Boolean ring
Now aR  a+aR
Therefore (a+a)2 = (a+a) (given)
 (a+a) (a+a) = a+a
 (a+a) a+(a+a) a=a+a (Left distributive Law)
 a2 + a2 + a2 + a2 = a + a (Right distributive law)
2
 (a+a) + (a+a) = a+a ( a  a )
 (a+a) = 0 (Left cancellation Law)
 Any Boolean ring is of characteristic 2.
Now (a+b)2 = (a+b)
 (a+b) (a+b) = (a+b)
 (a+b) a+ (a+b) b = a+b (Left dist. Law)
 a2 + ba + ab + b2 = a +b (right dist. law)
 (a + ba) + (ab + b) = a+ b (since a 2 = a; b 2 = b)
 (a + b) + (ba + ab) = ( a + b) + 0
(commutativity & associativity of addition)
 ba + ab = 0 (Left Cancellation law)
 ba + ab = ba + ba (Therefore R is of Characteristic 2)
 ab = b a (Left Cancellation law) ■
3. If R is a ring and a  R. Let r(A) = { xR : ax = 0  aA }. Show that r(A) is
a right ideal of R.
Solution
We know that 0  R is such that a0 = 0r(A)  
80
Let x1, x2  r(A)  ax1 = 0; ax2 = 0  aA.
Consider a (x1-x2) = ax1 – ax2 = 0 – 0 =0  aA
 x1 – x2  r(A)
 r(A) is a group under addition
In order to show that r(A) is a right ideal of R, we have to show that
x  r(A), yR  xy r(A). But x r(A)  ax = 0,  aA.
 a(xy) = (ax)y
= 0y
=0
Hence xy r(A)
Thus xr(A), yR  xyr(A)
r(A) is a right ideal of R ■
4. Let A be any ideal in a ring R and let [R:A] denote the set { xR : rxA, 
rR}. Prove that [R:A] is an ideal of R and that A [A.A]
Solution
Since 0R is such that r0 = 0A for all rR and hence [R:A] ≠ 
Let x1, x2 be two elements of [R : A]
Then rx1 A for all rR and rx2A for all rR
Since A is an ideal, rx1 A, rx2 A  rx1 – rx2 A
 r(x1-x2) A
 x1-x2  [R:A]
Let x be any elements of [R:A],r and s be any element of R. Then rx A
for all rR.  (rx) sA for all rR
r(xs) A for all rR
 xs  [R:A]
Also rxA for all rR
 (rs) xA (since sR)
 r(sx)A for all rR
 sx  [R:A]
Thus x [R:A], sR  xs [R:A] sx[R:A] Hence [R:A] is an ideal of R.
Also y  A  yr  A for all rR
 y[R:A]
A [R:A] ■
5. If R is a ring and A is a left ideal of let (A) = {xR: xa = 0, for all aA}.
Prove that (A) is a two sided ideal of R.
Solution
81
Since 0R is such that 0a = 0 aA. Therefore (A) 
Let x1, x2  (A). Then x1a = 0 & and x2a = 0  aA
Consider (x1-x2)a =x1a-x2a=0-0=0, aA.  x1-x2 (A)
Let x   (A): rR
Then xa=0,  aA
 r(xa) = r0, aA
 (rx)a = 0, aA
 rx  (A)
Also, since A is a left ideal and raA,  rR,  aA
x(ra) = 0 (x (A) )
 (xr) a = 0  aA
 xr(A)
Hence (A) is an ideal of R. ■
6. If R is a ring with unit element 1 and  is a homomorphism of R onto R,
prove that (1) is the unit element of R.
Solution
Since  is a homomorphism of R onto R, R is a homomorphic image
of R. If 1 is the unit element of R, then  (1) R.
Let a  R. Then a =  (a) for some aR ( is onto)
  (1) a =  (1) (a)
=  (1.a)
= (a) = a
Also a (1) = (a) (1)
=  (a.1)
=  (a)
= a
Hence  (1) is the unit element of R ■
7. If R is a ring with unit element 1 and  is a homomorphism of R into
an integral domain R such that Kernel of   R, prove that (1) is the unit
element of R.
Solution
Let  be a homomorphism of a ring R into an integral domain R.
Then, Ker  = {x : xR and (x) = 0 R}
Since Ker   R,  a  R such that (a)  0R
We have (1) (a) = (1a) = (a)
Let b be any element of R
82
Now (a) b = (a) b
 (1) (a) b = (a) b
 (a) [(1)b] = (a) b (R is commutative)
 (a) [(1)b] = (a)b = 0
 (a) [(1)b–b] = 0
 (1)b – b = 0 (since (a)  0 & R is without zero divisors)
 (1) b = b = b (1) (since R is commutative)
 (1) is the unit element of R ■
6.4. REVISION POINTS
Ring, Homomorphism of ring, Kernel of homorphism, Ideal, Subring, Quotient
ring
6.5. INTEXT QUESTIONS
1. Prove that any finite integral domain is a field
2. Prove that in a ring R the product of any two (and hence the product of
any finitely many) left (right) zero divisors is again a left (right)zero divisor.
3. If A, B are two left (right, two-sided) ideals in a ring R, not necessarily
commutative, then show that A+B = { a+b : aA, bB} is also a left (right
two-sided) ideal in R, which contains AUB.
4. A ring R is said to be a Boolean ring if a2 = a for all aR. Prove that any
Boolean ring is of characteristic 2 and is commutative.
5. If R is a ring and A is a left ideal of R and let (A) = {xR : xa = 0, for all
aA}. Prove that (A) is a two sided ideal of R.
6. If R is a ring with unit element 1 and  is a homomorphism of R into an
integral domain R such that Kernel of   R, prove that (1) is the unit
element of R.
6.6. SUMMARY
Ring, Types of rings, Homomorphism of ring, Kernel of Homomophism, Ideal,
Subring, Quotient ring are explained
6.7. TERMINAL EXERCISES
1. An element a in a ring R is said to be nilpotent if an = 0 for some positive
integer n. Does the Z108 of all residue classes modulo 108 have non-zero
nilpotents? If so, what are they?
2. If R is a non commutative ring with a unique left identity, i.e. an element e
such that ea = a for all a  R, then show that the left identity is also the
identity for R.
3. Let R be a ring with identity1. Define new operations  and  on R by
setting a  b = a+b+1, ab = a+b+ab and let R  { R, ,  }. Prove that R is
also a ring with identity and R is isomorphic with R .
83
6.8. SUPPLEMENTARY MATERIALS
1. Algebra by Micheal Artin, Prentice Hall of India, New Delhi, 1994.
2. Lectures in Abstract Algebra Volumes I, II and III by N. Jacobson, D. Van
Nostrand Co., New York, 1976.
6.9. ASSIGNMENT
1. Prove that the ring Zn is a field if and only if n is a prime
2. Theorem : Homomorphism of Ring
6.10. SUGGESTED READINGS / REFERECNE BOOK
Algebra – By Michel Artin – PHS Learning P Ltd
6.11. LEARNING ACTIVITIES
Students are requested to attend the P.C.P classes and work out the problems
given in the lesson.
6.12. KEYWORDS
Ring, Homonorphism, Kernel, Ideal, Subring, Quotiant ring

84
LESSON – 7

IDEALS
7.1. INTRODUCTION
In this Lesson we propose to study certain special types of ideals and rings
and discusses some of their important properties.
The theorem 1 on ring homomorphism (given in the last lesson) gives us a
complete information about all the possible homomorphic images of a given ring.
Applying this theorem and recalling the fact that the only ideals in any division ring
are the trivial ones, viz., the null ideal (O) and the unit ideal of R. We can state that
any division ring (and in particular, any field) has only trivial homomorphic images,
namely itself and the zero ring, consisting of only the zero element. Thus a field is
the most desirable kind of ring, in that it cannot be simplified further by applying a
homomorphism to it. Now we define two types of ideals in a commutative ring R
and indicate their inter relations, apart from suitable examples to illustrate these
ideals, it happens that the residue class ring modulo, one of these types of ideals, is
always a field.
7.2. OBJECTIVES
Ideal, Maximal ideal,Imbedding, Euclidean ring, Principla ideal ring, greatest
common divisor, Commutative ring with identity, prime, relatively prime are
discussed in detail.
7.3. CONTENTS
3.1 Definition – Ideal
3.2 Definition – Maximal ideal
3.3 Definition - Imbedded
3.4 Euclidean rings
7.3.1. DEFINITION - IDEAL
An ideal P in a commutative ring R is said to be a prime ideal if P is not a unit
ideal and whenever abP, where a, b  R then at least one of the factors a or b is
in P. ■
7.3.2. DEFINITION – MAXIMAL IDEAL
An ideal M in a ring R is said to be a maximal ideal of R, if M is not the unit
ideal and whenever N is an ideal in R such that M  N  R, then either N = M or
N = R. ■
The meaning of maximal ideal (given above) is that an ideal of R is maximal if
it is impossible to squeeze an ideal between it and the whole ring R. It can be
proved, by using Zorn’s Lemma or axiom of choice, that any commutative ring with
identity has at least one maximal ideal. In fact, any ideal is contained in a unique
maximal and this can be proved. Moreover, a ring may possess many maximal
ideals and this fact will be illustrated with the help of the ring Z of integers.
85
Theorem 3.2.1
If R be a commutative ring, then an ideal P of R is a prime ideal iff the residue
R
class ring is an integral domain.
P
Proof
R
Let be an integral domain. We have to prove that P is a prime ideal. It
P
suffices to show that xy  P implies that at least one of the factors x or y is in P (by
definition of a prime ideal). Now xy  P implies xy + P = P, which in turn implies
R
(x+P)(y+P)=P ( Here note that the elements of are residue classes x+P where xR
P
R
and that (x+P)(y+P) = xy+P ). But P is the zero element in the ring , which is
P
assumed to be an integral domain. Hence either x+P = P or y+P=P which means
that either xP or yP as desired. Therefore P is a prime ideal. Conversely, assume
R
that P is a prime ideal. We shw that is an integral domain. For this it is enough
P
R
if we show that (x+P)(y+P) =P (the zero element in ) implies that either x + P = P or
P
y + P = P. Now (x+P) (x+P) =P implies xy+P=P, and hence xyP. Since P is a prime
ideal, this means that either x or y is in P and hence either x+P or y+P is same as P,
R
as required. Thus becomes an integral domain. (Note that R being
P
R R
commutative, is also a commutative ring and so the ring is a commutative
P P
integral domain, as required in the definition of an integral domain). ■
Before obtaining the important result concerning the “iff” condition for an ideal
to be a maximal ideal of a commutative ring with identity (vide: Theorem 2), we list
a few examples of prime ideals and non prime ideals in the ring Z. Incidentally
these examples will justify the term “Prime ideal”.
Take the ring Z of integers. If A is a null ideal of Z, we show that A is a
principal ideal, generated by some suitable integer aA. Since A is a non-null ideal,
A contains non-zero integers. Of the two integers m and – m, one of them is
certainly positive. Thus A contains positive integers. Let a be the smallest of the
positive integers in A. (Such an integer a exists, as the set of all positive integers is
known to be well ordered). We now claim that this positive integer a generates the
ideal A. For, if mA then, by division algorithm available in Z, we can find two
integers q and r such that m = qa + r where o < r < a. Since m, a  A and A is an
ideal, we find that m - qaA. i.e., rA. Hence r = 0 for otherwise there exists a
positive integer r < a in A, contradicting the fact that s is the least positive integer
in A. Thus r = 0 and thus we get m = qa, which means that A = aZ is the principal
86
ideal in Z generated by a. Because of this property of Z viz. that any ideal in Z is a
principal ideal. Z is a principal ideal ring-this concept of a principal ideal ring will
be defined subsequently. Now we show that an ideal P in Z is a prime ideal iff it is
generated by a prime number. Suppose P = pZ is the ideal generated by the prime
p and suppose ab  P where a, bP. then p|b. This mean that either a or b is in
P = pZ. Hence, by definition, P becomes a prime ideal. On the other hand any
ideal A of Z generated by a non-prime is not at prime ideal. In fact, suppose A = aZ
is an ideal generated by a which is not a prime then we can choose two integers x
and y such that a|xy , a does not divide either x or y. For example, suppose a = 12
(a non prime) we can choose x =4, y =9 so that a|xy, though a divides neither x nor
y. Hence the ideal A cannot be a prime ideal, as xyP but for both x and y. Hence
the ideal A cannot be a prime ideal, as xyP but for both x and y are not in P. Thus
we see that any ideal in X is prime ideal iff it is generated by a prime number. This
property is one of the important properties of any Euclidean ring, a concept to be
defined subsequently and we will see later that Z is an Euclidean ring. Similarly in
the case of the ring Z[x] of all polynomials in x over Z.
Ring Z[x] will be considered in the next lesson – any ideal is a prime ideal iff it
is an ideal generated by an irreducible polynomial.
Now we will obtain the “iff” condition for an ideal M to be a maximal ideal in R.
In this connection we have the following important theorem.
Theorem 3.2.2
If R be a commutative ring with identity, then an ideal M in R is a maximal
R
ideal iff the residue class ring is a field.
M
Proof
R
Let M be an ideal of R such that the residue class ring is a field. Since a
M
field has only two ideals, viz. the null ideal and the unit ideal, the residue class ring
R
is . Now by theorem 2 in Lesson 5, there is one-to-one correspondance between
M
R
the set of ideals of and the set of ideals of R which contains M. The ideal M of R
M
R
corresponds to the zero ideal M of and the unit ideal R of R corresponds to the
M
R R
unit ideal of . Hence there is no ideal between M and R other than these two,
M M
which follows that M is a maximal ideal. Conversely, let M be a maximal ideal.
R
We must show that is a field. By the one-to-one correspondence mentioned
M
above, it follows that (since there is no ideal between M and R other than these two)
R
the residue class ring has only two ideals, namely the null ideal M and the unit
M
87
R R
ideal. Hence is a division ring and since R is a commutative ring turns out
M M
to be a field. This concludes the proof of the theorem. ■
Corollary 3.2.3
In any commutative ring with identity every maximal ideal is a prime ideal.
Proof
Let R be a commutative ring with identity and let M be a maximal ideal of R.
R
Then by the above theorem is a field and an integral domain. Therefore, by
M
theorem , M turns out to be a prime ideal and hence the corollary ■
Note: Even though any maximal ideal (in a commutative ring with identity) is
always a prime ideal, the converse of the statement is false. In fact, we can give an
example of a commutative ring with identity, in which not all prime ideals are
maximal. However, in the case of the ring Z of integers and also in the case of any
finite commutative ring (with identity), all prime ideals are also maximal. These
facts are proved in the following problems.
Problem 1
Give an example of a commutative ring with identity in which not all prime
ideals are maximal.
Solution
Let Z[x] (as usual) denote the set of all polynomials in the indeterminate (or
variable) x over Z (i.e. with coefficients from the ring Z of integers). One can easily
see that Z[x] becomes a commutative ring with identity, under the usual operations
of addition and multiplication of polynomials. Let P denote the sub set of Z[x]
consisting of all polynomials in Z[x] which are divisible by x ; i.e. polynomials which
lack the constant term. One can easily verify that P is an ideal in Z[x]. In fact it is
the principal ideal generated by the polynomial x. This ideal P is moreover a prime
ideal, because, whenever uv  P, u,v  R[x], it is easy to see that at least one of
the polynomials u or v must lack the constant term. Let N be the set of all
polynomials in Z[x], whose constant term n is either zero or an even integer. It is
not difficult to check that N is an ideal (even a prime ideal) in Z[x], with
the property that P  N  Z[x], so that N turns out to be a proper ideal in Z[x],

which contains P properly. Thus P cannot be a maximal ideal. Thus Z[x], is an


example of a commutative ring with identity, in which not all prime ideals are
maximal. ■
Remark
The ideal P, referred to above is properly contained in many other proper ideals
of Z[x]. In fact, if p is any prime and Np denotes the set of all polynomials in Z[x]
whose constant terms are either 0 or P, one can easily check that Np is a proper
88
ideal (even a prime ideal in Z[x] ) which contains P properly. This ideal Np is the
ideal generated by the set (xP) in Z[x].
Proplem 2
Prove that in the Ring of integers, prime ideals are maximal.
Solution
Let P be any prime ideal in the Ring Z of integers. Then there exists a prime
number p which generates the ideal P. i.e P = pZ. Now by example (iv) given in
Z
Lesson 5, we know that  Zp, where Zp, is the ring of residue classes modulo p.
pZ
Z
Since P is a prime, Zp is a field (vide: Lesson 5). Hence is a field and so, by
pZ
theorem 2, the ideal pZ =P becomes a maximal ideal. ■
Note: The above problem can also be proved from elementary properties,
without using theorem 2 or the basic homomorphism theorem of ring theory. ■
Problem 3
If R is any finite commutative Ring (with identity) then any prime ideal of R is
also maximal.
Solution
R
Let P be any prime ideal of the ring R. Then by theorem 1, is an Integral
P
R
Domain. Since R is a finite ring, is also a finite Integral Domain and hence
P
becomes a field (vide Lemma 3 in Lesson 5). Hence, by applying theorem 2, we get
that, P is even a maximal ideal, R being a commutative ring with identity. ■
Note: One can show, by means of a suitable example that theorem 2 is false
when the ring R is either not commutative or does not possess identity element.
However, one can show that, if R is a ring, not necessarily commutative, nor
R
possessing an identity element and if M is a maximal ideal of R then is a division
M
ring. ■
Problem 4
Let R be the ring of all the real valued continuous functions defined on the
closed unit interval. (See Example V. Lesson 5). Prove that Mr = { f(x)  R : f(x) = 0 }
is a maximal ideal of R, where 0 < r < 1.
[Note: It can be shown that every maximal ideal of R is of this form Mr, so that
the maximal ideals of R are in one-to-one correspondence with the points on the
unit interval]
89
Solution
Mr is an ideal of R, is easy and straight forward to ascertain. (By Example V
Lesson 5) Mr is the kernel of the epimorphism r : RF defined by setting r(f) = f(r),
where f  f(x)  R and F is the field of real numbers and so Mr is ideal of R. To show
that Mr is a maximal ideal of R, it is enough if we show that whenever N is an ideal
of R such the N contains Mr and N  Mr then N = R. Now since N Mr = g(r) 0; say
g(r) =  0. Then h(x) = g(x) - . Then h(x)  R such that h(r) = g(r) -  = 0, and
hence h(x)  Mr  N. Thus both g(x), h(x)  N so that =g(x) – h(x)  N, N being an
ideal of R. Since   0, -1 exists in R and 1 =  -1  N. Hence for any element t(x)
 R, t(x) = 1, t(x)  N, which means that N  R  N; i.e. N = R as required. Hence Mr
is a maximal ideal. ■
The above problems have furnished us with a large number of maximal ideals
in the ring Z of integers and the ring of all real valued continuous functions defined
on I, the closed unit interval.
The field of quotients of an integral domain:
We are familiar with the method by which the set Q of rational numbers is
constructed out of the set Z of integers. Remembering that Q is a field and Z is an
integral domain. We, therefore a construct of a field out of an Integral Domain such
that the constructed field contains Integral Domain as a sub-domain. A similar
procedure is adopted so as to “imbedd a given integral domain in a field”, as per the
following definition.
7.3.3 DEFINITION – IMBEDDED
A ring R is said to be imbedded in a ring R, if there exists an isomorphism of R
onto a sub-ring of R.
(In other words R contains an isomorphic imbedding of R into R. When R and
R possess identities e and e, we require that under this isomorphic imbedding e is
mapped into e.) We now have the following important and fundamental theorem.
Theorem 3.3.1
Every Integral Domain can be Imbedded in a Field.
Proof
Let R be any integral domain. Let M be the set of all ordered pairs (a, b) where
a, b  R and b  0. We now define a relation ~ on M by setting (a, b), ~ (c, d) if and
only if ad = bc.
We show that this relation ~ is an equivalence relation. For this we need to
check only, whether this relation is reflexive, symmetric and transitive; i.e. we have
to check for any (a, b), (c, d), (e, f)  M that (i) (a, b) ~ (a,b), is true?, (ii) if (a,b) ~
(c,d)  (c, d) ~ (a, b) is true ? and (iii) if (a, b) ~ (c, d) and (c, d) ~ (e, f)  (a, b) ~ (e, f)
is true ?. (i) and (ii) follow easily from the definition of this relation ~ and we know
that the integral domain is a commutative ring. To see the truth of (iii) we proceed
as follows: (a, b) ~ (c, d) and (c, d) ~ (e, f) imply ad = bc and cf = de. From ad = bc,
90
we get adf = bcf and from cf = de, we get bcf = bde. So we get adf = bde, and hence
using the commutativity of multiplication and cancellation law valid in the integral
domain, we obtain that af = be which implies (a, b) ~ (e,f) which is required. Thus
the relation ~ turns out to be an equivalence relation, and so defines a partition of
M into equivalence classes. We denote, as customary, any equivalence class
determined by the pair (a, b)  M in the form [a, b].
Thus [a, b] = { (c, d)  M: (a, b) ~ (c, d)}.
Let F be the set of all the equivalence classes [a, b], where (a, b)  M. We
define suitable operations of addition and multiplication on F and make F into the
desired field in which the given integral domain is imbedded. In fact, we define
addition ‘+’ and multiplication ‘.’ on F as below.
For [a, b], [c, d] F,
let [a, b] + [c, d] = [ad + bc, bd], [a,b] [c,d] = [ac, bd] ….(1)
Since both b and d are non-zero elements of R and R is an integral domain, we
get that bd  0 so that the symbols [ad + bc, bd] and [ac, bd] are permissible and
they belong to F. Since the above definitions of addition and multiplication on F
depend on the representatives that we choose in the equivalence classes, we need
to check that the above definitions are well defined. For this we need to examine
whether [a, b] = [a, b] and [c, d] = [c, d] imply [a, b] + [c, d] = [a, b] + [c’, d’] and
[a, b] [c, d] – [a, b] [c, d]. Now, [a, b] = [a, b] and [c, d] = [c,d] implies that
(a, b) ~ (a, b) and (c, d) ~ (c, d);
That is ab = ab and cd = cd .…(2)
To prove that [a, b] + [c, d] = [a, b] + [c, d] we need to show that [ab+ bc, bd]
= [a d + b c b d], i.e., (ad + bc) b d = (ad+bc) bd
That is ad bd + bcbd = adbd + bcbd … (3)
From (2) we get abdd = abdd and bbcd = bbcd, and hence, by adding and
using the commutativity of multiplication we obtain the required result (3).
Similarly, to prove [a, b]. [c, d] = [a, b]. [c, d] we need to show that
[ac, bd] = [ac, bd], i.e. acbd = acbd. This follows from (2) by multiplying the
equation in (2) and using commutativity of multiplication. Thus we have verified
that both the operations of addition and multiplication defined on F by equations
(1) are well defined.
Now we claim that F is a field under these operations of addition and
multiplication and that R can be imbedded in F. In the first place, we can easily
check that F is an abelian group under addition. In fact addition on F can easily
be seen to be commutative. It is also associative, for take [a, b], [c, d], [e, f]  F.
Then an easy computation could be done to get
([a, b] + (c, d]) + [e, f] = [(ad + bc)f+ (bd)e, (bd)f] and
[a, b] + ([c,d] + [e,f]) = [a(df) + b (cf + de), b(df)]
91
Since the right hand sides of the above equations are clearly equal, the
associatively of addition on F follows. Then the equivalence class [0, b]  F acts as
additive identity, as can be easily verified and for any [a, b]  F, (-a, b) acts as
additive inverse, since [a, b] + [-a, b] = [ab + b (-a), b2] = [0, b2] = [0, b] and also
[-a, b] + [a, b] = [0, b], the additive identity in F. Thus < F, + > is an abelian group.
Secondly, we examine the multiplicative structure of F and show that the non-
zero, form a group under multiplication. In fact, the commutativity and
associativity of multiplication and the absence of non-trivial zero divisors. In the
integral domain R force us to conclude that the non-zero elements of F form a
commutative semi-group. Further the element [a, a] in F acts as identity for
multiplication, since it can be verified that [a, a], [x, y] = [x, y] [a, a] = [x, y] where
[x, y]  F. Hence we may note that [a, a] = [c, d] iff c=d (using the cancellation laws
valid in the integral domain R).
Again for any non-zero elements [x, y] in F (which means that both, x, y are
non-zero elements of R), there exists an inverse, viz. (y, x) in F such that [x, y][ y, x]
= the identity [xy, xy]. Thus the non-zero elements of F form a multiplicative
abelian group. Thirdly, we can check that multiplication in F is distributive over
addition in F. Since multiplication in F is commutative, one needs to check one of
the distributive laws and we proceed to check the left distributive law:
[a, b] ([c, d] + [e, f]) = [a, b]. [c, d] + [a, b] . [e, f] … (4)
The left hand side of (4), on computation, reduces to [a (cf+de), b (df)] …(5)
The right hand side of (4) on the other hand, reduces to
(ac)(bf) + (bd)(ac), (bd)(bf)] = [b(acf + ade), b(bdf) …(6)
Using the commutative and associative properties of multiplication in R we
find that (5) and (6) equal. Thus the multiplication in F is distributive over addition
in F and we have therefore verified all the field axioms in the case of F. Hence
< F, +, > becomes a field as earlier claimed.
To conclude the proof of the theorem, it remains to prove that the given
integral domain R can be imbedded in this field F, For this purpose we need to
identify certain elements in F as elements in R. Let R = { [ ab, b]  F : a  R }.
We claim that R is a sub-ring of F which is isomorphic with R. Once we establish
this claim it follows that R is embeddable in F.
We note that [ab, b] = [x, y]  aby = bx  ay = x (ring cancellation law valid in R).
So that all the elements of the form [ab, b] where b is arbitrary and a is fixed,
are all equal and that only these are control. We identify the element [ab, b] of F
with the element of R, the identification being made possible by the isomorphism
: RR defined by (a) = [ab, d] that the above mapping  is one to one and onto.
Further (a+a) = [(a+a)b, b] = [ab, b] + [ab, b] as can be checked, so that
 (a+a) =  (a) +  (a). Again  (a a) = [aa b, b] = [ab, b] [ab. b] as can be verified
whence (a).(a) = (aa). Thus  preserves R, and hence  is an isomorphism of R
onto R1 in F, as required. ■
92
Note: The above proof of the theorem 3 appears to be very lengthy and it is so
because we have included the verification, in the proof of almost all the ring
axioms. In the case of F, though many of the verifications are routine and could
have been omitted. The field F constructed in the course of the proof is called the
field of quotients of R. When R=Z, the Ring of Integers, the field F is the familiar
field Q of rational numbers.
Problem 5
Let R be an Integral Domain and a, b  R. If an = b n and am = b m where m and
n are two relatively prime integers, then prove that a=b.
Solution
In the first place we note that, under the conditions stated in the problem, a=0
if and only if b = 0. This follows from the fact that xm = 0, for x  R implies x = 0
where R is an integral domain. Thus when a = 0, b = 0 then a = b is very obvious.
So let us assume that a  0, in which case b  0. In this case also we should prove
that a = b. Now since the integers m and n are relatively prime, there exist integers
p and q such that pm + qn = 1. As m and n are both positive integers, of the 2
integers p and q one of them is positive and the other is negative. Without loss of
generality, we may assume that p is positive and q is negative, Let q = – q where q
is positive. Now pm + qn = 1 gives that pm = 1+qn. So after some computation ,
we get b = a.
Thus the problem follows.
Note: The need for the introduction of the positive integer q arises in the
above proof, because aqn has no meaning (in general when q is negative.)
7.3.4. EUCLIDEAN RINGS
The type of rings we propose to study now is motivated by the properties of
several rings like the ring Z of integers, the ring Z[i] of Gaussian integers and the
polynomial rings (about which we will study in the next and final lesson on rings).
The Euclidean ring possesses some of the characteristic and important properties of
the above mentioned rings.
Definition 3.4.1
An Euclidean ring is an integral domain R in which for every element a  0 in
R a non-negative d(a) [called a semi-valuation or a d-value of a] is associated,
satisfying the following conditions,
(i). If a, b  R and both are non-zero then d(a) < d (ab),
(ii). If a, b  R and b  0, then there exists elements q, r  R such that b = qa+r
where either r = 0 or d(r) < d(a) (Division algorithm)
Note: For the zero element of R no d-value is assigned, in our definition. An
example of a Euclidean ring is the Ring Z of integers. In fact, it is known
that Z is an Integral Domain and for any mZ, |m|, i.e. the numerical
value of m serves as the d-value of m, as both the conditions (i) and (ii) in
the definitions are clearly satisfied. Later we will show that the ring Z[i] of
93
Gaussian Integers is also a Euclidean Domain. We will now prove that any
Euclidean ring is a Principal Ideal Ring. First we define the notion of a
principal ideal ring.
Definition 3.4.2
A Principal Ideal Ring is an integral domain R with identity in which every
ideal is principal. i.e. if A is any ideal in R then there exists an element aR which
generates A. i.e. every element in A is of the form ra with rR.
Theorem 3.4.3
Any Euclidean Ring is a Principal Ideal Ring.
Proof
Let R be a Euclidean ring. Since R is already an integral domain, in order to
prove that R is a principal ideal ring, it suffice to show that R has an identity
element and any ideal of R is a Principal Ideal. Let A be any ideal of R. If A is the
null ideal there is nothing to prove. So, assume that A is a non-null ideal. Then for
each aA, a0, there is a d-value d(a) which is a non-negative integer. Choose the
element say a whose d-value d(a) is the least among the d-values of the non-zero
elements in A. We claim that every element in A is of the form ra for some rR. In
fact if bA then by division algorithm which exists in Euclidean Ring R, there exists
elements q,rR such that b = qa + r where either r = 0 or d(r) < d (a). We assert
that r = 0 for, if not d(r) < d(a) and r=b-qaA (since A is an ideal and ba  A, rR).
This leads to the contradiction that data is the least among the d-values of the non-
zero elements of A. Hence r = 0 and we get b = qa. Thus every element in A is of
the form qa for some qR. Applying this fact to the unit ideal R, we find that there
exists an element say u in R such that any element of R is a multiple, say qu of u
for some qR. In particular u= eu for some eR. We claim that this element e is
the identity element of R. In fact for any element x in R, we can write x = qu =
q (eu) = (qe)u = (eq) u = e (qu) = ex. Hence, it follows that e is the identity element
of R. Now that we have proved that R has identity and for any ideal A of R there
exists a suitable element a  A such that every element of A is a multiple qa of a.
We can conclude that any ideal A of R is a Principal Ideal, generated by some
element a  A. Hence R becomes a Principal Ideal, generated by some element aA.
Hence R becomes a Principal Ideal Ring. ■
Note: Even though any Euclidean Ring is a principal ideal ring, as we have shown
just now, the converse is false. An example of a principal ideal ring which
is not a Euclidean ring can be given. We can show that the ring Z[x] of all
polynomials in x over the ring Z of integers cannot be a Euclidean ring. In
fact Z[x] contains an ideal viz. the ideal generated by { 2, x }, which is not
principal. This fact can be proved. Thus Z[x] is not a Euclidean ring, even
though it is an integral domain.
Some authors call a Euclidean Ring as Euclidean Domain and a Principal Ideal
Ring as a Principal Ideal Domain, since both these Rings are Integral Domains. In
94
order to obtain some further properties of Euclidean ring we introduce the familiar
notions of divisibility, prime element, greatest common divisor, relatively prime
elements etc. We will assume hereafter for this purpose that any element
considered in the Euclidean ring R is non-zero, unless otherwise stated.
If a, b  R then we say that a is a divisor of b, or b is a multiple of a, whenever
there exists an element in R such that b = ca. We exhibit this fact by writing a|b
One can easily check the following facts (i) For a, b, c  R, if a|b and b|c then a|c,
(ii) if a |b and a|c then a|(b  c), and (iii) if a|b then a|rb for all r  R.
Definition: 3.4.4 If a, b, c  R (where R is a ring) then an element d of R is
said to be a greatest common divisor (abbreviated as g.c.d.) of a and b if
(i) d|a and d|b (ii) whenever for cR, c|a and c|b then c|d.
We denote any g.c.d. of a and b by (a, b).
Lemma 3.4.5
Let R be an Euclidean Ring. Then any two elements a, b of R have a g.c.d. say
d and there exists elements ,   R such that d = a +b.
Proof
Let A = { ra + sb : r, s  R }. We claim that A is an ideal of R. For that, let
x, y  A, so that x = r1 a + s1b, y = r2 a + s2 b for some r1, r2, s1, s2  R. Then
x-y = (r1 – r2) a + (s1-s2) b and for any r R, rx = (rr1)a + (rs1) b so that both x – y and
rx are in A. Thus A turns out to be an ideal and hence, R being a Euclidean ring,
there exists an element dA which generates this ideal A, i.e. A=<d>. But both a
and b are in A, as we can write a = 1a + 0b, and b = 0a + 1b. Hence d|a and d|b.
Further since d  A, there exists ,   R such that d = a + b which shows that if
c|a, c|b for cR then c|d. Hence d is a greatest common divisor of a and b ■
Note: The above proof of the Lemma is valid obviously even in the case of a Principal
Ideal Ring. However in the case of Euclidean Ring there exists an algorithm
or a method for determining the g.c.d. (a,b) of any given pair of elements a, b
of the ring, while in the case of a principal ideal ring such an algorithm may
not exist. The following problem explains the algorithm for determining the
g.c.d. (a, b) of any given pair of element a, b of a Euclidean Ring.
Problem 6
Explain how the g.c.d. (a, b) of a given pair of elements a, b of a Euclidean ring
R can be found.
Solution:
Invoking the “division algorithm” which exists in R, we can determine
elements q0, r1, r2 …qn–1, rn etc. in R such
b = q0 a + r1 when d(r1) < d(a) …(1)
a = q1 r1 + r2 where d(r2) < d(r1) …(2)
r1=q2 r2 + r3 where d(r3) < d(r2) …(3)
95
…………………
…………………
rn-2=qn-1 rn-1 + rn where d(rn) < d (rn-1) …(n)
Thus we have strictly decreasing sequence of non-negative integers viz. d(a),
d(r1), d(r2) … d(rn-1), d(rn). This sequence must terminate at a finite stage, say at the
(n+1)th stage, so that we get the (n+1)th equation rn-1 = qnrn. Here we have taken that
rn+1=0. Now we assert that rn = (a, b). In fact the (n+1)th equation shows that rn|rn-1.
Using this fact that nth equation shows that rn|rn-1. Using this fact the nth equation
shows that rn|rn-2, and hence the (n-1)th equation, viz rn-1 = qn-2 rn-2+ rn-1 shows that
rn|rn-3. Proceeding in the same manner the (n-2)th equation shows that rn|rn-3.
Proceeding in the same manner the (n-3)th equation shows that rn|rn-4 etc. Finally we
get from the 2nd and the 1st equation that rn is a common divisor of both a and b. On
the other hand, if d in R is a common divisor of both a and b then from equation (1)
we find that d is a divisor of a1 so that d is a common divisor of r2. Arguing similarly,
we can show from equation (3) that d is divisor of r4, r5 etc. and finally from equation
(n) we could find that d is a divisor of rn. Thus we have established that, rn is a
common divisor of both a and b and that any common divisor d of both a and b is a
divisor of rn. Hence, by definition rn turns out to be a g.c.d of a and b. ■
Definition 3.4.8
Let R be a commutative ring with identity 1. An element u(0) in R is said to be a
unit (or an inevitable element) in R iff there exists an element v in R such that uv = 1.
Lemma 3.4.8
Let R be an integral domain with identity. Suppose a|b and b|a for any a,bR,
then a = ub, where u is a unit in R.
Proof
Since a|b we can write b = xa, for some x  R. Similarly since b|a, we have
a = yb for some y < r. This b = xa = x(yb) = (xy)b which shows that xy = 1, on
cancelling b which is permissible, R being an integral domain.
Hence y turns out to be a unit in R and a = yb ■
Such pairs of elements like a,b as above are called associates as per the
following definition.
Definition 3.4.9
Let R be a commutative ring with identity. Two elements a and b of R are said to
be associates if b = ua for some unit in R. Clearly, if a and b are associates in a ring,
then each is a divisor of the other, and they generate the same ideal, provided the ring
is a commutative ring with identity. Also, it can be shown that in a Euclidean Ring any
two given greatest common divisors of two elements are associates. ■
Lemma 3.4.10
If R is an Euclidean Ring and a, bR, then d(a) <d (ab), provided b is not a unit
in R.
96
Proof
We have seen earlier in the proof of Theorem 4 ( viz. every ideal in Euclidean
ring is a Principal Ideal Ring ) that for any non-null ideal A in R any element in A
whose d-value is the least among the d-values of the non-zero elements of A is a
generator of A. Suppose now A = <a> is the ideal generated by a, then ba  A and
by one of the properties of the d-value, we have d(a) < d (ba). We claim that
d(a)  d(ba). Suppose not, then d(ba) = d(a) implies that ba is also a generator of A.
This means a = x (ba) for some xR, and hence by canceling a (which is
permissible as R is an integral domain) we get 1=xb. This shows that b is a unit
which is a contradiction to the data b is not a unit in R. Thus d(a) < d(ba), as
required. ■
Definition 3.4.11
In a Euclidean ring R an element p is said to be a prime if
(i) p is not a unit and
(ii) whenever p=ab for a, b  R then either a or b is a unit. ■
The above definition shows that an element pR is a prime iff its only divisors
are units and associates of p. In other words any prime element of R admits only
trivial factorizations. We now prove an important property concerning the
factorization of any elements in a Euclidean ring and ultimately establish the
unique factorization property.
Lemma 3.4.12
Let R be a Euclidean Ring. Then every element of R is either a unit or can be
written as a product of a finite number of prime elements of R.
Proof
We prove the result by induction on d(a), where aR. Suppose d(a) = d(1),
then a must be a unit. In fact, for any non-zero bR, one of the properties of the
d-value shows that (since b = 1b), d(1) < d(b). Now by division algorithm in R there
exist elements q, rR such that 1 = qa + r where either r=0 or d(r) < d(a). Since,
d(1) is the least among the d-values, we conclude that r = 0 so that 1 = qa which
shows that a is a unit and the lemma follows. We assume therefore that the lemma
is true for all x  R for which d(x) < d(a) and prove that the lemma is true for a.
This would complete the induction and the lemma would have been proved. If a is
a prime element in R there is nothing to prove. So we may assume that a is not a
prime and that we can write a = bc where b,c R and both b and c are not units.
Now, by the last lemma, we get d(b) < d(bc) = d(a) and d(c) < d(bc) = d(a). Thus, by
our induction hypothesis, both b and c can be written, as a product of a finite
number of prime elements, say b = p 1, p 2, …pm and c = p1, p2 … pn where the p’s
and (p)’s are all prime elements. Hence we get a = be = p1 p2 – pm p1 p2 …..pn.
Thus a have been written as a product of a finite number of prime elements and
the lemma follows by induction. ■
97
Definition 3.4.13
Two elements a, b in a Euclidean ring are said to be relatively prime (or prime
to each other or primes), whenever the g c.d. of a and b is a unit. ■
Since any associate of g.c.d. is easily seen to be also a g.c.d. and since 1 and
any unit are associates, we could find that we can take (a,b) = 1 whenever a and b
are relatively prime. Further we will find that the prime elements in a Euclidean
ring play the same role that prime numbers play in number theory. In fact we
obtain familiar number theoretical results in the language of Euclidean rings and
we obtain many of them below.
Lemma 3.4.14
Suppose R is a Euclidean ring and a, b, c R and a| bc and (a,b) = 1 then a|c.
Proof
Since (a, b) = 1 we find, by Lemma 1 that there exists elements , R such
that a+b = 1. So we can write c = 1c = (a + b) c = ac + bc. But by data a|bc.
Therefore we can assume that bc = ad for some d  R. Hence from c = a (a+b) we
see that a|c as required. ■
Just as in the case of integers, we can show that if p is any prime element in a
Euclidean Ring R and a R then either pa or (p,a) = 1 (or a unit). In fact by the
definition of the g.c.d. (p,a) is a divisor of p and p being a prime this means that
(p, a) is neither p nor 1. So if (p, a)  1. then (p, a) = p which means that p|a.
Hence the assertion. ■
We use this fact in providing the following important property of any prime
element in R.
Lemma 3.4.15
If p is any prime element in a Euclidean ring R and p|ab where a, b  R then
p must be a divisor of atleast one of a or b.
Proof
Suppose p is not a divisor of a. We must show that p|b. Since p is prime
element and p does not divide a, we can say that (p, a) = 1. Hence by Lemma 5 we
can conclude that p|b, by data p|ab. ■
Corollary: 3.4.15
If a prime element p of a Euclidean ring is a divisor of a product a1a2…an then
p must be a divisor of atleast one of the elements a1, a2, … a n. ■
Unique Factorization Theorem:
Theorem 3.4.16
Let R be an Euclidean ring and a  0 be a non-unit in R. If a = p1p2 … p n
= p1p2…pm where the pi and pi are prime elements of R. Then n = m and each pi
for 1 < i < n is an associate of some pj for 1 < j < m. Conversely each pj is an
associate of some pi. (In short, if any element a of an Euclidean ring R has two
98
possible factorizations into prime factors then both are same but for order and unit
factors).
Proof
The given relation a = p 1 p 2… pn = p1p2pm shows that p must divide
p1p2 …. pm and hence by the above corollary p1 must divide some p1. Since both
p1 and pi are prime elements in R and p1 |pi, p1 and pi must be associate so that
we can take pi = ui p1 where ui is a unit in R. Thus p1p2 … pn = p1 p2 … pi …pm =
ui p 1p2 … pi-1 pi+1, … pm. After canceling of p1 we will be left with p2 … pn =
u1 p2 … pi-1 pi+1 … pm. Repeat the same argument with p2 in place of p1 and after
successively cancelling p 2, p3 etc. at the nth stage we could obtain 1 equal to the
product of certain units and a certain number viz. m-n of p’s provided n<m. But
since the p are not units, the above equality will force cancellation of all the p’s and
this means that m = n. Further, we have incidentally proved that every pi has some
pj as an associate and vice versa.
Hence the theorem follows ■
Applying the earlier proved lemma 4, the above theorem shows that every non-
zero element in a Euclidean ring R is either a unit or can be uniquely factorized as
a product of prime elements of R. Factorization being unique but for order and unit
factors. Because of this property any Euclidean ring turns out to be a unique
factorization domain (U.F.D.). The notion of a U.F.D. will be considered in the next
lesson.
We will close this lesson with two important results, one concerning the
maximal ideals of a Euclidean ring and the other providing an important example,
viz., Z[i] of a Euclidean ring.
Theorem 3.4.17
Prove that an ideal M of an Euclidean ring R is a maximal ideal iff the ideal M
is the principal ideal generated by a prime element of R.
Proof:
Since any ideal in a Euclidean Ring is a Principal Ideal, we can take ideal
M = < m >, the ideal generated by an element m  R. Intially we are going to prove
that when m is not a prime element then the ideal M is not maximal. Since m is
not a prime element we can write m = ab where a, b  R and both a and b are non-
units. We claim that, if A = <a>, the ideal generated by a, then M is properly
contained in the proper ideal A of R and hence it will follow that M is not a maximal
ideal. In fact, if xM then x = rm for some rR so that x = r(ab) = (rb)aA. Thus M
 A. However M  A. For otherwise, if aM and we should have a = m for some
R. This means that a = (a b) and hence b = 1 which is a contradiction, as b is
not a unit in R. Thus M is properly contained in the ideal A. Moreover this ideal A
is not the unit ideal in R. For other wise 1A and this implies 1 = a for some R.
But this again is impossible since a is not a unit. Thus M  A  R and the ideal A is
neither M nor R. Hence M cannot be a maximal ideal.
99
Conversely assume that mM is a prime element in R. We will prove that the
ideal M is maximal. In fact suppose N = <n>, the ideal generated by n  R, is an
ideal such that M  N  R, we will prove that either N = M or N = R and hence it
will follow that M is maximal. Since M  N, m  N = <n> so that we can write
m = xn for some xR. But m is a prime element and so either x is a unit in which
case n is an associate of m or n is a unit. When x is a unit, m and n are associates
and it can be easily shown that <m> = <n>, i.e. M = N. On the other hand when n
is a unit, we can show that N = R. In fact, n being a unit there exists pR such
that 1 = np. Since N = <n> and 1 = np, we get 1  N, N being the ideal generated by
n. Hence for r  R, r = 1r  N. Thus R  N  R. i.e. N = R as claimed. Thus we have
shown that if N is an ideal of R such that M  N  R the either N = M or M = R. This
means that M is a maximal ideal of R and the proof of the theorem is concluded. ■
Corollary 3.4.18
In an Euclidean ring R any prime ideal  R is also a maximal ideal.
Proof
Let P  R be a prime ideal and let P be generated by the element pR. We
claim that p must be a prime element. Suppose p is not a prime then, we can write
p = ab where a,b  R and both a and b are not units. ( Here we may note that p
cannot be a unit for then p = R, which is not the case ). Now a does not belong to P
for otherwise, if a P then a = pc for some c  R and we obtain p = ab pcb. This
means that cb = 1 which is not possible, since by our assumption b is not a unit in
R. Thus aP. Similarly we can show that b P. However ab = bP. Hence P
cannot be a prime ideal contradicting our assumption that P is a prime element.
Hence, by the last Theorem, P is also maximal and the corollary follows ■
Note: Since in any commutative ring with identity, any maximal ideal is also a
prime ideal. We conclude from the above corollary that in any Euclidean
ring R there is no difference between maximal and prime ideals.
Theorem 3.4.19
The ring Z[i] of Gaussian integers is an Euclidean ring.
Proof
We recall Z[i] = { + : ,   Z, i =  1 } is a commutative ring with identity,
under the usual operations of addition and multiplication of complex numbers. Since
Z[i] is clearly a sub-ring of the field of complex numbers, Z[i] is an integral domain.
Therefore, in order to show that Z[i] is a Euclidean ring we need to define a suitable
d-value d(a) for each non-zero element a in Z[i] possessing the required two properties,
viz. (i) d(ab) > d(a) & d(b) and (ii) there exist elements q, r  Z[i] ( given a, b  Z[i] with a
0 ) such that b = qa + r where either r = 0 or d(r) < d(a). We will now show that |a| =
(aa ) for a  Z[i] serves as a suitable d-value of a. Here a denotes, as usual, the
complex conjugate of a and |a| is the modulus of a. In fact, if we take d(a) = |a|
where, a Z[i]. Clearly d(a) is a non-negative integer such that for a, b  Z[i]
100
d(ab) = |a b |  |a| |b| i.e. (ab)  d(a)d(b). Thus the required property (i) of the
d-value in a Euclidean ring is satisfied. It remains for us to verify that, property (ii)
also holds good. Now Z[i] can be imbedded in the field Q(i) of all Gaussian numbers
of the form x+yi where x, y  Q. Given a, b  Z[i] with a  0, we need to
determine, q, r  Z[i] such that b = qa + r, where either r = 0 or d(b)< d(a).
b
Since Q(i) is the field of quotients for the integral domain Z[i], Q(i) and we can
a
b
take =x+yi where x, y  Q.
a
Let q1, q2 be the integers nearest to x and y respectively such that we can write
x = q1 + 1, y = q2 +  2 where 1, 2 Q and |1|  ½, |2| ½. Let q=q1+iq2  Z[i].
Then b=qa = r  Z[i], where r = (1+i2), a as can be easily checked. Now
(d (r))2 = |r|2 – | 1+i2|2|a|2=(12)+ ( 22)|a|2  (¼+¼) (d(r))2. Since |1| and |2|
are  ½. Hence d(r) < d(a). Thus we have determined elements q, p  Z[i] such that
b = qa + r where either r = 0 (which will be the case when  1 = 2 = 0 or d(r) < d(a).
This completes the proof of the theorem. ■
We will now illustrate the above procedure of finding the required q, r  Z[i], in
b 4  7i (4  7) (3 - 4i)
a particular case, say when a = 3 + 4i and b = 4 + 7i, then  
a 3  4i 25
40  5i 8 1 8 1
=   , so that in the notation of the earlier work, we have x = , y  ,
25 5 5 5 5
q = q1 +iq2 = 2. Then b-qa = (4+7i) – 2(3+4i) =-2-1=r and (d(r)) = (–2) + (–1) = 5
2 2 2

which is certainly < (d(a))2, since (d(a))2 = 32 + 42 = 25.


Thus in the next lesson we will use this fact that the ring Z[i] of Gaussian
integers is a Euclidean ring and establish a very important number theoretical
result, namely that if p is a prime number of the form 4n + 1 then there exist
integers a and b such that p = a2=b2.
SOLVED PROBLEMS
1) . Prove that if an ideal A of a ring R contains a unit then prove that the
ideal A must be the unit ideal R.
Solution:
Since A is an ideal of R, we say that, A  R
We know that 1  A, x  R  1 x  A (Since A is an ideal)
xA
Therefore R  A. Hence A = R. ■
2) . Let R be an Euclidean Ring and aR. Prove that a is a unit in R iff
d(a) = d(1), where 1 is the identity in R.
Solution:
Let a be a unit in R.
To prove: d(a) = d(1)
By the definition of Euclidean ring, d(1a)  d (1)
 d (a)  d (1)
101
Since a is a unit in R, a-1 exists & aa-1 = 1
Hence d(1) = d(aa-1)
 d (a)
Therefore d(a) = d(1)
For the Converse part let, d(a) = d(1).
Now we have to prove that a is a unit in R.
By division algorithm valid in A, we can find q, r  A such that
1= qa + r, where either r = 0 or d(r) < d(a)
Now, since d(a) = d(1) and d(r) = d (r.1)  d(1), we have d(r)  d(a).
r = 0 so that 1 = qa, and hence a is a unit in R. ■
3) . If R is a non-commutative ring with identity and if M is a maximal ideal
R
in R, prove that the residue class ring is a division ring.
M
Solution
R
Since R is a commutative ring with identity is also commutative
M
ring with identity.
R
The zero element of is M.
M
R
To prove that is a division ring.
M
R
Let M + r be any non-zero element of .
M
Then M + r  M i.e. r  M.
To prove that M + r is invertible
Let (r) be the principal ideal of R generated by r. Then 1+(r) is
also an ideal of R. Since r  M, M is properly contained in M + (r).
But M is a maximal ideal of R. Hence M + (r) = R, and since 1 R,
1 = m+r for mM  1 - r = m  M  M + 1 = M + r = (M+) (M+r)
M +  = (M+r)-1. Therefore (M + r) is invertible.
R
Hence is a division ring. ■
M
4) . Prove that in an Integral Domain R with identity and two elements a, b of
R generate some ideal iff a and b are associates.
Solution:
Given that (a) = (b), where (a) = the principal ideal generated by
a={ ax/xR } & ( b ) = the principal ideal of R generated by b.
To prove that a and b are associates
Now (a)  (b)
 a  (b)
 a = rb for some r  R
 b | a.
102
Similarly (b)  (a)
 b  (a)
 b = sa for some s  R
 a|b
Now a|b   c  R such that b = ac
b|a   d  R such that a = bd
 b = ac = (bd) c = b (dc)
 b.1 = b (dc)
 b (1-dc) = 0
 1-dc = 0 (since b  0 & R is without zero divisor)
 1 = dc.
Both c & d are units in R,
Thus a = bd where d is a unit.
Hence a and b are associates.
Conversely assume that a and b are associates
To prove that (a)=(b)
Since a and b are associates, a = bu where u is a unit in R.
Now a = bu  b|a
Again a = bu  b = au-1
a|b
Now a|b  b = pa for some p  R.
Let y  (b)  y = qb for some q  R
= q (pa)= (qp) a (a) (since q,p  R)
(b)  (a)
Similarly b|a  (a)  (b)
Hence (a) = (b). ■
7.4. REVISION POINTS
Ideal, Maximal ideal, Imbedding, Euclidean ring, Principal Ideal Ring, greatest
Common divisor, Commutative Ring with identity, Prime and Relatively prime.
7.5. INTEXT QUESTION
1. Give an example of commutative ring with identity in which not all prime
ideals are maximal
2. Prove that in the Ring of Integers all prime ideals are maximal
3. Let R be an Integral Domain and a,b  R. Suppose an = b n and am = bm
where m and n are two relatively prime integers. Then prove that a = b
4. Prove that in an Integral Domain R with two identity elements a, b of R
generate some ideal iff a and b, are associates.
7.6. SUMMARY
Ideal, Maximal ideal, Imbedding, Euclideen ring, principal ideal ring, greatest
common divisor, commulative ring with identify, prime and relatively prime are
explained.
103
7.7. TERMINAL EXERCISES
1. If R is a commutative ring and a  R, then show that aR = {ar : r R} is a
two-sided ideal.
2. If P is a prime number then prove that Jp, the ring of integers mod p, is a
field.
7.8. SUPPLEMENTARY MATERIALS
1. Algebra by Micheal Artin, Prentice Hall of India, New Delhi, 1994.
2. Lectures in Abstract Algebra Volumes I, II and III by N. Jacobson, D. Van
Nostrand Co., New York, 1976.
7.9. ASSIGNMENTS
1. If R be a commutative ring, then prove that an ideal P of R is a prime ideal
R
iff the Residue class Ring is an Integral Domain.
P
2. If R be a commutative ring with identity, then an ideal M in R is maximal
R
iff the residue class ring is a field.
M
3. Prove that every integral domain can be Imbedded in a Field.
4. Let R be an Euclidean ring. Then prove that any two elements a, b of R
have a gcd say d and there exists elements ,   R such that d = a +b.
5. If R is a Euclidean ring and a, bR, then prove that d(a) < d(ab), provided
b is not a unit in R.
6. If R be a Euclidean ring and a  0 be a non-unit in R. Suppose
a = p1 p2 … pn = p1 p2…pm where the pi and pi are prime elements of R.
Then n = m and each pi for 1 < i < n is an associate of some pj for
1 < j < m and conversely each pj is an associate of some p i. (In short, if
any element a of a Euclidean ring R has two possible factorizations into
prime factors then both are same but for order and unit factors).
7. An ideal M of an Euclidean ring R is a maximal ideal iff the ideal M is the
principal ideal generated by a prime element of R.
7.10. SUGGESTED READINGS / REFERENCE BOOK / SET BOOK
Algebra – By Michacl Artin – PHI Learning P Ltd
7.11. LEARNING ACTIVITIES
Students are requested to attend the PCP classes and work out the problems
given in the lesson.
7.12. KEYWORDS
Ideal, Maximal ideal, imbedding, Educlidean ring, Principal ideal ring, greatest
common divisor, Commutative ring with identity, prime, relative prime, integral
domain field.
104
LESSON – 8

FERMAT’S THEOREM
8.1. INTRODUCTION
In the last lesson, we introduced the concept of a Euclidean ring and
established some of its properties, many of which resemble the familiar properties
of integers. At the end of the lesson we have shown that the ring Z[i] of all
Gaussian integers a+ bi, where a, b  Z is a Euclidean domain, by prescribing a
suitable d-value. Now we apply the properties of a Euclidean ring to this ring Z[i]
and derive Fermat’s theorem in number theory , viz. that any prime number of the
form 4n + 1 can be exhibited as a sm of two squares. Before proving this theorem
we require lemmas.
8.2. OBJECTIVES
Polynomial rings, properties of polynomial rings, polynomials over the field of
rational numbers, primitive polynomial, Gauss Lemma , Unique factorization
domain and the Eisenstein’s Criterion are explained.
8.3. CONTENTS
8.3.1 Polynomial Rings
8.3.2 Properties of R(x)
8.3.3 Polynomials over the field of rational numbers
Lemma 3.1
Let p be a prime number. Suppose c is an integer relatively prime to p and
suppose there exist integers x and y such that cp = x2 + y2. Then p can be written
as the sum of two square. That is p = a2 + b 2 for some pair of integers a and b.
Proof
Since an (ordinary) integer can be considered as Gaussian integer with
magnify part absent, we find that the ring Z of integers is a sub-ring of the ring Z[i]
of Gaussian integers. It may happen that an integer p which is a prime element in
Z(i.e. prime number) need not be prime in Z[i] . For instance, the prime numbers 2
is not prime in Z[i] for we can write 2 = (1+i) (1-i) and both (1+i) and (1-i) are neither
units, nor associates of 2 in Z[i], since the only units in Z[i] are 1-1, i, -i as can be
easily shown. Now, we claim that the prime number p considered in the Lemma is
not a prime in Z[i] . Suppose not, then cp = x2 + y2 = (x+yi) (x-yi), this factorization
being possible in Z[i] , by Lemma 6 of the last lesson that since p is a prime divisor
of the L.H.S. and hence of the R.H.S., p must divide at least one of x + yi or x – yi.
If p|(x+yi), then (x+yi) = p(u+vi) which clearly implies (by taking complex conjugates
of both sides or otherwise) that (x-yi) = p(u-vi). Here (x+yi)(x-yi) = p 2 (u+vi)(u-vi).
Thus p 2|(x2+y2) = cp and it follows p|c. However this is not possible as, by
assumption, p and c are integers which are relatively prime. Hence p is not be a
divisor of x+yi. Similarly p cannot be a divisor of x-yi. Thus p cannot be a prime
element in Z[i] as claimed. We can therefore take p=(a+bi) (m+ni) where a+bi,
m+ni,  Z[i] and both a+bi, m+ni are not units. One can easily show that the
105
Gaussian integer  + i is a unit Z[i] iff 2 +  2 = 1. We find that a2+b2 and m2 + n2
are > 1, as both a + bi and m + ni are not units in Z[i] . Now from p = (a+bi) (m+ni),
by taking complex conjugates of both sides or otherwise, we get p = (a-bi)(m-ni).
Hence p2 = (a+bi)(m+ni)(a-bi)(m-ni) = (a2 + b2)(m2 + n2). Therefore (a2+b2) is a divisor
of p2. But the only possible (positive) integer divisors of p 2 are 1, p and p2, p being
a prime. Hence a2 + b2 = 1 or p or p2. But a2 + b2 > 1. So a2 + b2 > 1. So a2 + b 2 1.
Further a2+b 2 = 1 or p or p 2. But a2+b2 > 1. So a 2+b 2  1. Further a2+b 2 cannot be
p2 then m2+n2=1 which cannot be, since m2+n2 > 1. Thus the only possibility is
that a2+b 2 = p = m2+n2. Hence the lemma. ■
Now, if an odd prime number is divided by 4 the remainder can be either 1 or
3 only, as it cannot be clearly 0 or 2. Thus the set of odd numbers can be divided
into two disjoint classes viz. (i) those odd primes of the form 4n+1, which on
division by 4 leave remainder 1 and (ii) those odd primes of the form 4n+3, which
on division by 4 leave remainder 3. We prove in the main theorem that any odd
prime number of the form 4n + 1 can be exhibited as a sum of two squares and we
leave it as an exercise to show that no odd prime number we require the following
lemma which concerns number theory and in the proof of this lemma we make use
of the well-known Wilson’s theorem on number theory.
Lemma 3.2
If p is any prime number of the form 4n + 1 then, there exists an integer x
such that x2 +1  0 (mod p).
Proof
p 1 p -1
Take x = 1, 2, 3, … . Hence is an even number since p = 4n + 1, for
2 2
 p  1
some integer n. Therefore, we can also write x = (-1) (-2) (-3) …    . Now p – k
 2 
 – k (mod p), so that we get
 p  1  p -1
x2 = (-1) (-2) (-3) …    . 1.2.3 
 2   2 

 p  1  p -1
 (p-1) (p-2) (p-3) …  p    1,2,3 ...  (mod p)
 2   2 

 p  1  p -1
 (p-1) (p-2) (p-3) …    1.2.3 ...  (mod p)
 2   2 

 p  1 p  1 p  3
 1.2.3. …  . ...(p  2) (p - 1) (mod p) (by rearrangement)
 2  2 2
p 1
 (mod p)
2
 -1 (mod p). by Wilson’s theorem
106
|p-1  -1 (mod p). p being a prime.
Thus the integer x, has the property x2 + 1  0 (mod p) as required and the
lemma has been established. ■
As an illustration for the above lemma, take p = 4.4 + 1 = 17, a prime of the
form 4+1. Then, one can check by actual verification that x = 1.2.3.4.5.6.7.8.
 p 1 
 here  8  , viz, x = 40320  13 (mod 17) satisfies the congruence relation
 2 
x2  132  - 1 (mod 17)
Theorem 3.1
If p is any prime number of the frm 4n + 1, then d can be expressed as a sum
of two squares; i.e. p = a2 + b2 for some integers a, b.
Proof:
By the above lemma we can find an integer x such that x2  -1 (mod p). We
p
claim that, we can choose this integers x such that |x| < and also x2  - 1
2
(mod p). In fact, if the integer x we had initially taken, is > p, we can divide x by p
and obtain the remainder r. Then r < p and x  r (mod p), so that x2= r2. Let
p p p
r < . Suppose r > then (p-r) < and since p-r  -r (mod p), we get that (p-r)2 
2 2 2
(-r) (mod p). This integer (p-r) will then serve our purpose. Thus we can determine
2

p
an integer x such that |x| < and x2 + 1 is divisible by p i.e., x2 + 1 = cp for some
2
p
integer c. But x2 + 1 < ( )2 + 1, by our choice of x and hence we find that cp = x2 +
2
1 < p2. Thus c < p and p being a prime c and p are relatively prime. Therefore we
can now apply Lemma 1 and conclude that p = a 2+b2 for some integers a, b. The
theorem is therefore proved. ■
8.3.1 POLYNOMIAL RINGS
We have seen already in the last lesson that how the ring Z of the integers
and the ring Z[i] of Gaussian integers can be recognized as Euclidean rings. We are
now going to obtain yet another important example of a Euclidean ring. By Z[x] we
denote the set of all polynomials in x (an indeterminate) over Z. i.e., with
coefficients from Z and it is not difficult to show that Z[x] is a commutative ring
with identity (even an integral domain) under the usual (high school) operations of
addition and multiplication of polynomials. However this ring Z[x] is not a
Euclidean ring and this fact will be proved later. If instead of Z, we take either Q,
the field of rational on R, the field of reals or C, the field of complex numbers and
consider the set Q(x) or R(x) or C(x) of all polynomials in x over Q or R or C as the
case may be and we obtain rings, under the usual operations of addition and
multiplication of polynomials and these rings Q(x), R(x), C(x) are all Euclidean rings,
a suitable d-value being by the degree of the polynomial. More generally we can
107
construct the ring F[x] of all (so called) polynomials in x over a field F and prove
that F[x] is an Enclidean ring. So also we can construct the ring R[x] of all
polynomials over any ring R and study its properties. We propose to study about
these things in this lesson. These ideas will be necessary for the purpose of Galois
theory which will be treated in the last Lesson in Algebra.
Definition 3.1.1
Let R be any ring and let x denote any indeterminate (or variable). Let R[x]
denote the set of all symbols or sums of the form a0+a1+a2 x2+…+anxn where ai  R
for i = 1, 2,… ,n, n an arbitrary non-negative integer. The operations of addition ‘+’
and multiplication ‘.’ are defined on R[x] as follows: If f(x), g(x)  R[x] where
f(x) = a0+a1x+…+an, xn, g(x)= b0 + b1 x + …+ bm xm, then f(x) + g(x) = h(x) = c0 + c1x +
c2x2 + … + cpxp, with ci = ai + bi for all i and p is the maximum of the integers n and
m ( taking a i = 0 if t > n, bi = 0 if t>m) and f(x), g(x) = k(x) d0 + d1x + d2x2 + …+…+
dm+n xm+n where di = a0 bi + a1 b i-1 + … + at b i-t +…+ ajb o for all i (with the same
convention as earlier, viz. ai = 0 if t > n, b = 0 it t > m). Then it is possible to show
that R[x] is ring under the above defined operations of addition ‘+’ and
multiplication ‘.’. This ring R[x] is called the polynomial ring or more precisely, the
ring of polynomials in x over R. The elements of R[x] are called polynomials in x
over R. ■
Remark
In the above definition, it is understood that for the polynomials f(x) and g(x),
the coefficient an and bm are not zero in R. So that the definition of f(x) + g(x) can be
precisely given. We may also point out here that the expressions for elements in
R[x] , like a0 + a1 xn are only formal sums (or symbols) since no meanings as such
can be attached to the terms ai xi as also to the sum of such terms. To avoid this
difficulty, we may denote any element in R[x] as an infinite-tuple like
(a0, a1… …am …) where all ai  R and all but a finitely many ai are zero. With this
form for any element in R[x], addition and multiplication on R[x] are then defined
as follows:
(a0, a1, … au …) + (b0, b1, … bm …) = (c0, c1, … cp …) where ci = ai + bi for all i;
(a0, a1, … an …).(b 0, b1, … b m …) = (d0, d1, … dr …)
i
where d i =  a i b i - j for all i.
j 0
With these definition of addition and multiplication, one can show that R[x] is
a ring. Further, if we denote by x the infinite–tuple (0, 1, 0, 0, …) on the
assumption that the ring R has identity 1, we can easily check that for any positive
integer i, xi = (0, 0,…0, 1, 0, …) form a sub-ring of R[x] isomorphic with R. So if we
identify such an element (a0, 0, 0, …) with the element a0 we can write an
arbitrary element (a0, a1, … an, 0, 0, …) of R[x] in the form of polynomial
a0+a1x+a2x2 + … + anxn. However, we avoided this method of introducing the notion
of R[x] in order to present the ideas in a less abstract manner.
108
We will now derive certain familiar and important properties of the ring R[x]. We will
omit the proofs of some of the properties, as these properties can be easily established
from the definition of R[x] from the immediate consequences of the definitions.
8.3.2 PROPERTIES OF R(X)
1. The ring R can be embedded in R[x] for, all the so called constant
polynomials, say a0 x0 – a0 together with the zero polynomial 0 from a sub-
ring of R[x] which is obviously isomorphic with R.
2. R[x] is a commutative ring MR is and this fact is easily proved.
3. If R has identity 1, R[x] has also identity, viz. 1 x0  1 itself.
4. If R is an integral domain so is R[x] and conversely.
Proof of (4)
For the purpose of the proof of this property, we define the notion of degree of
any polynomial in R[x]. If f(x) is a non-zero polynomial and f(x)  a0 + a1 x + a2x 2 +
… + anxn, we say that f(x) is degree n, provided a n  0. For the zero polynomial we
do not assign any degree. We now make the claim that the degree of the product of
polynomials f(x) and g(x) which are of degrees n and m respectively cannot exceed
n+m and that it is precisely n + m when R is an integral domain. In fact, let f(x) be
the same as the polynomial earlier mentioned and g(x) be the polynomial b0 + b 1 x +
b2x2 + … + bmxm, where bm  0, the degree of g(x) being m. By the definition of
multiplication in R[x], we find that the coefficient of xn in the product f(x).g(x) is
t
 a i , b t -i . Now a i = 0 whenever i > n and b i = 0 whenever j > m. Let t>m + n.
i 0
If 0  i  n, then t – i > m so that bt-i = 0 hence, a1b t-i = 0. On the other hand, if
t  i > n then ai = 0, so that again aibt-i = 0. Thus when t>m+n all the terms in
t
 a i , b t - i are zero and hence the coefficient the degree of f(x)g(x) is atmost m+n
j 0
and the highest degree co-efficient in this product is obviously anbm. Even this
coefficient may be absent, if the ring R has non trivial divisors of zero and anbm are
chosen in R such that an  0, bm  0. But nbm = 0. In such cases the degree of the
product of two polynomials will actually be less than that the sum of the degree of
the two polynomials. However, when R is an integral domain the degree of f(x).g(x)
is precisely m+n, is an bm  0. Since both an and bm are  0. incidentally this shows
that the product of any two non-zero polynomials in R[x] can never the zero
polynomial. Hence R[x] is an integral domain. ■
Further it is evident that when R[x] is an integral domain, so is R. Thus we
have shown that the ring R[x] is an integral domain iff R is.
Now we have the following theorem, which incidentally furnishes us with many
examples of a Euclidean ring.
109
Theorem 3.2.1
For any field F, the ring F[x] of all polynomials in x over F is an Euclidean ring.
Proof
Since F is a field, it is also an Integral domain and hence, as we have seen earlier,
the ring F[x] is also an integral domain. So in order to show that F[x] is a Euclidean
ring it is enough if we are able to assign a d-value to each non-zero polynomial f(x) 
F[x] such that the d-value (which must be a non-negative integer) has the required two
properties mentioned in the definition of the Euclidean ring, namely (i) If f(x), g(x) be
any two non-zero polynomials in F[x] , then d(f(x). g(x))  both d (f(x)) and d(g(x)), and (ii)
given any pair of polynomials a(x), b(x) in F[x], where a(x) is non-zero, then there must
exist polynomials q(x), r(x) in F[x] such that we can write b(x) = q(x) a(x) + r(x) where
either r(x) is the zero polynomial or d(r(x)) < d(a(x)). For this purpose we take the degree
of any non-zero polynomial f(x) as d(f(x)). Then it is clear that d(f(x)) is a non-negative
integer and since d(f(x).g(x)) = d(f(x)) + d(g(x))  both d(f(x)) and d(g(x)) the property (i)
required of the d-value is satisfied. It remains for us to show that property (ii) viz.
division algorithm is also by this d-value. We establish this property (ii) by induction bn
d(b(x)). For the sake of fixing out ideals, let us assume that a(x) = a0 + a1x + … + anxn
with an  0 and b(x) = b0 + b1x + … + bmxm, with bm  0. Then d(a(x)) = n and d(b(x)) =
m. When m < n, the property is trivially satisfied, since we can obviously write b(x) =
q(x) + r(x) where q(x) is the zero polynomial and r(x) = b(x) so that d(r(x)) = d(b(x)) < d
(a(x)), as required. Therefore we have now a basis for the induction proof. So we may
assume as induction hypothesis that the property is true whenever d(b(x)) < m and
then prove the truth of the property when d(b(x)) = m. Now let m  n, where n is the
degree of a(x). Since an  0 in F, there exists the inverse an-1 in F such that an-1 an = 1,
the identity in the field F. A straight forward simplification will show that the
polynomial b(x) = bm.an–1 xm-n a(x) is of degree  m – 1. By using our induction
hypothesis, we can write b(x) = bman-1 xm-n. a(x) = q1(x) a(x) + r(x) where q1(x), r(x)  F(x)
and either r(x) is the zero polynomial or d(r(x)) < d (a(x)). This means that we can write
b(x) = q(x) + r(x) where q(x) = bman-1xm-n+q1(x) F[x] and r(x) is as earlier mentioned.
Thus the property of division algorithm has been proved when d(b(x)) = m. This
completes the induction and the theorem follows■
Now that we have proved that the ring F[x] is a Euclidean ring whenever F is a
field, F[x] has all the properties ( we obtained in last lesson ) of a Euclidean ring.
We state most of those properties of F[x] as below.
(i) F[x] is a principal ideal ring: That is, if A is any ideal in the ring F[x] then
there exists a polynomial p(x) F[x] such that every element in A is a multiple of
p(x) i.e. in the from (x) p(x) F[x].
(ii) Any two non-zero-polynomials f(x) and g(x) in F[x] have a greatest common
divisor d(x) F[x] and there exist polynomials (x), (x) in F[x] such that we can
write d(x) = (x) f(x) + (x) g(x). In fact this g.c.d. d(x) is the generator of the ideal
generated by f(x) and g(x).
110
(iii) It is easy to show that the (non-zero) constant polynomials (i.e. polynomials
of degree 0) in F[x] are the only units in F[x], so that we can define a prime element
in F[x] – the prime element being now called an irreducible polynomial – as the
polynomial p(x) with the property that whenever p(x) = a(x) b(x) then one of a(x) or
b(x) is a constant polynomial, the other being called an associate of p(x). For
instance in the case of the ring Q[x], x2-2 is an example of an irreducible
polynomial, so also in the case of the ring R[x], x2+1 is an irreducible polynomial.
However, one may note that the polynomial x2-2 which is irreducible over Q is
not irreducible over R. For that we can factories x2-2 as (x+ 2 ) (x– 2 ) over R.
With the definition of an irreducible polynomial having been given, we can
state that any polynomial f(x) in F[x] can be written uniquely as a product
of irreducible polynomials, unique but for the order of the factors and unit
factors.
(iv) Ideal A in F[x] is maximal (and hence prime) if and only if A is generated by
an irreducible polynomial p(x) in F[x]. Also any prime ideal in F[x] is maximal and
vice versa. Before proceeding further with our lesson we consider two important
and standard problems.
Problem 1
Let F and K are two fields and F  K. (i.e. K is an extension of the field F)
Suppose f(x), g(x)  F[x] are relatively prime in F[x], then prove that they are
relatively prime in K[x] also.
Solution
Since F  K, it is clear that F[x]  K[x]. Moreover both F[x] and K[x] are
Euclidean rings. Since f(x) and g(x) in F[x] are relatively prime in F[x], their g.c.d. in
F[x] is 1, the identity in F[x] being 1, the identity in F. So there exist polynomials
(x). (x) F[x] such that (x). f(x) + (x). g(x) = 1. This relation holds good in K[x]
also since all the polynomials (x), (x), f(x), g(x) are in K[x] also. Hence if d(x) is any
common divisor of f(x) and g(x) in K[x] is also 1 or any unit. Therefore f(x) and g(x)
continue to be relatively prime even in K[x] ■
Problem 2
F[x]
Let F be the field of real numbers. Prove that is a field isomorphic to
 x2 1 
the filed of complex numbers.
Solution 2
Let A denote the ideal <x2 + 1>, generated by x2+1, in the Euclidean ring F[x].
Then A = { (x2+1) f(x) : f(x)F(x) }. The residue classes modulo A in F[x], i.e. the
F[x]
elements in can be taken as A+(a+bx) where a, bF. In fact if p(x)  F[x] then
A
the residue class A+(a+bx) is same as A + (a+bx) where a + bx is the (unique)
remainder obtained on dividing p(x) by (x2 +1). Since F[x] is a Euclidean ring, the
division algorithm exists and we can determine q(x), r(x)  F(x) such that p(x) = q(x)
111
(x2+1) + r(x) where r(x) is the zero polynomial or the d(r(x)) = degree of r(x) is less
than 2, which is the degree of x2 + 1. Thus r(x) can be taken as a + bx where a, bF
and both a and b are zero, if r(x) is the zero polynomial. Now p(x) – r(x)  < x2 + 1 >
and so the residue classes p(x) + <x2 + 1 > and r(x) + <x2 + 1> are the same. Hence
F(x )
any residue class in can be taken as r(x) + <x2 + 1> where y(x) = a+bx, a,
 x2 1 
b  F. We may denote these residue classes modulo A as
F(x)
a  bx, so that  {a  bx} a, b F. Now we claim that the mapping
A
F( x )
:  C (where C denotes the field of complex numbers) defined by  (a  bx ) =
A
F( x )
a+ bi is an isomorphism of the field on to the field C. Clearly the mapping  is
A
‘onto’. Also  is one to one and well defined and this fact is seen by the following
“iff” conditions.

a  bx  c  dx  (a + bx) – (c+dx)  A = < x2 + 1|>


 (x2+1) is a divisor of (a-c) + (b-d) x
 a-c = 0 and b – d = 0
 a = c and b = d
 a + bi = c di
  (a  bx )  (c  dx)
Further  is a (ring) homomorphism.
In fact, (a  bx )  (c  dx )

=  ((a  c )  (b  d )x ) = (a+c) + (b+d)i = (a+bi) + (c+di)

=  (a  bx )  (c  dx )
which shows that  preservers addition.
And, ((a  bx ).(c  dx ))  [(a  bx )(c  dx )]

= [ac  (ad  bc )x  bdx 2 )

= [ac  bd )  (ad  bc )x  bd( x 2  1)

= [(ac  bd )  (ad  bc )x ] as bd (x2 + 1) A.


= (ac = bd) + (ad+bc) i
= (a+bi) (c+di)
=  (a  bx ). (c  cx ) .
112
which shows that  preserves multiplication also. Hence if we claim that  is an
F(x)
isomorphism of the field onto the field C, and the problem is answered ■
A
Note: Since x2 + 1 is obviously irreducible over the field F of real numbers, the
F( x )
ideal <x2+1> is a maximal, prime ideal and the residue class ring
 x2 1 
is a field.
8.3.3 POLYNOMIALS OVER THE FIELD OF RATIONAL NUMBERS:
In this section, unless otherwise stated let F denote the field Q of rationals, We
now specialize to the (Euclidean) ring F[x] of polynomials with rational coefficients.
In most of our results we obtain in the section, the polynomials are those in Z[x],
i.e. those with integer coefficients. The main result established here is Gauss
Lemma and this will find an application in proving an important theorem in the
next and final section of this lesson, viz., “If the ring R is a U.F.D so is the ring
R(x)”. We begin with two definitions.
Definition 3.3.1
A polynomial f(x)= a0+a1+a2x2+…anxn, where the coefficients ai are integers is
said to be primitive if the g.c.d. of the coefficients a0, a1, … a n is 1.
For instance, 3+4x+6x2+5x2 is a primitive polynomials while 3+6x-9x2+12x2-
3x4 is not a primitive polynomial.
Definition 3
If f(x) is any polynomial, as in the above definition, then the content of f(x) is
the g.c.d. of the coefficient a0, a1, … an.
The content of any primitive polynomial can be taken as 1 and it is easy to see
that if d is the content of a polynomials f(x) in Z[x] then we can write f(x) = dg(x)
where g(x) is a primitive polynomial.
Lemma 3.3.2
The product of any two primitive polynomials in Z[x] is also primitive.
Proof
Let the two primitive polynomials be f(x)= a0 + a1x + a2x2 + … + anxn, and
g(x) = b o+ b1x + b 2x2 + … + b mbm. Suppose the lemma is not true. Then all the
coefficients in the polynomial f(x).g(x) must be divisible by some integer > 1 and
hence by any prime divisor, say p of that integer. We will now show that this is
impossible. The coefficient of x in the product polynomial f(x) . g(x) is
t
 a i b i-1 for 0  t  m  n
i 0
Since f(x) is primitive, not all coefficients, a i in f(x) are divisible by this prime p.
Let ai be the first coefficient in f(x) which is not divisible by p. Similarly, since g(x)
is also primitive we can find the first coefficient, say bk in g(x) which is not divisible
113
by p. (Here 0  j  n and 0  k  m). Now the coefficient of x j+k in f(x) . g(x) is
cj+k = (a 0 bj+k + a1 bj+k-1 + … + aj-1 bk+1) + ajbk + (aj+1 bk-1 + aj+2 bk-2 + … + aj+k b 0). By
our choice, p is divisor of a 0, a1 … aj-1 and hence p| (a0 b j+k + a1bj+k-1 + … + aj-1 bk+1).
Similarly p is a divisor of bk-1, bk-2 … b 0 & so p|(aj+1 b k-1 + aj+2 bk-2 + … + aj+k b0).
But by our assumption p | cj+k. Hence the expression for cj+k shows that p|aj pk.
But since p is a prime the expression for cj+k shows that p| aj pk. But since p is a
prime, p must be a divisor of at least one of aj and b k which is a contradiction to
our choice of a j and b k. Thus f(x).g(x) must be primitive and the lemma is proved. ■
Now we are in a position to establish Gauss lemma which is stated as theorem
below.
Theorem 3.3.3
(Gauss’s lemma). If a primitive polynomial f(x)  Z[x] can be factored over F,
(i.e. can be factored as a product of two polynomials with rational coefficients) then
f(x) can be factored as the product of two polynomials with integer coefficients.
Proof
Assume that f(x) = a(x).b(x) where a(x).b(x)  F(x). Suppose d 1 is a common
multiple of the denominators in the coefficients of a(x), then we can write
a (x )
a(x) = 1 where a1(x) has integer coefficients. If c1 is the content of a1(x)
d1
c1
a1(x) = c1(x), where  (x) is a primitive polynomial and so a(x) = ( x ) . Similarly
d1
c2
treating b(x) we can write b(x) = ( x ) , where c2, d2 are integers and (x) is a
d2
c c1 c 2
primitive polynomial. Hence a(x)b(x) = c/d (x) . (x) where  . ,
d d1 d 2
c
i.e. f(x) = (x). (x), so that df(x) = c(x) (x). Now f(x). (x), (x) are all in Z (x) and
d
c, d  Z. Since both (x) and (x) are primitive, by the last lemma, (x).(x) is also
primitive. Also f (x) is primitive. Thus, equating the contents of both sides of the
c
equation df(x) = c(x).(x), we find d = c. Hence =1 and f(x) = a(x) b(x) = (x) (x).
d
Thus f(x) has been factored as a product of two polynomials, viz. (x) and (x) with
integer coefficients. Hence the theorem. ■
As an immediate consequence of the above theorem we have the following
corollary.
Corollary 3.3.4
Any irreducible polynomial over Z, continues to be irreducible over F, the field
of rational. ■
The following theorem gives a sufficient condition for the irreducibility of a
polynomial in Z[x]
114
Theorem 3.3.5
(The Eisenstein’s Criterion) Let f(x) = a0 + a1 x + a2 x2 + … + an xn be a
polynomial in Z[x] of degree n. If there exists a prime number p such that
(i) p is a divisor of ai for i = 0, 1, 2, …, n-1.
(ii) p is not a divisor of an and
( iii) p2 is not a divisor a 1.
then f(x) is irreducible over Z and hence irreducible over F.
Proof
Without loss of generality, we may assume that f(x) is primitive. For otherwise,
if d (>1) is the content of f(x) then we may divide each co-efficient in f(x) by d and
obtain a primitive polynomial f1(x) say. Since p is not a divisor of a n, p will not be a
prime divisor of d and so the conditions (i) , (ii), (iii) imposed on the coefficients f(x)
still hold good in the case of f1(x) also. We may then treat f1(x) instead of f(x) and
establish the theorem. Thus we may take f(x) to be a primitive polynomial. Suppose
now f(x) can be factorised as a product of two polynomials over F, then by Gauss
lemma f(x) can be factorised as a product of two polynomials over Z. Thus, if we
assume that the theorem is false, that is if f(x) is not irreducible then we can write
f(x) = (b 0 + b1x + …+ brxr). (c0 + c1x +…+ cr xr) where the b’s and c’s are integers and
r > 0, s > 0 such that r + s = n. Equating the constant terms on both sides of the
above identity we get a0 c0. Now by data the prime p is a divisor of a0 and hence is
a divisor of at least one of b0 and c0. But since p2 is not a divisor of a0, p cannot be
a divisor of both b 0 and c0. Now not all the coefficients b0, b1, … b r can be divisible
by p and this is false as p does not divide a n. So let b k be the first co-efficient that
is not divisible by p. By our assumption this means that p|b0 b1 … bk-1 and 0
< k  r < n. Now, from the factorisation of f(x), equating the coefficients of x k we
obtain ak = b0 ck + b1 c k-1 + … + bk cn. Since k < n, in the above equation we find
that p is a divisor of the integers, ak, b 0, ck, b 1 ck-1, … b k-1 c1 (by our assumption p
is a divisor of ai for i = 0, 1, 2, n-1 and also of b 0, b 1, … bk-1). This fact implies that
p|bk c0. But this is impossible since, by our assumption, p is neither a divisor of
bk, nor a divisor of c0. This contradiction proves the theorem. ■
As an application of the above theorem we prove in the following problem the
standard result viz. “Any cyclotonic polynomial 1 + x + x2 + … + xp-1 where p is a
prime is irreducible over Fn.
Problem 3
Prove that the polynomial 1 + x + x2 + … + xp-1 where p is a prime number, is
irreducible over the field F of rational numbers.
Solution
Let f(x) denote the polynomial 1 + x + x2 + … xp-1. Since all the coefficients in
f(x) are 1 the Eisentein’s criterion for irreducibility cannot be applied for f(x). So,
remembering the obvious fact that f(x) is irreducible, we apply the criterion for
115
f(x+1) and establish that f(x+f) is irreducible. Now f(x+1) = 1+ (x+1) + (x+1)2 + … +
( x  1) p  1
(x+1)p-1 which can also be written as equal to by summing up the G.P.
( x  1)  1
( x  1) p  1
Thus f(x+1) =
x
 p p p  p   p 
= x p   x p 1   x p  2  ...   x p  r  ...    x     1   x
 1 2 r  p  1  p 

 p p(p  1)(p  2)...(p  r  1) p


where,    p c r  . and    1
r 1.2.3....r p
p p p  p   p 
f(x+1)= x p 1    x p  2   x p  3  ...    x p  r 1  ...    x    …(1)
1  2 r  p  2  p  1

p p(p  1(p  2)...(p  r  1)


Consider the integer    where 1  r  p - 1.
r 1.2.3....r
Since p is a prime, none of the factors 2, 3, … r in the denominator can divide p
p
which is in the numerator. However since   is an integer, these factors must
r
cancel out in the remaining factors in the numerator, viz. (p-1) (p-2) … (p-r+1).
(p  1)(p  2)...(p  r  1) p
Thus is in integer and hence the integer   is divisible by
1.2.3.... r r
p for 1  r  p – 1. Thus the integer p, which is a prime, is a divisor of all the
coefficients in f(x+1), except the (leading) coefficient of xp1; further p2 is not a divisor
 p   p 
of   , the constant term since   =p. Thus the prime p satisfies all the
 p  1  p  1
conditions required for Eisentein’s Criterion. Hence we conclude that the
polynomial f(x+1) is irreducible over F and so f(x) is also irreducible over F. ■
Polynomial Rings over Commutative Rings:
This is the last section for the lesson and the main result proved in this
section. That is if a ring R is a U.F.D. (the definition for the U.F.D. is given in this
section) so is the ring R[x] of all polynomials. We already mentioned that whenever
R is an integral domain so is R[x]. Suppose now R denotes an integral domain with
identity, we can talk about divisibility, units, associates, prime or irreducible
elements. g.c.d., etc. in R, in exactly the same manner as we did for Euclidean
rings. In fact whenever a = bc where a, b, c  R, we say that b and c are divisors of
a and write b|a, c|a. An element uR is said to be unit in R iff there exists an
element v in R (viz. the inverse of u) such that uv =1, the identity in R. Again,
whenever b = au where u is a unit, the elements a and b of R are said to be
116
associates. One can easily show that a and b are associates in R iff each of a, b
divides the other. An element p  R is said to be a prime element (or an irreducible
element) if p is not a unit and if the only divisors of p are units and associates of p.
Now we come to the important definition of a U.F.D.
Definition 3.8.4
An integral domain R with identity is said to be a unique factorization domain
(abbreviated as U.F.D.) if (i) any non-zero element in R is either a unit or can be
factorized as a product of a finitely many prime (or irreducible) elements of R;
(ii) the factorization mentioned above is unique but for order and associates of the
prime elements. ■
In the last lesson, we shown (in a theorem) that the above two defining
properties of a U.F.D. are enjoyed by a Euclidean ring. Thus any Euclidean ring
and in particular Z, (the ring of integers), Z[i] (the ring of all Gaussian integers) and
F[x] (the ring of all polynomials in x over any field F) are all examples of a U.F.D.
However the converse is false. That is, a Euclidean ring need not be a U.F.D. In
fact, one can show that Z[x], the ring of all polynomials in x over Z, is only a U.F.D.
and not a Euclidean ring. So also, the ring F(x1,x2) of all polynomials in the two
indeterminate over a field F is only a U.F.D. and not a Euclidean ring. The symbol
F(x1,x2) denotes F(x1)(x2) which results from F by adjoining first x1 to obtain the ring
F1  F(x1) and then adjoin x2 to F1, so as to obtain F1 (x2)  F(x1)(x2). The fact Z[x] is
a U.F.D. follows from classical algebra in which we have studied standard
properties of polynomials in x with integer coefficients. So also, one can recognize
that the ring F(x1,x2) is a U.F.D. However both Z[x] and F(x1,x2) are not Euclidean
rings. For, in the case of Z[x], we have the example of a prime ideal, (viz, the ideal
of all polynomials  Z[x] which lack the constant term) which is not maximal, (Refer
Lesson 6) Thus Z[x] cannot be a Euclidean ring all prime ideals are maximal.
Again, in the case of the ring F(x1,x2) the ideal generated by <x1, x2> cannot be a
principal ideal, as can be shown. Hence F(x1, x2) cannot be a Euclidean ring. For
that it is known that it is an Euclidean ring with all ideals are principal ideals.
We list below some of the divisibility properties of a U.F.D. and these can be
easily established just as in the case of the ring Z. Let C denote a U.F.D. in all that
follows here. Then (i) for, a, b,  R, there exists a g.c.d., denoted by (a, b), for the
pair of elements a, b and it is unique but for associates. In fact, by the definition of
1  2  m 1  2  n
a U.F.D. we can write a =  ... , b   ... where the p’s and the q’s are
p1 p 2 p m q1 q 2 q n
distinct primes and the ’s, ’s are positive integers. It may or may not happen that
some or all of the primes pi are same as the primes qi and vice-versa. Let us
suppose that, after re-labelling, if necessary, the p’s and the q’s, pi, and q i are the
only primes which are associates for 1  i  k where k  min (m, n). With these
1  2  k
assumptions made one can easily check that (a, b) = . ... where i (i,  i) for
p1 p 2 p k
117
i = 1, 2, … k. If on the other hand, no prime pi is an associate of any qi and vice-
versa. Then g.c.d. (a, b) = 1. Thus the g.c.d. (a, b) can be determined.
(ii) If a and b are relatively prime in R, i.e. (a, b) = 1 then for c  R, a|bc  a|c.
(This fact may be easily proved).
(iii) If a is a prime element in R then for b, c  R, a|bc  a |b or a |c. (This is
a corollary from the result (ii)).
It is possible to establish an analogue of Gauss’ lemma for the case of R[x]
when R is a U.F.D. just as when R = Z. For this purpose we make the following
analogous definitions.
Definition 3.8.5
If f(x) = a0 + a1 x + a2 x2 + - + am xm is a polynomial in R[x] then the content of
f(x), denoted by c(f) is the g.c.d. of the coefficients a0, a1 … am.
One can note that c(f) is unique, but for unit factors.
Definition 3.8.6
A polynomial f(x) in R[x] is said to be a primitive polynomial if c(f) = 1 or a unit.
It is quite easy to see that for any polynomial f(x) in R[x] one can write
f(x) = c(f).f1(x) is a primitive polynomial. Moreover the above decomposition of f(x) as
a product of the element c(f) of R and the primitive polynomial f1(x) in R[x] is unique
but for unit factors.
The following lemma can be proved along the same lines as, the corresponding
Lemma 3 in the case of Z[x] was proved and so the proof is omitted. (You are
invited to supply the proof, as an exercise).
Lemma 3.8.4
If R is a U.F.D. then the product of any two primitive polynomials in R[x] is
also primitive.
Since we remarked in earlier that any polynomial f(x) in R[x] can be written as
f(x) = c(f).f1(x) where f1(x) is a primitive polynomial in R[x], the above lemma leads
us to the following corollary.
Corollary 3.8.2
If f(x), g(x)  R[x], then c(fg) = c(f) c(g) (upto units) and generally, if f1(x) f2(x) …
fn(x)  R[x], then c(f1, f2 …. fn) = c(f1) c(f2) … c(fn) (upto units).
Now we state and prove the analogue of Gauss lemma for the case of a U.E.D.
Lemma 3.8.5
Let R be a U.F.D. and f(x) in R[x] be primitive and suppose that F is the field of
quotients of R, in which R can be imbedded. Then f(x) is irreducible over F (i.e. as
an element in F[x], iff it is irreducible over R ( i.e. as an element in R[x]).
Proof :
Since any element of F can be taken as (a/b)where a, b  R and b  0 the ring
R[x] can be considered as a sub-ring of F[x]. Hence if the (primitive) polynomial f(x) in
118
R[x] is irreducible over F, then certainly f(x) is irreducible over R is clearly valid over F
also. Conversely suppose f(x) is irreducible over F also. Then we prove the lemma by
reducible absurdum. So assume that f(x) has a non factorization over F, i.e. let f(x) =
g(x) h(x) where g(x) an h(x) are non-constant polynomials over F. (Here we make use
of the easily provable fact that the only units in R[x], where R is any integral domain
with identity, are the units in R ). Since the coefficients in the g(x) are all quotients of
a
the form with a,b  R and b  0, we can replace all the coefficients in g(x) by
b
g (x )
quotients with the same denominator, say d1R and so, we can write h(x)  1
d1
h1( x )
where g1(x)  R[x]. Similarly we can write h(x)  where e 1  R h2 (x)  R. Let
e1
c(g1) = a1c(h1) = b1 so that we can take g1(x) = a1 g2(x) h1(x) = b 2 h2 (x) where g2(x), h2(x)
g1(x ) h1(x )
 R[x] and both are primitive. Now f(x) = g(x) h(x) = =
d1 e1
a1g 2 ( x ) b1 h2 (x) a 1b1
 g2(x) h2(x). Since both g2(x) and h2(x) are primitive, we get,
d1 e1 d1e 2
in view of the last Lemma 4, that the product g2(x). h2(x) is also primitive. Thus we
p
can write f(x) = f1(x) where p = a1 b 1. q0 = d1 e1 and f1(x) = g2(x) h2(x) is a primitive
q
polynomial in R[x]. Hence q0f(x) = pf1(x) and equating the contents of both sides, we
get q = p, since both f(x) and f1(x) are primitive. This means that f(x)  f1(x) = g2(x)
h2(x) and so f(x) admits a non-trivial factorization over R, as both g2(x) and h2(x) are
clearly non constant polynomials in R[x]. But this is a contradiction to our
assumption that f(x) is irreducible over R and therefore the lemma is proved.
Before to establish the main and final theorem of this section we need one more
lemma which proves the unique factorization of any primitive polynomial in R[x].
Lemma 3.8.6
If R is U.F.D. and f(x) in R[x] is a primitive polynomial then f(x) can be factored
in a unique manner as a product of irreducible polynomials in R[x].
Proof:
Here R[x] can be considered as a polynomial in F[x]. But, F being a field we
know that F[x] is a Euclidean ring and hence is a U.F.D. Therefore as an element of
F[x], f(x) admits a unique factorization into irreducible polynomials over F, say f(x) =
f1(x) f 2(x) where each f1(x) is a polynomial irreducible over F. As we mentioned in
c
the proof of the last lemma, each fi(x) can be written as 1 pi(x) where ci, di  R and
di
pi(x) is a primitive polynomial in R[x]. But since fi(x) is irreducible over F, so is p i(x)
and hence, by the last lemma, pi(x) is irreducible over R. This we are able to write
as
119
c1c 2 ...c k
f(x) = fi(x) f2(x) … f2(x) = p2(x) …pk(x)
d1d 2 ...d k

where each p i(x) is a primitive, irreducible polynomial in R[x] and ci, d i  R for
all i = 1, 2, … k. the above equation gives d 1d 2…dkf(x) = c1c2 …ck pi(x) p2(x)
…pk(x). Now by data f(x) is primitive and the product of primitive polynomials is
also primitive over R (by Lemma 4). Therefore equating the contents of both sides
of the last equation, we get d1d2 … dk = c1 c2 … ck and so we obtain f(x) = p1 (x) p2(x)
… p k(x). That is we have factored f(x) as a product of irreducible polynomials p i(x)
in R[x]. Further this factorization is unique to within order and unit factors, since
F[x] is a U.F.D. (as earlier mentioned) and each pi(x) in R[x] being primitive and
irreducible over F. Hence the lemma follows ■
Now we have all the necessary materials for the proof of our main theorem,
which we now state and prove.
[Note: For the proof of this main theorem, one must first establish the Lemma
5 and 6 and then prove the theorem; Lemma 4 may also be quoted].
Theorem 3.8.5
If R is a U.F.D then so is R[x].
Proof
Let f(x) be any arbitrary polynomial over R. Then we can write f(x) in a unique
way as f(x) = c(f) f1(x) where c(f) is the content of f(x) and f1(x)is a primitive
polynomial in R[x]. Now c(f) is in R and is unique but for unit factors. Also, since
f1(x) is primitive, by the last lemma f1(x) can be factored in a unique but for the
order of the factors and the associates of the factors. Again since c(f) is in R and R
is a U.F.D. c(f) admits a unique factorization as a product of prime elements in R
and we claim that this is the only factorization for c(f) when it is factored over R.
(i.e. as an element of R[x]. In fact, suppose we factorize any element of R, in
particular c(f) over R and obtain c(f) = a 1(x) . a2(x) … a m(x). Then equating the
degree of both sides, considering c(f) as an element of R[x], we get
0 = deg (a1(x)) + deg(a2(x)) + …+ deg. (am(x)). Since the degree of any non-zero
polynomial is a non-negative integer the above equation implies that each ai(x) is a
polynomial of degree 0. i.e. each a 1(x) is an element of R. Hence the only possible
factorization of c(f) are those it can have as an element of R. Hence we have
obtained the unique factorization of f(x) by putting together the unique factorization
of c(f) as an element of R and the unique factorization of the primitive polynomial
f1(x) over R. This concludes the proof of the theorem. ■
Just as we defined the ring R[x] of all polynomials in a single indeterminate
x over R, we can define the ring [x1 x2 … xn] of all polynomial in n in determinates
(or variables) x1, x2, … xn over R. This can be defined successively as follows.
Let R1  R(x1) be the ring obtained from R by the (ring) adjunction of one variable x1.
Again by adjoining another indeterminate x 2 to R1, we obtain another ring
R2 = R1[x2]  R[x1] [x2]. We denote this ring R2 by R[x1, x2] and call it the ring of all
120
polynomials in the two variables x1, x2 with coefficients in R. Similarly we may
adjoin yet another variable x2 to R2 and obtain the ring R3 = R2[x2]. We denote this
ring R2 by R[x1, x2] and call it the ring of all polynomials in the two variables x1, x 2
with coefficients in R. Similarly we may adjoin yet another variable x 3 to R2 and
obtain the ring R3 = R2[x2] and this ring R3 is denoted by R[x1, x2, x3] … and in
general R[x1, x2 … xn] are all integral domains.
Similarly by successive application of the last theorem the corollary follows.
Corollary 3.8.3
If R is a U.F.D. so is R [x1, x2, … xn].
Again when F is a field it is known that F[x1] is a Euclidean domain and so is a
U.F.D. Hence by successive application of the theorem we obtain the following
corollary also.
Corollary 3.8.4
If F is any field then the ring F[x1, x2, … x n] is a U.F.D.
Solved Problems
1. Find all the units of the ring Z[i] of Gaussian integers.
Solution
We know that Z[i] = {a+ib : a, b  Z} is the ring of Gaussian integers.
Here 1 + 0 i is the unit element.
Let x + i y be a unit & x + i y be its inverse.
Then (x + iy) (x + iy) = 1 + 0i  (x x – yy) + i (xy + yx) = 1 + oi
Equating real and imaginary parts, we get
x x - y y = 1 and x y  + yx = 0
Squaring and adding, we get
x2 x2 + y2 y2 + x2 x2 + y2 x2 = 1
 (x2 + y2) (x2+y2) =1
 (x2+y2) = 1 (Since the product of two positive integers can be
equal to 1 iff each of them = 1).
 x2 = 0, y2 = 1 or x2 = 1, y2 = 0
Thus x = 0, y =  1 or x =  1, y = 0
Therefore the only units of Z[i] are 0  i,  1 + 0 i. i.e. 1, -1, i, -i
2). If R is an integral domain and if F is its field of quotients, prove that any
g( x )
element f(x) in F[x] can be written as where g(x)  R[x] and a  R.
a
Solution

p 
We know that F =  : p  R, 0  q  R 
q 
121
a 0 a1 a
Let f(x)  F(x). Then f(x) = x  ...  n x n where a0, a1, … an  R. and
b 0 b1 bn
b 0, b1 … bn are non-zero elements of R. Now b0, b1, … bn are non-zero
elements of F. Therefore b0, b1, … b n is also a non-zero element of F and so is
invertible.

b 0 b1... b n  a 0 a1 a 
 f (x )    x ...  n x n 
b 0 b1 ... b n  b 0 b1 bn 

(a 0 b1 b 2 ... b n )  (b 0 a1 ... b n )x  ....  (b 0 b1 ...b na 0 )x n


=
b 0 b1 ... b n

g( x )
= where g(x) = (a0 b1 b2 … b n) + b 0 a1 … bn)x + …  R(x)
a
and a = b0 b1 … bn  R
3) Prove that x4+2x+2 is irreducible over the field of rational numbers.
Solution
Let f(x) = 2 + 2x + 0 x2 + 0 x3 + 1 x4
Now f(x) is a polynomial with integer coefficients. Also 2 is a prime number
such that 2 divides each of the coefficients of f(x) except the coefficient 1 of the last
term x4. Also 2 2 is not divisor of f(x) is irreducible over the field of rational
numbers.
8.4. REVISION POINTS
Polynomial ring, polynomial over the field of rational numbers, primitive
polynomial, Gauss Lemma, unique factorization domain, The Eisenstein’s
Criterion.
8.5. INTEXT QUESTION
1 Let F and K be two fields and FK. (i.e. K is an extension of the field F)
Suppose f(x), g(x)  F[x] are relatively prime in F(x), prove that they are
relatively prime in K[x] also.
F[x]
2 Let F be the field of real numbers. Prove that is a field isomorphic
 x2 1 
to the filed of complex numbers.
3 Prove that the polynomial 1 + x + x2 + … + xp-1 where p is a prime number, is
irreducible over the field F of rational numbers.
4 Find all the units of the ring Z[i] of Gaussian integers
5 If R is an integral domain and if F is its field of quotients, prove that any
g( x )
element f(x) in F[x] can be written as where g(x)  R[x] and a  R.
a
6 Prove that x4+2x+2 is irreducible over the field of rational numbers.
122
8.6. SUMMARY
Polynomial rings, polynomial over the field of rational numbers, primitive
polynomial, Gauss lemma, the Critenion unique factorization domain are discussed
in detail with relevant examples, Eisenstein’s Criterion, problems, lemmas and
theorem.

7. TERMINAL EXERCISES
1. Show that the units in a commutative ring with identity form a group
under multiplication.
2. Prove that the set N of all polynomials in Z[x] whose constant (terms are
either zero or a fixed prime p is a prime ideal in the ring Z[x]. Is this ideal
N, a maximal ideal?
3. Prove that, in an integral domain R with identity, two elements a, b of R
generate the same ideal iff a and b are associates.
4. Prove: (i) N is an ideal of R
R m
(ii) In R = , if x  0 for some m, then x =0
N
(Here x = N + x  R with x  R)
5. Let R be a commutative ring and suppose that A is an ideal of R.
Let N(A) = { x  R | xn  A for some n}. Prove that
(i) N(A) is an ideal of R which contain A.
(ii) N(N(A)) = (A)
Note: N(A) is often called the radical of A.
6. Prove that no prime of the form 4n + 3 can be written as a2 + b2 where a
and b are integers.
7. Prove that x2 + x + 1 is irreducible over F, the field of integers mod 2.
8. Prove that a principal ideal ring, is a unique factorization domain.
9. If Z is the ring of integers, show that Z[x1, x2, …, xn] is unique factorization
domain
8. SUPPLEMENTARY MATERIALS
1. Algebra by Micheal Artin, Prentice Hall of India, New Delhi, 1994.
2. Lectures in Abstract Algebra Volumes I, II and III by N. Jacobson, D. Van
Nostrand Co., New York, 1976.
9. ASSIGNMENTS
1. If p be a prime number. Suppose c is an integer relatively prime to p and
suppose that there exist integers x and y such that cp = x2 + y2. Then
prove that p can be written as the sum of two square. That is p = a2 + b2
for some pair of integers a and b.
123
2. If p is any prime number of the form 4n + 1 then prove that there exists an
integer x such that x2 +1  0 (mod p).
3. If p is any prime number of the form 4n + 1, then prove that d can be
expressed as a sum of two squares; i.e. p = a2 + b 2 for some integers a, b.
4. If F is any field, prove that the ring F[x] of all polynomials in x over F is a
Euclidean ring.
5. The product of any two primitive polynomials in Z[x] is also primitive.
6. If R is a U.F.D. then prove that the product of any two primitive
polynomials in R[x] is also primitive.
7. If R be a U.F.D. and f(x) in R[x] be primitive purpose F is the field of
quotients of R, in which R can be imbedded. Then prove that f(x) is
irreducible over F ( i.e. as an element in F[x] ), iff it is irreducible over R
(i.e. as an element in R[x] ).
8. If R is a U.F.D. so is R[x].
10. SUGGESTED READINGS / REFERENCE BOOK / SET BOOK
Algebra –By Michael Artin. PHI Learning P Ltd
11. LEARNING ACTIVITIES
Students are requested to attend the PCP classes and work out the problems
given in the lesson
12. KEYWORDS
Polynomialrings, Polynomial over the field of rational numbers, Primitive
polynomial, Gauss Lemma, The Eisenstein’s Criterion, Unique factorization
domain.

124
LESSON - 9

VECTOR SPACES
9.1. INTRODUCTION
Up to this point we have been introduced to groups and to rings, the former
has its motivation in the set of one- to – one mappings of a set onto itself, the latter
in the set of integer. The third algebraic model which we are about to consider
vector space can in large part, trace its orgins to topics in geomentry and physics.
Vector space differs from these previous two (group and ring) structures in
that one of the product defined on it uses elements outside of the set itself. It own
their importance to the fact so many models arising in the solutions of specific
problems turn out be vector spaces.
In this lesson, vector space over afield F, subspace homomorphism, Internal
direct sum, Linear combination over F, linear spen of G, Finite dimensional,
Linearly dependent over F, Basics of V, Dimension of the vector space, Dual space,
Second dual of V, annihilator of w, are defined with suitable examples, Lemmas,
theorems and problems are given.
9.2. OBJECTIVE
Vector space, subspace, homomorphism, Internal direct sum, Linear
Combination over F, Linear space of S, Finite dimensional over F, Basics of V,
Dimension of the vector space, dual space, second dual, annihilator the explained.
9.3. CONTENTS
9.3.1 Vector Space
9.3.2 Definition: Subspace
9.3.3 Definition : Vector Space
9.3.4 Theorem : Internal Direct Sum
9.3.5 Dual Spaces
9.3.1 DEFINITION – VECTOR SPACE
A non-empty set V is said to be a vector space over a field F, if V is an abelian
group under the operation denoted by ‘+’ and if for every F, vV, there is defined
an element, written as v in V satisfying the following conditions
(1)  (v + w) = v + w
(2) ( + )v = v + v
(3)  (v) = ()v
(4) 1.v = v
For all ,   F and v, w  V and 1 represents the unit element of F under
multiplication.
Note (1) The ‘+’ in condition (1) above is that of the group <V, +>, while the ‘+’ in the
left hand side , condition (2) is that of the field F and the ‘+’ in the right
hand side is that of the group <V, +>.
125
Note (2) If V is a vector space over F, the there is a mapping satisfying the above
four conditions. This mapping is usually called scalar multiplication. We
often refer to the elements of F as scalars and the elements of V as vectors.
We shall consistently use the following notations:
a. F will be a filed.
b. Lowercase Greek letters will be elements of F; we shall often refer to
elements of F as scalars.
c. Capital Latin letters will denote vector spaces over F.
d. Lowercase Latin letters will denote elements of vector space. We shall
often call elements of a vector spaces as vectors.
Example 3.1.1
Let F be a field and let K be a field which contains F as a sub-field. We
consider K as a vector space over F, using the ‘+’ of the vector space the addition in
the elements of K, and by defining, for F, vK, v to be the product of  and v as
elements in the field K. The axioms (1), (2), (3) for a vector space are the
consequences of the right distributive law, left distributive law and associative law
respectively which hold for K as a ring.
Example 3.1.2
Let F be a field and let V be the totality of all ordered n-tuples (1, 2, …n)
where i  F. Two elements (1, 2, …n) and ( 1,  2, … n) of V are declared to be
equal if and only if i =  i for each i = 1, 2, …n. We now introduce the operation ‘+’
in V as (1, 2, … n) + ( 1,  2, … n) = (1+ 1, 2+ 2, … n+ n). (Component wise
addition). Then V becomes an abelian group under this addition ‘+’. Now let us
define the map FV  V (i.e. scalar multiplication in V) by γ(1, 2, …n) = (γ1, γ2,
… γn) where γ  F. One can easily verify that V is a vector space over F and this
vector space is usually denoted by F(n).
Example 3.1.3
Let F be any filed and let V=F[x], the set of polynomials in x over F. We choose
to ignore , at present, the fact that in F[x] we can multiply any two elements, and
merely concentrate on the fact that two polynomials can be added and that a
polynomial can always be multiplied by an element of F . With these natural
operations F[x] is a vector space over F.
9.3.2 DEFINITION: SUBSPACE
If V is a vector space over F and if W  V, then W is a subspace of V if under
the same operation as for V, W itself , forms a vector space over F. Equivalently, W
is a subspace of V iff w1, w1 W and ,   F always imply that w1 + w2W.
Lemma 3.2.1
If V is a vector space over F, then
(1) . 0 = 0
(2) 0. v = 0
126
(3) (-) v = - (v)
(4) If v  0, then  v  0 implies that  = 0,
where   F , v  V, 0 represents the zero for addition in V and 0 represents the
zero for addition in F.
Proof
(1) Since  0= (0+0) = 0 + 0 we get 0 = 0.
(2) Since 0v = (0+0)v = 0v +0v we get 0v = 0.
(3) Since 0= [ + (-)] v = v + (-) v, (-) v = -(v).
(4) If  v = 0 and   0,then 0 = -1 0= -1 (v) = (-1) v = 1.v = v.
Let V be a vector space over F and let W is a subspace of the vector space V
(Over a field F) then since (V, +) is an abelian group, we can construct the quotient
group V/W; its element are the cosets of W in V viz, v+W where vV. The
commutativity of the addition, from group theory assures us that V is an abelian
W
group. We intented to make of it a vector space. If   F , v +W V , we define
W
( v +W) = v +W . Then we show that this product is also well defined; that is , if
v+W = v΄+W than (v+W) =( v΄+W).
Now because v+W = v΄+W then v- v΄ is in W; Since W is subspace of V it
implies that ( v- v΄) must also in W. Using previous lemma says that v  v   W
and also v  W  v   W . Thus  (v  W )  v  W  v   W   (v   W ) ; the
product has been shown to be well defined. The verification of the vector space
axioms for V is routine. Now V can be realised as a vector space over F, as the
W W
following lemma shows.
Lemma 3.2.2

If V is a vector space over F and if W is a subspace of V, then V is a vector


W
space over F where, for v1 + W, v2 + W V , and   F, we define
W
1. (v1 + W) + (v2+W) = (v1 + v2) + W,
2.  (v1+W) = v1 + W.
Proof

The commutativity of addition in V assures that V is an abelian group.


W
V V
Now let us show that the map F   given by  (v1+W) = v1 + W is well
W W
defined. If v  W  v   W then we must show that v  W  v   W . Since
v  W  v   W , v  v   W . Since W is subspace of V it implies that ( v- v΄) must
also in W. using previous lemma we can write that  v   v   W and also
127
v  W  v   W . Thus  (v  W )  v  W  v  W   (v   W ) ; so that
 (v  W )   (v  W ) . Now let us verify the four axious for the vector space.
1. [(v1+W)+ (v2+W)] = [(v1+v2) + W] =  (v1+v2) + W = v1 + v2 + W
= (v1+W) + (v2+W)
2. (+) (v1+W) = (+)v1+W = (v1+v1) + W= (v1 + W) + (v1 + W)
= (v1 + W) + (v1 + W)
3.  [(v1 + W)] = [v1+W] =v1 + W = (v1+W)

4. 1 (v1+W) = (1.v1+W) = v1 + W , where v1, v2  V, , ,  F. Hence V is a


W
vector space. This vector space V is called the quotient space V by W.
W
Just as for groups and rings, we now define the notion of (Vector space)
homomorphism between vector spaces.
9.3.3 DEFINITION : HOMOMORPHISM
If U and V are vector spaces over F, then a mapping T of U into V is said to be
a homomorphism, if
1. T(u1 + u2) = T(u1) + T(u2)
2. T(u1) = T(u1) for all u1, u2  U and all   F.
If T, in addition, is one-to-one, we call it an isomorphism. The kernel of T is

defined as u  U T (u )  0  where 0 is the identity element for addition in V. One
can verify that Kernel of T is a subspace of U and that T is an isomorphism if and
only if its Kernel is (0). Two vector spaces are said to be isomorphic if there is an
isomorphism of one onto the other. The set of all homomorphisms of U and V will
be written as Hom(U , V ) .
Theorem 3.3.1
If T is a homomorphism of U onto V with Kernel W , then V is isomorphic to
U . Conversely, if U is a vector space and W a subspace of U, then there is a
W
homomorphism of U onto U where Kernel is precisely W.
W
The proof is similar to the proof given in group theory.
Definition 3.3.2
Let V be a vector space over F and let U 1, U2, … Un be subspaces of V. Then V
is said to be the internal direct sum of U1, U2, … Un if every element vV can be
written in one and only way as v = u1 + u2 + ….+ un where ui  Ui.
Given any finite number of vector spaces over F say V1, V2, Vn. Consider the
set V of all ordered n-tuples (v1, v2, …vn) where viVi. We declare that two elements
(v1 … vn) and (v1 …vn) of V to be equal if and only if vi = v1 for each i. We add two
elements component wise i.e. (v1, v2, …vn) + (w1, … wn) = (v1+w1…vn+wn). Also if  
128
F and (v1, v2, … vn)  V, we define  (v1….vn) = (v1, …, vn). One can see that V is
vector space over F under these operations. We call V the external direct sum of
V1, V2, … V n and denote it by writing.
V = V 1  V2  …  Vn
9.3.4 THEOREM : INTERNAL DIRECT SUM
If V is the internal direct sum of U1, U2, … Un then V is isomorphic to the
external direct sum of U1, U2, … Un.
Proof
Given v  V, v can be written, by assumption, in one and only way, as v = u 1 +
u2 + …+un where ui  Ui. Define the mapping T:V  U1  U2  … Un by T(v) = (u1,
u2, … un). Since v has a unique representation T is well defined. It clearly is onto,
for the arbitrary elelment (w1, … wn)  U1  U2  … Un is T(w) where w =
w1+w2+……+wn  V . Then can easily verify that T is a one-to-one and
homomorphism. Hence the proof.
LINEAR INDEPENDENCE AND BASES
Definition 3.4.1
If V is a vector space over F and if v1, v2, … vn  V, then any element of the
form 1 v1 +2 v1 …+ nvn where the i  F is called linear combination over F of
v1, v 2 … vn.
Definition 3.4.2
If S is a nonempty subset of the vector space V, then L(S), the linear span of S,
is the set of all linear combinations of finite sets of elements of S.
Lemma 3.4.3
L(S) is a subspace of V.
Proof
If v and w are in L(S), then v =  1s1 + 2s2 +…. +  nsn and w = 1t1 + 2t2 + mtm
where  s and  s are in F and the si and tj (for i = 1 … n and j = 1, 2 m) are all in
S. Thus for ,  in F,
v + w =  ( 1s1 + …+  nsn) + ( 1 t1 + …+ m tm)
= (1)s1 + … + ( n)sn + (1) t1 + … + ( 1) t1 +…+ ( m)tm
and so is again in L(S). Hence L(S) is a subspace of V.
Note: It is not difficult to see that this subspace L(S) is the (set theoretically)
smallest subspace of V, which contains the set S.
Lemma 3.4.4
If S and T are subsets of V, then
(1) S  T implies L(S)  L(T)
(2) L(ST) = L(S) + L(T)
(3) L(L(S)) = L(S)
129
Proof
(1) Let us assume ST. then we show that L(S)  L(T). Let us take any
element w  L(S). Then w = 1s1 + … +  nsn where i  F and si  S for i = 1, 2, … n.
Now S  T  si  T which implies w  L(T). Hence L(S)  L(T).
(2) Now let us show that L(ST) = L(S) + L(T). Let s be any element in L(ST).
Then s  1s1  2 s2  .......  n sn where si  S  T and  i  F without loss of
generality we can assume that s1 , s 2 , ......s m  S and s m1 , s m 2 , ......s n  T .

Hence 1s1  2 s2  ....... m sm  L( S ) and m1sm1  m2 sm2  ....... n sn  L(T )


s  1s1  2 s2  .......  m sm  m1sm1  m2 sm 2  .......  n sn  L( S )  L(T )
which shows that s  L(S) + L(T), Hence L(ST)  L(S) + L(T).
By(1) L( S )  L ( S  T ) and L(T )  L( S  T ) Hence L( S )  L(T )  L( S  T )

Hence L( S )  L(T )  L( S  T ) .
Note: If W1 and W2 be two subspaces of the vector space v, it is easy to see that W1
+ W2 = (w1 + w2 / w1  w1, w2  W2) is a subspace of V, called the sum or join
of the two sub-spaces W1 and W2. One can also check that W1 + W2 is same
as L(W1W2) and that W1 W2 is not in general a subspace of V.
(3) Now let us show L(L(S)) = L(S). Let w be any element in L(S) then w 
L(L(S)) since w = 1.w where 1  F and w  L(S). Hence L(S)  L(L(S)). Let us take
any element v  L(L(S)) then v = 1w 1+ 2w2 + …+  nwn where  i  F and wi  L(S).
Since wi  L(S) w i =  i1 si1 +  i2 si2 + ….+  ik sik
where  ij  F and sij  S for j = 1, 2, … k
v =  1( 11 s11 + … +  1k s1k) + …+ n ( 1n s1n + … +  nk snk)
= (1 11) s 11 + … + ( 1  1k) s1k + … + (n  1n) s1n …+ (n  nk) s nk
Hence L(S), i.e., L(L(S))  L(S)
Hence L(S) = L(L(S)).
Definition 3.4.5
The vector space V is said to be finite dimensional (over F) if there is a finite
subset S in V such that V = L(S).
Note F(n) is finite – dimensional over F. For if S consists of the n vectors (1,0,0,….0),
(0,1,0,….0),……. (0,0,0,….1), then V = L(S)
Definition 3.4.6
If V is a vector space and if v1 … vn are in v, we say that they are linearly
dependent over F if there exist elements 1, 2, … n in F, not all of them 0, such
that
 1 v1 +  2v2 + … + nvn = 0
130
If the vectors v1…. vn are not linearly dependent over F, they are said to be
linearly independent over F.
It is clear from the definition of linear dependance that the vectors v1, v2, … vn
of the vector space V are linearly independent over F iff any linear relation of the
form 1v1 +  2v2 +… + nvn = 0 always implies that 1 = 2 = … = n = 0.
In F (3) it is easy 0 verify that (1, 0,0) (0, 1, 0) and (0, 0, 1) are linearly
independent.
But (1, 1, 0) (3, 1,3) (5, 3, 3) are linearly dependent because
–– 2 (1, 1, 0) –1 (3, 1, 3) + 1 (5, 3, 3) = (0, 0, 0)
Lemma 3.4.7
If v1, v2, … v n  V are linearly independent then every element in their linear
span has a unique representation in the form 1v1 + … +  n vn with i  F.
Proof
By definition, every element in the linear span is of the form 1 v1 + … + nvn.
To show uniqueness we must show that if 1 v1 + … + n vn = 1 v 1 + … +  n vn then
 1 = 1 ,  2 = 2 , …, n = n
If 1v1 + .. + nv n = 1 v1 + … + n v n then ( 1 - 1) v1 + … + (n -  n) vn = 0.
Since v1, v2, … v n are linearly independent, 1- 1 = 0,………….,  n-n = 0. So  i = i
for i = 1, 2, … n.
Theorem 3.4.8
If v1, v2 … v n are in V, then either they are linearly independent or some vk is a
linear combination of the preceding ones, v1, v2, …. vk-1.
Proof
If v1, v2, … vn are linearly independent, there is nothing to prove. Therefore
suppose that 1 v1 + 2 v 2 + … + n v n = 0 where not all the  s are 0. Let k be the
largest integer for which  k  0. since  i  0 for i > k,
1v1 + 2v2 + … + k vk = 0, since k  0.

vk =  k1 (-1 v1 - 2v 2 — …— k-1 v k-1).


1 1 1
= (   k 1) v1 + (   k 1) v 2 + … + (   k k-1) vk-1

Thus vk is a linear combination of its predecessors.


Corollary 3.4.9
If v1, v2 … vn in V have W as linear span and if v1, v2, … vk are linearly
dependent, then we can find a subset of v1, … vn of the form v1, v2, …. vk, vi1 vi2 ….

vir consisting of linearly independent elements whose linear span is also W.


131
Proof
If v1, v2, … vn are linearly independent we are done. If not, weed out from this
set the first vj ,which is a linear combination of its predecessors. Since v1, v2, …. vk
are linearly independent j>k. The subset so constructed, v1, v2 … vk …v j-1, vj+1 … v n
has n-1, elements. Clearly its linear span is contained in W. But we claim that it is
actually equal to W. For given wW, it can be written as a linear combination of v1,
v2, … vn. But in this linear combination, we can replace vj by a linear combination
of v1 … vj-1. Hence w is a linear combination of v1, … vj-1, vj+1, … vn. Continuing
this weeding out process, we reach a subset v1, v2, …. vk, vi1 vi2 …. vir whose linear
span is still W but in which no element is a linear combination of the preceding
ones. By previous theorem, the elements v1, v2, …. vk, vi1 vi2 …. vir must be linearly
independent.
Corollary 3.4.10
If V is a finite dimensional vector space then it contains a finite set v1 … vn of
linearly independent elements whose linear span is V.
Proof
Since V is finite dimensional it is the linear span of a finite number of elements
u1, … um. By the previous corollary we can find a subset of these denoted by v1,
…vn consisting of linearly independent elements whose linear span must be V.
Definition 3.4.11
A subset S of a vector V is called a basis of V if S consists of linearly
independent elements and V = L(S).
If V is a finite dimensional vector space and if u1 …. um span V, then some
subset of u1, … um forms a basis of V.
Lemma 3.4.12
If v1, v2, … vn be a basis of V over F and if w1, w2 … wm in V are linearly
independent over F then m  n.
Proof
Let v1, v 2, … vn be a basis for V, the vector wm is a linear combination of v1, v2,
… vn. Hence the vector wm, v1, v2 … vn are linearly dependent and they span V.
Thus some proper subset of these wm , vi1 , vi2 , …. vik with k  n – 1 forms a basis of
V. In doing so, we have included wm and removed at least one vi in the new basis.
Repeating this process, we will get a basis with the vectors wm 1 , wm , vi1 , vi2 , …. vik .
From this linear independent set , by previous corollary we can extract a basis of
the form wm 1 , wm , v j1 , v j2 , …. v js . where s  n-2. Keeping up this procedure we get a
basis of V of the form w 2, … wm-1, wm, v, v … ; Since w1 is not a linear combination
of w2, … wm-1 the above basis must actually include some v. To get to this basis we
have introduced (m-1) w’s each such introduction having cost us atleast one v and
yet there is a v left. Thus m-1  n-1 and hence m  n.
132
Corollary 3.4.13
If V is finite dimensional over F then any two bases of V have the same number
of elements.
Proof
Let v1, v2, …. vn be one basis of V over F and let w1, w2, … wm be another basis.
In particular w1, …. wm are linearly independent over F. Hence by the previous
theorem m  n. Now interchanging the roles of v’s and w’s we get n  m. Hence n = m.
COROLLARY

F (n ) is isomorphic F (m ) if and only if n  m.


COROLLARY

If V is finite-dimensional over F then V is isomorphic to F (n ) for a unique


integer n where n is the number of elements in any basis of V over F.
Definition 3.4.14
The number of vectors in a basis (which is unique) of a vector space V is called
the dimension of the vector space. we shall write the dimension of V over F as
dim V .
COROLLARY
Any two finite- dimensional vector spaces over F of the same dimensional are
isomorphic.
Lemma 3.4.15
If a vector space V is finite dimensional over F and if u1… um of V are linearly
independent, then we can find vectors um+1 … um+r in V such that u1, u2, … um,
um+1, …., um+r is a basis of V.
Proof
Since V is finite dimensional it has a basis. Let v1, v2, … vn be a basis of V.
Since these span V, the vectors u1, um, v1… vn also span V. Hence by previous
corollary there is a subset of the form u1, … um, v i1 , vi2 , …. vir which consists of
linearly independent elements which spans V. To prove the lemma put
um+1 = v i1 , …. um+r = vir

Lemma 3.4.16
If V is finite dimensional, and if W is a subspace of V , then W is finite-
dimensional , dim W  dim V and dim V  dim V  dim W .
W
Proof
Since V is finite dimensional. Let n = dim V. Then (n+1) elements in V are
linearly dependent. In particular, any n+1 elements in W are linearly dependent.
Thus we can find a largest set of linearly independent elements in W ; w 1, w2, … wm
and m < n.
133
If wW, then w1, w2, … wm, w is a linearly dependent set. Hence w+ 1w1 + …
+ m wm = 0 and not all the i ‘s are 0. If  = 0, by the linear independence of the wi
we would get that each i = 0 , a contradiction. Thus   0 and so w = - -1 (1 w 1 +
… + m wm). Since w is any vector in w1, w2, … wm span W. So W has a basis of m
elements where m < n. Hence dim W < dim V.
Now let w1, w2,…, w m be a basis of W. We can fill this out to a basis w 1, w2, …
wm, v1, … ,vr of V where m+r = dim V and m = dim W.
Let v1 , v2 ,...v r bet the images , in V W of v1 , v 2 ,...v r . Since any vector vV is of
the form v = 1 w1 + 2 w2 + … + m wm +  1 v1 +  r vr where ’s and ’s  F, then v ,
the image of v is of the form v  1v1   2 v 2  ...   r v r ,(since w1  w2  ...  wm  0 )
Thus v1 , v 2 ,...v r span V W . We claim that they are linearly independent. For if
 1v1   2 v 2  ...   r v r  0 then 1 v 1 + … + r vr  W and so 1 v1 + … + r v r = 1w1 +
… +  m wm which by the linear independence of the set w1, … wm, v1 … vr forces 1
= … = r = 1 = … = m = 0. Hence we have shown that V W has a basis of r

elements namely v1 , v 2 ,...v r and so dim V  r  dimV  m  dimV  dimW .


W
Corollary 3.4.17
If A and B are finite dimensional subspaces of a vector space V then A + B is
finite dimensional and dim (A+B) = dim (A) + dim (B) – dim (AB).
Proof
AB A
Since we can show, as in group theory, that  by equating the
B AB
dimensions of both sides, we get that
dim (A+B) — dim (B) = dim (A) — dim (AB)
Hence dim (A+B) = dim (A) + dim (B) — dim (AB).
9.3.5 DUAL SPACES
Notation: Hom (V, W)
Let V and W be two vector spaces over field F. Hom (V, W) denotes the set of
all vector space homomorphism of V into W.
Lemma 3.5.1
Hom (V, W) is a vector space over F under suitable operations.
Proof
Let us define the operation ‘+’ in Hom (V, W) as follows. For S and T in Hom
(V, W), S+T is defined by (S+T) (v) = S(v) + T(v) for any vector v  V. Hence S+T is a
mapping from V into W. In fact S+T is a homomorphism. For, if v1, v2  V then
(S+T) (v1+v2) = S (v1+v2) + T (v1+v2) (By definition)
= S (v1) + S (v2) + T (v1) + T (v2) (Since S, T are homomorphism)
= S(v1) + T(v1) + S(v2) + T(v2)
134
= (S+T) (v1) + (S+T) (v2)
Hence S+T preserves addition in V. Moreover for F, vV
(S+T) (v) = S(v) + T(v)
= S(v) + (T(v))
(S and T are vector space homomorphism).
=  [S(v) + T(v)]
=  [(S+T)(v)]
i.e., S+T preserves scalar multiplication
Hence (S+T)  Hom (V, W)
Under the operation of ‘+’ in Hom (V, W). How (V, W) becomes an abelian
group having the ‘0’ map defined by O(v) = 0 for all vV as additive identity and for
any S Hom (V, W), its inverse being (—S) given by (-S) (V) = — [S(v)] for all v in V.
To make this Hom (V, W) into a vector space over F, let us define scalar
multiplication by setting for any   F and for any S Hom (V, W) the map S from
V into W as (S) (v) =  (S(v)). We claim S is a homomorphism. In fact for v1, v2 
V,
S (v1 + v2) =  (S(v1+v2)) =  [S(v1) + S(v2)]
=  (S(v1)) +  (S(v2))
= S(v1) +  S(v2).
Further for   F, S(v) =  (S(v))
=  (S(v))
= () S(v)
= () S(v) (Since F is a field)
=  (S) (v)
Hence S preserves both addition and scalar multiplication and so S  Hom
(V, W).
Now we will show the Hom (V, W) is a vector space.
Let ,   F; T  How (V, W) and v  V.
Then ((S+T) (v) =  [(S+T) (v)] =  [S(v) + T(v)]
=  S(v) + T(v) = (S + T) (v)
Hence (S + T) = S + T
(( + ) S) (v) = ( + ) (S (v)) = (S(v)) +  (S(v))
= (S) (v) + (S) (v)
= (S + S) (v)
Hence ( + ) S = S + S
(()) (v) = () (sv) = ((S(v))) = ((S)) (v)
135
Hence ()S = (S)
(IS) (v) = 1 (S(v)) = S(v) i.e., I.S. = S.
Hence Hom (V, W) is a vector space over F.
The following important theorem gives a definite information about the
dimension of Hom (V, W) when both V and W have finite dimensions over F.
Theorem 3.5.2
If V and W are of dimensions m and n respectively over F, then Hom (V, W) is
of dimension mn over F.
Proof
We are going to prove the theorem by actually finding a basis of Hom (V, W)
over F consisting of mn elements.
Let v1, v2, …. , vm be a basis of V over F and w1, w2, … wn be a basis of W over
F. If vV then v = 1v1 + … +  m vm, where 1, 2, … m are uniquely defined
elements of F. Now we define the map T ij : V  W by Tij(v) = i w j where i = 1, … m
and j = 1 …n. Hence we have in all mn maps from V to W, We claim that they are
all homomorphism. For if x1, x2  V where x1 =  1 v1 + … + m vm : x2 =  1 v1 + …  m
vm then T ij (x 1 + x2) = (i +  i) wj = i wj +  i wj
= T ij(x1) + Tij (x2)
Further, for k  F, T ij (kx1) = (k i) w j = k ( i wj)
= k T ij (x 1)
Hence T ij  Hom (V, W).
Our claim is that these mn elements , Tij’s form a basis for Hom (V, W) . For
that first let us show that given any S  Hom (V, W) we can write S as a linear
combination of Tij’s. Since S(v1)  W and so S(v1) can be written as a linear
combination of w1, w2, … wn. Let us write S(v1) = 11 w 1 + 12 w 2 + … + 1n w n for
some suitable 11, 12, … 1n  F. Similarly we can write S(vi) = i1 w1 + i2 w2 + …
+ in wn for i = 1 … m where ij  F. Consider S0 = 11 T11 + 12 T 12 + … + 1n T1n +
21 T21 + … + i1 Ti1 + … + in Tin + … +m1 Tm1 +…+ mn T mn. Let us compute S0(vk)
for the basis vector vk. Now S0(v k) = 11 T11 (vk) + 12 T 12 (vk) + … +m1 Tm1 (vk)+…
mn Tmn (vk). By the definition of Tij we have Tij (vk) = 0; i  k and Tij (vk) = wj if i = k.
Hence S0 (v k) = k1 w1 + k2 w2 + … + kn w n which is nothing but S(vk). Hence
the homomorphism S0 and S agree on the basis of V. Hence S0 = S. Since S0 is the
linear combination of Tij ‘s we get S is also the same linear combination of Tij.
Hence T ij’s span Hom (V, W) over F.
In order to show that these Tij’s form a basis of Hom(V,W) over F. but we have
to show that they are linearly independent over F . Suppose that  11 T11 +  12 T12 +
… +  1n T1n + … +  i1 Ti1 + … +  in Tin +…+  m1 Tm1 + … + mn Tmn = 0 (zero
homomorphism) with  ij all in F. Applying this to vk we get 0= vk ( 11 T11 +  12 T12 +
… +  1n T 1n + … +  i1 Ti1 + … +  in Tin +…+  m1 Tm1 + … +  mn Tmn)= k1w 1+ k2 w 2 + …
+  kn wn since Tij(v k) = 0 if i  k = wj if i = k.
136
Since w1 … wn are linearly independent over F ,  k1 =  k2 = … =  kn = 0.
Similarly taking the vectors v1, v2, … vn we get  ki = 0 for i = 1 … m; j = 1 … n forms
a basis of Hom (V, W) over F. Hence the dimension of Hom (V, W) = mn.
Corollary 3.5.3
2
If dim F V  m then dimF Hom(V , V )  m . (or)
Dimension of Hom (V, V) = m2 if dimension of V is m.
Proof
In the theorem put V=W, and so m=n, and hence mn=m2.
Corollary 3.5.4
If dim F V  m then dim F Hom (V , F )  m . (or)
If dimension of W over F is m then dimension of Hom (V, F) = m.
Proof
As a vector space, F of dimension 1 over F. Applying the theorem and putting
W = F, we get dimension of Hom (V, F) = m.
Definition 3.5.5
If K is a vector space its dual space is Hom (V, F) and this space is denoted by
V .
Note

Any element in V is called a linear functional on V into F.


Lemma 3.5.6
If V is finite dimensional and 0 v V then there is an element f V such that
f(v)  0.
Proof
In fact the lemma is true if V is infinite dimensional. Since V is finite
dimensional over F and let v1, v2, … vn be a basis of V ;let vi be the element V
defined by vi (vj) = 0 if i  j, vi (vi) =1 if i = j

And vi (1 v1 + 2 v2 + … + n vn) = i

In fact the vi ’s are nothing but T ij is introduced in the previous theorem.

Hence v1 , v2 , v3 ,...vn forms a basis of V .


If 0  v  V, we can find a basis of the form v1 = v, v2 … vn and so that there is
an element in V namely v1 , such that v1 (v1) = v1 (v) = 1  0. Hence the result.
Definition 3.5.7

Since Hom (V, F) = V is a vector space over F we can form. How (V , F ) which

we denote by V and call this as the second dual of V.


137
Lemma 3.5.8

If V is finite dimensional, then there  is an isomorphism of V onto V .


Proof

Given an element v  V, we can associate with it an element Tv in V by the


setting for any f  V , T v(f) = f(v). One can easily see Tv  Hom (V , F ) . Now we

define the map  V  V as follows.


For any v  V,  (v) = Tv for every v  V. We claim  is a homomorphism. For
 (v1 + v2) = Tv1 + v2. But Tv1 + v2 (f) = f(v1+v2) = f (v1) + f(v2) = T v1 (f) + Tv2 (f).

Hence T v1+v2 = Tv1 + Tv2. Therefore  (v1+v2) =  (v1) +  (v2). Similarly we can
show that for   F,  ( v) = Tv = T v = (v). Hence  is a vector space

homomorphism of V into V . Since V is finite dimensional it is enough if we show


that the kernel of  trivial.
If  (v) = 0 then Tv = 0. i.e. Tv(f) = f(v) = 0 for all f  V which implies, that v = 0.
Hence  is an isomorphism. Since V is finite dimensional, and V is also finite
dimensional. Hence  is onto.
Definition 3.5.9
If W is a subspace of V then the annihilator of W, A(W) = {f  V | f(w) = 0 for
all w  W }.
Note

We can show that A(W) is a subspace of V . In fact, for any f and g  V and ,
  F we find that  f +  g  A (W). Since for w  W; (f + g) (w) = (f) (w) + (g)
(w) = f(w) + g(w)
= 0 + 0 (since f, g  A(W))
=0+0=0
Theorem 3.5.10
If V is finite dimensional and W is a subspace of V, then W is isomorphic to
V
and dim A(w) = dim V – dim W.
A(W )
Proof

Let W be a subspace of V, where V is finite –dimensional. If f  V , let f be the


restriction of f to W. Thus f is defined on W by f (w) = f(w), for every w  W.

Since f  V clearly f  W . Let us define the map T: V  W defined by T(f) =

f for every f  V . Clearly T is a homomorphism of V in to W . For T(f+g) =


f  g  T(f )  T(g ) and T(f) = f = T(f). Now let us find the Kernel of T and if f is in
138
the Kernel of T, then the restriction of f to W must be 0. That is f (W )  0 for all
w  W . Also , conversely if f (W )  0 for all w  W then f is in the kernel of T. Hence
the Kernel of T = A(W).

We now claim that the mapping T is onto W . For that we must show that
given any element h W , then h is the restriction of some f  V so that h = f .
If w1 … wm is a basis of W, that it can be expanded to a basis of V of the form
w1 … wm, v1 … v r where r + m = dim V. Let W1 be the subspace of V spanned by v1
… vr. Thus V = W  W1. If h  W , define f  V as follows: Given any v  V which
can be written uniquely as v = w + w 1 where w  W and w1  W1 define f(v) = h(w).
Then it is clear that f is in V and that f = h. Thus h=T(f) and so T maps V on to W .
Since the kernel of T is A(W).
V
Hence T is onto So W
A(W)

Hence dim (W)  dim (V ) - dim [A(W)]

But we have already proved that dim (V ) = dim (V)


Hence dim (W) = dim (V) - dim (A(W))
So dim (A(W)) = dim (V) - dim (W)
Corollary 3.5.11
A( A(W ))  W .
Proof

Since W  V and A( A(W ))  V , we have identified V with V . Now


W  A( A(W )) , for if w  W then  (w)  Tw acts on V by Tw ( f )  f ( w) and so is 0
for all f  A(W ) . Since dim A( A(W ))  dim V  dim A(W ) . So that

dim A( A(W ))  dim V  dim A(W )  dim V  (dim V  dim W )  dim W . Since
W  A( A(W )) and they are of the same dimension, it follows that W  A( A(W )) .
We close this lesson after providing solutions to some standard problems.
Problem 1
Let V be a vector space over a field F. Let A & B be subspace of V. Then
AB B
is isomorphic to
A AB
Solution
We know that A+B is a subspace containing A (by problem)
AB
Hence is also vector space over F.
A
139
AB
An element of is of the form A+ (a+b) where a  A
A
But A+a = A if a  A.
AB
 an element of is of the form A+b
A
AB
Consider T : B  defined by
A
T(b) = A + b
clearly T is onto.
T(b1+b2) = a + (b1+b2)
= (A + b 1) + (A+b2)
= T(b1) + T(b2)
T(b) =A+b
=  (A+b)
= T(b)
Therefore T is a linear transformation
Let K be the kernel of T.
Then K = { b: bB, A+B = A}
= { b : b B, B is contained in A} ( Since bA  A+b=A)
= AB
B AB
By the homomorphism Theorem, =
AB A
9.4. REVISION POINTS
Vector space over a field, subspace, homomorphism, internel direct sum,
linear combination over F, linear span of S, Finite dimensional, Linearly dependent
over F, Basis of V, Dimension of the vector space, dual space, second dual of V,
annihilator of w.
9.5. INTEXT QUESTION
1. If T is an isomorphism of a vector space V into a vector space W (both over
the same field F), prove that T maps any basis of K onto a basis of W.
2. If V is a finite dimensional vector space and T is an isomorphism of V onto V,
prove that T must map V onto V.
3. Let Vn = [p(x)  F(x)] deg p (x) < n]. Define T by T(0 + 1x + … + n-1 x n-1) =
0 + 1 (x+1) + 2 (x+1)2 + …n-1 (x+1)n-1. Prove that T is an isomorphism of
Vn onto itself.
4. Prove that the community of addition and the axiom (4), viz 1, v = are
consequences of other axioms in the definition of a vector space V, provided
given any v  V, there exist scalars  1,  2, … n in F and vectors v1, v1, … vn
in V (not necessarily unique) such that v =  1v1,  2v2 + … + nvn.
5. If F is the field of real numbers, show that the set of real valued continuous
functions on the closed interval [0, 1] forms a vector space over F.
140
6. Prove that the intersection of two subspaces of V is a subspace of V, but the
union of two subspaces of V need not be a subspace of V.
7. If U and W are subspaces of V, prove that
U + W is a subspace of V
U + W = (UW)
8. Let V be a vector space over a field F. Let A & B be subspace of V. Then
AB B
is isomorphic to
A AB
6. SUMMARY
Vector space, subspace, homomorphism, Internel direct sum, Linear
combination over F, Linear space of S, Finite dimensional over F, Basic of V,
Dimension of the vector space, Deal space, second dual, annihilator, are explained
in detail.
7. TERMINAL EXERCISES
1. If V is finite-dimensional and W is a subspace of V such that dim V = dim
W, prove that V = W.
2. If V is finite dimensional and T is a homomorphism of V into itself which is
not onto prove that there is some v  0 in V such that T(v) = 0.
3. Let W = [0 + 1 x + … + n-1 xn-1  F(x) / 0 + 1 + … + n-1 = 0). Show
that W is a subspace of Vn and find a basis of W over F.
4. Let v1, vn be a basis of V and let w1, … wn be any n elements in V define T
on by T (V1 v1 + … + Vn vn) =  1 w1 + …+ nWn.
(a) Show that T is a homomorphism of V into itself.
(b) When is T an isomorphism?
5. If V is finite dimensional and W is a subspace of V, prove that there is a
subspace W1, of V such that V = W  W1.
(Hint: V being finite dimensional and W being a subspace of V, W is also
finite dimensional with a basis, say, [w1, w2, … wr) where r  n, the
dimension of V. Extend this basis to a basis of V, by adjoining n-r
elements, say v1, v2, …vn-r. Then the subspace of V spanned by these n-r
elements will be the required subspace W1)
6. If S, T  (V, W) and S(vi) = T(vi) for elements, vi of a basis of V, prove that
S = T.

7. If V is finite – dimensional and v1  v2 are in V prove that there is an f 


V such that f(v1)  f(v2)
8. If W1 and W2 are subspaces of V, which is finite dimensional, describe
A (W1 + W2) in terms of A (W1) and A(W2).
9. If V is finite-dimensional and W1 and W2 are subspaces of V, describe
A (W1W2) in terms of A (W1) and A(W2).
141

10. Prove that A(W ) is a subspace of V .

11. If f and g are in V such that f (v)  0 implies g (v)  0 , prove that g  f
for some   F .
9.8. SUPPLEMENTARY MATERIALS
1. Algebra by Micheal Artin, Prentice Hall of India, New Delhi, 1994.
2. Lectures in Abstract Algebra Volumes I, II and III by N. Jacobson, D. Van
Nostrand Co., New York, 1976.
9.9. ASSIGNMENT
1. If n > m, prove that there is a homomorphism of F(n) onto F(m) with a
Kernel W which is isomorphic to F(n-m).
2. If v  0  F(n) prove that there is an element T  Hom (F(n), F) such that
T(V)  0. (Hint: Apply Lemma 10).
3. If T is a homomorphism of V onto V with Kernel W prove that there is a
one-to-one correspondence between the subspaces of V and the subspace
of U which contains W. (This Problem is similar to Lemma 9 in Lesson 3)
4. Let T be a defined on F(2) by T(x1, x2) = (x1+x2, x1, + x2 where , , , 
are some fixed elements in F.
(a) Prove that T is a homorphism of F(2) into itself.
(b) Find necessary and sufficient conditions on , , ,  so that T is
an isomorphism
9.10. SUGGESTED READINGS / REFERENCE BOOK / SET BOOK
Algebra – By Michael Artin – PHI Learning P Ltd
9.11. LEARNING ACTIVITIES
Students are requested to attend the P.C.P classes and workout the problems
given the Lesson
9.12. KEYWORDS
Vecotr space, subspace, Homomorphism, Internal direct sum, Linear
combination, Linear Span, Finite dimensional Linearly dependant, Basis, Dueal
space, Second dual

142
LESSON – 10

INNER PRODUCT SPACES


10.1. INTRODUCTION
When we studied in the last lesson about vector space V over a field F, the field
F has played no important role. But in this lesson we take F to be either the field of
real numbers or the field of complex numbers. In the first case V is called a real
vector space , in the second , a complex vector space.
In this lesson, Inner product space, Length, Orthogonal, orthogonal
complement, orthonormal, Gramschmidt orthogomalization process, module, right
R.Module, Cyclic, finitely generated, minimal generasing sets, rank, are discussed
in detail with suitable examples, Lemmas, Theorems, Problems are given.
10.2. OBJECTIVE
Inner product space, Length, orthogonal, orthogonal complement orthonormal
are defined with example, some Lenmas, Theorems are given. Gram-Schimdt
orthogonalization process, module, right R. Module, Cyclic, fimitely generated,
minimal generateing sets, ranks are discussed in detail.
10.3. CONTENTS
10.3.1 Definition : Inner Product Space
10.3.2 Definition : Length
10.3.3 Definition : Orthogonal
10.3.1 DEFINITION : INNER PRODUCT SPACE
The vector space V over F is said to be a an Inner product space if there is
defined for any two vectors u, v in V an element (u, v) in F, called the inner product
of u and v, such that
1. (u, v) = (v, u ) where (v, u ) denotes the complex conjugate of (u, v).
2. (u, u )  0 and ( u, u) = 0 if and only if u = 0
3. (u + v, w) =  (u, w) +  (v, w) for any u, v, w in V and ,  in F .
Note 1
If we take the field F to be the field of real numbers the condition (1) simply
means that (u, v) = (v, u) If we take the field F to be the field of complex numbers,
then condition (1) assures that (u, u) = (u, u) which implies that (u, u) is real hence
condition (2) makes sense even if we take F as the field of complex numbers.
Note 2
Let us find the value of (u, v + w) where u, v, are in V and ,  are in F.

(u, v + w)  (v   w, u ) (by condition (1)

= α(v, u)  β(w, u) (By condition (3))

= (v, u)  (u, w)

=  (u, v) +  (u, w)
143
Example 3.1
In F(2), let us define for u = (1, 2) and v = ( 1, 2)

(u, v) = 1  1 + 1  2

Now (v, u)  (1 1  1 1 )  1 1   2  2  1 1   2  2  (u, v ) .


Hence condition (1) is satisfied.

So (u, u) = 1 1   2  2 | 1 |2  |  2 |2  0 and (u, u) = 0 iff both 1 and 2 are


zero i.e. iff u = (0, 0) the zero in F(2).
Hence the property (2) is also satisfied.
Again, if ,   F and w = (1, 2) in F(2), consider
(u + v, w) = (( 1 +  1, 2 +  2), (1, 2))

= (1 +  1)  1 + (1 +  1)  2

=  (1  1 + 2  2 ) +  ( 1  1 +  2  2 )

=  (u, w) + (v, w)
This property (3) is also verified So (u, v) as taken in this example serves as an
inner product for F(2).
Here-after let V denote an inner product space over F.
10.3.2 DEFINITION : LENGTH
If v is in V, then the length of v (or norm of v) written as v is defined by

v = (v, v)

In the example if v = (1, 2) then v = |  |


1
2
 | 2 |
2

Lemma 3.2.1
For u1, u2, v1, v2  V and 1 2,  1  2  F, we have
(1 u1 +  1v1, 2 u2 +  2v2) =  1 1 (u1, u 2) + 1  2 (u1 , v 2 ) +  12 (v1, u2) +  1 2
(v1, v2). In particular (u + v, u + v) =   (u, u) +   (u, v) +   (v, u) +   (v,
v).
Proof
(1u1 +  1v1, 2u2 +  2v2) = 1 (u1, 2 u2 +  2v2) +  1(v1,  2 u2 +  2v2)
by condition (3).

= 1 ( 2 (u1 , u 2 )   2 (u1 , v2 ))  1 ( 2 (v1 , u2 )   2 (v1 , v2 )) (by note 2)

= ( 1  2 (u1 , u 2 )  1  2 (u1 , v2 ))  1 2 (v1 , u 2 )  1  2 (v1 , v2 )


Thus the first result is proved. The particular case follows by taking 1 = 2 =
,  1=2=, u1=u2=u and v1=v2=v.
144
Note: The first result in this Lemma can be proved to be equivalent to condition (3)
in Definition 1. So it can replace the condition (3) and we have an
equivalent definition for an inner product space.
Corollary 3.2.1

u    u
Proof
2
u  (u ,u )
   (u , u ) (By lemma 1, taking  = 0 in the particular case)
2 2
 u
Hence u    u
Lemma 3.2.2

If a, b, c are real numbers such that a > 0 and a2  2b  c  0 for all real
2
numbers  then b  ac .
Proof
a2  2b  c can be written as follows:
1  b2 
a2  2b  c = (a  b )2   c  
a  a 
b
since the value is  0 for all values of , (in particular when  = - ) we have
a
b2
c- >0. Since a > 0, we get b2 < a c.
a
Theorem 3.2.1

If u , v  V then u, v   u  v

Proof
Let  denote the zero vector in V. By the property (3) of an inner product, we
find that for any v in V, (, v) = (+,v) = (, v) + (, v) whence we see that (, v) =
0. Similarly (v, ) = 0. therefore if u = , or v = , then (u, v) = 0 and u  v  0 .
Hence the result is true when u =  or v = . Now let assume that u   and v  
and for the moment let us assume (u, v) to be real.
We have 0 < (u+v, u+v)
= 2(u, u) + 2 (u, v) + (v, v) (By lemma 1)
Let a = (u, u), b = (u, v) and c = (v, v) Hence by lemma 2,
b2 < ac Hence
(u, v)2 < (u, u) (v, v)
2 2
 u v
145

Hence u, v   u  v

u
Suppose now (u, v) =  (say) is not real. By assumption (u, v)  0 so that is

meaningful. Hence by the property (3) of the inner product, we get
u  1 (u, v )
 , v   u, v   1
   (u, v )
Since we have proved the inequality for reals.

u  u
1=  ,v  .v
  
1
= .u . v

Hence   u . v i.e. | (u, v) |  u . v


10.3.3 DEFINITION : ORTHOGONAL
If u, v are in V, then u is said to be orthogonal to v if (u, v) = 0
We note that if u is orthogonal to v, then v is orthogonal to u.

For (v, u) = u, v   0  0 .


Definition 3.3.1
If W is a subspace of V, the orthogonal complement of W denoted by W┴, is
defined by W┴ = [xv / (x, w) = 0 for all w  W].
Lemma 3.3.1
W┴is a subspace of V.
Proof
To show W┴ is a subspace, we must show that if a, b  W┴ and ,   F then
(a + b)  W┴ Now, for w in W, ( a + b, w) =  (a, w) + (b, w) = 0 + 0 = 0
since a, b  W┴, (a, w) = 0 = (b, w)
We note that W  W┴ = [0]. For if w  W W┴ it must be self – orthogonal ,
then (w, w) = 0. Hence w = 0.
Definition 3.3.2
The set of vectors {vi } in V is said be an orthonormal set if

1. each vi is of length 1 (i.e. (vi , vi) = 1)


2. For i  j (vi, vj) = 0
Lemma 3.3.2
If {vi } is an orthonormal set then the vectors {vi } are linearly independent.
Further if w = 1 v1+…….. + n vn , then i = (w, vi) for i = 1, 2, …,n.
146
Proof
In order to show that the {vi } are linearly independent then suffice to show
that if 1v1 + 2v 1 + … + nvn = 0 then all ’s are zero.
So let 1v1+ 2v2+ … + nvn = 0 then for each i,
0 = (1 v1 + 2v2 + … + nvn,vi)
= 1 (v1, vi) + 2 (v 2, vi) + … + n (vn, vi)
= i (vi, vi) (since {vi } is an orthonormal set, (vi, vj) = 0 for j  i)

= i
Thus i = 0 for each i, as required.
If w = 1 v1 + … +nvn then we find (w, vi) = i, as above,
Lemma 3.3.3
If {v1, v2, … vn } is an orthonormal set in V and if w in V then
u = w – (w, v1) v1 - (w, v2) v2 - …- (w,vi)vi - … - (w, vn)vn is orthogonal to each of
v1, v 2, … vn.
Proof
Consider
(u, vi)= (w, vi) – (w, v1) (v1, vi) – (w, v2) (v2, vi) – …– (w, vi) (vi, vi) – … – (w, vn) (vn, vi)
= (w, vi) – 0 – 0 – … -(w, vi)- … - 0.
(since (v1, v2, … vn) is an orthonormal set)
= (w, vi) – (w, vi) = 0
Hence u is orthogonal to vi. Similarly u is orthogonal to all vi’s.
Theorem 3.3.1
Let V be a finite – dimensional inner product space. Then V has an
orthonormal set as a basis.
(Gram – Schmidt orthogonalization process).
Proof
Since V is of finite dimension let us assume that the dimension of V over F is n
and let v1, v2, … vn be a basis of V. Using this basis let us construct an
orthonormal set w1, w2, … wn such that each wi is of length 1 and (wi, wj) = 0 if i  j.

v1  v1 v 1 
Let w1 = . Then (w 1 , w 1 )   , 
v1  v 
 1 v1 
2
1 v1
= 2
(v 1 , v 1 )  2
1
v1 v1

Now with the help of w1 and v2 we construct w2 such that w1 =1 and


(w2, w1) = 0. So we try to find the value of  such that w 1 + v2 is orthogonal to w1.
147
(i.e. ( w1 + v2, w1) = 0
(i.e.) ( w1 + v2, w1) =  (w1, w1) + (v2, w1) =  + (v2, w1) = 0.
Hence  = - (v2, w1). So let us take u2 = – (v2 w1) w1 + v2. Then u2 is
orthogonal to w1. Since v1, v2 are linearly independent, w1 and v2 are must be
linearly independent as w1 is a multiple of v1. Hence u2  0.
u2
Now let w2 = . Then { w1, w2 } is an orthonormal set.
u2
Let us take now u3 = – (v3, w1) w1 – (v3, w2) w2 + v3.
Then (u3, w1) = - (v 3, w1) (w1, w1) – (v3, w2) (w2, w1) + (v3, w 1)
= – (v3, w1) (1) – 0 + (v3, w1)
=0
Similarly (u3, w2) = 0. Hence u3 is orthogonal to both w1 and w2 . Since v1, v2,
v3 are linearly independent, and since w1, w2 are in the linear span of v1, v2 we have
u3
w1, w2 and v3 are linearly independent. Hence u3  0. Now let w3 = . Then {
u3
w1, w2, w3 } is an orthonormal set. Similarly proceeding we can constract w1, w2,
…wi in the linear span of v1… vi such that { w1, w2… wi } — is on orthonormal set,
To construct w i+1, let us proceed as before.
Let ui+1 = - (vi+1, w1) w1 – (vi+1, w 2) w 2 – … + vi+1
(ui+1, wi) = (vi+1, w1) (w1, wi) – (vi+1, w 2) (w2, wi) – … – (vi+1 wi) (wi, wi) +… + (vi+1, wi)
= – (vi+1, wi) (wi, wi) + (vi+1, wi)
(since {w1, w2, .. wi } is an orthonormal set (wi, wi) = 0 if i  j)
= – (vi+1, wi) + (vi+1, wi) [since (w j, wi) = 1].
=0
Hence ui+1 is orthogonal to each of w2, w2, … wi. Since v1, v2, … vi+1 are
linearly independent and since w1, w2 … wi, vi+1 are linearly independent.
u i 1
Hence ui+1  0. Let wi+1 =
u i 1

Thus { w1, w2, …, wi+1 } is an orthonormal set. Hence from the given set { v1, v2,
… vn }. Since we have already shown that an orthonormal set is linearly
independent, we have thus formed the basis { w1, w2, … wn } which is orthonormal.
Theorem 3.3.2
If V is a finite dimensional inner product space and if W is a subspace of V,
then V = W  W┴ (i.e. V = W + W┴ and W W┴ = (0)). We say that V is the direct
sum of W and W┴.
Proof
Since V is finite dimensional, W is also finite dimensional and hence by the
previous theorem, we can construct an orthonormal set { w1, … wr } in W which is a
148
basis of W. Given any v in V we show that we can write v as a sum of two elements
one from W and the other from W┴. In fact, let
v0 = v – (v1 w1) w1 – (v, w2) w2 - … - (v, wr) wr
Then v0 is orthogonal to each wi. Hence v0 is orthogonal to every element in W.
Hence v0  W┴
Now we write
v = [ (v, w1) w1 + (v, w2) w2 + … + (v, wr) wr] + v0
= v1 + v0 where v1 = (v, w1) w1 + … + (v, wr) wr  W and v0  W┴. We have
already shown that W W┴ = (0).
Hence V = W  W┴.
Solved Problems
1. Let V be the set of all continuous complex valued functions on the closed
interval (a, b).
b

show that (f, g) =  f (t) g(t) dt defines an inner product in V.


a

Solution
b

i. (f , g) =  f(t) g(t) dt
a

=  f(t) g(t) dt
a

=  g(t) f(t) dt
a

= (g, f)
b

ii. f  g, h =  (f (t)  g(t)) h(t) dt


a

=  f (t) h(t)  g(t)


a

h(t) dt

b b

=   f (t ) h(t) dt    g(t) h(t) dt


a a

=  (f, h) + (g, h)
iii. f(t) f (t ) = [a0 + ib0) + (a1+ib 1) t + …] [a0-ib 0)+(a1-ib 1) t+…
= (a02 + b02) + 2(a0 a1 + b0 b 1) t + …
= F(t) say.
Where F(t) is real valued continuous function in t
149
b b

 (f, f)   f (t ) f(t) dt   F(t) dt  0


a a

Also (f, f) = 0  F(t) = 0  ai = 0, b i, = 0 i


 f(t) = 0
 f is a zero function
Hence it is an inner product.
2. Let an inner product in V over the real feld be defined as (p, q) =
1

 p(t) q(t) dt starting from the basis (1, t, t2) in V, obtain an orthonormal
0

basis.
Solution
Let [1, 2,  3] = (1, t, t2)
Choose  1 = 1 = 1
1
2 2
1  1  ( 1 ,  1 )  (1,1)   1.1 dt  1
0

( 2 ,  2 )
Let  2   2  2
1
1
1
1
Here 2 = t,  2 = 1  (2,  2) = (t, 1) =  t.1 dt  2
0

1 1
2  t - 1  t 
2 2
1 1
1 1 1 1 1 1 1
 2 2   2 . 2   (t  , t  )   (t  )(t  )dt  (t  )3  
2 2 0 2 2 2 2 
 0 12
(3, 1) (3, 1)
3  3  2
1  2
2
1 1
3 = t2 .  1 = 1,  2 = t-½
1
2 1
(3,  1) = (t2, 1) =  (t .1) dt 
0 3
1
2 1
(3,  2) = (t2, t-½) =  (t (t  ) dt
0 2
1
t4 1 t3  1 1 1
=     
4 2 3  0 4 6 12
1 1/12  1  1 1 1
 3  t 2  1  2 2
t    t  t    t  t 
3 1/12  2 2 3 6
2  1 1 1
3   3 ,  3    t 2  t    t 2  t  dt 
 6  6 180
150
Hence
1  2  3  1 2 1
, ,  i.e. (1,2, 3(t - ), 6 5 (t  t  )
1  2  3  2 6

3. In F(2) define for u  (1 ,  2 ) and v  ( 1 ,  2 ) ,


(u , v )  21 1  1  2   2 1   2  2 .
Solution:

(i) RHS : (v, u )  2 11  1 2   2 1   2  2

(v, u )  211  1 2   2 1   2  2


(v, u )  211  1 2   2 1   2  2
(v , u )  21 1  1  2   2 1   2  2  (u , v ) : LHS
(ii) (u, u )  211  1 2   2 1   2  2
2 2
 2 1  1 2   2 1   2  0 and (u, u) = 0 iff both 1 and 2 are zero i.e. iff u
= (0, 0) the zero in F (2).
(iii) ( u   v , w)  (( (1 ,  2 )   ( 1 ,  2 )), ( 1 ,  2 ))
 ((( 1 ,  2 )  ( 1 ,  2 )), ( 1 ,  2 ))
 (( 1  1 ,  2   2 ), ( 1 ,  2 ))
 2( 1  1 ) 1  ( 1  1 ) 2  ( 2   2 ) 1  ( 2   2 ) 2
  (21  1  1 2   2  1   2  2 )   ( 21  1  1 2   2  1   2  2 )
  (u , w)   (u , v)
There fore the above defines an inner product on F(2).
10.4. REVISION POINTS
Inner product space, Length, Orthogonal, Orthogonal complement, Gram
Schmidt orthogonalization process, module, right R – module, Cycli, Fimitely
generated, minimal generating set, rank.
10.5. INTEXT QUESTION
Let V be the set of all continuous complex valued functions on the closed
interval (a, b).
b

show that (f, g) =  f (t) g(t) dt defines an inner product in V.


a

10.6. SUMMARY
Studetns will get clear idea about, Inner product space, Length, orthonal,
orthogonal complement, orthonormal, Gram Schmidt. Orthogonolization process,
module – right R – Module, cyclic, fimitely generated minimal generating sets
rank.
151
10.7. TERMINAL EXERCISES
1. Let V be the set of all continuous complex valued functions on the closed
interval (a, b).
b

show that (f, g) =  f (t) g(t) dt defines an inner product in V.


a

2. Let an inner product in V over the real feld be defined as (p, q) =


1

 p(t) q(t) dt starting from the basis (1, t, t2) in V, obtain an orthonormal
0

basis.
10.8. SUPPLEMENTARY MATERIAL
1. Algebra by Micheal Artin, Prentice Hall of India, New Delhi, 1994.
2. Lectures in Abstract Algebra Volumes I, II and III by N. Jacobson, D. Van
Nostrand Co., New York, 1976.
10.9. ASSIGNMENT
1. Prove that the following defines an inner product in the respective vector
space:
(a) In F(n) define, for u = (1, … n) and v = ( 1, … n)
(u, v) = 1 1   2  2  ....... n  n
(b) In F(2), given u = (1, 2) and v = ( 1,  2) define
(u, v) = 2 1 1  1  2   2 1 2 2
(c) If F is the real field, find all 4-tuples of real numbers (a, b, c, d) such
that f u = (1, 2) v = ( 1,  2), v = ( 1,  2) in F(2) define (u, v) = a1  1 +
b2  2 + c1  2 + d2  1.
2. If dim V = n and if (w1, w2, … wm) is an orthonormal set in V prove that
there exist vectors wm+1, … Wn such that (w1, w2, … wm, wm+1 … Wn) is an
orthonormal set.
d2y
3. Let V be real function y = f(x) satisfying  9y  0
dx 2
(a) Prove that V is two-dimensional real vector space.

(b) In V define (y, z) =  yz dx Find an orthonormal basis in V.


0

10.10. SUGGESTED READINGS / REFERENCE BOOK / SET BOOK


Algebra – By MichaelArtin – PHI Learing P Ltd
10.11. LEARNING ACTIVITIES
Students are resulted to workout the problems given in the book.
10.12. KEYWORDS
Inner product space, Orthogonal, Orthonormal set, Module.

152





























018E1110
ANNAMALAI UNIVERSITY PRESS : 2021-22

You might also like