Algebra Abstract and Modern
Algebra Abstract and Modern
and Modern
Dr U. M. Swamy
Dean, Faculty of Science (Retired)
Department of Mathematics
Andhra University
Visakhapatnam
Dr A. V. S. N. Murty
Professor of Mathematics
Srinivasa Institute of Engineering and Technology
Amalapuram
This eBook may or may not include all assets that were part of the print version.
The publisher reserves the right to remove any material present in this eBook at
any time.
ISBN 9788131758922
eISBN 9789332509931
Head Office: A-8(A), Sector 62, Knowledge Boulevard, 7th Floor, NOIDA
201 309, India
Registered Office: 11 Local Shopping Centre, Panchsheel Park, New Delhi
110 017, India
and
My family members
A. V. S. N. Murty
Preface ix
Part I: Preliminaries
1. Sets and Relations 1-3
1.1 Sets and subsets 1-3
1.2 Relations and functions 1-11
1.3 Equivalence relations and partitions 1-21
1.4 The cardinality of a set 1-27
The concept of a set was used even by the ancient mankind without having
an exact idea of what it was. In modern mathematics, the notion of a set is
most basic. In fact, almost all the mathematical systems are certain collec-
tion of sets and their theories can be categorised as parts of set theory. We do
not intend to discuss axiomatic development of set theory. But, any person
with an intention of starting to learn the present day algebra must necessarily
possess certain elementary knowledge of set theory. This chapter provides a
fairly good platform to refresh with those elementary notions of sets, rela-
tions, functions and the cardinality of a set.
Example 1.1.1. Let us call a positive integer, a prime number if it has exactly
two positive divisors, namely 1 and itself. Clearly, 1 is not a prime number,
since 1 has only one positive divisor. Let C be the collection of all prime
numbers. We shall argue that C is a well-defined collection of objects. Let a
be any object. If a is not a positive integer, then we can immediately say that
a does not belong to the collection C. Suppose that a is a positive integer,
we can evaluate all the positive divisors of a and see whether these are two in
Example 1.1.2. Let C be the collection of all sets A satisfying the property
that A is not an object in A (or A does not belong to A). We shall argue that
C is not a well-defined collection. Suppose on the contrary that C is a well-
defined collection, that is, C is a set. Then, if C is an object in C, it follows
that C is not an object in C. On the other hand, if C is not an object in C,
then it follows that C is an object in C. Either way, it leads to a contradiction.
Therefore, we cannot decide whether C is an object in C. Therefore, C is not
a well-defined collection.
Example 1.1.3
1. The collection of all intelligent persons in India is not a set, since, if we
select a person from India, we cannot say with certainty whether he/
she belongs to the collection or not, as there is no standard scale for the
evaluation of intelligence.
2. For a similar reason, as detailed above, the collection of all tall persons
in India is not a set.
3. The collection of all prime numbers is a set, as discussed in Example
1.1.1.
4. The collection of all positive integers, which are not prime, is a set.
In this book, it is convenient to represent a set with the help of certain
property or properties satisfied only by the elements of the set. In order to
represent a set by this method, we write between the brackets { } a variable
x which stands for each of the set followed by the property or properties
p ossessed by each element of the set and these two are separated by a symbol
‘:’ or ‘|’, read as ‘such that’. Therefore, we write
{x : p(x)} or {x | p(x)}
to represent the set of all objects x that satisfy the statement p(x). For example,
the set of all prime numbers is represented by
{x : x is a prime number}.
{x : x is an integer and x2 1 2 5 0}
Example 1.1.5
1. Let P be the statement, ‘x is an integer and x2 5 0’ and Q be the state-
ment, ‘x 5 0’. Then, we have P ⇔ Q since, for any integer x, x2 5 0 if
and only if x 5 0.
2. Let P be the statement, ‘x is a real number and x2 5 x’ and Q be the state-
ment, ‘x 5 0 or x 5 1’. Then, P ⇔ Q since, for any real number x, x2 5
x if and only x 5 0 or x 5 1.
x∈A⇒x∈B and x ∈ B ⇒ x ∈ A.
Definition 1.1.4. For any set S, the collection of all subsets of S is again a set
and is called the power set of S and is denoted by P(S).
Note that the power set P(S) of any set S is always nonempty, since the
empty set is a subset of every set S. In fact, if S is the empty set , then
P() 5 {},
a set consisting of only one element. It can be easily proved that, for any non-
negative integer n, a set S has exactly n elements if and only if the power set
P(S) has exactly 2n elements.
Definition 1.1.5. A set whose element are sets is called a class of sets or
f amily of sets.
Class of sets will be usually denoted by script letters, such as !, @, #, etc.
For any set S, the power set P(S) is a class of sets. A class # of sets is called
an indexed class if there exists a set I such that, for each i ∈ I, there is a unique
member Ai in # associated with i and the class # is equal to the class of all Ai,
i ∈ I; in this case, we write
# 5 {Ai : i ∈ I} or # 5 {Ai}i∈I
1
An x : x is a real number and 0 ≤ x ≤ .
n
Then, {An }n∈Z is an indexed class of sets and the set Z1 of positive integers
is the index set.
Definition 1.1.6. For any indexed class of sets {Ai}i∈I , we define the set as
∩ A {a : a ∈ A
i∈ I
i i for all i ∈ I }.
This set is called the set intersection of Ai’s, i ∈ I. In particular, if A1, A2, …, An
are sets, we define
n
∩ A {a : a ∈ A
i 1
i i for i 1, 2, … , n}
A ∩ B 5 {x : x ∈ A and x ∈ B}.
Two sets A and B are said to be disjoint if A ∩ B 5 , that is, there are no
common elements of A and B.
Definition 1.1.7. For any indexed class {Ai}i∈I of sets, we define the set as
∪ A {a : a ∈ A
i ∈I
i i for some i ∈ I }.
This set is called the set union of Ai’s, i ∈ I. In particular, for any sets A1, A2,
…, An, we define
n
∪ A {a : a ∈ A
iI
i i for some 1 ≤ i ≤ n}
and this is also denoted by A1 ∪ A2 ∪ … ∪ An. For any sets A and B, we have
A ∪ B 5 {x : x ∈ A or x ∈ B}.
1
An x : x is a real number and 0 ≤ x ≤ .
n
[0, 1] 5 A1 ⊃ A2 ⊃ A3 ⊃ … ⊃ An ⊃ An11 ⊃ …
1
∩ +
An {x : x is a real number and 0 x
n
for all n ∈ Z + } {0}
n∈Z
Theorem 1.1.1. The following holds good for any sets A, B and C.
1. A ∪ B ⊆ C ⇔ A ⊆ C and B ⊆ C
2. A ⊆ B ∩ C ⇔ A ⊆ B and A ⊆ C
3. A∩B⊆A⊆A∪B
4. A∪A5A5A∩A
5. A ∪ B 5 B ∪ A and A ∩ B 5 B ∩ A
6. (A ∪ B) ∪ C 5 A ∪ (B ∪ C) and (A ∩ B) ∩ C 5 A ∩ (B ∩ C)
7. A5A∩B⇔A⊆B⇔A∪B5B
8. A ∩ (A ∪ B) 5 A 5 A ∪ (A ∩ B)
9. A ∩ (B ∪ C) 5 (A ∩ B) ∪ (A ∩ C)
10. A ∪ (B ∩ C) 5 (A ∪ B) ∩ (A ∪ C)
11. A ∩ (i∪ A ) i∪ (A ∩ Ai ) for any indexed class {Ai}i∈I of sets.
∈I i ∈I
12. A ∪ ( ∩ Ai ) ∩ (A ∪ Ai ) for any indexed class {Ai}i∈I of sets.
i∈I i∈I
13. A ⊆ B ⇒ A ∩ C ⊆ B ∩ C and A ∪ C ⊆ B ∪ C
14. A ∩ B ⊆ A ∩ C and A ∪ B ⊆ A ∪ C ⇔ B ⊆ C
Definition 1.1.8. For any two sets A and B, the difference of A with B is
defined as
A 2 B 5 {x : x ∈ A and x ∉ B}.
Theorem 1.1.2 (De Morgan Laws). For any indexed class {Bi}i∈I of sets and
for any sets A, B and C, the following holds good.
1. A ( i ∪ B ) i∩ (A Bi )
∈I i ∈I
2. A ( i ∩ B ) i∪ (A Bi )
∈I i ∈I
3. B ⊆ C ⇒ A 2 C ⊆ A 2 B and B 2 A ⊆ C 2 A
4. ( ∪ Bi ) A ∪ ( Bi A)
i∈I i∈I
5. ( i ∩ B ) A i∩ ( Bi A)
∈I i ∈I
6. (A ∪ B) 2 C 5 (A 2 C) ∪ (B 2 C)
7. (A ∩ B) 2 C 5 (A 2 C) ∩ (B 2 C)
8. A 2 (B ∪ C) 5 (A 2 B) ∩ (A 2 C)
9. A 2 (B ∩ C) 5 (A 2 B) ∪ (A 2 C)
10. (A 2 B) 2 C 5 A 2 (B ∪ C) 5 (A 2 C) 2 B
11. A 2 (B 2 C) 5 (A 2 B) ∪ (A ∩ C)
12. A∩B5⇔A⊆A2B⇔B⊆B2A
13. A25A
14. 2A5
Definition 1.1.9. For any sets A and B, the symmetric difference of A and B
is defined as
A ⊕ B 5 (A 2 B) ∪ (B 2 A).
Theorem 1.1.3. The following holds good for any sets A, B and C.
1. A ⊕ B 5 B ⊕ A
2. (A ⊕ B) ⊕ C 5 A ⊕ (B 1 C)
5 (A ∩ B ∩ C) ∪ ((A 2 B)2C) ∪ ((B 2 C)2A)
∪ ((C 2 A)2B)
3. A ⊕ 5 A
4. A ⊕ A 5
A ∩ (B ⊕ C) 5 (A ∩ B) ⊕ (A ∩ C).
EXERCISE 1(a)
1. Express each of the following sets in the form {x : P(x)} and specify the prop-
erty P(x).
(i) The set of all rational numbers, whose denominators are not divisible by 5.
(ii) The set of all integer multiples of 5 in between 296 and 96.
(iii) The set of all points in the three-dimensional Euclidean space, whose
distance from (0, 0) is a rational number.
(iv) The set of all pairs of real numbers, whose sum of their squares is nonzero.
(v) The set of all even primes.
(vi) The set of all subsets of {1, 2, 3, 4} not containing 3.
5. Let X 5 {a ∈ R : 21 a 1} and
Y 5 {r ∈ R : r 5 sin t 2 cos t for some t ∈ R}.
Is X 5 Y ?
6. Let X 5 {1, 2, 3, …, 100}, A 5 {a ∈ X : a 5 b2, b ∈ Z},
B 5 {a ∈ X : a is odd} and for each 1 i 96,
Ci 5 {i, i 1 1, i 1 2, i 1 3, i 1 4}. Write explicitly all elements in each of the
following sets.
(i) A∩B
(ii) A ∪ B ∪ C2
96
(iii) ( ∪ Ci ) ∩ A
i1
25
(iv) B ∩ ( ∪ Ci )
i20
(v) X 2 (A ∪ B)
90
(vi) X ( ∪ Ci )
i6
96
(vii) A ( ∪ Ci )
i1
(viii) A2B
7. For any two sets A and B, prove that
A 5 A ∩ B ⇔ A ⊆ B ⇔ A ∪ B 5 B.
12. Prove or disprove each of the following for any sets X and Y.
(i) P(X ∩ Y) 5 P(X) ∩ P(Y)
(ii) P(X ∪ Y) 5 P(X) ∪ P(Y)
(iii) P(X 2 Y) 5 P(X) 2 P(Y)
(iv) P(X) 5 P(Y) ⇔ X 5 Y
Definition 1.2.1. A pair of elements (not necessarily in the same set) writ-
ten in a particular order is called an ordered pair and is written by listing its
elements in a particular order, separated by a comma, and enclosing the pair
in brackets. In the ordered pair (x, L), x is called the first component (or first
coordinate) and L is called the second component (or second coordinate).
The ordered pairs (x, L) and (L, x) are different even though they consist of
the same pair of elements. For example, the pairs (2, 5) and (5, 2) represent
two different points in the plane.
Definition 1.2.2. Let A and B be any two sets. Then, the set of all ordered
pairs (a, b) with a ∈ A and b ∈ B is called the Cartesian product of A and B
and is denoted by A 3 B. That is,
Definition 1.2.3. For any sets A1, A2, …, An, we define the Cartesian product
of A1, A2, …, An as the set
In particular, for any set A and for any positive integer n, we define
Definition 1.2.4. Let A and B be any sets. Then, any subset of A 3 B is called
a relation from A to B. For any relation R from A to B (that is, R ⊆ A 3 B), if
(a, b) ∈ R, then we say that ‘a is R-related to b’ or ‘a is related to b with respect
to R’ or ‘a and b have relation with R’ and is usually denoted by a R b.
Example 1.2.2. Let Z be the set of all integers and n a positive integer.
Define
R 5 {(a, b) ∈ Z 3 Z : n divides a 2 b}
S 5 {(a, b) ∈ Z 3 Z : a 5 nb}.
Definition 1.2.7. For any relation R from a set A to a set B, the inverse of R
is defined by
Dom(g o f ) 5 Dom(f )
and Codom(g o f ) 5 Codom(g).
Two functions f and g are said to be equal if their domains are equal and f (a) 5
g(a) for all the elements a in the common domain. For two functions f and g,
both f o g and g o f may be defined but still they may not be equal, consider
the following example.
f o IA 5 f 5 IB o f.
In this case, g is unique and is called the inverse of f and is denoted by f −1.
Note that, for any a ∈ A and b ∈ B,
f (a) 5 b ⇔ a 5 f −1(b)
Theorem 1.2.4. The following are equivalent to each other for any function
f:X→Y
1. f is an injection.
2. A 5 f −1(f (A)) for any A ⊆ X.
3. f (A1 ∩ A2) 5 f (A1) ∩ f (A2).
EXERCISE 1(b)
is a bijection.
24. Let A be an n-element set and B be an m-element set. Find the number of injec-
tions of A into B in each of the following cases.
(i) n 5 m, (ii) n . m, (iii) n , m.
25. Define f : Z1 → Z1 by f (a) 5 2a 2 1. Prove that there exist infinitely many func-
tions g : Z1 → Z1, such that g o f 5 IZ1 and there is no function h : Z1 → Z1
such that f o g 5 IZ1.
26. Prove Theorem 1.2.4.
27. Prove Theorem 1.2.5.
Theorem 1.3.2. For any equivalence relation R on a set X, the class of all
R-equivalence classes in X is a partition of X and is denoted by X/R; that is,
X {R(x ) : x X }.
R
X {R(x ) : x X }.
R is called the partition on X induced by R or the quotient of X by R.
The converse of the above result is also true, in the sense that, for any
partition # of X, there exists an equivalence relation R# on X such that the
partition of X induced by R# is precisely equal to the given partition #.
j(X) . Part(X),
Example 1.3.2. Consider the relation R given in Example 1.3.1 (4), we have
R 5 {(a, b) ∈ Z 3 Z : a 5 0 5 b or ab . 0}.
Rn(a) 5 {a 1 nx : x ∈ Z}.
Theorem 1.3.5. Let R and S be two equivalence relations on a set X and X R {R(x ) : x X }
and X S be partitions corresponding to R and S, respectively. Then, R ⊆ S if
and only if X S is a refinement of X R.{R(x ) : x X }.
Now, since x ∈ S(x), we get that x ∈ R(z) for some z ∈ Z and hence (x, z) ∈ R.
Since (x, y) ∈ R also, we have that (y, z) ∈ R so that y ∈ R(z) ⊆ S(x). Therefore,
(x, y) ∈ S. Thus, R ⊆ S. b
The following theorem is a simple verification.
X
R
X
S
X
R ∩S
1. R o S is an equivalence relation on X.
2. R o S is symmetric.
3. R o S is transitive.
4. RoS⊆SoR
5. SoR⊆RoS
6. RoS5SoR
7. S o R is symmetric.
8. S o R is transitive.
f5goh
Also, define
g : X/R → Y by g(R(x)) 5 f (x). If R(x) 5 R(x9), then (x, x9) ∈ R and hence
f (x) 5 f (x9). Therefore, g is a well-defined function and clearly g is an
injection. Also, it is clear that g(h(x)) 5 f (x) for all x ∈ X. Thus, f 5 g o h,
g is an injection and h is a surjection. b
EXERCISE 1(c)
12. Let R be a binary relation on a nonempty set X. Then, prove that R is an equiva-
lence relation on X if and only if R is reflexive on X and
(a, b) ∈ R and (b, c) ∈ R ⇒ (c, a) ∈ R.
13. Describe the equivalence relations on Z corresponding to the following parti-
tions of Z:
(i) {…, – 5, – 1, 3, 7, …}, {…, – 6, – 2, 2, 6, …},
{…, – 7, – 3, 1, 5, …}, {…, – 8, – 4, 0, 4, …}
(ii) {2n : n ∈ Z}, {2n 1 1: n ∈ Z}
(iii) Z2, {0}, Z1
(iv) {…, – 3, 0, 3, 6, …}, {…, – 2, 1, 4, 7, …}, {…, – 1, 2, 5, 8, …}.
Definition 1.4.1. For any set X, let |X| denote the class of all sets that are
equivalent to X (that is, bijective with X). Then, |X| is called the cardinality of
X or the cardinal number of X or, simply, a cardinal number.
If we define, for any two sets A and B, A . B whenever there is a bijec-
tion of A onto B, then . is actually an equivalence relation on the class of
all sets. The following is a direct consequence of the discussion made after
Definition 1.2.3.
Theorem 1.4.1. Let A, B and C be any sets. Then, the following holds good.
1. |A| 5 |B| ⇔ A . B ⇔ A ∈ |B| ⇔ B ∈ |A|
2. A ∈ |B| and B ∈ |C| ⇒ A ∈ |C|.
Definition 1.4.2. For any nonnegative integer n, let In be the set of positive
integers less than or equal to n. That is,
In 5 {1, 2, 3, …, n}.
Theorem 1.4.2. The following are equivalent to each other for any nonnegative
integers n and m.
1. |In| 5 |Im|
2. In . Im
3. n 5 m
In view of the above theorem, we denote the cardinality of In by simply n.
Note that, for any set A, |A| 5 n if and only if there is a bijection of A onto the
set {1, 2, …, n} and, for this reason, we say that A has n-elements or A is an
n-element set if |A| 5 n.
Definition 1.4.4. A cardinal number is said to be f inite if any (and hence all)
of its members are finite sets.
Example 1.4.1. The set Z1 of positive integers is an infinite set, for we can
easily check that there cannot be a bijection of Z1 onto In for any nonnegative
integer n. If f : In → Z1 is a function, we can choose m ∈ Z1 such that f (a) ,
m for all a ∈ In.
Theorem 1.4.3. Let n be a nonnegative integer and X be a set, such that |X| 5 n.
Then, for any subset Y of X, |Y| 5 m for some 0 m n.
Definition 1.4.5. Let a and b be two cardinal numbers and X and Y be sets,
such that |X| 5 a and |Y| 5 b. Then, we define a is less than or equal to b (and
express this by a b) if there is an injection of X into Y.
First of all, we have to prove that is a well-defined relation on the cardi-
nals, in the sense of the following.
Theorem 1.4.4. Let X, Y, A and B be sets, such that |X| 5 |A| and |Y| 5 |B|. Then,
there is an injection of X into Y if and only if there is an injection of A into B.
Proof: Since |X| 5 |A| and |Y| 5 |B|, there are bijections f : X → A and g : Y
→ B. If h : X → Y is an injection, then g o h o f 21 is an injection of A into B.
On the other hand, if p : A → B is an injection, then g21 o p o f is an injection
of X into Y. b
Thus, is a well-defined binary operation on the set of cardinals. Since
A A for any set A, it follows that is reflexive on the set of cardinals.
Also, since the composition of injections is again an injection, we have that
is a transitive relation. In addition to the reflexivity and transitivity of the
relation , we have another important property, namely the anti-symmetricity;
that is, a b and b a are possible only if a 5 b. The proof of this is not
that straight forward and requires a skilled proof.
X1 5 X and Z1 5 Z
X 5 X1 ⊇ Z1 ⊇ X2 ⊇ Z2 ⊇ X3 ⊇ Z3 ⊇ X4 ⊇ …
Define p : X → Z by
Theorem 1.4.6. Let X and Y be any nonempty sets. Then, there is an injection
of X into Y if and only if there is a surjection of Y onto X.
x if f (x) y
g ( y ) .
x0 otherwise (that is, y ∉ f (X ))
Since f is an injection, for each y ∈ f (X), there exists unique x ∈ X such that
f (x) 5 y. Therefore, g is a well-defined function of Y into X. Also, for any x ∈
X, f (x) ∈ Y and g(f (x)) 5 x and hence g is a surjection of Y onto X.
Conversely suppose there is a surjection g : Y → X. For each x ∈ X, consider
the set
Definition 1.4.6. If a is the cardinal of a set A, then the cardinal of the power
set P(A) is denoted by 2a, for the simple reason that P(A) . {0,1}A under the
Proof: Let a be a cardinal and A be a set such that |A| 5 a. Since the cardinal of
the empty set Ø is 0 and P(Ø) 5 {Ø} which is a nonempty set, we get that |Ø| 5
0 , 1 5 20 5 |P(Ø)| therefore, we can suppose that A is a nonempty set. Define
a 5 |A| # |P(A)| 5 2a
Now, we prove that |A| |P(A)| or, equivalently, |P(A)| |A|. By Theorem
1.4.6, it is enough if we can prove that there is no surjection of A onto P(A).
Suppose, if possible, that there is a surjection g : A → P(A). Then, for each a
∈ A, g(a) is a subset of A and every subset of A is of the form g(a) for some a
∈ A (since g is a surjection). Now, consider the set B defined by
B 5 {a ∈ A : a ∉ g(a)}.
Definition 1.4.7. Let X be any set and Z1 be the set of positive integers. Then,
X is said to be a countable set and |X| said to be a countable cardinal if |X| 5
|Z1|; that is, if X is equipotent with Z1 and if f : Z1 → X is bijection, then X can
be expressed as X 5 {f (1), f (2), …, f (n), …} or, simply X 5 {x1, x2, …}.
If X is not a countable set, then X is called an uncountable set and |X| is called
an uncountable cardinal.
Theorem 1.4.8. The following are equivalent to each other for any non-
empty set X:
1. There is an injection f : X → Z1
2. X is at most countable.
3. X is a subset of a countable set.
4. There is a surjection g : Z1 → X.
Proof: (1) ⇒ (2): Suppose that there is an injection f : X → Z1, put Y 5 f (X).
Then, X . Y ⊆ Z1. Suppose that X is not finite, then Y is an infinite subset of
Z1. Define g : Z1 → Y as follows.
Let g(1) be the least element in Y (use the well-ordering principle in Z1).
Having defined g(1), …, g(n 2 1), let g(n) be the least element in Y – {g(1),
g(2), …, g(n – 1)}, for any n . 1. Since Y is infinite, Y – {g(1), …, g(n – 1)}
for any n . 1 and hence g is welldefined. Now, we have
f (n) 5 0 xn xn xn …, where xn 5 0 or 1.
1 2 3 i
y 5 0 y1y2y3…
Then, since yn ? xn for each n, we get that y ? f (n) for all n and hence y ∉ X,
n
which is a contradiction. Thus, X is uncountable and thus so is R. b
For any cardinal number a, we have proved that a , 2a (Theorem 1.4.7)
Z Z
and, in particular Z�
2 which automatically implies that 2 is an
uncountable cardinal. In the following, we prove that the cardinal number of
Z�
the set R of real numbers is precisely 2 . The cardinal number of Z1 will
be usually denoted by N0 and that of R by c.
Theorem 1.4.9
c 2N0
Proof: Recall that P(Z ) . 2Z , the set of all mapping of Z1 into the two-
element set {1, 0}. We prove that there is a bijection of 2 onto R. Define
f : 2Z → R by
f (g) 5 0 g(1) g(2)… for any g ∈ 2 ,
where 0 g(1) g(2)… is the real number in the interval [0, 1) whose decimal
places are g(1), g(2), …. Then, clearly f is an injection. On the other hand,
noted that any real number x can be represented in the binary scale in the
form
x 5 …x7x5x3x1 x2x4x6x8….
+
where each xn 5 0 or 1 for every n ∈ Z1. Now, define h : R → 2Z by
Z
2N0 .
c |R| 2Z� 2 b
∞
Theorem 1.4.10. Let X1, X2, … be finite sets. Then, ∪ X n is at most countable.
n1
∞
Proof: Let X ∪ X n . Without loss of generality, we can assume that each
n1
Xn is nonempty. Let |Xn| 5 mn and X n {xn1 , xn2 , …, xnmn }. Then,
Proof: Let {X1, X2, …, Xn, …} be a countable class of countable sets. Then,
∞
for each n ∈ Z1, there exists a bijection f n : Z1 → Xn. Now, let X n∪ X n and
1
define f : Z1 3 Z1 → X by
f (n, m) 5 f n(m).
Then, clearly f is a surjection. Also, since Z1 3 Z1 is countable, there is a bijec-
tion g : Z1 → Z1 Z1. Now, f o g is a surjection of Z1 onto X. Therefore, X is
at most countable. But, since X is infinite, it follows that X is countable. b
Note that countable product of at most countable sets may not be count-
able. For consider the following.
Example 1.4.2. The set 2 5 {0, 1} is a finite set and Z1 is a countable set.
Here, 2Z1 ( R) is uncountable.
EXERCISE 1(d)
1. Prove that the cardinal numbers of Z1, Z2, Z, Q1 and Q are all equal to each
other.
5. Prove that a set X is infinite if and only if |X| 5 |Y| for some proper subset Y of X.
7. If X is a set such that |X| 5 |P(P(X))|, then prove that there exists a surjection
f : X → P(X).
8. Deduce from Exercise 7 above that |X| , |P(P(X))| for any set X.
10. For any sets X and Y, prove that |X| 5 |Y| if and only if |P(X)| 5 |P(Y)|.
12. Give an example of a set # of circles in the plane such that every circle with
positive radius properly contains a member of #.
13. Prove that |(a, b)| 5 |(c, d)| for any intervals (a, b) and (c, d) in R with a , b and
c , d.
14. Let In 5 {1, 2, …, n} for any n ∈ Z1. Prove that |Z1| 5 |Z1 – In| for any n ∈ Z1.
15. Let X be a countable set and PF(X) be the set of all finite subsets of X. Then,
prove that
∞
∪X
n1
n
X PF ( X ) .
16. Prove that the set of polynomials in the indeterminate x over the set of rational
numbers is countable.
such that f (a) 5 0. Prove that the set of algebraic numbers is countable.
18. A real number is said to be transcendental if it is not algebraic. Prove that the set
of transcendental numbers is uncountable.
20. Prove that the set of complex numbers, whose real and imaginary parts are ratio-
nal numbers, is countable.
This chapter is meant to review some of the important properties of the set of
positive integers, the set of integers, the set of rational numbers, the set of real
numbers and the set of complex numbers. We do not discuss any axiomatic
development of these systems. We simply assume familiarity with addition
and multiplication of these and their usual properties. Also, we briefly dis-
cuss the concept of a partial order on a set in general and the usual ordering
on the real number system, in particular, these facilitate us in facing several
encounters with these throughout this book. Further, we recall the notion of
a matrix over the number systems and some important elementary properties
of the matrices and their determinants.
2.1 INTEGERS
In this section, we review certain important elementary properties of integers,
by assuming familiarity with the addition, subtraction, multiplication and the
usual ordering in these (that is, m # n if and only if n 2 m is nonnegative).
As mentioned in the beginning of the book, we follow the notations given
below.
Z : The set of integers {…, 22, 21, 0, 1, 2, …}
Z1 : The set of positive integers {1, 2, 3, …}
0 , … , an , an21 , … , a2 , a1 , a
b # n [ S ⇒ n 1 1 [ S.
b # m0 and m0 S.
n( n1)
Example 2.1.1. Let us prove that 1 1 2 1 … 1 n 5 for all n [ Z1.
n( n1) 2
Let P(n) be the statement ‘1 2 n ’ and S 5 {n [ Z1 : P(n) is
2
true}. Then, clearly 1 [ S # Z .
1
n [ S ⇒ P(n) is true
n( n1)
⇒1121…1n5
2
n( n 1) (n 1)(n 2)
⇒ 1 2 n n 1 n 1
2 2
⇒ P(n 1 1) is true
⇒ n 1 1 [ S.
for all n $ 7. To prove this, let us apply the first principle of induction (Theorem
2.1.3). We have 7 [ S and
7 # n [ S ⇒ 3n , n!
⇒ 3n11 5 3n ? 3 , n!(n11)
⇒ 3n11 , (n11)!
⇒ n 1 1 [ S.
S 5 {n [ Z1 : 7 # n}.
The first principle of induction does not work sometimes when we need to
know the truth of one or more smaller cases and not necessarily the immedi-
ately preceding one. To handle situations like this, we need another form of
induction given below.
b, b 1 1, …, n [ S ⇒ n 1 1 [ S.
b, b 1 1, …, m0 2 1 [ S.
m [ S for all m , n ⇒ n [ S.
Then, S 5 Z1.
Proof: The proof of Theorem 2.1.3 is precisely the proof of (1) ⇒ (2).
(2) ⇒ (3): Assume that the first principle of induction holds.
Let b [ S # Z1 and, for any n $ b,
b, b 1 1, …, n [ S ⇒ n 1 1 [ S.
b # n [ A ⇒ n 1 1 [ A.
From the induction hypothesis, a and b [ S and hence a and b can be expressed
as products of primes and therefore so is n 1 1. Thus, n 1 1 [ S. By the second
principle of induction, it follows that m [ S for all m $ 2. Thus, any n . 1 can
be expressed as a product of primes.
We prove the uniqueness of the factorization also by using induction prin-
ciple. Let
p1 p2 … pr 5 n 5 q1 q2 … qs
where pi’s and qj’s are prime numbers. Suppose that pi 5 qj for some i and j.
We can suppose, by renumbering of pi’s and qj’s, that p1 5 q1. Then, p2 p3 …
pr 5 q2 q3 … qs , n and hence, by the induction hypothesis, r 5 s and each
pi is equal to some qj and vice versa. Next suppose that pi qj for all i and j.
Without loss of generality, we can suppose that p1 . q1. Then,
n . (p1 2 q1) p2 … pr 5 p1 p2 … pr 2 q1 p2 … pr
5 q1 q2 … qs 2 q1 p2 … pr
5 q1 (q2 … qs 2 p2 … pr).
Theorem 2.1.7 (The Division Algorithm in Z). Let a and b be any integers
and b . 0. Then, there exist unique integers q and r such that
a 5 bq 1 r and 0 # r , b.
(Here q is called the quotient and r is called the remainder of a modulo b.)
S 5 Z1 ∩ {a 2 bx : x [ Z}.
a 5 bx 1 (a 2 bx) 5 bx 1 m0
Definition 2.1.1. For any m and n [ Z1, let CD(m, n) be the set of all com-
mon divisors (factors) of m and n in Z1. That is,
a divides b ⇒ a # b
and hence every member of CD(m, n) is less than or equal to both m and
n. This implies that CD(m, n) is finite and has a largest (greatest) member,
which is called the greatest common divisor of m and n and is denoted by
g.c.d.{m, n} or, simply (m, n). The following is an interesting property of the
g.c.d.’s.
Theorem 2.1.8. Let m and n be positive integers. Then, the following are
equivalent to each other for any d [ CD(m, n).
1. d 5 g.c.d.{m, n}
2. d 5 ma 1 nb for some a and b [ Z
3. Every member of CD(m, n) divides d.
Definition 2.1.2. Two positive integers m and n are said to be relatively prime
(or, prime to each other) if (m, n) 5 1. This is equivalent to saying that CD(m,
n) 5 {1}.
Note that m and n are relatively prime if and only if there is no prime
number dividing both m and n. The following is an important consequence
of Theorem 2.1.8.
Proof: Suppose that p does not divide m. Then, (p, m) 5 1 and therefore, by
Theorem 2.1.9, p divides n.
1! 5 1
and s(n 1 1) 5 f (n, s(n)), since (n, s(n)) [ T and hence (n 1 1, f (n, s(n))) [ T.
Thus, s is a sequence satisfying the required properties.
Next, we prove the uniqueness of s. Let s and t be sequences such that
s(1) 5 x1 5 t(1)
and s(n 1 1) 5 f (n, s(n)) and t(n 1 1) 5 f (n, t(n))
| a |
{
a if a0
a if a0
.
Note that |a| $ 0 for all a [ Z. The following can be proved by straight
forward verification.
(i) |a| 5 0 ⇔ a 5 0
(ii) |ab| 5 |a||b|
(iii) |a 2 b| 5 |b 2 a|
(iv) |a 1 b| |a| 1 |b|
(v) |a 1 b| 5 |a| 1 |b| if and only if either both a and b are nonpositive or
nonnegative.
(vi) |a 2 b| |a 2 c| 1 |c 2 b|
(vii) ||a| 2 |b|| |a 2 b|
(viii) ||a| 2 |b|| 5 |a 2 b| if and only if |a 1 b| 5 |a| 1 |b|.
EXERCISE 2(a)
( )
2
n(n1)
(ii) 13 1 23 1 33 1 … 1 n3 5
2
(iii) 1?1! 1 2?2! 1 3?3! 1…1 n?n! 5 (n 1 1)! 2 1
(iv) If X is a set and |X| 5 n, then |P(X)| 5 2n
(v) xn 2 1 5 (x 2 1)(xn21 1 xn22 1 … 1 x 1 1)
n n
(vi) (x 1 y)n 5 x nr y r
r
r0
n
(vii) 4 nn 3 2n 3
r1 3 3n
n
(viii) 1 n
r1 r (r1) n1
n
n(n1)(n2)
(ix) r (r 1)
r1 3
(x) 1 1 3 1 5 1 … 1 (2n 2 1) 5 n2.
2. Find all n [ Z1 for which 2n 1 1 , 2n21.
3. Let X be a set such that |X| 5 n $ 2. Prove that these are exactly n(n1) subsets
2
each with exactly two elements.
4. For any n and r [ Z1 such that 1 r n, prove that
n i n 1
∑ r r 1
ir
5. Use the Binomial theorem given in Exercise 1 (vi) above to prove the
following.
n n
(i) 3 n for all n ∈ Z
r0 r
(ii) (a 1 b)n [ aZ1 1 bn for all a, b and n [ Z1
n n
(iii) =
r even
r r odd r
0 r n 0 r n
where p1, p2, …, pr are distinct primes and n1, n2, …, nr are nonnegative
integers.
10. Let a p1n1 p2n2 ... prnr , where pi’s are distinct primes and ni’s are nonnegative inte-
gers and let b [ Z1. Then, prove that b divides a if and only if
b p1m1 p2m2 ... prmr , where mi [ Z and 0 mi ni.
r r
11. Let a pini and b pimi , where pi’s are distinct primes and ni’s are non-
i1 i1
negative integers. Then, prove that
r
ki
g.c.d.{a, b} 5 pi , where k 5 minimum of n and m
i1 i i i
r
and l.c.m.{a, b} pidi , where di 5 maximum of ni and mi.
i1
12. For any positive integers a and b, prove that the product of a and b is equal to the
product of their g.c.d. and l.c.m.
13. Let a, b and c be positive integers such that a divides both b and c. Then, prove
that a divides mb 1 nc for any integers m and n.
14. Let 1 , n [ Z1. Prove that either n is a prime or has a prime divisor which
is n .
15. Let a1, a2, …, ar [ Z1 and a 5 g.c.d.{a1, a2, …, ar}. Then, prove the following:
(i) d 5 b a 1 b a 1…1b a for some b , b , …, b [ Z
1 1 2 2 r r 1 2 r
A 1 B 5 {a 1 b : a [ A and b [ B}
nZ, 1 1 nZ, …, (n 2 1) 1 nZ
form a partition of Z.
a 5 nq 1 r
a [ r 1 nZ for some r 5 0, 1, …, n 2 1.
Proof: For any a and b [ Z, a n b if and only if a and b belong to the same
set in the partition. Therefore,
a b (mod n) for a n b.
That is, for any integers a and b and for any positive integer n,
Theorem 2.2.3. Let 1, n [ Z. Then, the following holds for any a, b, c and
d [ Z.
(i) a b (mod n) and c d (mod n) ⇒ a 1 c b 1 d (mod n)
(ii) a b (mod n) ⇒ 2a 2b (mod n)
(iii) a b (mod n) ⇒ ac bc (mod n)
(iv) a b (mod n) and c d (mod n) ⇒ ac bd (mod n)
(v) a b (mod n) ⇒ am bm (mod n) for all m [ Z1
(vi) ac bc (mod n) and (c, n) 5 1 ⇒ a b (mod n)
Proof:
(i) a n b and c n d ⇒ n divides |a 2 b| and |c 2 d|
⇒ nx 5 a 2 b and ny 5 c 2 d for some x, y [ Z
⇒ n(x 1 y) 5 (a 1 c) 2 (b 1 d), x 1 y [ Z
⇒ n divides |(a 1 c) 2 (b 1 d)|
⇒ (a 1 c) n (b 1 d)
(ii) and (iii) can be proved similarly.
(iv) a n b and c n d ⇒ ac n bc and bc n bd (by (iii))
⇒ ac n bd (since n is transitive)
(v) is a simple consequence of (iv) and the principle of induction.
(vi) ac n bc and (c, n) 5 1 ⇒ n divides |ac 2 bc| 5 c|a 2 b|
⇒ n divides |a 2 b| (since (c, n) 5 1)
⇒ a n b.
Corollary 2.2.1. Let 1 , n [ Z and a, b, a1, a2, …, ar, b1, b2, …, br [ Z. Then,
the following holds:
r r r r
(i) ai bi (mod n) for all 1 i r ⇒ ai ≡ bi (mod n) and ai ≡ bi (mod n).
r r i1 i1 i1 i1
ai ≡ bi (mod n).
i1 i1
(ii) If f (x) is a polynomial with integer coefficients and
a b (mod n), then f (a) f (b) (mod n).
Proof: To prove (i), use Theorem 2.2.3 (i) and apply induction on r. (ii) is a
consequence of (i), (iii) and (v) of Theorem 2.2.3.
Observe that, when x y (mod 8), either of x and y may be replaced by
the other in any polynomial congruence modulo n (by Corollary 2.2.1 (ii))
and this idea can be used in solving linear congruences ax b (mod n)
when a and n are relatively prime. Before proving this, let us have the
following.
Proof: Since (|a|, n) 5 1, there exists u and v [ Z such that u|a| 1 vn 5 1 (by
Theorem 2.1.8) and hence
sa 1 vn 5 1 (*)
s 5 nq 1 r and 0 r , n.
(nq 1 r)a 1 vn 51
or ra 1 (v 1 qa)n 5 1 (**)
Theorem 2.2.5. Let n [ Z1 and a and b [ Z such that (|a|, n) 5 1. Then, the
linear congruence equation
ax b (mod n)
has a unique solution r in {0, 1, 2, …, n 2 1} and the set of all integer solu-
tions of this is precisely equal to the congruence class r 1 nZ.
Example 2.2.1. Let us find all integer solutions of 55x 65 (mod 80). First
observe that, if m is a common divisor of a, b and n, then ax b (mod n) if
and only if ma x ≡ mb (mod mn ). Since 5 is a common divisor of 55, 65 and 80,
55x 65 (mod 80) if and only if 11x 13 (mod 16). Note that (16, 11) 5 1.
A quick check reveals that 311 1 (mod 16). Therefore, the integer solutions
of 11x 13 (mod 16) are all y 313 (mod 16) or, equivalently, all y 7
(mod 16) (since 39 7 (mod 16)). Thus, integer solutions of 55x 65 (mod 80)
are members of 7 1 16Z and those in {0, 1, 2, …, 79} are precisely 7, 23,
39, 55, 71.
The method that is followed in Example 2.2.1 above is formalized in the
following.
Proof: Let d 5 (a, n). Suppose that ax b (mod n) has an integer solution. Let s
be an integer solution of ax b (mod n). Then, as b (mod n) and hence nr 5
as 2 b for some r [ Z. Now, b 5 as 2 nr, d divides both a and n and hence
and hence nq 5 ax0 2 b, which implies that ax0 b (mod n). Thus, ax b
(mod n) has an integer solution.
We close this section by developing tests for the divisibility of integers by
various primes. These tests are easy for small primes, but these are not practi-
cal for large primes. The following is easy, since any integer is divisible by 2
if and only if the last digit in it is one of 0, 2, 4, 6, and 8.
Theorem 2.2.7. Let a [ Z1, then a is even if and only if a r (mod 10) for
some r [ {0, 2, 4, 6, 8}. Also a is divisible by 5 if and only if a 0 (mod 10)
or a 5 (mod 10) (that is, the last digit in a is 0 or 5).
r
Theorem 2.2.8. Let a [ Z1 and a 5 ar ar 2 1 … a1a0 5 ai 10i, where ai’s are
i0
integers such that 0 ai 9. Then,
r r
a ∑ ai (mod 3), a ∑ ai (mod 9)
i0 i0
Example 2.2.2
1. Let us test the divisibility of 62354 by 7. Here, a 5 62354 5 10k 1 a0,
where k 5 6235 and a0 5 4 since 10.5 1 (mod 7), m7 5 5.
7 divides 62354 ⇔ 7 divides k 1 m7a0
⇔ 7 divides 6235 1 5.4 (5 6255)
⇔ 7 divides 625 1 5.5 (5 650)
⇔ 7 divides 65 1 5.0 (5 65)
Since 7 does not divide 65, it follows that 7 does not divide 62354.
2. Consider a 5 5876438 and test its divisibility by 7 we have 10.5 1
(mod 7) and hence m7 5 5.
7 divides a ⇔ 7 divides 587643 1 5.8 (5 587683)
⇔ 7 divides 58768 1 5.3 (5 58783)
Example 2.2.3
1. Test 7892654 for its divisibility by 11. Since 10.10 1 (mod 11), we
have m11 5 10. Also, since 10 2 1 (mod 11), we can take s 5 21 in
Theorem 2.2.9 (iii).Therefore,
11 divides 7892654 ⇔ 11 divides 789265 1 (21) 4 (5 789261)
⇔ 11 divides 78926 1 (21)1 (5 78925)
⇔ 11 divides 7892 1(21)5 (5 7887)
⇔ 11 divides 788 1 (21)7 (5 781)
⇔ 11 divides 78 1 (21)1(5 77),
which is true.
Thus, 11 divides 7892654. In this context, note that Corollary 2.2.2 (iii)
is a better test for the divisibility by 11.
2. Test 7892654 for the divisibility by 13.
Since 10.4 1 (mod 13), we have m13 5 4. Therefore,
13 divides 7892654 ⇔ 13 divides 789265 1 4.4 (5 789273)
⇔ 13 divides 78927 1 4.3 (5 78939)
⇔ 13 divides 7893 1 4.9 (5 7929)
⇔ 13 divides 792 1 4.9 (5 828)
⇔ 13 divides 82 1 4.8 (5 114)
⇔ 13 divides 11 1 4.4 (5 27)
Since 13 does not divide 27, it follows that the given number 7892654 is
not divisible by 13.
3. Test whether 7892654 is divisible by 23.
Since 10.7 1 (mod 23), m23 5 7.
We have 23 divides 7892654 ⇔ 23 divides 789265 1 7.4 (5 789293)
⇔ 23 divides 78929 1 7.3 (5 78950)
⇔ 23 divides 7895 1 7.0 (5 7895)
⇔ 23 divides 789 1 7.5 (5 824)
EXERCISE 2(b)
3. Prove that, for any prime p, (p 2 1)! 1 1 0 (mod p). (This is known as Wilson’s
theorem.)
5. For any prime p and for integer a, prove that ap a (mod p). (This is known as
Fermat’s theorem.)
6. Let n [ Z1. A set {a0, a1, …, an21} of distinct integers is called a transversal for
congruence mod n if ai [ i 1 nZ for each 0 i n 2 1.
Prove the following for any transversal {a0, a1, …, an21} for congruence
modulo n.
(i) ai aj (mod n) for any i j
(ii) ai i (mod n) for any 0 i n 2 1
(iii) For any a [ Z, a ai (mod n) for some 0 i n 2 1.
8. For any n [ Z1, prove that any n consecutive integers form a transversal for
congruence mod n.
9. Characterise all n in each of the following cases that satisfy the given
condition
(i) 1 1 2 1 3 1 … 1 (n 2 1) 0 (mod n)
(ii) 12 1 22 1 32 1 … 1 (n 2 1)2 0 (mod n)
(iii) 13 1 23 1 33 1 … 1 (n 2 1)3 (mod n)
10. For any n [ Z1, prove that 3 divides n implies 3 divides m for m any rearrange-
ment of the digits in n.
n
11. For any n [ Z1, prove that 103 ≡ 1 (mod 3n2 ).
12. Find all the digits x (0 x 9) for which 12x, 527, 846, 531 is divisible by 3;
9; or 11.
14. Prove the following for any relatively prime positive integers m and n:
(i) For any a and b [ Z1, a b (mod mn) ⇔ a b (mod m)
and a b (mod n)
(ii) If c and d are both integer solutions of x a (mod n) and of x b (mod
m), then c d (mod mn).
Definition 2.3.1. Let Z be the set of all integers and Z* 5 Z 2 {0}. We define
a binary relation R on the set Z 3 Z* as follows: For any (a, b) and (c, d) in
Z 3 Z*,
(a, b) R (c, d) ⇔ ad 5 bc (that is, the products ad and bc are equal).
The following is a straight forward verification.
Definition 2.3.2. For any (a, b) [ Z 3 Z*, the equivalence class of (a, b)
corresponding to R will be denoted by a . That is,
b
a
5 R(a, b) 5 {(c, d) [ Z 3 Z* : (a, b) R (c, d)}
b
i.e., a 5 {(c, d) [ Z 3 Z* : ad 5 bc}.
b
2
For example, represents the set of all pairs (c, d) of integers, with d 0,
3
such that 2d 5 3c. Note that, for any (a, b) and (c, d) [ Z 3 Z*,
a c
⇔ (a, b) R (c, d) ⇔ ad 5 bc.
b d
Definition 2.3.3. For any (a, b) [ Z 3 Z*, the R-equivalence class a is called
b
a rational number and the set of all rational numbers is denoted by Q. That is,
a
: a and b ∈ and b ≠ 0
b
ad a
bd b
0 0 0
b d 1
a 0
⇔ a0
b d
ab a
b 1
r
a ( ) ad
b , if c ≠ 0.
s c ( )
d
bc
0
4. r 1 (2 r) 5 0 (5 )
1
5. r ? (s ? t) 5 (r ? s) ? t
6. r ? (s 1 t) 5 r ? s 1 r ? t
7. r?s5s?r
1 b
8. r ? 1 5 r, where 15 (5 for any 0 b [ Z)
1 b
9. r ? ( 1r ) 5 1
Definition 2.3.5. Any nonempty set together with the operations 1, 2, ? and /
satisfying the properties (1) to (9) above is called a field.
We will be discussing about fields in great detail later in Part III and Part
IV of this book. We just want to highlight here that the set Q of rational num-
a
bers is a field. For any integer a, consider the rational number and, we can
a 1
see that the map a is an injection of the set Z of integers into the set
1
Q of rational numbers. If we identify a with a/1, then we can see that Z is a
subset of Q and, for any a and b [ Z,
a b a b a a a b ab
, and ⋅ .
1 1 1 1 1 1 1 1
These demonstrate that the usual arithmetical operations addition, subtrac-
tion and multiplication on the integers are simply the restrictions of those on
rational numbers to Z. Thus, for all practical purposes, we can treat integers
as rational numbers by means of the identification of a with a .
1
As we have constructed rational numbers from integers, we can construct
real numbers from rational numbers. However, the procedure is not as simple
as the construction of rational numbers. We need some more techniques from
analysis to construct real numbers from rational number. However, for the
benefit of an enthusiastic reader, a brief sketch of the construction of real
numbers is given in the exercises. The proofs are not very difficult, but require
care, attention and some elementary knowledge about sequences, Cauchy
sequences, convergent sequences and their limits. The real number system is
denoted by R and it is known that R is a field.
Next we construct the system of complex numbers. Consider the Cartesian
product R 3 R, where R is the set of real numbers. We define addition, sub-
traction, multiplication and division to make R 3 R a field. For any z 5 (a, b)
and w 5 (c, d) in R 3 R, let us define
z 1 w 5 (a, b) 1 (c, d) 5 (a 1 c, b 1 d)
2 z 5 2(a, b) 5 (2a, 2b)
z ? w 5 (a, b) ? (c, d) 5 (ac 2 bd, ad 1 bc)
1 1 a b
, if a 0 or b 0
z (a, b) a 2 b 2 a 2 b 2
z (a, b) ac bd bc ad
and , if c 0 or d 0.
w (c, d ) c 2 d 2 c 2 d 2
Theorem 2.3.2. R 3 R, together with the operations defined above, is a field.
For any a [ R, consider (a, 0) in R 3 R. Then, a (a, 0) is an injection
of R into R 3 R and satisfies the following for any a and b [ R.
(a 1 b, 0) 5 (a, 0) 1 (b, 0)
(2a, 0) 5 2(a, 0)
(ac, 0) 5 (a, 0)?(c, 0)
1
, 0 1 if a 0
a ( a, 0)
c (c, 0)
, 0 if a 0.
a ( a, 0)
Put (0, 1) 5 i.
R 3 R 5 {a 1 ib : a and b [ R}
where i is the element (0, 1). a 1 ib is the usual familiar form of complex
numbers and let us agree to call any element of R 3 R as a complex number.
C 5 {a 1 ib : a and b [ R}
Z1 ⊂ N ⊂ Z ⊂ Q ⊂ R ⊂ C
in such a way that the usual arithmetic operations addition, subtraction and
multiplication on each of these are precisely restrictions of those on the next
system. Moreover, Q, R and C are fields while the others are not.
We close this section with an additional operation, namely the complex
conjugation, on C.
z a ib ( a i( b)).
If z 5 a 1 ib, and a and b [ R, then a and b are called real part and imaginary
part of z, respectively.
The following are easy verifications.
Theorem 2.3.3. The following holds for any complex numbers z and z:
1. z z z z
2. zz
3. zz z z
4. zz The real part of z.
2
5. z−z The imaginary part of z
2i
6. z z ⇔ z ∈ R ⇔ The real part of z 0.
z z a2 b2 ,
|z|2 5 z z 5 a2 1 b2.
EXERCISE 2(c)
A sequence {an} of rational numbers is said to be a Cauchy sequence if, for each posi-
tive rational number [, there exists n0 [ Z1 such that
|an 2 am| , [ for all n and m $ n0.
A sequence {an} in Q is said to be convergent if there exists r [ R such that, for
each rational [ . 0, there exists n0 [ Z1 such that |an 2 r| , [ for all n $ n0 and in
this case we write an → r and r 5 limit of an.
Prove the following:
3. If a, b, c and d [ Q with a , b and c , d, then (a, b)Q is bijective with (c, d)Q,
where
(a, b)Q 5 {r [ Q : a , r , b}.
6. { 1n } is a Cauchy sequence.
7. Every convergent sequence in Q is a Cauchy sequence.
10. For each r [ R, there exists {an} [ CS(Q) such that an → r and, if {bn} is
another Cauchy sequence in Q such that bn → r, then {an} ~ {bn}.
12. For any a [ Q, the sequence {an}, such that an 5 a for all n, is called a con-
stant sequence and is denoted by {a}. Then, a {
a} is an injection of Q into
CS(Q)/~.
13. If {an} and {bn} [ CS(Q), then {an 1 bn} and {an bn} [ Q.
{a n} ~ {an} and {b n} ~ {bn} ⇒ {an bn} ~ {an bn} and {anbn} ~ {anbn}.
{
an}{
bn} {a
n bn }
{
an} {
an}
where {an} is the ~-equivalence class of {an} in CS(Q). Then, the operations 1, 2 and ?
are well-defined on CS(Q)/~.
2.4 ORDERING
The well-ordering property of positive integers is with respect to the natural or
usual ordering. This natural ordering is there on the rational number system and
the real number system also. However, there is no such ordering on the complex
number system. In this section, we introduce the abstract concept of a partial
ordering on a given set and discuss its elementary properties in general and those
of the natural ordering on R in particular. Let us begin with the following.
a # b ⇔ b $ a.
Example 2.4.1
1. (Z1, #), (Z, #), (Q, #) and (R, #) are all partially ordered sets, where
# is the natural ordering.
2. For any nonempty set X, the equality relation is a partial order on X. That
is, for any a and b [ X, if we define a # b if and only if a 5 b, then # is
a partial order on X. Note that this is the only binary relation on X which
is both an equivalence relation and a partial order on X.
3. Let P(X) be the set of all subsets of a given set X. For any A and B [
P(X), define
A # B if and only if A is a subset of B.
Definition 2.4.2. A partial order # on a set X is called a total order if, for
any a and b [ X, either a # b or b # a and, in this case, (X, #) is called a
totally ordered set.
R together with the natural ordering is totally ordered set. In Examples
(2) and (3), the partial orders are not total orders, except when X has at most
one element.
Example 2.4.2
1. In (R, ), Z1 is bounded below and not bounded above, while the set Z2
of negative integers is bounded above and not bounded below.
2. 1 is the lub of the interval (0, 1) and 0 is the glb of (0, 1) in (R, ).
Example 2.4.3
1. The interval (0, 1) in (R, ) has neither a minimal element nor a maxi-
mal element.
2. Let X be a set with more than one element and consider the poset
(P(X), #) of all subsets of X. Let Y be the set of all nonempty subsets
of X. Then, Y has minimal elements; in fact, for any x [ X, {x} is a
minimal element in Y and is not the least element, since {x} {y} for
any y x in X. Also, let Z be the set of all proper subsets of X. Then,
Z has maximal elements; in fact, for any x [ X, X 2 {x} is a maximal
element in Z and is not greatest in Z.
Example 2.4.4
1. Z is a chain in (R, ). In fact, R itself is a chain in (R, ) and hence any
nonempty subset of R is a chain.
2. If X 5 {a, b, c, d}, then
A 5 {, {a}, {a, b}, {a, b, c}, X}
Zorn’s lemma 2.4.1. Let (X, ) be a poset in which each chain has an upper
bound in X. Then, (X, ) has a maximal element.
The following is an equivalent form of Zorn’s lemma and, in this form only
the Zorn’s lemma is used several times in this book.
Example 2.4.5
1. The natural order on Z1 is a well-order (by Theorem 2.1.1).
2. The natural order on Q is a total order, but not a well-order; for, the
interval (0, 1) ∩ Q has no least member, since for any 0 , a , 1, there
is a rational number r such that 0 , r , a.
3. The division order | on Z1 (that a|b if a divides b) is not a total order (for
example, if p and q are distinct primes, then p q and q p) and hence
not a well-order.
product of infinite class of sets. That is, if {Ai}i[I is an infinite class of non-
empty sets, then their Cartesian product can be defined as
∏ A {c : I → ∪ A
i∈ I
i
i∈ I
i : c(i ) ∈ Ai for all i ∈ I }.
The axiom of choice, given below, say that the Cartesian product of any non-
empty class of nonempty sets is a nonempty set.
The Axiom of Choice 2.4.1. Given any nonempty class {Ai}i[I of nonempty
sets, there is a choice function c : I → ∪ Ai (that is, c is a function such that
i∈I
c(i) [ Ai for all i [ I).
EXERCISE 2(d)
1. List all the partial orders on a 2-element set, a 3-element set and a 4-element
set.
2. Prove that the number of partial orders on an n-element set is less than or equal
n ( n1)
to 2 2 .
6. Let (X1, ), …, (Xn, ) be posets and X 5 X1 3 X2 3 … 3 Xn. Prove that the
lexicographic ordering on X is a well-order if and only if the partial order on
each of the Xi’s is a well-order.
2.5 MATRICES
Though matrices are originated from the study of solutions of certain systems
of linear equations and are later found to be in one-to-one correspondence with
linear transformations of a finite dimensional linear space into another finite
dimensional linear space, but these have acquired an independent status and
form one of the most important areas of study in modern abstract algebra. In
particular, matrices are a rich source of examples and counter examples of sev-
eral concepts in noncommutative algebraic structures which we come across
throughout this book. Actually, we study later in detail about matrices over an
abstract ring. However, in this section, we briefly discuss matrices over the real
number system or complex number system. Let us begin with the following.
Definition 2.5.1. For any positive integer n, let In denote the set of integers
from 1 to n; that is,
In 5 {1, 2, …, n}.
A : Im 3 In → R (or C)
Here, aij is called the ijth entry in the matrix A and m 3 n is called the size
of A. Actually, the size of A is not an integer, but it is a pair (m, n) (which is
usually written as m 3 n) of integers. An m 3 n matrix A 5 (aij) and an r 3
s matrix B 5 (bij) are said to be equal if m 5 r, n 5 s and aij 5 bij for all 1 #
i # m and 1 # j # n; that is, A and B have equal number of rows and equal
number of columns and have the same ijth entry for each i and j. The n-tuple
(ai , aj , …, ai ) is called the ith row and the m-tuple (ai , a2 , …, am ) is called
1 2 n j j j
the jth column of the m 3 n matrix A 5 (aij). A 1 3 n matrix is called a row
matrix and an m 3 1 matrix is called a column matrix.
on the diagonal, all other entries are 0. A diagonal matrix A 5 (aij) is called a
scalar matrix if a11 5 a22 … 5 ann
Example 2.5.1
2 3 1 1
1. 1 0 2 3 is a 3 3 4 matrix over R (over C also, since R ⊆ C).
0 1 1 1
1 2 1
2. 2 3 0 is a square matrix of order 3 over C.
2 i i
1 2 0
3. 0 2 3 is an upper triangular matrix of order 3.
0 0 1
2 0 0 0
3 1 0 0
4. is a lower triangular matrix of order 4.
1 2 2 0
5 2 0 4
2 0 0
5. 0 1 0 is a diagonal matrix.
0 0 1
2 0 0
6. 0 2 0 is a scalar matrix.
0 0 2
Note that, for any positive integer n, if we define f : R → Mn(R) by f(a) 5
(aij), where
a if i j
aij ,
0 if i ≠ j
Definition 2.5.4. Let m and n be any positive integers and A 5 (aij) and B 5
(bij) be any m 3 n matrices over R or C. Then, we define
A 1 B 5 (cij), where cij 5 aij 1 bij for all 1 i m and 1 j n and define
2A 5 (2aij)
Example 2.5.2
2 3 1 4 3 1 2 5
Let A 5 1 2 0 3
and B 5 2 3 1 4 M3 3 4(R).
4 5 2 2 1 2 0 1
4 5
2 3 31 1 2
Then, A 1 B 1 2 23 0 (1) 3 4
4 (1) 5 (2) 2 0 2 1
5 4 3 1
1 5 1 1
3 3 2 3
2 3 1 4
and A 1 2 0 3 .
4 5 2 2
Definition 2.5.5. Let m, n and r be any positive integers and A 5 (aij)
Mm3n(R) and B 5 (bij) Mn 3 r(R). Then, we define the product AB as an
m 3 r matrix given by
n
A ? B 5 (cij) where cij aik bkj
k1
Note that for the product AB to be defined it is necessary that the number
of columns in A must be equal to the number of rows in B. Therefore, even
if AB is defined, BA may not be defined. If we define the dot product of two
n-tuples, a 5 (a1, a2, …, an) and b 5 (b1, b2, …, bn) by
then the ijth entry in the product AB is precisely the dot product of ith row in A
and jth column in B. Also, note that the product of any two square matrices of
the same order is always defined.
3 2 1
2 1 3 2
2 0 1
Example 2.5.3. Let A 3 2 1 0 M3 3 4(R) and B
1 1 0
M4 3 3(R) 6 4 2 1
4 2 3
Then, A ? B 5 (cij), where
C11 5 2 ? 3 1 1 ? 2 1 3 ? 1 1 2 ? 4 5 19
C12 5 2 ? 2 1 1 ? 0 1 3 (‑1) 1 2 2 5 5
C13 5 2 ? (21) 1 1?1 1 3?0 1 2?3 5 5
C21 5 3 ? 3 1 2?2 1 (21)?1 1 0 ? 4 5 12
C22 5 3 ? 2 1 2?0 1 (21) ? (21) 1 0 ? 2 5 7
C23 5 3 ? (21) 1 2 ? 1 1 (21) ? 0 1 0 ? 3 5 21
C31 5 6 ? 3 1 4 ? 2 1 2 ? 1 1 1 ? 4 5 32
C32 5 6 ? 2 1 4 ? 0 1 2?(21) 1 1 ? 2 5 12
and C33 5 6 ? (21) 1 4 ? 1 1 2 ? 0 1 1 ? 3 5 1
19 5 5
and hence AB 5 12 7 1. Similarly, we can compute BA and see that
32 12 1
AB BA. Here, note that AB is a 3 3 3 matrix and BA is a 4 3 4 matrix.
Even if AB and BA are of same size, they may not be equal; for, consider the
matrices
0 2 1 1 0 2
A 0 1 2 and B 2 1 0.
3 0 1 3 2 1
7 4 1 6 2 3
Then, AB 8 5 2 and BA 0 5 4 and therefore AB BA.
6 2 7 3 8 8
Definition 2.5.6
1. For any m and n Z1, the m 3 n matrix all of whose entries are zero is
called the zero matrix and is denoted by Om3n or, simply O, when there is
ambiguity about the size of the matrix.
2. For any n Z1, the square matrix (dij) is called the identity matrix of
order n, where dij is defined by
1 if i j
dij .
0 if i j
The identity matrix of order n is denoted by In3m or, simply I, when there is
no ambiguity about the order of the matrix.
Theorem 2.5.2. The following holds for any matrices A, B and C, in the sense
that whenever one side of an equation is defined, then the other side is also
defined and both sides of that equation are equal.
1. A(BC) 5 (A B)C
2. A(B 1 C) 5 AB 1 AC
3. (A 1 B)C 5 AC 1 BC
4. AI 5 A 5 IA, where I is the identity matrix of appropriate order.
In addition to the operations addition and multiplication of matrices, we
have yet another operation of matrices, namely the scalar multiplications. The
real or complex numbers are called scalars and we multiply any matrix by any
scalar as defined below.
aA 5 (aaij).
Theorem 2.5.3. The following holds for any matrices A and B and for any
scales a and b.
Theorem 2.5.4
At 5 (aji).
Example 2.5.4
1 2
1 2 3
1. If A , then A 2 1
t
2 1 0
3 0
3 1 4
3
3 1 2
1 0 2
2. If A 1 0 2 4, then A
t .
2 2 3
4 2 3 0
3 4 0
3. If A is a square matrix, then At is also a square matrix of order same
as of A.
AB 5 I 5 BA.
Theorem 2.5.6
1. If A is an m 3 n matrix and B is an n 3 r matrix, then (AB)t 5 BtAt
2. If A is a nonsingular square matrix, then there exists a unique square
matrix B such that
AB 5 I 5 BA
EXERCISE 2(e)
(i) A1B
(ii) (A 1 B) 1 C
(iii) B1C
(iv) A 1(B 1 C)
(v) AB
(vi) At
(vii) Bt
(viii) BtAt
(ix) BA
(x) AtBt
2. For any two matrices A and B, prove that both AB and BA are defined if and only
if A and Bt are of the same size and that, in this case, both AB and BA are square
matrices.
6. For any scalar a, let Sa be the n 3 n scalar matrix in which all the diagonal entries
are a and other entries are 0. Prove that SaA 5 aA 5 ASa for all n 3 n matrices A.
n( n 1)
1 1 0 1 n
n
2
(ii) 0 1 1 0 1 n
0 0 1 0 0 1
where, for any square matrix A, An is defined inductively by A0 5 I and An 5
An21 ? A for any n . 0.
10. Prove that the sum and product of two upper triangular matrices are again upper
triangular matrices and that the same statement for lower triangular matrices is
also true.
12. Prove that the diagonal entries of a skew-symmetric matrix are all zero.
14. Prove that any square matrix can be expressed as the sum of a symmetric matrix
and a skew-symmetric matrix.
2.6 DETERMINANTS
In this section, we briefly discuss an important function known as determinant
function which maps square matrices into scalars. The term ‘determinant of A’
is conventionally used to call the value of this function at a given square matrix A.
Determinants have definite importance as a theoretical tool, besides their
effectiveness as a device for computations. For example they provide us with
simple criterion for the nonsingularity; namely, a square matrix is nonsingular
if and only if its determinant is nonzero.
There are several ways of defining the determinant function. However, we
prefer the classical definition which uses permutations. In view of this, we
first have a brief discussion on permutations. We begin with the following.
Definition 2.6.1. For any positive integer n, let In 5 {1, 2, …, n}. Any bijec-
tion of In onto itself is called a permutation on In. The set of all permutations
on In is denoted by Sn.
Any permutation f on In can expressed by means of an array (a 2 3 n matrix)
1 2 3 n
f (1) f (2) f (3) f ( n)
symbolising that each i is mapped to f (i). Note that the order of the columns
in this representation of f is immaterial. For example
1 2 3 4 5 3 4 2 1 5
and
3 4 1 5 2 1 5 4 3 2
f 5 (i1 i2 … ir)
That is, f (i1) 5 i2, f (i2) 5 i3, …, f (ir21) 5 ir and f (ir) 5 i1 and f (i) 5 i for all
i In 2{i1, i2,…, ir}. Observe that (i1 i2 … ir), (i2 i3 … iri1), …, (ir i1 i2 … ir21)
are all represent the same cycle. A 2-cycle is called a transposition. Note that,
if f is an r-cycle, the f r(5 f ο f ο … ο f, r times) is the identity map on In and
r is the least such positive integer. In particular, if f is a transposition, then f 2
is the identity map and f interchanges two elements in In and keeps all other
elements fixed.
Two cycles (a1 a2 … ar) and (b1 b2 … bs) are said to be disjoint if ai bj for
all 1 i r and 1 j s. It can be easily proved that f o g 5 g o f for any
disjoint cycles f and g and that any permutation on In can be expressed, in an
essentially unique way, as a product of disjoint cycles. Further, any cycle is a
product of transpositions (since (a1 a2 … ar) 5 (a1ar) o (a1ar21) o … o (a1a2))
and hence any permutation can be expressed as a product of transpositions,
although not necessarily uniquely. For example (2 4) ? (4 5) ? (1 3) 5 (1 3) o
(2 4) o (1 3) o (4 5) o (1 3). However, it can be proved (see Corollary 6.4.2)
that, if a permutation can be expressed as a product of even number of trans-
positions, then it cannot be expressed as a product of odd number of transpo-
sitions. In view of this, a permutation is called an even (odd) permutation if
it is a product of even (odd, respectively) number of permutations. If f and g
are even permutations, then clearly f o g, f21 and g21 are also even (since f 5
f1 ο f2 ο … ο fr implies that f21 5 fr 21 ο f21r21 ο … ο f221 ο f121). Note that an
r-cycle is even if and only if r is odd.
1 if f is even
sgn f .
1 if f is odd
∑ (sgn f ) a
f ∈Sn
a
1f (1) 2f (2) … anf ( n )
Examples 2.6.1
a a12
1. Let A 11 be a 2 3 2 matrix. Since S2 has only two elements,
a21 a22
namely the identity e which is an even permutation and the transposition
s 5 (1 2) which is odd, we have
det A 5 (sgn e)a11a22 1 (sgn s)a12a21
5 a11a22 2 a12a21
a11 a12 a13
2. Consider a 3 3 3 matrix A a21 a22 a23
a31 a32 a33
S3 has 3! elements; these are
1 2 3 1 2 3 1 2 3
e , f , g
1 2 3 2 3 1 3 1 2
1 2 3 1 2 3 1 2 3
a , b ,
2 1 3 1 3 2 3 2 1
R1
R
A = 2 or A 5 (C1, C2, …,Cn)
Rn
a1j
where Ri 5 (ai1 ai2 … ain) and Cj 5 a2j .
anj
R1 R1 R1
R
2
Ri1
1. det Ri Si det Ri det Si
Ri1 Ri1
Rn Rn Rn
R1 R1
Ri−1 Ri1
2. det aRi = a det Ri
Ri+1 Ri1
Rn Rn
Proof:
1. The left hand side of the equation is
∑ (sgn f )a
f ∈Sn
1f (1) … ai1f (i1) ( aif (i ) bif (i ) )ai+1f (i+1) … anf (n )
Theorem 2.6.3. If two rows of a square matrix A are equal, then det A 5 0.
∑ (sgn f )[a1f (1) … arf (r ) … asf (s ) … anf (n ) − a1fg (1) … arfg (r ) … asfg (s ) … anfg (n ) ]
f ∈A
∑ (sgn f )[a1f (1) … arf (r ) … asf (s ) … anf (n ) − a1fg (1) … arfg (r ) … asfg (s ) … anfg (n ) ]
f ∈A
∑ (sgn f )[a1f (1) … arf (r ) … asf (s ) … anf (n ) − a1f (1) … arf (s ) … asf (r ) … anf (n ) ]
f ∈A
R f (1)
det (sgn f ) det A
R f (n )
R f (1) R1
det (1) det (sgn f ) det A.
m
R f (n ) Rn
Proof: Let A 5 (aij) and B 5 (bij) be two n 3 n matrices and AB 5 (cij). Then,
n
cij air brj for any 1 i, j n.
r1
n n
∑ (sgn f ) ∑ a1r1 br1 f (1) …∑ anrn brn f ( n )
f ∈Sn r11 r =1
n
∑ ( a1r1 a2r2 … anrn ) ∑ (sgn f ) br1 f (1) br2 f (2) … brn f ( n )
1 ≤ r1 , r2 ,…, rn ≤ n f ∈Sn
In the above summation, if r1, r2, …, rn are not all distinct, then, by
Theorem 2.6.3,
∑ (sgn f ) b
f ∈Sn
r1 f (1) br2 f (2) … brn f (n ) 0
det(AB) ∑ ( a1g (1) a2g (2) … ang ( n ) ) ∑ (sgn f ) bg (1) f (1) … bg ( n ) f ( n )
g ∈ Sn f ∈Sn
∑ (sgn g ) a1g (1) … ang ( n ) (det B) (by Theorem 2.6.10)
g ∈Sn
1
det A1
det A
Aij ∑ (sgn f ) ∏ a
f ∈Sn r ≠i
r f (r )
f (i ) j
∑ (sgn f ) a
f ∈Sn
1f (1) … ai-1f (i-1) ⋅ ai+1f (i+1) … anf (n ) .
f (i )= j
that contain a given entry aij as a factor. These are corresponding to those
permutations f for which f (i) 5 j. Therefore, the sum of all the summands in
the summation for det A involving aij as a factor is
∑ (sgn f ) a
f ∈S n
1f (1) … a2f (2) anf (n ) = aij Aij
f (i ) j
and hence
n
∑ ∑ ( sgn f ) a 1f (1) a2f (2) … anf (n )
j1 f ∈Sn
f (i ) = j
n
∑ aij Aij for each 1 i n.
j1
n
Similarly, det A aij Aij for each 1 j n.
i1
Next, let 1 r i n and consider
n n
∑a
j1
rj Aij ∑ arj
j1
∑ (sgnf) ∏ a
f ∈ Sn s ≠i
s f (s )
f (i ) j
Proof: Let us first find the cofactor A11 of a11 in det A. By Definition 2.6.4,
we have
A11 ∑
f ∈Sn
( sgn f ) a2f (2) a3f (3) … anf (n ) .
f (1) 1
Next, to find the value of a general cofactor Aij of aij, let us bring aij to the (1, 1)
position by performing some row and column interchanges on A. To bring aij
to the (1, 1) position, we move the jth column to the left to j21th column (that
is, interchanging jth column and j21th column), then to j22th column, …,
to 1st column, so that after j21 interchanges of the columns, the jth column
becomes the first column. Next, in a similar way, we move the ith row up to
the 1st row in i21 interchanges of the rows. Now, we have a matrix B that
is obtained from A by j21 interchanges of columns and i21 interchanges
of rows. Therefore, by Theorem 2.6.4 and Corollary 2.6.1, the determinant
of the new matrix B is (21)(j21)1(i21) det A 5 (21)ij det A. If B 5 (bij), then
b11 5 aij and the matrix obtained by deleting the 1st row and 1st column in B is
precisely Aij. Thus,
n
det A ∑ ( 1) i j
aij det Aij , for each 1 ≤ i ≤ n
j1
n
∑ ( 1) i j
aij det Aij , for each 1 ≤ j ≤ n.
i1
Now,
det A 5 (21)111 a11 det A111 (21)112 a12 det A121 (21)113 a13 det A13
a22 a23 a a23 a a22
a11 a12 21 a13 21
a32 a33 a31 a33 a31 a32
a11 ( a22 a33 a32 a23 ) a12 ( a21 a33 a31 a23 ) a13 ( a21 a32 a31 a22 ).
This is the expansion of det A with respect to the 1st row. Notice that we get
the same value for det A by expanding it with respect to any other row or any
column.
The following result characterizes nonsingular matrices in terms of the
value of their determinants.
1
A21 5 (A )t
det A ij
where (Aij) is the matrix whose ijth entry is the cofactor Aij of the ijth entry in A.
Proof: Let A 5 (aij) be an n 3 n matrix and Aij be the cofactor of aij in det A.
Let B be the transpose of (Aij).
That is, B 5 (Aij)t 5 (bij); say.
Then, bij 5 Aji for all i and j. By Theorem 2.6.6,
n n
∑a
j1
rj b ji ∑ arj Aij ri det A
j1
AB 5 det A?I
1 1
A B I
det A det A
B A
Definition 2.6.5. For any n 3 n matrix A 5 (aij), the transpose of the matrix
(Aij) is called adjoint of A and is denoted by adj A, where Aij is the cofactor
of aij in det A.
EXERCISE 2(f)
1 x1 x12 x1n-1
1 x2 x22 x2n-1
4. Prove that det ∏ ( x j − xi ).
1 ≤ i ≤ j ≤ n
i xn xn2 xnn-1
1 a a3
5. Prove that det 1 b b3 (b c)(c a)( a b)( a b c).
1 c c 3
6. If A is a nonsingular square matrix such that A2 5 A, then prove that det A 5 1.
a b c 2b 2c
(ii) det 2a bc a 2c ( a b c)3
2a 2b c a b
8. If A is a n 3 n matrix such that Am 5 On3n for some m Z1, then prove that A
is singular.
9. Let A 5 (aij) and B 5 (bij) be n 3 n matrices such that bij 5 (21)ij aij for all i
and j. Then, prove that det A 5 det B.
12. Prove that for any n 3 n nonsingular matrix A, det (adj A) 5 (det A)n21.
Definition 3.1.1. Let S be a nonempty set and S 3 S be the set of all ordered
pairs of elements of S. That is,
Example 3.1.1
1. The usual multiplication ‘?’ is a binary operation on the set Z of integers.
Quite often, we simply write ab for a ? b.
2. The usual addition 1 is a binary operation on the set Z of integers.
3. Let us define the mapping 2 : Z 3 Z → Z by 2 (a, b) 5 a 2 b, the
usual difference of b with a, for any integers a and b. Then, 2 is a binary
operation on Z. Note that 2 is not a binary operation on the set Z1 of
positive integers, since a 2 b need not be positive for any two positive
integers a and b. Likewise, 2 is not a binary operation on the set Z2 of
negative integers. Note that both the multiplication and addition given in
(1) and (2), respectively are binary operations on Z1. The operation 2
is called the difference operation. Note that each of addition, difference
and multiplication is a binary operation on the set Q of rational numbers
and on the set R of real numbers.
4. Let R be the set of real numbers and, for any real numbers a and b,
define
a ∧ b 5 The minimum of a and b
and a ∨ b 5 The maximum of a and b.
Then, both ∧ and ∨ are binary operations on R. In fact, these are binary
operations on any nonempty subset of R and, in particular, on Q, Z and Z+.
5. Consider the set Z+ of positive integers. For any a and b in Z+, define
a g b 5 (a, b), the greatest common divisor of a and b
and a b 5 [a, b], the least common multiple of a and b.
Then, g and are both binary operations of Z+. Usually, we write (a, b)
and [a, b] to denote respectively the greatest common divisor and the
least common multiple of any positive integers a and b.
6. Let X be any set and P(X), the power set of X; that is, P(X) is the set of
all subsets of X. For any A and B P(X), define
A ∩ B 5 {x : x A and x B}
A ∪ B 5 {x : x A or x B}
A 2 B 5 {x : x A and x B}
A 1 B 5 (A 2 B) ∪ (B 2 A).
Then, ∩, ∪, 2 and 1 are all binary operations on the set P(X) and are
respectively called intersection, union, difference and symmetric differ-
ence. Note that P(X) is not empty even if X is empty.
7. For any positive integers m and n, let Mm3n(R) be the set of all m 3 n
matrices over the real number system R. For any A 5 (aij) and B 5 (bij)
in Mm3n(R), define
A 1 B 5 (cij),
where cij 5 aij 1 bij and 1 is the usual addition of real numbers. Then, 1
is a binary operation on Mm3n(R) and is called the addition of matrices
(of same order).
8. For any positive integer n, let Mn(R) be the set of all n 3 n matrices
(square matrices of order n). For any A 5 (aij) and B 5 (bij) in Mn(R),
define
n
A ? B 5 (dij), where dij 5 ∑ aik bkj
k =1
for any 1 # i, j # n. That is,
dij 5 ai1blj 1 ai2b2j 1 … 1 ainbnj.
(The operations involved in defining dij above are the usual addition and
multiplication of real numbers.) Then, ‘?’ is a binary operation on Mn(R)
and is called the multiplication of square matrices (of the same order).
Note that, we can define addition and multiplication (as in (7) and (8)
above) on the sets Mm3n(Z) and Mn(Z) of matrices over the set Z of
integers or on the sets Mm3n(Q) and Mn(Q) of matrices over the set Q of
rational numbers.
9. Let X be any set and M(X) be the set of all mappings of X into itself. For
any mappings f and g in M(X), define
f o g : X → X by (f o g)(x) 5 f (g(x)) for any x X.
Then, 1 and ‘?’ are binary operations on C and are called the usual addi-
tion and multiplication of complex numbers, respectively.
11. Let n be a positive integer and
Zn 5 {0, 1, 2, …, n 2 1}.
Note that the 1 and ? on the right hand sides of the above are the usual
addition and multiplication in the real number system R. Then, 1 and ?
are binary operations on RX and are respectively called the point-wise
addition and point-wise multiplication.
16. The above example can be generalized as follows. Let * be a binary oper-
ation on a nonempty set S and X be a nonempty set. Let SX be the set of
all mappings of X into S. For any f and g in SX, define f * g : X → S by
(f * g)(x) 5 f (x) * g(x) for all x X.
Then, * is a binary operation on SX and is called the point-wise operation
on SX with respect to the operation * on S.
Note that, in (16) above (and so is in (15)), we have denoted the operations
on SX and in S with the same symbol *. There should not be any confusion.
The * on the left sides is the one we are defining on SX and that on the right
sides is the given binary operation on S.
Note 3.1.1. In defining a binary operation on a set S, one should observe the
following:
(i) For each ordered pair of elements in S, the element assigned to it must
be again an element of S.
(ii) Exactly one element of S must be assigned to each ordered pair of
elements in S.
For example, consider the set R of real numbers and, for any a and b in R,
define a * b a . Then, * is not a binary operation on R, since * is not defined
b
for all ordered pairs of elements in R. Note that 2 * 0 is not defined, while 0
* 2 is defined. However, this * is a binary operation on a smallest set, namely,
the set R 2 {0} of nonzero real numbers.
Let us consider another example. Let S be the set of all people in a par-
ticular village and define, for any a and b in S, a * b 5 c where c is a person
whose height is equal to the minimum of the heights of a and b. Then, * is
not a binary operation on S, since a * b may not be an unique element in S;
there can be more than one person in the village whose height is equal to the
minimum of those of a and b.
Note 3.1.2. Let S be a finite set with n elements. Then, the number of elements
in S 3 S is n2. Any binary operation S is simply a mapping
2
of S 3 S into S; that is,
an element of Ss3s. Therefore, there are exactly n n many binary operations on S.
* a1 a2 a3 … aj … an
a1 a1 * a1 a1 * a2 a1 * a3 …a1 * aj … a1 * an
a2 a2 * a1 a2 * a2 a2 * a3 … a2 * aj… a2 * an
a3 a3 * a1 a3 * a2 a3 * a3 … a3 * aj … a3 * an
ai ai * a1 ai * a2 ai * a3 ai * a j ai * an
an an * a1 an* a2 an * a3 …an * aj … an * an
In the example given below, we shall construct the table representing the
binary system (Z9, 19) where Z9 5 {0, 1, 2, …, 8} and 19 is the addition
modulo 9 (see Example 3.1.1 (11)).
for any a and b in Z9. The following table represents the binary system (Z9, 19).
19 0 1 2 3 4 5 6 7 8
0 0 1 2 3 4 5 6 7 8
1 1 2 3 4 5 6 7 8 0
2 2 3 4 5 6 7 8 0 1
3 3 4 5 6 7 8 0 1 2
4 4 5 6 7 8 0 1 2 3
5 5 6 7 8 0 1 2 3 4
6 6 7 8 0 1 2 3 4 5
7 7 8 0 1 2 3 4 5 6
8 8 0 1 2 3 4 5 6 7
Here, 2 19 7 5 0, 2 19 5 5 7, 4 19 5 5 0, 5 19 7 5 3, 6 19 8 5 5 , 8 19
8 5 7, etc.
Example 3.1.3. Let S be the set of all positive divisors of 36 and, for any a
and b in S, define
(see Example 3.1.1 (5)). Then, (S, g) is a binary system which is represented
by the table given below. We have
Let (S, *) be a binary system. For any elements a, b and c in S, the expres-
sion a * b * c has no meaning, since * is a binary operation and hence * is
defined for pairs of elements. For example, 1 2 2 2 3 has no meaning and we
should specify whether it is (1 2 2) 2 3 (this is what we usually take) or 1 2
(2 2 3). Note that (1 2 2) 2 3 ≠ 1 2 (2 2 3). For arbitrary elements a, b and
c in a binary system (S, *), a * b and c are two elements in S and hence (a * b)
* c is defined and so is a * (b * c). In general, (a * b) * c and a * (b * c) may
be different. In this context, we have the following definition.
(a * b) * c 5 a * (b * c)
Example 3.1.4. The binary operations given in Example 3.1.1, except those
in (3), (6) and (16), are all associative and therefore, these together with the
corresponding underlying sets, are semigroups. (Z, 2) is not a semigroup,
since 2 is not associative. For any set X, ( P(X), ∩), (P(X), ∪) and (P(X), 1),
given in Example 3.1.1 (6), are all semigroups. However, (P(X), 2) is not a
semigroup; For, consider X 5 {a, b, c, d}, A 5 {a, b, c}, B 5 {c, d} and c 5
{b}. Then, A, B and C P(X) and
(a * b) * (c * d)
a * (b * (c *d))
a * ((b *c) * d)
((a * b) * c) * d
(a * (b * c)) * d.
One can easily prove that these expressions represent one single element, if *
is associative. In fact, we can generalize and extend the associativity for any
finite sequence of elements. First, let us have the following definition.
meaningful product of a1, a2, …, an, in this order. The one given in the follow-
ing definition is a meaningful product.
Definition 3.1.7. Let (S, *) be a binary system and a1, a2, …, an S. The
standard product in=1ai of a1, a2, …, an, in this order, is defined inductively
as follows.
a1 if n 1
n
n1
∏ ai
∏ ai * an if n 1
.
i1
i =1
For example,
∏a a * a
i =1
i 1 2
3 2
∏ a ∏ a * a
i =1
i
i =1
i 3 ( a1 * a2 ) * a3
∏ a (((a * a ) * a ) * a ) * a .
i =1
i 1 2 3 4 5
x5s*t
where s and t are meaningful products of a1, a2, …, ar and ar+1, ar+2, ..., an, in
these orders, respectively. By the induction hypothesis, we get that
r n
s ∏ ai and t a j
i1 jr1
Now, x s * t
r n
∏ ai * ∏ a j
i =1 jr1
r n1
∏ ai * ∏ a j * an
i1 jr1
r n − 1
∏ ai * ∏ a j * an (since * is associative)
i1 j = r + 1
n1
∏ ai * an (by the induction hypothesis)
i1
n
∏ ai
i1
Thus, x is the standard product of a1, a2, ..., an, in this order. b
Example 3.1.5. Except in (3), (6), (8), (9), (14) and (16), all other binary oper-
ations given in Example 3.1.1 are commutative. The operations ∩, ∪ and 1 on
P(X), given in Example 3.1.1 (6), are all commutative and the operation 2 on
P(X) is not commutative. Also, the operation * on SX, given in Example 3.1.1
(16) is commutative if and only if the operation * on S is commutative.
n n
∏a ∏a
i1
i
i1
s (i )
for any permutation s of {1, 2, …, n} (by Theorem 3.1.1). We shall prove theo-
rem using induction on n. If n 5 1, the theorem is trivial, if n 5 2, the theorem
follows from the commutativity of *. Let n . 2 and assume that the theorem is
n n1
∏a ∏ as ( i ) * as ( n )
i1
s (i )
i1
k1 n
∏ ai * ∏ ai * ak (by induction hypothesis)
i1 ik1
k1 n
∏ ai * ak * ∏ ai (by associativity and commutativity of *)
i1 ik 1
n
∏ ai b
i1
n2 n n2 n
n
2 2
n2 n
| S X | n 2
.
Worked Exercise 3.1.2. Let X be a nonempty set and M(X) be the set of all
mappings of X into itself. Let o be the composition of mappings on M(X).
Then prove that o is a commutative operation on M(X) if and only if X has
exactly one element.
Answer: If X has exactly one element, then M(X) also has only one element
and the result is trivial. Conversely suppose that X has more than one element,
choose a ≠ b X and define f and g : X → X by
52 5
5 2
515.
EXERCISE 3(a)
3. Fill in the blanks in the following table such that the binary operation * repre-
sented by the table is commutative.
* a b c d e f
a d e c b
b f d f
c a c a d
d c e d
e b b d
f d c b a
7. Prove that the associativity and the commutativity are independent of each
other.
3.2 GROUPS
The integer 0 has a special property in the binary system (Z, 1) and is unique
satisfying this property; namely
a 10 5 a 5 0 1 a for all a Z.
Similarly, the integer 1 is the unique element in the binary system (Z, ?)
satisfying the property
Likewise, the identity map IX, defined on any set X by IX(x) 5 x for all
x X, is the unique element in the binary system (M(X), o) satisfying the
property
where M(X) is the set of mappings of X into itself and o is the composition of
mapping. An abstraction of these ideas is made in the following definition.
Example 3.2.1
1. 0 is the only identity in (Z, 1).
2. 1 is the only identity in (Z, ?), where ‘?’ is the usual multiplication of
integers.
3. Let X be any nonempty set and IX : X → X be defined by IX(x) 5 x for all
x X. Then, IX is the identity in (M(X), o).
4. Let m and n be any positive integers and Mm3n(R) be the set of all m3n
matrices over R. Let Om3n be the m 3 n matrix in which each entry is the
number 0. Then,
O mn A A A O mn
for all matrices A and hence Om3n is the identity in (Mm3n(R), 1), where
1 is the usual addition of matrices.
5. Let S be any nonempty set and define a * b 5 b for all a and b in S. Then,
every element of S is a left identity in the binary system (S, *).
6. Let S be any nonempty set and define a * b 5 a for all a and b S. Then,
every element of S is a right identity in (S, *).
7. For any set X, the empty set [ is the identity in (P(X), ∪) and also in
(P(X), 1), where ∪ is union operation and 1 is the symmetric difference
operation.
8. Also for any set X, the whole set X is the identity in (P(X), ∩), where ∩
is the intersection operation.
From examples (7) and (8) above, the concept of identity is depending on
the binary operation of the system. Also, it depends on the underlying set. For
example, 0 is the identity in the system (Z, 1) where as it is not the identity
in (Z+, 1), since 0 is not an element in the underlying set Z+.
Also, from examples (5) and (6), observe that a binary system can possess
any number of right identities without having any left identity and vice versa.
However, if e is a right identity and f is a left identity, then e must be equal to
f. This is proved in the following theorem.
Theorem 3.2.1. Let (S, *) be a binary system and e be a left identity in (S, *).
Then, every right identity in (S, *) is an identity and coincides with e.
Corollary 3.2.1. There can be almost one identity in any binary system.
Proof: Let (S, *) be a binary system and e and f be identities in (S, *). Since e
is a left identity and f is a right identity, e 5 f by the above theorem.
Example 3.2.2
1. Let E be the set of even integers and is the usual multiplication of inte-
gers. Then, there is no identity in the binary system (E, ?).
2. Let S be any set and define a * b 5 b for all a, b S. Then, every ele-
ment of S is a left identity in (S, *). However, (S, *) has no right identity,
unless S is a singleton set (one element set).
3. Similarly, if we define a * b 5 a for all a, b S then every element of S
is a right identity in (S, *) and there are no left identities in (S, *), unless
S is a singleton set.
4. The integer 0 is the only identity in (Z, 1).
Example 3.2.3
1. The set Mn(R) of all n 3 n square matrices over R together with the
matrix multiplication is a monoid. Here, the matrix In, in which all the
diagonal entries are 1 and the others are 0, is the identity element in
Mn(R). In is called the identity matrix. Note that In 5 (aij), where aij 5 1
or 0 according as i 5 j or i ≠ j.
2. (Z, 1), (Z, ?) and (Z+ , ?) are all monoids, where 1 and ‘?’ are the usual
addition and multiplication, respectively. 0 is the identity in (Z, 1) and
1 is the identity in (Z, ?) and in (Z+, ?).
a 1 x 5 b.
a 1 (2a 1 b) 5 (a 1 (2a)) 1 b 5 0 1 b 5 b.
This is to say that 2a 1 b is the unique real number satisfying the equa-
tion a 1 x 5 b. In this process finding the unique solution of a 1 x 5 b, we
have skipped one step by not explaining what 2a is. It is obvious that 2a is
the unique real number x satisfying the equation x 1 a 5 0. This concept is
abstracted in the following definition.
Example 3.2.4
1. In the monoid (R, 1), the number 0 is the identity and every element of
R has inverse.
2. In the monoid (Z, ?), 1 is the identity, where ‘?’ is the usual multiplica-
tion. Here, 1 and 21 are the only elements having inverses.
3. Consider the set M(Z) of all mappings of Z into itself. Then, (M(Z), 0)
is a monoid, in which the identity map I, defined by I(x) 5 x for all x
Z, is the identity element. Define f : Z → Z by f (x) 5 2x for all x Z.
For each integer a, define ga : Z → Z by
x if x is an even integer
g a (x ) 2
a if x is an odd integer.
which is false, since we cannot get an integer g(1) such that 2g(1) 5 1.
From the examples (3) and (4) above, we have noticed that an element in
a monoid can have several left inverses without having any right inverses.
However, if an element has both left inverse and right inverse, then they must
be equal. This is proved in the following theorem.
5 ar (since e is identity)
Example 3.2.5. Let X be any set and P(X) be the power set of X. For any A
and B P(X), define
A 1 B 5 (A 2 B) ∪ (B2A).
Then, (P(X), 1) is a monoid with the empty set [ as the identity element.
Here, for any A P(X),
A 1 A 5 (A2A) ∪ (A2A) 5 [ ∪ [ 5 [
Theorem 3.2.3. Let X be any nonempty set and M(X) be the set of all map-
pings of X into itself. Then, (M(X), o) is a monoid in which the following
holds for any f M(X).
1. f has a left inverse in M(X) if and only if f is an injection.
2. f has a right inverse in M(X) if and only if f is a surjection.
3. f is invertible in M(X) if and only if f is a bijection.
Proof: We know that (M(X), o) is a monoid (see Example 3.2.3 (4)) in which
o is the composition of mapping and the map IX : X → X, defined by IX(x) 5 x
for all x X, is the identity. Let f be an arbitrary element of M(X); that is,
f : X → X is a mapping.
1. Suppose that f has a left inverse in M(X). Then, there exists g M(X)
such that
g o f 5 IX.
For any a, b X, we have
f (a) 5 f (b) ⇒ g(f (a)) 5 g(f (b))
⇒ (g o f )(a) 5 (g o f )(b)
⇒ IX(a) 5 IX(b)
⇒ a 5 b.
Therefore, f is an injection.
Conversely, suppose that f is an injection. Define
g : X → X by
a if x f (a) for some a ∈ X
g (x ) ,
s otherwise
Example 3.2.6
1. (Z, 1), (Q, 1), (R, 1) and (C, 1) are all groups in which 1 is the usual
addition, 0 is the identity and 2a is the inverse of any element a.
2. (Q 2{0}, ?), (R 2{0}, ?) and (C 2{0}, ?) are all groups in which ‘?’ is the
1
usual multiplication, 1 is the identity and a is the inverse of any element a.
3. Neither (Z, ?) nor (Z2{0}, ?) are groups, since not all elements are
invertible.
4. For any set X, (P(X), 1) is a group in which 1 is the symmetric differ-
ence operation, is the identity and, every element is inverse of itself.
5. Let X be a nonempty set and S(X) the set of all bijections of X onto itself.
Then, (S(X), o) is a group in which o is the composition of mappings, IX
is the identity and f –1 is the inverse of any bijection f. Recall that, for any
bijections f and g, the composition f o g is also a bijection.
6. The set Mm3n(R) of all m 3 n matrices over R together with the addition
of matrices is a group. Here the zero matrix, in which all the entries are 0,
is the identity and, for any A 5 (aij) , the matrix (2aij) is the inverse of A.
9. For any points a 5 (a1, a2) and b 5 (b1, b2) in the two-dimensional
Euclidean space R 3 R, d(a, b) be the usual Euclidean distance between
a and b; that is,
Let X be the set of all points on a given geometrical figure (in fact, X may
be any nonempty subset of R 3 R).
A bijection f of X onto itself is called a symmetry of X if
d(f (a), f (b)) 5 d(a, b) for all a, b X.
Theorem 3.2.4. Let (M, *) be a monoid and G be the set of all invertible ele-
ments in (M, *). Then, (G, *) is a group.
Proof: First we shall observe that G is a nonempty set, since the identity e in
(M, *) is always invertible and hence e G. Also, if a and b G and a9 and
b9 are inverses of a and G, respectively, then
1. an+m 5 an * am
2. (a9)n 5 a–n
3. (an)m 5 anm 5 (am)n
Worked Exercise 3.2.2. Let G be the set of all rotations of the plane about
the origin in the plane and o the composition of mappings. Thus, prove that
(G, o) is a group.
Answer: The rotation about the origin through an angle u can be represented
analytically as the map fu : R 3 R → R 3 R defined by
f u if u 2p
f u ο f
f u2p if u 2p
Worked Exercise 3.2.3. For any real numbers a and b with a ≠ 0, define
let G 5 {fa,b : 0 ≠ a R and b R}. Then prove that (G, o) is a group, where
o is the composition of mappings.
and hence fa,b o fc,d 5 fac, ad1b. Also, if a ≠ 0 and c ≠ 0, then ac ≠ 0. Therefore,
0 is an associative binary operation on G. Further, f1,0 is the identity in
(G, o) , since f1,0(x) 5 1 x 1 0 5 x for all x R. Also, f 1 ,− b is the inverse
a a
of fa,b. Thus, (G, o) is a group.
We shall conclude this section with two more important examples of
groups given below.
Example 3.2.7. Let G 5 {f1, f2, f3, f4, f5, f6}, where each f i is a function of
R 2 {0, 1} into itself as defined below.
1
f1 (x ) x, f 2 (x ) , f 3 (x ) 1 x,
x
1 x 1 x
f 4 (x ) , f 5 (x ) and f 6 (x ) .
1 x x x 1
0 f1 f2 f3 f4 f5 f6
f1 f1 f2 f3 f4 f5 f6
f2 f2 f1 f4 f3 f6 f5
f3 f3 f5 f1 f6 f2 f4
f4 f4 f6 f2 f5 f1 f3
f5 f5 f3 f6 f1 f4 f2
f6 f6 f4 f5 f2 f3 f1
Example 3.2.8. The group discussed here is called the group of symmetries
of the square (see Example 3.2.6 (9)). Let X be the set of all points in a square
of unit side. Recall that a symmetry of X is a bijection f of X into itself such
that
d1
1 2
h
r3 r1
4 3
r2
d2
o e r1 r2 r3 h v d1 d2
e e r1 r2 r3 h v d1 d2
r1 r1 r2 r3 e d1 d2 v h
r2 r2 r3 e r1 v h d2 d1
r3 r3 e r1 r2 d2 d1 h v
h h d2 v d1 e r2 r3 r1
v v d1 h d2 r2 e r1 r3
d1 d1 h d2 v r1 r3 e r2
d2 d2 v d1 h r3 r1 r2 e
EXERCISE 3(b)
1. Determine the following in which 1n and ?n are the addition and multiplication
modulo n, for a given positive integer n.
(i) 7 112 11
(ii) 8 110 7
(iii) 7 ?12 11
(iv) 8 ?10 7
(v) 4 ?6 5
(vi) (7 110 6) ?10 8
(vii) 77 in (Z8, ?8)
(viii) 5–6 in (G7,?7)
(ix) 7–8 in (G11,?11)
(x) 68 in (Z9,?9)
3. Determine which of the following gives a group structure on the given set.
(i) For any a, b Z, a * b 5 a 1 b 1 ab.
(ii) For any a, b R+ , a * b ba .
(iii) a * b ab for any a, b R+.
2
(iv) a * b 5 |ab| for any a, b C.
(v) a * b 5 a 1 b 2 2 for any a, b Z.
(vi) For any a, b, R+ , a * b 5 5ab.
(vii) For any a, b Q+ , a * b 5 |ab|.
(viii) a * b 5 a 1 b 1 ab, for any a, b Z.
form a group under the matrix multiplication. What is identity element? Determine
the inverse of each element.
6. Let (G, *) be a group. Prove that the identity e is the only element satisfying
x * x 5 x.
7 Let n be a positive integer and G be the set of all nth root of unity; that is,
8. Let (G, *) be a group and X be any nonempty set. Let GX be the set of all map-
pings of X into G. For any f, g GX, define f * g : X → G by
Prove that (G , *) is a group. What is the identity in this group? Determine the
X
9. For any positive integer n, prove that (Zn, ?n) is a monoid, where Zn 5 {0, 1, 2,
…n 2 1} and ?n is the multiplication modulo n.
11. For any 1 # a , n, prove that a is invertible in the monoid (Zn, ?n) if and only if
a is relatively prime to n.
12. For any prime number p, prove in detail that (Gp, ?p) is a group, where
Gp 5 Zp 2 {0}.
13. For any positive integer n, give an example of a group having exactly 2n
elements.
16. Prove that the following are equivalent to each other for any integer n . 1.
(1) ?n is a binary operation on Zn 2 {0}.
(2) (Zn 2 {0}, ?n) is a group.
(3) n is a prime number.
(4) any 1 # a , n is relatively prime with n.
17. Let G be the set of all rotations about the origin in the plane and reflections in the
lines through origin. Then prove that (G, o) is a group, where o is the composi-
tion of mappings.
18. Consider the regular n-gon (polygon of n equal sides and equal internal angles)
inscribed in the unit circle in the plane, so that one of the vertices is (1, 0). Let
Rn be the set of all rotations about the origin which maps this regular n-gon into
itself. Prove that (Rn, o) is a group, where o is the composition of mappings.
How many elements are there in this group?
19. Let Dn be the set of all rotations and reflections which the regular n-gon, given in
18 above, into itself. Then prove that (Dn, o) is a group, where o is the composi-
tion of mappings. How many elements Dn has? The group (Dn, o) is called the
dihedral group of degree n (see Theorem 6.4.8). The elements of Dn are called
the symmetries of the regular n-gon.
20. For any rational numbers r and s, define r ~ s if and only if r 2 s is an integer.
Then prove that ~ is an equivalence relation on the set Q of rational numbers.
Let Q denote the set of equivalence classes w.r.t. ~ in Q and, for any classes
Z
s. Then prove that (Q , 1) is a group.
r, s, define r s r
Z
Theorem 3.3.1. Let (G, *) be a group and a, b and c G. Then, the following
holds.
1. a * b 5 e ⇔ a–1 5 b ⇔ b–1 5 a, where e denotes the identity in the
group.
2. (a–1)–1 5 a
3. (a * b) –1 5 b–1 * a–1
4. a * b 5 c ⇔ a 5 c * b–1 ⇔ b 5 a–1 * c
Proof:
1. a * b e ⇒ a1 a1 * e a1 * ( a * b) ( a1 * a) * b e * b b
a1 b ⇒ a * b a * a1 e
a * b e ⇒ b1 e * b1 ( a * b) * b1 a * (b * b1 ) a * e a
b1 a ⇔ a * b b1 * b e
2. Since a–1 * a 5 e, it follows from (1) that (a–1)–1 5 a
3. Since (a * b) * (b–1 * a–1) 5 a * (b * b–1 )* a–1 5 a * e * a–1 5 a * a–1 5 e,
again from (1) it follows that (a * b)–1 5 b–1 * a–1
4. a * b c ⇒ c * b1 ( a * b) * b1 a * (b * b1 ) a * e a
a c * b1 ⇒ a * b (c * b1 ) * b c * (b1 * b) c * e c
a * b c ⇒ a1 * c a1 * ( a * b) ( a1 * a) * b e * b b
b a−1 * c ⇒ a * b a * ( a1 * c) ( a * a1 ) * c e * c c
Note that if we take e for c in (4), we get (1).
Let us recall that a semigroup is a pair (S, *) where S is a nonempty set and *
is an associative binary operation on S and that a semigroup with identity is
called a monoid and also that a monoid is called a group if every each of its
elements is invertible.
Theorem 3.3.2. Let (S, *) be a semigroup. Then, (S, *) is a group if and only
if the following conditions are satisfied.
1. (S, *) has a right identity e. That is, there exists e S such that a * e 5
a for all a S.
2. For each a S, there exists a9 S such that a * a9 5 e
Proof: If (S, *) is a group, then clearly (1) and (2) are satisfied. Conversely sup-
pose that the conditions (1) and (2) are satisfied. By (1), there exists e S, such
that a * e 5 a for all a S. We shall prove that this e is actually the identity in (S,
*). Let a be an arbitrary element in S, By (2), there exists a9 and x in S such that
Consider
( a * a) * ( a * a) a * ( a * a) * a
( a * e) * a (by (i)) (ii)
a * a
Now,
e ( a * a) * x (by (i))
(( a * a) * ( a * a)) * x (by (ii))
( a * a) * (( a * a) * x ) (by associativity)
( a * a) * e (by (i))
a * a
a9 * a 5 e 5 a * a9 (iii)
Theorem 3.3.3. A semigroup (S, *) is a group if and only if, for any elements
a and b in S, the equation
a*x5b and y * a 5 b
are solvable in S (in the sense that there are elements x and y in S satisfying
these equations).
a * (a–1 * b) 5 (a * a–1) * b 5 e * b 5 b
and (b * a–1) * a 5 b * (a–1 * a) 5 b * e 5 b
Now, b * e 5 (s * a) * e 5 s * (a * e) 5 s * a 5 b.
Thus, e is a right identity in (S, *). Also, since a * x 5 e is solvable in S, we
get that, for each a S, there exists a9 S such that a * a9 5 e. Thus, by the
above Theorem 3.3.2, (S, *) is a group.
Recall that, in the elementary school mathematics, one is used to conclude
b 5 c whenever a 1 b 5 a 1 c for some a and we were used to give reasoning
for this by saying ‘subtracting a from both sides’ which amounts to adding
−a both sides.
This is abstracted in the following theorem.
Proof: Consider
a * b a * c ⇒ a−1 * (a * b) a1 * (a * c)
⇒ (a−1 * a) * b ( a1 * a) * c
⇒ e*be*c
⇒ bc
A semigroup may satisfy both the left and right cancellation laws without
being a group. This is to say that the converse of the above theorem is not
true. For, consider the following examples.
Example 3.3.1
1. Consider the semigroup (Z1, 1), where Z1 is the set of positive integers
and 1 is the usual addition. Since (Z, 1) is a group, (Z, 1) satisfies both
the cancellation laws. Since Z1 is a subset of Z, (Z1, 1) also satisfies
both the cancellation laws. Nevertheless, (Z1, 1) is not a group, since
this has no identity.
2. A monoid may satisfy the cancellation laws without being a group. Con-
sider the set W of all nonnegative integers. Then, for the same reason
given above, (W, 1) is a monoid satisfying both the cancellation laws
and this is not a group, since no element, except 0, has inverse.
Even though the converse of Theorem 3.3.4 is not true in general, we prove
the converse in the case of finite semigroups. Recall that a semigroup (S, *) is
called finite if the underlying set S is finite.
Theorem 3.3.5. Let (S, *) be a finite semigroup satisfying both the cancella-
tion laws. Then, (S, *) is a group.
Proof : Since S is a finite set, we can enumerate the elements of S. Let a1, a2,
…, an be all the distinct elements of S. That is,
a * S 5 S.
a * x 5 b and y * a 5 b
Example 3.3.2
1. (Z, 1), (Q, 1), (R, 1) and (C, 1) are all abelian groups, since the
addition 1 is commutative.
2. (Q2{0},?), (R2{0},?) and (C2{0},?) are abelian groups, since the
multiplication is commutative.
3. For any set X, (P(X), 1) is an abelian group, since, for any A and B in
P(X),
A 1 B 5 (A 2 B) ∪ (B2A) 5 (B2A) ∪ (A 2 B) 5 B 1 A.
4. Let X be a set with atleast three elements and S(X) the set of all bijections
of X onto itself. Then, (S(X), o) is a group which is not abelian. For, con-
sider three distinct elements a, b and c in X and define f and g : X → X by
f (a) 5 b, f (b) 5 a and f (x) 5 x for all x ≠ a, b
and g(b) 5 c, g(c) 5 b and g(x) 5 x for all x ≠ b, c.
Then, (f o g)(a) 5 f (g(a)) 5 f (a) 5 b
and (g o f )(a) 5 g(f (a)) 5 g(b) 5 c ≠ b 5 (fog)(a).
Therefore, f o g ≠ g o f. Thus, (S(X), o) is an abelian group.
5. The matrix multiplication is not commutative. Let NSMn(R) be the set
of all nonsingular n 3 n matrices over R. Then, (NSMn(R), ? ) is a group
which is not abelian if n > 1.
6. The addition of matrices is a commutative operation. (Mm3n(R), 1) is an
abelian group for any positive integers m and n, where Mm3n(R) is the set
of all m 3 n matrices over R.
Theorem 3.3.6. The following are equivalent to each other for any group
(G, *).
1. (G, *) is an abelian group.
2. (a * b)–1 5 a–1 * b–1 for all a and b G.
3. (a * b)2 5 a2 * b2 for all a and b G.
Proof:
(1) ⇒ (2): If (G, *) is an abelian group and a and b G, then, by Theorem
3.3.1 (1),
(2) ⇒(3): Suppose that (a * b)–1 5 a–1 * b–1 for all a and b G.
Then, for any a and b G, we have
(a * b)2 5 (a * b) * (a * b)
5 (a * (b–1)–1) * ((a–1)–1 * b)
5 a * ((b–1)–1 * (a–1)–1) * b
5 a * (a–1 * b–1)–1 * b
5 a * ((a * b)–1)–1 * b
5 a * (a * b) * b
5 (a * a) * (b * b) 5 a2 * b2
(a * b)2 5 a2 * b2 ⇒ (a * b) * (a * b) 5 (a * a) * (b * b)
⇒a * (b * a) * b 5 a * (a * b) * b
⇒ b * a 5 a * b (by cancellation laws)
Worked Exercise 3.3.1. Let X be any nonempty set and S(X) be the set of all
bijections of X onto itself. Then prove that (S(X), o) is an abelian group if and
only if |X| , 3.
Answer: If |X| ≥ 3, then we have proved in Example 3.3.2 (4) that the group
(S(X), o) is not abelian.
On the other hand, suppose that |X| , 3. Then, |X| 5 1 or 2. If |X| 5 1,
then S(X) has only one element, namely the identity map and S(X) 5 {IX}
is clearly abelian. If |X| 5 2, say X 5 {a, b}, then there are exactly two
bijections, namely the identity map IX and the function f : X → X defined
by f (a) 5 b and f (b) 5 a and therefore S(X) 5 {IX, f} which is clearly an
abelian group.
Worked Exercise 3.3.2. Let (G, *) be a group. Then prove that (G, *) is abe-
lian if and only if there exist three consecutive integers n such that (a * b)n 5
an * bn for all a, b G.
(a * b)0 5 e 5 e * e 5 a0 * b0
(a * b)1 5 a * b 5 a1 * b1
and (a * b)2 5 (a * b) * (a * b)
5 a * (b * a) * b
5 a * (a * b) * b
5 (a * a) * (b * b) 5 a2 * b2
a * (an–1 * b) * bn–1 5 an * bn
5 (a * b)n
5 (a * b) * (a * b)n–1
5 (a * b) * an–1 * bn–1
5 a * (b * an–1) * bn–1
bn–1 * (a * b) 5 (bn–1 * a) * b
5 (a * bn–1) * b (by (iv))
5 a * (bn–1 * b)
5 a * bn
5 bn * a (by (v))
5 (bn–1 * b) * a
5 bn–1 * (b * a)
Worked Exercise 3.3.3. Let (G, *) be a group such that a2 5 e for all a G.
Then prove that (G, *) is an abelian group.
Worked Exercise 3.3.4. Let (G, *) be a group such that x2 ≠ e for all x ≠ e in
G. Then prove that (G, *) is an abelian group if and only if
am 5 bm and an 5 bn.
Answer: Since m and n are relatively prime there exist integers r and s
such that
rm 1 sn 5 1
Now, consider
a 5 a1 5 arm 1 sn
5 (am)r * (an)s
5 (bm)r * (bn)s
5 brm1sn
5 b1 5 b
Worked Exercise 3.3.6. Let (G1, *), (G2, *), …, (Gn, *) be groups and
G 5 G1 3 G2 3 … 3 Gn.
For any a 5 (a1, a2, …, an) and b 5 (b1, b2, …, bn) G, define
Then prove that (G, *) is a group and (G, *) is abelian if and only if each
(Gn, *) is abelian.
For all a, b G and hence (G, *) is abelian. Conversely suppose that (G, *)
is abelian. Fix 1 ≤ i ≤ n. For any ai and bi Gi, consider
Worked Exercise 3.3.7. Describe the group of symmetries of the set X of all
points on the perimeter of an equilateral triangle.
5 4
1 6 2
for all a and b X, where d(a, b) is the usual Euclidean distance between
a and b. Therefore, a symmetry of X should map each vertex to a vertex only
and hence we can identify the group (Sym(X), o) with the group (S(V), o),
where S(V) is the set of bijections of the set V of vertices onto V 5 {1,
2, 3}. It follows that Sym(X) has exactly b elements which are described
below. Each of these map each of 1, 2, 3 to the number given vertically
below that.
a o a 5 b; a o a o a 5 e 5 c o c 5 d o d 5 s o s 5 b o b o b
c o d 5 a; d o c 5 a o a 5 b; d o s 5 a
s o d 5 b; c o s 5 b; s o c 5 a
a o c 5 s; c o a 5 d; a o d 5 c
d o a 5 s; s o a 5 c; a o s 5 d
o e a b c d s
e e a b c d s
a a b e s c d
b b e a d s c
c c d s e a b
d d s c b e a
s s c d a b e
EXERCISE 3(c)
1. Prove the following for any elements a, b and c in a group G in which e is the
identity.
(i) a*b5e⇔b*a5e
(ii) (a * b) *c 5 e ⇔ (b * c) * a 5 e
2. Give an example of a finite semigroup satisfying the left cancellation law, but
not satisfying the right cancellation law.
5. Prove that a group (G, *) is abelian if and only if (a * b)n 5 an * bn for all a and
b in G and for all integers n.
6. In any finite semigroup, prove that there exists an element e such that e2 5 e.
7. Let m and n be relatively prime positive integers and (G, *) a group such that
am * bm 5 bm * am and an * bn 5 bn * an
for all a and b G. Then prove that (G, *) is an abelian group.
8. Let (G, *) be a finite group and suppose that the number of elements in G is
even. Then prove that there exists an element a, other than the identity, in G
such that a2 5 e.
9. For any element a in a finite group (G, *), prove that there exists a positive inte-
ger n such that an 5 e, the identity in G.
10. For any finite group (G, *), prove that there exists a positive integer n such that
an 5 e for all a G
where e is the identity in (G, *).
11. For any elements a and b in a group (G, *) and for any positive integer n, prove that
(a * b * a–1)n 5 a * b * a–1 ⇔ bn 5 b.
12. Let (G, *) be a group and X be any nonempty set. Let GX be the set of all map-
pings of X into G. For any f and g GX, define f * g : X → G by
(f * g)(x) 5 f (x) * g(x) for all x G.
14. Prove that any group with fewer than six elements is abelian.
15. For any integer n . 1, prove that the set of all nonsingular n 3 n matrices over
R forms a nonabelian group under the multiplication of matrices.
* e
e e
Next, we consider a two element group G. Then, there should be only one
element in G other than the identity e and therefore G 5 {e, a}, where a ≠ e,
we have a * e 5 a 5 e * a and e * e 5 e. What could be a * a? It cannot be a,
for, if a * a 5 a then a * a 5 a * e and, by the cancellation law, a 5 e which
is false. Therefore, the only possibility is a * a 5 e. The table for the group
(G, *), where G 5 {e, a}, is given below.
* e a
e e a
a a e
Next, we shall take up a 3-element group. In this case, there are exactly two
elements, say a ≠ b, in G other than the identity e. That is,
The table representing this group (G, *) should be like the one given below
* e a b
e e a b
a a
b b
Let us search for the possible entries for the vacant places in the table. First
observe that a * b ≠ a (since a * b 5 a 5 a * e ⇒ b 5 e, which is false).
Similarly, a * b ≠ b (since a ≠ e). Therefore, the only possibility is a * b 5 e
and b * a 5 e.
* e a b
e e a b
a a e
b b e
a*a5e⇒a*a*b5e*b5b
⇒a*e5b
⇒ a 5 b, which is false.
* e a b
e e a b
a a b e
b b e a
The above procedure for arriving at the full table representing the group
{e, a, b} yields the fact that the table of any three element group looks like
the same, except the interchanging of the elements a and b or relabeling the
elements a and b as b and a.
Consider the group (Zn, 1n), where Zn 5 {0, 1, 2, …, n21} and 1n is the
addition modulo n. For n 5 2 and n 5 3, the tables representing (Z2, 12) and
(Z3, 13) look like exactly the above tables representing a 2-element group
and a 3-element group. In these cases, 0 is the identity, Z2 5 {0, 1} and
Z3 5 {0, 1, 2}.
13 0 1 2
12 0 1 0 0 1 2
0 0 1 1 1 2 0
1 1 0 2 2 0 1
(Z2, 12) (Z3, 13)
Proof: Consider the table representing the operation * on G. For any ele-
ments a and b in G, the equation a * x 5 b is solvable in G if and only if b
appears in the row with a at the extreme left. Also, the equation y * a 5 b is
solvable in G if and only if b appears in the column, with a at the very top.
Now, the theorem is a direct consequence of Theorem 3.3.3.
Example 3.4.1. The table representing the group (Z9, 19) is given below.
Recall that Z9 5 {0, 1, 2, 3, 4, 5, 6, 7, 8} and 19 is the addition modulo 9.
19 0 1 2 3 4 5 6 7 8
0 0 1 2 3 4 5 6 7 8
1 1 2 3 4 5 6 7 8 0
2 2 3 4 5 6 7 8 0 1
3 3 4 5 6 7 8 0 1 2
4 4 5 6 7 8 0 1 2 3
5 5 6 7 8 0 1 2 3 4
6 6 7 8 0 1 2 3 4 5
7 7 8 0 1 2 3 4 5 6
8 8 0 1 2 3 4 5 6 7
Example 3.4.2. Let X 5 {1, 2, 3} and S(X) be the set of all bijections of X
onto itself. S(X) has six elements and these are
o e a b c d s
e e a c d s
a b e c d
b b e a d c
c c d e a b
d d e
s s d a b e
The vacancies in the above table can be filled in using Theorem 3.4.1.
Worked Exercise 3.4.1. Let X 5 {1, 2, 3} and P(X) be the set of all subsets
of X. Construct the table representing the group (P(X), 1), where 1 is the
symmetric difference.
1 e a b c p q r X
e e a b c p q r X
a a e p r b X c q
b b p e q a c X r
c c r q e X b a p
p p b a X e r q c
q q X c b r e p a
r r c X a q p e b
X X q r p c a b e
(P(X), 1).
Answer:
?7 1 2 3 4 5 6
1 1 2 3 4 5 6
2 2 4 6 1 3 5
3 3 6 2 5 1 4
4 4 1 5 2 6 3
5 5 3 1 6 4 2
6 6 5 4 3 2 1
(G7, ?7).
Worked Exercise 3.4.3. Examine whether the table below represent a group
structure.
* a b c d e
a b c d e a
b a d c a b
c d c b e c
d c a a b d
e a b c d e
EXERCISE 3(d)
1. Examine the following tables representing binary systems and determine which
of them represent a group structure.
(i) (ii)
* 0 1 2 3 4 * a b c
0 0 0 0 0 0 a a b c
1 0 1 2 3 4 b b c a
2 0 2 4 1 3 c c a b
3 0 3 1 4 2
4 0 4 3 2 1
(iii) (iv)
* a b c d e * 1 2 3 4
a b c d e a 1 1 2 3 4
b c d e a b 2 2 3 4 1
c d e a b c 3 3 4 2 3
d e a b c d 4 4 3 2 1
e a b c d e
(v) (vi)
i 2i 1 1 * 1 2 3 4
i 21 1 2i i 1 1 2 3 4
2i 1 21 i 2i 2 2 3 4 1
21 2i i 1 21 3 3 4 1 2
1 i 2i 21 1 4 4 1 2 3
3. Let n be a positive integer greater than n and S 5 {1, 2, …, n 2 1}. Prove that
the multiplication modulo n is a binary operation on S if and only if n is a prime
number.
5. By observing the tables in 4 above, prove that every 2-element groups, 3-element
groups, 4-element groups and 5-element groups is abelian.
8. For any positive integer n, give an example of an abelian group with exactly n
elements and construct a table representing it.
9. Let G be the set of all rational numbers with odd denominators. Prove that (G,
1) is a group, where 1 is the usual addition of rational numbers.
0 1 0 i
10. Let A and B , where i is the complex number such that i2 5 1.
1 0 i 0
Let
Q8 5 {(AnBm) : n, m Z}.
Prove that (Q8, ?) is a group, where ‘?’ is the usual multiplication of complex
numbers and that Q8 has exactly 8 elements. This group is called the Quaternion
group of order 8.
4.1 SUBGROUPS
Recall that a binary operation * on a set S is a mapping of S 3 S into S. If
A is a subset of S, then A 3 A is a subset of S 3 S and if the restriction of * to
A 3 A is a binary operation on A, then the restriction also will be denoted by
* and is called the operation A induced by the operation * on S. In particular,
if (G, *) is a group and A is a subset of G such that a * b A whenever a and
b A, then * can be treated as a binary on A with respect to which A can be
a group and, in this case, we say that A is a group under *.
Theorem 4.1.1. The following are equivalent to each other for any nonempty
subset H of a group G:
1. a and b H ⇒ ab H and a21 H.
2. a and b H ⇒ ab21 H.
3. a and b H ⇒ a21b H.
4. H is a subgroup of G.
e 5 h21h H
and hence aa1, aa2, …, aan are all distinct. Also, since a and ai H, we have
aH ⊆ H and H and aH have the same number of elements. From the finiteness
of H, it follows that aH 5 H. In particular,
a H 5 aH
Example 4.1.1
1. If H is a subgroup of a group K and K is a subgroup of a group G,
then clearly H is a subgroup of G. If 1 denotes the usual addition
of numbers, Z is a group of (Q, 1), Q is a subgroup of (R, 1) and
R is a subgroup of (C, 1) and hence Z, Q and R are all subgroups
of (C, 1).
2. If e is the identity in a group G, then clearly {e} and G are subgroups of
G and are called the trivial subgroups of G. Any subgroup other than {e}
and G is called a nontrivial subgroup. A group G is called nontrivial if
G {e} and trivial if G 5 {e}.
3. If ? is the usual multiplication of numbers, then Q 2 {0} is a subgroup
of (R 2 {0}, ?) and R 2 {0} is a subgroup of (C 2 {0}, ?). Therefore,
both Q 2 {0} and R 2 {0} are subgroups of (C 2 {0}, ?).
4. Let X be any nonempty set and (S(X), o) be the group of bijections of X
onto itself, where o is the composition of mappings. Let x0 be an arbi-
trary element of X and
H x0 { f ∈ S ( X ) : f ( x0 ) x0 }.
e if n 0
a a n1a
n
if n 0
1 n
( a ) if n 0
and hence an H for all integers n and for all subgroups H of G containing
a. Also, by Worked Exercise 3.2.18, <a> is a subgroup of G and contains a.
Thus, <a> is the smallest subgroup containing a.
Definition 4.1.3. For any subset S of a group G, let <S > be the intersection
of all subgroups of G containing S. Then, by the above theorem, <S > is a sub-
group of G containing S and is called the subgroup generated by S in G.
Note 4.1.1
1. For any subset S of a group G, <S > is the smallest subgroup of G con-
taining S.
2. <> 5 {e} and <G> 5 G, for any group G.
3. <{a}> 5 <a> 5 {an : n Z} for any a G.
4. For any nonempty subset S of a group G, S is a subgroup of G if and only
if S 5 <S >.
In the following, we describe the elements of <S >, for any nonempty sub-
set S of a group. The description of elements of <{a}> is already given in
Theorem 4.1.3. This is generalised in the following theorem.
Theorem 4.1.5. Let S be a nonempty subset of a group G and <S > be the
smallest subgroup of G containing S. Then,
n
S ∏ si : n ∈ and, for each i, si or si1 ∈ S .
i=1
Proof: Let A be the set defined on the right side of the required equality. If
H is any subgroup of G containing S, then s and s21 H for all s S and
therefore any product of elements of S and their inverses must be in H. This
is to say that A ⊆ H for all subgroups H containing S. Also, S ⊆ A and,
since
We have proved in Theorem 4.1.4 that the intersection of any class of sub-
groups is again a subgroup. A similar statement is not true for unions of
subgroups. In this context, we have the following theorem.
b ∈ B ⇒ a and b ∈ A ∪ B (since a ∈ A)
⇒ ab ∈ A ∪ B (since A ∪ B is a subgroup)
⇒ ab ∈ A or ab ∈ B
⇒ ab ∈ A (since ab ∈ B ⇒ a ( ab)b1 ∈ B )
⇒ b a1 ( ab) ∈ A
Therefore, B ⊆ A.
From the above, it follows that, for any subgroups A and B of a group G,
A ∪ B is not a subgroup in general. However, we have noticed earlier that
A ∩ B is always a subgroup and this is the largest subgroup contained in both
A and B. Also, there is a smallest subgroup containing both A and B (which
need not be A ∪ B). In certain cases, we can describe the elements of this
elegantly. First, we have the following definition.
Note that, from the associativity of the operation in G, we get that (AB)
C 5 A(BC) for any subsets A, B and C of G. Also, observe that a nonempty
subset A of G is a subgroup of G if and only if AA21 5 A.
Worked Exercise 4.1.1. Prove that any subgroup of the group (Z, 1) is the
subgroup generated by a single nonnegative integer.
Answer: Recall that, when the symbol 1 is used for the binary operation in
a group, then we write na for an and as such
Now, let A be a subgroup of (Z, 1). If A 5 {0}, then clearly A 5 <0>. Sup-
pose that A {0}. Then, there exists a 0 such that a A. Since A is a
subgroup, 2a also is in A. Since a or 2a is positive, it follows that A ∩ Z1
is a nonempty subset of Z1. By the well-ordering principle, A ∩ Z1 has the
smallest member, say m. Then, since m A, we get that <m> ⊆ A. On the
other hand, let x A. By the division algorithm, we can write
x 5 qm 1 r, q, r Z and 0 # r , m.
Worked Exercise 4.1.2. Compute all subgroups of the group (nZ, 1) for any
positive integer n.
Worked Exercise 4.1.3. Compute all the subgroups of (Zn, 1n) for any posi-
tive integer.
where qm 5 n.
EXERCISE 4(a)
1. Determine whether the set given is a subgroup of the group in each of the
following.
(i) {0, 1, 2, 3, 4} in (Z8, 18)
(ii) {0, 3, 6, 9} in (Z12, 112)
(iii) R1 in (R, 1)
(iv) R1 in (R2{0}, ?)
(v) Q1 in (R1, ?)
(vi) pQ in (R, 1)
(vii) {z C | |z| 5 1} in (C 2 {0}, ?)
(viii) 5Z in (8Z, 1)
3. Let # be a nonempty class of subgroups of a group G such that, for any A and B
in #, there is a member C in # containing both A and B. Then prove that ∪ C
C ∈
is a subgroup of G.
4. Let G be a group such that G 5 <a> for exactly, one element a in G. Then prove
that G has at most two elements.
5. Let G be a group having exactly two subgroups. Then prove that G 5 <a> for
some a G.
6. Let n be positive integer and consider the group (Zn, 1n) of integers modulo n. For
any 0 , d , n, prove that Zn 5 <d> if and only if d is relatively prime with n.
7. Prove that there is a bijection between the set of subgroups of (Zn, 1n) and the
set of positive divisors of n.
12. For any positive integer n, prove that (Zn, 1n) has exactly two subgroups if and
only if n is prime.
14. For any finite subgroup H of a group G and a G prove that H and aHa21 has
equal number of elements.
15. Determine all the subgroups of the group (S(X), ?) of bijections of X onto itself,
where X is a 3-element set.
G 5 {an : n is an integer}.
Example 4.2.1
a 5 qn 1 r and 0#r,n
q and r are respectively called the quotient and the remainder obtained by
dividing a with n.
a
q q 1.
n
qn # a , qn1n
and hence 0 # a 2 qn , n.
Therefore, we have a 5 qn 1 r and 0 # r , n. For proving the uniqueness, let
q0 and r0 be any integers such that a 5 q0n 1 r0 and 0 # r0 , n.
r r
Then, an q0 n0 and 0 n0 1. Therefore,
a a
q0 q0 1 and hence q0 q
n n
Example 4.2.2
1. The quotient and remainder when 46 is divided by 8 are respectively 5
and 6, since
46 5 5 ? 8 1 6 and 0 # 6 , 8.
an H ⇔ a2n 5 (an)21 H,
it follows that there exists a positive integer n such that an H. By the well-
ordering principle, there exist least positive integer m such that am H. Since
H is a subgroup and am H, we have <am> ⊆ H. On the other hand, let x H.
n 5 qm 1 r and 0 # r , m.
Let x and y G. Then, there exist integers m and n such that x 5 am and
y 5 an. Now,
Example 4.2.3. Consider the cyclic group (Z, 1) in which Z 5 <1>. Following
the proof of Theorem 4.2.2, one can prove that any subgroup of (Z, 1) must be
of the form <n> 5 {mn : m Z} 5 Zn for some nonnegative integer n.
Example 4.2.4. Let n be a positive integer and consider the group (Zn, 1n)
of integer modulo n. Here again one can prove that any subgroup of (Zn, 1n)
must be of form <d> 5 {0, d, 2d, 3d, …, (m 2 1)d}, where d is a positive
divisor of n and md 5 n (see Worked Exercise 4.1.3).
1. O(a21) 5 O(a)
2. For any integer n, there exists 0 # r , O(a) such that an 5 ar.
3. For any integer n, an 5 e if and only if O(a) divides n.
4. <a> 5 {e, a, a2, …, am21}, where m 5 O(a).
Proof:
Since 0 # r , O(a) and O(a) is the least positive integer such that aO(a)
5 e, it follows that r 5 0 and n 5 qO(a). Thus, O(a) divides n.
4. <a> 5 {an : n is an integer}
5 {ar : 0 # r , O(a)}
5 {e 5 a0, a, a2, ….., aO(a)21}.
O(a) 5 |<a>|.
That is, the order of the element a is precisely equal to the order of the cyclic
group <a>.
Also, if ar 5 as, then ar2s 5 e 5 as2r and hence O(a) divides r 2 s and s 2 r
and; if 0 # r and s , O(a), then r 5 s. Thus, e, a, a2, …, aO(a)21 are distinct.
Thus, |<a>| 5 O(a).
It is well known that the greatest common divisor (g.c.d.) of two positive
integers m and n can be written as a linear combination m and n. We shall
prove this using Theorem 4.2.2.
Theorem 4.2.5. Let m and n be two positive integers and (m, n) be the great-
est common division of m and n. Then, there exist integers r and s such that
(m, n) 5 rm 1 sn.
H 5 <d> 5 Zd.
and hence m and n <q> so that am 1 bn <q> for all integers a and b and,
in particular, d 5 rm 1 sn <q> 5 Zq. Therefore, q divides d. Thus, d is the
greatest common divisor of m and n.
The converse of the above therefore is also true, in the sense that, if d is a
common divisor of m and n and d is of the form rm 1 sn, r, s Z, then d is
m
O(a r )
( m, r )
Proof: Let 0 # r , m be fixed and d 5 (m, r). Then, by Theorem 4.2.5, there
exist integers s and t such that
d 5 sm 1 tr.
Put b 5 ar. Since d divides both m and r, m and r are integers and
d d
m r
1 s t
d d
m m rm r
b d ( a r ) d a d ( a m ) d e.
bq e ⇒ (ar )q e
⇒ a rq e
⇒ O(a) divides rq
⇒ m divides rq
m r
⇒ divides ⋅ q
d d
m
⇒ divides q since m , r 1.
d
d d
m m
∴ O (a r ) O (b) .
d (m, r )
m O (a)
O (a d ) .
d d
Proof: This follows from the above theorem and the fact that, for any posi-
tive divisor d of m, (m, d) 5 d.
Let us recall that, if G 5 <a>, then a is called a generator of the cyclic
group G. In the next result, we derive formulae to determine the number of
generators of a cyclic group. First, we have the following definition.
Definition 4.2.4. For any positive integer n, let f(n) be the number of
positive integers less than or equal to n and relatively prime with n. Then,
f : Z1 → Z1 is a function and is called the Euler–Totient function, which is
an important arithmetical function in the theory of numbers.
Note that f(1) 5 1 5 f(2), f (3) 5 2 5 f(4), f(5) 5 4, f(6) 5 2 and
f(7) 5 6. In fact, for any prime number p, f(p) 5 p 2 1, since any positive
integer less than p is relatively prime with p.
Theorem 4.2.7. Let G be a cyclic group and a G such that G 5 <a>. Then,
G is infinite if and only if an am for all n m Z and, in this case, a and
a21 are the only generators of G.
G 5 <a> 5 <a21>
and therefore a and a21 are two distinct generators of G. Now, suppose that b
is any generator of G. Then,
<b> 5 G 5 <a>
Example 4.2.5. In the group (Z3, 13) of integers modulo 3, 1 and 2 (5 21)
are the only generators and (Z3, 13) is a finite group.
2. This follows from the definition of f(n) and from (1) above. Note that,
since f(1) 5 1, (2) is trivial when n 5 1.
Example 4.2.6
1. (Z, 1) is a cyclic group with 1 and 21 as the only generators.
2. (Zn, 1n) is a finite cyclic group with f(n) generators.
3. There are exactly two generators in each of the groups (Z3, 13), (Z4, 14)
and (Z6, 16), since f(3) 5 2 5 f(4) 5 f(6).
4. Since f(8) 5 4, there are four generators for (Z8, 18) and these are 1, 3,
5 and 7.
5. For any prime number p, there are p 2 1 generators for the group (Zp, 1p),
since f(p) 5 p 2 1. That is, any nonzero (nonidentity) element Zp is a
generator.
24 24
O(16) 3.
(16, 24) 8
Worked Exercise 4.2.2. For any positive integers a and b, prove that
where (a, b) and [a, b] are respectively the g.c.d. and l.c.m. of a and b.
Answer: These follow from the fact that, for any n and m Z1,
nZ ⊆ mZ ⇔ n mZ ⇔ m divides n
Answer: Since 36Z 1 24Z5 (36, 24)Z 5 12Z and 12Z is an infinite cyclic
group generated by 12, we get that 12 and 212 are the only generators of
12Z 5 36Z 1 24Z.
Worked Exercise 4.2.4. If a cyclic group has exactly two generators, then
what can we say about the order of G?
EXERCISE 4(b)
1. State whether each of the following are true and substantiate your answer.
(i) Every finite abelian group is cyclic.
(ii) An infinite group is abelian if and only if it is cyclic.
(iii) (Q, 1) is a cyclic group, where Q is the set of rational numbers.
(iv) (R, 1) is a cyclic group.
(v) (C, 1) is a cyclic group.
(vi) If G1 and G2 are cyclic groups, then G1 3 G2 is also a cyclic group.
(vii) (C 2 {0}, ?) is an abelian group which is not cyclic.
(viii) (Q 2 {0}, ?) is a cyclic group.
(ix) Any group of order 5 or 7 is cyclic.
(x) The group S(X) of bijections of a set X onto itself is a cyclic group under
the composition of mappings.
2. Which of the following are cyclic groups? Substantiate your answers.
(i) (R 2 {0}, ?)
(ii) (P(X), 1), where X is a set.
(iii) The group of quaternions which is of order 8.
(iv) (Q1, ?)
(v) (R1 , ?)
(vi) For any positive Integer n, the group of all nth roots of unity of under the
usual multiplication of complex numbers.
3. What can we say about a cyclic group having exactly one generator?
4. List all the elements in each of the following subgroups of the groups mentioned.
(i) <7> in (Z18, 118)
(ii) <5> in (Z20, 120)
(iii) <3> in (Z12, 112)
(iv) <3> in (Z16, 116)
(v) <i> in (C 2 {0}, ?)
(vi) < 2> in (R1 , ?)
(vii) < 2> in (R, 1)
(viii) < 2 > in (C 2 {0}, ?)
(ix) <e> in any group G, where e is the identity in G.
(x) <12> in (3Z, 1).
5. Prove that O(a) is finite for any element a of a finite group and give an example
of an infinite group in which every element is of finite order.
6. If G is a group in which O(a) is finite for all a G, then can G be finite?
7. Let K and H be finite cyclic subgroups of orders m and n respectively in an abe-
lian group G. If m and n relatively prime, prove that G has a cyclic subgroup of
order mn.
8. In Exercise 7 above, if the least common multiple of m and n is K, then prove that
G has a cyclic subgroup of order K.
9. If A and B are subgroups of a group G and one of A and B is cyclic, then prove
that A ∩ B is a cyclic subgroup of G.
10. If a and b are elements of a group G, such that O(a) and O(b) are relatively prime
positive integers, then prove that <a> ∩ <b> 5 {e}.
11. Let G be a finite cyclic group of order n and d be a positive divisor of n. Then
prove that the equation xd 5 e has exactly d solutions in G.
12. Prove that the set {4, 8, 12, 16} is a group under the multiplication modulo 20. What
is the identity element? Is this a cyclic group? If so, what are its generators?
13. Let G 5 {7, 35, 49, 77}. Then prove that (G, ?84) is a group, where ?84 is the
multiplication modulo 84. What is the identity in G? Is this a cyclic group? If so,
what are its generators?
14. Is Z 3 Z a cyclic group, where the operation is a coordinate-wise addition?
15. Is Z9 3 Z16 cyclic? If so, what are the generators?
16. Let G and H be finite cyclic groups of orders m and n, respectively. Then prove
that G 3 H is cyclic if and only if m and n are relatively prime.
17. If G is a finite cyclic group and H is an infinite cyclic group, then can G 3 H be
cyclic?
18. Let G be a group and a G such that O(a) 5 24. Then find a generator of the
group <a9> ∩ <a10>. In general, find a formula for the generator of <an> ∩ <am>
for any 1 # n, m , 24.
19. Let G be a finite
n
cyclic group of order n. Then, for each positive divisor nr of n,
prove that < a r > is the only subgroup of order r and that the map r → <a r > is a
bijection of the set of positive divisors of n onto the set of all subgroups of G.
20. Find all the subgroups of (Z24, 124) and (Z30, 130).
21. Given a positive integer m, give an example of a cyclic group having exactly m
subgroups.
22. List all the subgroups of (Z625, 1625).
23. Let Un be the group of all nth-roots of unity under the usual multiplication of
complex numbers. Any generator of Un is called a primitive nth-root of unity.
Determine all the primitive nth-roots of unity for each of n 5 5, 7 and 11.
24. Prove that any group having only a finite number of subgroups must be finite.
25. Give an example of a nonabelian group such that every proper subgroup is cyclic.
26. Let G be an abelian group and n be a positive integer. Then prove that the set {a
G : O(a) divides n} is a subgroup of G.
27. Let a be an element of order n in a group G. Then prove that O(ar) 5 O(an2r) for
all 1 # r , n.
28. Let G be a cyclic group of order 24 and a G such that a8 e and a12 e. Then
prove that a is a generator of G.
29. For any elements a and b of a group, prove that O(ab) 5 O(ba).
30. Let G be an abelian group. Prove that the set of elements of finite order in G is a
subgroup of G.
Theorem 4.3.1. Let H be a subgroup of a group G and define two binary rela-
tions LH and RH on G as follows:
(a, b) LH if a2lb H
(a, b) LH ⇒ a2lb H
⇒ (a2lb)21 H (since H is a subgroup of G)
⇒ b2la H
⇒ (b, a) LH
aH 5 {ax : x H}
and Ha 5 {xa : x H}.
u(x) 5 {y X : (x, y) u}
and that the equivalence class of u form a partition of X in the sense that any
two distinct equivalence classes of u are disjoint and the union of all equiva-
lence classes of u is equal to the whole set X. In the following, we prove that
the equivalence classes of LH (respectively RH) are precisely the left (right)
cosets of H. Note that, if G is an abelian group then aH 5 Ha for all a G.
x LH(a) ⇔ (a, x) LH
⇔ a2lx H
⇔ x 5 a(a2lx) aH
x RH(a) ⇔ (a, x) RH
⇔ ax2l H
⇔ xa2l 5 (ax2l) 2l H
⇔ x 5 (xa2l)a Ha
aH 5 bH ⇔ a2lb H
and Ha 5 Hb ⇔ ab H. 2l
In particular, aH 5 H ⇔ a H ⇔ Ha 5 H.
Corollary 4.3.2. For any subgroup H of a group G, any two left (right) cosets
of H in G are either equal or disjoint.
Corollary 4.3.3. For any subgroup H of a group G, the left (right) cosets of H
in G form a partition of G, then note that, if we use 1 to denote the operation
on the group G, then write a 1 H for aH and H 1 a for Ha.
Example 4.3.1
1. If H 5 {e}, then aH 5 {a} 5 Ha for any a G and any subgroup H of G.
2. Consider the group (Z, 1) of integers under the usual addition and let
H be a subgroup of (Z, 1). Then, by Worked Exercise 4.1.1, H 5 nZ
for some nonnegative integer n. If H 5 {0}, then for any a Z, the left
coset a 1 H 5 {a} 5 H 1 a. Suppose that H {0}. Then, n > 0. For
any a Z, choose q and r Z such that
a 5 qn 1 r and 0#r,n
are all the left (right) cosets of H in Z. One can observe that these are all
distinct. Thus, there are exactly n cosets of nZ 5 H in Z.
3. Let X 5 {1, 2, 3} be a 3-element set and S(X) be the group of all bijec-
tions of X onto itself under the composition of mappings. We know from
o e a b c d s
e e a b c d s
a a b e s c d
b b e a d s c
c c d s e a b
d d s c b e a
s s c d a b e
Let H 5 {e, s}, then H is a subgroup of (S(X), o). The left and right
cosets of H in S(X) are given below.
eH 5 {ee, es} 5 {e, s} 5 H
aH 5 {ae, as} 5 {a, d} 5 dH
bH 5 {be, bs} 5 {b, c} 5 cH.
Therefore, there are exactly three distinct left cosets of H in S(X) and
each of these contain exactly two elements. Also
He 5 {ee, se} 5 {e, s} 5 H
Ha 5 {ea, sa} 5 {a, c} 5 Hc
Hb 5 {eb, sb} 5 {b, d} 5 Hd.
Therefore, again there are exactly three right cosets of H in S(X) and
each of these contain exactly two elements. Note that, even though the
number of left cosets of H is equal to the number of right cosets, a left
coset may not be a right coset and vice versa. For example, aH is not
equal to any right coset.
4. Consider the group (Z24, 124) of integers modulo 24. Let us compute all
the subgroups of Z24 and their cosets. This being an abelian group, any
left coset of a subgroup is a right coset. We know that the subgroups of
Z24 correspond to the positive divisors of 24 which are precisely 1, 2, 3,
4, 6, 8, 12 and 24.
For any divisor d of 24, let
24
Hd 5 {0, d, 2d, …, (q21)d}, where q .
d
Then H1 5 {0, 1, 2, …, 2421}5 Z24
H2 5 {0, 2, 4, 6, 8, 10, 12, 14, 16, 18, 20, 22}
H3 5 {0, 3, 6, 9, 12, 15, 18, 21}
The above eight are all the subgroup of Z24. Coming to the cosets, it
is clear that Z24 is the only coset of H1. There are two cosets of H2
namely
0 24 H 2 ( H 2 ) and 124 H 2 ;
Worked Exercise 4.3.1. Let H be a subgroup of a group G such that there are
exactly two left cosets of H in G. Then prove that every left coset of H in G is
a right coset and vice versa.
Answer: Since H(5 eH) is a left coset, it is given that there is only one more
left coset of H in G and let this be aH. Then, aH H and a H. Now,
aH ∩ H 5 and aH ∪ H 5 G
EXERCISE 4(c)
1. Determine all the subgroups of each of the following and list their cosets.
(i) (Z50, 150)
2. Consider the subgroup Z of the group (R, 1). Prove that there is a bijection
between the set of left cosets of Z in R and the interval [0, 1).
4. For any subgroup H of a group G, prove that aH → Ha21 is a bijection of the set
of left cosets of H in G onto the set of right cosets of H in G.
6. Let X 5 {1, 2, 3} and (S(X), o) be the group of bijections of X onto itself. Let f
S(X) be such that f(1) 5 2, f(2) 5 3 and f(3) 5 1. Then prove that H 5 {e, f,
f 21} is a subgroup of S(X). Compute all the left and right cosets of H in S(X).
7. Let A 5 <2p> be the cyclic subgroup generated by 2p in the group (R, 1).
Prove that the trigonometric functions sine and cosine are constant on any left
coset of A in (R, 1).
8. Let (S(X), o) be the group of all bijections of a nonempty set X onto itself and
x y X. Let
Ax 5 {f S(X) : f(x) 5 x}
and Ax,y 5 {f S(X) : f(x) 5 y}.
Prove that Ax is a subgroup of S(X) and that Ax,y is not a subgroup of S(X).
10. Let A be a subgroup of a group G and x, y and z G such that xyA 5 xzA. Then,
prove that yA 5 zA. Also, prove that Ayx 5 Azx implies Ay 5 Az.
11. Give an example of a subgroup A of a group G such that the product of two left
cosets of A in G is not a left coset of A in G.
12. For any subgroup A of a group G, prove that the only left (right) coset of A in G,
which is also a subgroup of G, is A itself.
13. Let A and B be two subgroups of a group G and x and y G such that Ax 5 By.
Then prove that A 5 B.
14. Prove that a subset S of a group G cannot be a left coset of two distinct sub-
groups of G.
Proof: Since G is a finite set and H is a nonempty subset of G, |G| and |H| are
positive integers. Again, since G is finite, the number of left (right) cosets of
H in G is also finite. Let a1H, a2H, …, anH be all the distinct left cosets of H
in G. By the above theorem,
Corollary 4.4.1. For any subgroup H of a finite group G, the number of left
cosets of H in G is equal to the number of right cosets of H in G which is
|G|
same as .
|H |
Proof: In the proof of the above theorem, we have proved that |G| 5 n|H| and
hence
G
5 n 5 The number of left cosets of H in G.
H
In the above proof, we can replace left cosets with right cosets and prove
similarly that |G| 5 m|H|, where m is the number of right cosets of H in
G. Now,
G
n m. b
H
Definition 4.4.1. For any subgroup H of a finite group G, the number of left
(right) cosets of H in G is called the index of H in G and is denoted by iG(H).
G
iG ( H ) .
H
Note that, even when G is an infinite group, a subgroup can have only
finitely many cosets and one can talk about the index of such a subgroup.
Consider the following example.
Example 4.4.2.
1. Let n be a positive integer and H 5 nZ 5 {na : a Z}. Then, H is a
subgroup of (Z, 1) and, by Example 4.3.1 (2),
H, 1 1 H, 2 1 H, …, (n 2 1)H
Therefore, H , 1 112 H and 2 112 H are all the distinct left cosets of H
in Z12.
3. Consider the example given in Example 4.3.1 (3) in which H 5 {e, s}
and G 5 the group S(X) of bijection of a 3-element set X onto itself.
|G| 6
Here, |G| 5 3! 5 6, |H| 5 2 and hence iG(H) 5 5 3 and therefore
|H | 2
there are exactly 3 distinct left(right) cosets of H in G.
For any element a in group, recall that the order of a is defined as the least
positive integer n such that an 5 e (if all there is one such) and is denoted
by O(a).
Theorem 4.4.3. Let a be an element in a finite group G. Then, O(a) divides |G|.
Proof: First note that, since G is finite and an G for all integers n, an 5 am
for some n m and hence ar 5 e for some positive integer r. Therefore, a
is an element of finite order. Let O(a) 5 n. Then, by Corollary 4.2.1, O(a) is
equal to the order of the subgroup <a> generated by a in G. By the Lagrange’s
Theorem 4.4.2, |<a>| divides |G|. Thus, O(a) divides |G|.
Proof: Let |G| 5 p be a prime number. For any a e in G, O(a) > 1 and O(a)
is a divisor of p. Since p is prime, O(a) 5 p and therefore
Definition 4.4.2. Let n be any positive integer. For any integers a and b, a is
said to be congruent to b modulo n if n divides a 2 b and we denote this by
a b (mod n).
Theorem 4.4.5. Let n be a positive integer. The following holds for any inte-
gers a, b and c.
1. a a (mod n)
2. a b (mod n) ⇒ b a (mod n)
3. a b (mod n) and b c (mod n) ⇒ a c (mod n)
4. a b (mod n) ⇒ a 1 c b 1 c (mod n)
5. a b (mod n) ⇒ ac bc (mod n)
6. a 0 (mod n) ⇔ a is an integral multiple of n.
7. For each a Z, there exists an integer r such that 0 # r , n and a r
(mod n).
8. For any 0 # r s , n, r s (mod n).
Proof: (1) to (6) are direct implications of the above definition. For any a Z,
we can use the division algorithm to get integers q and r such that
a 5 qn 1 r and 0#r,n
and now, a 2 r 5 qn and hence a r (mod n). This proves (7). To prove 8,
consider 0 # r, s , n. Then, |r 2 s| , n and hence r 2 s cannot be a multiple
of n, unless r 5 s.
Recall that, for any n Z1, we have defined f(n) to be the number of
positive integers which are less than or equal to n and relatively prime with n
and that the function f : Z1 → Z1 is called the Euler–Totient function.
af(n) 1 (mod n)
for all integers a which are relatively prime with n, where f is the Euler–
Totient function.
ap a (mod p).
| A || B |
| AB | | BA | .
A∩ B
Proof: First note that it is quite possible that ab 5 a1b1 for distinct elements
a and a1 in A and distinct elements b and b1 in B. We shall find out how often
does an element ab appear as a product of an element in A and an element in
B. For any x A ∩ B, we have
ab 5 (ax) (x21b)
| A || B | | B || A |
| AB | | BA |,
| A∩ B | | B ∩ A|
since A ∩ B 5 B ∩ A.
A B
|G| $ |AB| 5 A B (by the above theorem)
A∩ B
and hence | G | |A| or | G | |B|.
For any subgroup A of a finite group G, we have |A|. iG(A) 5 |G| and hence
both the order and index of A in G are divisors of the order of |G|. In the
following, we derive some more properties of the index of a subgroup of a
finite group. First note that if A and B are subgroups of a group G and A ⊆ B,
then A is a subgroup of B also and we can talk of the index of A in B also.
Worked Exercise 4.4.1. Prove that the following holds for any subgroups
A and B of a finite group G.
1. If A ⊆ B, then iG(A) 5 iB(A) ? iG(B)
2. iA(A ∩ B) # iG(B)
3. iA(A ∩ B) 5 iG(B) if and only if G 5 AB.
4. iG(A ∩ B) 5 iG(A) iG(B) if and only if G 5 AB.
Answer:
1. Suppose that A ⊆ B. Then, by Corollary 4.4.2, we have
G G B
iG(A) 5 . iG ( B) iB ( A)
A B A
A AB G
2. iA(A∩B) 5 iG ( B) .
A∩ B B B
A B
3. Consider |AB| 5 iA ( A ∩ B) B . Now
A∩ B
iA(A∩B) 5 iG(B) ⇒|AB| 5 iG(B)|B| 5 |G| ⇒ G 5 AB
G G G G
iG(A∩B) 5 iG ( A) iG ( B)
A∩ B A B AB A B
and, conversely, if iG(A∩B) 5 iG(A)∙iG(B), then
G G G A B
and hence G AB
A∩ B A B A∩ B
Worked Exercise 4.4.2. Let G be a group of order prn, where p is a prime not
dividing n. Let A and B be subgroups of G of orders pr and ps, respectively. If
B A, prove that AB is not a subgroup of G.
Answer: We are given that |A| 5 pr and |B| 5 ps. Since |B| divides |G| and pr
is the largest power of p dividing |G|, we get that 0 # s # r. Since |A ∩ B| is
a divisor of |A|, |A ∩ B| 5 pt for some t $ 0. Suppose, AB is a subgroup of G,
| A||B|
then |AB| is a divisor of |G| 5 prn. But |AB| 5 , since |A|, |B| and |A ∩ B|
A∩B
are all powers of p, |AB| 5 p for some a $ 0 and a # r (since pr is the largest
a
Worked Exercise 4.4.3. Let A and B be finite subgroups of a group such that
|A| 5 pr and |B| 5 qs, where p and q are distinct primes and r and s are positive
integers. Then prove that A ∩ B 5 {e}.
EXERCISE 4(d)
1. State whether the following are true or false and justify your answer.
(i) There is no subgroup of order 9 in (Z24, 14).
(ii) There is a subgroup of order 36 in (Z120, 1120).
(iii) In any group of order 240, there is a subgroup of index 36.
(iv) If X is a 5-element set, then the group (S(X), o) of bijections of X onto
itself has a subgroup of order 24.
(v) For any positive integer n, there is an element of order n in (R, 1).
(vi) If X is a 6-element set, then the group S(X) has an element of order 27.
(vii) The order of any element in a finite group is finite.
(viii) 43 divides 24221.
(ix) 19 divides 91821.
(x) 8 divides 729421.
2. If A and B are subgroups of a group G such that |A| is a prime and A ∩ B {e},
then prove that A ⊆ B.
4. Prove that the following are equivalent to each other for any subgroup A of a
group G.
(i) iG(A) 5 2.
(ii) x21y A for all x and y G 2 A.
(iii) xy21 A for all x and y G 2 A.
8. Let A and B be subgroups of finite index in a group G such that AB 5 BA. Then
prove that
iAB(A ∩ B) 5 iAB(A) iAB(B).
9. If an abelian group has two subgroups of orders n and m, then prove that it has a
subgroup whose order is the least common multiple of n and m.
10. Let a and b be elements of a group such that a5 5 e and ab21a 5 b2, then find O(b).
12. Let G be an abelian group of order 2n, where n is an odd positive integer. Prove
that G contains exactly one element of order 2.
13. Prove that any group of order 4 is abelian and give an example of a noncyclic
group of order 4.
14. Prove that any nonabelian group has atleast six elements and give an example of
a 6-element nonabelian group.
Theorem 4.5.1. The following are equivalent for any subgroup N of a group G.
1. N is a normal subgroup of G.
2. aN 5 Na for all a G.
3. aNa21 ⊆ N for all a G.
4. Every right coset of N in G is a left coset of N in G.
5. The product of any two left cosets of N in G is a left coset of N in G.
6. (aN) (bN) 5 abN for all a and b in G.
7. (Na) (Nb) 5 Nab for all a and b in G.
8. LN 5 RN (see Theorem 4.3.1); that is, for any a and b G, a21b N if
and only if ab21 N.
Proof:
(1) ⇒ (2): Suppose that N is a normal subgroup of G. Then, for any a G,
aN 5 Nb for some b G, aN 5 Nb for some b G which implies that
a (Nb) ∩ (Na) and hence Nb 5 Na. Therefore, aN 5 Na.
(2) ⇒ (3): If aN 5 Na, then aNa21 5 N.
(3) ⇒ (4): Suppose that aNa21⊆ N for all a G. Then,
aN 5 (aNa21)a ⊆ Na
and, since a21N(a21)21 ⊆ N, we get that a21N ⊆ Na21 and hence Na 5 a(a21N)
a ⊆ a(Na21)a 5 aN. Thus, Na 5 aN for all a G.
(4) ⇒ (5): Suppose that every right coset of N in G is a left coset of N in G.
For any a and b G, first choose c G such that Nb 5 cN and now
(aN)(bN) 5 xN.
a21b N ⇒ b 5 a(a21b) aN
⇒ e 5 ebb21 N (aN) b21 5 (Na)(Nb21) 5 Nab21 (by (7))
⇒ e 5 xab21, for some x N
⇒ ab21 5 x21 N.
x Na ⇔ xa21 N
⇔ x21a N (by (8))
⇔ a21x 5 (x21a)21 N
⇔ x aN.
Example 4.5.1
1. In any group G, the trivial subgroup {e} and G are always normal in G.
2. Consider the group (S(X), o) of all bijections of a 3-element set X onto
itself (see Example 3.4.2). Following the notation given in Example
3.4.2, let H 5 {e, s}. Then, H is a subgroup of S(X). Also,
bsb21 5 bsa 5 ca 5 d H.
Therefore, H is not a normal subgroup of S(X).
Example 4.5.2. Consider the set G 5 {1, 21, i, 2i, j, 2j, k, 2k}. Define the
binary operation on G as just multiplication of numbers, treating i, j and k as
numbers satisfying the following rules.
i2 5 j2 5 k2 5 21
ij 5 k and ji 5 2k
jk 5 i and kj 5 2i
ki 5 j and ik 5 2j
i
k j
Worked Exercise 4.5.1. Prove that any subgroup of index 2 in any group is
normal.
Worked Exercise 4.5.2. Prove that every subgroup of the Quaternion group
of 8 elements is normal.
A 5 {1, 21}.
xA 5 {x, 2x} 5 Ax
Worked Exercise 4.5.3. Prove that the centre of any group is normal.
Answer: Recall that, for any group G, the centre of G is defined as the set
Thus, for any x G, xZ(G) 5 Z(G)x and hence Z(G) is a normal subgroup of G.
Worked Exercise 4.5.4. Let X be any nonempty set X and (S(X), o) be the
group of bijections of X onto itself. For any x X, let
Then prove that Hx is a normal subgroup of S(X) for all x X if and only if X
has at most two elements.
Answer: Suppose that X has three distinct elements, say x, y and z. Let f and
g : X → X be defined by
On the other hand, suppose X has at most two elements. If |X| 5 1, then
S(X) 5 {e} 5 Hx and hence Hx is normal. If |X| 5 2, then X 5 {x, y}, where
x y, and Hx 5 {e} 5 Hy, which is clearly normal.
EXERCISE 4(e)
2. For any normal subgroup A of a group G and for any subgroup B of G, prove that
A ∩ B is a normal subgroup of B.
i 0 0 i
, , i 0 and 0 i
0 i i 0 0 i i 0
Prove that G is a group under the multiplication of matrices over the complex
numbers. Also prove that G is not abelian and that every subgroup of G is nor-
mal. Compare this with the example given in Worked Exercise 4.5.2.
4. Determine all the normal subgroups of the group S(X) of bijections of a 3-ele-
ment set X onto itself.
10. Let A be a normal subgroup of a finite group G such that |A| and iG(A) are rela-
tively prime. For any x G, prove that x A if and only if x|A| 5 e.
N ∩ ( a A a1 ).
a∈G
16. Let A be a subgroup of a group G such that x2 A for all x G. Then prove that
A is normal in G.
17. For any real numbers a and b, define Tab : R → R by Tab(x) 5 ax1b for all x
R. Prove that the set {Tab : a and b R} is a group under the usual composition
of mappings in which {T1b : b R} is a normal subgroup.
18. Prove that the intersection of any class of normal subgroups of a group G is
again a normal subgroup of G.
19. Prove that, for any group G, the centre Z(G ) ∩ N (a), where N(a) is the
a ∈G
normaliser of a in G.
20. For any set X with |X| $ 3, prove that Z(S(X)) 5 {e}.
21. Let G1 and G2 be any groups and A1 and A2 be subsets of G1 and G2 respectively.
Then prove that A1 3 A2 is a (normal) subgroup of the group G1 3 G2 if and only
if A1 and A2 are (normal) subgroups of G1 and G2, respectively.
22. If G1 and G2 are groups, prove that {e1} 3 G2 and G1 3 {e2} are normal sub-
groups of G1 3 G2, where e1 and e2 are identities in G1 and G2, respectively.
Then, AB is called the product of A and B, in this order, induced by the binary
operation on G. Then, ? is a binary operation on the set P(G) of all subsets of
G and is called the multiplication of subsets of the group G.
Then, G/N is a group under the multiplication of subsets of the group; that is,
for any aN and bN G/N,
Proof: First of all recall that, aN∙bN 5 abN for all a and b G (since N is a
normal subgroup of G) and hence the multiplication of subsets of the group
G is a binary operation on the set G/N of all left cosets of N in G.
For any a, b and c G, we have
(aN?bN)?cN 5 (abN)?cN
5 ((ab)?c)N
5 (a(bc))N
5 aN?(bN?cN)
for all aN G/N and hence eN (5 N) is the identity in G/N. Further, for any
aN G/N with a G, consider
and therefore a21N is the inverse of aN in G/N. Thus, G/N is a group under the
multiplication of subsets of the group G.
Definition 4.6.2. For any normal subgroup N of a group G, the group G/N
constructed above is called the quotient group of G by N. Whenever we refer
to a G/N as a group, we only mean the set of left(right) cosets of N in G
together with the multiplication of subsets of the group G.
Proof: This follows from the facts that |G/N| 5 iG(N) and |G| 5 |N|?iG(N).
Example 4.6.1. Let (Z, 1) be the group of all integers under the usual addi-
tion and let n be any positive integer and <n> 5 {na : a Z} be the subgroup
of (Z, 1) generated by n. Since 1 is the operation on the group Z, the ele-
ments of Z/<n> are of the form a 1 <n> with a Z. From Example 4.3.1
(2), we know that any coset of <n> in Z is of the form r 1 <n>, where 0 #
r , n. Thus,
Example 4.6.2. Let E be the Euclidean plane and pick any point P in E and
fix a coordinate system with P as origin. Then, any point in E can be expressed
uniquely as an ordered pair (x, y) of real numbers and P corresponds to (0, 0).
For any points Q1 5 (x1, y1) and Q2 5 (x2, y2), define
Q1 + Q2
Q2(x2,y2)
Q1(x1,y1)
Let L be a straight line in E passing through P. Then, one can easily check
that L is a subgroup of E. Let us describe the cosets of L in E. If Q is any point in
E, then one can prove that the coset Q 1 L is precisely the line passing through
Q and parallel to L. Therefore, the quotient group E/L consists precisely all the
straight lines parallel to L, including L. For any points Q1 and Q2 in E, we have
Q2
Q1 + Q2
Q1
L Q1 + L Q2 + L Q1 + Q2 + L
[a, b] 5 ab a21b21.
One can easily verify that ab 5 [a, b]ba and hence we can view the com-
mutator [a, b] as a measure of the extent to which ab differs from ba. In fact,
the elements a and b commute (that is, ab 5 ba) if and only if [a, b] 5 e,
the identity in G. The commutators of elements of a group G may not form a
subgroup of G, in general. However, we have the following theorem.
n
[G , G ] ∏[ai , bi ] : ai and bi ∈G .
i1
and therefore [G, G] is precisely the subgroup of G generated by the set {[a, b] :
a and b G} (see Theorem 4.1.5). Thus, [G, G] is a subgroup of G. Next, let
A be any subgroup of G such that [G, G] ⊆ A. For any a A and x G,
xax21 5 (xax21a21)a A,
Definition 4.6.4. For any group G, the subgroup [G, G] is called the derived
subgroup or commutator subgroup of G and the quotient group G/[G, G] is
called the commutator quotient group or abelianized group. The reason for
the latter terminology is the following theorem.
a A ⇒ aN A/N 5 B/N
⇒ aN 5 bN for some b B
⇒ a21b N
⇒ ab21 N ⊆ B (since N is normal)
⇒ a 5 (ab21) b B
Remark 4.6.1. Note that, in the above, A/N is normal in G/N if and only if A
is normal in G. Also, for any subgroups A and B containing N, A/N ⊆ B/N if
and if only A ⊆ B.
If N 5 {e}, then G/N G and, if N 5 G, then G/N 5 {N}, the trivial
group.
EXERCISE 4(f)
1. Describe the quotient group of each of the following in the groups mentioned
against them.
(i) {0, 4, 8, 12} in (Z16, 116).
(ii) The set E of even integers in (Z, 1).
(iii) Z in (Q, 1).
(iv) {1, 21} in ({1, 21, i, 2i}, ∙).
(v) R in (C, 1).
(vi) Q in (R, 1).
3. For any group G, determine the quotient groups of the trivial normal subgroups
{e} and G.
4. Let A be a subgroup of a group G such that x2 A for all x G. Then prove that
A is normal in G and the quotient group G/A is abelian.
7. Let N be a normal subgroup of a group G. Then prove that G is finite if and only
if both N and G/N are finite.
9. Let G be a group and S 5 {a2 : a G}. Then prove that <S > is a normal sub-
group of G and that G/<S > is an abelian group.
10. List all normal subgroups of the group (S(X), o) of bijections of a 3-element set
X onto itself and construct tables representing the quotient group of each normal
subgroup S(X).
11. Let G 5 {(a, b) R 3 R : a 0} and, for any (a, b) and (c, d) G, define
(a, b) * (c, d) 5 (ac, ad 1 b).
Then prove that (G, *) is a group. If K 5 {(1, b) : b R}, then prove K is a
normal subgroup of G.
12. Let G be a group of order 2n, where n is odd. Prove that G contains a normal
subgroup of index 2.
and hence these two groups look same as we have the correspondence f : Z4 → G
defined by f (0) 5 0, f (1) 5 i, f (2) 5 i2 and f (3) 5 i3. This correspondence is
compatible with the operations 14 on Z4 and ‘?’ on G and is a bijection of Z4
onto G. This helps us in proving that any property of Z4 gives a similar prop-
erty in G9. Such correspondences are called homomorphisms.
GxG * G
fxf f
G′ x G′ G′
ο
where * and o are the binary operations in the groups G and G9, respectively.
Examples given below make these things clear.
Example 5.1.1
1. Consider the groups (Z, 1) and (R1, ?) where 1 and ‘?’ are the usual
addition and multiplication of real numbers. Let m be any positive inte-
ger and define f : Z → R1 by
f (a) 5 ma for all a Z.
0, if a is even
f : Z → Z2 by f ( a)
1, if a is odd
for any a Z. Then, since a 1 b is even if and only if both a and b are
even or both a and b are odd, we get that
f (a 1 b) 5 f (a) 12 f (b) for all a and b G.
ar bt asbu
f ( AB) f
crdt csdu
( ar bt ) (cs du ) ( as bu ) (cr dt )
arcs ardu btcs btdu ascr asdt bucr budt
ardu btcs asdt bucr
( ad bc) ( ru st )
f ( A) f ( B).
Proof:
1. Let f (e) 5 x. Then, x is an element in G9 and hence
xe9 5 x 5 f (e) 5 f (ee) 5 f (e)f (e) 5 xx.
Therefore, xx 5 xe9 and, by the left cancellation law, x 5 e9. Thus,
f (e) 5 e9.
2. For any a G, we have
f (a)f (a21) 5 f (aa21) 5 f (e) 5 e9 5 f (e) 5 f (a21a) 5 f (a21) f (a)
and hence f (a21) is the inverse of f (a) in G9; that is,
f (a21) 5 f (a)21.
Proof:
1. Let A be a subgroup of G. Then,
f (A) 5 {f (a) : a G} ⊆ G9.
First of all, since A is a subgroup of G, we have e A and hence
e9 5 f (e) f (A)
so that f (A) is a nonempty subgroup of G9. Also,
x and y f (A) ⇒ x 5 f (a) and y 5 f (b) with a and b A.
⇒ xy21 5 f (a)f (b)21 5 f (a) f (b21) 5 f (ab21)
f (A) (since A is a subgroup).
Proof:
1. Let f be a surjection and A be a normal subgroup of G. Then, we have
already proved that f (A) is a subgroup of G9. Now,
x f (A) and y G9 ⇒ x 5 f (a) and y 5 f (b) for some a A
and b G
⇒ y x y21 5 f (b)f (a)f (b)21 5 f (bab21) f (A).
Remark 5.1.1. Note that in Theorem 5.1.3 (1), it is necessary that f is a sur-
jection; for, consider a group G9 and a subgroup G of G9 such that G is not
normal in G9. Let f : G → G9 be the inclusion map defined by f (x) 5 x for all
ker f 5 f21({e9}),
Theorem 5.1.5. Let N be a normal subgroup of a group G and G/N, the quo-
tient group of G by N. Define
f (ab) 5 abN
5 aN ? bN (since N is normal in G)
5 f (a)f (b)
Example 5.1.2
{Z, if m0
{0}, if m0
.
6. f : Z → Z2 is defined by f ( a)
{ 0, if a is even
1, if a is odd
.
Example 5.1.3
1. If m > 1, then the map f : Z → R1, defined by f (a) 5 ma, is a monomor-
phism of (Z, 1) into (R1, ?).
2. The homomorphism given in (2), (3) and (5) (if m 0) of Example
5.1.2 are monomorphisms, while those given in (2), (3), (6) and (8) are
epimorphisms.
Note that a homomorphism f : G → G9 is an epimorphism if and only if
f (G) 5 G9. In the following, we give a characterization of monomorphisms
in terms of their kernels.
and hence ker f contains e alone. Thus, ker f 5 {e}. Conversely, suppose that
f is not a monomorphism, then f is not an injection and hence there exists
a b G such that f (a) 5 f (b). Now, ab21 e and
f o g 5 IB and g o f 5 IA;
that is, f (g(b)) 5 b for all b B and g(f (a)) 5 a for all a A.
This unique g is called the inverse of f and is denoted by f21. Also, in this case,
for any a A and b B, we have
f (a) 5 b ⇔ a 5 f21(b).
Theorem 5.1.8. The following holds for any groups G, G9 and G:
1. G G
2. G G9 ⇒ G9 G
3. G G9 and G9 G ⇒ G G
Proof:
1. follows from the fact that the identity map IG on G is an isomorphism of
G onto itself.
2. is consequence of Theorem 5.1.7 and
3. follows from the fact that the composition of two isomorphisms is again
an isomorphism.
Example 5.1.4. Consider the groups (Z4, 14) and (G, ?), where G 5 {1, 21, i,
2i} and ‘?’ is the multiplication of complex numbers. Define f : Z4 → G by
Therefore, g o f is a homomorphism.
aK 5 {ax : x K}
5 {ax : f (x) 5 e9}
5 {y G : f (y) 5 f (a)},
since f (ax) 5 f (a)f (x) 5 f (a)e9 5 f (a) and, if y G is such that f (y) 5 f (a), then
y 5 ax, where x 5 a21y ker f 5 K. Thus, aK 5 f21{f (a)} for any a G.
a b ( q q)n ( r s)
( q q)n ( r s), if r s n
( q q1)n ( r s n), if r s n
5 tn 1 (r 1n s) and 0 r 1n s < n
Worked Exercise 5.1.5. Let G be an abelian group of order m and let n be any
positive integer relatively prime to m. Define f : G → G by f (a) 5 an for all a
G. Then, prove that f is an automorphism of G.
am 5 e for all a G.
and hence |f (G)| 5 |G| so that f (G) 5 G (since G is finite and f (G) ⊆ G).
Therefore, f is a surjection also. Thus, f is an isomorphism of G onto G; that
is, f is an automorphism.
EXERCISE 5(a)
3. Let C 2 {0} and R 2 {0} be the groups of nonzero complex numbers and nonzero
real numbers, respectively, under the usual multiplications. Prove that the map
f : C 2 {0} → R 2 {0} defined by f (z) 5 |z| is a homomorphism.
4. Let G be a group and define f : G → G by f (a) 5 a21 for any a G. Prove that f
is an endomorphism if and only if G is an abelian group.
5. Determine the Kernels of the homomorphisms (if they are) given in Exercise 1
above.
6. Determine all the homomorphisms of (Z, 1) into (Z2, 12).
7. Determine all the endomorphisms of the group (Z, 1) into itself.
8. Prove that every nontrivial endomorphism of (Z, 1) into itself is a
monomorphism.
9. Prove that there is no epimorphism of (Z, 1) onto itself, except the identity map.
10. Consider the groups (Z, 1) and (R, 1) and, for any real number a, define fa : Z → R
by fa(x) 5 ax for all x Z. Prove that a mapping f : Z → R is a homomorphism
if and only if f 5 fa for some a R.
11. Let G and G9 be finite groups and f : G → G9 be a homomorphism. Prove that
the index of the Kernel of f in G is a divisor of |f (G)|.
12. Let f : G → G9 be a homomorphism of groups and a G. If O(a) is finite, then
prove that O(f (a)) is also finite and is a divisor of O(a). Give an example where
O(f (a)) is a proper divisor of O(a).
13. Let f and g : G → G9 be homomorphisms of groups and A 5 {a G : f (a) 5
g(a)}. Then, prove that A is a subgroup of G.
14. Let f : G → G9 be a homomorphism and G be a finite group of prime order. Then,
prove that f is either trivial or a monomorphism.
15. Let f : G → G9 be a homomorphism of groups and [G, G] be the commutator
subgroup of G (see 4, 6, 12). Then, prove that f (G) is an abelian group if and
only if [G, G] ⊆ ker f.
16. Let G be a group and a and b G. Consider the group Z 3 Z under coordinate-
wise addition and define f : Z 3 Z → G by
f (m, n) 5 ambn for any (m, n) Z 3 Z.
Obtain a necessary and sufficient condition, in terms of a and b, for f to be a
homomorphism.
G/ker f f (G)
G/ker f G9.
aK 5 bK ⇒ a21b K 5 ker f
⇒ f (a)21f (b) 5 f (a21)f (b) 5 f (a21b) 5 e9
⇒ f (a) 5 f (b)
This clears the ambiguity and it follows that g is well defined. For any aK and
bK G/K, we have
h
G g
→ G / K →G′
Example 5.2.1. Let n be a positive integer and (Zn, 1n) be the group of inte-
gers modulo n. As in Worked Exercise 5.1.4, define f : Z→Zn by f (a) 5 r,
where r is the remainder obtained by dividing a with n; that is, if q and r are
integers such that a 5 qn 1 r and 0 r < n, f (a) 5 r.
In Worked Exercise 5.1.4, we have proved that f is an epimorphism. By the
Fundamental Theorem of Homomorphisms, Z/ker f Zn, we have
Then, f1 and f2 are epimorphisms and ker f1 5 N1 and ker f2 5 N2; For
for any x G since a1, b1, a2, b2 are all homomorphisms, is also a homo-
morphism. Also, for any x G,
Therefore, ker 5 {e} and hence is a monomorphism. For any (x1, x2)
G13G2, choose a1, a2 G such that a1(a1N1)5 x1 and a2(a2N2) 5 x2. Since N1
and N2 are normal subgroups and N1 ∩ N2 5 {e}, we get that ab 5 ba for all
a N1 and b N2 (for, consider aba21b21 N1 ∩ N2 5 {e}).
Now since a1 and a2 G 5 N1N2, we get that
and hence
aN 5 xN, Also,
a21y 5 (st)21(tu) 5 (t21s21t) u M (since s, u M)
and hence aM 5 yM. Therefore, f (a) 5 (aN, aM) 5 (xN, yM). Thus, f is an
epimorphism. By the Fundamental Theorem of Homomorphisms,
Answer: Consider the group (Z, 1) of integers and let N 5 nZ and M 5 mZ.
Since (n, m) 5 1,
In the following, we shall classify the cyclic groups and prove that (Z, 1)
and (Zn, 1n) are the only (up to isomorphism) cyclic groups.
Z/ker f G.
Since ker f is a subgroup of (Z, 1), ker f 5 nZ for some nonnegative integer
n (by Worked Exercise 4.1.1),
Worked Exercise 5.2.2. For any positive integers n and m, prove that the fol-
lowing are equivalent to each other.
1. n and m are relatively prime.
2. Zn 3 Zm is cyclic.
3. Zn 3 Zm Znm
Answer: (1) ⇒ (3) is proved in Worked Exercise 5.2.1 and (3) ⇒ (2) is triv-
ial, since Znm is cyclic and, for any group G and G9 such that G G9, G is
cyclic if and only if G9 is cyclic. We are left with only (2) ⇒ (1). Assume that
Zm 3 Zn is cyclic. Since the order of Zm 3 Zn is mn, it follows from Theorem
5.2.5 that
Zm 3 Zn Zmn.
k 5 mn 5 kd
where d is the g.c.d. of m and n. Therefore, d 5 1; that is, m and n are rela-
tively prime. b
EXERCISE 5(b)
2. For any positive integers m and n whose least common multiple is k, prove that
Zk is isomorphic to a subgroup of Zn 3 Zm.
5. For any finite group G, prove that there exists a prime number p such that Zp is
not a homomorphic image of G.
6. Prove that the order of any homomorphic image of a finite group G must be a
divisor of the order of G.
M/M ∩ N MN/N.
M/ker f MN/N.
5 {m M : mN 5 N}
5 {m M : m N} 5 M ∩ N
Example 5.3.1. Consider the group (Z, 1) of integers and let M 5 <3> 5 3Z
and N 5 <5> 5 5Z. Since 1 is commutative on Z, M and N are normal
subgroups of Z. Also,
M ∩ N 5 15Z and M 1 N 5 Z
3Z/15Z Z/5Z.
A closer examination of the cosets and operation in the quotient group 3Z/15Z
reveals this to be none other than the group ({0, 3, 6, 9, 12}, 115). The above
isomorphism is a disguised version of the isomorphism
While applying the above theorem, one frequently starts with a normal
subgroup of G9 and use its inverse images, rather than starting with a normal
subgroup of G containing the kernel. In this context, recall the one-to-one
correspondence between normal subgroups of G containing the kernel and the
normal subgroups of G9. In view of this, the above theorem can be rephrased
as follows.
G/f2l(M) G9/M.
f (N) 5 f (f2l(M)) 5 M.
The following is another special case which is of interest on its own and
is popularly called the Third Isomorphism Theorem. The reader is cautioned
that there seems to be no universally accepted agreement on the numbering
of these three Isomorphism theorems. However, the Fundamental Theorem
of Homomorphisms deserves to be called as the First Isomorphism Theorem,
since the other two are proved using this.
Therefore,
(G / M ) G
.
(N / M ) N
EXERCISE 5(c)
f1
G G1
f2 f
G2
3. Let G be the group of nonzero real numbers under the usual multiplication and N
5 {1, 21}. Then, prove that N is a normal subgroup of G and the quotient group
G/N is isomorphic to the group of positive real numbers under multiplication.
5. Let G8 be the group of symmetries of a square (see Example 3.2.8) and define
f : G8 → Z2 3Z2 by
f (e) 5 f (r2) 5 (0, 0), f (r1) 5 f (r2) 5 (1, 0),
f (h) 5 f (0) 5 f (0, 1) and f (d1) 5 f (d2) 5 (1, 1).
Prove that f is an epimorphism and deduce that
G8/Z(G8) Z2 3 Z2, where Z(G8) is the centre of G8.
7. For any (a, b) R 3 R with a 0, define Tab : R→R by Tab (x) 5 ax 1 b for all
x R and let
G {Tab : ( a, b) ∈ R R and a 0}
and N 5 {T1b: b R}.
Prove that G is a group under the composition of mappings and N is a normal
subgroup of G. Further, prove that the quotient group G/N is isomorphic to the
group of nonzero real numbers under the multiplication.
iG ( A) i ( f ( A))
G
iG ( f 1 ( A)) = i ( A).
G
11. Prove that any group of order 4 is isomorphic to either Z4 or Z2 3 Z2 and hence
is abelian.
12. Prove that any group of prime order p is isomorphic to Zp and hence is cyclic.
f
G A
g f
16. Let A1, A2 and N be subgroups of a group G such that N is normal in G and A1N
5 A2N. Then, prove that A1 / A1 ∩ N A2 / A2 ∩ N .
5.4 AUTOMORPHISMS
Let us recall that a homomorphism of a group G into itself is called an
endomorphism of G and a bijective endomorphism of G is called an auto-
morphism of G. Among the endomorphisms of a group G, the automor-
phisms of G need special attention, for the reason that they form a group
on their own under the composition of mappings and that the structure of
this group reveals that of the group G itself. Even though the following is a
repetition, we prefer to give an independent status for convenience and for
its importance.
Theorem 5.4.1. For any group G, the set Aut(G) of all automorphisms of G
forms a group under the composition of mappings.
f ο IG f IG ο f .
f o f21 5 IG 5 f21 o f.
Example 5.4.1
1. If G is an abelian group, then the map f : G → G, defined by f (x) 5 x21
for any x G, is an automorphism of G.
2. Consider the group (Z12, 112) of integers modulo 12 and define f : Z12 →
Z12 by
f ( a) 5a ( a 12 a 12 a 12 a 12 a).
Note that f (a) 5 r, where 5a 5 12q 1 r, where q and r are integers and
0 r < 12.
Then, f is clearly an endomorphism of Z12
Also,
ker f {a ∈ Z12 ; f ( a) 0}
{a ∈ Z12 ; 5a 12q, q ∈ Z}
{0} (since (5, 12) 1)
and therefore f is a monomorphism of Z12 into Z18 .Since Z12 is a finite set, it
follows that f is an surjection also. Thus, f is an automorphism of Z12.
In the following, we discuss about a special subgroup of Aut(G) consisting
of certain special class of automorphisms which are important in the case of
an abelian group.
Ta ( x ) Ta ( y ) ⇒ axa1 aya1 ⇒ x y
G / Z (G ) G / ker a a(G ) I (G ).
It can be easily seen that a group G is abelian if and only if Ta(x) 5 x for
all a and x G (that is, Ta 5 IG for all a G and the group I(G) is trial). If G
is a nonabelian group, then there exists an automorphism Ta IG. In the fol-
lowing, we prove that a group G has a nonidentity automorphism if and only
if G has atleast 3 elements.
Proof: Suppose that |Aut(G)| > 1. Then, the group Aut(G) has an element
other than its identity. That is, there exists an automorphism f of G such that f
IG, the identity map. Now, we can choose an element a G such that f (a)
IG(a) 5 a. Since f (e) 5 e, it follows that a e and, since f is injective,
f (a) f (e) 5 e. Therefore, e, a and f (a) are three distinct elements in G and
hence |G| > 3 > 2.
Conversely suppose that |G| > 2. If G is not abelian, then ax xa for some a
and x G and hence
Ta(x) 5 axa21 x
Note that (A, f ) and hence is a nonempty set. For any (B, g) and (C,
h) G, define
Proof: Let G 5 <a> be a cyclic group. Then, by the above theorem f f (a)
is an injection of Aut(G) into the set gen(G) of generators of G. Further, if b
is any generator of G, then we can define automorphism f such that f (a) 5 b
(that is, f (an) 5 bn for any n Z). Thus, f f (a) is a bijection of Aut(G) onto
gen(G). From Theorems 4.2.7 and 4.2.8, gen(G) is a finite set and hence so is
Aut(G) and |Aut(G)| 5 |gen(G)|.
Corollary 5.4.2
1. For any infinite cyclic group G, |Aut(G)| 5 2.
2. For any finite cyclic group G of order n, |Aut(G)| 5 f (n), where f is the
Euler-Totient function.
Proof: These follow from Corollary 5.4.1, Theorems 4.2.7 and 4.2.8.
(f o g) 5 a o (f o g) o a21
5 (a o f o a21) o (a o g o a21)
5 (f ) o (g).
Worked Exercise 5.4.2. List all the automorphisms of the group (Zn, 1n) for
any positive integer n and, in particular, of the group (Z12, 112).
Consider Z12. We have f (n) 5 4 since 1, 5, 7 and 11 are the only integers
r such that 1 r < 12 and (r, 12) 5 1. Therefore, there are exactly four auto-
morphisms of Z12 and are given below.
f1 5 I, the identity map of Z12,
f5 : Z12 → Z12 defined by f5(m) 5 5m (mod 12),
f7 : Z12 → Z12 defined by f7(m) 5 7m (mod 12),
f11 : Z12 → Z12 defined by f11(m) 5 11m (mod 12).
The following table gives a complete description of all the four automor-
phisms of Z12.
0 1 2 3 4 5 6 7 8 9 10 11
f1 0 1 2 3 4 5 6 7 8 9 10 11
f5 0 5 10 3 8 1 6 11 4 9 2 7
f7 0 7 2 9 4 11 6 1 8 3 10 5
f11 0 11 10 9 8 7 6 5 4 3 2 1
EXERCISE 5(d)
1. For any endomorphism f of a finite group G, prove that the following are equiva-
lent to each other:
(i) f is an epimorphism.
(ii) f is an automorphism of G.
(iii) f is a monomorphism.
2. Give an example of an infinite group G and of an isomorphism of G onto a
proper subgroup of G.
3. Let G 5{e, a, b, ab} be a group of order 4 in which a2 5 e 5 b2 and ab 5 ba.
Then, determine Aut(G).
4. Let A be a subgroup of a group G, such that f (A) ⊆ A for all f Aut(G). Then,
prove that A is a normal subgroup in G.
5. Let G be a group and f Aut(G). Prove that the set {a G : f (a) 5 a} is a sub-
group of G.
6. For any group G, prove that {a G : f (a) 5 a for all f Aut(G)} is a normal
subgroup of G.
7. Let G be a finite group and f Aut(G), such that, for any xG,
f (x) 5 x if and only if x 5 e.
9. Let G be a finite group and f Aut(G) such that f sends more than three-quarters
of the elements of G onto their inverses. Then, prove that f (a) 5 a21 for all a
G and that G is abelian.
10. Let C be the commutator subgroup of a group G and f Aut(G). Prove that
f (C) ⊆ C.
11. If a group G has a nonidentity automorphism, then prove that G has atleast three
elements.
15. Find all the automorphisms of (Z, 1) and (Zn, 1n) for any n Z1.
16. Let G be a finite cyclic group of order n and define fm : G → G by fm(a) 5 am for
all a G and m Z1. Prove that fm is an automorphism of G if and only if m is
relatively prime with n.
17. Let G be a finite group of order n > 2. If a2 e for some a G, then prove that
G has a nonidentity automorphism.
18. If G is a noncyclic finite abelian group, prove that Aut(G) is not abelian.
19. Let G be a finite group, such that |Aut(G)| 5 p, where p is a prime number. Then,
prove that |G| 3.
For any nonempty set X, the set M(X) of all mappings of X into itself forms a
monoid under the composition of mappings in which the invertible elements
(elements possessing inverses) are precisely the bijections of X onto itself.
In fact, we have observed that a mapping f : X → X has left (right) inverse in
M(X) if and only if f is an injection (respectively, surjection) and, as such, the
set of all bijections of X onto itself forms a group under the composition of
mappings. Before the advent of the abstract form of a group, mathematicians
were only interested in the group structure of certain sets of bijections, which
were also known as permutations, when the set X is finite. In this chapter, we
discuss thoroughly the structure of this type of groups.
G9 5 {fa : a G}.
for all x G and hence fab 5 fa ? fb. Also, if e is the identity in the group G,
and hence fa21 5 f a−1 G. Therefore, G9 is a subgroup of the group (S(G), o);
that is, G9 is a group of permutations on the set G. Now, define
fa 5 fb ⇒ fa(e) 5 fb(e) ⇒ a 5 ae 5 be 5 b
gx : X → X by gx(aH) 5 (xa)H
aH 5 bH ⇒ a21b H
⇒ (xa)21(xb) 5 a21x21xb 5 a21b H
⇒ (xa)H 5 (xb)H
gx((x21b)H) 5 (x(x21b))H 5 bH
Theorem 6.1.3. Let H be a subgroup of a finite group G such that |G| is not a
divisor of iG(H)! Then, H contains a nontrivial normal subgroup of G.
Proof: Let X be the set of all left cosets of H in G. Then, |X| 5 iG(H). By the
above theorem, there exists a homomorphism : G → S(X) such that ker is
Worked Exercise 6.1.2. Let G be a group of order 187. Prove that any sub-
group of order 17 in G must be normal.
| G | 187
iG ( A) 11.
17 17
Since 17 is a prime and 17 . 11, we get that 17 does not divide 11! 5 iG(A)!
and hence |G| does not divide iG(A)! By Theorem 6.1.3, A contains a non-
trivial normal subgroup of G. Let N be a nontrivial normal subgroup of G
contained in A. Then, 1 , |N| and, by Lagrange’s theorem, |N| is a divisor of
|A| 5 17. Since 17 is a prime, |N| 5 17 5 |A| and hence N 5 A. Thus, A is a
normal subgroup of G.
EXERCISE 6(a)
3. If X is a finite set with |X| 5 n, prove that S(X) is a finite group of order n!
4. For any set X, prove that X is finite if and only if S(X) is finite.
6. Let A be a subgroup of a finite group G and iG(A) 5 m. If A does not contain any
nontrivial normal subgroup of G, then prove that |G| divides m!
7. Let G be a group of order 396. Prove that any group of order 11 in G is normal
in G.
8. Let G be a finite group of order n and p be a prime number such that p2 does not
divide n. Prove that any subgroup of order p in G is normal in G.
9. For any elements a and b in a set X, prove that there is a permutation f on X such
that f (a) 5 b and f (b) 5 a.
The above theorem says that two finite sets X and Y are equipotent if and only
if the groups (S(X), o) and (S(Y), o) are isomorphic. In particular, if X is a set
with n elements, then S(X) S(In), where In 5 {1, 2, …, n}. For this reason,
it is enough if we study the permutations on the set {1, 2, …, n}. We begin
this with the following formal definition.
In 5 {1, 2, …, n}
1 2 3 … n
f
f (1) f (2) f (3) … f ( n)
which symbolizes that each 1 # i # n is mapped onto f (i), the integer that is
written just below i in the array. As usual, let e denote the identity in the group
Sn. Note that e is the identity mapping on In.
Example 6.2.1
1. Consider I6 5 {1, 2, 3, 4, 5, 6} and define f : I6 → I6 by f (1) 5 3, f (2) 5 5,
f (3) 5 1, f (4) 5 2, f (5) 5 4 and f (6) 5 6. Then, f is denoted by
1 2 3 4 5 6
f .
3 5 1 2 4 6
2. Let f S9 be given by
1 2 3 4 5 6 7 8 9
f .
4 6 8 5 2 3 7 9 1
f (i ), if 1 i m
( f )(i ) .
i, if m i n
Answer: There are 3! (56) elements in the group S3, which are given below
(recall Example 3.4.3).
o e a b c d s
e e a b c d s
a a b e s c d
b b e a d s c
c c d s e a b
d d s c b e a
s s c d a b e
EXERCISE 6(b)
1. Consider the following elements in S8 and compute the expressions in (i) to (viii)
given below.
1 2 3 4 5 6 7 8
a
3 7 4 6 8 5 1 2
1 2 3 4 5 6 7 8
b
4 5 6 7 8 1 2 3
and
1 2 3 4 5 6 7 8
c
8 7 6 5 1 2 3 4
(i) a2b
(ii) ab2
(iii) abc
(iv) ab2c
(v) a2bc
(vi) abc2
(vii) b2ca
(viii) c3a
2. For any positive integer n, prove that the order of any element in Sn is finite and
is a divisor of n!
4. Determine all the elements in the cyclic subgroups ,a., ,b. and ,c., where
a, b and c are as given above.
8. Compute aba21, bcb21 and cac21 for the elements a, b and c given in Exercise 1.
6.3 CYCLES
Consider the permutation f in S5 given by f (1) 5 3, f (2) 5 1, f (3) 5 5, f (4) 5 2
and f (5) 5 4. In the array form, f can be expressed as
1 2 3 4 5
f .
3 1 5 2 4
1 3 5 4 2
f ↓ ↓ ↓ ↓ ↓
3 5 4 2 1
2 3
4 5
Definition 6.3.1. Let n be a positive integer and i1, i2, …, ir be distinct ele-
ments in the set In 5 {1, 2, …, n}. Define f : In → In by
i j1 , if i i j , 1 j r
f (i ) i1 , if i ir .
i, if i i j , 1 j r
That is, f (i1) 5 i2, f (i2) 5 i3, …, f (ir21) 5 ir, f (ir) 5 i1, and f (i) 5 i for all
i {i1, i2, …, ir}.
The action of f is cyclic on the set {i1, i2, …, ir} and f is identity on the comple-
ment of {i1, i2, …, ir}.
i1
ir
i2
ir1 i3
Example 6.3.1
1. a 5 (2 5 3 4 6) is a 6-cycle in S6 and hence in Sm for any m 6. a is
defined by
a(2) 5 5, a(5) 5 3, a(3) 5 4, a(4) 5 6, a(6) 5 2
and a(i) 5 i for all i {2, 3, 4, 5, 6}.
1 2 3 4 5 6 7 8
2. Let a 5 . Let us express this in cyclic form.
3 6 5 7 2 4 8 1
We have a(1) 5 3, a(3) 5 5, a(5) 5 2, a(2) 5 6, a(6) 5 4, a(4) 5 7,
a(7) 5 8 and a(8) 5 1. Therefore,
a 5 (1 3 5 2 6 4 7 8)
The example given in (2) above can be extended to any r-cycle as given
in the following.
Theorem 6.3.1. Let a be a cycle in Sn and a(i) (i) for some i In. Then,
a 5 (i c(i) c2(i) … cr21(i)) for some r . 1.
Proof: Let a 5 (i1 i2 … ir). Since a(i) i, we get that i 5 ik for some 1 # k
# r. Then,
and ar(i) 5 i.
From the above theorem, it follows that ar(ik) 5 ik for all 1 # k # n and hence
ar coincides with the identity permutation. Therefore, ar 5 e. Also, for any
1 # k , r,
ak(i1) 5 ik11 i1
and hence ak e for all 1 # k , r. Thus, r is the smallest positive integer such
that ar 5 e and hence the order of a is r, which gives (1).
(2) Since ar 5 e, we get that a21 5 ar21. By the above theorem, we get that
ar21(ir) 5 ir21
ar21(ir21) 5 ir22, etc.
and hence ar21 5 (ir ir21 … i2 i1).
(3) This is clear, since O(a) 5 r. b
Note: If a 5 (i1 i2 … ir) is an r-cycle, then ak(ij) 5 is, where s j 1 k (mod r).
Proof: Let a 5 (i1 i2 … ir). If k In 2 {i1, i2, …, ir}, then a(k) 5 k 5 (i1 ij)
(k) for all 2 # j # r and hence
a(k) 5 ((i1, ir) o (i1 ir21) o … o (i1 i2))(k).
On the other hand, for 1 # j , r
a(ij) 5 ij11 5 ((i1 ir) o (i1 ir21) o … o (i1 i2))(ij)
((i1 ir) o (i1 ir21) o … o (i1 i2))(ij) 5 ((i1 ir) o (i1 ir21) o … o (i1 ij)) (ij)
5 (i1 ir) o … o (i1 ij11)(i1)
5 (i1 ir) o … o (i1 ij12)(ij11)
5 ij11 5 a(ij)
and (i1 ir) o (i1 ir21) o … o (i1 i2)(ir) 5 (i1ir)(ir) 5 i1 5 a(ir).
a 5 (2 4) o (2 1) o (2 6) o (2 3) o (2 5).
Definition 6.3.3. For any permutation f in Sn, the support of f is defined as the
set {i Sn : f (i) i} of all elements in In which are not fixed by f. The support
of f will be denoted by supp(f ).
Note that if a is the r-cycle (i1 i2 … ir), then the support of a is precisely
the set {i1, i2, …, ir}.
Examples 6.3.3
1 2 3 4 5 6 7 8 9
1. Let f 5 .
6 5 3 7 4 1 2 8 9
(since f (i) i and f is an injection, f (f (i)) f (i) and hence f (i) A so that
f (i) B and g(f (i)) 5 f (i)). Similarly, if i B, then g(i) i and f (i) 5 i and
hence (f o g)(i) 5 (g o f )(i). Thus, f o g 5 g o f. b
In the following, we prove a fundamental theorem on permutations which
will be an important tool in the study of permutations.
f 5 a1 o a2 o … o as
where a1, a2, …, as are pair-wise disjoint cycles each of length atleast two. This
expression of f is unique except for the order of occurrences of the cycles ai.
f ( k ), if k ∈ ij
a j ( k ) .
k , if k ∉
i
j
The restrictions of aj and f to ij are equal and hence, by (3) above, aj is a cycle
of length atleast two and clearly
Thus, f 5 a1 o a2 o … o as.
Also, since distinct equivalence classes are disjoint, a1, a2, …, as are pair-wise
disjoint cycles, each of length atleast two.
The uniqueness of this representation of f is a direct consequence of the
facts that aj and f coincide on ij and aj is the identity outside ij . b
Proof: This is a consequence of the above theorem and the fact that any cycle
is a product of transpositions (see Theorem 6.3.3). Note that the identity per-
mutation e can be expressed as
e 5 (i j) o (i j)
There should exist least r1 . 1 (since all these are elements in the finite set A)
at which f r1 (i1 ) i1 .
Clearly r . 1 and f r1 (i1 ) 5 i1. Next choose i2 A 2 {i1, f (i1), …, f r11 (i1 )}
and consider
By the same argument given by, there exists r2 . 1 such that f r2 (i2 ) 5 i2 and
Next choose i3 A 2 {i1, f (i1), …, f r11 (i1 ), i2, f (i2), …, f r21 (i2 )} and con-
tinue the above process. This process terminates when all the elements of
the support of f exhaust. Then, we get disjoint cycles a1, a2, …, as such that
f 5 a1 o a2 o … o as. Since ai’s are pair-wise disjoint, they commute with each
other and hence
2, f (2) 5 6, f 2(2) 5 2.
f 5 (1 3 5) o (2 6) o (7 8 9).
(1 3 5) 5 (1 5) o (1 3)
and (7 8 9) 5 (7 9) o (7 8)
and hence f 5 (1 5) o (1 3) o (2 6) o (7 9) o (7 8).
Worked Exercise 6.3.2. Let i1, i2, …, ir be given distinct elements in In. How
many distinct r-cycles can be formed using all the i1, i2, …, ir.
Proof: Let O(f i) 5 ri, r 5 l.c.m. of {r1, r2, ..., rs} and r 5 riti, ti Z+. Since
f i’s are pair-wise disjoint, we get that
fi o fj 5 fj o fi for all 1 # i, j # s.
Now,
f r 5 (f1 o f2 o … o fs)r
5 f1r o f2r o … o fsr (since fi o fj 5 fj o fi)
f1r1t1 ο f 2r2t2 ο ο f srsts
5 e (since O(fi) 5 ri, f i ri 5 e).
EXERCISE 6(c)
2. Which of the following are cycles? If they are cycles, then express them in cyclic
representation.
(i) 1 2 3 4 5 6 7 8
4 6 8 2 3 5 1 7
1 2 3 4 5 6
(ii)
2 3 4 1 6 5
1 2 3 4 5 6 7 8
(iii)
2 3 4 5 6 7 8 1
1 2 3 4 5 6 7 8 9
(iv)
8 6 7 9 5 1 2 3 4
3. Express the following as products of disjoint cycles, each of length atleast two.
Also express each of following as a product of transpositions.
1 2 3 4 5 6 7 8 9
(i)
4 3 2 1 5 8 6 7 9
1 2 3 4 5 6 1 2 3 4 5 6
(ii) ο
3 2 4 1 6 5 5 1 4 6 2 3
1 2 3 4 5 6 7 8 1 2 3 4 5 6
(iii) ο
2 3 1 4 6 5 7 8 5 6 2 3 4 1
(iv) 1 2 3 ο 1 2 3 4 5 6
2 3 1 3 1 2 5 6 4
(v) (1 3 7 4 6) ο (2 3 5 6 4) ο (8 7 6 2 4 3 5)
(vi) (1 2 3 4) ο (2 3 4 5) ο (3 4 5 6)
4. Determine the order of each of the following permutations.
1 2 3 4 5 6 7 8 9
(i)
2 4 6 1 7 3 8 9 5
(ii) 1 2 3 4 5 6 ο 1 2 3 4 5
4 3 2 1 6 5 2 3 4 1 5
(iii) (5 4 3 2) ο (1 2 3 4 5 6) ο (2 4 6 1 3 5)
(iv) (8 7 6 9 3 4) ο (4 3 9 6 7 8) ο (3 4 5 6 7 8 9)
5. For any permutations f and g in Sn, prove that supp(f o g) ⊆ supp(f ) ∪ supp(g).
10. For any permutation f in Sn, prove that supp(f ) 5 if and only if f 5 e.
11. Let f Sn. Prove that f is a transposition if and only if supp(f ) is a 2-element set.
12. Prove that f is a 3-cycle if and only if supp(f ) is a 3-element set. Can this be
extended for any r-cycle?
13. Let e f Sn. Prove that f 2 5 e if and only if f is a product of disjoint
transpositions.
14. For any f and g Sn, prove that O(f ) 5 O(g f g21).
15. For any 1 # r # n, prove that there is an element in Sn whose order is r.
16. If a is an r-cycle and r is odd, prove that a2 is also a cycle.
17. If r is even in Exercise 1b, then can a2 be a cycle? Substantiate your answer.
18. If a is an r-cycle and 1 # s < r such that r and s are relatively prime, then prove
that as is also an r-cycle.
19. For any r-cycle a, prove that f o a o f21 is also an r-cycle for any permutation f.
20. If a and b are disjoint cycles, prove that f o a o f21 and f o b o f21 are also disjoint
cycles.
21. If f 5 a1 o a2 o … o as is a representation of a permutation f in Sn as a product of
disjoint cycles and g is any permutation in f, then prove that
g o f o g21 5 (g o a1 o g21) o (g o a2 o g21) o … o (g o as o g21)
is a representation of g o g o g21 as a product of disjoint cycles.
22. For any positive integer n, a partition of n is defined to be a finite sequence r1, r2,
…, rs of positive integers such that r1 # r2 # … # rs and r1 1 r2 1 … 1 rs 5 n.
List all the partitions of 4 and 5.
23. For any permutation f in Sn, let |f | denote the number of elements in the support
of f. Let f 5 a1 o a2 o … o as where a1, a2, …, as are disjoint cycles, each of length
greater than 1, such that |a1| # |a2| # … # |as|. Then prove that |a1|, |a2|, …, |as|
is a partition of |f |.
24. Prove that any permutation in Sn determines a partition of n such that f and g
determine the same partition of n if and only if g 5 h o f o h21 for some h Sn.
25. Prove that Sn is generated by the n 2 1 transpositions (1 2), (1 3), (1 4), …, (1 n).
27. Prove that Sn is generated by the transpositions (1 2), (2 3), (3 4), …, (n 2 1 n).
28. If f 5 (i1 i2 … ir) is an r-cycle in Sn, then prove that g f g21 5 (g(i1) g(i2) … g(in))
for all g Sn.
29. For any f and g Sn, prove that f is an r-cycle if and only if g f g21 is an r-cycle.
(i j) o (k l) o (j i) 5 (k l)
(i j) o (j i) 5 e 5 (k l) o (l k).
f 5 a1 o a2 o … o as
where a1, a2, …, as are pair-wise disjoint cycles. Then, the Cauchy index of f
is defined as
and is denoted by CI(f ). Also, for the identity permutation e, we define CI(e)
to be 0.
Since any f e can be uniquely, except for the order of occurrences of the
cycles, expressed as a product of disjoint cycles, the Cauchy index of f is
well-defined and CI(f ) is always a positive integer for any f e. Note that,
s
CI( f ) ∑ Ο( ai ) s | f | s
i1
Examples 6.4.1
1 2 3 4 5 6 7 8 9
1. Let f 5
4 1 3 2 7 9 8 5 6
Then, f 5 (1 4 2) o (5 7 8) o (6 9) 5 a1 o a2 o a3
Therefore, CI(f ) 5 O(a1) 1 O(a2) 1 O(a3) 2 3
5 3 1 3 1 2 2 3 5 5
2. Let f 5 (1 2) o (2 5) o (4 5) o (3 7) o (9 3) o (6 9) o (8 9). Then, we
should express f as a product of disjoint cycles, to find the Cauchy index
of f. We have
f 5 (1 2 5 4) o (3 9 8 6 7) 5 a1 o a2
Therefore, CI(f ) 5 O(a1) 1 O(a2) 2 2
54152257
a1 5 (i k1 k2 … kr).
(i j) o f 5 (i j) o (i k1 k2 … kr) o a2 o a3 o … o at
5 (i k1 k2 … kr j) o a2 o a3 o … o at
CI ((i j) o f ) 5 (r 1 2) 1 r2 1 … 1 rt 2 t
5 (r1 1 1) 1 r21 … 1 rt 2t
5 (r1 1 r2 1 … 1 rt 2 t) 1 1
5 CI(f ) 1 1
0 5 CI(e) 5 CI(f2l o f )
5 CI((c1 o c2 o … o cs)21 o f )
5 CI(cs o cs21 o … o c2 o c1 o f )
5 CI(cs21 o … o c2 o c1 o f ) 6 1
5…
5 CI(f ) 6 1 6 … 6 1
5 CI(f ) 1 p 2 q
CI(f ) 1 p 5 q
and CI(f ) 1 s 5 CI(f ) 1 p 1 q 5 2q.
Corollary 6.4.1. Let a1 a2, …, ar and b1, b2, …, bs be transpositions such that
a1 o a2 o … o ar 5 b1 o b2 o … o bs.
Then, r 1 s is even and hence either both r and s are even or both r and s
are odd.
So that r 1 s is even. The later assertion follows from the fact that r 1 s is odd
if and only if one of r and s is odd and the other is even. b
Examples 6.4.2
1 2 3 4 5 6 7 8 9
1. The permutation f 5 can be expressed as
4 6 5 7 2 3 1 9 8
f 5 (1 4 7) o (2 6 3 5) o (8 9)
5 (1 7) o (1 4) o (2 5) o (2 3) o (2 6) o (8 9)
Theorem 6.4.2. For any integer n . 1, the set of all even permutations in Sn
is a normal subgroup of Sn and is of index 2 in Sn.
Consider the group G 5 {1, 21} under the usual multiplication of real num-
bers. Define
1, if f is even
: Sn → G by ( f ) .
1, if f is odd
e 5 (1 2) o (1 2)
Sn / An G.
Therefore, we have
| Sn |
iSn ( An ) | Sn An | | G | 2.
| An |
a2 5 b, a3 5 e, b2 5 a and b3 5 e
o e a b
e e a b
a a b e
b b e a
Worked Exercise 6.4.2. List all the elements of A4 and construct a table
representing the group A4.
a1 5 (1 2 3), a2 5 (1 3 2)
b1 5 (2 3 4), b2 5 (2 4 3)
c1 5 (1 2 4), c2 5 (1 4 2)
d1 5 (1 3 4), d2 5 (1 4 3)
Also, p 5 (1 2) o (3 4)
q 5 (1 3) o (2 4)
and r 5 (1 4) o (2 3)
are also even permutations. There are only 12 even permutations and there-
fore the above together with the identity form A4. That is,
Here, we have
p2 q2 r 2 e
The following table represents the group A4.
o e a1 a2 b1 b2 c1 c2 d1 d2 p q r
e e a1 a2 b1 b2 c1 c2 d1 d2 p q r
a1 a1 a2 e p c1 q d2 b1 r d1 b2 c2
a2 a2 e a1 d1 q b2 r p c2 b1 c1 d2
b1 b1 q c2 b2 e d1 p r a1 a2 d2 c1
b2 b2 d2 p e b1 r a2 c1 q c2 a1 d1
c1 c1 r d1 a1 p c2 e q b2 d2 a2 b1
c2 c2 b1 q r d2 e c1 a2 p b2 d1 a1
d1 d1 c1 r q a2 p b1 d2 e a1 c2 b2
d2 d2 p b2 c2 r a1 q e d1 c1 b1 a2
p p b2 d2 c1 a1 b1 d1 c2 a2 e r q
q q c2 b1 a2 d1 d2 a1 b2 c1 r e p
r r d1 c1 d2 c2 a2 b2 a1 b1 q p e
Proof: Let S 5 {(i j k) : k In 2 {i, j}} Clearly S ⊆ An. Any even permuta-
tion must be a product of terms of the form
(a b) o (c d) or (a b) o (a c)
(a b) o (c d) 5 (a c b) o (a c d)
and (a b) o (a c) 5 (a c b),
it follows that An is generated by the set of all 3-cycles in Sn. Next, we prove
that any 3-cycle can be expressed as product of 3-cycles in S. Any 3-cycle is
of the form (i j a) or (i a j) or (i a b) or (j a b) or (a b c), where a, b and c are
distinct elements In 2 {i, j}. Now, we have
(i a j) 5 (i j a) o (i j a)
(i a b) 5 (i j b) o (i j a) o (i j a)
(j a b) 5 (i j b) o (i j b) o (i j a)
and (a b c) 5 (i j a)2 o (i j c) o (i j b)2 o (i j a)
Proof: Suppose that N contains a 3-cycle (i j k), where i, j and k are some
distinct elements in In.
For any a In 2 {i, j, k}, we have
(i j a) 5 (i j) o (k a) o (i j k)2 o (k a) o (i j)
5 (i j) o (k a) o (i j k)2 o ((i j) o (k a))21 f N f21,
where f 5 (i j) o (k a) An.
Example 6.4.3
1. Any finite group of prime order is simple; for if |G| 5 p, where p is a
prime number and H is a subgroup of G, then, by the Lagrange’s theo-
rem, |H| divides |G| and hence |H| 5 1 or p which implies that H 5 {e}
or H 5 G.
2. A4 is a not a simple group, since
N 5 {e, p q, r}
( )
Proof: Since A3 is a group of order 3 32! , A3 is simple (see Example 6.4.3 (1)).
Also A4 is not simple by Example 6.4.3 (2). Now, let n . 4. We shall prove
that An is simple. Let N be a normal subgroup of An and N {e}. By
Theorem 6.4.4, it is enough if we can prove that N contains a 3-cycle. We
prove this by distinguishing the following cases.
Case (1): Suppose that N contains an element f such that f is the product of
disjoint cycles, atleast one of which is of length r 4. Then,
f 5 (i1 i2 … ir) o b
where r 4 and (i1 i2 … ir) and b are disjoint. Now, put a 5 (i1 i2 i3). Then, a
An and, by the normality of N in An, a o f o a21 N. Now, we have
(i1 i2 ir) 5 b2l o (i1 ir ir21 … i2) o a o (i1 i2 … ir) o b o (i1 i3 i2)
5 f21 o (a o f o a21) N
Case (2): Suppose that N contains an element f such that f is the product of
disjoint cycles, atleast two of which are 3-cycles. Then,
(i1 i4 i2 i6 i3) 5 b21 o (i4 i6 i5) o (i1 i3 i2) o (i1 i2 i4) o (i1 i2 i3)
o (i4 i5 i6) o b o (i1 i4 i2)
5 f21 o (a o f o a21) N
f 5 (i1 i2 i3) o b
where (i1 i2 i3) and b are disjoint and b is a product of disjoint 2-cycles. Now
we have
Now, put 5 (i1 i3) o (i2 i4). We have N. since n $ 5, we can choose
j In 2 {i1, i2, i3, i4}. Put 5 (i1 i3 j).
(i1i3 j ).
Theorem 6.4.6. For any n . 1, the alternating group An is the only subgroup
of index 2 in the symmetric group Sn.
Proof: Let H be a subgroup of index 2 in Sn. Then, there is only one left coset
of H other than H. Then, fH 5 Sn 2 H and hence fH 5 gH for any f and g
Sn 2 H. In particular, fH 5 f21H for any f Sn 2 H (note that f H if and
only if f21 H) and hence f2 H for all f Sn.
If a 5 (i j k) is any 3-cycle in Sn, then a3 5 e and hence
a21 5 a2 H and therefore a H.
Therefore, H contains all 3-cycles. By Corollary 6.4.4, An ⊆ H. But
| Sn | |S |
iSn ( H ) 2 iSn ( An ) n
|H | | An |
5 the n-cycle (1 2 3 … n)
1 2 3 4 … n
and
1 n n 1 n 2 … 2
(1 2 3 … n)
1 2 3 4 n
and f .
1 n n 1 n 2 2
f ∏
2in2i
(i, n 2 i ) (2 n) ο (3 n 1) ο (4 n 2) ο
o o 5 .
If we choose a and b G such that f (a) 5 and f (b) 5 , then we get that G
is generated by a and b, O(a) 5 n, O(b) 5 2 and aba 5 b.
Conversely suppose that G is generated by two elements a and b, such that
O(a) 5 n, O(b) 5 2 and aba 5 b (or ab 5 ba21 or ba 5 a21b). Then, for any
integers j and k,
From these relations and from the hypothesis that G is generated by a and b,
it follows that every element of G can be expressed as ak bj for some integers
k and j. Since O(a) 5 n,
e, if j is even
b j .
b, if j is odd
a kr b s j .
for all 0 # k , n and j 5 0 or 1. Using the fact that u and satisfy the same
conditions (1), (2) and (3) in Dn as a and b in G, it can easily checked that f is
an isomorphism. Thus, G . Dn. b
Theorem 6.4.8. Let n 3 and D9n be the group of all symmetries of a regular
n-gon (a polygon of n equal sides). Then,
Dn D9n.
2 2 2 2
a( x, y ) x cos y sin , x sin y cos
n n n n
and b(x, y) 5 (x, 2y).
2�
n
Worked Exercise 6.4.3. Prove that the dihedral group Dn, n 3, is not simple.
EXERCISE 6(d)
1. State whether the following are true or false and substantiate your answers.
(i) The Cauchy index of a permutation f is equal to the number of elements
in the support of f.
(ii) The order of A5 is even.
(iii) A3 is an abelian group.
(iv) A4 is an abelian group.
(v) |A5| 5 120
(vi) A2 is trivial.
19. Find all maximal subgroups M of Sn such that f 2 M for all f Sn.
21. Let H be a subgroup of Sn containing an odd permutation. Then prove that exactly
half of the number of elements in H are even.
22. Prove that the order of any subgroup of Sn containing an odd permutation is
even.
23. If f is an odd permutation in Sn, then prove that fAn is the set of all odd permuta-
tions in Sn.
26. Prove that there is no subgroup of order 6 in A4. What does this say about the
Lagrange’s theorem.
33. Prove that {f S6 : f (3) 5 3 and f (5) 5 5} is a subgroup of S6. What is its
order?
36. If f is a 10-cycle, then find all the integers m between 2 and 10 for which f m is
also a 10-cycle.
37. For any f and g in Sn, prove that g is even if and only if f o g o f 21 is even.
38. Prove that the set of all odd permutations in Sn is a coset of An in Sn.
39. Prove that the centre of the group Sn is trivial for any n # 3.
41. For any 1 , r # n, derive a formula for the number of r-cycles in Sn.
43. For any 1 , r # n, derive a formula for the number of r-cycles in An.
44. Let f Sn, such that the order of f in the group Sn is odd. Prove that f is an even
permutation.
45. Let n 3. Let G be the multiplicative group of matrices over complex numbers
generated by
e 2 i n 0 0 1
A and B .
0 e
2 i n 1 0
46. If a is the generator of order n in Dn, prove that ,a. is normal in Dn and
Dn/,a. Z2.
48. For any n . 1, prove that the alternating group An is the only subgroup of index
2 in Sn.
50. Prove that Z(Dn), the centre of the dihedral group of degree n, is of the order
1 or 2 according as n is odd or even.
Before the concept of an abstract group took its present shape, the theory of
groups dealt only with permutation groups. Abstract groups were introduced
much later in order to focus attention on those properties of permutation
groups that concern the resultant composition (the binary operation in the
permutation groups) only and do not refer to the set on which the permuta-
tions act. However, we have seen that any group can be identified (isomor-
phic) with a group of permutations on some set. Switching back from the
abstract point of view to the concrete case of a permutation group is often
useful in the abstract theory. The use of permutation groups provides cer-
tain counting techniques which play an important role in the theory of finite
groups. In this chapter, we provide a procedure for passing from the abstract
point of view to the concrete case of permutations, by introducing the con-
cept of ‘a group acting on a set’ and develop certain counting techniques in
finite groups.
(i) e(x) 5 x
(ii) (a ? b)(x) 5 a(b(x))
for all x X and a and b G, where e is the identity in G. This is abstracted
in the following definition.
Example 7.1.1
1. Let X be a nonempty set and G be a subgroup of the group S(X) of per-
mutations on X (bijections of X onto itself). Define
: G 3 X → X by (a, x) 5 a(x)
for any a G and x X. Then, it can be easily verified that is an action
of G on X. Note that the identity in G is the identity mapping and the
binary operation in the group G is the composition of mappings. This
action is called the natural action of G.
2. Let G be a group and X be the set G itself. Define : G 3 X → X by (a,
x) 5 a x, where a x is the product of a and x in the group G. Then
clearly is an action of G on itself and is called the action of G on itself
by left translation. The action of an element a in G on an element x in X
(5 G) is simply multiplying x by a on the left.
3. Again let G be a group and X be the set G. Define : G 3 X → X by
(a, x) 5 xa21. Then,
(e, x) 5 xe21 5 xe 5 x
and (ab, x) 5 x(ab)21 5 x(b21a21) 5 (xb21)a21 5 (a, (b, x))
(e, x) 5 exe21 5 x
and (ab, x) 5 (ab)x(ab)21 5 a(bxb21)a21 5 (a, (b, x)).
for all a G and xH X with x G. First note that is well defined; for
xH 5 yH ⇒ x21y H
⇒ (ax)21(ay) 5 x21a21ay 5 x21y H
⇒ axH 5 ayH
1. ex 5 x for all x X
2. (ab)x 5 a(bx) for all a, b G and x X.
The condition (2) is not the associative law, for we are not dealing with a
binary operation on a set. a and b are elements of the group G and x is an ele-
ment of X. There should not be any confusion with this notation. One should
understand as per the context.
There may be several actions of the same group on the same set, as given
in the examples (2), (3) and (4) above. In the following, we establish an inter-
relation between the actions of a given group G on a given set X and the
homomorphisms of G into the group S(X) of permutations on X.
for all x X and therefore f(a, b) 5 f(a) ? f(b). This means that f is a
homomorphism of G into S(X). If 1 and 2 are two actions of G on X, such
that f 1 f 2 then for any (a, x) G 3 X,
Therefore, an action is effective if and only if ker 5 {e}; that is, e is the
only element a in G such that (a, x) 5 x for all x X. For example, if is
the action of G on itself by left translation (see Example 7.1.1 (2)), then is
effective. On the other hand, if is the action of G on itself by conjugation
(that is, (a, x) 5 axa21), then is effective if and only if the centre Z(G) of
G is trivial, since
Worked Exercise 7.1.1. Let X be the set of all complex number with unit
modulus and G be the additive group of real numbers. Define : G 3 X → X
by (a, x) 5 eiax for any a G and x X. Then prove that is an action of
G on X. Is effective?
Answer: Note the 0 is the identity in G (5 (R, 1)) and that the usual addition
1 is the binary operation on G.
Then prove that is an action of G on the power set of X and that ker
5 ker .
(e, A) 5 {(e, x) : x A} 5 A
and (a, (b, A)) 5 {(a, y) : y (b, A)}
5 {(a, (b, x)) : x A}
5 {(ab, x) : x A}
5 (ab, A)
EXERCISE 7(a)
3. Let X be the set of vertices 1, 2, 3, 4 of a square and G 5 D4, the dihedral group
of degree 4. Define : D4 3 X → X by (f, i) 5 f(i). Then prove that is an action
of D4 on X.
8. Let G be the group of symmetries of a cube. Prove that there are nontrivial
actions of G on each of the set of edges, the set of faces, the set of vertices and
the set of diagonals of the cube.
10. Let G be a group and X1 and X2 be nonempty sets, let 1 and 2 be actions of G
on X1 and X2, respectively. Define (132) : G 3 (X1 3 X2) → X1 3 X2 by
11. Let a group G act on itself by left translation. Prove that the action is effective
and deduce the Cayley’s theorem.
13. Let H be a proper subgroup of a finite group G. Then prove that G ∪ aHa1.
a∈G
14. Let p be the smallest prime dividing the order of a finite group G. Then prove
that any subgroup of index p in G is normal in G.
15. Derive from the above that any subgroup of index 2 in a group is normal.
16. Prove that any subgroup of order 539 in a group of order 2695 is normal.
17. Let p and q be distinct primes such that p , q. Prove that any subgroup of order
q in a group of order pq is normal.
18. Let G be a group of odd order. Then prove that any subgroup of index 3 is
normal in G.
Theorem 7.2.1. Let a group G act on a set X. Then, the orbits of elements
of X form a partition of X. That is, any two orbits are either equal or disjoint
subsets of X and the union of all orbits is equal to X.
Proof: For any x X, we have O(x) 5 {ax : a G}. Clearly O(x) is a subset
of X for each x X. Also, since
x 5 ex O(x)
We have X ∪ O( x ). Now, let x and y X such that the orbits O(x) and
x∈ X
O(y) are not disjoint. Then, choose z O(x) ∩ O(y). Thus,
z 5 ax 5 by for some a and b G
and hence x 5 (a21b)y and y 5 (b21a)x so that O(x) ⊆ O(y) and O(y) ⊆ O(x)
and therefore O(x) 5 O(y). Then,
O(x) ∩ O(y) 5 or O(x) 5 O(y). b
Definition 7.2.2. Let a group G act on a set X and x X. The stabilizer of x
is defined to be the set
St(x) 5 {a G : ax 5 x}.
Theorem 7.2.2. Let a group G act on a set X and x X. Then, the stabilizer
St(x) is a subgroup of G and there is a bijection of the orbit O(x) onto the set
of left cosets of the stabilizer St(x) in G.
Proof: Since ex 5 x, where e is the identity in the group G,
e St(x) 5 {a G : ax 5 x}
and hence St(x) is a nonempty subset of G. Also,
a and b St(x) ⇒ ax 5 x 5 bx
⇒ (ab)x 5 a(bx) 5 ax 5 x
⇒ ab St(x)
and a St(x) ⇒ ax 5 x
⇒ a21x 5 a21(ax) 5 (a21a)x 5 ex 5 x
⇒ a21 St(x).
Therefore, the stabilizer St(x) is a subgroup of G. Next, let A be the set of all
left cosets of St(x) in G. That is,
A 5 {aSt(x) : a G}.
Define a : O(x) → A by a(ax) 5 aSt(x).
For any a and b G, we have
ax 5 bx ⇔ a21bx 5 x
⇔ a21b St(x)
⇔ aSt(x) 5 bSt(x).
Therefore, a is well defined and is an injection. Clearly a is a surjection also.
Thus, a is a bijection of O(x) onto the set of left cosets of St(x) in G. b
Example 7.2.1
1. Let X be any nonempty set and G be any group. Define : G 3 X →
X by
(a, x) 5 x for all a G and x X.
Then clearly is an action of G on X and is called the trivial action.
Here, for any x X,
the orbit O(x) 5 {x}
and the stabilizer St(x) 5 G.
2. Consider the action of any group G on itself by left translation (see
Example 7.1.1 (2)). Here, X 5 G and, for any x X (5 G),
the orbit O(x) 5 {ax : a G} 5 X 5 G
since any b G can be written as b 5 (bx21)x, bx21 G.
Also, the stabilizer St(x) 5 {a G : ax 5 x} 5 {e}.
3. Consider the action of a group on itself by conjugation (see Example
7.1.1 (4)), where : G 3 G → G is defined by (a, x) 5 axa21, for all a
and x G. Here, for any x X,
the orbit O(x) 5 {(a, x) : a G}
5 {axa21 : a G}
and the stabilizer St(x) 5 {a G : (a, x) 5 x}
5 {a G : axa21 5 x}
5 {a G : ax 5 xa}
The orbit O(x) is called the conjugacy class of x in G and is usually
denoted by C(x). Also, the stabilizer St(x) is called the centralizer of x in
G and is usually denoted by CentG(x). That is,
C(x) 5 {axa21 : a G}
and CentG(x) 5 {a G : ax 5 xa}.
4. Let H be a subgroup of a group G and X be the set of left cosets of H in
G. Define : G 3 X → X by
(a, xH) 5 (axa21)H
Example 7.2.2
1. The action of a group G on itself by left (right) translation (see Example
7.1.1 (2) and (3)) is transitive, since, for any x and y G, yx21 G and
y21x G and
(yx21)x 5 y and x(y21 x)21 5 y.
2. The action of a group G on itself by conjugation (see Example 7.1.1 (4))
is not transitive, in general. This is transitive if and only if there is only
one conjugacy class or, equivalently, any two elements of the group or
conjugates to each other.
A group may act on the same set (or on two different sets) differently. In
the following, we define equivalence of two such actions in a natural way.
Definition 7.2.4. Let a group G act on two sets X and X. These two actions
are said to equivalent if there is a bijection a : X → X such that
a(ax) 5 aa(x)
a ? f(a) 5 f(a) ? a
f�(a)
X X
� �
X′ X′
f�′(a)
for all a G, where f(a) and f(a) are the permutations on X and X cor-
responding to the actions and , respectively (see Theorem 7.1.1). In other
words,
(see Example 7.1.1 (2) and (3)). Define a : G → G by a(x) 5 x21 for any x G.
Then, a is a bijection and, for any a and x G,
Proof: Let be the given transitive action on X and Y be the set of left
cosets of H in G. Let be the action of G on Y by left translation; that is,
: G 3 Y → Y is defined by
Since the action is transitive, we get that the orbit O(x) 5 X. Let a : X → Y
be the bijection given in the proof of Theorem 7.2.2; that is,
Clearly the whole of X 3 X and the diagonal X5 {(x, x) : x X} are equiva-
lence relation X which are compatible with every action of G on X. These two
equivalence relations are called trivial relations.
and hence ((a, x), (a, z)) c so that (x, (a, x)) c. Now, it can be easily
verified that (a, Y) 5 Y. Conversely suppose that there is a proper subset Y of
X such that |Y| . 1 and, for any a G, either (a, Y) 5 Y or (a, Y) ∩ Y 5 .
From this it follows that, for any a and b G,
Example 7.2.4. Let a group G act on itself by left translation. (see Example
7.1.1 (2)) and H be a nontrivial proper subgroup of G. The action of G on
itself is defined by (a, x) 5 ax, the product of a and x in G. For any x and y
G, we have y 5 (yx21)x 5 (yx21, x) and yx21 G and therefore the action
is transitive. Define a relation c on G by
(x, y) c ⇒ x21y H
⇒ (ax)21(ay) 5 x21a21ay 5 x21y G
⇒ (ax, ay) c
and hence c is compatible with the action . Also since H {e}, c H and
since H G, c G 3 G. Therefore, the action is imprimitive.
Note that in the above example, any equivalence class of c is simply
a left coset of H in G. If G has no nontrivial proper subgroups, then the
above action is primitive and vice versa. In other words, the action of G
on itself by left translation is primitive if and only if {e} is a maximal
subgroup of G.
This is generalised in the following theorem. First, let us call a proper sub-
group K of a group G maximal if there is no proper subgroup of G containing
K properly; that is, for any subgroup H of G, K ⊆ H ⊆ G implies that either
H 5 K or H 5 G.
Theorem 7.2.7. Let a group G act transitively on a set X with |X| 2. Then,
the action is primitive if and only if the stabilizer of any x X is a maximal
subgroup of G.
either aY 5 Y or aY ∩ Y 5 .
Choose x Y. We shall prove that the stabilizer St(x) is not a maximal sub-
group of G.
Put H 5 {a G : aY 5 Y}.
a St(x) ⇒ ax 5 x
⇒ aY ∩ Y
⇒ aY 5 Y
⇒ a H.
St(x) H G,
Worked Exercise 7.2.1. Let G be a finite group of prime order. Suppose that
G acts on a set X and x X such that ax 5 x for some a e in G. Then prove
that bx 5 x for all b G.
EXERCISE 7(b)
N ∩ x H x1.
x ∈G
for any (a1, a2) G1 3 G2, x1 X1 and x2 X2. Prove that this is an action,
which is not transitive.
10. An action of a group G on a set X is called doubly transitive if, for any x1, x2, y1
and y2 X, there exists a G such that ax1 5 y1 and ax2 5 y2. Prove that any
doubly transitive action is primitive. Is the converse true?
Theorem 7.3.1. Let G be group acting on a finite set X and O(x1), O(x2), …,
O(xn) be all the distinct orbits in X. Then, the number of elements in X can be
obtained by the formula
n n 1
| X | ∑ | O( xi ) | | G | ∑ .
i1
i1 | St( xi ) |
n
∪ O( xi ) X and O( xi ) ∩ O( x j ) for i j.
i1
Therefore, the total number of elements in X is equal to the sum of the num-
bers of elements in the orbits. That is,
n 1
| G | ∑ .
i1 | St( xi ) | b
the class equation. First let us recall that (a, x) axa21 is an action of a group
G on itself and is called the action of G on itself by conjugation. Here, for any
x G, the orbit of x is simply the conjugate class C(x) of x in G; that is,
St(x) 5 {a G : axa21 5 x}
5 {a G : ax 5 xa}
where Z(G) is the centre of G and x1, …, xn are elements of G such that C(x1),
C(x2), …, C(xn) are all the distinct conjugacy classes, each with more than one
element. This equation is known as the class equations of G.
Now, we shall distinguish two types of conjugacy class, namely classes each
with only one element and classes each with more than one element. Note
that, for any x G, the conjugacy class of x is
n
| G | | Z (G ) | ∑ | C ( xi ) |
i1
n
| Z (G ) | ∑ iG (Cent G ( xi ))
i1
n 1
| Z (G ) | | G | ∑ . b
i1 | Cent( xi ) |
For any subset A of a group G and for any element x in G, the set
xAx21 5 {xax21 : a A}
C(A) 5 {xAx21 : x G}
CentG(A) 5 {x G : xAx21 5 A}
5 {x G : xA 5 Ax}.
C(A) and CentG(A) are respectively called the conjugacy class of A in P(G)
and the centralizer or normalizer of A in G. Clearly, for any subset A of a
finite group G,
|G |
| C ( A) | iG (Cent G ( A)) .
| Cent G ( A) |
Proof: We have to prove that |X| 2 |X0| is a multiple of p. Observe that the
orbit of an element x X is a singleton set if and only if x X0. Therefore,
there are exactly |X0| number of singleton orbits in X. Let O(x1), O(x2), …,
O(xn) be all the distinct orbits each with more than one element. Then, by
Theorem 7.3.1, we have
n n
| X | | X 0 | ∑ | O( X i ) | | X 0 | ∑ iG (St( xi )).
i1 i1
Note that the stabilizer St(xi) is a subgroup of G and |G| 5 pn. By the Lagrange’s
theorem, |St(xi)| is a divisor of pn and
|G |
iG (St( xi )) | O( xi ) | 1
| St( xi ) |
Theorem 7.3.6. Let p be a prime number. Then, any group of order p2 is abelian.
Proof: Let G be a group of order p2 and Z(G) its centre. By Theorem 7.3.5,
|Z(G)| . 1 and, by the Lagrange’s theorem |Z(G)| is a divisor of |G| 5 p2.
Therefore, |Z(G)| 5 p or p2. Suppose that |Z(G)| 5 p. Then, Z(G)| G and
hence we can choose a G such that a Z(G). Then, there exists x G such
that ax xa. Consider the centralizer of x
CentG(x) 5 {y G : xy 5 yx}.
Xa 5 {x X : ax 5 x}.
That is, Xa is the set of all elements of X which are fixed by the action of a.
Note that Xe 5 X, where e is the identity in group G.
Theorem 7.3.7 (Burnside’s theorem). Let a finite group G act on a finite set
X and n be the number of orbits in X. Then,
1
n ∑ |X a |.
|G| a∈G
A 5 {(a, x) G 3 X : ax 5 x}.
Note that, for any fixed element x in X, the number of pairs (a, x) in A is pre-
cisely equal to the order of the stabilizer St(x). Also, for any fixed element a
in G, the number of pairs (a, x) in A is exactly equal to |Xa|, where Xa 5 {x
X : ax 5 x}. Therefore, we have
∑ | St( x) | | A | ∑ | X
x∈ X a∈G
a |⋅
By Corollary 7.2.1,
|G |
| St( x ) | ,
| O( x ) |
where O(x) is the orbit of x. Note that O(x) 5 O(y) for all y O(x). Since X
is finite, the number of orbits in X is finite. Let O(x1), O(x2), …, O(xn) be all
the distinct orbits in X. Then,
|G |
∑ | St( x) | ∑ |O( x) |
x∈ X x∈ X
n
1
| G | ∑ ∑
i1 x∈O( xi ) | O( x) |
n
1
| G | ∑ ∑
i1 x∈O( xi ) | O( xi ) |
n
1
| G | ∑ | O( xi ) |
i1 | O( xi ) |
| G | ⋅ n
Answer: Let (Zp, 1) be the additive group of integers modulo p and X be the
set of all possible necklaces. Since there are p beads in each of the necklaces
and each bead can have any of n different colours, |X| 5 np.
xp x1 xi +1
x2 xi +2
i =
x3
Let Zp act on X as shown in the figure, where the subscripts are modulo p. The
action of any i Zp on any given necklace yields the same necklace; only
the beads are permuted cyclically. Therefore, the number of orbits in X is
same as the number of different necklaces, which can be computed by using
Xi 5 {x X : ix 5 x},
5 {x X : (i 1p j)x 5 x for all 0 j p}
5{x X : jx 5 x for all j Zp},
since Zp is a cyclic group of prime order and hence any nonidentity element
generates Zp. Therefore, for and 0 i Zp, Xi consists those necklaces which
are unchanged by permutation and hence Xi consists of those necklaces in
which all the beads are of same colour. Since we are given with n different
colours, it follows that |Xi| 5 n for all 0 i Zp. Thus, by the Burnside’s
theorem, the number of different necklaces (the number of orbits in X) is
1 1 p1
∑
| Z p | i∈ p
| X i | ∑| X i |
p i0
1 p
( n ( n n))
p
1 p
n ( p 1)n
p
n p1
( n p 1)
p
Worked Exercise 7.3.2. Let G be a group and a G such that O(a) . 1. Sup-
pose that G has exactly two conjugacy classes. Then prove that |G| 5 2.
Since O(a) 5 O(xax21), it follows that O(b) 5 n for all b e and, in particular
bn 5 e for all b e. Now, we prove that n is a prime. Since n . 1, we can
O( a)
choose a prime p dividing n. Then, O( a p ) p np n and hence ap 5 e,
so that p 5 O(a) 5 n. Thus, n is a prime. Next, consider a2. If a2 e, then a2
C(a) and hence a2 5 xax21 for some x G, so that (a2)m 5 xmax2m for all
m . 0 and
Since p divides both |N| and |N| 2 |N0|, it follows that |N0| is a multiple of p.
Also, since e Z(G) ∩ N 5 N0, |N0| . 0 and hence |N0| p . 1. Thus, Z(G)
∩ N is nontrivial.
EXERCISE 7(c)
1. Determine all the distinct conjugacy classes in each of the following and verify
that the number of elements in each conjugacy class is a divisor of the order of
the group
(i) The symmetric group S3 of degree three.
(ii) The alternating group A4 of degree four.
(iii) The symmetric group S4 of degree four.
(iv) The dihedral group D4 of degree four.
2. Find the number of different (distinguishable) dice that can be made by marking
the faces of a cube using one to six dots.
3. How many different tetrahedral dice can be made by marking the faces of a
regular tetrahedron using one to four dots?
4. How many different ways can seven people be seated at a round table, where
there is no distinguishable leader to the table?
5. Find the number of different ways the edges of an equilateral triangle can be
painted if four different colours of paint are available, assuming only one colour
is used on each edge, and the same colour may be used on different edges.
6. Repeat Exercise 5 above with the assumption that a different colour is used on
each edge.
10. Let N be a normal subgroup of a finite group G such that the order and the index
of N are relatively prime. If a is an element of G, such that O(a) divides |N|, then
prove that a N.
11. Let G be a group and H 5 {a G : C(a) is finite}, where C(a) is the conjugacy
class of a. Then prove that H is a subgroup of G.
12. Let N be a normal subgroup of order 3 in a group G such that N Z(G). Then
prove that G has a subgroup of index 2.
13. Find the number of different necklaces that can be formed with five beads and
two colours.
14. Determine the number of different necklaces that can be formed with six beads
and two colours.
15. Find the number of neckties having n strips (of equal width) of K distinct
colours.
C(S) 5 {aSa21 : a G}
and N(S) 5 {a G : aSa21 5 S}.
Prove that N(S) is a subgroup of G and that C(S) is bijective with the set of left
cosets of N(S) in G. N(S) is called the normalizer of S in G.
17. Let G be a group of order pn, where p is a prime and n Z1. If A is a proper
subgroup of G, then prove that A is properly contained in the normalizer of
A in G.
19. Prove that any subgroup of order 343 in a group of order 2401 is normal.
20. Find the number of different necklaces formed by 11 beads, where each bead can
have any of the five given different colours.
Then, |X | 5 |G|p21, since, for any (x1, x2, …, xp21) G p21, (x1, x2, …, xp21, xp)
X, where xp 5 (x1, x2, …, xp21)21 and vice versa. Since p divides |G| and
p − 1 . 0, it follows that p divides |X|.
Consider the group Zp of integers modulo p. We shall define an action of
Zp on X as follows: for any x 5 (x1, x2, …, xp) X and i Zp 5 {0, 1, 2, …,
p21}, define
and hence the above defines a mapping of Zp 3 X into X. It can be easily veri-
fied that this is an action of Zp on X. Therefore, by Theorem 7.3.4,
Since p divides |X| and |X| 2 |X0|, it follows that p divides |X0|. Also, since (e,
e, …, e) X0, we get that |X0| . 0. Therefore, |X0| 5 pn for some n Z1 and,
in particular,
|X0| p . 1.
Thus, there exists (x, x, …, x) X0 other then (e, e, …, e) and hence there
exists x e in G such that xp 5 e. Then, O(x) . 1 and O(x) is a divisor of p.
Since p is prime, it follows that O(x) 5 p. Thus, x is an element of order
p in G. b
Corollary 7.4.1. Let G be a finite group and p be a prime divisor of the order
of G. Then, G has a subgroup of order p.
is a subgroup of order p in G.
Proof: Let G be a finite group and p be a prime number. Suppose that |G| 5 pm,
0 m Z. For any a G, O(a) divides |G| 5 pm and hence O(a) 5 pm for
some 0 n m. Therefore, G is a p-group. Conversely suppose that |G| pm
for any m 0. Then, there exists a prime q p such that q divides |G|. By
the Cauchy’s theorem, there exists an element a of order q in G and therefore
G is not a p-group. b
In the proof of the above theorem, the finiteness of G is necessary. For any
prime p, there are infinite p-groups. Consider the following example.
a
H n : a and n ∈ Z and n 0.
p
1 1 1 1
∉ Z and hence n Z m Z.
pn pm p p
a
x Z for some a and n Z, n 0.
pn
Since pnx 5 p n (( a / p n ) Z) a Z Z, it follows that O(x) is a divisor of
pn and hence O(x) 5 pm for some 0 m n. Thus, G is a p-group.
NG(H) 5 {a G : aHa21 5 H}
H is normal in A ⇔ A ⊆ NG(H).
NG (H )
iNG ( H ) ( H ) ps for some s ∈ Z .
|H |
Since pn divides |G| 5 |H| ? iG(H) 5 pn21 ? iG(H), it follows that p divides iG(H)
and therefore p divides iNG ( H ) ( H ). Recall that NG(H) 5 {a G : aHa21 5
H} and hence H is a normal subgroup of NG(H) and therefore we have the
quotient group NG(H)/H whose order is iNG ( H ) ( H ). Now, NG(H)/H is a group
whose order is divisible by p. Therefore, by Corollary 7.4.1, NG(H)/H has a
subgroup K of order p. Then, K5 A/H where A is a subgroup of NG(H) con-
taining H. Now,
Proof: In the proof of the above theorem, we have H ⊆ A ⊆ NG(H) and hence
H is a normal subgroup of A (since a A ⇒ a NG(H) ⇒ aHa21 5 H).
Corollary 7.4.4. Let G be a finite group and p be a prime number. Then, every
p-subgroup of G is contained in a maximal p-subgroup.
By the Sylow Theorem 2 I (7.4.4), for any prime p and a finite group
G, maximal p-subgroups exist in G. If p does not divide |G|, then p0(51)
is the largest power of p dividing |G| and hence {e} is the only Sylow p-sub-
group of G.
Since p divides |X| 2 |X0| and p does not divide |X|, it follows that p does not
divide |X0| and hence |X0| 0; that is, X0 is a nonempty set. Thus, there exists x
G such that H ⊆ xSx21. In particular, if H is also a Sylow p-subgroup of G,
then |H| 5 |S| and H ⊆ xSx21 and hence H 5 xSx21 for some x G. b
Corollary 7.4.6. Let G be a finite abelian group. Then, for each prime num-
ber p, G has a unique Sylow p-subgroup.
Sylow Theorem 2 II describes all Sylow p-subgroups in terms of a given
Sylow p-subgroup. However, it does not give the exact number of Sylow
p-subgroups. In the following, we derive certain formulae to find the exact
number of Sylow p-subgroups.
Theorem 7.4.6 (Sylow Theorem 2 III). Let G be a finite group. For any prime
p, let np be the number of Sylow p-subgroups of G. Then,
1. np divides |G| / pn, where pn is the largest power of p dividing |G|,
2. np divides |G| and
3. np 5 mp 1 1 for some nonnegative integer m.
Proof:
1. Let X be the set of all subgroups of G. Then, G acts on X by conjugation.
Let P be a Sylow p-subgroup of G. By the Sylow Theorem 2 II (7.4.5),
the orbit of P in X is precisely the set of all Sylow p-subgroups of G. We
know that
np 5 |O(P)| 5 i(NG(P)),
where NG(P) 5 {a G : aPa21 5 P}. Let pn be the largest power of p
dividing |G|. Then, |P| 5 pn and, by Worked Exercise 4.4.1 (1),
|G |
iNG ( P ) ( P ) ⋅ n p iNG ( P ) ( P ) ⋅ iG ( N G ( P )) iG ( P )
pn
and hence np is a divisor of |G| / pn.
2. This a single consequence of (1).
3. Let Y be the set of all Sylow p-subgroups of G and S Y. Let S act on Y
by conjugation. Then, by Theorem 7.3.4,
|Y| |Y0|(mod p),
where Y0 5 {P Y : aPa21 5 P for all a S}
5 {P Y : S ⊆ NG(P)},
where NG(P) 5 {a G : aPa21 5 P}. We shall prove that Y0 is a single-
ton set. Consider P Y0. Then,
aPa21 5 P and hence aP 5 Pa for all a S
Proof: The theorem is trivial if |G| 5 1. Therefore, we can assume that |G| . 1.
Let us suppose that
where p1, p2, …, pk are distinct primes and r1, r2, …, rk are positive integers.
Then, since d is a divisor of |G|,
where s1, s2, …, sk are integers such that 0 si ri for all 1 i k. Now, for
each i, pisi is a divisor of d and d is a divisor of |G| and hence pisi is a divisor
of |G|. Therefore, by Sylow Theorem 2 I (7.4.4) there exists a subgroup Ai of
G such that |Ai| 5 pisi . For any i j, Ai ∩ Aj is a subgroup of Ai as well as of Aj
and hence, by Lagrange’s theorem |Ai ∩ Aj| is a common divisor of |Ai| ( pisi )
s s
and |Aj| ( p j j ). Since pi and pj are distinct primes, pisi and p j j are relatively
prime and hence |Ai ∩ Aj| 5 1 for all i j. Also, since G is an abelian group,
AiAj is a subgroup of G and
| Ai || Aj | s
| Ai Aj | pisi p j j .
| Ai ∩ Aj |
|G |
n3 divides 5 5 and n3 5 3m 1 1, m 0
3
|G |
n5 divides 5 3 and n5 5 5s 1 1, s 0.
5
These imply that n3 5 1 5 n5. Therefore, there is a unique subgroup A of
order 3 in G (note that 31 is the largest power of 3 dividing |G| and 51 is the
largest power of 5 dividing |G|) and hence A is normal. Similarly, there is a
normal subgroup B of order 5 in G. Since |A| and |B| are relatively prime, we
get that A ∩ B 5 {e}. From this, we have
| A || B |
| AB | | A || B | 3 5 15 | G |
| A∩ B |
Thus, G is cyclic.
The above result is extended to any groups of order pq, where p and q
are primes, p . q and q does not divide p 2 1. In the above result, we have
15 5 5 ? 3 and 3 does not divide 5 2 1.
Worked Exercise 7.4.2. Let G be a group of order pq, where p and q are dis-
tinct primes, p . q and q does not divide p 2 1. Then prove that G is cyclic.
Answer: Let np and nq be the number of Sylow p-subgroups and the number
of Sylow q-subgroups, respectively. Then, we have
np 5 mp 1 1, m 0, nq 5 sq11, s 0,
Thus, G is cyclic.
Worked Exercise 7.4.3. Prove that there are no simple groups of order 63.
a NG(N) ⇒ aNa21 5 N
⇒ aSa21 ⊆ aNa21 5 N
⇒ aSa21 5 S
⇒ a NG(S) 5 N.
EXERCISE 7(d)
1. State whether each of the following is true or false and substantiate your
answer.
(i) For any prime p and for any finite group G, there is a Sylow p-subgroup
of G.
(ii) The order of a Sylow 3-subgroup of a group of order 108 is 27.
(iii) Any Sylow 3-subgroup of a group of order 54 is normal.
(iv) There exists a subgroup of order 16 in a group of order 216.
(v) Any group of order 159 is simple.
(vi) Any group of order 159 is cyclic.
(vii) A group of prime power order has no Sylow p-subgroups.
(viii) Every p-subgroup of a finite group is a Sylow p-subgroup.
(ix) Any group of order 121 is abelian.
(x) Any group of order 8 is abelian.
2. Determine all the Sylow p-subgroups of the following groups for all the primes p.
(i) Z24, the group of integers modulo 24.
(ii) S3, the symmetric group of degree 3.
(iii) S4, the symmetric group of degree 4.
(iv) A4, the alternating group of degree 4.
(v) Z3 3 Z3.
6. For any fixed prime p, prove that the intersection of all Sylow p-subgroups of a
group G is a normal subgroup of G.
8. If p is a prime and r and n are positive integers such that n , p, then prove that
there are no simple groups of order pr n.
9. Let G be a group of order pn, where p is a prime and n Z1. Prove that there
are normal subgroups Ai for 0 i n such that |Ai| 5 pi and Ai ⊂ Ai11 for all 0
i n.
10. Deduce from above that there are no simple groups of order pn, for any prime p
and n $ 2.
13. Prove that there is exactly one, up to isomorphism, group of order 323.
17. Determine all the conjugacy classes in S4 and write down the class equation of S4.
19. Let G be a group of order 341. Prove that any subgroup of order 31 is normal in G.
22. Let p be a prime and n Z1 such that p . n. Prove that any subgroup of order
p in a group G of order pn is normal in G.
23. If a group G contains a proper subgroup of finite index, then prove that G con-
tains a proper normal subgroup of finite index.
25. Let G be an infinite p-group. Then prove that either G has a subgroup of order
pn for each positive integer n or there exists a positive integer m such that every
finite subgroup of G is of order pm.
27. Let p and q be distinct primes and p . q. Prove that any group of order pnq con-
tains a unique normal subgroup of index q.
28. Prove that any finite abelian group of square-free order is cyclic (an integer is
said to be square-free if it is not divisible by m2 for any integer m . 1).
30. Use the above to find all Sylow p-subgroups of Z30, for each prime p.
It is well known that any cyclic group is abelian and the product of any class
of abelian groups is abelian. In this chapter, we prove the celebrated theo-
rem known as the Fundamental Theorem of finitely generated abelian groups
which states that any finitely generated abelian group is a product of finite
number of cyclic groups. This amounts to saying that the cyclic groups are
like ‘building blocks’ for the finite or finitely generated abelian groups. Since
any cyclic group is isomorphic to the group Z of integers or the group Zn
of integers modulo n for some positive integers, the Fundamental Theorem
implies that any finitely generated abelian group is isomorphic to the product
of a finite number of copies of Z and Zn’s. This facilitates to a great extent
the study of finitely generated abelian groups and, in particular, finite abelian
groups. In fact, we derive a precise formula for the number of abelian groups
of a given order n in terms of the partitions of n.
for any (a1, a2, …, an) and (b1, b2, …, bn) G. In this group, (e1, e2, …, en)
is the identity, where ei is the identity in Gi and, for any (a1, a2, …, an) G,
( a11 , a2 1 , …, an 1 ) is the inverse of (a1, a2, …, an), where ai−1 is the inverse
of ai in Gi. This group G is called then direct product, or simply, the product
of G1, G2, …, Gn and is denoted by Gi or G1 G2 Gn . If a group H
i1
is isomorphic to G1 3 G2 3 … 3 Gn, then H is said to be decomposed into
product of groups G1 3 G2 3 … 3 Gn. In this section, we obtain equivalent
conditions for the decompositions of a group G into products of groups in
terms of normal subgroups of G and the corresponding quotient groups.
If A and B are normal subgroups of a group G such that AB 5 G and
A ∩ B 5 {e}, then we have proved (see 7. …) that the map (a, b) ab is
an isomorphism of A 3 B onto G. This is extended further in the following
theorem.
x 5 (x1, e2, …, en)(e1, x2, e3, …, en) … (e1, e2, …, en–1, xn)
Define g : N1 3 N2 3 … 3 Nn → G by
G N1 N 2 N n G1 G2 Gn .
Therefore, ab(ba)–1 5 ab(a–1b–1) 5 e so that ab 5 ba. Now, for any (a1, …, an)
and (b1, …, bn) N1 3 … 3 Nn,
g (( a1 , …, an )(b1 , …, bn )) g ( a1b1 , …, an bn )
( a1b1 )( a2 b2 ) ( an bn )
a1 (b1a2 )(b2 a3 ) (bn1an )bn
a1a2 an b1b2 bn
g ( a1 , …, an ) g (b1 , …, bn ).
g(a1, …, an) 5 e ⇒ a1 a2 … an 5 e
G1 3 G2 3 … 3 Gn N1 3 N2 3 … 3 Nn G. b
Corollary 8.1.2. Let N1, N2, …, Nn be subgroups of a group G. Then, the map-
ping (a1, a2, …, an) a1a2…an is an isomorphism of N1 3 N2 3 … 3 Nn onto
G if and only if the following are satisfied.
1. Each Ni is a normal subgroup of G.
2. N1N2 … Nn 5 G.
3. Ni ∩ (N1…Ni–1 Ni11…Nn) 5 {e} for each 1 # i # n.
We obtain another characterization of a decomposition of a group G in terms
of its quotient groups. First recall that quotient groups of G are precisely (up to
isomorphism) the homomorphic images of G. If N is a normal subgroup of a
group G, then the natural map x xN is an epimorphism of G onto the quotient
group G/N. If N1 and N2 are two normal subgroups, then clearly the map x
(xN1, xN2) is a homomorphism of G into G/N1 3 G/N2. We obtain a necessary and
sufficient condition for this map to be a surjection in the following theorem.
where a1 and b1 N1 and a2 and b2 N2. Now, put x 5 b1a2. Then, we have
Then, aN1 5 N1 and aN2 5 xN2 and hence a N1 and a–1x N2 so that
x 5 a(a–1x) N1N2. Thus, N1N2 5 G. b
G N1 ∩ N 2 G N1 G N 2
N i ∩ N j G.
j≠i
Therefore, xNj 5 aNj for all j i and Ni 5 aNi and hence a Ni and a–1x
Nj for all j i, so that
(
x a( a−1 x ) ∈ N i . ∩ N j .
j ≠i
)
Thus, G N i ( ∩ N j ).
j ≠i
Conversely suppose that N i ( ∩ N j ) G for each 1 # i # n. Then, NiNj 5 G
j ≠i
for all i j and
m−1
∩ N i N m G for all 1 m n
i1
G G G
f: n
→
N1 Nn
∩N
i=1
i
by
n
f x ∩ N i ( xN1 , , xN n ).
i =1
(
N i ⋅ ∩ N j G.
j ≠i
)
x 5 (x1, …, xi–1, ei, xi11, …, xn) (e1, …, ei–1, xi, ei11, .., en)
and hence a 5 e. From this we get that, for any i j, ai Ai and aj Aj,
Worked Exercise 8.1.2. Let G be a finite nontrivial group such that a2 5 e for
all a G. Then prove that G C1 3 C2 3 … 3 Cn, where n . 0 and each Ci
is a cyclic group of order 2 and deduce that |G| 5 2n.
C1 C1 C2 C1 C2 C3 …
Since G is finite, the process should terminate at a finite stage and then
G C1 3 C2 3 … 3 Cn.
Worked Exercise 8.1.3. Prove that any group of order p2 where p is a prime,
is isomorphic to either Z p 2 or Zp 3 Zp.
| A || B | | A | | B |
| AB | = = = p2 = | G |
| A∩ B | 1
G A 3 B Zp 3 Zp,
G A1 3 A2 3 … 3 An.
EXERCISE 8(a)
1. State whether each of the following is true or false and substantiate your
answer.
(i) Any group of order 25 is cyclic.
(ii) If G is a group of order 9 and G is not cyclic, Then, G Z3 3 Z3.
(iii) Any group of order 121 is abelian.
(iv) Any group of order 8 is abelian.
(v) Any cyclic group of order 180 can be decomposed as a product of non-
trivial groups.
(vi) Z36 Z9 3Z4.
(vii) If G is a group of order 36, than, G Z9 3 Z4.
(viii) Z Z .
12Z 4Z 3Z
2. Prove that Z4 cannot be decomposed as a product of groups of order 2.
5. Let G be a group of order pq, where p and q are distinct primes. If A and B are
normal subgroups of orders p and q, respectively, then prove that G is cyclic.
7. Let G be a cyclic group of order mn, where m and n are relatively prime positive
integers. Prove that there exist subgroups A and B of orders m and n, respectively
such that G A 3 B.
8. Is Z 3 Z cyclic?
9. Let G be a finite abelian group of order p1r1 p2 r2 … pn rn , where p1, p2, …, pn are
distinct primes and r1, r2, …, rn are positive integers. For each 1 # i # n, let
Si 5 {a G : O(a) 5 ps for some 0 # s Z}.
10. Let m p1r1 p2 r2 … pn rn , where pi’s are distinct primes ri’s are positive integers.
Prove that
Z m Z pr1 Z pr2 Z prn .
1 2 n
12. Prove that the group of symmetries of a square, the group (Z, 1) and the group
(Q, 1) are all indecomposable.
13. Prove that the group (Zn, 1n) is indecomposable if and only if n 5 pr for some
prime p and r . 0.
14. Let A and B be normal subgroups of a finite group G such that |A| ? |B| 5 |G| and
|A| and |B| are relatively prime. Then prove that G A 3 B.
15. Let A and B be subgroups of a group G such that G A 3 B. For any normal
subgroup N of G, prove that either N is contained in the centre of G or N has
nontrivial intersection with A or B.
18. Let G1, G2, …, Gn be groups and G 5 G1 3 G2 3 … 3 Gn. For any group H,
prove that a mapping f : H → G is a homomorphism if and only if pi ? f : H → Gi
is a homomorphism for each 1 # i # n, where pi : G → Gi is the ith projection.
19. Let A and B be any normal subgroups of a group G. If the natural map f : G →
G (defined be f (x) 5 xA) induces an isomorphism of B onto G/A, then prove
A
that G A 3 B.
20. Prove the following for any groups G1, G2 and G3:
(i) G1 3 G2 G2 3 G1
(ii) (G1 3 G2) 3 G3 G1 3 (G2 3 G3)
(iii) G1 G2 ⇒ G1 3 G3 G2 3 G3
(iv) G1 3 G2 is abelian ⇔ G1 and G2 are abelian.
(v) G1 3 G2 is cyclic ⇒ G1 and G2 are cyclic.
(vi) The converse of (v) is not true.
We first consider the case of finite abelian groups. Let us recall that a
group G is called a p-group, where p is a given prime number, if the order of
any element of G is a power of p and that a finite group is a p-group if and
only if it is of order pn for some nonnegative integer n. The following is a
central topic in the structure theory of finite abelian groups.
Proof: The trivial group {e} is a p-group, for any prime p (since its order is
1 5 p0). Let G be a nontrivial abelian group of order n and n . 1. We can
write
n p1r1 p2 r2 … pk rk ,
where p1, p2, …, pk are distinct primes and r1, r2, …, rk are positive integers.
From Sylow Theorem I (7.4.4), there exist subgroups A1, A2, …, Ak of G such
that | Ai | pi ri for 1# i # k. Since the group G is abelian, each Ai is a normal
subgroup of G. Also, since pi’s are distinct primes, it can be easily verified
that Ai ∩ Aj 5 {e} (for, the order of Ai ∩ Aj is a common divisor of |Ai| and
|Aj|) for i j.
Further the order of Ai ∩ (A1…Ai21Ai11…Ak) is a common divisor of
|Ai| 5 (5 piri ) and |A1…Ai21Ai11…Ak| ( p j j ) and, since piri and p j j are
s s
j ≠i j ≠i
relatively prime, it follows that |Ai ∩ (A1…Ai21Ai11…Ak| 5 1 and hence
Ai ∩ (A1…Ai21Ai11…Ak) 5 {e}.
G A1 3 A2 3 … 3 Ak
Definition 8.2.1. Let G be a finite abelian group. For any prime p, let
G G p1 G p2 G pk ,
where p1, p2, …, pk are all the distinct primes dividing the order of G.
The two results proved above reduce the study of arbitrary finite abelian
groups to the study of finite abelian p-groups. The basic result on p-groups
from which the whole structure theory can be pinned down is proved in the
following theorem.
Theorem 8.2.2. Let p be an arbitrarily fixed prime number. Then, any finite
abelian p-group is isomorphic to a product of a finite number of cyclic
p-groups.
Proof: Let G be a finite abelian p-group. Then, |G| 5 pn for some nonnega-
tive integer n. We shall use induction on n. If n 5 0, there is nothing to prove,
since G becomes trivial. If n 5 1, then G is a group of order p, which is a
prime, and hence G is itself a cyclic p-group. Now, let n . 1 and suppose that
the theorem holds good for all groups of order pm with m , n.
Since G is a p-group, the order of any element of G is a power of p. Let a be
an element of maximal order in G and O(a) 5 pk, where k # n. Put H 5 ,a.,
the cyclic subgroup of G generated by a. If k 5 n, then H 5 G and hence G
is itself a cyclic p-group. Suppose that k , n and consider the quotient group
G/H whose order is pn2k and n 2 k , n. By the induction hypothesis,
G/H G1 3 G2 3 … 3 Gr ,
Ai ∩ (A1…Ai21Ai11…Ar) 5 {H}.
Since Gi and hence Hi/H is cyclic, we get a coset biH generating Hi/H. Let |H/
Hi| 5 p ji for each 1 # i # r. Then, the order of biH in G/H is equal to p ji .
In the next step, we produce a representative ci of biH such that the order of ci
in G is equal to the order of biH in G/H. j
For convenience, let us write tempo-
rarily b for bi and j for ji. Since (bH ) p 5 H, we have
j
b p ∈ H a
j
and hence b p 5 as for some s. Since kthe order of a is maximal and O( a ) = p k ,
we have O(b) # pk and therefore b p e which implies that
k− j k− j j k− j k
a sp ( a s ) p (b p ) p b p e.
Therefore, O(a) divides spk2j ; that is, pk divides spk2j and hence pj divides s.
Let s 5 tpj and c 5 ba2t. Then, b21c 5 a2t ,a. 5 H and
j j j j
c p (bat ) p b p at p a s as e
Let K be the subgroup of G generated by c1, c2, …, cr. We shall prove that
HK 5 G and H ∩ K 5 {e}, so that G H 3 K.
Consider an arbitrary element x G. Then, xH G/H and hence
Now,
H xH (c1 H ) m1 (c2 H ) m2 ( cr H ) mr
(b1 H ) m1 (b2 H ) m2 (br H ) mr .
Since the order of biH in G/H is equal to the order of ci (which is equal to p ji ),
we get that p ji divides mi and hence (ci ) mi 5 e for all 1 # i # r, so that x 5
e. Thus, H ∩ K 5 {e}. Again by Theorem 8.1.1, we have
G H 3 K.
From the induction hypothesis, there exist cyclic p-groups K1, K2, …, Kw
such that
K K1 3 … 3 Kw.
G H 3 K H 3 K1 3 … 3 Kw.
where p1, p2, …, pr are (not necessarily distinct) primes and n1, n2, …, nr are
positive integers.
We derive a formula to find the exact number of distinct (nonisomorphic)
abelian groups of a given order n. For example,
Z4 and Z2 3 Z2
Z8, Z4 3 Z2 and Z2 3 Z2 3 Z2
are all the distinct abelian groups of order 8. Before going for the derivation
of the formula, we collect few miscellaneous facts about abelian groups that
will be used in the derivation of the formula.
For any abelian group G and for any integer m, the sets
Gm 5 {am : a G}
and G(m) 5 {a G : am 5 e}
are subgroups of G. For any prime p and a positive integer n, it can be easily
verified that
m
Z ppn Z pnm for any m n
and Z pn ( p) Z p .
and G ( m) G1 ( m) G2 ( m) Gr ( m)
Definition 8.2.2. Let n be any positive integer. A finite sequence {n1, n2, …, nr}
of positive integers is said to be a partition of n if
n1 # n2 # … # nr and n1 1 n2 1 1 nr 5 n.
Theorem 8.2.4. Let p be a prime and n be a positive integer. Then, there are
exactly P(n) number of distinct (nonisomorphic) abelian groups of order pn,
where P(n) is the number of partitions of n.
Proof: For any partition {n1, n2, …, nr} of n, consider the product
Z pn1 Z pn2 Z pnr
is a bijection between the set of all partitions on n and the set of all distinct
(nonisomorphic) abelian groups of order pn. If G is an abelian group of order
pn, then, by Theorem 8.2.2,
G (p)
(Z pm1 Z pm2 Z pms )(p)
Z p Z p Z p
(s times)
nj nj
G p (Z pn1 Z pn2 Z pnr ) p
nj nj nj
Z ppn1 Z ppn2 Z ppnr
Z n j +1n j ×Z n j +2n j Z nr n j
p p p
nj
Gp Z m j n j Z m j +1n j Z mr n j
p p p
Corollary 8.2.3. Let m p1n1 p2n2 … prnr , where p1, p2, …, pr are distinct primes
and n1, n2, …, nr are positive integers. Then, the number of distinct (noniso-
morphic) abelian groups of order m is equal to p(n1) p(n2) … p(nr), where
p(ni) is the number of partitions of ni.
G1 3 G2 3 … 3 Gr ,
Corollary 8.2.4. Let m be any square-free positive integer (that is, m is not
divisible by any perfect square greater than 1). Then, any abelian group of
order m is cyclic and hence Zm is the only (up to isomorphism) abelian group
of order m.
m 5 p1 p2 … pr
where p1, p2, …, pr and distinct primes. By Corollary 8.2.3, the number of
distinct abelian groups of order m is p(1) p(1)…p(1) 5 1. We know that the
group Zn of integers modulo n is an abelian group of order n. Thus, Zn is the
only (up to isomorphism) abelian group of order n. b
Worked Exercise 8.2.1. Find the number of abelian groups of order 7,200.
are the only partitions of 5 and hence p(5) 5 7. Thus, there are exactly 7 3 2 3
2 (5 28) abelian groups of order 7,200.
Worked Exercise 8.2.2. Prove that any abelian group of order 2,310 is cyclic.
Worked Exercise 8.2.3. List all (up to isomorphism) abelian groups each of
order 240.
240 5 24 ? 31 ? 51
Z2 3 Z2 3 Z2 3 Z2 3 Z3 3 Z5
Z2 3 Z2 3 Z 22 3 Z3 3 Z5
Z2 3 Z 23 3 Z3 3 Z5
Z 22 3 Z 22 3 Z3 3 Z5
Z 24 3 Z3 3 Z5
Z2 3 Z2 3 Z2 3 Z30 ( Z2 3 Z2 3 Z6 3 Z10)
Z2 3 Z2 3 Z60 ( Z2 3 Z6 3 Z20)
Z2 3 Z120 ( Z8 3 Z30)
Z12 3 Z20
Z240
for some integers n1, n2, …, nr. In fact, for any subset S of G, we have
∑a
i1
i for a1 1 a2 1 1 an,
where a1, a2, …, an are any elements of an abelian group (G, 1).
G A1 3 A2 3 3 Ak,
where each Ai is a nontrivial cyclic group such that either all of the Ai’s are
infinite or for some s, 1 # s # k, A1, A2, …, As are of orders m1, m2, …, ms
respectively with mi divides mi11 for each 1 # i , s and As11, …, Ak are
infinite.
Proof: Since (G, 1) is a finitely generated abelian group, there are finite sets
generating G. Let k be the least positive integer such that G has a k-element
generating set. We shall use induction on k. If k 5 1, then G is generated by
a single element and hence G itself is cyclic. Suppose that k . 1 and assume
that the theorem is true for all abelian groups generated by a set of k 2 1
elements.
If G has a generating set {a1, a2, …, ak} such that
G Z 3 3 Z (r copies).
Next, suppose that G has no generating set {a1, …, ak} satisfying the property
(*). Then, for any generating set {a1, …, ak}, there exists integers n1, …, nk,
not all zero, such that n1a1 1 1 nkak 5 0. Since
k k k
∑ n a 0 ⇒ ∑ (n )a ∑ n a 0.
i1
i i
i1
i i
i1
i i
There is an equation of the form n1a1 1 1 nkak 5 0 with one of ni’s posi-
tive. Let T be the set of all positive integers occurring in equations of the form
n1a1 1 1 nkak 5 0, where {a1, …, ak} is a generating set for G. The above
discussion implies that T is a nonempty set of positive integers. Let m1 be the
least positive integer in T. We can assume that
for some generating set {a1, a2, …, ak} for G and integers n2, …, nk. We shall
prove that m1 divides each ni, 2 # i # k. By the division algorithm, let us
write
Therefore, b1 0. Also,
a1 5 b1 2 q2a2 2 2 qkak
and therefore {b1, a2, …, ak} is a k-element generating set for G and r2, …, rk
are nonnegative integers less than m1 occurring in Equation (2). By the least
property of m1, it follows that r2 5 r3 5 5 rk 5 0 and hence mi 5 qim1 for
2 # i # k. Also, from Equation (2), we have
m1b1 5 0.
H1 A2 3 3 Ak
where each Ai is a cyclic group such that either all the A2, A3, …, Ar are infi-
nite or for some s, 2 # s # r, A2, …, As are of orders m2, …, ms respectively
with mi divides mi11 for all 2 # i , s and As11, …, Ar are infinite. The proof
is complete, if we can show that m1 divides m2 also. To do this, let Ai 5 ,bi.
and bi be of order mi for 2 # i , k. Then, {b1, b2, …, bk} is a generating set
for G and
G A1 3 H1 A1 3 A2 3 3 Ak
where A1, A2, …, Ak are cyclic groups satisfying the required properties. This
completes the proof of the theorem. b
Since any infinite cyclic group is isomorphic to the group (Z, 1) and any
finite cyclic group of order n is isomorphic to the group (Zn, 1n), the follow-
ing is an immediate consequence of the above theorem.
G Z Z Z m1 Z ms
(r components)
where r and s are nonnegative integers and m1, …, ms are positive integers
such that mi divides mi11 for all 1 # i , s.
Worked Exercise 8.2.5. State and prove the converse of Corollary 8.2.5.
Z m1 Z ms
G 5 G1 3 G2 3 … 3 Gn
since each Gi is abelian, the product G is also abelian. Let ai be a generator for
Gi and ei be the identity in Gi. Put xi be the element in G whose ith component
is ai and other components are identities. That is,
g ( g1 , ..., g n ), gi ∈ Gi
( a1r1 , ..., anrn ), ri ∈
x1r1 x2r2 ... xnrn ∈< x1 , ..., xn > .
Answer: If G is finite and a G, then O(a) # |G| (in fact, O(a) is a divisor of
|G|) and hence O(a) is finite. Conversely, suppose that O(a) is finite for every
a G. By Corollary 8.2.5,
G Z Z Z m1 Z ms
(r components),
where r and s are nonnegative integers and m1, …, ms are positive integers. If
r . 0, then consider
a 5 (2, 0, …, 0, 0, …, 0)
G Z m1 Z ms
EXERCISE 8(b)
1. State whether each of the following is True or False. Substantiate your answer.
(i) Z2 Z2 Z4 .
(ii) Z3 Z12 is cyclic.
(iii) Z6 Z 25 is cyclic.
(iv) There is exactly one abelian group of order 105.
(v) There is exactly one group of order 30.
(vi) Any abelian group of order 165 is isomorphic to Z165.
(vii) The number of abelian groups of order 24 is 3.
(viii) The number of groups of order 24 is 3.
(ix) Any abelian group of order divisible by 7 contains a cyclic subgroup of
order 7.
(x) Any abelian group of order divisible by 9 contains a cyclic subgroup of
order 9.
2. Prove that the following are equivalent to each other for any positive integers m
and n.
(a) Z m n is cyclic.
(b) Z m Z n Z mn .
(c) m and n are relatively prime.
(d) Z m, n , the subgroup generated by m and n in the group (Z, 1).
3. Prove that any abelian group of order 8 is isomorphic to one of the following.
Z8 , Z 4 Z 2 , Z 2 Z 2 Z 2 .
4. Let G1, G2, …, Gn be groups and ai Gi, 1 # i # n. Let a 5 (a1, a2, …, an). Prove
that O(a) is finite in the product G1 3 G2 3 … 3 Gn if and only if O(ai) is finite
in Gi for each 1 # i # n and, in this case, O(a) 5 l.c.m. {O(a1), …, O(an)}.
6. Let Z(G) denote the centre of any group G. For any groups G1, G2, …, Gn, prove
that Z(G1 3 G2 3 … 3 Gn) 5 Z(G1) 3 Z(G2) 3 … 3 Z(Gn).
8. Let N1 and N2 be normal subgroups of groups G1 and G2, respectively. Prove that
N1 3 N2 is a normal subgroup of G 1 3 G2 and that
G1 N1 G2 N 2 ≅ (G1 × G2 ) N1 × N 2 .
11. Prove that any finite abelian p-group is generated by its elements of highest
order.
12. Show that any homomorphic image of a p-group is a p-group and product of
p-groups is also a p-group.
13. Let G be a finite cyclic group and p be a prime dividing the order of G. Prove that
there are exactly p 2 1 elements in G each having order p.
14. Prove that a nontrivial finite abelian group is cyclic if and only if it is isomorphic
to Z p n1 Z p n2 Z p nr for some distinct primes p1, …, pr and positive integers
1 2 r
n1, …, nr.
15. Determine all (up to isomorphism) abelian group of order 144, 625 and
1,94,481.
16. Prove that a cyclic group is indecomposable if and only if it is either infinite or
of prime power order.
17. Describe all the positive integers n for which there is exactly one (up to isomor-
phism) abelian group of order n.
18. Let {a1, a2, …, ar} be a generating set for an abelian group G. Prove that the
following are equivalent to each other.
r
(1) ( n1 , n2 , ..., nr ) ni ai is an isomorphism of Z Z Z ( r copies)
i1
onto G.
(2) Any element a of G can be uniquely expressed as
a 5 n1a1 1 … 1 nrar, for integers n1, n2, …, nr.
(3) For any integers n1, n2, …, nr,
n1a1 1 n2a2 1 … 1 nrar 5 0 ⇒ n1 5 n2 5 … 5 nr 5 0
19. Let G be a finitely generated abelian group. Prove that G is finite if and only if G
is isomorphic with a product of finitely many cyclic groups.
20. Let G be a nontrivial abelian group. Prove that G is finite if and only if
G Z m Z m
1 r
for some positive integers m1, …, mr such that mi . 1 and mi divides mi11 for all
1 # i , r.
1 , m1|m2| … |mr.
In general, if a and b are integers, we write a to say that a divides b; that is,
b
ac 5 b for some integer c.
Theorem 8.3.1. Let G be a nontrivial finite abelian group. Then, there exists
a unique division sequence
1 , m1|m2| … |mr
Proof: By Theorem 8.2.6, there exists nontrivial finite cyclic groups G1, G2,
…, Gr of orders m1, m2, …, mr, respectively such that mi divides m1, m2, …, mr
respectively such that mi divides mi11 for each 1 # i , r and
G G1 3 G2 3 … 3 Gr .
Since any cyclic group of order m is isomorphic to Zm, we get that 1 , m1|m2|
… |mr is a division sequence and
G Z m1 Z m2 Z mr . (1)
G Z n1 Z n2 Z ns . (2)
We shall prove that r 5 s and mi 5 ni for each 1 # i # r. First note that the
order of any element of Z mi is a divisor of mi and hence that of G is a divisor
of the l.c.m. {m1, m2, …, mr} which is equal to mr (since mi divides mr for all
1 # i # r). Therefore, the order of any element of G is a divisor of mr. Since
the element (0, …, 0, 1) is of order ns in Z n1 Z ns G , there exists an
element of order ns in G and therefore ns divides mr. Similarly, by symmetry,
we can prove that mr divides ns and hence mr 5 ns. From the decompositions
(1) and (2), we have
G Z m1 Z m2 Z mr
then G is said to be of type (m1, m2, …, mr) and the integers m1, m2, …, mr
are called the invariants of G. The following is an immediate consequence of
Theorem 8.3.1.
Corollary 8.3.1. Let n be a positive integer greater than 1. Then, the number
of abelian groups of order n is equal to the number of division sequences 1 ,
m1|m2| … |mr such that n 5 m1 m2 … mr.
G Z 6 Z8 Z 5 Z 2 Z 3 Z8 Z 5
Z 2 Z120
G Z 2 Z 4 Z 9 Z 27 Z81 .
G Z 9 Z54 Z324
and 9|54|324 is a division sequence. Therefore, 9, 54, 324 are the invari-
ants of G.
The following provides an algorithm to find the invariants of a given finite
abelian group. Let us recall that a partition of a positive integer n is a finite
sequence {n1, n2, …, nr} of positive integers such that n1 # n2 # … # nr and
n1 1 n2 1 … 1 nr 5 n. For the convenience and for the purpose of proving
the following result, we relax the definition of a partition of n by including
certain zeros in the beginning. Accordingly, a partition of n is a sequence of
nonnegative integers {n1, n2, …, nr} such that 0 # n1 # n2 # … # nr and
n1 1 n2 1 … 1 nr 5 n.
Theorem 8.3.2. Let G be an abelian group of order n p1m1 p2m2 pkmk where
p1, p2, …, pk are distinct primes and m1, m2, …, mk are positive integers. For
each 1 # i # k, let {ni1 , ni2 , ..., nir } be a partition of mi such that
n n
For each 1 # j # r, let s j p1 1 j … pk kj . Then, s1, s2, …, sr are the invari-
ants of G.
Gj Z n Z n … Z n , for 1 # j # r.
p11 j p22 j pk kj
n n n
Since p1, p2, …, pk are distinct primes, it follows that p1 1 j , p2 2 j , ..., pk kj are
pair-wise relatively prime and hence
n n n
G j Z s j , where s j = p1 1 j p2 2 j … pk kj .
Example 8.3.2. Let G be the set of all positive integers less than 100 and rela-
tively prime to 100. Then, G is an abelian group under multiplication modulo
100. Let us find the invariants of G.
First, we shall list all elements of G and find the order G. We have |G| 5
f (100) 5 40 5 2335
G 5 {1, 3, 7, 9, 11, 13, 17, 19, 21, 23, 27, 29, 31, 33,
37, 39, 41, 43, 47, 49, 51, 53, 57, 59, 61, 63, 67,
69, 71, 73, 77, 79, 81, 83, 87, 89, 91, 93, 97, 99}.
One can easily verify the G 5 ,3., the cyclic subgroup generated by 3.
Therefore, G is a cyclic group of order 40 and hence G Z40. 40 is the only
invariant of G.
EXERCISE 8(c)
2. Let X be a set with 5-elements and G 5 (P(X), ⊕), where P(X) is the power set of
X and ⊕ is the symmetric difference operation. Determine the invariants of G.
3. Let G be the group of all positive integers less than 47 and relatively prime to 47,
under the multiplication modulo 47. Then find the invariants of G.
4. Let G be the group of all mappings of a 4-element set into the group (Z4, 14)
under point-wise operation. Determine all the invariants of G.
5. Determine the invariants of each abelian group of order less than or equal to 30.
| A || B | p⋅q
| AB | = | G |
| A∩ B | 1
Since the solutions of xp 1 (mod q), x 1 (mod q), are r, r2, …, rp21, it fol-
lows that G G9, since replacing a by aj as a generator of ,a. replaces r by
rj. This completes the proof.
Corollary 8.4.1. Let G be a group of order pq, where p and q are primes such
that p , q. Then, either G is cyclic or G is a nonabelian group generated by
two elements a and b satisfying the properties (1) and (2) above.
Now, we list all groups of order less than or equal to 20. Some of the
proofs involved in the listing are left as exercises to the reader.
a 5 (1 2) and b 5 (1 2 3),
a2 5 e 5 b3, a e, b e and
a21ba 5 (1 2) (1 2 3) (1 2) 5 (1 3 2) 5 b2 (see Theorem 8.4.1).
D6 S3 3 Z2 D3 3 Z2.
It is well known that there are two familiar binary operations, namely the
addition 1 and the multiplication ? on the set Z of integers and that (Z, 1) is
an abelian group where as (Z, ?) is only a semigroup. We have earlier worked
with algebraic systems, namely semigroups, monoids and groups, where
there is only one binary operation in each. Now, in this chapter, we initiate
the study of abstract algebraic systems having two binary operations as in the
case of integers. Also, we have the rational number system, the real number
system, the complex number system, the set of all n 3 n matrices, where in
each of these cases we have two binary operations satisfying certain connec-
tive properties in addition to the properties satisfied by the individual opera-
tions. We introduce a common abstraction of these in the form of a ring and
develop a general elementary theory of rings. A ring is basically a combina-
tion of an abelian group and a semigroup and therefore a previous knowledge
of groups and semigroups will be of considerable help. Most of the important
concepts in group theory have natural extensions to ring theory.
Recall from the group theory that the element 0 in (I.3) above is unique
and is called the additive identity or the zero element in R. Also, for any
a R, the element 2a in (I.4) above is unique and is called the additive
inverse of a.
In order to minimize the use of parentheses (brackets) in expressions
involving both the operations 1 and ?, let us stipulate that multiplication is to
be performed before addition. Accordingly, the expression a ? b 1 c stands
for (a ? b) 1 c and not for a ? (b 1 c). Also, because of the generalised asso-
ciative law, parentheses can also be omitted when we write sums or products
of more than two elements. For example, we write
a1 ? b1 1 a2 ? b2 1 a3 1 a4 ? b4
instead of ((a1 ? b1) 1 (a2 ? b2)) 1 (a3 1 (a4 ? b4)).
Example 9.1.1
1. Let Z be the set of all integers, Q be the set of all rational numbers and
R be the set of all real numbers. Then, (Z, 1, ?), (Q, 1, ?) and (R, 1, ?)
are all rings, where 1 and ? are the usual addition and multiplication of
real numbers. In each of these cases, the number 0 is the zero element
(that is, the additive identity).
2. Let (G, 1) be an abelian (commutative) group in which 0 is the identity
element. Define a ? b 5 0 for all a and b G. Then, (G, 1, ?) is a ring,
since ? is clearly associative and distributive over 1, for,
a ? (b 1 c) 5 0 5 0 1 0 5 (a ? b) 1 (a ? c)
and (a 1 b) ? c 5 0 5 0 1 0 5 (a ? c) 1 (b ? c)
for any a, b and c G. The multiplication here is called the trivial mul-
tiplication in an abelian group. Rings of this type are called zero rings.
3. Let M2(R) be the set of all 2 3 2 matrices over R (that is, with entries as real
a a b b
numbers).For any matrices A 11 12 and B 11 12 in M 2 ( R),
define a21 a22 b21 b22
a b a12 b12
A B 11 11
a21 b21 a22 b22
a b a b a11b12 a12 b22
and A B 11 11 12 21
a21b11 a22 b21 a21b12 a22 b22
4. Let C denote the set of all complex numbers; that C is the set of all
expressions of the form a 1 ib, where a and b are arbitrary real numbers.
For any x 5 a 1 ib and y 5 c 1 id in C, define
x 5 y ⇔ a 5 c and b 5 d
x 1 y 5 (a 1 b) 1 i(b 1 d)
x ? y 5 (ac 2 bd) 1 i(ad 1 bc),
where a 1 b, ac, etc. are the sums and products in the ring (R, 1, ?).
Then, (C, 1, ?) is a ring and is called the ring of complex numbers. Note
that 0 1 i0 is the zero element in C, which is also denoted simply by 0.
Further, the additive inverse 2x of x 5 a 1 ib is given by
2x 5 (2a) 1 i(2b)
As usual, we denote a 1 i0 by a, 0 1 ib by ib and 0 1 i1 by i. As per this
notation, note that ii 5 21.
5. In the above example, the operation 1 is defined coordinate wise. If
multiplication is also defined as coordinate wise (considering a and b
as first and second coordinates of a 1 ib), then C together with these
coordinate wise addition and multiplication forms a ring.
6. The procedure in 5 above can be generalised as follows. Let (R1, 1, ?),
(R2, 1, ?), …, (Rn, 1, ?) be any rings and
R 5 R1 3 R2 3 … 3 Rn 5 {(a1, a2, …, an) : ai Ri for 1 i n}.
For any a 5 (a1, a2, …, an) and b 5 (b1, b2, …, bn) in R, Define
Then, (R, 1, ?) is a ring and is called the product of R1, R2, …, Rn and is
n
denoted by Ri or, simply, R1 3 R2 3 … 3 Rn. Note that, (0, 0, …, 0)
i1
is the zero element in the product, where 0 stands for the zero element in
each Ri. The additive inverse of any a 5 (a1, a2, …, an) is given by
7. Let (R, 1, ?) be any ring and X be any nonempty set. Let RX be the set
of all mappings of X into R. For any f and g RX, define f 1 g and f ? g :
X → R by
(f 1 g)(x) 5 f (x) 1 g(x)
and (f ? g)(x) 5 f (x) ? g(x), for all x X,
where the operations 1 and ? on the right side are those in R. Then, (RX,
1, ?) is a ring. The operations 1 and ? on RX defined above are called
point-wise addition and point-wise multiplication. The constant map
which maps each element of X onto the zero element in R will be the
zero element in RX and the additive inverse of f is defined by (2f )(x) 5
2f (x) for all x X.
8. Let n be any positive integer and consider the group (Zn, 1n), where
Zn 5 {0, 1, 2, …, n 2 1}
a b if a b n
a n b ,
a b n if a b n
for any a and b Zn. Note that a 1n b is precisely the remainder obtained
by dividing the usual sum a 1 b by n. Now, extend this to the multiplica-
tion also, by defining
a n b 5 r, where 0 r , n, ab 5 qn 1 r, q and r Z,
for any a and b Zn. Note that a n b is precisely the remainder obtained
by dividing the usual product ab by n. This operation ?n is called the
multiplication modulo n. Zn is a finite set with n elements and the addi-
tion 1n and multiplication ?n modulo n are given in the following table
for n 5 6.
16 0 1 2 3 4 5
0 0 1 2 3 4 5
1 1 2 3 4 5 0
2 2 3 4 5 0 1
3 3 4 5 0 1 2
4 4 5 0 1 2 3
5 5 0 1 2 3 4
(Z6, 16)
?6 0 1 2 3 4 5
0 0 0 0 0 0 0
1 0 1 2 3 4 5
2 0 2 4 0 2 4
3 0 3 0 3 0 3
4 0 4 2 0 4 2
5 0 5 4 3 2 1
(Z6, ?6)
It can be proved that (Zn, 1n, ?n) is a ring and is called the ring of integers
modulo n.
9. Let R be a set consisting of only one element, say R 5 {a}. Then, the
only way of defining binary operation on R is a 1 a 5 a and a ? a 5 a
and (R, 1, ?) is a ring in which a itself is the zero element and 2a 5 a.
This ring is called the trivial ring. When we say that R is a nontrivial
ring, it means that R contains atleast two elements.
10. Let X be any set and P(X) be the set of all subsets of X. For any A and B
in P(X), define
A 1 B 5 (A 2 B) ∪ (B 2 A)
and A ? B 5 A ∩ B.
Then, we shall prove that (P(X), 1, ?) is a ring. Recall that the empty set
is the zero element and that (P(X), 1) is an abelian group. Clearly ?
is associative. Also, for any A, B and C in P(X), we have
A ? (B 1 C) 5 A ∩ ((B 2 C) ∪ (C 2 B))
5 (A ∩ (B 2 C)) ∪ (A ∩ (C 2 B))
5 ((A ∩ B) 2 (A ∩ C)) ∪ ((A ∩ C) 2 (A ∩ B))
5 (A ∩ B) 1 (A ∩ C)
5 (A ? B) 1 (A ? C)
Since ∩ is a commutative operation, there is no need to verify that ? is
right distributive over 1. Thus, (P(X), 1, ?) is a ring.
In the following, we prove certain important elementary properties of
rings.
Theorem 9.1.1. Let (R, 1, ?) be a ring. Then, the following holds good for
any elements a, b and c in R.
1. 0a 5 0 5 a0, where 0 is the zero element in R.
2. a(2b) 5 2(ab) 5 (2a)b
3. (2a)(2b) 5 ab
4. a(b 2 c) 5 ab 2 ac
5. (a 2 b)c 5 ac 2 bc.
Proof:
1. We have
0 1 0a 5 0a 5 (0 1 0)a 5 0a 1 0a
and, by the cancellation law in the group (R, 1), it follows that 0 5 0a.
Also,
0 1 a0 5 a0 5 a(0 1 0) 5 a0 1 a0
and hence 0 5 a0.
2. Consider
ab 1 a(2b) 5 a(b 1 (2b)) 5 a0 5 0 (by (1))
and hence a(2b) is the additive inverse of ab. That is, 2(ab) 5 a(2b).
Similarly
ab 1 (2a)b 5 (a 1 (2a))b 5 0b 5 0 (by (1))
and therefore, (2a)b 5 2(ab).
3. We have (2a)(2b) 5 2((2a)(b))52(2(ab)) 5 ab
4. a(b 2 c) 5 a(b 1 (2c)) 5 ab 1 a(2c) 5 ab 2 ac
5. (a 2 b)c 5 (a 1 (2b))c 5 ac 1 (2b)c 5 ac 2 bc. b
Note that the additive operation 1 is always commutative in any ring R and
therefore, by a commutative ring R, we only mean that the multiplication is
commutative.
Definition 9.1.3. A ring (R, 1, ?) is said to be a ring with unity or a ring with
identity if there exists an element e in R such that
a ? e 5 a 5 e ? a for all a R.
Theorem 9.1.2. Let (R, 1, ?) be a ring with unity. Then, R is trivial if and
only if 0 5 1 in R (that is, the additive identity coincides with the multiplica-
tive identity in the ring R).
a 5 1a 5 0a 5 0
Example 9.1.2
1. Each of the rings (Z, 1, ?), (Q, 1, ?), (R, 1, ?), (C, 1, ?), (Zn, 1n, ?n) for
any n Z1 and (P(X), 1, ∩ ) is a commutative ring with unity. X is the
unity element in (P(X), 1, ∩), while 1 is the unity element in all these
other rings.
2. Any zero ring (see Example 9.1.1 (2)) is commutative and has no unity
element, unless it is trivial.
1 0
3. The ring M2(R) of 2 3 2 matrices is with unity, where
0 1
is the unity. However M2(R) is not commutative, for consider
0 2 1 0
A and B .
0 1 2 0
Worked Exercise 9.1.1. Prove that (Zn, 1n, ?n) is a commutative ring with
unity for any positive integer n, where 1n and ?n are addition and multiplica-
tion modulo n, respectively.
a ?n b 5 r, r ?n c 5 s
b ?n c 5 t and a ?n t 5 u.
Then, ab 5 qn 1 r, rc 5 q1n 1 s
bc 5 pn 1 t and at 5 p1n 1 u,
By the uniqueness of the quotient and the remainder in the division algo-
rithm, it follows that
qc 1 q1 5 ap 1 p1 and s 5 u.
In particular, (a ?n b) ?n c 5 r ?n c 5 s 5 u 5 a ?n t 5 a ?n (b ?n c).
and b 1n c 5 x, a ?n (b 1n c) 5 y
a ?n b 5 z and a ?n c 5 v
so that b 1 c 5 qn 1 x, ax 5 q9n 1 y,
ab 5 pn 1 z and ac 5 p9n 1 v,
a(b 1 c) 5 ab 1 ac
a ?n (b 1n c) 5 y 5 t 5 z 1n v 5 (a ?n b) 1n (a ?n c).
Thus, ?n is distributive over 1n. Therefore, (Zn, 1n, ?n) is a commutative ring.
If n 5 1, then Zn is trivial. If n 1, then 1 is the unit element in Zn. Thus,
(Zn, 1n, ?n) is a commutative ring with unity.
Worked Exercise 9.1.2. Prove that, for any set X, (P(X), 1, ∩) is a commuta-
tive ring with unity.
x 5 a 1 ib, y 5 c 1 id and z 5 r 1 is
Thus, ? distributes over 1. Also, 1(5 1 1 i0) is the unity in Z[i]. Thus, (Z[i],
1, ?) is a commutative ring with unity. Z[i] is called the ring of Gaussian
integers.
Recall the following from group theory.
0 if n 0
na ( n 1)a a if n 0.
(n)(a) if n 0
a if n 1
a n n1 .
a a if n 1
EXERCISES 9(a)
1. Which of the following are rings? Substantiate your answers (here 1 and are
usual addition and multiplication of numbers).
(i) (Z1, 1, ?)
(ii) (6Z, 1, ?)
(iii) (E, 1, ?), where E is the set of even integers.
(iv) (O, 1, ?), where O is the set of odd integers.
(v) (P(X), ∩, ∪), where P(X) is the power set of a set X.
(vi) (P(X), ∪, ∩)
(vii) (P(X), 1, ∪)
(viii) (Z[ 2], , ), where Z[ 2] {a b 2 : a, b Z}.
(ix) (Q[ 2], , ), where Q[ 2] {a b 2 : a and b are rational numbers}.
(x) (Zn, ?n, 1n)
(xi) (Q 2 {0}, 1, ?)
(xii) (R 2 Q, 1, ?)
4. Let (R, 1, ?) be a ring. Prove that the following are equivalent to each other.
(i) 1 distributes over ?.
(ii) R is trivial; that is, R 5 {0}.
(iii) (R, ?, 1) is a ring.
(iv) a 1 b 5 ab for all a and b in R.
5. Let a and b be two elements of a ring such that a b 5 b a. Prove the following
for any positive integer n.
(a 1 b)n 5 an 1 nc1 an21b 1 nc2 an22b2 1 … 1 ncn21 abn21 1 bn
n
n!
∑ ncr a nr b r , where ncr .
r0 r !( n r )!
6. For any prime p and a and b Zp, prove that (a 1 b)p 5 ap 1 bp.
a ⊕ b 5 a 1b 1 1
a b 5 a ? b 1 a 1 b.
Prove that (R, ⊕, ) is a ring with unity and that (R, 1, ) is commutative if and
only if (R, ⊕, ) is commutative.
9. Let (R, 1, ?) be a ring such that (a2 1 a)x 5 x(a2 1 a) for all a and x R. Then
prove that (R, 1, ?) is a commutative ring.
ab 5 ca ⇒ a 5 0 or b 5 c.
11. Prove that a ring (R, 1, ?) is commutative if the group (R, 1) is cyclic.
12. Let (R, 1, ?) be a ring and n be an integer such that n 1 and xn 5 x for all x R.
Then prove that, for any a and b R,
ab 5 0 ⇔ ba 5 0.
13. Prove that the set {0, 2, 4} is a commutative ring with unity with respect to addi-
tion and multiplication modulo 6.
Example 9.2.1
1. The zero element 0 and the unity, if it exists, in any ring are idempotents.
2. 3 and 4 are idempotents in (Z6, 16, ?6), since 3 ?6 3 5 3 and 4 ?6 4 5 4; 5
is not an idempotent, since 5 ?6 5 5 1 in Z6.
3. In the ring Z of integers, 0 and 1 are the only idempotents.
Example 9.2.2
1. (Z2, 12, ?2), the ring of integers modulo 2 is a Boolean ring.
2. For any set X, (P(X), 1, ∩) is a Boolean ring, since A ∩ A 5 A for all
A ⊆ X.
3. For any set X, the set Z 2X of all mappings of X into Z2 is a Boolean ring
under the point-wise operations (see Example 9.1.1 (7)), since Z2 is a
Boolean ring.
Note that examples (2) and (3) given in Example 9.2.2 are not different.
They appear to be the same in the sense given below.
1 if x A
A : X → Z2 by A ( x ) .
0 if x A
Theorem 9.2.2. For any elements a and b in a Boolean ring (R, 1, ?),
a 1 a 5 0; that is, a 5 2a
and ab 5 ba
a 1 b 5 (a 1 b)2 5 (a 1 b) (a 1 b)
5 a2 1 ab 1 ba 1 b2
5 a 1 ab 1 ba 1 b.
0 5 ab 1 ba.
0 5 aa 1 aa 5 a 1 a
Example 9.2.3
1. The zero element 0 in any ring R is nilpotent, since 01 5 0
2. 6 is a nilpotent element in Z8, since
63 5 (6 ?8 6) ?8 6 5 4 ?8 6 5 0
3. Except 0, no element in the ring of integers is nilpotent.
Example 9.2.4
1. The ring (Z, 1, ?) of integers has no zero divisors, since, for any integers
a and b, ab 5 0 only when a 5 0 or b 5 0.
2. 2 and 3 are zero divisors in (Z6, 16, ?6).
3. If X is a set and A is a nonempty proper subset of X, then A ∩ (X 2 A) 5 ,
A and X 2 A and hence A is a zero divisor in (P(X), 1, ∩).
4. Consider the ring M2(R) of 2 3 2 matrices. Let
1 0 0 0
A and B .
1 0 1 1
0 0 0 0 0 0
Then A B and B A .
0 0 2 0 0 0
Example 9.2.5
Theorem 9.2.3. Let (R, 1, ) be a ring with unity. If a and b are units in R,
then so is their product ab and (ab)21 5 b21a21. Also, the set U(R) of all units
in R forms a group under multiplication.
Worked Exercise 9.2.2. For any n 1, prove that any nonzero element in the
ring Zn is either a unit or a zero divisor.
n a
a = n n 0.
d d
Worked Exercise 9.2.3. Let a and b be elements in a ring R such that ab 5 ba.
Prove the following:
1. a 1 b is a nilpotent if a and b are nilpotents.
2. ab is a nilpotent if a or b is a nilpotent.
Answer:
1. Suppose that a and b are nilpotents. Then, there exist positive integers n
and m such that an 5 0 5 bm. Now, since ab 5 ba, we have
(a 1 b)m1n 5 am1n 1(m 1 n)c1am1n21b 1 … 1 (m 1 n)cm1nbm1n
mn
∑ ( m n)Cr a mnr b r .
r0
since as 5 0 for all s ≥ n and bt 5 0 for all t ≥ m and since, for any 0 r
m 1 n, either m 1 n 2 r ≥ n or r ≥ m (otherwise (m 1 n 2 r) 1 r ,
n 1 m, an absurd), we get that am1n2r 5 0 or br 5 0 for all 0 r m 1
n and therefore (a 1 b)m1n 5 0, so that a 1 b is a nilpotent.
2. If an 5 0, then (ab)n 5 anbn 5 0bn 5 0. Therefore, if a is a nilpotent, then
ab (5 ba) is also a nilpotent.
Worked Exercise 9.2.4. Prove that 1, 21, i and 2i are the only units in the
ring Z[i] of Gaussian integers.
EXERCISE 9(b)
1. Determine all the zero divisors, nilpotents, idempotents and units in each of the
following rings
(i) The ring R of real numbers.
(ii) Z 3 R, where Z is the ring of integers.
(iii) Z 3 Z.
(iv) The ring Z24 of integers modulo 24.
(v) Z15.
(vi) Z12 3 Z.
(vii) (P(X), 1, ∩), for any set X.
(viii) Z4 3 Z9.
2. In any nontrivial ring with unity, prove that no zero divisor is a unit.
3. Let a be a nonzero element in a commutative ring R. Prove that a is not a zero divi-
sor if and only if a satisfies the following cancellation law for any b and c in R:
ab 5 ac ⇒ b 5 c.
4. Let R be a Boolean ring with unity. Prove that the unity is the only unit in R0 and
the zero is the only nilpotent in R.
5. Let n be any integer greater than 1. The content of n is defined to be the product
of all distinct primes dividing n and is denoted by c(n). Prove that a Zn is a
nilpotent if and only if c(n) divides a.
6. Using 5 above, derive a formula for the number of nilpotents in the ring Zn of
integers modulo n.
7. For any integers a and n with 0 a , n, prove that a is an idempotent in Zn if
and only if a(a 2 1) is a multiple of n.
8. Let R1, R2, …, Rn be rings and R 5 R1 3 R2 3 … 3 Rn. For any a 5 (a1, a2, …,
an) R, prove that a is a nilpotent (idempotent) in R if and only if each ai is a
nilpotent (respectively idempotent) in Ri for 1 i n.
9. In 8 above, when each Ri is a ring with unity, prove that (a1, a2, …, an) is a unit
in R1 3 R2 3 … 3 Rn if and only if ai is a unit in Ri for each 1 i n.
10. If a and b are idempotents in a commutative ring R, prove that ab is also an
idempotent. Can a 1 b be an idempotent? If not, give a counter example.
11. Find all the solutions of x2 1 2x 1 4 5 0 in the ring Z6.
12. Prove that 0 is the only solution of x2 5 0 in a ring R if and only if R has no
nonzero nilpotents.
13. For any prime number p, prove that the set of all nonzero elements in Zp forms a
group under multiplication modulo p.
14. Prove that any nonzero nilpotent in any ring R is a zero divisor in R.
15. Let f be the Euler-totient function. Prove that the multiplicative group of units
in Zn is of order f (n), for any integer n 1.
16. Let R be a ring with unity and a R be a nilpotent. Then prove that 1 1 a is a unit.
17. Let R be a ring with unity and without zero divisors. For any a and b in R, prove
that ab 5 1 if and only if ba 5 1 and that a2 5 1 if and only if a 5 1 or 21.
a if n 1
na .
( n −1)a a if n 1
Theorem 9.3.1. Let (R, 1, ?) be ring and char(R) 5 n 0. Then, for any
integer m,
ma 5 (nr)a 5 r(na) 5 r0 5 0.
Since n is the least positive integer such that na 5 0 for all a R and since
r , n, it follows that r cannot be positive. Since 0 r, we get that r 5 0 and
hence m 5 qn. Thus, n divides m. b
Theorem 9.3.2. Let (R, 1, ?) be a ring with unity 1. Then, the characteristic
of R is precisely the order of the unity in the group (R, 1).
Proof: This follows from the fact that, for any a R and n Z,
na 5 n(1a) 5 (n1)a
Example 9.3.1
1. The characteristic of each of the rings Z, Q, R and C is zero, since for any
integer n 0 and for any nonzero real or complex number a, na 0.
2. char(Zn) 5 n for any positive integer n, where Zn is the ring of integers
modulo n.
3. char(Zn 3 Zm) is the least common multiple of m and n for positive inte-
gers m and n.
r 5 ms and r 5 nt
for some positive integers s and t. Then, for any element (a, b) in R 3 S,
we have
EXERCISE 9(c)
9.4 SUBRINGS
In this section, we deal with the situation where a subset of a ring constitutes a
ring again. Recall the set Z of integers is a subset of the ring (R, 1, ?) of real
numbers and Z itself is a ring under the addition and multiplication of real
numbers. This is abstracted in the following definition.
a and b S ⇒ a 2 b S and ab S.
Clearly {0} and the whole of R are subrings of any ring R and are called
trivial subrings and all other subrings (if they exist) are called nontrivial sub-
rings. A subring S of R is called a proper subring if S R.
Example 9.4.1
1. Z is a subring of the ring (Q, 1, ?) of rational numbers, Q is a subring of
the ring (R, 1, ?) of real numbers and R is a subring of the ring (C, 1, ? )
of complex numbers.
2. Let Y be a subset of a set X. Then, P(Y), the power set of Y, is a subring
of (P(X), 1, ∩).
3. For any nonnegative integer n, the set nZ of all integral multiples of n, is
a subring of the ring (Z, 1, ?) of integers. In particular, the set E of even
integers is a subring of (Z, 1, ?).
Note that a ring R may possess the unity (multiplicative identity) while a
subring may not possess and, even when a subring possesses unity then it may
be different from that of R. Consider the following examples.
Example 9.4.2
1. The set E of even integers is a subring of the (Z, 1, ?) of integers. Z has
unity, while E has no unity (there is no even integer e such that ea 5 a
for all even integers).
2. Let X be a set and Y be a proper subset of X. Then, P(Y) is a subring of (P(X),
1, ∩). X and Y are unit elements in P(X) and P(Y), respectively and Y X.
C(a) 5 {x R : ax 5 xa}.
a(x 2 y) 5 ax 2 ay 5 xa 2 ya 5 (x 2 y)a
and a(xy) 5 (ax)y 5 (xa)y 5 x(ay) 5 x(ya) 5 (xy)a
Therefore, C ( R) ∩ C ( a).
aR
EXERCISE 9(d)
3. Prove that S is a subring of the ring Z of integers if and only if S 5 nZ for some
nonnegative integer n.
4. Let n be a positive integer and Zn be the ring of integers modulo n. Prove that S
is a subring of Zn if and only if S 5 {0, m, 2m, …, (r 2 1)m} for some divisor m
of n and rm 5 n, r 0.
5. Let X be a subset of a ring R and (X) be the intersection of all subrings of R con-
taining X. Prove that (X) is the smallest subring of R containing X. (X) is called
the subring of R generated by X.
10. Let S and T be subrings of a ring R. Then prove that S ∪ T is a subring of R if and
only if either S ⊆ T or T ⊆ S.
11. Let # be a class of subrings of a ring R such that, for any S1 and S2 #, there exist
S3 # containing both S1 and S2. Then prove that ∪ S is a subring of R.
S
12. Let S be a nonempty subset of a finite ring R. Prove that S is a subring of R if and
only if
a and b S ⇒ a 1 b and ab S.
13. Let R be a ring which has no nonzero nilpotent elements. Prove that every idem-
potent in R is in the centre C(R).
14. Let R be a ring such that a2 1 a C(R) for all a R. Prove that R is a commuta-
tive ring.
15. Let (R, 1, ) be a ring of characteristic n 0 and (Zn, 1n, ?n) the ring of integers
modulo n. Define the operations 1 and on R 3 Zn by
(x, a) 1 (y, b) 5 (x 1 y, a 1n b)
and (x, a) (y, b) 5 (xy 1 ay 1 bx, a ?n b).
f (a 1 b) 5 f (a) 1 f (b)
and f (a b) 5 f (a) f (b) for all a and b R.
Note that the symbols 1 and occurring on the left sides of the above
equations denote the addition and multiplication in R where as 1 and occur-
ring on the right sides denote those in R9. This use of the same symbols for
the operations of addition and multiplication in two different rings need cause
no confusion provided the reader gives careful attention to the context if the
notation is employed. The following is the usual terminology we apply, as in
the case of group theory.
Definition 9.5.2
1. An injective homomorphism is called a monomorphism or an embedding.
2. A surjective homomorphism is called an epimorphism.
3. A bijective homomorphism is called an isomorphism.
4. A homomorphism of a ring R into itself is called an endomorphism of R.
5. An isomorphism of a ring R onto itself is called an automorphism of R.
6. A ring R is said to be isomorphic with a ring R9 and denote this by
R > R9 if there is an isomorphism of R onto R9. The following examples
should help us for a better understanding of the above concepts.
Example 9.5.1
1. Let R and R9 be any rings and define f : R → R9 by f (x) 5 0 for all x R.
Then, for any a and b in R,
f (a 1 b) 5 0 5 0 1 0 5 f (a) 1 f (b)
f (ab) 5 0 5 0 0 5 f (a) f (b)
and therefore f is a homomorphism, which is called the trivial or zero
homomorphism. Note that this is not a monomorphism unless R 5 {0}
and is not an epimorphism unless R9 5 {0}.
2. The identity mapping IR of a ring R onto itself is an automorphism of R.
3. Let R be any ring and X be any nonempty set. Consider the ring RX of
all mappings of X into R under the point-wise operations (refer Example
9.1.1 (7)). For any x X, define ax : RX → R by
and hence f (1) is the multiplicative identity in R9 so that f (1) 5 1, the unity in
R9. Next, let a be a unit in R. Then, there is an element a21 in R such that aa21
5 1 5 a21a. By applying f to these, we get that
Proof: Let S and S9 be subrings of R and R9, respectively since f (0) 5 0 and S
and S9 contain zero elements, it follows that f (S) f and f21(S9) f. Now,
a 5 qn 1 r and b 5 q9n 1 s,
a 1 b 5 qn 1 r 1 q9n 1 s 5 (q 1 q9)n 1 (r 1 s)
( q q)n ( r n s) if r s n
( q q1)n ( r n s) if r s n
If rs 5 un 1 t, 0 # t , n, then
(g o f )(a 1 b) 5 g(f (a 1 b)) 5 g(f (a) 1 f (b))5 g(f (a)) 1 g(f (b))
and (g o f )(ab) 5 g(f (a)f (b)) 5 g(f (a))g(f (b))
0 if a is even
f (a) .
1 if a is odd
and hence f is the zero homomorphism. On the other hand, if f (1) 5 1, then
for any 0 , a Z,
f (a) 5 f (1 1 1 1 … 1 1) (a times)
5 f (1) 1 f (1) 1 … 1 f (1) (a times)
5 1 1 1 1 … 1 1 (a times)
5a
EXERCISE 9(e)
For any rings R and R9, let Hom(R, R9) denote the set of all ring homomorphisms of R
into R9 and End(R) denote the set of all endomorphisms of the ring R.
3. For any ring R, prove that (End(R), 1, ο) is a ring, where 1 is the point-wise
addition and ο is the composition of mappings.
10. Let R be a ring with unity. Prove that there is a subring S of R such that S is iso-
morphic to Z or to Zm depending on whether char(R) 5 0 or m.
Proof: Note that, for any f and g End(G), f 1 g and f ο g are defined by
(f 1 g)(x 1 y) 5 f (x 1 y) 1 g(x 1 y)
5 f (x) 1 f (y) 1 g(x) 1 g(y)
5 f (x) 1 g(x) 1 f (y) 1 g(y) (since G is abelian)
5 (f 1 g)(x) 1(f 1 g)(y)
and therefore f 1 g and f ο g are endomorphisms of the group (G, 1). Thus,
1 and o are binary operations on End(G). The associativity of 1 in End(G)
follows from that of 1 in G. The zero endomorphism will be the zero element
in End(G). Also, for any f End(G), the map 2f, defined by (2f )(x) 5 2f (x)
for all x G, is the additive inverse of f in End(G). Thus, (End(G), 1) is an
abelian group. Also, clearly the composition o is associative and,
(f ο (g 1 h))(x) 5 f (g 1 h)(x)
5 f (g(x) 1 h(x))
5 f (g(x)) 1 f (h(x))
5 (f ο g 1 f ο h)(x)
f o (g 1 h) 5 f o g 1 f o h
and (f 1 g) o h 5 f o h 1 g o h
for any f, g and h End(G) Thus, (End(G), 1, ο) is a ring. Further, the iden-
tity homomorphism IG of the group G is the multiplicative identity in the ring
End(G). Thus, (End(G), 1, ο) is a ring with unity. b
In general the ring End(G) of endomorphisms of an abelian group G is not
commutative. For, consider the following example.
It can be easily verified that f and g are endomorphisms of G(5 Z 3 Z). Now
consider
Definition 9.6.2. Let (R, 1, ) be a ring and n Z1. For any matrices
and A B 5 (dij), where dij 5 ai1 b1j 1 ai2 b2j 1 … 1 ain bnj for any 1 # i,
j # n. Note that the 1 and on the right sides are those in the ring R.
A105A501A
and A 1 (2A) 5 0 5 (2A) 1 A,
nn
tij ∑ ∑ air brk ckj
k1 r1
n n
∑∑ ( air brk )ckj
k1 r1
n n
∑∑ air (brk ckj )
k1 r1
n n
∑ air ∑ brk ckj
k1
r1
n
A ( B C ) ∑ air (brj crj )
r1
n n
∑ air brj ∑ air crj
r1 r1
5 A B 1 A C
Note:
1. If R is a ring with unity 1 and En is the n 3 n matrix defined by
1 if i j
En 5 (eij), where eij
0 if i j
then En A 5 A 5 A En for any matrix A in Mn(R). Therefore, if R is
with unity, then Mn(R) is a ring with unity.
2. If n 5 1, then a 1 3 1 matrix (a) can be identified with a itself and, hence
M1(R) is isomorphic with R.
3. If n 1, then Mn(R) may not be commutative even when R is commuta-
tive; for consider the following theorem.
Theorem 9.6.3. For any n 1, the ring Mn(R) of n 3 n matrices over the real
numbers is a noncommutative ring with unity.
Proof: Let n 1. Since the real number system R forms a ring with unity,
by Theorem 9.6.2 (1), Mn(R) is a ring with unity. Let A 5 (aij) and B 5 (bij)
be the matrices defined by
1 if i 1 j
aij
0 otherwise
1 if j 1 and i 1 or 2
and bij .
0 otherwise
Definition 9.6.3. The algebraic system (R4, 1, ), where R4 is the set of all
quadruples of real numbers and 1 and are the binary operations defined
as follows, is called the system of real quaternions and each element of this
system is called a real quaternion.
For convenience, let us write a quadruple by (a0, a1, a2, a3). For any
a 5 (a0, a1, a2, a3) and b 5 (b0, b1, b2, b3) in R4,
we define
k j
i j 5 k, j k 5 i, k i 5 j.
i i 5 j j 5 k k 5 21
i j 5 k, j k 5 i, k i 5 j (A)
j i 5 2k, k j 5 2i, i k 5 2j
(a0 1 a1i 1 a2j 1 a3k) (b0 1 b1i 1 b2j 1 b3k) 5 c0 1 c1i 1 c2j 1 c3k,
To multiply a0 1 a1i 1 a2j 1 a3k by b0 1 b1i 1 b2j 1 b3k on the right, we first
multiply each ‘term’ in the first quaternion with each term in the second on
the right, use the laws given in (A) and collect the terms with each of i, j and
k and without any of them. Recall that the elements 1, 21, i, 2i, j, 2j, k and
2k form a nonabelian group of order 8 under the above multiplication rules
(A) and is called the group of quaternions. If we write
(1, 0, 0, 0) 5 1, (0, 1, 0, 0) 5 i
(0, 0, 1, 0) 5 j and (0, 0, 0, 1) 5 k,
Theorem 9.6.4. The real quaternions form a noncommutative ring with unity
under the addition and multiplication given above. This ring is denoted by QR
and is called the ring of real quaternions.
Answer: Consider the ring Z2 of integers modulo 2. Then, Z2 5 {0, 1}. Then,
the ring M2(Z2) of 2 3 2 matrices over Z2 has exactly 2232 (516) elements and
is not commutative. For, consider
1 1 1 0 11 0 0 0
0 0 1 0 0 0 0 0
1 0 1 1 1 1
and .
1 0 0 0 1 1
Worked Exercise 9.6.2. Prove that every nonzero element in the ring of real
quaternions is a unit.
Answer: Let 0 a 5 a0 1 a1i 1 a2 j 1 a3k QR. Put s a02 a12 a22 a32 .
Since a 0, atleast one ai must be nonzero and hence s 0. Now, consider
a0 a1 a a
b i 2 j 3 k.
s s s s
s
Then, ab ba 1. Therefore, a is a unit in QR.
s
EXERCISE 9(f )
1. Evaluate the following products a b and b a in the rings mentioned against them.
(i) a, b End(R3) defined by
a( r1 , r2 , r3 ) ( r1 r2 , r1 r2 ,0) and
.
b( r1 , r2 , r3 ) ( r3 r1 , r3 r2 , r2 r3 )
1 0 0
0 2 1
(ii) a 0 2 3 and b 1 2 3 in M 3 (Z) .
2 1 0 3 1 2
(iii) a 5 2 1 3i 1 4j 1 5k and b 5 1 1 2i 1 3j 1 k in QR.
(iv) a 5 1 1 2i 1 5j 2 3k and b 5 3 1 2i 2 2j 1 k in QR.
2. Give an example of a noncommutative ring with unity having exactly 81 elements.
11. Prove in detail that QR is a ring under the addition and multiplication of real
quaternions.
12. Find the centre of the ring of Mn(R) of all n 3 n matrices over the real number
system R.
13. Extend the above Exercise 12 for an arbitrary ring R in place of the ring R of real
numbers.
15. If S is a subring of a ring R, prove that Mn(S) is a subring of Mn(R) for any posi-
tive integer n.
16. If R is a ring such that Mn(R) is a ring with unity, then prove that R is a ring with
unity.
Definition 9.7.1. A nontrivial commutative ring with unity and without zero
divisors is called an integral domain.
Recall that a ring R with unity 1 is nontrivial or nonzero (that is, R {0})
if and only if the additive identity 0 and the multiplicative identity 1 are dif-
ferent. In the following, we obtain some other simple equivalent conditions
for a ring to be an integral domain.
Theorem 9.7.1. The following are equivalent to each other for any nontrivial
commutative ring (R, 1, ) with unity.
1. (R, 1, ) is an integral domain.
2. For any elements a, b and c in R,
ab 5 ac ⇒ a 5 0 or b 5 c.
3. For any elements a and b in R,
ab 5 0 ⇒ a 5 0 or b 5 0.
a(b – c) 5 ab – ac 5 0
Example 9.7.1
1. The ring Z of integers, the ring Q of rational numbers, the ring R of real
numbers and the ring C of complex numbers are all integral domains
with respect to usual addition and multiplication. In each of these, the
product of any two nonzero elements is again nonzero and hence there
are no zero divisors.
2. The ring Z[i] of Gaussian integers is an integral domain with respect to
the addition and multiplication of complex numbers. Note that Z[i] is a
subring of the ring C of complex numbers.
3. (Zp, 1p, ·p) is an integral domain for any prime number p.
Example 9.7.2
1. The rings Q, R and C are all fields, since for any nonzero number a,
there is 1/a for which a 1a 1 and hence a is a unit.
2. The ring Z of integers is not a field, since 2 is not a unit in Z. In fact, 1
and 21 are the only units in Z.
3. For any prime number p, (Zp, 1p, ·p) is a field, since any 0 , a , p is
relatively prime with p and hence a unit in Zp.
4. The ring QR of real quaternions is not a field, even though every nonzero
element is a unit in it; because it is not commutative.
Theorem 9.7.2. Every field is an integral domain and the converse is not true.
ab 5 0 ⇒ a21(ab) 5 0 ⇒ (a21a)b 5 0 ⇒ b 5 0.
where a1, a2, …, an are all the distinct nonzero elements in R. Since R is an
integral domain, a 0 and each ai 0, we get that aai 0 for each 1 i n.
Therefore, the set
is a subset of R 2 {0}. Further, aai aaj for all i j (since a 0 and ai aj).
S is an n-element subset of R 2 {0}, which also has n elements. Therefore,
Corollary 9.7.1. The following are equivalent to each other for any positive
integer n.
1. n is a prime number.
2. (Zn, 1n, ·n) is an integral domain.
3. (Zn, 1n, ·n) is a field.
Proof: First note that, for each of these, n must be necessarily greater than 1
(for, if n 5 1, Zn is trivial).
(1) ⇒ (2) follows from the fact that, for any a and b Zn,
ab 5 0 in Zp ⇒ n divides ab
n divides ab ⇔ n divides a or b.
Corollary 9.7.2. For any prime number p, (Zp, 1p, ·p) is a field.
Definition 9.7.3. A nontrivial ring with unity is called a division ring if every
nonzero element is a unit.
Note that any field is a division ring and the converse is not true, since
the ring QR of real quaternions is a division ring, but not a field (see Worked
Exercise 9.6.2). However, any commutative division ring is a field.
Worked Exercise 9.7.1. Prove that any integral domain has exactly two idem-
potents.
Answer: Let R be an integral domain. Then, 0 1 and clearly these two are
idempotents. If a is any idempotent in R, then a2 5 a and hence
a(a 2 1) 5 0 so that a 5 0 or a 5 1
Answer: We have
Here, the elements are added and multiplied as in the complex number system,
except that the coefficients are reduced modulo 3. In particular, note that
21 5 2, 2i 5 2i, 2 ·3 2 5 1, 2i ·3 2i 5 2
Q001-Algebra-111001_CH 09.indd 48
2 2 0 1 21i i 11i 2 1 2i 2i 1 1 2i
i i 11i 21i 2i 1 1 2i 2 1 2i 0 1 2
11i 11i 21i i 1 1 2i 2 1 2i 2i 1 2 0
21i 21i i 11i 2 1 2i 2i 1 1 2i 2 0 1
2i 2i 1 1 2i 2 1 2i 0 1 2 i 11i 21i
1 1 2i 1 1 2i 2 1 2i 2i 1 2 0 11i 21i i
9-48 Algebra – Abstract and Modern
2 1 2i 2 1 2i 2i 1 1 2i 2 0 1 21i i 11i
3 0 1 2 i 11i 21i 2i 1 1 2i 2 1 2i
0 0 0 0 0 0 0 0 0 0
1 0 1 2 i 11i 21i 2i 1 1 2i 2 1 2i
2 0 2 1 2i 2 1 2i 1 1 2i i 21i 11i
i 0 i 2i 2 21i 2 1 2i 1 11i 1 1 2i
11i 0 11i 2 1 2i 21i 2i 1 1 1 2i 2 i
21i 0 21i 1 1 2i 2 1 2i 1 i 11i 2i 2
2i 0 2i i 1 1 1 2i 11i 2 2 1 2i 21i
1 1 2i 0 1 1 2i 21i 11i 2 2i 2 1 2i i 1
2 1 2i 0 2 1 2i 11i 1 1 2i i 2 21i 1 2i
9/21/2011 4:52:41 PM
Rings 9-49
Answer: We have
12 0 1 i 11i
0 0 1 i 11i
1 1 0 11i i
i i 11i 0 1
11i 11i i 1 0
·2 0 1 i 11i
0 0 0 0 0
1 0 1 i 11i
i 0 i 1 11i
11i 0 11i 11i 0
It can be easily verified that (Z2[i], 12, 2) is a commutative ring with unity.
Since
(1 1 i)(1 1 i) 5 1 1 2i 1 (21) 5 0,
Worked Exercise 9.7.4. Let R be a nontrivial ring such that, for each 0 a
R, there exists unique element x in R such that axa 5 a. Prove that R is a
division ring.
Answer: We first prove that R has no zero divisors. Suppose that a and b R
such that ab 5 0 and a 0. Choose x R such that axa 5 a. Then,
EXERCISE 9(g)
3. Prove that Q[i] 5 {a 1 bi : a and b are rationals} is a field under the addition
and multiplication of complex numbers.
4. Let R be a field. Prove that R is a Boolean ring if and only if R has exactly two
elements.
a
R : a and b Z and p does not divide b.
b
Prove that R is an integral domain under the usual addition and multiplication of
rational numbers.
13. Let R be a nontrivial finite ring without zero divisors. Then prove that R is with
unity and that (R 2 {0}, ?) is a group.
14. Prove that any finite commutative ring without zero divisors is a field.
15. For any simple abelian group (G, 1), prove that the ring End(G) of all endomor-
phisms of (G, 1) is a division ring.
18. Let R be a nontrivial finite ring without zero divisors. Then prove that R is a divi-
sion ring.
19. Let R and S be two rings. Then prove that the product ring R 3 S is an integral
domain if and only if one of R and S is an integral domain and the other is the
trivial ring.
In the study of finite groups, we have proved several results using the concept
of a normal subgroup, quotient construction and induction on the group order.
Homomorphic images of groups are identified with quotient groups with the
help of the kernel of the homomorphism which is a normal subgroup. The
role of normal subgroups in groups is played by ideals in rings. The concepts
of ideal and quotient rings are important in the structure theory of rings. A
special kind of subrings, which are most suitable (ideal) for the study of the
structure of rings, are popularly called ideals.
10.1 IDEALS
In this section, we introduce the notion of an ideal in a ring and discuss sev-
eral important elementary properties of ideals.
Clearly any left ideal or right ideal of a ring R is a subring of R. But a sub-
ring of R may be neither a left ideal nor a right ideal. For example, the set Z
of integers is a subring of the ring R of real numbers and Z is not an ideal of
R, since 1 1 Z. If R is a commutative ring, there is no difference between
2
a left ideal, a right ideal and an ideal. Sometimes, we refer to an ideal as a
two-sided ideal.
Example 10.1.1
1. For any ring R, clearly {0} and R are ideals of R and are called trivial
ideals. {0} is called the zero ideal. Ideals other than {0} and R are called
proper ideals.
2. If (R, 1, ?) is a ring with trivial multiplication, that is, ab 5 0 for any a
and b in R, then every subgroup of (R, 1) is an ideal of R.
3. Consider the ring M2(R) of 2 3 2 matrices over the real R and let
0 a
I : a and b are real numbers .
0 b
r s
r s 0 a 0 ra sb I for all M 2 ( R).
0 b 0 ta ub
t u t u
a b
4. Let J : a and b are real numbers. Then, J is a right ideal of
0 0
M2(R) and is not a left ideal.
5. For any nonnegative integer n, let
n {na : a is an integer}.
a b
6. Let K : a, b, c and d are even integers . Then, K is an ideal
c d
of M2(Z). Later, we shall prove that M2(R) has no nontrivial ideals, since
R is so. However, Z has several ideals and so is M2(Z).
7. Let X be any set and consider the ring (P(X), 1, ∩) of all subsets of X,
where 1 and ∩ are defined by
A + B = ( A − B) ∪ ( B − A) and A ∩ B = The intersection of A and B.
IY { f R X : f ( y ) 0 for all y Y }.
Theorem 10.1.1. Let R be a ring with unity and U(R) be the set of all units
in R. Then, the following are equivalent to each other for any left (right or
two-sided) ideal I of R.
1. I5R
2. U(R) ⊆ I
3. I ∩ U(R)
4. 1I
Proof: Let I be a left ideal of R. (1) ⇒ (2) and (2) ⇒ (3) are trivial.
(3) ⇒ (4): Suppose that a I ∩ U(R). Then, a has multiplicative inverse a21
in R and 1 5 a21 ? a I (since a I and I is a left ideal of R).
(4) ⇒ (1): If 1 I, then, for any r R, r 5 r ? 1 I and therefore R ⊆ I so
that I 5 R.
In the following, our discussion is restricted to ideals of rings. Some of
the results proved for ideals can be extended to left or right ideals easily.
However, we are more interested in two-sided ideals, since these lead to the
construction of quotient rings. First, we discuss certain standard methods of
constructing new ideals from given ones.
Proof: First note that every ideal contains the zero element of the ring; for, an
ideal I is nonempty and hence there exists a I so that 0 5 0a I. Therefore,
0 Ia for all a . Put I ∩ I a Then, 0 I and hence I is a nonempty
a
subset of R. Since the intersection of any family of subgroups of (R, 1) is
again a subgroup, it follows that I is a subgroup of (R, 1). Also,
Theorem 10.1.3. Let {Ia}a be a class of ideals of a ring. Suppose that, for
any a and b , there exists such that Ia ⊆ I and Ib ⊆ I and Ib ⊆ I
(such classes are called directed above). Then, ∪ I a is an ideal of R.
a
Corollary 10.1.1. Let {Ia}a be a chain of ideals of a ring R (that is, given
any two members in the class, one of them is contained in the other). Then,
∪ I a is again an ideal of R.
a
Proof: Let A be the set of all elements of the form given in the theorem. We
shall prove that A is the smallest ideal of R containing a. By taking r 5 0 5 s;
m 5 0 and n 5 1, we have a A. If I is any ideal of R containing a, then ra,
as, xay and na I for any r, s, x, y R and n Z and hence A ⊆ I. Thus, we
are left with only verifying that A is an ideal of R. Using the commutativity of
1 and the distributivity of the multiplication over the addition, one can easily
prove that A is an ideal of R. Thus, A 5 <a>.
In certain special cases, <a> turns out to be much simpler.
m
< a > ∑ xi ayi : 0 m Z, xi and yi R.
i1
Proof: Let B be the set given on the right hand side. Then, by taking m 5 1,
x1 5 1 5 y1 (the unity in R), we get that a B. Since B is closed under 1, it
follows that na B for all n Z. Also, by taking m 5 1 and x1 5 1, we get
m
m
ra 1 as 1 na 1 ∑ x ay = r s ∑ xi ayi a 1 na.
i=1
i i
i1
I 1 J 5 {a 1 b : a I and b J}.
r(a 1 b) 5 ra 1 rb I 1 J
and (a 1 b)r 5 ar 1 br I 1 J
∑I
i1
i I1 I 2 I n {a1 a2 an : ai I i }.
n n
Then, I i is the smallest ideal containing ∪ I i .
i1 i1
∑I
a
a ∪I
a
a .
m
<S > ∑ ra : m 0, ri R and ai S .
i1 i i
For any ideals I1, I2, …, In of a ring R, we have proved that any element of
n
∪ I i can be expressed as a sum a1 1 a2 1 1 an with ai Ii, 1 # i # n.
i1
However, there is no guarantee that this expression is unique, unless the ide-
als I1, I2, …, In satisfy certain additional conditions.
I i ∩∑ I j {0} for all 1 i n.
ji
( )
By the uniqueness, a 5 0. Thus, I i ∩ I j {0}. Conversely, suppose that
j ≠i
the given condition is satisfied. Since I 5 I1 1 I2 1 1 In, any element of I
can be expressed as a1 1 a2 1 1 an, with ai Ii. Now, suppose that
a1 1 a2 1 1 an 5 b1 1 b2 1 1 bn,
ai bi ∑ (b j a j ) I i ∩∑ I j {0}
j ≠i j≠i
Corollary 10.1.9. Let I and J be ideals of a ring R. Then, any element of R can
be uniquely expressed as a 1 b with a I and b J if and only if
I J R and I ∩ J {0}.
In a ring with unity, we can have yet another beautiful description of direct
summands. Before going to this, let us define the following definition.
Proof: First note that, for any central idempotent e in R, the principal ideal
generated by e is of the form
(1 2 e)25 1 2 e 2 e 1 e2 = 1 2 e 2 e 1 e = 1 2 e
and (1 2 e)x 5 x 2 ex 5 x 2 xe 5 x(1 2 e)
x 5 ex 1 (1 2 e)x I 1 J
From 1 5 e 1 s, we have
e 5 e(e 1 s) 5 e2 1 es 5 e2 (since es 5 0)
ex 1 sx 5 (e 1 s)x 5 x 5 x(e 1 s) 5 xe 1 xs
Thus, I 5 eR 5 <e>.
The central idempotents in any ring have certain nice properties. They form
a Boolean ring under suitable operations, defined in the following theorem.
Theorem 10.1.8. Let (R, 1, ?) be a ring and B(R) be the set of all central
idempotents in R. For any a and b B(R), define
Proof: First observe that, for any a and b B(R), a ∗ b and a b are central
idempotents of R; i.e.,
It is a straight forward verification to prove that all the axioms of rings are
satisfied in B(R) with ∗ as addition and ? as multiplication. Note that a a 5
a and a ∗ a 5 0 for all a B(R) and 0 is the zero element in B(R) also. If R
has unity 1, then 1 B(R) and 1 is the unity in B(R) also.
In addition to the two binary operations ∩ and 1 on the set of ideals of a
ring, we introduce yet another binary operation, which is denoted by juxtapo-
sition, in the following definition.
n
IJ ∑ ai bi : n Z+ , ai I and bi J .
i1
n n
rx ∑ ( rai )bi and xr ∑ ai (bi r )
i1 i1
and rai, ai I and bi, bir J and hence rx and xr belong to IJ. Thus, IJ is an
ideal of R. Clearly IJ ⊆ I and J.
There is an important observation that an ideal of an ideal need not be an
ideal; that is, if I is an ideal of a ring R, then I can be treated as a ring on its
own (since I is a subring of R) and, if J is an ideal of I, then J need not be an
ideal of the ring R. This is illustrated in the following example.
Example 10.1.2. Let R be the ring of real numbers under the usual addition
and multiplication. Consider the ring RR of all mapping of R into R under the
point-wise addition and multiplication. Let
S 5 {f RR : f is continuous},
c onstant map of R which maps every element of R onto the real number 1 .
2
The assumption 12 i2 J leads to a contradiction; for, let 12 i2 J. Then,
1 2
i 5 i2f 1 ni2 for some f S with f (0) 5 0 and n Z.
2
Therefore, fi2 5 ( 12 2 n)i2; that is, f (x)x2 5 ( 12 2 n)x2 for all x R. If x 0,
f (x) 5 12 2 n. Therefore, f is a nonzero constant function on R 2 {0} and
f (0) 5 0. This is a contradiction to the fact that f is continuous.
Next we discuss a characterization theorem for ideals of a matrix ring
Mn(R), where R is an arbitrary ring with unity.
Proof: It can be easily verified that Mn(J) is an ideal of Mn(R) for any ideal J
of R. Conversely suppose that I is an ideal of Mn(R).
For any 1 # i, j # n, let Eij be the n 3 n matrix over R such that the ijth
entry is 1 and all other entries are 0. Then, any n 3 n matrix (aij) can be
expressed as
n
( aij ) ∑ aij Eij
i , j1
1 if j k
and EijEkr 5 jkEir, where jk .
0 if j k
Now, consider the given ideal I of Mn(R) and define
That is, J is the set of all 11 entries (entries in the first column and first row) of
the matrices belonging to I. Since I , J is a nonempty subset of R. Clearly
a 2 b J for any a and b J. Next, suppose that a J and r R. Then, a
5 a11 for some (aij) I. Since I is an ideal of Mn(R), we have
n
ra E11 r E11 ∑ aij Eij E11 rE11 ( aij ) E11 I
i , j1
n
and ar E11 E11 ∑ aij Eij rE11 E11 ( aij )rE11 I
i , j1
and hence ra and ar J (since ra is the 11th entry of raE11 and ar is the 11th
entry of arE11). Thus, J is an ideal of R. We shall prove that I 5 Mn(J).
n
Let A 5 (aij) I. Then, A aij Eij . For any 1 # i, j # n, consider
i , j1
n n
E1i AE j1 E1i ∑ ars Ers E j1 ∑ ars ir E1s E j1
r ,s1 r ,s1
n
∑ ars ir sj E11 aij E11.
r ,s1
n n
Ei1 BE1 j Ei1 ∑ brs Ers E1 j ∑ brs Ei1 Ers E1 j
r ,s1 r , s1
n
∑ brs 1r Eis E1 j
r , s1
n
∑ brs 1r s1 Eij b11 Eij aij Eij .
r , s1
The above theorem is false for rings without unity. This is illustrated in the
following example.
Example 10.1.3. Consider the ring 2Z of even integers. Note that 2Z has no
unity. Consider the ring M2(2Z) of 2 3 2 matrices over 2Z. Let
It can be easily checked that I is an ideal of M2(2Z). Note that I M2(J) for
any ideal J of 2Z.
Theorem 10.1.10 is useful only to the extent that we can describe the ideals
of a ring R, which is not usually easy to do, although it is easy for the ring Z
of integers (recall that any ideal of Z, being a subgroup of (Z, 1), is generated
by a nonnegative integer). However, the ideals of fields are easy to describe.
Proof: Clearly {0} and R are ideals of R and these are distinct, since R is
nontrivial. If R is a field and I {0} is an ideal of R, then there exists 0
a I and hence 1 5 a21a I, so that I 5 R. Conversely suppose that {0}
and R are the only ideals of R. Let 0 a R and I 5 aR 5 <a>. Then, I is a
nonzero ideal of R and hence I 5 R. In particular, 1 R 5 I 5 aR and hence
1 5 ab for some b R. Therefore, a is a unit. Thus, R is a field. b
Corollary 10.1.10. The ring R of real numbers (or the ring C of complex
numbers) has only two ideals, namely {0} and the whole ring.
Corollary 10.1.11. For any positive integer n, the ring Mn(R) of n 3 n matri-
ces over R has exactly two ideals namely {0} and the whole ring Mn(R).
Since Mn(R) is a noncommutative ring for n > 1, Mn(R) is not a field (and
not a division ring) even though it has only two ideals. This says that the
commutativity of the ring R in Theorem 10.1.11 cannot be dropped from the
hypothesis. Nontrivial rings having only two ideals are called simple rings.
Mn(R) is a simple ring for all n Z1.
Worked Exercise 10.1.1. Let I and J be ideals of a ring R with unity. Then
prove that R 5 I ! J if and only if there are central idempotents a and b in R
such that a 1 b 5 1, ab 5 0, I 5 aR and J 5 bR.
ax 1 bx 5 (a 1 b)x 5 x 5 x(a 1 b) 5 xa 1 xb
and hence ax 2 xa 5 xb 2 bx I ∩ J 5 {0}
x 5 (a 1 b)x 5 ax 1 bx 5 ax 1 0 5 ax aR.
x 5 (a1b)x 5 ax 1 bx I 1 J
I 1 (J ∩ K) 5 (I 1 J) ∩ K.
b 5 x 2 a K (since x K and a I ⊆ K)
Worked Exercise 10.1.3. Consider the ring Z of integers. For any positive
integers n and m, let I 5 nZ and J 5 mZ. Then, compute I 1 J, I ∩ J and IJ.
Answer: Let (m, n) and [m, n] be the greatest common divisor and least com-
mon multiple of m and n, respectively. Note that a nZ if and only if a is a
multiple of n (or n divides a). Therefore,
a I 1 J ⇔ a nZ 1 mZ
⇔ a 5 nx 1 my for some x, y Z
⇔ (m, n) divides a
⇔ a (m, n)Z
a I ∩ J ⇔ a nZ ∩ mZ
⇔ mn divides a
⇔ a mnZ.
Now, we have
I1 {a R : ( a, b) I for some b R}
Then, I1 and I2 are ideals of R, since I1 5 p1(I) and I2 5 p2(I), where p1 and
p2 : R 3 R → R are the first and second projections, respectively and since
p1 and p2 are epimorphisms of rings. Also,
(a, b) I ⇒ a I1 and b I2
⇒ (a, b) I1 3 I2.
(a, b) I1 3 I2 ⇒ a I1 and b I2
⇒ (a, c) I1 and (d, b) I2 for some c, d R.
⇒ (a, 0) 5 (1, 0)(a, c) I and (0, b) 5 (0, 1)(d, b) I
⇒ (a, b) 5 (a, 0) 1 (0, b) I
EXERCISE 10(a)
1. Determine all the ideals in each of the following rings under the operations.
(i) The ring Z of integers.
(ii) The ring Q of rational numbers.
4. Let R be a ring with unity. If R is a division ring, prove that R has only two ideals.
Is the converse true?
10. For any ideals I and J of a commutative ring R, prove the following:
(i) r(I ∩ J) 5 r(I) ∩ r(J)
(ii) I ⊆ J ⇒ r(I) ⊆ r(J)
(iii) r(I) 1 r(J) ⊆ r(I 1 J)
(iv) r(I 1 J) 5 r(r(I) 1 r(J))
11. For any ideas I and J of a ring R, let
(I : J) 5 {x R : xa J for all a I}.
Prove that (I : J) is a left ideal of R.
12. Express each of the following in the form of nZ for a suitable n.
(i) 8Z ∩ 12Z
(ii) 6Z 1 9Z
(iii) (6Z : 9Z)
(iv) r(12Z)
(v) (12Z)*
(vi) (9Z : 6Z)
13. For any positive integers m and n, express each of the following in the form of
dZ for a suitable integer d:
(i) mZ ∩ nZ
(ii) mZ 1 nZ
(iii) r(mZ)
(iv) (nZ : mZ)
(v) (mZ)*
14. Consider the ring Z[ 3] {a b 3 : a and b Z}, under the usual addition
and multiplication of real numbers. Prove that the set
I {a b 3 : a, b Z and a b is even}
is an ideal of Z[ 3].
15. Let R be a ring without zero divisions. If every subring of R is an ideal of R, then
prove that R is commutative.
17. Define the notion of the right annihilator Annr(S) of S and formulate and prove a
statement similar to the Exercise 16 for right annihilators.
18. Prove that every ideal of Zn is a principal ideal.
19. Let R be commutative ring in which {0} and R are the only ideals. Then prove that
R is field or R is a finite ring with |R| as a prime and ab 5 0 for all a and b R.
21. In the ring Z[i] of Gaussian integers, prove that the set
I {a bi : a and b are even}
is an ideal and find the annihilator I*.
22. If R is a simple ring with unity, then prove that the ring Mn(R) of all n 3 n matri-
ces is a simple ring.
(a 1 I) 1 (b 1 I) 5 (a 1 b) 1 I.
R/I 5 {a 1 I : a R}.
(a 1 I) 1 (b 1 I) 5 (a 1 b) 1 I
and (a 1 I) (b 1 I) 5 ab 1 I.
abelian group. Next, with regard to the multiplication in R/I, we should first
prove that the operation on R/I is well defined.
To do this, for any a, b, a9 and b9 R, we have
a I a I
⇒ a a I and b b I
b I b I
⇒ ( a a) b I and a(b b) I
⇒ ab ab ( a a)b a(b b) I
⇒ ab I ab I .
Therefore, the multiplication on R/I depends on the cosets, but not on their
representatives. For any a, b and c R, we have
Therefore, ? distributes over 1 from left and, similarly from right also. Thus,
(R/I, 1, ?) is a ring.
Definition 10.2.1. For any ideal I of a ring R, the ring (R/I, 1, ?) constructed
above is called quotient ring of R by I or factor ring of R by I.
Note 10.2.1
1. The zero element in the quotient ring R/I is 0 1 I 5 I, where 0 is the zero
element in R.
2. If the ring R has unity 1, then R/I also has unity, namely 1 1 I. The con-
verse may not hold good. That is, R/I may have unity while R has no unity.
For example, R is an ideal of R and the quotient R/R is the trivial ring
which obviously has unity (when a ring has only one element, then that
element is the additive identity as well as the multiplicative identity).
3. If R is a commutative ring, then the quotient R/I is also commutative ring
for any ideal I or R.
Theorem 10.2.3. Let I be an ideal of a ring R. Then, there exists a ring S and
a homomorphism f : R → S such that I 5 ker f.
Proof: Consider the quotient ring R/I and define f : R → R/I by f (a) 5 a 1 I
for any a R. Then, f is a homomorphism of rings; for,
f (a 1 b) 5 (a 1 b) 1 I 5 (a 1 I) 1 (b 1 I) 5 f (a) 1 f (b)
and f (ab) 5 ab 1 I 5 (a 1 I)(b 1 I) 5 f (a)f (b)
R / ker f f ( R).
Proof: The proof is same as that of Theorem 5.2.1, except that the map g :
R/K → f (R) defined by
g(a 1 k) 5 f (a)
I J
I .
I ∩J J
R S
.
I f (I )
R S
1
.
f (J ) J
Theorem 10.2.8. Let I and J be ideals of a ring and I ⊆ J. Then, J/I is an ideal
of R/I and
(R/I)/(J/I) R/J.
J J/I
f : Z → Zn by f (a) 5 r,
Z/nZ Zn.
F F/ker f R.
Worked Exercise 10.2.2. For any ideal I and J of a ring R, prove that R/I ∩ J
is isomorphic to a subring of R/I 3 R/J.
It can be easily verified that f is a homomorphism of the ring R into the prod-
uct ring R/I 3 R/J. By the fundamental theorem of homomorphisms, f (R) is
a subring of R/I 3 R/J and R/ker f f (R). Since I 5 (0 1 I) and J 5 (0 1 J)
are the zero elements of R/I and R/J, respectively, (I, J) is the zero element in
the ring R/I 3 R/J. Therefore,
Since 3Z/12Z is an ideal of Z/12Z Z12 (by Example 10.2.1), there must be
an ideal J of Z12 such that
Worked Exercise 10.2.4. Let I and J be ideals of array R such that I ∩ J 5 {0}
and I 1 J 5 R. Then prove that
J R/I, I R/J.
Answer: We are given that char(R) 5 n and hence n is the least positive
integer such that na 5 0 for all a R. Now, for any a 1 I R/I, a R,
we have
Worked Exercise 10.2.6. Let P(X) be the power set of any set X. Let Y be a non-
empty proper subset of X. Prove that P(Y) is an ideal of the ring (P(X), 1, ∩)
and describe the quotient ring P(X)/ P(Y).
Therefore, P(Y) is an ideal of the ring (P(X), 1, ∩). Any element of the quo-
tient ring P(X)/P(Y) is of the form Z 1 P(Y) for some Z P(X); that is,
Z ⊆ X. Now, we can write
Z 5 (Z ∩ Y) ∪ (Z ∩ (X 2 Y))
5 (Z ∩ Y) 1 (Z ∩ (X 2 Y))
and hence Z 1 P(Y) 5 ((Z ∩ Y) 1 P(Y)) 1 ((Z ∩ (X 2 Y)) 1 P(Y))
5 (Z ∩ (X 2 Y)) 1 P(Y), since Z ∩ Y P(Y).
EXERCISE 10(b)
2. State whether each of the following is true or false. Substantiate your answers.
(i) For any ring R, R has unity if and only if R/I has unity for all ideals I of R.
(ii) For any ideal I of a commutative ring, R/I is commutative.
(iii) A ring R is commutative if and only if R/I is commutative for all ideals
I of R.
(iv) For any integral domain R, R/I is an integral domain for any ideal I of R.
(v) A ring R is an integral domain if and only if R/I is an integral domain for
any ideal I of R.
(vi) Any quotient ring of a field is a field.
(vii) For any ideal I of a ring R, R/I is a field if and only if R is a field.
(viii) A nontrivial ring R is a field if R/I is a field for all proper ideal, I of R.
3. Prove that a homomorphism f : R → S of rings is an injection if and only if
ker f 5 {0}.
4. For any subring S of a ring R, prove that the multiplication of additive cosets of
S in R is well defined by the equation
(a 1 S)(b 1 s) 5 ab 1 S
if and only if S is an ideal of R.
5. For any ideals I and J of a ring R, prove that the set
J {a I : a J }
is an ideal of R/I.
6. Determine all the idempotents, nilpotents and units in each of the following
quotient rings
(i) Z/6Z
(ii) Z/8Z
(iii) Z/7Z
7. Let I be an ideal of the ring Z such that Z/I is an integral domain. Then prove that
I 5 {0} or I 5 pZ for some prime number p.
8. Prove that the following are equivalent to each other for any positive integer n.
(i) n is a prime number.
(ii) Z/nZ is a field.
(iii) Z/nZ is an integral domain.
9. Let R be a commutative ring and N be the set of all nilpotents in R. Then prove
that N is an ideal of R and the quotient ring R/N has no nonzero nilpotents.
10. Prove that Zn’s, {0} and Z are the only homomorphic images of the ring Z of
integers.
11. Let S be a subring and I be an ideal of a ring R such that S ∩ I 5 {0}. Prove that
S is isomorphic to a subring of the quotient ring R/I.
12. For any two subsets A and B of a ring R, let the product of A and B be defined by
the set
AB 5 {ab : a A and b B}.
Give an example of an ideal I of a ring R such that the product (x 1 I) (y 1 I) of
two cosets (x 1 I) and (y 1 I) is properly contained in the coset (xy 1 I).
14. For any pair of relatively prime positive integers m and n, prove that Zm/nmZ
Z n.
15. Let R1, …, Rn be rings and I1, …, In be ideals of R1, …, Rn, respectively. Then
prove that
R1 3 … 3 Rn/I1 3 … 3 In R1/I1 3 … 3 Rn/In.
16. For any positive integer n, determine all the ideals of the quotient ring Z/nZ and
all the homomorphic images of Z/nZ.
a I 5 nZ ⇔ n divides a.
The classical version of the Chinese Remainder Theorem is that ‘given dis-
tinct primes p1, p2, …, pn and integers a1, a2, …, an, one can always find an
integer a such that
a 1 Ii 5 ai 1 Ii for all 1 i n.
Theorem 10.3.1. Let R be a ring with identity and I1, I2, …, In be ideals of
R. If Ii 1 Ij 5 R for any i j, then, for any x1, x2, …, xn R there exists x
R such that
Proof: Recall that an ideal I of R is the whole of R if and only if the unity 1
belongs to I. Suppose that
Ii 1 Ij 5 R for all i j.
I i ∩ I j R.
ji
1 R 5 Ii 1 Ij and hence 1 5 aj 1 bj
1 ∏ ( a j b j ) si ti ,
ji
Therefore,
n
x xi ∑ x j t j xi (si ti ) (by (*))
ji
∑ x j t j − xi si I i
ji
Theorem 10.3.2 Let I1, I2, …, In be ideals in a ring R such that, for any ele-
ments x1, x2, ..., xn in R, there exists x R with x 2 xi Ii for all 1 i j.
Then, Ii 1 Ij 5 R for all i j.
x 2 a 5 x 2 xi Ii and x 5 x 2 xj Ij
Theorem 10.3.3. Let I1, I2, …, In be ideals of a ring with unity and R/I1,
R/I2, …, R/In be the corresponding quotient rings. Define f : R → R/I1 3 R/I2
3 … 3 R/In by
Also, f is a surjection if and only if, for any x1, x2, …, xn in R, there exists
x R such that
x 1 Ii 5 xi 1 Ii for all 1 i n
R R R R
f : n
→ is an isomorphism.
I1 I 2 In
∩I
i1
i
Corollary 10.3.1. Let R be a ring with unity and I1, I2, …, In be ideals of R
n
such that Ii 1 Ij 5 R for all i j and ∩ I i {0}. Then, R R/I1 3 R/I2 3
i1
… 3 R/I .
n
Corollary 10.3.2. Let R be a ring with unity and R be the direct sum of ideals
I and J. Then,
R R/I 3 R/J.
and hence mi divides a 2 ai, so that a ai(mod mi). Also, if b is any other
integer with this property, then
where p1, p2, …, pn are distinct primes and r1, r2, …, rn are positive integers.
Then,
Z m Z p r1 Z p r2 Z p rn .
1 2 n
Z12 Z 22 Z31 Z 4 Z3 .
EXERCISE 10(c)
(iv) P(X) P(Y) 3 P(X 2 Y) for any subset Y of a set X, where P(X) is the
ring. (P(X), 1, ∩)
(v) Z Z3 3 Z2
(vi) Z3ZZ
3. Let I and J be ideals of a ring R with unity and S be the ring given in Exercise 2.
Let
a b
K : a I , c J , b R.
0 c
∏ R { f
j J
j : J → ∪ R j : f ( j ) R j for all j J }.
j J
Then prove that R j is a ring under the point-wise addition and multiplication.
j J
IJ ⊆ P ⇒ I ⊆ P or J ⊆ P.
Example 10.4.1
1. {0} is a prime ideal of the ring Z of integers.
2. A nonzero ideal I of Z is prime if and only if I 5 pZ for some prime
number p.
3. Consider the ring Z 3 Z under co-ordinate wise addition and multiplica-
tion. Then, Z 3 {0} and {0} 3 Z are prime ideals of Z 3 Z.
4. {0} is not a prime ideal in the ring M2(R) of 2 3 2 matrices over the real
number system.
Theorem 10.4.1. Let P be a proper ideal of a ring R such that, for any a and
b in R,
ab P ⇒ a P or b P.
Proof: Let I and J be ideals of R such that I P and J P. Then, there exist ele-
ments a and b such that a I, a P, b J and b P. Then, by the hypothesis,
ab P. Since ab [ IJ, it follows that IJ P. Thus, P is a prime ideal of R.
Conversely suppose that R is a commutative ring and P is a prime ideal of R. Let
a and b [ R such that a P and b P. Consider the ideals <a> and <b>. We
have <a> P and <b> P and hence <a><b> P. Therefore, there exists an
element x <a><b> such that x P. x is a finite sum of elements of the form
ab [ P ⇔ a [ P or b [ P.
ab [ P ⇒ (a 1 P)(b 1 P) 5 ab 1 P 5 P
⇒ (a 1 P)(b 1 P) 5 The zero in R/P
⇒ a 1 P 5 P or b 1 P 5 P
⇒ a [ P or b [ P.
J
5 {a 1 I : a [ J}
I
is an ideal of the quotient ring R/I and that J J/I is a one-to-one cor-
respondence between the ideals of R containing I and the ideals of R/I (see
Theorem 10.2.9). This correspondence can be carried to prime ideals also,
as proved in the following theorem.
Proof: Suppose that P is a prime ideal of R. Then, P/I is a proper ideal of R/I,
since P is proper in R. Let A and B be ideals of R/I such that AB ⊆ P/I. Then,
A 5 J/I and B 5 K/I for some ideals J and K of R containing I. Also,
Conversely suppose that P/I is a prime ideal of R/I. Then, P/I is proper in R/I
and hence P is a proper ideal of R. Let J and K be ideals of R such that JK ⊆
P, Then, J/I and K/I are ideals of R/I and
Since P/I is a prime ideal of R/I, it follows that J/I ⊆ P/I or K/I ⊆ P/I and
hence J ⊆ P or K ⊆ P. Thus, P is a prime ideal of R.
Using the commutativity of the ring R, one can easily prove that I is an
ideal of R containing I. In fact, we prove in the following that I is the inter-
section of all prime ideals of R containing I. Before going to the proof of this,
let us recall an axiom of the theory of sets, which is popularly known as the
Zorns lemma. Though it is called a lemma, it is actually an axiom. There are
several equivalent formulations of this axiom. We present a convenient form
of this in the following lemma.
Zorn’s Lemma 10.4.1. Let X be any set and be a nonempty class of subsets
of X. Suppose that, for any subclass # of in which any two members of
# are comparable (that is, for any A and B [ #, either A ⊆ B or B ⊆ A), the
union of all the members of # is a member of . Then, has a maximal
member; that is, there exists M [ such that M is not properly contained in
any member of .
Proof: Let I be the nil radical of I and J be the intersection of all prime
ideals of R containing I. We shall prove that I J . If P is any prime ideal
of R containing I, then
x I ⇒ x n I # P for some n Z
⇒ x n P , n Z
⇒ x P (since P is prime)
S 5 {xn: n [ Z1}.
Example 10.4.3. In the ring Z of integers, let n p1a1 p2a2 prar , where p1, p2,
…, pr are distinct primes and a1, a2, …, ar are positive integers. Then, p1Z,
p2Z, …, prZ are all the prime ideals containing nZ and hence
nZ ( p1Z) ∩ ( p2 Z) ∩ ∩ ( pr Z) ( p1 p2 pr )Z.
24Z 6Z (since 24 23 31 )
Corollary 10.4.3. The nil radical of {0} in any commutative ring R is pre-
cisely the set of nilpotents in R; that is,
R I
N .
I I
Worked Exercise 10.4.1. Prove that the following are equivalent to each other
for any ideal I of a commutative ring R.
1. I 5 I
2. I 5 J for some ideal J of R.
3. I is the intersection of a class of prime ideals of R.
x I ⇒ x n I ∩ pa for some n Z
a
⇒ x n pa for all a
⇒ x pa for all a (since pa is prime)
⇒ x ∩ pa I
a
a I ⇒ a n I for some n Z
⇒ ( a I ) n a n I I , n Z
R
⇒ a I is a nilpotent in
I
⇒ a I I
⇒ a I.
Worked Exercise 10.4.3. Let P and Q be prime ideals of a ring R. Prove that
P ∩ Q is a prime ideal if and only if P ⊆ Q or Q ⊆ P.
b [ Q ⇒ ab [ P ∩ Q
⇒ a [ P ∩ Q or b [ P ∩ Q
⇒ b [ P ∩ Q (since a Q)
⇒b[P
and hence Q ⊆ P.
Worked Exercise 10.4.4. Let R and S be commutative rings with unities and
f : R → S be an epimorphism of rings. Prove that S is an integral domain if
and only if ker f is a prime ideal of R.
Exercise 10(d)
2. Let {Pa}a[D be a class of prime ideals of a ring such that, for any a and b [ D,
there is g [ D such that Pg ⊆ Pa and Pg ⊆ Pb. Then prove that ∩ Pa is a prime
a∆
ideal of R.
3. Let {Pa}a[D be class of prime ideals of a ring R with unity such that, for any a and
b [ D, there is g [ D such that Pa ⊆ Pg and Pb ⊆ Pg. Then prove that ∪ Pa is a
a∆
prime ideal of R.
4. Let {Pa}a[D be a chain of prime ideals of a ring R with unity (that is, for any a
and b [ D, Pa ⊆ Pb or Pb ⊆ Pa). Then prove that ∩ Pa and ∪ Pa are prime
a∆ a∆
ideals of R.
(iv) I I
M⊆I⊆R⇒M5I or I 5 R.
Example 10.5.1
1. Let us recall that any ideal of Z is of the form nZ for some nonnegative
integer n and that nZ ⊆ mZ if and only if m divides n. Therefore, nZ is
a maximal ideal of Z if and only if n is a prime number.
2. In the ring R of real numbers, {0} is a maximal ideal. In fact, {0} is a
maximal ideal in any field, since a field has two ideals, namely {0} and
the whole field.
3. Let R 5 R 3 R under co-ordinate wise addition and multiplication.
Then, R 3 {0} and {0} 3 R are maximal ideals of R.
4. Consider the ring Mn(R) of n 3 n matrices over the real number system
R. Then, {0} is a maximal ideal of Mn(R), since, by Theorem 10.1.10,
Mn(R) is the only nonzero ideal of Mn(R).
Let us recall that a commutative ring R with unity is a field if and only if R
has exactly two ideals, namely {0} and R. This, together with the fact that the
ideals of R/I are in one-to-one correspondence with the ideals of R containing
I, imply the following important result. However, we prefer to give an inde-
pendent proof in view of its technicality.
Proof: First note that M is a proper ideal of R if and only if R/M is nontrivial.
Suppose that M is a maximal ideal of R. Then, M is a proper ideal and hence
R/M is a nontrivial commutative ring with unity (since so is R). Now, let a 1
M be a nonzero element of R/M. Then, a 1 M M and hence a M. Let I
be the ideal defined by
I 5 M 1 <a> 5 {x 1 ra : x R}.
ab 1 M 5 (a 1 M)(b 1 M) 5 1 1 M
1 5 (1 2 ab) 1 ab I
Proof: This follows from the fact that R/{0} R and from Theorem
10.5.3.
Theorem 10.5.2. Let R be a ring with unity. Then, every maximal ideal of R
is a prime ideal and the converse is not true.
n
a ∑ xi ayi ar sa : xi , yi , r , s R, n 0.
i1
n
1 x ∑ xi ayi ar sa
i1
n
b J ⇒ b 5 1b 5 x ∑ xi ayi ar sa b
i1
n
⇒ b 5 xb 1 ∑ (x a) (y b) (ar ) b ( sa) b
i1
i i
⇒ b M 1 IJ, since x M, a I, b J
⇒ b M, since IJ ⊆ M and M 1 IJ 5 M.
2 5 2 ? 1 5 2(a 2 2n) 5 2a 2 4n I
2 2 M and 2 M.
Proof: This is a consequence of the above and of the fact that M 1 <a> 5 R
if and only if 1 5 x 1 ar for some x M and r R. b
Next, we obtain a general result which assures the existence of suitably
many maximal ideals. The crucial step here is again the Zorn’s Lemma; an
equivalent form of which is given in Zorn’s Lemma 10.4.1.
Theorem 10.5.4. Let R be a ring with unity. Then, any proper ideal of R is
contained in a maximal ideal of R.
Proof: Recall that an ideal of R is proper if and only if it does not contain the
unity of R. Let I be a proper ideal of R. Consider the class
Theorem 10.5.5. Let R be a commutative ring with unity. Suppose that R has
exactly one maximal ideal. Then, 0 and 1 are the only idempotents in R.
and hence a and a 2 1 are zero divisors in R. Therefore, a and a 2 1 are both non-
units and, by Corollary 10.5.3, a M and a 2 1 M. From this, we get that
1 5 a 2 (a 2 1) M
which is a contradiction to the fact that M R. Thus, either a 5 0 or
a 5 1. b
Recall from Corollary 10.4.3, the prime radical of R is defined as the set
of all nilpotents in R and, by Theorem 10.4.4, it is precisely the intersection of
all prime ideals of R. In the following, we introduce another type of radical,
which plays an important role in the structure theory of commutative rings.
Example 10.5.3
1. Recall that the maximal ideals of the ring Z are precisely of the form pZ
for some prime number p. Now, the Jacobson radical of Z is given by
J (Z) ∩ pZ {0},
pP
∩M
x X
x { f F X : f ( x ) 0 for all x X } {0}
1 5 (1 1 a) 2 a M
Theorem 10.5.7. Let R be a commutative ring with unity. Then, the Jacobson
radical J(R) is given by
Proof: If a J(R), then <a> ⊆ J(R) and hence every element of 1 1 <a>
is a unit in R; that is, 1 1 ar is a unit for all r R. Conversely suppose that
1 1 ar is a unit for all r R. Then, every element of 1 1 <a> is a unit and
therefore, by the above theorem, <a> ⊆ J(R) which is equivalent to saying
that a J(R).
Corollary 10.5.4. The following holds for any commutative ring R with
unity.
Proof:
1. Let a J(R) be an idempotent. Then, by Theorem 10.5.7, 1 2 a is a unit
in R and hence there exists b R such that (1 2 a)b 5 1. Now,
a 5 a1 5 a(1 2 a)b 5 (a 2 a2)b 5 0 (since a2 5 a)
Thus, 0 is the only idempotent in J(R).
2. Let a R. Suppose that a 1 J(R) is a unit in R/J(R). Then, there exists
b R such that
(a 1 J(R))(b 1 J(R)) 5 1 1 J(R)
and therefore 1 2 ab J(R). By Theorem 10.5.7, 1 2 (1 2 ab) is a unit
in R and therefore ab is a unit. Thus, a is a unit in R. The converse is
trivial.
3. a N(R) ⇒ a belongs to every prime ideal of R
⇒ a belongs to every maximal ideal of R (since every maximal
ideal is prime)
⇒ a J(R).
Thus, N(R) ⊆ J(R). b
Theorem 10.5.8. Let R be a commutative ring with unity and J(R) be the
Jacobson radical of R. Then, the quotient ring R/J(R) is semisimple.
Proof: We have to prove that the Jacobson radical of R/J(R) is trivial. Let a 1
J(R) be an element in the Jacobson radical of R/J(R). Then, by Theorem
15.5.15,
Mi 5 {a R : ai 5 0}.
Now,
1 1 1
(1, 1, …, 1) ( a1 , a2 , …, an ) , , …,
a1 a2 an
1. I is a prime ideal of B.
2. I is a maximal ideal of B.
3. For any a B, either a I or b 2 ab I for all b B, but not both.
a(1 2 a) 5 a 2 a2 5 0 I.
1 5 a 1 (1 2 a) J
and hence J 5 B.
(2) ⇒ (3): Suppose that I is a maximal ideal of B and a B. Suppose that a
I. Then, I 1 <a> 5 B and hence, for any b B,
b 5 (b 2 ab) 1 ab I.
Worked Exercise 10.5.3. Consider the ring Z[i] of Gaussian integers. Let
since a and b are even and hence as, bt, at and bs are all even. Thus, I is an
ideal of Z[i]. Note that Z[i] is a commutative ring with unity. We shall prove
that I is not a maximal ideal of Z[i]. Let
J 5 {a 1 bi Z[i] : a2 1 b2 is even}.
Observe that a2 1 b2 is even if and only if either both a and b are even or both
a and b are odd. We verify that J is an ideal of Z[i].
Let x 5 a 1 bi and y 5 c 1 di J. Then, a2 1 b2 and c2 1 d2 are even and
x 2 y 5 (a 2 c) 1 (b 2 d)i J
since (a 2 c)2 1 (b 2 d)2 5 (a2 1 b2) 1 (c2 1 d2) 2 2(ac 1 bd), which is
even. Also, for any z 5 s 1 ti Z[i],
I ⊂ J ⊂[i ],
x 2 y 5 (a 2 c) 1 (b 2 d)i I
xz 5 (a 1 bi)(s 1 ti)
5 (as 2 bt) 1 (at 1 bs)i
g.c.d. (3, b2) 5 1 and therefore there exist c and d Z such that 3c 1
b2d 5 1. Now, b2d J and 3c I # J and hence
1 5 3c 1 b2d J
which implies that J 5 Z[i].
2. Suppose that 3 does not divide a and 3 divides b. In this case, using the
technique of (1) above we can prove that J 5 Z[i].
3. Suppose that 3 divides neither a nor b. Then, a 5 3k 1 1 or 3k 1 2 for
some k Z and b 5 3s 1 1 or 3s 1 2 for some s Z. Then,
a2 5 3(3k2 1 2k) 1 1 or 3(3k2 1 4k 1 1) 1 1
and b2 5 3(3s2 1 2s) 1 1 or 3(3s2 1 4s 1 1) 1 1
which imply that a2 1 b2 5 3t 1 2 for some t Z and hence 3 does not
divide a2 1 b2. Put c 5 a2 1 b2. Then, g.c.d.(3, c) 5 1 and hence there
exist integers n and m such that 3m 1 cn 5 1. Now,
3 I # J and c 5 a2 1 b2 5 (a 1 bi)(a 2 bi) J
since a 1 bi J. Therefore, 3 and c J and hence
1 5 3m 1 cn J
which implies that J 5 Z[i].
Thus, in all cases, we have proved that J 5 Z[i]. Therefore, I is a maximal
ideal of Z[i].
x M ∩ N ⇒ x 5 x1 5 xa 1 xb MN
EXERCISES 10(e)
1. Determine all maximal ideals and the Jacobson radicals in each of the following
rings.
(i) R
(ii) Q
(iii) Z
(iv) Z120
(v) Z3Z
(vi) Q3Z
(vii) M3(R)
(viii) Zn, n [ Z1.
5. Let R be a finite commutative ring with unity. Prove that an ideal of R is prime if
and only if it is maximal.
6. Let X be a nonempty finite set with n elements. Prove that there are exactly n
maximal ideals in the ring (P(X), 1, ∩).
7. Let #(X, R) be the set of all real valued continuous functions defined on a topo-
logical space X. Then prove that #(X, R) is a commutative ring with unity under
the point-wise addition and multiplication.
9. Let X be a Compact Hausdorff space and #(X, R) be the ring given in 7 above.
Prove that x M x is a bijection of X onto the set of maximal ideals of #(X, R),
where Mx is the set given in 8 above.
10. Let Z[i] be the ring of Gaussian integers. Prove that the set
I 5 {a 1 bi [ Z[i] : a 2 b is even}.
is a maximal ideal of Z[i] and find the number of elements of the quotient ring
Z[i]/I.
is an ideal of Z[i] and is not maximal. Is this a prime ideal? Estimate the number
of elements in the quotient ring Z[i ] / I .
13. Prove that a proper ideal M of a ring R is maximal if and only if, for any ideal I
of R, either I ⊆ M or I 1 M 5 R.
20. Prove the following for any commutative ring R with unity.
(i) R is semisimple if and only if, for each a [ R, 1 2 ra is a nonunit for
some r [ R.
(ii) If R is regular, then it is semisimple.
(iii) If I is an ideal of R such that R/I is semisimple, then the Jacobson radical
of R is contained in I.
Theorem 10.6.1. Let R be any ring. Then, there exists a ring S with unity
satisfying the following properties:
1. R is embedded in S.
2. R is isomorphic to an ideal of S.
3. R is commutative if and only if S is commutative.
4. char(R) 5 char(S).
S 5 Z 3 R.
(m, a) 1 (n, b) 5 (m 1 n, a 1 b)
and (m, a) ? (n, b) 5 (mn, mb 1 na 1 ab).
S 5 Zn 3 R,
(i, a) 1 (j, b) 5 (i 1n j, a 1 b)
and (i, a) (j, b) 5 (i n j, ib 1 ja 1 ab).
Then, similar to the above case, we can easily prove that (S, 1, ?) is a ring
with unity satisfying the properties (1) through (4). Note that char(S) 5 n 5
char(R).
We have proved earlier that any field is an integral domain and that any
finite integral domain is a field. Also, any subring (with unity) of a field is
an integral domain. In the following theorem, we prove that integral domains
arise as subrings of fields only.
F 5 {[a, b] : (a, b) S}
and hence [ad 1 bc, bd] 5 [a′d′ 1 b′c′, b′d′]. Thus, 1 is well defined. Simi-
larly, we can prove that ? is well defined. In the following, let [a, b], [c, d] and
[s, t] be arbitrary elements of F.
Therefore, 1 is associative.
Therefore, 1 is commutative.
Therefore, F is nontrivial. Also for any [a, b] [0, 1], we have a 0 and
hence [b, a] F and
Definition 10.6.2. For any integral domain R, the field F constructed above
is called the field of quotients of R.
Example 10.6.1. The ring Z of integers is an integral domain and the field
of quotients of Z is precisely the field Q of rational numbers. A rational
number is usually written as a/b which is precisely [a, b]. Recall from the
high school mathematics that two rational numbers a/b and c/d are equal if
and only if ad 5 bc and that a/b represents a class of pairs (c, d) for which
ad 5 bc.
By means of the monomorphism f of an integral domain R into the field of
quotients F defined by f (a) 5 [a, 1], we can identify an element a in R with
the element [a, 1] in F. With this identification, we can treat R as a subring
of F. Also, any element [a, b] of F can be expressed as ab21 with a R and
0 b R, since
[a, b] 5 [c, d] ⇔ ad 5 bc
⇔ ab21 5 cd21
Answer: We can treat R as a subring of F and the unities in R and F are the
same. In Theorem 9.3.2, we have proved that the characteristic of any ring
with unity is precisely the order of the unity in the additive group of the ring.
Therefore,
Worked Exercise 10.6.2. Let Z[i] be the ring of Gaussian integers. Deter-
mine the field of quotients of Z[i].
and that Z[i] is an integral domain under the addition and multiplication
of complex numbers. Let C be the field of complex numbers. Then, Z[i] is
a subring of C. By Theorem 10.6.3, the field of quotients of Z[i] is equal
(isomorphic) to
a bi ( a bi )(c di )
st 1
c di (c di )(c di )
ac bd (bc ad )i
c2 d 2
ac bd bc ad
2 i
c d 2 c2 d 2
Since a, b, c and d are all integers, it follows that st21 belongs to the set
Worked Exercise 10.6.3. Prove that any isomorphism between two integral
domains can be extended to their fields of quotients.
ab21 5 cd21 ⇔ ad 5 bc
⇔ f (ad) 5 f (bc)
⇔ f (a)f (d) 5 f (b)f (c)
⇔ f (a)f (b)21 5 f (c)f (d)21
This shows that g is well defined and is an injection. Also, if xy21 F′ with x,
y R′ and y 0, we can choose elements a and b in R such that f (a) 5 x and
f (b) 5 y (since f is a bijection). Then, b 0 and g(ab21) 5 f (a)f (b)21 5 xy21.
Therefore, g is a surjection also. Further, for any ab21, cd21, F,
5 f (ac)f (bd)21
5 f (a)f (c)(f (b)f (d))21
5 f (a)f (b)21f (c)f (d)21
5 g(ab21)g(cd21).
EXERCISE 10(f)
6. Prove that every field contains a subfield, that is, isomorphic to Q or Zp (such
subfields are called prime subfields).
7. Prove that any automorphism of a field F fixes every element of the prime
subfield.
9. If F is a subfield of a field K, then prove that F and K have the same prime
subfields.
10. If a field F has exactly 9 elements, then prove that the prime subfield of F is
isomorphic to Z3.
11. If Z5 is the prime subfield of a field F, then prove that there exists a F such
that a 1 a 0.
12. Prove that any automorphism of an integral domain can be extended uniquely to
an automorphism of its field of quotients.
13. Let R be a commutative ring with no zero divisions. Then prove that R can be
embedded in an integral domain.
14. Prove that a commutative ring can be embedded in a field if and only if it has no
zero divisors.
a
S 1R : a ∈ R and s S .
s
Then prove that (S21R, 1, ?) is a commutative ring with unity. This ring is called
the ring of fractions of R by S.
16. Prove that the field of quotients of an integral domain R is the ring of fractions
of R by R 2 {0}.
17. Let P be a prime ideal of a commutative ring R with unity. Then prove that
(R 2 P)21R has a unique maximal ideal.
18. Let S be a multiplicative subset of a commutative ring R with unity. Prove that
there is a one-to-one correspondence between the prime ideals of R disjoint with
S and the prime ideals of the ring S21R of fractions of R by S.
We are very familiar with polynomials which are introduced to us very early
in our mathematical education, in fact, in high school itself, we are thoroughly
drilled in adding, multiplying, dividing, factoring and simplifying them. We
have learnt the remainder theorem in eighth or ninth standard. Later, at higher
level, polynomials appear as functions and we were concerned with their con-
tinuity, derivatives and integrals and their maxima and minima. Now, we are
interested in polynomials, but from neither of the above view points. Here,
polynomials will simply be elements of a certain ring and we shall be con-
cerned with the algebraic properties of this ring.
At the secondary school level, we have studied polynomials with integer
coefficients, rational coefficients, real coefficients and, may be even complex
coefficients. Notice that, in each case, the set of coefficients is a ring and the
set of polynomials also forms a ring under suitable addition and multiplica-
tion, with which we are all familiar. In this chapter, we make an abstraction
of these cases and study polynomials with coefficients from a given abstract
ring.
a0 a1 x a2 x 2 an x n .
( a0 , a1 , a2 , …, an , …)
of elements of R such that an’s are zero for all but finite number of n’s; equiva-
lently, there exists a nonnegative integer k such that an 5 0 for all n k. The
set of all polynomials over R will be denoted by Poly(R).
Recall that an infinite sequence of elements of R can be viewed as a
mapping of the set of nonnegative integers into R. Let us agree that two
polynomials
f ( a0 , a1 , a2 , …) and g (b0 , b1 , b2 , …)
Definition 11.1.2. For any polynomials a 5 (a0, a1, a2, …) and b 5 (b0, b1,
b2, …) over a ring (R, 1, ?), define
a b ( a0 b0 , a1 b1 , a2 b2 , …)
and a ⋅ b (c0 , c1 , c2 , …),
n
where cn a0 bn a1bn1 an b0 ar bnr ar bs .
r0 rsn
Note that the additive operation 1 and the multiplicative operation ? on the
right sides of the above defining equations are those in the (R, 1, ?). Since a
and b are in Poly(R), there exist nonnegative integers m and n such that
ai bi 0
i
and ci ar bir 0 for all i max{m, n}
r0
Theorem 11.1.1. For any ring (R, 1, ?), (Poly(R), 1, ?) is a ring, where 1 and ?
are the operations defined above.
a ( a0 , a1 , a2 , …)
b (b0 , b1 , b2 , …)
and c (c0 , c1 , c2 , …)
with ai’s, bi’s and ci’s are elements in the given ring R. Then, using the asso-
ciatively and commutativity of 1 in R, we can prove that
( a b) c a (b c) and a b b a.
Therefore, 0 is the identity element for 1. Also, for any a 5 (a0, a1, a2, …) in
Poly(R), the polynomial 2a defined by
a (a0 , a1 , a2 , …)
a ⋅ b ( d0 , d1 , d2 , …)
and ( a ⋅ b) ⋅ c ( x0 , x1 , x2 , …).
Then,
n n
dn ∑ ar bnr ∑ ar bs
r0 rs
n
and xn ∑ ds cns ∑ ds ct
s0 stn
∑ ar bu ct
stn
r∑
us
∑
rutn
ar bu ct
∑ a ∑ b c
rsn
r
uts
u t
( a ⋅ b) ⋅ c a ⋅ (b ⋅ c).
yn ∑ a (b c )
rsn
r s s
∑ab ∑ac
rsn
r s
rsn
r s
Definition 11.1.3. For any ring R, (Poly(R), 1, ?) is called the ring of polyno-
mials over R and is simply denoted by Poly(R).
In the following, we prove that any given ring R is isomorphic to a subring
of the ring of polynomials over R.
f ( a b) ( a b, 0, 0, …)
( a b, 0 0, 0 0, …)
( a, 0, 0, …) (b, 0, 0, …)
f ( a) f ( b)
and f ( ab) ( ab, 0, 0, …)
( a, 0, 0, …) (b, 0, 0, …)
f ( a) f (b).
f ( a) f (b) ⇒ ( a, 0, 0, …) (b, 0, 0, …)
⇒ a b.
Theorem 11.1.3. For any ring R, the ring Poly(R) is commutative if and only
if R is commutative.
∑ba
srn
s r
th
n term in b a
Theorem 11.1.4. A ring R is with unity if and only if the ring Poly(R) is with
unity.
nth term in e ⋅ a
Proof: First note that, from Theorems 11.1.1 and 11.1.4, R is a nontrivial
commutative ring with unity if and only if so is Poly(R). Therefore, we can
∑
rsnm
ar bs an bm 0,
Proof: Let R be a ring and suppose that Poly(R) is a field. Then, Poly(R) is an
integral domain and hence, by Theorem 11.1.5, R is also an integral domain.
Consider the element
x a a x 1 (1, 0, 0, …)
hence ( a, a0 , a1 , a2 , …) (1, 0, 0, …).
a0 a1 x1 a2 x 2 an x n ,
where a0, a1, a2, …, an are real numbers and x is an indeterminate. Though we
are familiar with this, we did not know what x is what ax is for any a R.
Further, we should give a mathematically valid explanation for the operation
symbol 1 in the above expression of a polynomial. We shall give satisfactory
answers to these questions in the following result. Recall that R can be identi-
fied with a subring of Poly(R) and any element a in R can be identified with
the polynomial (a, 0, 0, …).
Theorem 11.1.7. Let (R, 1, ?) be a ring with unity and x be the polynomial
over R given by
x (0, 1, 0, 0, …).
a0 a1 x a2 x 2 an x n ,
where a0, a1, …, an are elements of R, identified with the elements of Poly(R).
x (0, 1, 0, 0, …)
x 2 x ⋅ x (0, 0, 1, 0, 0, …)
x 3 x 2 ⋅ x (0, 0, 0, 1, 0, 0, …)
x n x n1 ⋅ x (0, 0, …, 0, 1, 0, 0, …)
( n 1) th term
for any positive integer n. Also, for any a R, by identifying a with (a, 0,
0, …) in Poly(R), we have
a ⋅ x (0, a, 0, 0, …)
a ⋅ x 2 (0, 0, a, 0, 0, …)
a ⋅ x 3 (0, 0, 0, a, 0, 0, …)
n
a ⋅ x (0, 0, …, 0, a, 0, 0, …)
( n 1) th term
p ( a0 , a1 , a2 , …, an , 0, 0, …).
Now, we have
Remarks 11.1.1
1. If we identify a R with (a, 0, 0, …) in Poly(R) and identify R as a
subring of Poly(R), then Poly(R) is the subring generated by R and x. For
this reason, we prefer to write R[x] for Poly(R).
2. The expression a0 1 a1x 1 a2x2 1 … 1 anxn for a polynomial looks
simple and elegant for two reasons. The first one is that we are familiar
with this right from our school days. The second is that we can straight
away multiply two polynomials by treating x also as a real number or as
an element in the ring containing ai’s.
3. As mentioned in the above theorem and its proof, x is not an element
in the ring R and it is an element in Poly(R); that is, x is a polynomial
over R.
4. x is usually called an indeterminate.
5. When we are completely aware of what Poly(R) and R[x] are, we prefer
to use familiar notation for polynomials over a given ring R and for the
ring of polynomials over R. R[x] is the most standard notation used to
denote the ring polynomials over R.
6. The expression a0 1 a1x 1 … 1 anxn for a polynomial is known as poly-
nomial in an indeterminate form. Though x is called an indeterminate, it
is actually a polynomial by itself and the operations 1 and ? in the above
expression are the addition and multiplication of polynomials only.
7. Often, for
n
convenience, a polynomial a0 1 a1x 1 … 1 anxn is also writ-
ten as ai x i , with the assumption that x0 5 1.
i1
Example 11.1.1
p 5 a0 1 a1x 1 … 1 anxn
The ai’s involved in this expression are called the coefficients in the poly-
nomial p.
Example 11.1.2
1. The degree of 2 1 3x4 in R[x] is 4, since
2 1 3x4 5 2 1 0x 1 0x2 1 0x3 1 4x4
2. In Z4[x], the degree of (1 1 2x)2 is 0, since
(1 1 2x)2 5 1 1 (2 14 2)x 1 (2 ?4 2)x2
5 1 1 0x 1 0x2 5 1.
Definition 11.1.5. The zero polynomial and the polynomials of degree 0 are
called constant polynomials.
For any ring R, the set of constant polynomials over a ring R form a sub-
ring of R[x] and is isomorphic to R. In the following, we discuss how the
degrees of polynomials vary when we take sums and products.
Definition 11.1.6. Let R be a ring and R[x] be the ring of polynomials over R.
The following holds for any nonzero polynomials f and g R[x]:
1. Either f 1 g 5 0 or deg(f 1 g) max{deg(f ), deg(g)}.
2. Either f ? g 5 0 or deg(f ? g) deg(f ) 1 deg(g).
3. If R is an integral domain, then f ? g 0 and deg(f ? g) 5 deg(f ) 1 deg(g).
Cmn ∑
i jmn
ai b j am bn 0 (since R is an integral domain)
Proof: We have proved (2) ⇔ (3) in Theorem 11.1.5 and (2) ⇒ (1) by
Definition 11.1.6 (3).
(1) ⇒ (3) follows from the fact that the degree is defined for nonzero poly-
nomials only and hence, by (1), f ? g 0 for all nonzero f and g in R[x]. b
Worked Exercise 11.1.3. Let R be a commutative ring with unity. Then, prove
that R and R[x] have the same characteristic.
Answer: First note that both R and R[x] have the same unity, namely 1. Now,
it follows from Theorem 9.3.2 that
EXERCISE 11(a)
12. Let R be a commutative ring with unity. Prove that a0 1 a1x 1 … 1 anxn R[x] is
a unit in R[x] if and only if a0 is a unit in R and a1, a1, …, an are nilpotents in R.
14. Let R be a commutative ring with unity and I be a proper ideal of R. Then prove
that I is a prime ideal of R if and only if I[x] is a prime ideal of R[x].
16. Let R be a commutative ring with unity. Then prove that R[x]/x R, where
x is the ideal generated by x in R[x].
Example 11.2.1
1. If f (x) 5 2 1 3x 1 x2 1 6x3 Z[x], then degree of f (x) is 3 and 6 is the
leading coefficient in f (x).
2. If f (x) 5 2 2 x2 Z[x], then 21 is the leading coefficient of f (x).
3. 3 1 x 2 2x3 1 x4 is a monic polynomial over Z.
Theorem 11.2.1 (Division Algorithm for Polynomials). Let f (x) and g(x) be
polynomials over a commutative ring R with unity such that g(x) 0 and the
Proof: If f (x) 5 0 or f (x) 0 such that deg(f (x)) deg(g(x)), then we can
take q(x) 5 0 and r(x) 5 f (x). Therefore, we can assume that f (x) 0 and
deg(f (x)) $ deg(g(x)).
We apply induction on the degree of f (x). First, let deg(f (x)) 5 0. Then, since
deg(f (x)) deg(g(x)) 0, it follows that deg(g(x)) 5 0 and hence both f (x)
and g(x) are constant polynomials, so that f (x) and g(x) are elements of R. In
particular, the leading coefficient of g(x) is g(x) itself and is invertible in R (by
the hypothesis). Now, put
f1 ( x ) f ( x ) an bm1 x nm g ( x ).
f ( x ) ( q1 ( x ) an bm1 x nm ) g ( x ) r ( x )
5 q(x)g(x) 1 r(x),
where r(x) and r9(x) satisfy the requirements of the theorem. Then, we get that
which is a contradiction. The last in equality is based on the fact that the
degrees of both r(x) and r9(x) are less than deg(g(x)). Thus, it is necessary
that q9(x) 2 q(x) 5 0 and hence r(x) 2 r9(x) 5 0. Therefore, q(x) 5 q9(x) and
r(x) 5 r9(x).
Definition 11.2.2. The polynomials q(x) and r(x) in the above theorem and
called, respectively, the quotient and remainder on dividing f (x) by g(x). The
proof of the above theorem actually provides an algorithm to find the quotient
and remainder and hence the theorem is called the division algorithm. Let us
take up an example.
Now, deg(f3(x)) 5 1 deg(g(x)) and the process stops, we have, from (1),
(2) and (3),
f1 ( x ) f ( x ) an bm1 x nm g ( x ),
where an and bm are leading coefficients of f (x) and g(x), respectively. Let c1
be the leading coefficient of f1(x) and n1 5 deg(f1(x)). Let
f 2 ( x ) f1 ( x ) c1bm1 x n1m g ( x )
and continue the process of obtaining polynomials f0(x) 5 f (x), f1(x), f2(x),
f3(x), … of degrees n0 5 n, n1, n2, n3, … with leading coefficients c0 5 an, c1,
c2, c3, …, respectively. Then,
At some stage, we should get fs(x) such that deg(fs(x)) m 5 deg(g(x)) and
let fs(x) be first such stage. Then,
f (x ) an bm1 x nm g ( x ) f1 ( x )
c0 bm1 x n0 m g ( x ) c1bm1 x n1m g ( x ) f 2 ( x )
…
s1
∑ bm1ct x nt m g ( x ) f s ( x )
t0
f (x) 5 q(x)g(x) 1 r(x),
s1
where q(x) bm1ct x nt m and r ( x ) f s ( x ) and these are the quotient and
t0
remainder, respectively.
The following are an immediate consequences of the division algorithm
(Theorem 11.2.1).
Corollary 11.2.1. Let f (x) and g(x) be polynomials over a commutative ring
R with unity. If g(x) is a nonzero monic polynomial, then there exist unique
q(x) and r(x) in R[x] such that
Corollary 11.2.2. Let F be a field and f (x) and g(x) F[x] with g(x) 0.
Then, there exists unique q(x) and r(x) in R[x] such that
fa : R[x] → S
Proof: First of all, observe that a S and a0, a1, …, an R ⊆ S and hence
a0 1 a1a 1 … 1 anan S. Also, a0 1 a1x 1 … 1 anxn 5 b0 1 b1x 1 … 1
bnxn implies that ai 5 bi for all 1 i n and hence a0 1 a1a 1 … 1 anan 5
b0 1 b1a 1 … 1 bnan. Therefore, fa is a well-defined mapping of R[x]
into S.
For any a R and the evaluation homomorphism fa, we often write f (a)
for fa(f ), where f 5 f (x) is a polynomial over R. This is to say that fa(f ) is
just the element of R obtained by substituting a for x in the polynomial f 5
f (x). The following is an important result to which we were exposed at school
level itself and now, we supplement a proof of the most general version of the
remainder theorem.
where f (a) is the element in R obtained by substituting a for x in f (x); that is,
f (a) 5 a(f (x)).
Definition 11.2.5. For any polynomial f (x) over R and a R, we say that a is
a root (or zero) of f (x) if f (a) 5 0.
In other words, a is a root of f (x) if and only if f (x) 5 (a 2 x)g(x) for some
g(x) R[x]. The next theorem is about the number of roots of a polynomial
in a integral domain.
Proof: We apply induction on the degree of f (x), If deg(f (x)) 5 0, then the
theorem is trivial, since f (x), being nonzero, cannot have any root in R.
Next, suppose that deg(f (x)) 5 n 0 and assume that any nonzero polyno-
mial of degree m n can have at most m distinct roots in R. If f (x) has a root
in R; that is, if a R is such that f (a) 5 0, then, by Theorem 11.2.5,
0 5 f (b) 5 (a 2 b)g(b)
Corollary 11.2.4. Let f (x) and g(x) be polynomials of degree n over an inte-
gral domain R and a1, a2, …, an11 be distinct elements of R such that f (ai) 5
g(ai) for 1 i n. Then, f (x) 5 g(x).
Proof: Consider h(x) 5 f (x) 2 g(x) R[x]. If h(x) 0, then deg (h(x))
n and a1, …, an11 are distinct roots of h(x) in R, which is a contradiction to
Theorem 11.2.5. Therefore, h(x) 5 0 and f (x) 5 g(x).
Proof: If f (x) is a nonzero polynomial, then deg(f (x)) 5 n and f (x) can have
at most n district roots in R which is a contradiction to the assumption that
each element of the infinite set A is a root of f (x).
Worked Exercise 11.2.1. Find the quotient and remainder when f (x) 5 3 1
4x 1 3x2 1 2x3 1 2x4 1 x5 is divided by g(x) 5 5 1 3x 1 4x2 1 x3 in Z6[x].
3 4 x 3 x 2 2 x 3 2 x 4 x 5 2
5 3x 4 x 2 x 3 x
(5 x 2 3x 3 4 x 4 x 5 )
3 4 x 4 x 2 5 x 3 4 x 4
4x
(2 x 0 4 x 3 4 x 4 )
3 2 x 4 x 2 x 3
1
(5 3 x 4 x 2 x 3 )
4 5x
R[x ] R.
ker fa
so that f ker fa. Using the division algorithm (whose proof is given in the
next section), one can prove that fa(f ) 5 0 implies f 5 (a 2 x)g for some g
R[x]. Therefore,
ker fa 5 a 2 x
Thus, R[x ]/ a x R.
Exercise 11(b)
1. Find the remainder and quotient when f (x) is divided by g (x) in the rings men-
tioned in each of the following:
(i) f (x) 5 2 1 3x 1 x2 1 x4 1 2x5 and g(x) 5 2 1 x2 2 x3 in Z[x].
(ii) f (x) 5 1 1 x 1 x2 1 x4 1 x6 and g(x) 5 1 1 x 1 x2 in Z2[x].
1 2 3 1 1
(iii) f ( x ) x x 2 x 3 and g ( x ) x in Q[ x ].
2 3 4 3 2
(iv) f (x) 5 2 1 3x 1 4x2 1 5x3 and g(x) 5 1 2 x in Z[x].
(v) f (x) 5 1 1 2x 1 3x2 1 4x3 and g(x) 5 1 1 x2 in Z5[x].
(vi) f (x) 5 (1 1 i) 1 (2 1 3i)x 1 (1 2 2i)x2 1 (1 1 3i)x3 and g(x) 5 i 1
(2 1 i)x 2 2x2 in C[x].
2. Evaluate each of the following for the indicated evaluation homomorphism fa :
Z5[x] → Z5.
(i) f0(2 1 3x 1 4x2 1 x3)
(ii) f1(1 1 x 1 x2 1 x3 1 x4 1 x5)
(iii) f2(1 1 2x 1 3x2 1 4x3 1 x4)
(iv) f3(2 1 3x 1 4x2 1 x5 1 x6)
(v) f4(2 1 x 1 3x2 1 x4 1 x7)
3. Find eight elements in the Kernel of the evaluation homomorphism f5: Q[x] → R.
4. For any subfield F of any field E, prove that the Kernel of the evaluation homo-
morphism fa: F[x] → E is an infinite set for each a [ E.
5. Find all the roots in Z5 of the polynomial 2 1 4x 1 3x2 1 4x3 1 x4 in Z5[x].
6. Prove that 1 1 4x is a unit in Z8[x].
7. Let F be an infinite field and f (x) [ F[x]. Prove that f (a) 5 0 for infinitely many
elements a in F if and only if f (x) 5 0.
8. Let R be an integral domain and f (x) and g(x) [ R[x]. Prove that {a [ R : f (a) 5
g(a)} is infinite if and only if f (x) 5 g(x).
9. Determine all the roots of x 2 x5 in Z5.
10. For any prime number p, prove that every element of Zp is a root of the polyno-
mial x 2 xp in Zp[x].
Proof: Let F be a field, 0 f (x) [ F[x] and deg(f (x)) 5 n. We shall use
induction on n. If n 5 0, there is nothing to prove. If n 5 1, then f (x) 5 a0 1
a1 x and a1 0 and hence a0 a11 is the only root of f (x).
Suppose that n 1 and that the theorem is true for all polynomials of degree
less that n. If f (x) has no roots in F, then we are done. Let a be a root of f (x)
in F. Then,
0 5 f (b) 5 (a 2 b)mg(b)
Theorem 11.3.2. Let F be a field and F[x] be the ring of polynomials over F.
Then, F[x] is an integral domain in which every ideal is principal.
since g(x) and f (x) [ I and I is an ideal. By the least property of n, it follows
that r(x) 5 0 and hence
Theorem 11.3.3. Let R be a ring such that R[x] is an integral domain in which
every ideal is principal, Then, R is a field.
a, x 5 f (x).
First of all, note that f (x) 0, since 0 a [ f (x). Since both a and x [
f (x), we get that
a 5 f (x)g(x) (1)
and x 5 f (x)h(x) (2)
Therefore, h(x) 5 b0 1 b1x for some b0 and 0 b, in R and hence, from (2),
we have
x 5 a0(b0 1 b1 x).
1 5 af1(x) 1 xf2(x)
which implies that 1 5 ac0 where c0 is the constant term in f1(x). Thus, a is a
unit in R. Therefore, R is a field.
Answer: Since Z is not a field, it follows from Definition 13.3.2 that there
must be an ideal of Z[x] which is not principal. Let
Then, one can easily verify that I is an ideal of Z[x]. We prove that I is not a
principal ideal. On the contrary, suppose that
Then, 0 5 deg(2) 5 deg(f (x)g(x)) 5 deg(f (x)) 1 deg(g(x)), so that deg(f (x)) 5
0 5 deg(g(x)). Let f (x) 5 b and g(x) 5 c [ Z. Then, 2 5 bc, we can assume
that b 0 and c 0 (since ,2f (x) 5 f (x)). Then, b 5 1 or 2.
But b 1, since 1 I. Therefore, b 5 2. This implies that I 5 2, which
is a contradiction, since 2 1 x [ I and 2 1 x 2. Thus, I f (x) for
any f (x) [ Z[x]. That is, I is not a principal ideal in Z[x].
∩M
a∈F
a {0}.
∩M
a∈F
a {0} ⇔ F is infinite.
Therefore, an 5 a for all a [ F. Put f (x) 5 x 2 xn. Then, 0 f (x) [ F[x] and
f (a) 5 0 for all a [ F. Therefore, 0 f (x) [ ∩ M a . Thus, ∩ M a {0}.
a∈F a∈F
Converse of this is proved in Worked Exercise 11.3.2.
f (x) [ M0 ⇔ f (0) 5 0
⇔ x divides f (x) (by Corollary 11.2.3)
⇔ f (x) [ x.
R[x]/x R.
Exercise 11(c)
2. Let R be a commutative ring with unity in which every ideal is principal. Then
prove that an ideal in R is maximal if and only if it is prime.
3. Let F be a field, 0 a [ F and f (x) [ F[x], prove that f (x) 5 af (x).
5. Let F be a field and I be the set of all polynomials over F for each of which the
sum of the coefficients is zero. Then prove that I is an ideal of F[x] and determine
a generator of I.
6. Prove that there are infinitely many polynomials f (x) in Z3[x] for each of which
every element of Z3 is a root.
9. Let f (x) [ R[x], a [ R and f 9(x) be the derivative of f with respect to x. Then
prove that f (a) 5 0 5 f 9(a) if and only if (a 2 x)2 divides f (x).
10. In Z[x], prove that x is a prime ideal but not a maximal ideal.
11. Let P be a prime ideal of a commutative ring R with unity. Then prove that P[x]
is a prime ideal of R[x]. If M is a maximal ideal of R, is M[x] a maximal ideal
of R[x]?
12. Prove that R[x]/1 1 x2 is isomorphic to the field of complex numbers.
13. Prove that Z[x]/1 1 x2 is isomorphic with the ring of Gaussian integers
Z[i].
16. Let f (x) [ Z[x] be a monic polynomial and a be a rational number such that
f (a) 5 0. Then prove that a must be integer.
If f (x) is not irreducible and deg(f (x)) 0, we say that f (x) is reducible.
Note that the above definition applies only to polynomials of positive
degree and as such the constant polynomials are neither reducible nor
irreducible. Also, the irreducibility of a polynomial f (x) [ R[x] depends
much on the integral domain R; that is, a given polynomial may be irre-
ducible when viewed as a polynomial over one domain, yet reducible when
viewed as a polynomial over another domain. For, consider the following
example.
Example 11.4.1
1. The polynomial 1 1 x2 is irreducible over R the field R of real numbers;
but it is reducible over the field C of complex numbers, since 1 1 x2 5
(1 1 ix)(1 2 ix) and 1 1 ix and 1 2 ix [ C[x].
2. 1 1 x2 is reducible in Z2[x], since 1 1 x2 5 (1 1 x)(1 1 x) in Z2[x]; but
1 1 x2 is irreducible in Z3[x].
Thus, to ask merely whether a polynomial is irreducible, without specifying
the coefficient ring involved, is incomplete and meaningless. More often, it
is a formidable task to decide when a given polynomial is irreducible over
a specific ring. The following provide certain simple tips in finding a given
polynomial to be irreducible over a given field or an Integral Domain.
Theorem 11.4.1
1. Let R be an integral domain and f (x) [ R[x] with deg(f (x)) 5 1. Then,
f (x) is irreducible over R.
2. Let F be a field and f (x) [ F[x] with deg(f (x)) 5 2 or 3. Then, f (x) is
irreducible if and only if f (x) has no root in F.
Proof:
1. Recall that the degree of any nonzero polynomial is a nonnegative inte-
ger. If f (x) 5 g(x)h(x) and g(x);h(x) [ R[x], then
1 5 deg(f (x)) 5 deg(g(x)) 1 deg(h(x))
and hence deg(g(x)) 5 0 or deg(h(x)) 5 0. Therefore, f (x) cannot be
expressed as a product of two polynomials of positive degree. Thus, f (x)
is irreducible over R.
2. We shall prove that f (x) is reducible over F if and only if f (x) has a root
in F. First note that any polynomial a0 1 a1x, a1 0 of degree over F
one has a root, namely, a11a0 in F. Suppose that f (x) is reducible over
F. Then,
Theorem 11.4.2. Let f (x) be a nonzero polynomial over a field F. Then, the
following are equivalent
1. f (x) is irreducible over F.
2. f (x) is a maximal ideal of F[x].
3. f (x) is a prime ideal of F[x].
Proof: (1) ⇒ (2): Suppose that f (x) is irreducible over F. Then, f (x) is not
a constant polynomial and hence f (x) is not a unit in F[x], so that f (x)
is a proper ideal of F[x]. Let I be any ideal of F[x] containing f (x). By
Theorem 11.3.2, I 5 g(x) for some g(x) [ F[x]. Since f (x) ⊆ I 5
g(x), we get that f (x) 5 g(x)h(x) for some h(x) [ F[x]. Since f (x) is irre-
ducible, it follows that either g(x) or h(x) is a constant, If g(x) is a constant,
then g(x) is a unit (note that f (x) 0 and h(x) 0, since deg(f (x)) 0 and
hence f (x) 0) so that I 5 g(x) 5 F[x]. If h(x) is a constant, then h(x) is
a unit in F[x] and g(x) 5 f (x)h(x)21 [ f (x) so that I ⊆ f (x) and hence
I 5 f (x). Thus, f (x) is a maximal ideal of F[x].
(2) ⇒ (3): This is trivial, since any maximal ideal of a commutative ring with
unity is a prime ideal.
(3) ⇒ (1): Suppose that f (x) is a prime ideal of F[x]. Then, f (x) is a
proper ideal and hence f (x) is not a unit. Therefore, deg(f (x)) 0. For any
g(x) and h(x) F[x],
Corollary 11.4.3. Let f (x) be a nonconstant monic polynomial over the real
number system R. Then, f (x) can be expressed as a product of polynomials
over R, each of degree 2 or 1.
Proof: Let deg(f (x)) 5 n. Since f (x) R[x] ⊆ C[x], it follows from Corollary
11.4.1 that
gj(x) 5 s 2 x, s C.
f (s) 0 0.
g j ( x ) g k ( x ) ( s x )( s x )
5 (a 1 bi 2x)(a 2 bi 2 x)
5 (a2 1 b2) 2 (a 1 bi 1 a 2 bi)x 1 x2
5 a2 1 b2 2 2ax 1 x2 R[x].
Answer: If there are any factors f (x), then atleast one factor must be of degree
one, say a 1 bx (with a and b Z2 and b ≠ 0). In this case, 2b21a is a root of
a 1 bx and hence of f (x). Therefore, if f (x) is reducible, then f (x) must have
a root in Z2; but it can be easily verified that f (0) 5 1 5 f (1) (sense 1 1 1 5
0 in Z2) . Therefore, f (x) is irreducible over Z2.
Worked Exercise 11.4.2. Let f (x) R[x] and s be a complex number. Then,
prove that s is a root of f (x), if and only if s is a root of f (x), where s is the
complex conjugate of s.
f (s ) a0 a1 s an (s ) n
a0 as a n s n
a0 a1s an s n
f (s) 0 0.
Thus, s is a root of f (x). The converse follows from the fact that s 5 s.
Exercise 11(d)
2. Let f (x) be a nonconstant polynomial over a field F. Then, prove that f (x) is
irreducible if and only if F [ x ]/ f ( x ) is a field.
3. Prove that Z2[x]/1 1 x 1 x2 is a field and determine the number of elements in it.
6. Let F be a field, f (x) F[x] and 0 ≠ a F. Then prove that f (x) is irreducible
over F if and only if so is f (x).
10. Let R[x] be the ring of polynomials over a commutative ring R with unity. Then,
the ring of polynomials over R[x] will be denoted by R[x, y]; that is, R[x, y] 5
R[x][y]. By induction, we define
R[x1 x2, …, xn] 5 R[x1 …, xn21][xn]
14. Prove that Z4[x] has infinitely many units and infinitely many nilpotent elements.
6,300 5 2 3 2 3 3 3 3 3 5 3 5 3 7
and 2, 3, 5 and 7 are prime numbers. In this chapter, we extend the factoriza-
tion theory of the ring Z and, in particular, the above-mentioned Fundamental
Theorem of Arithmetic, to a more general setting. Naturally, any reasonable
abstraction of these number theoretic ideals depends on a suitable interpreta-
tion of prime elements (the building blocks for the study of divisibility prob-
lems in Z). All the topics discussed in this chapter are more concerned with
integral domains. We proceed from the most general results about divisibility,
prime elements and factorization to stronger results concerning certain spe-
cific classes of integral domains.
First, let us recall that an integral domain is a nontrivial commutative ring
with unity and without zero divisors (or equivalently, product of two nonzero
elements is again nonzero).
Example 12.1.1
1. In the ring Z of integers, 1 and 21 are the only units and hence a ~ b if
and only if |a| 5 |b|, for any a and b Z, where |a| is the absolute value
of a.
2. In a field F, each nonzero element is a unit and hence a ~ b (since a 5
b(b21a)) for any nonzero elements a and b in F. Therefore, ã 5 R 2 {0}
for any 0 a R.
3. Consider the ring Z[i] of Gaussian integers in which 1, 21, i and 2i are
the only units. For any x 5 a 1 bi Z[i],
x 5 {x, 2x, ix, 2ix}
5 {a 1 bi, 2a 2 bi, 2b 1 ai, b 2 ai}.
4. Let R[x] be the ring of polynomials over an integral domain R. Then, the
units of R[x] are precisely the units in R. For any f (x) R[x],
f (x ) = {uf ( x ) : u is a unit in R}.
Proof:
1. a|b ⇔ ax 5 b for some x R
⇔ b aR 5 <a>
⇔ <b> ⊆ <a>.
if and only if the ideal <a1, a2, …, an> generated by the set {a1, a2, …, an} in
R is the principal ideal.
Proof: Suppose that d 5 a1r1 1 a2r2 1…1 anrn (ri R) is a g.c.d. of a1, a2,
…, an in R. Then, d|ai and hence ai <d> for all 1 i n. Therefore,
where <a1, a2, …, an> is the ideal generated by a1, a2, …, an in R. Also, since
Conversely, suppose that <a1, a2, …, an> 5 <d> for some d R. Then, ai
<d> and hence d|ai for all 1 i n. If c is a common divisor of a1, a2, …,
an, then c|ai and hence ai <c> for all 1 i n, so that
it follows that d 5 a1r1 1 a2r2 1…1 anrn for some r1, r2, …, rn R.
It is well known that the ring Z of integers is an integral domain in which
every ideal is a principal ideal. This together with the above theorem implies
the following corollary.
Corollary 12.1.1. Any nonzero integers a1, a2, …, an have g.c.d. and
numbers. For any x a b 5, let |x| be the usual modulus of the complex
number x a ib 5; that is,
|x| = a 2 + 5b 2 .
3 3 9 ( 2 5 )( 2 5 ) p( 2 5 ).
Theorem 12.1.6. The following holds for any nonzero and nonunit p in an
integral domain R.
1. p is a prime element in R if and only if <p> is a prime ideal of R.
2. p is an irreducible element in R if and only if <p> is a maximal principal
ideal of R.
Proof:
1. is trial because of the fact that x <p> if and only if p divides x.
2. Suppose that p is an irreducible element in R and <p> ⊆ <x>, x R.
Then, <p> R (since p is a nonunit) and p <x> and hence p 5 xy for
some y R. Since p is irreducible, either x or y is a unit in R. If x is a
unit, then <x> 5 R. If y is a unit, then x 5 py21 <p> and hence <x> ⊆
<p>, so that <x> 5 <p>. Thus, <p> is a maximal principal ideal of R.
Conversely, suppose that <p> is a maximal principal ideal of R. Then, <p> R
and hence p is not a unit. Suppose a and b R such that p 5 ab. Then, p
<a> and hence <p> ⊆ <a>. By the maximality of <p>, either <p> 5 <a> or
<a> 5 R. If <a> 5 R, then a is a unit. If <p> 5 <a>, then a 5 pc for some
c R and hence
p 5 ab 5 (pc)b 5 p(cb)
Thus, q is prime. Converse follows from the fact that p ~ q if and only
if q ~ p.
2. Suppose that p is irreducible. Then, p is nonzero and nonunit and hence
so is q. Let a and b R such that q 5 ab. Then, p 5 qu 5 abu. Since p
is irreducible, either a or bu is a unit. Therefore, a or b is a unit. Thus, q
is irreducible.
Worked Exercise 12.1.2. Prove that an integral domain is a field if and only
if there are exactly two associate classes.
EXERCISE 12(a)
1. State whether the following are true and justify your answers.
(i) 5 is an irreducible element in Z.
(ii) 10 is an irreducible element in Z[i].
(iii) 13 is a prime element in Z[i].
(iv) Any prime element in Z is a prime element in Z[i].
(v) 2 2 is an associate of 2 in Z[ 2].
(vi) 2 is irreducible in R.
(vii) 5 is a prime element in R.
(viii) 25 is a prime element in Z.
3. Determine all the associates of each of the following in the rings mentioned
against them
(i) 4 in Z
(ii) 1 1 x in R[x]
(iii) 1 1 i in Z[i]
(iv) 2 1 x in Z5[x]
(v) 1 1 2x in Z3[x]
(vi) 1 1 x 1 x2 in Z2[x].
7. Let a and b Z such that |a2 2 10b2| is prime in Z. Then prove that a b 10
is irreducible in Z[ 10].
Example 12.2.1
1. Z is a PID.
2. Any field is a PID, since <0> and <1> are the only ideals of a field.
3. The ring F[x] of polynomials over a field F is a PID (refer Theorem
11.3.2).
4. The ring Z[x] of polynomials over Z is an integral domain, but not a
PID; for, we have exhibited, in Worked Exercise 11.3.1 an ideal of Z[x]
which is not principal.
The following is an immediate consequence of Theorems 12.1.3 and
12.1.4.
Theorem 12.2.1. In a PID, any finite number of nonzero elements have both
g.c.d. and l.c.m.
This identity is known as Bezout’s identity. The following is one of the most
useful applications of Bezout’s identity.
Theorem 12.2.2. Let R be a PID and a, b and c R such that a and b are
relatively prime and a divides bc. Then, a divides c.
Proof: Since a and b are relatively prime, there exist r and s R such that
ar 1 bs 5 1
Theorem 12.2.3. The following are equivalent to each other for any nonzero
element p in a PID R.
1. p is a prime element.
2. p is an irreducible element.
3. <p> is a maximal ideal of R.
4. <p> is a prime ideal of R.
Proof: First observe that, to satisfy any of the conditions (1) through (4)
above, it is necessary that p is a nonunit in R. (1) ⇒ (2) follows from Theorem
12.1.5.
Proof:
1. Let I1 ⊆ I2 ⊆ I3 ⊆ … be an ascending sequence of ideals of R. Put
I ∪ I n . Then, I is an ideal of R. Since R is a PID, there exists a R
n1
such that
∞
∪I
n1
n I a .
for all k Z1, so that In 5 In1k for all k Z1. Thus, the sequence termi-
nates at a finite stage.
2. Let # be a nonempty class of ideals of R and suppose, if possible, that #
has no maximal member. Since # is nonempty, choose I1 # Then, I1 is
not maximal in # and hence there exists I2 # such that I1 I2. Again,
since I2 is not maximal, there exists I3 # such that I2 I3. Continuing
this procedure, we get an ascending sequence
I1 I2 I3 …
of ideals of R which does not terminate at any finite stage. This is a con-
tradiction to (1) above. Thus, # must contain a maximal member. b
Proof: Since a is a nonunit, the principal ideal <a> is a proper ideal and
hence <a> is contained in a maximal ideal M of R. Since R is a PID, there
exists p R such that M 5 <p>. Also, since a 0. We have
a 5 p1p2 … pn21pnan
Worked Exercise 12.2.1. Prove that the ring Z[i] of Gaussian integers is
a PID.
A 5 {a2 1 b2 : 0 a 1 bi I}.
y (c di )( a bi ) ( ac bd ) ( ad bc)i
a bi,
x ( a bi )( a bi ) a2 b2
ac bd ad bc
a and b .
a2 b2 a2 b2
1 1
| ma | and | n b | .
2 2
Also,
y 5 (m 1 ni)x <x>.
Worked Exercise 12.2.2. Let R be a PID. Then prove that any nonzero proper
ideal of R can be expressed as a finite product of maximal ideals of R.
a 5 p1p2pn,
where p1, p2, …, pn are prime elements. Put Mi 5 <pi> for each 1 i n. By
Theorem 12.2.3, each Mi is a maximal ideal of R. Now,
I 5 <a> 5 <p1p2pn>
5 <p1><p2><pn>
5 M1M2Mn.
Worked Exercise 12.2.3. Determine all the units of the ring Z[i] of Gaussian
integers.
Answer: Let a 1 bi be a unit in Z[i]. Then, there exist integers c and d such that
(a 1 bi)(c 1 di) 5 1.
a2 1 b2 5 1 5 c2 1 d2
Worked Exercise 12.2.4. Let I be a nonzero ideal of the ring Z[i]. Then prove
that the quotient ring Z[i]/I is finite.
Answer: Since Z[i] is a PID, I is a nonzero principal ideal and hence I 5 <x>
for some 0 x Z[i]. Let x 5 a 1 bi. Then, a 0 or b 0 and hence a2 1 b2
is a positive integer. Consider an element y 1 I Z[i ] / I with y Z[i]. As
in Worked Exercise 12.2.1, we can write
y 5 (m 1 ni)x 1 r
is finite.
EXERCISE 12(b)
a 5 p1p2 … pn,
Examples 12.3.1
1. 6 5 2 ? 3 and 6 5 (22)(23) are factorizations of 6 in Z, since 2, 3, 22
and 23 are irreducible in Z.
2. 20 5 2 ? 2 ? 5, 20 5 (22)2(25) and 20 5 2(22)(25) are factorizations
of 20 in Z.
3. 1 1 2x 1 x2 5 (1 1 x)(1 1 x) is a factorization of 1 1 2x 1 x2 in Z[x],
since 1 1 x is irreducible in Z[x].
4. 1 1 x2 5 (1 1 x)(1 1 x) is a factorization of 1 1 x2 in Z2[x], since 1 1 x
is an irreducible element in Z2[x].
Examples 12.3.2
1. Z is a FD.
2. Any PID is a FD (recall Theorem 12.2.6). Note that an element in a PID
is prime if and only if it is irreducible.
3. The ring Z[i] of Gaussian integers is a FD, since it is a PID (see Worked
Exercise 12.2.1).
4. The ring F[x] of polynomials over a field F is a FD, since F[x] is a PID.
In the following, we give sufficient condition on an integral domain for it
to be a FD. This helps us as a tool to quickly ascertain that a given integral
domain is a FD.
A 5 {(a) : a S}.
Proof: The map a |a| from Z 2 {0} into Z satisfies the properties men-
tioned in Theorem 12.3.1 and hence Z is a FD. b
Then, satisfies (1), (2) and (3) of Theorem 12.3.1 and hence Z[i] is a FD. b
{
Z n a b n : a and b ∈ .
}
Then, Z [ n ] is a FD.
Proof: Define : Z [ n ] → Z by
( a b n ) a 2 nb 2 .
Proof: We know that every prime element in any integral domain (and hence
in R) is irreducible. Conversely, suppose that p is an irreducible element in R.
Let b and c R such that p|bc. We can assume that b and c are both nonzero
(since p|0). If b is a unit, then p|c. Similarly, if c is a unit, then p|b. Suppose that
neither b nor c is a unit. Since p|bc, there exists a R such that pa 5 bc. Then,
a 0 (since b 0 and c 0) and a is nonunit (otherwise, if a is a unit, then
p 5 b(ca21) and, since p is irreducible, b or ca21 is a unit which is not true).
Thus, a, b and c are nonzero nonunits in R. Since R is a UFD, we get that
where pi’s, qi’s and ri’s are irreducible elements in R. Therefore, we have
Theorem 12.3.3. Let R be a FD. Then, R is an UFD if and only if every irre-
ducible element in R is prime.
p1p2pn 5 q1q2qm.
p1 p2 pn qs (1) ( ∏ q ) = p u ( ∏ q ).
j ≠ s (1)
j
1 1
j ≠ s (1)
j
p2 pn u1 ( ∏ q ).
js (1)
j
We can repeat the above process, with p2 in place of p1, to get s(2) {1, 2, …,
m} 2 {q(1)} such that p2u2 5 qs(2) for some unit u2 in R. Then,
p2 pn u1 p2 u2 ∏
j ≠ s (1), s ( 2 )
qj
This process can be continued for m steps (since n > m) to exhaust all qj’s and
then we get
pm11pn 5 u1u2um,
where u1, u2, …, um are units and hence their product u1u2um is also a unit.
Now, pm11 divides the unit u1u2um and hence pm11 itself is a unit which
is a contradiction to the fact that an irreducible element is necessarily a
nonunit. Thus, n 5 m and we have permutation s on {1, 2, …, n} such
that pi ~ qs(i) for all 1 i n. Thus, R is an UFD, converse is proved in
Theorem 12.3.2. b
Proof: We have proved in Theorem 11.3.2 that, for any field F, F[x] is a PID
and hence, by the above corollary, F[x] is an UFD. b
Unique factorizations in an UFD help us in determining the g.c.d. and
l.c.m. of two elements. In this direction, we have the following theorem whose
proof is a routine verification.
r r
m : m ∑ ai and ∏ pi I ,
ai
A i1 i1 .
where p1 , …, pr are irreducible in R
Since I has atleast one nonzero nonunit, if follows that A is a nonempty set of
positive integers. Let m be the least member in A. Then, there exist irreducible
elements p1, p2, …, pr in R and positive integers a1, a2, …, ar such that
r r
d ∏ piai I and m ∑ ai .
i1 i1
Now, we claim that I 5 <d>. Since d I, we have <d> ⊆ I. On the other hand,
suppose that 0 x I. We can assume that each pi is not an associate of any
a
other pj, j i. It is enough if we can prove that each pi i divides x. Suppose
a1
that p1 does not divide x. Then, we have
z 5 z1 5 zay 1 zbp1
p1n y( ap1a11n p2a2 … prar ) db
x( ap1a11n p2a2 … prar ) db I
⇔ (a 1 b 3 )(a 2 b 3 ) 5 1
⇔ a 1 b 3 is a unit.
where p1, p2, …, pr are irreducible elements in R and a1, a2, …, ar are positive
integers. Since x P and P is a prime ideal, pi P for some i.
EXERCISE 12(c)
2. Prove that the l.c.m. of any finite subset of an UFD exists and is unique up to
associates.
3. Let R be an UFD and F be its field of quotients. For any prime p in R, let
a
R( p ) F : g.c.d.{a, b} 1 and p b .
b
Prove that R(p) is a subring of F and that R(p) is a PID and hence an UFD.
4. Let S be a multiplicative set in an UFD R and S21R the ring of fractions of R by
S. Prove that S21R is an UFD.
5. Let R be an UFD and {an} be a sequence of elements in R such that an11 divides
an for all n Z1. Then prove that there exists n Z1 such that an is an associate
of an1k for all k Z1.
6. Let p be a prime element in an UFD R such that any prime element in R is an
associate of p. Prove that every nonzero proper ideal of R is of the form <pn> for
some n Z1.
7. Let R be an UFD and P be the only nonzero prime ideal of R. Then prove that
any nonzero proper ideal I of R is of the form Pn for some n Z.
8. Prove that any increasing sequence of principal ideals in an UFD terminates at a
finite stage.
9. If P is a nonzero prime ideal of an UFD R, then is R/P a UFD?
10. Let R be an UFD and (a, b) denote the g.c.d. {a, b} for any a and b R. Prove
the following for any nonzero elements a, b and c R.
(i) (a, (b, c)) ~ ((a, b), c)
(ii) (a, 1) ~ 1
(iii) (ca, cb) ~ c(a, b)
c(f (x)) 5 1,
Note 12.4.1
1. Any monic polynomial over any UFD is primitive.
2. If R is an UFD and 0 f (x) R[x], then
f (x) 5 c(f (x))g(x)
In the following, we prove that primitive polynomials over any UFD are
closed under multiplication.
Theorem 12.4.1. Let R be an UFD and f (x) and g(x) be primitive polynomials
over R. Then, f (x)g(x) is primitive.
If deg(f (x)) 5 n 5 0, then c(f (x)) 5 a0 and hence a0(5 f (x)) is a unit so that
c(f (x)g(x)) ~ c(g(x)) ~ 1. Therefore, we can assume that n > 0 and m > 0.
We prove that there is no prime element in R that divides all the coefficients
in f (x)g(x). To prove this, let p be a prime in R. Since f (x) is primitive, p does
not divide some coefficient ai in f (x). Let ai be the first coefficient in f (x) which
is not divisible by p. Similarly, let bj be the first coefficient in g(x) which is not
divisible by p. Let ci1j be the coefficient of xi1j in f (x)g(x). Then,
Since p|ar for all 0 r < i and p|bs for all 0 s < j, it follows that ci1j 2 arbs
is divisible by p. Since p does not divide aibj, we get that p does not divide ci1j.
Therefore, there is no prime dividing all the coefficients in f (x)g(x) and hence
Corollary 12.4.1. For any nonzero polynomials f (x) and g(x) over an UFD,
c(f (x)g(x)) 5 c(f (x))c(g(x)).
Proof: Let R be an UFD and f (x) and g(x) R[x] 2 {0}. Then, there exist
primitive polynomials f1(x) and g1(x) in R[x] such that
a0 a1 a
f ( x) x n x n
b0 b1 bn
Proof: Suppose that f (x) is irreducible in F[x]. Suppose that f (x) 5 g(x)h(x)
where g(x) and h(x) R[x]. If g(x) and h(x) are both of positive degree then
they are nonunits in R[x] and in F[x], which is a contradiction to the irreduc-
ibility of f (x) in F[x]. Therefore, g(x) or h(x) is of degree 0. Let deg(g(x)) 5 0.
Then, g(x) R. Since f (x) is primitive in R[x], we get that g(x) is a unit in R.
Similarly, if deg(h(x)) 5 0, then h(x) is a unit in R. Thus, f (x) is irreducible
in R[x].
Conversely, suppose that f (x) is reducible in F[x]. Then, f (x) 5 g(x)h(x) where
g(x) and h(x) are nonunits in F[x] and hence g(x) and h(x) are polynomials of
positive degree (since F is a field). Let
a0 a1 a
g( x) x n x n , an 0, n 0
b0 b1 bn
c0 c1 c
and h( x ) x m x m , cm 0, m 0
d0 d1 dm
for some primitive polynomials g1(x) and h1(x) in R[x]. Now, we have
and hence f (x) 5 ug1(x)h1(x), which implies that f (x) is reducible in R[x]. b
g(x) 5 g1(x)g2(x)gn(x)
for some irreducible polynomials g1(x), g2(x), …, gn(x) in F[x]. Then, deg(gi(x))
> 0 for each 1 i n. We can write gi ( x ) ai bi1 hi ( x ) where ai, bi R
(bi 0) and hi(x) is primitive in R[x]. Then, we have
b1b2bng(x) 5 a1a2anh1(x)h2(x)hn(x).
Since each hi(x) is primitive, so is their product. Also, since g(x) is primitive,
it follows, by taking contents both sides that
b1b2bn ~ a1a2an in R
If c(f (x)) is a unit in R, c(f (x))uh1(x), h2(x), …, hn(x) are irreducible elements
in R[x] and hence we have a factorization of f (x) in R[x]. If c(f (x)) is not a unit
in R, then, by the factorization property in R,
where p1, p2, …, pm are irreducible elements in R and hence in R[x]. Then,
f (x) 5 p1p2pm21(pmu)h1(x)h2(x)hn(x)
is another factorization of f (x) in R[x]. Since each of gi(x) and hi(x) are
irreducible polynomials of positive degree in R[x], they are primitive and
hence these are irreducible in F[x]. Also, by taking contents in (1) and (2),
we get that
h1(x)h2(x)hs(x) 5 ug1(x)g2(x)gm(x).
Corollary 12.4.2. If R is an UFD, then so is R[x1, x2, …, xn], where R[x1, x2,
…, xn] 5 R[x1, …, xn21][xn].
Proof: First, we assume that f (x) is primitive and prove that f (x) is irreduc-
ible in R[x]. Suppose, if possible, f (x) is not irreducible. Since f (x) is a non-
constant primitive polynomial, there exist two nonzero nonunit polynomials
f1(x) and f2(x) in R[x] such that
f (x) 5 f1(x)f2(x).
Similarly, s 0. Since n 5 r 1 s, it follows that 0 < r < n and 0 < s < n. Now,
f (x) 5 f1(x)f2(x) implies that a0 5 c0d0. Since p|a0 and p2a0, it follows that p
divides exactly one of c0 and d0. Without loss of generality, we can assume
that p|c0 and pd0. Also, we have an 5 crds and pan and hence pcr and pds.
Therefore, we have p|c0 and pcr. Let i be the least positive integer such that
pci. Then, p|cj for all 0 j < i. Equating the coefficients of xi in f (x) and f1(x)
f2(x), we get that
where cj 5 0 for all j > r and dj 5 0 for all j > s. Now, since p|c0, p|c1, …,
p|ci21, we get that
p | ai 2 cid0.
Also, since 0 < i r < n, p|ai, by (1). Therefore, p|cid0 which is a contradiction
since p is prime, pci and pd0.
Thus, f (x) is irreducible in R[x]. Also, since f (x) is primitive, the Gauss
Lemma (Theorem 12.4.2) implies that f (x) is irreducible in F[x].
Now, let us take up the general case. There exists a primitive polynomial g(x)
in R[x] such that
where c(f (x)) is the content of f (x). If c(f (x)) is a unit, then f (x) is primitive
and hence, by the first case, f (x) is irreducible in R[x] and in F[x]. Suppose
that c(f (x)) is a nonunit. Let
Then, deg(f (x)) 5 deg(g(x)) and hence bn 0. Since f (x) 5 dg(x), we get
that
ai 5 dbi for all 0 i n.
Since pan, we get that pd and pbn. Also, for any 1 i < n, p|ai implies that
p|bi. Further, p2b0, since p2a0. Thus, by the first case, g(x) is irreducible in
R[x] and hence in F[x]. Since d is a unit in F, we get that f (x) is an associate
of g(x) in F[x]. Thus, f (x) is irreducible in F[x]. b
Since Z is a UFD and the rational number field is the field of quo-
tients of Z, the following is an immediate consequence of the Eisenstein’s
Criterion.
x p 1
f ( x ) 1 x x 2 x p1 .
x 1
(x 1) p 1
Therefore, f (x 1)
(x 1) 1
1
( x p px p1 (p2 )x p2 px )
x
p1
∑ ( pr ) x pr 1
r0
and hence
n
a0 q n p ∑ ar p r q nr
r1
nn1
aan0 pq n qp aarr pprrqqnnrr.
and ∑ 01
rr
n1
an p n q ∑ ar p r q nr .
r0
Therefore, p divides a0qn and q divides anpn. Since p and q are relatively
prime, it follows that p divides a0 and q divides an. b
Worked Exercise 12.4.1. Let f (x) 5 5 1 11x 2 7x2 1 9x3 Z[x]. Prove that
f (x) is irreducible over Q as well as over Z.
Answer: Since deg(f (x)) 5 3, f (x) is irreducible over Q if and only if f (x)
has a root in Q. Suppose that p/q is a root of f (x), we can assume that p and q
are relatively prime integers and q > 0. Then, by the above Theorem 12.4.5, p
should divide 5 and q should divide 9. Therefore, p 5 1 or 5 and q 5
1 or 3 or 9 and hence
p
q
{ 1 5 1 5
1, 5, , , , .
3 3 9 9
}
But, none of the elements in this set is a root of f (x). Therefore, f (x) has no root
in Q. Thus, f (x) is irreducible over Q. Since f (x) is primitive in Z[x], it follows
from Gauss Lemma (Theorem 12.4.2), that f (x) is irreducible over Z also.
Worked Exercise 12.4.2. Let f (x) 5 22 1 15x 2 9x2 1 x3 Z[x]. Prove that
f (x) is irreducible over neither Q nor Z.
p
±1 or ± 2.
q
By a physical verification, we can see that 2 is a root of f (x); that is, f (2) 5 0.
Thus, x 2 2 is a factor of f (x); in fact,
Exercise 12(d)
3. Let R be an UFD and f (x) R[x]. Prove that f (x) is irreducible over R if and only
if f (x 1 a) is irreducible over R for some a R.
4. Which of the following polynomials are irreducible over the UFDs mentioned
against them?
(i) 15 2 9x2 1 6x3 1 2x4 over Z
(ii) 3 1 2x2 1 x3 over Q
(iii) 4 1 2x 1 x3 over Z5
(iv) 1 1 x2 1 x5 over Z2
(v) 9 2 x3 over Z31
(vi) 1 1 x3 1 x6 over Q
(vii) 5 1 10x 1 15x3 1 2x5 over Z
(viii) 2 1 2x 1 x4 over Q
(ix) 9 2 x3 over Z11
(x) 14 2 7x 1 10x4 over Q.
6. For any prime number p, prove that p 2 xn is irreducible over Q for any positive
integer n.
7. Prove that 1 1 x4 is irreducible over Q and reducible over Zp for any prime
number p.
8. Determine all irreducible polynomials of degree 2 in Z2[x].
9. Give an example of a polynomial which is irreducible over Z but not irreducible
over Z2.
10. For any prime p, prove that there are exactly (p(p 2 1))/2 irreducible monic
polynomials of degree 2 in Zp[x].
11. Let p be a prime number and
f (x) 5 1 2 x 1 x2 2 x3 1 … 1 (21)p21xp21.
Then prove that f (x) is irreducible over Z.
12. Let R be an UFD and F be its field of quotients. Prove Theorem 12.4.5 with R
and F in place of Z and Q, respectively.
Example 12.5.1
1. The ring Z of integers is an Euclidean domain, in which the gauge func-
tion is defined by
g(a) 5 |a|, the absolute value of a,
for any a Z 2 {0}. Clearly, |ab| 5 |a||b|. To prove the division algo-
rithm, let a and b Z and b 0. Without loss of generality, we can
assume that b > 0. Let q be the integral part of the rational number a/b;
that is, q is an integer such that
a a
q and q 1 .
b b
Proof:
1. follows from the facts that g(1) is a positive integer and g(1) 5 g(1 ? 1)
5 g(1)g(1).
2. Let 0 a R. If a is a unit in R, then there exists b R such that ab 5 1.
Then, b 0 and
1 5 g(1) 5 g(ab) 5 g(a)g(b).
Since g(a) and g(b) are positive integers, we get that g(a) 5 1 5 g(b).
Conversely, suppose that g(a) 5 1. By the Euclidean division algorithm,
there exist q and r R such that
1 5 qa 1 r, where r 5 0 or g(r) < g(a) 5 1.
If r 0, then g(r) Z1 which is not true, since g(r) < 1. Therefore,
necessarily r 5 0 and 1 5 qa. Thus, a is a unit in R.
A 5 {g(a) : 0 a I}.
Proof: This follows from the fact that every PID is an UFD and from the
above theorem. b
Also, for any f (x) and h(x) F[x] with h(x) 0, by the division algorithm for
polynomials, there exist q(x) and r(x) F[x] such that
where r(x) 5 0 or deg(r(x)) < deg(h(x)) and hence r(x) 5 0 or g(r(x)) <
g(h(x)). Thus, F[x] is an Euclidean domain. b
The following is a generalization of the well-known algorithm to find the
g.c.d. of any two positive integers.
Then, there exists n such that rn11 5 0 and rn21 5 qn11rn. If n is the least such
integer, then
Proof: First observe that, since g(a) > g(r1) > g(r2) > …, the above process
of getting qn’s and rn’s should terminate (at most after g(a) number of steps)
and hence rn11 5 0 for some n. Writing from bottom to top of the above equa-
tions, we have
rn21 5 qn11rn
rn22 5 qnrn21 1 rn, 0 rn 5 rn22 2 qnrn21
rn23 5 qn21rn22 1 rn21, 0 rn21 5 rn23 2 qn21rn22
r1 5 q3r2 1 r3, 0 r3 5 r1 2 q3r2
a 5 q2r1 1 r2, 0 r2 5 a 2 q2r1
b 5 q1a 1 r1, 0 r1 5 b 2 q1a.
Tracing from top to bottom of the left hand side equations, we get that
rn|rn21, rn|rn22, …, rn|a and rn|b. Therefore, rn is a common divisor of a and b.
Also, if d is any common divisor of a and b, then tracing from bottom to top
of the equations on the right hand side above, we get that
aR 1 bR 5 <d> 5 dR.
g : Z[ 2]{0} → Z by g ( a b 2) | a 2 2b 2 | .
g ( xy ) g (( a b 2) (c d 2))
g ( ac 2bd ( ad bc) 2)
| ( ac 2bd ) 2 2( ad bc) 2 |
| a 2c 2 4b 2 d 2 4 acbd 2a 2 d 2 2b 2c 2 4 adbc |
| ( a 2 2b 2 ) || (c 2 2d 2 ) |
g ( x ) g ( y ).
x ( a b 2) (c d 2)
a b 2, say
y c 2 2d 2
where a and b are rational numbers. Choose integers m and n such that
1 1
| m a| and | n b| .
2 2
g ( r ) | (a m) 2 2( b n) 2 || c 2 2d 2 |
1 2
| c 2 2d 2 || c 2 2d 2 | g ( y ).
2 4
Answer: We know that R[x] is an Euclidean domain. Let us follow the algo-
rithm given in Theorem 12.5.4.
Let f (x) 5 1 1 x 1 x2 and g(x) 5 1 1 2x 1 3x3 1 x5. Then,
Theorem 12.5.5. We have the following inclusions among the above classes
F ⊂ ED ⊂ PID ⊂ UFD ⊂ FD ⊂ ID
Proof:
1. In Example 12.5.1 (2), we have proved that every field is an Euclidean
domain and therefore F ⊆ ED.
Z is an Euclidean domain, but not a field and hence F ED.
2. In Theorem 12.5.2, we have proved that every Euclidean domain is a
PID and therefore ED ⊆ PID. Consider the ring R given by
{ b
R a (1 i 19 ) : a and b Z .
2
}
Then, R is a PID, but not an Euclidean domain (the proof of this is little
bit involved and hence we skip the proof ). Therefore, ED PID.
3. In Corollary 12.3.4, we have proved that every PID is an UFD and hence
PID ⊆ UFD. The ring Z[x] is an UFD but not a PID (see Theorem 12.4.3).
Therefore, PID UFD.
ID
FD
UFD
PID
ED
F
EXERCISE 12(e)
1. Let R be an Euclidean domain with gauge function g. For any a and b R 2 {0},
prove that a and b are associates if and only if a divides b and f(a) 5 f(b).
3. Prove that Q[x], R[x] and C[x] are all Euclidean domains.
4. If R is an integral domain which is not a field, prove that R[x] is not an Euclidean
domain.
6. In any integral domain with gauge function g, prove that g(a) 5 g(2a) for any
nonzero element a.
7. Prove that, for any n Z1, g : Z 2 {0} → Z1 defined by g(a) 5 |a|n is a gauge
function on Z.
10. Let x 102 10 3 and y 1 7 3. Find q and r in Z[ 3] such that x 5 yq 1 r
where either r 5 0 or r a b 3 with |a2 2 3b2| < 146.
for some rational numbers a and b. Now, choose integers m and n such that
1 1
|am | and |bn |
2 2
r 5 x 2 (m 1 ni)y Z[i].
1 1
g ( y ) g ( y).
4 4
Theorem 12.6.2. Let p Z1 be prime and n Z such that p does not divide
n. Suppose that we can find integers x and y such that np 5 x2 1 y2. Then, p
can be expressed as a sum of two squares of integers; that is, p 5 a2 1 b2 for
some a and b Z.
Proof: First observe that any integer m can be treated as a Gaussian inte-
ger m 1 0i and that Z is a subring of Z[i]. Next, we prove that p cannot
be a prime element in Z[i]. Suppose, if possible, that p is prime in Z[i].
Since
pn 5 x2 1 y2 5 (x 1 iy)(x 2 iy),
p(s 1 it) 5 x 1 iy
for some s and t Z and hence ps 5 x and pt 5 y, so that p(s 2 it) 5 x 2 iy.
Therefore,
from which it follows that p2 divides pn and hence p divides n, which is a con-
tradiction to our hypothesis. Therefore, p is not a prime in Z[i]. Since Z[i] is
an UFD, an element in Z[i] is prime if and only if it is irreducible. Therefore,
p is not an irreducible element in Z[i]. Therefore, there exists two nonunits
a 1 bi and c 1 di in Z[i] such that
p 5 (a 1 bi)(c 1 di)
and hence p2 5 g(p) 5 g(a 1 bi)g(c 1 di),
x2 ≡ 21 (mod p).
p 1 p 1
x 1 2 3 (1)(2)(3)
2 2
p 1 p 1
x 2 1 2 3 (1)(2)(3)
2 2
p 1 p 1
1 2 3 ( p 1)( p 2)( p 3) p (mod p)
2 2
p 1 p 1 p 3
1 2 3 ⋅ ⋅ ( p 2)( p 1)
2 2 2
( p 1)! 1 (mod p).
x 5 qp 1 r, where 0 r < p.
Worked Exercise 12.6.1. Prove that any integer of the form 4n 1 3 cannot be
expressed as the sum of two perfect squares.
4n 1 3 5 a2 1 b2
Worked Exercise 12.6.2. Let I be a nonzero ideal of the ring Z[i] of Gaussian
integers. Then prove that the quotient ring Z[i ] / I is finite.
Proof: By Corollary 12.6.1, Z[i] is a PID and hence I 5 <x> for some x
Z[i]. Let x 5 a 1 bi with a and b Z. Since I is nonzero, we have x 0
and hence a 0 or b 0, so that g(x) 5 a2 1 b2 > 0, where g is the gauge
function on Z[i]. Any element of Z[i ] / I is of the form y 1 I, y Z[i]. Now,
Exercise 12(f )
1. In each of the following two elements x and y of Z[i] are given. Find q and r in
Z[i] such that x 5 qy 1 r; with r 5 0 or g(r) < g(y).
(i) x 5 3 1 2i and y 5 2 2 3i
(ii) x 5 5 and y 5 2i
(iii) x 5 1 1 i and y 5 2 1 i
(iv) x 5 2 1 3i and y 5 1 2 i
(v) x 5 4 2 5i and y 5 5 1 4i.
10. Find x and y Z[i] such that g.c.d. {25 1 10i, 3 1 i} 5 (25 1 10i)x 1
(3 1 i)y.
Example 13.1.1
1. Let (M, 1) be any abelian group and R 5 End(M), the ring of all endo-
morphisms of (M, 1), then M is a left R-module, where the map (f, x)
fx of R 3 M into M is defined simply by fx 5 f (x), the image of x under f.
2. Let (M, 1) be any abelian group and consider the ring Z of integers. For
any n Z and x M, define
0 if n 0
nx ( n 1) x x if n 0.
(n)(x ) if n 0
Theorem 13.1.1. Let M be a left R-module. Then, the following holds for any
x M and a R:
1. 0x 5 0, where 0 on the left is the zero element in the ring R and 0 on the
right is the zero element (identity element) in the group (M, 1).
2. a0 5 0, where 0 is the zero element in M.
3. (2a)x 5 2(ax) 5 a(2x).
Proof:
1. Since 0x 1 0 5 0x 5 (0 1 0)x 5 0x 1 0x, it follows that 0x 5 0.
2. Also, since a0 1 0 5 a0 5 a(0 1 0) 5 a0 1 a0, we get that a0 5 0.
3. We have 0 5 0x 5 (a 2 a)x 5 ax 1 (2a)x and hence (2a)x 5 2(ax). Also,
0 5 a0 5 a(x 1(2x)) 5 ax 1 a(2x)
and therefore a(2x) 5 2(ax).
Note that we are using the same symbol 0 to denote the additive identity (zero
element) in the ring as well as the identity element in the group M. This need
not create any ambiguity and we should take it as per the context. Further,
throughout our discussions in this chapter, all modules are assumed to be left
R-modules, unless otherwise stated. Further, a module over R will be denoted
by (M, 1, R) or simply by M when there is no ambiguity about the group
operation 1 on M and about the ring R over which M is a module.
Theorem 13.1.2. Let R be a ring with unity and M be an R-module. Then, the
following are equivalent to each other for any subset N of M:
1. N is an R-submodule of M.
2. N [ and ax 2 by N for all a, b R and x, y N.
3. 0 N and ax 2 by N for all a, b R and x, y N.
Example 13.1.2
1. As in Example 13.1.1 (3), any ring R can be treated as an R-module and
the R-submodules of R are precisely the left ideals of R.
2. Let R be any ring and R[x] be the set of all polynomials over R. Then,
R[x] is an R-module under the usual addition and multiplication of
polynomials (recall that elements of R can be treated as polynomials of
degree zero). For any n $ 0, let
Rn[x] 5 {f (x) R[x] : f (x) 5 0 or deg f (x) # n}.
Thus, N is a submodule of M.
However, as usual, the union of submodules may not be a submodule. In
fact, as in the case of subgroups of a group, for any submodules N and K of an
R-module M, N ∪ K is a submodule of M if and only if either N ⊆ K or K ⊆ N.
∪N
i∈ I
i {
x1 x2 xn : x j N i ,
j
}
i1 , … , in I .
{
Proof: Let N x1 x2 x n : x j N i , i1 , i2 , …, in I .
j
}
We shall prove that N is the smallest R-submodule of M containing i ∪
I
Ni .
Clearly, Ni ⊆ N for all i I and hence i ∪ N i ⊆ N . Also,
I
N ∪N
i∈ I
i .
N 1 K 5 {x 1 y : x N and y K}.
n
x1 , x2 , …, xn x1 x2 xn ∑ xi .
i1
M is said to be finitely generated if M 5 <x1, x2, …, xn> for some x1, x2, …,
xn M. The elements x1, x2, …, xn are said to be generators of M and the set
X 5 {x1, x2, …, xn} is said to generate M.
M 5 Rx 5 {ax : a R}
for some x M.
Note that the set of generators of a module need not be unique. For exam-
ple, the Z-module Z is cyclic, since Z 5 <1> and also Z 5 <2, 3>. In fact,
we have following exercise.
xi p1 p2 pi1 pi1 pn ∏p
j ≠i
j
and X 5 {x1, x2, …, xn}. Then, x1, x2, …, xn are relatively prime (since there is
no prime dividing all xi’s).
Hence, there exist integers y1, y2, …, yn such that
Thus, Z 5 <x1, x2, …, xn>. Further, for each 1# i # n, pi divides all xj, j i
and, in fact
g.c.d. {xj : j i} 5 pi
Worked Exercise 13.1.2. Let F be any field and n be a positive integer. Let M
be the set of all polynomials over F of degree less than n. Then, prove that M
is a finitely generated F-module and exhibit two distinct sets of generators of
M, each with n elements.
and therefore {1, x, x2, …, xn21} is a generating set for M and has n elements.
Also, for any a 0 in F,
EXERCISE 13(a)
1. State whether each of the following is true or false and justify your answer:
(i) Any ring R is a finitely generated R-module.
(ii) Any ring R with unity is a finitely generated R-module.
(iii) If R is a finite ring, then any R-module is finitely generated.
(iv) Any finite R-module is finitely generated.
(v) Any module over a finite ring is finite.
(vi) Any finite R-module has only a finite number of R-submodules.
(vii) Any finitely generated R-module has only a finite number of R-submodules.
(viii) If an R-module M is cyclic and M 5 <x>, then M 5 <ax> for each 0
a R.
(ix) If a is a unit in a ring R and M is an R-module, then <x> 5 <ax> for
each x M.
(x) If M is a finite R-module, then R is finite.
2. Consider the real member field R and the R-module R4. Which of the following
are R-submodules of R4?
(i) {(a1, a2, a3, a4) R4 : a2 5 0}
(ii) {(a1, a2, a3, a4) R4 : a1 5 a2 5 a3 5 a4}
(iii) {(a1, a2, a3, a4) : a1 1 a2 5 0}
(iv) {(a1, a2, a3, a4) : a1 5 a2}
is an ideal of R.
5. Let R be a ring and M be the set of all mappings of R into itself. Prove that M is
an R-module under the operations defined by
is an R-submodule of M.
Proof:
1. Since f (0) 5 0, 0 ker f. If a and b R and x and y ker f, then f (x)
5 0 5 f (y),
f (x 2 y) 5 f (x) 2 f (y) 5 0 2 0 5 0
and f (ax) 5 af (x) 5 a0 5 0.
Example 13.2.1
1. Let R be any ring and M and N be R-modules. Define f : M → N by f (x)
5 0 for all x M. Then, f is an R-homomorphism and is called the zero
homomorphism. Note that ker f 5 M and f (M) 5 {0}.
Then, fa is an R-endomorphism of M.
Fundamental theorem of homomorphisms and other isomorphism theo-
rems are analogous to those of groups and rings. Before going to these, we
formally define the notion of quotient module.
x 1 N 5 {x 1 s : s N}.
M
{x N : x ∈ M }
N
(x 1 N) 1 (y 1 N) 5 (x 1 y) 1 N
and a (x 1 N) 5 ax 1 N
on M/N is well defined and (M/N, 1) is an abelian group. Also, for any
x and y M and a R,
x1N5y1N⇒x2yN
⇒ ax 2 ay 5 a(x 2 y) N
⇒ ax 1 N 5 ay 1 N
and therefore the scalar multiplication defined in M/N is also well defined. It
is routine to verify all the axioms of an R-module for M/N. Thus, M/N is an
R-module.
Definition 13.2.4. The R-module M/N defined above is called the quotient
R-module (or simply quotient module) of M by N.
The proofs of the following two theorems are similar to those in groups
and rings.
M
> f ( M ).
ker f
Worked Exercise 13.2.1. For any R-module M, prove that EndR(M) is a ring
with unity under the point-wise addition and the composition of mappings as
multiplication.
(f 1 g)(x 1 y) 5 f (x 1 y) 1 g(x 1 y)
5 f (x) 1 f (y) 1 g(x) 1 g(y)
5 f (x) 1 g(x) 1 f (y) 1 g(y)
5 (f 1 g)(x) 1 (f 1 g)(y)
f ο (g 1 h) 5 f ο g 1 f ο h
and (f 1 g) ο h 5 f ο h 1 g ο h
( M A) > M .
( B A) B
Proof: By Theorem 13.2.5, B/A is an R-submodule of M/A and hence we can
form the quotient module (M/A)/(B/A). Define f : M/A → M/B by f (x 1 A) 5
x 1 B for any x 1 A M/A, x M. For any x and y M,
x1A5y1A⇒x2yA⊆B
⇒x2yB
⇒ x 1 B 5 y 1 B.
M
M
ker f x A ∈ : f ( x A) The zero in
A B
M
x A ∈ : x B B
A
M
x A ∈ : x ∈ B
A
B
.
A
( M A) > M .
( B A) B
Worked Exercise 13.2.3. For any R-submodules A and B of an R-module M,
prove that ( A B) / A > B /( A ∩ B).
Proof: Clearly, A 1 B is an R-module (being an R-submodule of M) and A
is an R-submodule of A 1 B. Also, A ∩ B is an R-submodule of B. Define
f : B → ( A B) / A by f (b) 5 b 1 A for any b B.
Since B ⊆ A 1 B, f is well defined. Clearly, f is an R-homomorphism. Also,
for any x 5 a 1 b A 1 B with a A and b B, we have
x2b5aA
and hence x 1 A 5 b 1 A 5 f (b), b B. Therefore, f is an R-epimorphism.
Further
A B
ker f b ∈ B : f (b) zero in
A
5 {b B : b 1 A 5 A}
5 {b B : b A}
5 A ∩ B.
B B A B
> .
A ∩ B ker f A
Proof: Let I be a left ideal of R. Then, clearly R/I is a left R-module. Put
x 5 1 1 I. Then, for any a R,
a 1 I 5 a(1 1 I) 5 ax <x>
ker f 5 {a R : f (a) 5 0}
5 {a R : ax 5 0}.
It can be easily verified that ker f is a left ideal of R. By the fundamental theo-
rem of R-homomorphisms,
R R
> M.
I ker f
EXERCISE 13(b)
4. For any R-module M, prove that the set EndR(M) of all R-endomorphisms of M
is a subring of the ring End(M) of all endomorphisms of the group (M, 1).
5. Let R be a ring with unity and consider R as a right R-module. Then, prove that
the ring R is isomorphic to the ring EndR(R) of all R-endomorphisms of R.
6. Let f : M → N be an R-homomorphism of R-modules and A and B be R-submod-
ules of M and N, respectively. Then, prove that f (A) and f21(B) are R-submodules
of N and M, respectively.
7. Prove that an R-homomorphism of R-modules is an R-monomorphism if and
only if its kernel is trivial (zero).
8. Let M be the set of all differentiable real valued functions defined on the set R of
real numbers. Then, prove that M is an R-module under the point-wise addition and
scalar multiplication. Prove that the derivative operator is an R-homomorphism of
M into RR and determine its kernel.
9. If M is a cyclic R-module, prove that the quotient M/N is also a cyclic R-module
for any R-submodule N of M.
10. Let M and N be R-module and M > N. Prove that M is cyclic (finitely generated)
if and only if N is so.
11. A sequence (finite or infinite) of R-modules and R-homomorphisms
f n1 fn f n1 f n2
… → M n1 → M n → M n1 →… is called exact if f i(Mi21) 5
ker fi11 for all i.
Suppose that the following diagram of R-modules and R-homomorphisms is
commutative.
A f B g C
� � �
D � E � F
12. Consider the group (Q, 1) of rational numbers as a Z-module. Prove that the
ring EndZ(Q) of Z-endomorphisms of Q is isomorphic to the ring Q of rational
numbers.
13. If R is a ring with unity and I is a left ideal of R such that R / I > R as R-modules,
then prove that there is an idempotent e in R such that Re 5 I. Is the converse true?
14. For any R-submodules A and B of an R-module M such that A 1 B 5 M, prove that
M M M
> .
A∩ B A B
Definition 13.3.1. Let R be a fixed given ring and {Mi}iI be any nonempty
class of R-modules. Let
M a : I → ∪ M i : a(i ) ∈ M i for all i ∈ I .
i∈ I
By the choice axiom, we observe that M is a nonempty set. For any and
M and a R, define
( 1 )(i) 5 (i) 1 (i)
and (a)(i) 5 a(i)
for all i I. Then, M is an R-module under these addition and scalar multipli-
cation. This M is called the direct product of {Mi}iI and is denoted by M i .
i∈ I
Each Mi is called a direct factor of the direct product M.
The direct product of any class of R-modules satisfies the following cat-
egorical property.
pi
M
Mi
q
qi
Proof:
1. For each i I, define pi : M → Mi by
pi(a) 5 a(i) for all a M.
Then, pi(a 1 ) 5 (a 1 )(i) 5 a(i) 1 (i) 5 pi(a) 1 pi()
and pi(aa) 5 (aa)(i) 5 aa(i) 5 api(a)
for all a and M and a R. Therefore, each pi, i I, is an R-homo-
morphism.
2. Let N be any R-module and, for each i I, let qi : N → Mi be an
R-homomorphism. Then, define q : N → M by
q(x)(i) 5 qi(x)
for all x N and i I. Then, q(x) M for all x N and q is a well-
defined mapping of N into M. For any x and y N and a R, we have
pi
M
Mi
q
qi
Also, since the direct product N satisfies the properties (1) and (2) (from
Theorem 13.3.1), there exists a unique R-homomorphism p : M → N such that
qi ο p 5 pi for all i I.
qi
N
Mi
p
pi
pi ο (q ο p) 5 (pi ο q) ο p
5qi ο p
5 pi 5 pi ο (IdM)
pi
M
Mi
qοp
pi
q ο p 5 IdM.
Similarly, we can prove that p ο q 5 IdN. Therefore, p and q are bijections and
inverses to each other. In particular, p : M → N is an R-isomorphism.
Thus,
M > N ∏ Mi .
i∈ I
x if j i
( j ) i .
a( j ) if j ≠ i
M
> Mi .
ker pi
∏M ∏M
i∈ I
i
i1
i M1 M 2 M n
which is the usual set of n-tuples (x1, x2, …, xn) with xi Mi.
Next let us turn our attention to the concept of direct sum of a given family
of R-modules over a given fixed ring R.
g
gi
Proof: Since M and {fi}iI satisfy (2) with M9 in place of N and f i9in place of gi,
there exists an R-homomorphism f 9: M → M9 such that f 9 ο fi 5 fi for all i I.
fi M
Mi
f′
f ′i
M′
Also, since M9 and {f i9}iI satisfies (2) with M in place of N and f i in place of
gi, there exists an R-homomorphism f : M9 → M such that
f ο f i9 5 fi for all i I.
f ′i M′
Mi
f
fi
M a ∈ ∏ M i : a is finite.
i∈ I
x if j i
f i ( x )( j )
0 if j ≠ i
for any x Mi and j I. Note that |fi(x)|⊆{i} and hence fi(x) M for all x Mi. It
can be easily verified that fi is an R-homomorphism. Now, let N be any R-module
and gi : Mi → N be an R-homomorphism for each i I. Define g : M → N by
Since || is finite, a(i) 5 0 and hence gi(a(i)) 5 0 for all but finite number of
i’s. Therefore,
g (a) ∑ gi (a(i ))
i∈|a|
and the summation in the definition of g(a) is meaningful. For any a and
M and a R, we have
g (a ) ∑ gi ((a )(i ))
i∈ I
∑ gi (a(i ) (i ))
i∈ I
∑ gi (aa(i ))
i∈ I
∑ agi (a(i ))
i∈ I
(g ο fi )( x ) g ( fi ( x )) ∑ g j ( fi ( x )( j )) gi ( x ),
j ∈I
g (a) ∑ gi (a(i ))
i∈ I
∑ (g ′ ο f i )(a(i ))
i∈ I
∑ g ′ ( f i (a(i )))
i∈ I
g ′ ∑ f i (a(i )) 5 g9(),
i∈I
Proof: From Theorems 13.3.4 and 13.3.5, we can take M as the R-submodule
of the direct product M i given by
i∈ I
M a ∈ ∏ M i : a is finite
i∈ I
x if j i
f i ( x )( j )
0 if j ≠ i
a 5 x1 1 x2 1 … 1 xn
x ∑ xi ,
i∈ I
with xi Ni for all i I and xi 5 0 for all but a finite number i’s, then M is
called the internal direct sum of {Ni}iI and denote this by M ⊕ N i . In this
iI
case, each Ni is called a direct summand of M. If I is a finite set, say I 5 {1,
2, …, n}, then ⊕ N i will be written as N1 ⊕ N2 ⊕ … ⊕ Nn.
iI
If M is the external direct sum of R-modules {Mi}iI, then we have proved in
Corollary 13.3.1 that there are R-submodules {Ni} of M such that Ni > Mi and
M is the internal direct sum of {Ni}iI. On the other hand, if M is the internal
direct sum of R-submodules {Mi}iI, then we prove below that M is (isomorphic
to) the external direct sum of {Mi}iI.
Proof: Let N be the external direct sum of {Mi}iI. That is, by Theorem 13.3.5,
N a ∈ ∏ M i : a is finite.
i∈ I
Define f : M → N as follows.
x j if i i j , 1 j n
f ( x )(i ) .
0 otherwise
( )
3. M i ∩ M j {0} for each i ∈ I .
i ≠ j ∈I
xi ∈ M i ∩ ∑ M j ⇒ xi x j1 x jn , jk ≠ i for 1 k n and x jk ∈ M jk
i ≠ j ∈I
⇒ (xi ) x j1 x jn 0.
⇒ xi 5 0 (by (2)).
(
Therefore, M i ∩ M j 5 {0} for all i I.
i ≠ j ∈I
)
(3) ⇒ (1): We are given that M = M i .
iI
If x1 1 x2 1 … 1 xn 5 y1 1 y2 1 … 1 yn and xj, y j ∈ M i j , then
x1 y1 ( y2 x2 ) ( yn xn ) ∈ M i1 ∩ ∑ M i {0}
i ≠i1
a1 a
x x1 a2 x2 3 x3 x1 x2 x3
2 3
e1 5 e1e1 Re1 ⊆ M
and e2 5 e2e1 1 e2(e2 2 e2e1) M1 1 M2,
Worked Exercise 13.3.3. Let R be a ring with unity and I be a left ideal of R.
Consider R as a left R-module. Then, prove that I is a direct summand of R if
and only if I 5 Re for some idempotent e in R.
x 5 xe 1 x(1 2 e) I 1 J
e 2 e2 5 ef I ∩ J 5 {0}.
x I ⇒ x 5 x(e 1 f) 5 xe 1 xf
⇒ x 2 xe 5 xf I ∩ J 5 {0}
⇒ x 2 xe 5 0
⇒ x 5 xe Re.
EXERCISE 13(c)
1. Consider the R-module R4 and determine whether R4 is the direct sum of <x1>,
<x2>, <x3> and <x4> in each of the following cases:
(i) x1 5 (0, 2, 0, 3), x2 5 (0, 3, 0, 4), x3 5 (2, 0, 0, 5) and x4 5 (0, 6, 0, 9)
3. Let I be any nonempty set and R be any ring. Then, RI, the set of all mappings of
I into R, is an R-module under the usual point-wise operations. Prove that there
exists a family {Mi}iI of R-modules such that Mi is R-isomorphic to R for each
i I and R I > M i .
i I
9. Let {Mi}iI and {Ni}iI be two families of R-modules and, for each i I, let
f i : Mi → Ni be an R-homomorphism. Then, prove that there is a unique
R-homomorphism f : M i → N i such that f()(i) 5 fi((i)) for all a M i
i∈ I i∈ I i∈ I
and i I.
11. State results similar to the above two exercises for external or internal direct
sums of R-modules and prove them.
12. Let R be a ring with unity and I be a left ideal of R such that ( R / I ) > R regarded
as R-modules. Then, prove that I 5 Re for idempotent e in R and deduce that I is
a direct summand of R.
n
RM ri xi : ri R and xi ∈ M .
i1
Example 13.4.1
1. Let R be a field or a division ring and consider R as left R-module. Then,
R has no left ideals except {0} and R. Therefore, R is a simple R-module.
2. Let F be a field and R 5 Mn(F), the ring of all n 3 n matrices over F. For
each 1 # i, j # n, let eij be the matrix whose ijth entry is 1 and all the other
entries are zero. Fix 1# k # n and consider M 5 Rekk. Then, M is a left
R-module. We prove that M is a simple R-module. Clearly, RM {0}. Let
N {0} be an R-submodule of M and let 0 ≠ ( aij ) ∈ N .
Then, since
n
( aij ) ∑ aik eik ,
i1
Theorem 13.4.1. The following are equivalent to each other for any module
M over a ring R with unity:
1. M is a simple R-module.
2. There exists a maximal left ideal I of R such that M > (R/I).
3. M {0} and M 5 Rx for any 0 x M.
fx 5 f(x)
Theorem 13.4.2. (Schur’s lemma) EndR(M) is a division ring for any simple
R-module M.
Proof: Let M be a simple R-module. We have already observed that the set
EndR(M) of all R-endomorphisms of M is a ring with unity under the point-
wise addition and composition of mappings as multiplication. Note that the
identity map is the unity element in the ring EndR(M). We have to only prove
that every nonzero element in EndR(M) has multiplicative inverse in EndR(M).
Let 0 f EndR(M). Then, ker f M and f(M) {0}. Both ker f and f(M)
are R-submodules of M. Since M is simple, it follows that ker f 5 {0} and
f(M) 5 M. These imply that f is a bijection and hence f is an R-isomorphism,
so that the inverse map f21 exists and is an R-endomorphism. Therefore, f has
multiplicative inverse in EndR(M). Thus, EndR(M) is a division ring.
Example 13.4.2
1. Clearly any simple R-module is completely reducible.
2. Let {M} be any family of simple R-modules and M be the (external)
direct sum of {M}. Then, M is completely reducible.
In fact, we prove below that any completely reducible module is neces-
sarily a direct sum of a family of simple R-modules.
P J ⊆ : ∑ M a is a direct sum and N ∩ ∑ M a {0} .
a∈ J
a∈ J
a∈I (
A N ⊕ ∑ M a N ⊕ ⊕ M a .
a∈ I
)
We prove that A 5 M. For this it is enough if we can prove that M ⊆ A for all
∆. Let ∆ be arbitrarily fixed. If I, then M ⊆ M a ⊆ A. Suppose
a I
that ∉ I. Now M ∩ A is an R-submodule of M. Since M is simple, we get
that M ∩ A 5 {0} or M. Suppose, if possible, M ∩ A 5 {0}. Then
( )
M ∩ ⊕ M a {0}
a∈I
and hence M a is a direct sum and has zero intersection with N. This
aI ∪{ }
implies that I ∪ {} P, which is a contradiction to the fact that I is a maxi-
mal member of P, since ∉ I and I I ∪ {}. Therefore, M ∩ A {0} and
hence M ∩ A 5 M so that M ⊆ A. Thus, M ⊆ A and hence A 5 M.
Thus, M a is a direct sum and M N ⊕ ( ⊕ M a ).
aI aI
M N ⊕ ⊕ Ma .( a∈ I
)
Then,
N>
M
(( ⊕ M ) ⊕ ( ⊕ M )) >
a∈ I
a
a∈ − I
a
⊕ Ma .
(⊕ M )
a∈ I
a (⊕ M )
a∈ I
a
a∈I
EXERCISE 13(d)
1. Which of the following are simple modules and which of them are completely
reducible?
(i) Z, as a Z-module
(ii) R, as a R-module
(iii) R, as a Q-module
(iv) Q, as a Z-module
(v) Q, as a Q-module
(vi) For any field F, F[x] as an F-module.
2. Let X be any nonempty set and R be the ring of all real numbers. Let
M 5 {a : X → R : |a| is finite}.
Prove that M is a completely reducible R-module and the M is simple if and only
if X is a singleton set.
3. Let R be ring of 2 3 2 matrices over a field F. Prove that R is a completely reduc-
ible R-module.
6. Prove that the direct sum of any family of completely reducible R-modules is
completely reducible.
Example 13.5.1
1. Consider R as an R-module. Then, clearly {1} is linearly independent.
2. For any positive integer n, consider Rn as an R-module. For each 1# i # n,
let ei denote the n-tuple in which the ith coordinate is 1 and all other coor-
dinates are 0. Then, {e1, e2, …, en} is linearly independent.
3. Let X be any nonempty set and consider RX as an R-module. For each x
X, define ex : X → R by
1 if y x
ex ( y ) .
0 if y ≠ x
Example 13.5.2
1. Any ring R is a free R-module, since {1} is a basis.
2. Also, as in Example 13.5.1 (2), Rn is a free R-module. The set {e1, e2, …, en}
is a basis for Rn. Any element x 5 (a1, a2, …, an) can be expressed as
This basis {e1, e2, …, en} is called the standard basis of Rn.
3. The set R[x] of polynomials over a ring R is a free R-module, since {1, x,
x2, …} is a basis for R[x].
4. Consider the R-module RX as in Example 13.5.1 (3). Then, RX is not a
free R-module. However, the R-submodule M of RX given by
M 5 {f RX : | f | is finite}
Proof: Suppose that G is finite. Then, there exists a positive integer n such
that na 5 0 for all a G (for example, we can take n 5 |G|) and hence there
is no linearly independent set in G; in particular, G has no basis. Therefore, G
is not a free Z-module. Conversely, suppose that G is an infinite group. Since
G is given to be cyclic, we get G is isomorphic to Z as Z-module. Since Z is
a free Z-module, it follows that G is also a free Z-module.
The following result suggests an alternate proof of the above theorem.
Theorem 13.5.2. Let M be a free R-module with a basis {x1, x2, …, xn}. Then,
M > Rn, as R-modules.
Proof: Define f : Rn → M by
x j ∑ a ji ei , a ji ∈ R
i∈ I
in which all but finite number of aji are zero. For each 1# j # n, let
Sj 5 {i I : aji 0}
n
and S ∩ S j . Then, each Sj and hence S are finite subsets of I.
j1
Let D 5 {ei : i s}.
Then, D is a linearly independent finite set and generates M and hence D is a
basis. Since D ⊆ B and B is also a basis of M, if follows that D 5 B. Thus, B
is a finite basis of M.
Let C be any other basis of M and |C| 5 m and |B| 5 n. Then, by Theorem
13.5.2, M > Rm and M > Rn. Therefore, there exists an R-isomorphism g :
Rm → Rn and let h 5 g21. Suppose, if possible, that m n. Without loss of
generality, we can suppose that m < n. Let {e1, e2, …, em} and {f1, f2, …, fn} be
the standard basis of Rm and Rn, respectively. Let
n
g (ei ) ∑ a ji f j for each 1 i m
j1
and
m
g1 ( f j ) ∑ bij ei for each 1 j n.
i1
n
(ei ) g −1 g (ei ) ∑ a ji g1 ( f j )
j1
n m
∑ a ji ∑ bkj ek .
j1 k1
m n
Therefore, ei bkj a ji ek for each 1 # i # m. Since {ei} are linearly
k1 j1
independent, we get that
n 1 if k i
∑b a ki .
j1
kj ji
0 if k ≠ i
B
A [ A 0 ] and B .
0
These are n 3 n matrices, where each of the 0 blocks is a matrix of appropri-
ate size. Then,
I 0
A9B9 5 In and B A m .
0 0
This implies that det(A9B9) 5 1 and det(B9A9) 5 0. This is a contradic-
tion, since A9 and B9 are n 3 n matrices over the commutative ring R
and det(A9B9) 5 det(B9A9). Thus, m < n is impossible. Similarly, n < m is
impossible. Thus, m 5 n.
n
M ′ X ∑ ai xi : ai ∈ R and xi ∈ X .
i1
M M
f :X → by f(x) 5 M9, the zero element in .
M′ M ′
Then, the natural map g : M → M defined by g(y) 5 y 1 M9 and the zero
M
homomorphism h : M → M defined by h(y) 5 M9 for all y M are both
M
R-homomorphisms such that
1 if x xi
f : X → R by f ( x )
0 if x ≠ xi .
Then, there exists an R-homomorphism f : M → N such that
f ( x ) f ( x ) for all x X.
Now, we have
0 f ( a1 x1 an xn ) a1 f ( x1 ) an f ( xn ) ai since f ( xi ) 1 and
f ( x j ) 0 for all j i. Therefore, ai 5 0 for each 1# i # n. Thus, X is lin-
early independent. Thus, X is a basis of M.
EXERCISE 13(e)
3. Prove that the direct sum of a family of free R-modules is again a free R-module.
Is this true for direct products?
4. If M1, M2, …, Mn are free R-modules, prove that the direct product M1 3 M2 3
… 3 M is a free R-module.
n
is called a linear combination of x1, x2, …, xn. For any subset X, the subspace
generated by X is called linear span of X in V and is denoted, as usual, by <X>.
Recall that
Theorem 13.6.1. Any nonzero vector space over any field F is free (i.e., a
free F-module).
P 5 {X ⊆ V : X is linearly independent}.
ax 5 0 ⇒ a21(ax) 5 0 ⇒(a21a)x 5 0⇒ x 5 0
ax 5 0 ⇒ a 5 0.
Corollary 13.6.1. Let V be a vector space over a field F. Then, any linearly
independent subset of V can be extended to a basis of V.
On the lines of the proof given above, we can prove that P has a maximal
member, say B. Then, B is a basis of V and S ⊆ B.
Theorem 13.6.2. Let V be a vector space over a field F. Then, any subspace
of V is a direct summand of V.
n m
x W ∩ W ⇒ x ∑ ai xi ∑ b j y j ,
i1 j1
Proof: Let V be a vector space over a field and let B be a basis of V. Then, it
can be proved that
V ⊕ Fx.
x∈ B
as vector spaces over F, where HomF(V, W) is the vector space of all linear
transformations of V into W.
Proof: Fix bases B 5 {e1, e2, …, em} and C 5 {f1, f2, …, fn} for V and W,
respectively. If : V → W is a linear transformation, then, for each 1# i # m,
(ei) w 5 <C> and hence
n
a(ei ) ∑ aij f j .
j1
Then, the m 3 n matrix A 5 (aij) is called the matrix of with respect to the
bases B and C. Conversely, if A 5 (aij) is a m 3 n matrix over F, then we can
define : V → W by
n
a(ei ) ∑ aij f j for each 1# i # n
j1
m m
and a( x ) ∑ ai a(ei ) if x ∑ ai ei .
i1 i1
n
ei ∑ bij e j for some bij F.
j1
Now, consider
n n n
ei ∑ bij ej ∑ bij ∑ a jk ek
j1 j1
k1
n n
∴ ei ∑ ∑ bij a jk ek .
k1
j1
k 1 if i k
∑b a ik .
j1
ij jk
0 if i ≠ k
Similarly, by considering
n n n n n
ei ∑ aij e j ∑ aij ∑ b jk ek ∑ ∑ aij b jk ek ,
j1 j1 k1 k1
j1
Then,
n n
∑ r ∑ a e 0
j1
j
i1
ji i
n
n
r a e 0.
∑ ∑ j ji i
i1 j1
∑r a
j1
j ji 0 for all 1 # i # n.
Let
r1
r
P 2
rn
0
0
AP .
0
Therefore,
0
0
A AP
0
and hence
0
0
P
0
Note: The matrix A 5 (aij) given in the above theorem is called the matrix of
transformation from the basis B9 to the basis B.
The following describes the effect of a change of a basis on the matrix of a lin-
ear transformation. First, recall that if B 5 {e1, e2, …, em} and C 5 {f1, f2, …, fn}
are basis of vector spaces V and W, respectively, and : V → W is a linear
transformation such that
n
a(ei ) ∑ aij f j for each 1 i m,
j1
then the m 3 n matrix A 5 (aij) is called the matrix of with respect to the
bases B and C.
1. Let B9 5 {e19, e29, …, em9} and C9 5 {f19, f29, …, fn9} be new bases of V
and W, respectively. Then, the matrix of with respect to the bases B9
and C9 is of the form PAQ21 where P and Q are matrices of transforma-
tions from B9 to B and C9 to C, respectively.
2. Conversely, if P and Q are m 3 m and n 3 n invertible matrices, respec-
tively, then there exist bases B9 and C9 of V and W, respectively, such that
PAQ21 is the matrix of with respect to the bases B9 and C9.
n
Proof: (1) We have a(ei ) aij f j for each 1# i # m.
j1
Let A9 5 (aij)9 be the matrix of with respect to the bases B9 and C9. Then,
we have
n
a(ei ) ∑ aij f j for each 1# i # m.
j1
n
f j ∑ qjk f k for 1 # j # n
k1
m
and ei ∑ pil el for 1# i # m.
l1
Therefore,
m
a(ei ) a∑ pil el
l1
m
∑ pil a(el )
l1
m n
∑ pil ∑ alj f j
l1 j1
m n n
∑ pil ∑ alj ∑ q ′jk f k′
l1 j1 k1
n m n
∑ ∑ pil ∑ alj q ′jk f k′.
l1 j1
k1
Thus, the matrix with respect to the bases B9 and C9 is equal to PAQ21.
This proves (1). The proof of (2) is similar to the above and to the proof
of Theorem 13.6.5.
Worked Exercise 13.6.1. Let F be any field and consider the vector spaces
F3 and F2 over F. Define : F3 → F2 by
(a, b, c) 5 (a 1 b 1 c, b 1 c)
Answer: For any x 5 (a, b, c) and y 5 (a9, b9, c9) F3 and r and s F, we
have
Therefore, is a linear transformation of F3 into F2. Let {e1, e2, e3} and
{f1, f2} be standard bases of F3 and F2, respectively. Then
Therefore,
1 0
1 1
1 1
is the matrix of a with respect to the bases {e1, e2, e3} and {f1, f2}.
EXERCISE 13(f)
4. Let F be a field and V 5 {f(x) F[x] : deg(f(x)) # 4}. Prove that B 5 {1, x, x2,
x3, x4} and C 5 {1, 1 1 x, 1 1 x 1 x2, 1 1 x 1 x2 1 x3, 1 1 x 1 x2 1 x3 1 x4}
are bases of V. If D is the differentiation operator on V, then determine the matrix
of D with respect to each of the bases B and C. Also determine the matrix of
transformation from B to C and from C to B.
5. For any subset X of a vector space V over a field F, let <X> be the linear span of
X in V. Prove the following for any subsets X and Y in V:
(1) X ⊆ <X>
(2) X 5 <X>
(3) <X ∪ Y> 5 <X> 1 <Y>
(4) <X> 5 ∪ {<Y> : Y is finite subset of X}.
It is well known that the field Q of rational numbers is a subfield of the field
R of real numbers and, in this case, we say that R is an extension field of Q.
Likewise, the field C of complex numbers is an extension of R. Let us recall
that the polynomial 1 1 x2 has no root in R. However, there is an extension
field, namely C, containing a root of 1 1 x2. In this chapter, we discuss in
detail about the existence of an extension field containing roots of a given
polynomial over a given field.
The field R of real numbers has a deficit that not all polynomials over R
have roots in R. The field C of complex number is an extension of R con-
taining all the roots of any polynomial over C. Such fields like C are called
algebraically closed fields. We discuss these and similar concepts in the pres-
ent chapter.
Here afterwards F denotes an arbitrary field, unless otherwise stated. Also,
a homomorphism of one field F into another field K is always assumed to be a
ring homomorphism of F into K carrying the unity in F onto the unity in K.
Example 14.1.1
1. For any field F, [F : F] 5 1.
2. The degree of the field C of complex numbers over the field R of real
numbers is 2, since {1, i} is a basis of C over R.
3. Let F be any field and F[x] be the ring of polynomials over F. Let K
be the field of quotient of F[x]. Then, K is a field extension of F. Also,
consider the set {1, x, x2, …}. If a01 1 a1x 1 a2x2 11 anxn 5 0, then
a0 5 a1 5 5 an 5 0 and hence {1, x, x2, …} is an infinite linearly
independent subset of K. Therefore, K is an infinite extension of F.
4. Let Q[ 2] 5 {a b 2 : a and b Q}. Then, Q[ 2] is a field extension
of Q and is of degree 2 over Q, since {1, 2} is a basis of Q[ 2] over Q.
The following theorem about the degrees of finite extensions of fields is
very important and useful.
[L : F] 5 [L : K] [K : F].
∑ a x y 0.
1im
ij i j
1 jn
m
For each 1 i m, let bi ∑ aij y j K .
j1
m m n
∑ b x ∑ ∑ a y j xi 0.
i i ij
i1 i1 j1
Since x1, x2, …, xm are linearly independent over K and bi K, it follows that
bi 5 0 for each 1 i m. Again, for each 1 i m,
∑a j1
ij y j bi 0
and, since {y1, y2, …, yn} is linearly independent over F and aij F, it follows
that
n
K i = ∑ aij y j , for each 1 ≤ i ≤ m
j =1
m m n
x ∑ ki xi ∑ ∑ aij y j xi ∑ aij xi y j .
i1 i1
j1 1im
1 jn
[L : F] 5 | B | 5 mn 5 [L : K] [K : F].
n
[ Fn : F1 ] ∏[ Fi : Fi1 ].
i2
Answer: Suppose that [Fn : F1] is finite. Then, for each 1 , i n, [Fi : F1]
is finite and [Fi : Fj] is finite for all j , i. Conversely, suppose that [Fi : Fj] is
finite for all i . j. Then, [Fi : Fi21] is finite for all 1 , i n and, in particular,
[Fn : F1] is finite. Also, in this case,
Q( p ) {a b p : a and b ∈ Q}.
Proof: First, let us recall that p is a real number which is not rational. For
any a, b, c and d Q, we have
( a b p ) (c d p ) ( a c) (b d ) p
1 ab p a b
and 2 2 p ∈ Q( p ).
ab p
2
a pb 2
a pb a pb 2
2
F( 3) 5 {a 1 b 3 : a and b F}.
[F( 3) : Q] 5 [F( 3) : F] [F : Q] 5 2 ? 2 5 4.
EXERCISE 14(a)
2. Consider two distinct prime numbers p and q and let K be the field given in
Exercise 1 above with p and q in place of 2 and 3, respectively. Then, prove that
K is a finite extension of Q and find the degree and basis of K over Q.
3. Let p1, p2, p3, … be the sequence of all prime numbers. Define Fn recursively as
follows:
Then, prove that, for each n Z1, Fn is a finite extension of Fn21 and is of degree
2 over Fn21.
Theorem 14.2.1. Let F be any field and p(x) F[x] be irreducible over F.
Then, there exists a field extension E of F such that E contains a root of p(x).
Proof: Since p(x) is given to be irreducible, the principal ideal ,p(x). gen-
erated by p(x) in the ring F[x] is a maximal ideal and hence the quotient ring
F[x]/,p(x). is a field. Define s : F → F[x]/,p(x). by
n
p( x p( x ) ) ∑ ai ( xp( x ) )i
i0
n
∑ ai ( x i p( x ) )
i0
n
∑ ai ( x i )p( x )
i=0
p( x )p( x ) 0 in E .
Proof: We prove this by using induction on the degree of f (x). Let deg(f (x)) 5
n. If n 5 1, then f (x) 5 a 1 bx, a and 0 b F and 2ab21 is the only root of
f (x) and hence F itself is the extension F containing all the roots of f (x). Let n > 1
and assume the theorem for all nonconstant polynomials of degree less then n.
By the above Corollary 14.2.1, there exists a field extension K of F such that K
contains a root a of f (x) in K. Now, x 2 a divides f (x) in K[x] and hence
Proof: Consider f (x) 5 f1(x), f2(x), …, fm(x) and use Theorem 14.2.2 above.
Example 14.2.1
1. For any field F, every element a of F is algebraic over F, since x 2 a
F[x] and a is a root of x 2 a.
2. 2 is algebraic over Q, since x2 2 2 Q[x] and 2 is a root of x2 2 2.
3. The complex number i is algebraic over Q since i is a root of 1 1 x2
Q[x].
4. It is known that the real numbers e and are transcendental over Q. The
proofs of these are beyond the scope of this book.
Let us recall that a nonzero polynomial over a field is called monic if its
leading coefficient is the unity element of the field.
It can be easily verified that I is an ideal of the ring F[x]. Since F[x] is a prin-
cipal ideal domain, there exists p(x) I such that
Without loss of generality, we can assume that p(x) is monic; for, if p(x) 5
a0 1 a1x 1…1 anxn, an 0 and g ( x ) an 1a0 an 1a1 x … an 1an1 x n1 x n ,
then p(x) 5 ang(x) and hence p(x) and g(x) are associates in F[x] so that
,p(x)> 5 ,g(x)>. Thus, we can assume that p(x) is a monic polynomial in
F[x] such that
From this, it follows that p(x) is irreducible; for, let p(x) 5 g(x)h(x) for some
nonconstant g(x) and h(x) F[x]. Then, 0 5 p(a) 5 g(a)h(a) and hence g(a)
5 0 or h(a) 5 0, so that g(x) or h(x) I. But, since deg(p(x)) 5 deg(g(x))
1 deg(h(x)) and both deg(g(x)) and deg(h(x)) are positive, it follows that
deg(g(x)) , deg(p(x)) and deg(h(x)) , deg(p(x)) which is a contradiction to
(*). Thus, p(x) is an irreducible monic polynomial in F[x] such that p(a) 5 0.
To prove the uniqueness of p(x), let
Example 14.2.2
1. We have Q ⊆ R and 2 R. x2 2 2 is the minimal polynomial of 2
over Q.
2. Let v be a root of 1 1 x 1 x2 1…1 xp21 in C, where p is a given prime
number. Then, using the Eisenstein criterion, we have proved that 1 1 x
1 x2 1…1 xp21 is irreducible over Q and hence it is the minimal poly-
nomial of v over Q.
Let F ⊆ K be fields and a K. Then, the smallest subfield of K contain-
ing F and a will be demoted by F(a). Note that F(a) is the intersection of all
subfields of K containing F ∪ {a} and it can be easily verified that
F(a) 5 {f (a) g(a)21 : f (x) and g(x) F[x] and g(a) 0}.
and hence each of an, an+1, an+2, … can be expressed in the form b0 1 b1a 1…1
bn21an21 with bi F and hence so is f (a) for any f (x) F[x]. Now, put
F [ x] E.
p(x )
[F(a) : F] 5 n 5 deg(p(x)). b
The converse of the above theorem is not true. That is, an algebraic
extension of a field F need not be a finite extension of F. For, consider the
following example.
Example 14.2.3. It is well known that the set of prime numbers is a countably
infinite set and hence we can express this set in a sequential form. Let p1, p2, …,
pn, … be all the distinct primes. For each n 0, En be the subfield of R defined
recursively by
Proof: We have f (x) F[x] and deg(f (x)) 5 n 1. Let p(x) be an irreducible
polynomial R over F such that p(x) divides f (x). By Theorem 14.2.1, there
exists a field extension K of F such that K contains a root of p(x). Let a K
be a root of p(x) and hence of f (x). By Theorem 14.2.4,
for each 1 , i n.
Note that any finite extension of F is finitely generated over F; but a finitely
generated extension may not be a finite extension. For, if a is a transcendental
number, then Q(a) is finitely generated over Q and it is not a finite extension
of Q. However, we have the following theorem.
Since an is algebraic over F, it is algebraic over F(a1, a2, …, an21) also and
hence
[E(a) : F] 5 [E(a) : E] [E : F] , ∞.
L 5 {a K : s(a) 5 t(a)}.
Then L is a subfield of K containing F and a1, a2, …, an. Therefore, F(a1, a2,
…, an) ⊆ L ⊆ K 5 F(a1, a2, …, an) and hence L 5 K. Thus, s(a) 5 t(a) for
all a K; that is, s 5 t. b
s1: F [ x ] → F ( a)
p( x )
s 2 : F [ x] → F ( b)
p( x )
s s 2 ο s1 1 : F ( a) → F (b).
EXERCISE 14(b)
Theorem 14.3.1. The following are equivalent to each other for any field K.
1. K is algebraically closed.
2. Any nonconstant polynomial over K can be factored completely into
linear factors in K[x].
3. If f (x) is a nonconstant polynomial over K, then all the roots of f (x)
belong to K.
4. Every nonconstant polynomial over K has atleast one root in K.
We prove later any two algebraic closures of a field F are isomorphic under
an isomorphism that keeps each element of F fixed. The following theorem is
in the direction of proving the existence of an algebraic closure of any field.
For any field F, we know that for any given indeterminates x1, x2, …, xn,
the set F[x1, x2, …, xn] of all polynomials in x1, x2, …, xn with coefficients in F
is an integral domain and we identify xixj with xjxi; that is, the indeterminates
are commuting with each other. For any set S 5 {xa}a∆ of commuting inde-
terminates, we define
F [S ] ∪
T ⊆S
F [T ].
T is finite
Consider the polynomial ring F[S], which is an integral domain. Let A be the
ideal in F[S] generated by all polynomials f (xf) of positive degree in F[S].
We claim that A is a proper ideal of F[S]. Suppose, if possible that A 5 F[S].
Then, 1 A and therefore
1 g1 f1 ( x f1 ) g 2 f 2 ( x f2 ) g x f x ( x fn ),
where g1, g2, …, gn F[S]. Note that g1, g2, …, gn will involve only a finite
number of variable (indeterminates). Write x fi xi for each f i F[x]. After
reindexing, we can assume that x f1 x1 ,, x fn xn and the variables occur-
ring in all the gi, 1 i n, are in the set {x1, x2, …, xn, …, xm}. Therefore,
we have
n
1 ∑ gi ( x1 ,…, xm ) f i ( xi ) (*)
i1
Now, let E be an extension of F in which each of the polynomials f1, f2, …, fn has a
root and let ai be a root of fi in E, for each 1 i n. If we substitute xi 5 ai for 1
i n and xn+1 5 5 xm 5 0 in (*), we get that 1 5 0, which is absurd.
Thus, A is a proper ideal of F[S] and hence, using the Zorn’s lemma, we get a
maximal ideal M of F[S] containing A. Then, there is an embedding a a 1
M of F into F [S ]/ M and hence F [S ]/ M can be regarded as a field exten-
sion of F. Also, each nonconstant polynomial f 5 f (x) F[x] has a root in
F [S ]/ M . Thus, we have constructed a field K1 ( F [S ] M ) which is an exten-
sion of F and in which every nonconstant polynomial in F[x] has a root.
Now, inductively, we can form a chain of fields
F 5 K0 ⊂ K1 ⊂ K2 ⊂
such that any nonconstant polynomial over Kn has a root in Kn+1 for all
n 0. ∞
Put K ∪ K n .
n0
Then, K is a field extension of F. If
is a polynomial over K, then there exists n such that b0, b1, …, bm Kn and
therefore g(x) Kn[x] so that g(x) has a root in Kn+1 ⊆ K. Thus, F has an alge-
braically closed extension. b
The following will be useful in proving the uniqueness (up to isomor-
phism) of algebraic closure of a given field.
Then, ps(x) L[x]. Since L is algebraically closed, ps(x) has all the roots in
L. Let p L be a root of ps(x). By Theorem 14.2.4,
Define hp : E → L by
Since p(x) is the polynomial of least degree for which a is a root, it follows
that any element of E ( 5 F(a)) can be uniquely expressed as b0 1 b1a 11
Then, it can be easily verified that is a partial order on P. We prove that the
partially ordered set (P, ) satisfies the hypothesis of the Zorn’s lemma. Let
{(Ka, ua)}a∆ be a chain of elements in P. Put
Theorem 14.3.7. Let f (x) be a nonconstant polynomial over a field F and let
E and K be two splitting fields of f (x) over F. Then, there exists an isomor-
phism s : E → K which is identity on F.
and b1, b2, …, bn E be all the roots of f (x) (not necessarily distinct). Then,
E 5 F(b1, b2, …, bn) is an algebraic extension of F. By Theorem 14.3.4, there
exists an embedding l* : E → K such that l*/F 5 Id. Put
Now, l*(b1), l*(b2), …, l*(bn) are the roots of f l*(x) in K , and if b1, b2, …,
bn K are the roots of f (x), then
since K ⊆ K . Also,
Worked Exercise 14.3.1. Find the splitting field of x3 – 2 over the field Q of
rational numbers and its degree over Q.
Q[ x ] 1 1
Q(2 3 ) and [Q(2 3 ) : Q] 3.
f ( x)
1
However, Q(2 3 ) is not the splitting field of f (x), since
1 1 2
f (x) 5 x3 2 2 5 (x 2 2 3 ) (x2 1 2 3 x 1 2 3 )
and hence f (x) has two complex roots, say w and w. Let
1 2 1
p(x) 5 x2 1 2 3 x 1 2 3
Q(2 3 )[x].
1
Then, p(x) is irreducible over Q(2 3 ) and hence
1
Q(2 3 ) [ x ] 1
Q (2 3 ) (w ) Q (2 3 , w )
1
p(x )
1 1
and [Q(2 3 , w) : Q(2 3 )] 5 2, which is the degree of p(x).
1 1
Further w Q(2 3 , w). Thus, Q(2 3 , w) is the splitting field of x3 – 2
over Q and
1 1 1 1
[Q(2 3 , w ) : Q] [Q(2 3 , w ) : Q (2 3 )][Q(2 3 ) : Q]
3.2 6.
Answer: 3i and 3i are the roots of f (x) and hence Q ( 3i) is the splitting
1 3 i 1 3
field of f (x). Further, are the roots of g(x). Put w .
2 2
Then, 3 5 2w 1 1 and hence 3 Q(w). Therefore, Q( 3 ) ⊆
Q(w). Also,
1 1
w 3 Q ( 3)
2 2
and hence Q(w) ⊆ Q( 3). Thus, Q(w) 5 Q( 3 ) is the splitting field of f (x)
and g(x). Also, f (x) and g(x) are both irreducible over Q and [Q(w) : Q] 5 2.
EXERCISE 14(c)
3. Construct splitting fields of the following polynomials over the field Q of ratio-
nal numbers.
(i) x4 1 1
(ii) x3 2 1
(iii) x6 2 1
5. Construct a splitting field for x3 1 x 1 1 over the field Z2 and list all its elements.
7. Prove that the degree of a splitting field of a polynomial of degree n (> 0) over a
field F is almost n!
8. Let p(x) F[x] be an irreducible polynomial over F. If p(x) has one root in
a splitting field E of a polynomial f (x) F[x], then prove that p(x) has all its
roots in E.
10. For any prime p, find the splitting field of xp 2 1 over Q and its degree over Q.
Example 14.4.1
1. Let f (x) 5 2 1 3x 1 x2 1 4x3 1 2x4 Q[x]. Then,
f 9(x) 5 3 1 2x 1 12x2 1 8x3
Theorem 14.4.1. Let F be any field, f (x) and g(x) F[x] and a F. Then
the following holds.
1. (f (x) 1 g(x))9 5 f 9(x) 1 g9(x)
2. (a f (x))9 5 a f 9(x)
3. (f (x)g(x))9 5 f 9 (x)g(x) 1 f (x) g9(x)
4. If char(F) 5 0 and deg(f (x)) 5 n > 0, then deg(f 9(x)) 5 n 2 1
5. If char(F) 5 p and f (x) 5 xp, then f 9(x) 5 0.
Definition 14.4.2. Let F be any field and f (x) F[x]. Let E be any extension
of F and a E. For any positive integer m, if (x 2 a)m divides f (x) in E[x] and
(x 2 a)m+1 does not divide f (x), then a is called a root of f (x) of multiplicity m
and m is called the multiplicity of the root a. A root of multiplicity 1 is called
a simple root and a root of multiplicity m > 1 is called a multiple root.
Earlier in Ring Theory, we have proved that a polynomial degree n over F
can have at most n roots in any extension of F. For this counting purpose, we
shall count a root a as m roots if a is a root of multiplicity m and not as one
root. In the following, we obtain a necessary and sufficient condition in terms
of the derivative of f (x) for a root of f (x) to be a multiple root.
Corollary 14.4.1. Let f (x) F[x]. Then, f (x) has only simple roots in any
extension E of F if and only if the g.c.d.{f (x), f 9(x)} is a unit in E[x].
Theorem 14.4.3. Let f (x) be an irreducible polynomial over F. Then, f (x) has
a multiple root in some field extension of F if and only if f 9(x) 5 0.
Theorem 14.4.4. The following holds for any irreducible polynomial f (x)
over a field F.
1. If char(F) 5 0, then f (x) has no multiple roots.
2. When char(F) 5 p 0, f (x) has a multiple root if and only if f (x) 5
g(xp) for some g(x) F[x].
Proof:
1. Let char(F) 5 0 and f (x) 5 a0 1 a1x 11 anxn.
Since f (x) is irreducible, deg(f (x)) 5 n > 0. Then,
f 9(x) 5 0 ⇒ a1 1 2a2x 11 nanxn21 5 0
⇒ a1 5 a2 5 5 an 5 0 (since char(F) 5 0)
⇒ f (x) 5 a0, which is not true.
Therefore f 9(x) 0 and hence, by the above theorem, f (x) has no mul-
tiple roots.
2. Let char(F) 5 p 0 and f (x) 5 a0 1 a1x 11 anxn with an 0 and n >
0. Then, f 9(x) 5 a1 1 2a2x 11 nanxn21. Suppose that f (x) has a multiple
root. Then, by Theorem 14.4.3, f 9(x) 5 0 and hence a1 5 2a2 5 3a3 5
5 nan 5 0. Since char(F) 5 p, it follows that, for each 1 K n, either
aK 5 0 or p divides K (that is, K 5 K1p for some integer K1 > 0).
Thus, we have
f (x) 5 b0 1 b1xp 1 b2x2p 11bmxmp 5 g(xp)
for some positive integer m, where g(x) 5 b0 1 b1x 11 bmxm.
Conversely suppose that f (x) 5 g(xp) 5 b0 1 b1xp 11 bmxmp. Then,
f (x ) pb1 x p1 pmbm x mp1 0 (since char(F) 5 p) and hence,
again by Theorem 14.4.3, f (x) has a multiple root. b
Note that a( p(x )) p( x ). Since a(x 2 a)m 5 (x 2 b)m, we get that (x 2 b)m is a
factor of p(x) and hence m m9. By interchanging the roles of a and b, we get
that m9 m. Thus, m 5 m9 and hence a and b are of same multiplicities. b
Worked Exercise 14.4.1. Prove that a polynomial f (x) F[x] has a multiple
root if and only if f (x) and f 9(x) have a nonconstant common factor.
Conversely suppose that f (x) and f 9(x) have a nonconstant common factor,
say p(x) with deg(p(x)) > 0.
Suppose, if possible, that all the roots of f (x) are distinct.
Let a1, a2, …, an be all the distinct roots of f (x) in any extension of F. Then,
n
f (x ) a∏ (x ai ), for some a ∈ F .
i1
n
and f (x ) a∑ ∏ (x a j ).
i1
ji
Therefore, for each 1 i n, f ( ai ) (ai a j ) 0 and hence no root of
j ≠i
f (x) is a root of f 9(x). Thus, f (x) and f 9(x) have no nonconstant common factor,
which is a contradiction to the hypothesis. Thus, f (x) has a multiple root.
Answer: We will be proving later in the next section that |F| 5 pn, where p is
prime, char(F) 5 p and n Z+ and a ap is an automorphism of the field F.
Suppose, if possible, that f (x) has a multiple root. Then, by Theorem 14.4.4 (2),
EXERCISE 14(d)
3. Let f (x) F[x] and deg(f (x)) 5 n > 0. If char(F) 5 p and f 9(x) 5 0, then prove
that p divides n and f (x) has at most np distinct roots.
We have proved above that the number of elements in any finite field is of
the form pn, for some prime p and n Z1. On the other hand, we prove in
the next two results below that, for any prime p and a positive integer n, there
exists unique (up to isomorphism) finite field with exactly pn elements. First,
we take up the uniqueness.
Proof: First recall that the prime subfield Fp is isomorphic to Zp and hence
Zp can be considered as the prime subfield of F. Let F * 5 F – {0}. Then, F *
pn 1
is a group under multiplicationn
and |F | 5 p – 1. Therefore, a
* n
1 for all
a F *. This implies that a p na for all a F. Therefore, every element of
F is a root of the polynomial x p x, which can have at most pn roots in any
pn
extension of F. Thus, F is precisely n
equal to the set of all roots of x x and
hence F is the splitting field of x p x, over Zp. b
Since any two splitting of a polynomial are isomorphic, the following is an
immediate consequence of the above theorem.
Corollary 14.5.1. If E and F are finite fields and |E| 5 pn 5 |F|, where p is a
prime and n is a positive integer, then E F.
and hence all the roots of f (x) are distinct. We have to prove that the roots of
f(x) form a subfield of K. If a and b are roots of f (x), then
n n n
( a b) p a p b p a b (since p divides pnc r )
n
(−a) p a (if p 5 2, note that a 5 –a)
n n n
( ab) p a p b p ab
n n
and ( a1 ) p ( a p )1 a1, if a 0.
Thus, the set E of all roots of f (x) forms a field and, since the roots are distinct,
|E| 5 deg(f (x)) 5 pn. Note that E is the splitting field of f (x) over Zp. b
Corollary 14.5.2. For any prime p and for any positive integer n, there exists
unique (upn to isomorphism) field E such that |E| 5 pn and this is the splitting
field of x p x over Zp.
mn
f ( x ) x p x F [ x ].
n
Since the multiplicative group F * 5 F – {0} is of order pn –1, we have a p 1 1
for all a F * Also, since
nm
a p a for all a F.
Zp Fp ⊂ F ⊂ E ⊂ F
For any field F, let us recall that the set Aut(F) of all automorphisms of F is
a group under the composition of mappings. In the following, we prove that
Aut(F) is a cyclic group for any finite field F.
Theorem 14.5.6. Let F be a finite field of characteristic p and |F| 5 pn. Then,
Aut(F) is a cyclic group of order n.
ap 5 bp ⇒ (a2b)p 5 0 ⇒ a2b 5 0 ⇒ a 5 b.
Further,
n
by Theorem 14.5.2, F is equal to the set of all roots of the polynomial
x p x ∈ Fp [x ]. If s : Fp(a) → F is an embedding, then s(a) F 5 Fs(a)
n
(since s(a) is a root of x p x), which implies that s(F) ⊆ F. Since F is finite
and s is an injection, we get that s(F) 5 F.
By Theorem 14.3.3, we know that the number of extensions s : Fp(a) → F
is equal to the degree of the irreducible polynomial satisfied by a. Therefore,
Aut(F) contains precisely n elements. Since O(f) 5 n, f is a generator of
Aut(F). Thus, Aut(F) is a cyclic group of order n. b
Worked Exercise 14.5.1. Let F be a finite field and |F| 5 pn, p a prime and
n Z1. For each divisor m of n, prove that F has exactly one subfield E of
F such that |E| 5 pm.
Answer: Let us recall from group theory that any cyclic group of order n has
a unique subgroup of order d for each divisor d of n. Let m be a divisor of n.
Consider the cyclic group F * 5 F – {0} of order pn – 1. Then, pm – 1 divides
pn – 1; for, if md 5 n, then
EXERCISE 14(e)
2. Find a generator for the multiplicative groups of nonzero elements of a field with
8 elements.
4. Let F be a field such that |F| 5 4. Then find irreducible polynomials over F of
degree 2, 3 and 4.
n
5. Prove that for n 3, the polynomial x 2 x 1 is irreducible over the field Z2
of integers modulo 2.
6. Let F be a finite field. Prove that any element of F can be expressed as the sum
of two squares.
7. Let a and b be two elements of a finite field F. Then prove that there exist ele-
ments a and b in F such that a 1 aa2 1 bb2 5 0.
8. Let F be a finite field and char(F) 5 p. Prove that each element a of F has a
p
unique pth root a in F.
Galois theory of fields is one of the most elegant theories in Abstract Algebra
and it is an excellent combination of group theory and the theory of algebraic
field extensions. In general, Galois theory and, in particular, the fundamental
theorem of Galois theory has several applications to the theory of equations
and geometry. Although we are not making a detailed study of Galois theory,
we shall discuss its fundamental concepts and give certain simple applica-
tions like proving the fundamental theorem of algebra and the nonconstruc-
tability of certain geometric figures using straight-edge and compass alone.
In this chapter, we introduce the concepts of separable extensions and normal
extensions and prove certain important properties of these. Also, we prove the
celebrated theorem ‘the fundamental theorem of Galois theory’ which estab-
lishes a one-to-one correspondence between the subfields of a splitting field
E of a separable polynomial in F[x] and the set of subgroups of the group of
F-automorphisms of E. This one-to-one correspondence transforms certain
problems of subfields of fields into more simpler and amenable problems
about the subgroups of groups. Certain applications of Galois theory are dis-
cussed in the next chapter.
Example 15.1.1
1. Clearly, any polynomial of degree one is separable.
2. x2 1 1 R[x] is an irreducible polynomial and its roots i and 2i are
distinct and therefore x2 1 1 is a separable polynomial.
3. The polynomial 1 1 x 1 x2 Q[x] is separable, since its roots are
(1) ( 3 )i and it is an irreducible polynomial.
2 2
4. The polynomial 1 1 2x 1 x2 is separable, since its irreducible factors are
1 1 x alone. In fact, for any a F, (x 2 a)n is a separable polynomial.
f (y) 5 y3 2 x K[y].
We prove that f (y) is irreducible over K and has multiple roots. If f (y) is
reducible over K, there must exist a root (g ( x ) h( x ))( h( x ) ≠ 0) in K which
implies that
(g ( x))
3
x
(h( x))
3
b13 x b23
which implies that (b1 b2 )3 b13 b23 0 (note that char(K) 5 3) and hence
b1 2 b2 5 0. This proves that f (y) has only one root whose multiplicity is
three. If a is a root of f (y) in any extension of K, then K(a) is an algebraic
extension of K and K(a) is not a separable extension.
ai a
aij = for 1 i n and 1 < j m.
b bj
Since F is infinite, we can choose 0 a F such that a aij for all 1
i n and 1 < j m. Then,
ai a
a≠ and hence ab 2 abj ai 2 a
b bj
for all 1 i n and 1 < j m. Take c 5 a 1 ab.
Then, clearly F(c) ⊆ F(a, b) (since a F).
Now, put h(x) 5 f (c 2 ax) F(c)[x].
Then, h(b) 5 f (c 2 ab) 5 f (a) 5 0.
Also, h(bj) 5 f (c 2 abj) 0 for all j > 1, since c 2 abj ai for all
i and a1, a2, …, an are the roots of f (x). Therefore, x 2 b is the only common
factor of g(x) and h(x) in F(c)[x]. Now, b is algebraic over F and hence over
F(c). If p(x) F(c)[x] is the minimal polynomial of b over F(c), then p(x)
divides g(x). Also, p(x) divides h(x) and hence p(x) divides x 2 b which
implies that p(x) 5 x 2 b. Therefore, b F(c) and hence ab F(c), since
a F. Now, a 5 c 2 ab F(c). Thus, a and b F(c) and hence F(a, b)
⊆ F(c). Thus, F(a, b) 5 F(c). If E 5 F(a1, a2, …, an), then by using induc-
tion on n, we can prove that E 5 F(c) for some c E. Thus, E is a simple
extension of F. b
Since any finite extension of a field is an algebraic extension and since any
algebraic extension of a field of characteristic zero is a separable extension,
the following is an immediate consequence of the above theorem.
be the minimal polynomial of a over K. Then, clearly q(x) divides p(x) in K[x]
and [K(a) : K] 5 n. Consider K9 5 F(a0, a1, …, an21). Then,
and hence we get that K 5 K9. Thus, we have proved that, for any intermediate
field K between F and E,
A 5 {[F(a) : F] : a E}.
{F(cd 1 b) : d F} is finite.
Example 15.1.3
1. The field C of complex numbers is a normal extension of the field R of
real numbers and [C : R] 5 2.
2. R is not a normal extension of the field Q of rational numbers; for, x3 2
2 Q[x] is irreducible over Q and has a root 3 2 in R, but it does not
split into linear factors in R, since it has complex roots.
Definition 15.1.5. Let {fa(x)}a be a family of nonconstant polynomials
over a field F. An extension E of F is called a splitting field of {fa(x)}a if
every fa(x) splits into linear factors in E(x) and E is generated over F by all the
roots of the polynomials fa(x), a .
If {f1(x), f2(x), …, fn(x)} is a finite family of polynomials over F, then the
splitting field of the family {f1(x), …, fn(x)} over F is same as the splitting
n
field of the simple polynomial f ( x ) f i ( x ) over F. On the same lines of
i1
the proof of Theorem 14.3.7, we can prove that any two splitting fields of a
family {fa(x)}a of polynomials over F are isomorphic under an isomor-
phism that keeps each element of F fixed. The following gives equivalent
statements for an extension to be normal.
F [ x]
s1 : → F ( a)
< p( x )>
F [ x]
and s2 : → F ( b)
p( x )
given by s1(f (x) 1 < p(x)>) 5 f (a) and s2(f (x) 1 <p(x)>) 5 f (b).
Put s s 2 ο s1 1; F(a) → F(b). Then, s is an isomorphism such that
s(a) 5 b and s(c) 5 c for all c F. Then, by Theorem 14.3.4, s can be
extended to an embedding s* : E → F . Now, by (2), s* maps E onto E; that
is, s* is an automorphism of E. Therefore,
f (x) and hence s(ai) 5 aj for some j which implies that s(ai) E for all
1 i n. Therefore, s(E) ⊆ E and s is an automorphism of E. By Theorem
15.1.4, E is a normal extension of F. b
Worked Exercise 15.1.1. Let F ⊆ E be fields such that [E : F]52. Then prove
that E is a normal extension of F.
[E : F(a)][F(a) : F] 5 [E : F] 5 2.
Also, since F F ( a), [F(a) : F] > 1 and hence it follows that [F(a) : F] 5 2 and
[E : F(a)] 5 1. Therefore, F(a) 5 E and
Since deg(p(x)) 5 2 and since E has a root of p(x), it follows that the other
root of p(x) must also be in E. Thus, E is the splitting field of p(x) F[x] and
hence E is a normal extension of F.
F ⊂ F(c) ⊂ F(a, c)
and F ⊂ F(a) ⊂ F(a, c).
EXERCISE 15(a)
2. Ifz is the complex number such that z 1 and z3 5 1, then find a such that
Q( 2, z ) Q( a).
3. Prove that any extension of Q is separable.
7. If E is an extension of a field F, then prove that the set of all elements in E which
are separable over F forms a subfield of E containing F.
E
G {s Aut( E ) : s ( a) a for all a F}.
F
E
G [ E : F ]
F
[E : F] 5 [F(a) : F] 5 deg(p(x)) 5 n.
E
G n [ E : F ]. b
F
F ⊆ EH ⊆ E.
Theorem 15.2.2 (Dedikind Theorem). Let F and E be fields and s1, s2, …, sn
be distinct embeddings of F into E. Let a1, a2, …, an E such that
Proof: Suppose, if possible, that there exist a1, a2, …, an E, not all zero,
such that
Then, we can find such a relation having as few nonzero coefficients as pos-
sible. On renumbering, we can assume that this relation is
for any a E. Multiplying (1) by s1(c) and subtracting the result from (2),
we get that
[E : EH] 5 |H|.
for all 1 j n. We can choose these y1, …, yn11 such that as few of them as
possible are nonzero and renumber them such that
for all 1 j n. Now, let g H and operate on (2) with g. Then, we get the
system of equations
Since H 5 {gg1, gg2, …, ggn} 5 {g1, g2, …, gn}, (3) is equivalent to the sys-
tem of equations
for all 1 j n. This is a system of equations like (2), but with fewer terms,
which becomes a contradiction to our assumption, unless
If this happens, then g ( yi y11 ) yi y11 for all g H and hence yi y11 EH
for all 2 i r. Therefore, there exist z1, z2, …, zr EH such that y1zi 5 yi
−1
for 1 i r (take z1 5 1 and zi = yi y1 for 2 i r). Then, Equation (2)
with j 5 1 becomes
g1(a1)y1z1 1 1 g1(ar)y1zr 5 0
g1(a1z1 1 1 arzr) 5 0
and hence a1z1 1 1 arzr 5 0, since g1 is an embedding. Now, since {a1, …, ar}
are linearly independent over EH, it follows that
z1 5 z2 5 = zr 5 0
g1(aj)x1 1 1 gn(aj)xn 5 0, 1 j m,
in n unknowns x1, …, xn. Since m < n, this system has a nontrivial solution.
Therefore, there exist y1, y2, …, yn E, not all zero, such that
for all 1 j m. Since {a1, a2, …, am} is a basis of E over EH, any element
a E can be uniquely written as a 5 s1a1 1 1 smam with si EH, and
hence
n n m
∑ g (a) y ∑ g ∑ s a y
i1
i i
i1
i
j1
j j i
m n
∑∑ s j gi ( a j ) yi
j1 i1
m n
∑ s j ∑ gi ( a j ) yi
j1
i1
0 (by (6))
which is a contradiction to the fact that g1, g2, …, gn are distinct embeddings
of E into E (refer Dedikind’s Theorem 15.2.2). Therefore, m < n is impossible
so that [E : EH] n. Thus,
[E : EH] 5 n 5 |H|. b
E E
G H and [E : EH] 5 G
E H E H
E
| H | G [ E : E H ] | H |
E H
E
[ E : E H ] G . b
E H
[E : F] 5 [F(a) : F] 5 n.
E
[ E : E0 ] G .
F
n
f ( x ) ∏ ( x s i ( a)).
i1
s *i ( a0 a1 x ar x r ) s i ( a0 ) s i ( a1 ) x s i ( ar ) x r .
n
Now, s *i f ( x ) (x s i (s j ( a))).
j =1
Since sis1, sis2, …, sisn are distinct members of G ( E / F ) and |G(E/F)| 5 n,
we get that G ( E / F ) = {s i s1 , s i s 2 , …, s i s n } and hence s *i ( f ( x )) f ( x )
for all 1 i n. By expanding f (x), we have
E
G The number of distinct roots of p( x ) n.
F
E
[ E : F ] G .
F
Worked Exercise 15.2.1. Let f (x) F[x] has r distinct roots in its splitting
field E over F. Then prove that G ( E / F ) is isomorphic to a subgroup of the
symmetric group Sr of degree r.
Answer: Let a1, a2, …, ar be all the distinct roots of f (x) in its splitting field
E over F. For any s G ( E / F ), s(ai) is again a root of f (x) in E. Also,
s(ai) s(aj) for ai aj. Thus, s(a1), s(a2), …, s(ar) is a permutation of a1,
a2, …, ar and let us denote this permutation by s. Therefore, s Sr for
each s G ( E / F ) . Define : G ( E / F ) → Sr by (s) 5 s. For any s and
G ( E / F ),
EXERCISE 15(b)
1. Let f (x) 5 x4 2 2 Q[x] and E be the splitting field of f (x) over Q. Prove that
G ( E / Q) is isomorphic to the group of symmetries of a square.
6. Let a 1 and a5 5 1. Then prove that Q(a) is a normal extension of Q and that
G (Q( a) Q) is isomorphic to Z4, the group of integers modulo 4.
Proof:
1. Define : → and u : → by
E E
K1 ⊆ K 2 ⇒ G ⊆ G ⇒( K 2 ) ⊆ ( K1 )
K 2 K1
E E
and G ⊆ G ⇒ E E ⊆ E E ⇒ K1 ⊆ K 2.
K 2 K1 G
K1
G
K2
E s
E
ker f s G : Id
G .
F K
K
E
G
F K
Im f ⊆ G .
E
F
G
K
E E K
G [ E : F ] [ E : K ][ K : F ] G G
F K F
Now,
E
G
F K
Im f G
E F
G
K
E
G
F K
G .
E F b
G
K
Worked Exercise 15.3.1. Prove that the Galois group of x3 2 2 Q[x] is the
group of symmetries of the triangle.
E
G {Id, s , s 2 , t , st , s 2 t}
Q
Id s s2 t st s2t
21/ 3 21/ 3 v21/ 3 v 2 21/ 3 21/ 3 v21/ 3 v 2 21/ 3
v v v v v2 v2 v2
Worked Exercise 15.3.2. Prove that the Galois group of x4 2 2 Q[x] is the
group of symmetries of a square (that is, the octic group).
E
G [ E : Q] 8.
Q
Therefore, the Galois group G (E Q) is a group of order 8. Note that {1, 21/4,
21/2, 23/4, i, i21/4, i21/2, i23/4} is a basis of E over Q. Now, if b E, then
s1 s2 s3 s4 s5 s6 s7 s8
i i i i i i i i i
1/4 1/4 1/4 1/4
2 2 i2 2 i 21/4 21/4 i21/4
21/4
i 21/4
a3 a1 d1
a4
s1 : rotation by 0°
s2 : rotation by 90°
s3 : rotation by 180°
s4 : rotation by 270°
s5 : rotation by d1
s6 : rotation by l1
s7 : rotation by d2
s8 : rotation by l2
These are precisely all the symmetries of the square whose vertices are a1, a2,
a3 and a4. Thus, the Galois group G (E Q) of the polynomial x42 2 over Q
is isomorphic with the group of symmetries of a square and hence G (E Q)
is the octic group.
f (x)5(x 2 1)(x4 1 x3 1 x2 1 x 1 1)
EXERCISE 15(c)
1. In each of the following, find all the subgroups of the Galois group of f (x) and
the corresponding fixed fields.
(i) f (x) 5 x3 2 2 Q[x]
(ii) f (x) 5 x4 1 1 Q[x]
(iii) f (x) 5 x4 2 2 Q[x]
(iv) f (x) 5 x2 2 a F[x], where F is a field of characteristic 2.
2. Prove that the Galois group of x4 1 1 Q[x] is the Klein four-group and is
isomorphic to Z23 Z2.
4. Let a be a real number such that Q(a) is a normal extension of Q for which [Q(a)
: Q] 5 2m, where m ≥ 0. Prove that there are fields E0 5 Q ⊂ E1 ⊂ E 2 ⊂ … ⊆
Em 5 Q(a) such that [Ei : Ei21] 5 2 for each 1 i m.
5. If K is the splitting field of x4 2 3x2 1 4 over Q, then find the Galois group
G (K Q).
6. Let a cos (2 3) i sin (2 3). Then, find the Galois group G (( a) Q) and
all its subgroups and the corresponding fixed fields.
Now,
( )( )
p(x ) x b b 2 a x b b 2 a ,
Theorem 16.1.2. Let f (x) R[x] be of odd degree. Then, f (x) has a real root.
Proof: Without loss of generality, we can assume that f (x) is a monic poly-
nomial. Suppose
g (x ) (x 2 1) f (x )f (x )
( x 2 1)( a0 a1 x an x n )( a0 a1 x an x n ).
Then, g(x) R[x]. Let E be the splitting field of g(x) over R. Then, R ⊆ C
⊆ E, since i is a root of g(x). We prove that E 5 C. First observe that [C : R]
5 2 and is a divisor of [E : R]. Therefore, [E : R] is an even positive integer.
Suppose that
f ( x ) f ( x ) ( g ( x ) an x n )( g ( x ) an x n )
g ( x ) g ( x ) ( an g ( x ) an g ( x )) x n an an x 2n
n1
g ( x ) g ( x ) ∑ ( an ai an ai ) x ni an an x 2n
i0
Example 16.2.1 2 p i
( ) ( )
n
1. For any n Z , e 1
cos 2np isin 2np is primitive nth root of
unity in C.
2. Let F be a field of characteristic zero. Then, for any positive integer n,
xn 2 1 has n distinct roots in its splitting field and form a cyclic group
under multiplication. If a is a primitive nth root of unity, then am, where 0
, m , n and (m, n) 5 1, is also a primitive nth root of unity and we know
that the number of such m is f(n).
1 i 3
3. is a primitive cube root of unity.
2
( ) ( )
4. cos 65p isin 65p is a primitive 5th root of unity.
Examples 16.2.2
1. Recall that any finite extension of a finite field is separable. If E is the
splitting field of a polynomial f (x) over a finite field, then E is a Galois
extension of F and, by Worked Exercise 14.5.1, G(E/F) is a cyclic group
and hence E is a cyclic extension of F.
2. Let p be a prime and a be a primitive pth root of unity. Then, E 5 Q(a) is
the splitting field of xp 2 1 Q[x] and E is a cyclic extension of Q.
Theorem 16.2.2. Let F be a field and suppose that F contains a primitive nth
root of unity. Then, E is a finite cyclic extension of F of degree n if and only it
E is the splitting field of an irreducible polynomial xn 2 b F[x].
c, cv, cv2, …, cvn–1 are the roots of xn 2 b. However sr(c) 5 cvr implies
that c, cv, cv2, …, cvn21 are also roots of the minimal polynomial f (x) of c
over F. Therefore, xn 2 b divides f (x). Since c is a root of xn 2 b, it follows
that f (x) divides xn 2 b and hence f (x) 5 xn 2 b. Since f (x) is the minimal
polynomial of c over F, f (x) is irreducible over F; that is, xn 2 b is irreducible
over F. Also,
and hence F(c) 5 E. Therefore, E is the splitting field of the irreducible poly-
nomial xn 2 b F[x] over F.
Conversely suppose that xn 2 b F[x] is an irreducible polynomial over F
and E is its splitting field over F. Let c E be a root of xn 2 b; that is, b 5
cn. Then, clearly c, cv, cv2, …, cvn21 are the n distinct roots of xn 2 b, where
v F is a primitive nth root of unity. Therefore, xn 2 b is a separable irreduc-
ible polynomial and hence E 5 F(c) is a Galois extension of F. For each s
G(E/s), let the set As be defined by
As 5 {r Z : s(c) 5 vrc}.
As 5 r 1 nZ,
since vrc 5 vsc if and only if r s(mod n). Further, for any s and G(E/F),
Therefore, As 5 As 1 A, where the sum on the right side is interpreted as
the binary operation on the additive group Z/nZ (Zn) of integers modulo
n. Finally, if As 5 nZ, the zero in Z/nZ, then s(c) 5 c. Therefore, s is the
identity on E (since E 5 F(c) and s/F 5 Id). Consequently, s As is an
isomomorphism of G(E/F) onto a subgroup of Z/nZ. Also,
such that the quotient Gi11/Gi is an abelian group for 0 # r , n; and, in this
case the series {e} 5 G0 ⊆ G1 ⊆ … ⊆ Gn 5 G is called a solvable series.
Examples 16.3.1
1. Any abelian group G is a solvable group, since {e} 5 G0 ⊂ G1 5 G is a
solvable series.
2. Let S3 be the symmetric group of degree 3 and let H 5 {e, (1 2 3), (3
2 1)}, where (1 2 3) is the 3-cycle mapping 1 → 2, 2 → 3, and 3 → 1.
Then, H is a subgroup of S3. Now,
{e} ⊂ H ⊂ S3
is a solvable series in S3, since H is an abelian normal subgroup and S3/H
is a group of order 2 and hence abelian. Therefore, S3 is a solvable group.
Recall from group theory that for any elements a and b of a group G, the
element aba21b21 is called a commutator in G and the subgroup G9 generated
by the set of all commutators in G is called the derived subgroup of G. For
any positive integer n, we define the nth derived subgroup of G, denoted by
G(n), is defined recursively as follows:
Theorem 16.3.1. A group G is solvable if and only if G(n) 5 {e} for some
positive integer n.
{e} 5 G0 ⊆ G1 ⊆ … ⊆ Gn 5 G
is a solvable series in G. For each i, Gi11/Gi is an abelian group and hence its
derived subgroup is trivial. This implies that
By induction, we can prove that G(i) ⊆ Gn21 and hence G(n) ⊆ G0 5 {e}. Thus,
G(n) 5 {e}.
Conversely suppose that G(n) 5 {e} for some n . 0. Then,
is a solvable series, since H/H, is abelian for any group H. Thus, G is a solv-
able group. b
Theorem 16.3.2. Any subgroup and any quotient group of a solvable group
is solvable.
Proof: Let G be a solvable group. Then, there exists a positive integer n such
that G(n) 5 {e}. If H is any subgroup of G, then H(n) ⊆ G(n) 5 {e} and hence
H(n) 5 {e} so that H is solvable. Let G/N be a quotient group of G, where N is
a normal subgroup of G. Now,
Proof: Recall that any subgroup of G/N is of the form H/N, where H is a
subgroup of G containing N. Now, since G/N is a solvable group, there exist
a sequence of subgroups
N ⊆ H1 ⊆ H2 ⊆ … ⊆ Hn 5 G
such that {N} ⊆ H1/N ⊆ H2/N ⊆ … ⊆ Hn/N 5 G/N is a solvable series in G/N.
In particular, each Hi is a normal subgroup of Hi11 (since Hi/N is a normal
subgroup of Hi11/N). Also,
(Hi11/N)/(Hi/N) Hi11/Hi
and hence Hi11/Hi is an abelian group. Since N is also solvable, there exists
solvable series
{e} 5 N0 ⊆ N1 ⊆ … ⊆ Nm 5 N
{e} 5 N0 ⊆ N1 ⊆ … ⊆ Nm ⊆ H1 ⊆ H2 ⊆ … ⊆ Hn 5 G
Gi H i ,0 ⊆ H i ,1 ⊆ ⊆ H i , mi Gi1
such that Hi, j is a normal subgroup of Hi, j11 and Hi, j11/Hi, j is a group of prime
order. This is true for each 0 # i , n and hence, by clubbing all these sequences
we get a required sequence of subgroups in G. The converse is trivial. b
Answer: Let A4 be the alternating group of degree 4 (that is, the subgroup of
even permutations in S4). Then, A4 is a normal subgroup of S4. Put
a 5 (1 2) o (3 4), b 5 (1 3) o (2 4) and c 5 (1 4) o (2 3)
Worked Exercise 16.3.2. Prove that the dihedral group Dn is solvable for any
positive integer n . 1.
Worked Exercise 16.3.3. Prove that the symmetric group Sn is not solvable
for any n . 4.
F 5 F0 ⊆ F1 ⊆ … ⊆ Fn 5 K
such that, for each 0 , i # n, Fi 5 Fi21(ai) for some ai Fi with the property
that airi Fi1 for some ri 1.
Observe that, if Fi 5 Fi21(ai) and airi Fi1 then ai is a root of the polynomial
x ri airi Fi1[ x ] and hence Fi is a simple algebraic extension of Fi21 and
therefore, [Fi : Fi 21] is finite. Also, since
F 5 E0 ⊆ E1 ⊆ E2 ⊆ … ⊆ Er 5 E
f1 (x ) ∏
F ( v )
( x n1 s (c1 ))
s G
F
Since K is a normal extension of F, f2(x) F(x) and hence g2(x) F[x]. Take
L to be the splitting field of g2(x) over F. Then, L contains a2 and K and hence
E2 5 E1(a2) ⊆ L. Therefore, L is a normal extension of F containing E2. Also,
because of the nature of the polynomial g2(x), it is clear that there exists a
finite ascending sequence of intermediate fields between K and L such that
any member of the sequence is a splitting field of a polynomial of the form
Theorem 16.4.2. Let n be a positive integer and F be a field containing all the
nth roots of unity. Let E be the splitting field of the polynomial xn 2 a F[x].
Then, the Galois group G(E/F) is abelian.
Proof: If b is any root of xn 2 a, then b, bc, bc2, …, bcn21 are the roots of xn
2 a, where c is a primitive nth root of unity and, also E 5 F(b). For any s1 and
s2 G(E/F), s1(b) 5 bci and s2(b) 5 bc j for some i and j and hence
We prove that the later is a solvable series in G(E/F), so that G(E/F) is a solv-
able group. Since F(c) is the splitting field of the polynomial xn 2 1 F[x], we
get that F(c) is a normal extension of F. By the fundamental theorem of Galois
theory, G(E/F(c)) is a normal subgroup of G(E/F). Also, since F(c) contains the
primitive nth root of unity c and since E is the splitting field of the polynomial xn
2 a F(c)[x], we get from Theorem 16.4.2 that F(E/F(c)) is an abelian group,
Further, by the fundamental theorem of Galois theory, we have
E
G
F F (c)
G Z*n ,
E F
G
F (c)
Proof: Suppose that f (x) is solvable by radicals over F. That is, the splitting
field E of f (x) over F is contained in a radical extension of F. Hence we can
find a sequence of fields
such that viri Fi1 for some integers ri 1 and E ⊆ Fm. By Theorem 16.4.1,
we can suppose that Fm is a normal extension of F and Fi is the splitting field
of x ri viri Fi1[ x ]. Now,
G(Fm/Fi21)/G(Fm/Fi) G(Fi/Fi21).
where Hi21/Hi is a solvable group. Now, Hm (5{e}) and Hm21/Hm are solvable
and hence, by Theorem 16.3.3, Hm21 is a solvable group. Also, since Hm22/
Hm21 is solvable, so is Hm22. Continuing this process, we get that H0 is a solv-
able group. That is, G(Fm/F) is a solvable group. Since E ⊆ Fm and
G(Fm/F)/G(Fm/E) G(E/F),
[E : F] 5 |G(E/F)| 5 n, say.
We divide the proof into two parts (1) and (2) as given below.
1. First we assume that F contains a primitive nth root of unity. Then, F
contains primitive mth roots of unity for all positive integers m which
divide n. since G(E/F) is a finite solvable group, there exists a sequence
of subgroups
{e} 5 G ⊆ G ⊆ G ⊆ … ⊆ G 5 G(E/F)
0 1 2 r
Corollary 16.4.1. Let f (x) F[x] such that the Galois group of f (x) is
isomorphic to Sn for some n . 4. Then, f (x) is not solvable by radicals
over F.
Recall from Worked Exercise 15.2.1 that the Galois group of a polynomial
f (x) F[x] having r distinct roots is isomorphic to a subgroup of Sr which is
the group of permutations of the r distinct roots of f (x). Before we give some
more applications of Theorem 16.4.4, let us have the following definition.
Proof: Let E be a splitting field of f (x) over F and a1, a2, …, an be the roots
of f (x) in E. Let G be the Galois group of f (x) over F. Then, G G ( E / F ).
For each s G, s(a1), s(a2), …, s(an) are roots of f (x) in E and hence a per-
mutation of a1, a2, a3, …, an. As in Worked Exercise 15.2.1, we can consider
G as a subgroup of Sn.
Now, suppose that G is a transitive permutation group. Let p(x) be the mini-
mal polynomial of a1 over F. For each root ai, there exist s G such that
s(a1) 5 ai (since G is transitive). Then,
Therefore, each ai is a root of p(x). Since p(x) divides f (x), it follows that
f (x) 5 cp(x) for some c F. Since p(x) is irreducible over F, f (x) is also
irreducible over F.
Conversely suppose that f (x) is irreducible over F. Then, ,f (x)., the ideal
generated by f (x) in F(x), is a maximal ideal of F(x) and hence F [ x ] f (x )
is a field and
(i 1 1 i 1 2) 5 si ο a ο si H for all 0 # i # n 2 2.
(i i 1 1) ο (i 1 1 i 1 2) ο (i i 1 1) 5 (i i 1 2)
and (i i 1 2) ο (i 1 2 i 1 3) ο (i i 1 2) 5 (i i 1 3) and so on.
Also 0 is not a root of f (x). Thus, there are at most three real roots of
f (x). Also, the roots occur in conjugate pairs. Thus, all these imply that f (x)
has exactly two nonreal roots in C. By Theorem 16.4.6, the Galois group of
f (x) over Q is isomorphic with S5, which is not a solvable group. Thus, by
Theorem 16.4.4, f (x) is not solvable by radicals over Q.
Mathematicians proved that all of the above impossible using the techniques
of Galois theory. In this section, we present proofs of the impossibilities of the
above.
Before we take up the proofs, let us be clear that a ruler (or straight edge)
is an instrument through which we can draw a line segment joining two given
points in the Euclidean space and that a compass is an instrument by which we
can draw a circle with a given point as centre and passing through another given
point. Also, let us understand that ‘construction by using ruler and compass’
means ‘construction by using ruler and compass only in finite number of steps’.
Let us imagine that we are given a line segment which we shall define to
be one unit in length.
Theorem 16.5.1. If a and b are constructible real numbers, then so are a1b,
a 2 b, ab and a/b if b ≠ 0.
Proof: Let a and b be constructible real numbers. We can suppose that a and b
are not zero. Then, there are line segments of lengths |a| and |b| available to us.
Now, extend the line segment of length |a|, and lay off on the extension the length
b with compass. This constructs a line segment of length a 1 b. Similarly, we
construct a 2 b. Notice that, clearly 2a and 2b are constructible.
Draw a line OA of length |a| and extend it. Draw a line through O not
containing A. Suppose OB is of length |b|. Take the point P such that OP is
of unit length, by using the compass. Draw a line parallel to PA and passing
through the point B which cuts the line OA at Q. By the property of similar
triangles, we get that
B
P
O A Q
1 b
a OQ
and hence |OQ| 5 |a||b| 5 |ab|. Thus, ab is constructible. Next, let |OA| 5 |a|
and draw a line OB of length |b| through O not containing A. Suppose P is a
point on OB such that OP is of unit length. Draw BA and a line parallel to BA
passing through P which cuts the line OA at Q, say. Then, again by a property
|a|
of similar triangles, we have | OQ | a . Thus, a/b is constructible. b
|b| b
B
O Q A
Corollary 16.5.1. The set of all constructible real members forms a subfield
F of the field R of real numbers.
Since Q is the prime subfield of R, it follows that the field F of con-
structible real numbers is a field extension of Q. Now, in the two-dimensional
Euclidean plane, we can locate any point (q1, q2) whose both coordinates are
rational numbers.
Proof: First notice that starting from Q 3 Q, any further point in the plane
which can be located by using ruler and compass can be found in one of the
following three ways:
1. As an intersection of two constructible lines (that is, lines passing
through two given points having rational coordinates).
2. As an intersection of a constructible line and a constructible circle.
3. As an intersection of two constructible circles.
Equations of lines and circles of the type mentioned in (1), (2) and (3) above
are of the form.
ax 1 by 1 c 5 0
and x2 1 y2 1 dx 1 ey 1 f 5 0,
where a, b, c, d, e and f are all rational numbers. It is clear that, for case
(1) above, a simultaneous solution of two linear equations with rational
coefficients can only lead to rational values of x and y, which give us a
new constructible point. For case (2) above, upon substitution in a qua-
dratic equation using linear equation and when solved by the quadratic
formula, gives solutions involving square roots of numbers which are pos-
sibly not square in Q. In case (3), the intersection of two circles, whose
equations are
x2 1 y2 1 d1x 1 e1y 1 f 1 5 0
and x2 1 y2 1 d2x 1 e2y 1 f 2 5 0,
is same as the intersection of the first circle and the line whose equation is
which reduces to case (2). In any case, the new constructed numbers lie in a
field Q( a ) for some 0 , a Q. If E is the smallest field containing those
real numbers constructed so far, the above argument shows that the next new
number constructed, lie in a field E ( a ) for some a H. Therefore, starting
from Q, we get
In a similar way, we get Q2 from Q1 and so on. In this way, we get an ascend-
ing chain of subfields of R
Q 5 Q0 ⊆ Q1 Q2 ⊆ …
It follows from all these that, if a real number a is constructible from Q, then
there exists an ascending chain of subfields of R
Q 5 Q0 ⊆ Q1 ⊆ Q2 ⊆ … ⊆ Qm
In fact, the converse of the above result is also true. But before this, we first
prove the following theorem.
OQ OQ OP 1
.
a OA OQ OQ
P O A
Proof: Using the fundamental theorem of Galois theory, we can get a1, a2,
…, ar such that
where [Q(a1, …, ai) : Q(a1, …, ai21)] 5 2 and a1, a2, …, ar are constructible.
Therefore, a is constructible. b
Theorem 16.5.5. Doubling the cube is impossible; that is, when a cube is
given, it is not always possible to construct with ruler and compass alone the
side of a cube which has double the volume of the given cube.
Proof: Without loss of generality, we can assume that the side of the given
cube is of unit length and hence its volume is 1. Let the side of the cube to be
constructed, if possible, be x. Then, x3 5 2 (double the volume of the given
cube) and hence x3 2 2 5 0. Therefore, we have to construct the number 21/3.
But x3 2 2 is an irreducible polynomial over Q and hence [Q(21/3) : Q] 5 3
which is not of the form 2n. Thus, by Corollary 16.5.2, 21/3 is not constructible
by ruler and compass alone. b
Proof: Without loss of generality, we can assume that the given circle has
radius 1 and hence its area is p. We need to construct a square of side p .
But p is not algebraic over Q (the proof of this is not given in this book) and
hence p is not algebraic over Q. Therefore, [Q( p ) : Q] is not finite. Thus,
p is not constructible. b
Theorem 16.5.7. Trisecting the angle is impossible; that is, there exists an
angle which cannot be trisected with ruler and compass alone.
1
8a3 2 6a 2 1 5 0 (since cos 60 5 ).
2
C
�
O B
Now, the polynomial 8x3 2 6x 2 1 is irreducible over Q and has a root a and
therefore
This implies that a is not constructible. Thus, 20° is not constructible and
hence 60° cannot be trisected. b
sin 4u 5 2sin 3u
a3 1 a2 2 2a 2 1 5 0.
x3 1 x2 2 2x 2 1
EXERCISES 16
1. Prove that the identity map is the only automorphism of the field R of real num-
ber and hence the Galois group of R/Q is trivial.
3. Find the Galois group of the splitting field of the polynomial x4 2 x2 1 1 over Q.
5. Prove that the polynomial x7 2 10x5 1 15x 15 is not solvable by radicals over Q.
7. Prove that it is impossible to construct a regular 9-gon using ruler and compass
alone.
8. Prove that a regular 17-gon is constructible with ruler and compass alone.
2. (i) {0, 1, 2, 3, 4, 5}
(ii) {(21, 0) ∪ (0, 1)} ∩ Q
(iii) {, {a}, {c}, {a, c}}
(iv) {012, 013, …, 019, 023, …, 029, …, 089, 123, 124, …, 129,
134, …, 139, …}
(v) {(0, 0)}
(vi) , the empty set.
4. X 5 {1, 2, 3}, P(X) 5 {, {1}, {2}, {3}, {1, 2}, {2, 3}, {1, 3}, X}
6. X 5 {1, 2, 3, …, 100}
A 5 {1, 4, 9, 16, 25, 36, 49, 64, 81, 100} and B 5 {1, 3, 5, …, 99}
Ci 5 {i, i 1 1, i 1 2, i 1 3, i 1 4} for each 1 i 96
(i) A ∩ B 5 {1, 9, 25, 49, 81}
(ii) A ∪ B ∪ C2 5 {1, 2, 3, 4, 5, 6, 7, 9, 11, 13, 15, 16, 17, …}
96
(iii) i∪
1
Ci X and X ∩ A A
25
(vi) X ( i∪
6
Ci ) {1, 2, 3, 4, 5, 95, 96, 97, 98, 99, 100}
(vii)
(viii) A 2 B 5 {4, 16, 36, 64, 100}
10. (i) A B 5 (A 2 B) ∪ (B 2 A) 5 (B 2 A) ∪ (A 2 B) 5 B A
(ii) (A B) C 5 [((A 2 B) ∪ (B 2 A)) 2 C] ∪ [C 2 ((A 2 B) ∪
(B 2 A))]
5…
5 (A ∩ B ∩ C) ∪ ((A 2 B) 2 C) ∪ ((B 2 C) 2 A) ∪
((C 2 A) 2 B)
The other equality is by symmetry.
12. (i) True; since A ⊆ X ∩ Y ⇔ A ⊆ X and A ⊆ Y.
(ii) False; for, if X 5 {1, 2} and Y 5 {3, 4}, then the set A 5 {2, 3}
belongs P(X ∪ Y), but A is in neither P(X) nor P(Y).
(iii) False; for the empty set P(X 2 Y), but P(X) 2 P(Y).
(iv) True; since X P(X) 5 P(Y) ⇒ X ⊆ Y.
EXERCISE 1(B)
2. (i) Clearly, A 5 B ⇒ A 2 C 5 B 2 C.
But A 2 C 5 B 2 C A 5 B; for, if A and B are any subsets of
C, then A 2 C 5 5 B 2 C.
(ii) A ∩ C 5 B ∩ C A 5 B; for, if C is any subset of A ∩ B, then A
∩ C 5 C 5 B ∩ C.
(iii) A ∪ C 5 B ∪ C A 5 B; for, if A and B are any subsets of C, then
A ∪ C 5 C 5 B ∪ C.
(iv) (A 2 B) 3 (C 2 D) ⊆ (A 3 C) 2 (B 3 D). The other inclusion
may not be true; for, take A 5 {1, 2}, B 5 {3, 4}, C 5 {2, 3}
and D 5 {2, 5}. Then, (1, 2) (A 3 C) 2 (B 3 D) and (1, 2)
(A 2 B) 3 (C 2 D), since 2 C 2 D.
(v) (A 3 C) ∪ (B 3 D) ⊆ (A ∪ B) 3 (C ∪ D). The other inclusion
may not be true; for, in (iv) above, (2, 5) (A ∪ B) 3 (C ∪ D), but
(2, 5) (A 3 C) ∪ (B 3 D).
4. For each a A, there are exactly m b’s in B such that (a, b) A 3 B.
Since A is n elements, A 3 B has nm elements. Also, the number of rela-
tions from A to B is 2nm.
6. The number of functions from A into B is mn.
8. If f (a) 5 f (b), then a 5 b (since f |{a, b} is an injection). Therefore, f is
an injection if f |Z is an injection for any subset Z of X. The converse is
clear. This statement is not valid for surjections. The map f : R → R1 ∪ {0}
30. (i) ⇒ (ii): If f has a left inverse, then f is an injection (by Ex. 14 of
Exercise 1(B)). Suppose that f (X) Y. Choose x1 x2 X and define
x if y f ( x ), x X
g1, g2 : y → x by g1 ( y )
x1 if y f ( X )
x if y = f (x ), x X
and g 2 (y ) .
x2 if y f (X )
Then, g1 g2 (since Y 2 f (X) ) and g1 o f 5 IX 5 g2 o f, which is a
contradiction to the uniqueness of the left inverse of f.
(ii) ⇒ (i) and (ii) ⇒ (iii) are clear.
(iii) ⇒ (ii): If f has a right inverse, then f is a surjection. (by Ex.
14 of Exercise 1(B)). Suppose a and b X such that f (a) 5
f (b) 5 y0, say. For each y y0 in Y, choose xy X such that f (xy) 5 y.
Define h1, h2 : Y → X by
x y if y y0 x y if y y0
h1 (y ) and h2 (y ) .
a if y y0 b if y y0
EXERCISE 1(C)
4. Easy verification.
14. {Q 1 r : r R}
(i)
{Z 1 r : r Q}
(ii)
(iii)
R is not reflexive and hence not an equivalence relation.
(iv){{(a, b) R2 : a2 1 b2 5 r} : 0 r R}.
(v)is not equivalence relation (it is not symmetric).
(vi)is not reflexive, since (1, 1) R
(vii)is not reflexive, since (2¼, 2¼) R
(viii)is not transitive.
(ix){{A B : B F} : A P(X)}, where F is the class of all finite
subsets of X.
(x) {rQ1 : r R*}
EXERCISE 1(D)
18. If A and T are the sets of algebraic numbers and transcendental numbers,
respectively, then A ∪ T 5 R. If T is countable, than A ∪ T (5 R) is
countable, since A is countable.
20. The given set is bijective with Q 3 Q Z1.
CHAPTER 2
EXERCISE 2(A)
12. Let a p1r1 p2r2 ptrt and b p1s1 p2s2 ptst , where pi’s are distinct primes
and ri 0, si 0.
14. Suppose that n > 1 and n is not a prime. Let p be the least prime dividing
n. Then, pm 5 n for some m > 1. Any prime dividing m should divide n
and hence p m. Now, p2 pm 5 n and hence p n .
16. Follows from the fact that a positive integer n divides a and b if and only
if n divides b and d (since a 5 bc 1 d and a 2 bc 5 d).
t
a
c ∏ piri si
(a, b) ik1
k
b
and d ∏ pisi ri .
(a, b) i1
EXERCISE 2(B)
4. Use Theorem 2.2.8 for (i) and (ii) and, for others, use Theorem 2.2.9
(note that m7 5 5, m11 5 10, m17 5 12, m19 5 2, m23 5 7 and m29 5 3).
6. (i) For any i j {0, 1, …, n21}, i j (mod n) and, since ai i 1
nZ, ai i (mod n) and hence ai aj (mod n).
(ii) Clear.
(iii) For any a Z, choose i such that a 5 qn 1 i, 0 i n 2 1 and
then a i ai (mod n) and hence a ai (mod n).
8. Let m, m 1 1, …, m 1 (n 2 1) be n consecutive integers, where m Z.
Choose 0 i n 2 1 such that m i (mod n). Then, m 1 1 i 1 1
(mod n), …, m 1 k i 1 k (mod n) for all k. This implies that {m, m 1
1, …, m 1 (n 2 1)} is a transversal for congruence modulo n.
10. If n 5 a1a2 … ar and m 5 as(1)as(2) … as(r), where s is any rearrangement
r r
of {1, 2, …, r}, then 3 divides n ⇔ 3 divides ai as (i ) ⇔ 3 divides
i1 i1
m.
12. 3 divides 12x 527846531 ⇔ 3 divides 1 1 2 1 x 1 5 1 2 1 7 18 1 4
1 6 15 1 3 11
⇔ 3 divides 44 1 x
⇔ x {1, 4, 7}.
EXERCISE 2(C)
EXERCISE 2(D)
2. If a < b, then b < a is not possible. The number of pairs (a, b) with a <
n( n1)
b can be atmost n 2n 2 . Therefore, the number of partial orders
2
n ( n1)
on X can be at most 2 2 .
4. For any a, b X, the set {a, b} has least element, that is, a b or b a.
6. Let each (Xi, ) be well ordered. Consider
A⊆X5X 3X 3…3X. 1 2 n
EXERCISE 2(E)
Let A 5 (aij), B 5 (bij) and C 5 (cij) and suppose that A(BC) (and hence
(AB)C) is defined. For any 1 i m and 1 j u,
n n
the ijth entry in A( BC ) ∑ aip ∑ bpq cqj
p1 q1
n n
∑∑ aip (bpq cqj )
q1 p1
n n
∑ ∑ aip bpq cqj
q1
p1
∑s a
r1
ir rj aaij , for any 1 i, j n.
where bis and asj are the isth and sjth entries in Bt and At, respectively.
(ii) If B1 and B2 are such that AB1 5 I and B2A 5 I, then B1 5 IB1 5
(B2A)B1 5 B2(AB1) 5 B2I 5 B2.
(iii) (AB)(B21A21) 5 A(BB21)A21 5 AIA21 5 AA21 5 I and, similarly
(B21A21)(AB) 5 I.
(iv) If AB 5 I 5 BA, then AtBt 5 (BA)t 5 It 5 I 5 BtAt and hence
(At)21 5 Bt 5 (A21)t.
10. Straight-forward verification.
EXERCISE 2(F)
det A det 2 1
1 xn x1 xn ( xn 1) xn ( xn x1 )
n2
and now expand with respect to the 1st row, factor out xj 2 x1 and apply
induction on n.
6. A2 5 A ⇒ (det A)2 5 det A ⇒ det A 5 0 or det A 5 1. If A is nonsingular,
then det A 0.
8. Am 5 On3n ⇒ (det A)m 5 0 ⇒ det A 5 0.
10. Follows from Theorems 2.6.2 and 2.6.8.
12. By Ex. 11, det A 5 2det A and hence det A 5 0.
14. (i) det(AB) 5 det A det B 5 det B det A 5 det(BA).
(ii) det(A At) 5 det A det At 5 det A det A.
(iii) det(ABA21) 5 det A det B det A21
5 det A det B (det A)21 5 det B.
CHAPTER 3
EXERCISE 3(A)
* a b c d e f
a d e a c a b
b e f d c f d
c a d c a d c
d c c a e d b
e a f d d b d
f b d c b d a
6. Let S3 be the set of bijections of {1, 2, 3} onto itself and o be the com-
position of mappings in S3. Then, o is associative, but not commutative;
for consider f and g defined by f (1) 5 1, f (2) 5 3, f (3) 5 2 and g(1) 5
2, g(2) 5 1 and g(3) 5 3. Then, (f o g)(1) 5 3 and (g o f )(1) 5 2 and
hence f o g g o f.
8. No, the statement is false. Consider the operation * on {a, b} given by
the table
* a b
a b a
b a a
Here, (a * a) * b 5 b * b 5 a
and a * (a * b) 5 a * a 5 b.
2
12. The number of binary operations on an n-element set is n2n . The number
n +n
of commutative binary operations on an n-element is n 2 .
∴ The number of noncommutative binary operations on an n-element
2 n2 +n
set is nn n 2
and, on a 3-element set, it is 39 2 36 5 36 3 26.
14. By definition of 1, e 1 x 5 x 5 x 1 e for all x S9.
EXERCISE 3(B)
EXERCISE 3(C)
2. Let S be any finite set with more than one element and define x * y 5 y
for all x and y S. Then, (S, *) is a finite semigroup which satisfies the
left cancellation law and does not satisfy the right cancellation law.
4. Imitate the proof of Theorem 3.3.2.
6. Let (S, ) be a finite semigroup and a S. Then, {an : n Z1} ⊆ S and
hence there exist m < n in Z1 such that am 5 an. Put a 5 n 2 m Z1.
Then,
an 5 am1a 5 am
a2m 5 am am 5 am am1a 5 a2m1a
By induction on k, akm 5 akm1a for all k Z1. Also, akm12a 5 akm1a aa 5
akm aa 5 akm1a 5 akm.
By induction on r,
8. Let G be a finite group and |G| 5 2m. For any a G, let Aa 5 {a, a21}.
Then, |Aa| 2 for all a G. Also, Aa ∩ Ab 5 or Aa 5 Ab for each
pair of all elements a and b in G. {Aa : a G} is a partition of G and
hence
2m 5 |G| 5 ∑|Aa|.
Since Ae 5 {e}, there must be atleast one more a e such that |Aa| 5 1;
that is, a 5 a21 or a2 5 e and a e.
14. If |G| 5 1, then G 5 {e}. If |G| 5 2, then G 5 {e, a}, where a2 5 e. Let
|G| 5 3. If e a G, then a2 e (otherwise, a2 5 e implies that there
exists b G 2 {e, a} such that ab, b, e, a are distinct elements of G)
and hence {e, a, a2} 5 G. Let |G| 5 4. If a2 5 e for all a G, then a 5
a21 for all a G and hence ab 5 (ab)21 5 b21a21 5 ba for all a and b
G. Therefore, we can assume that there exists a G such that a2 e.
If a3 5 e, then there exists b G 2 {e, a, a2} such that ba is an
element of G other than e, a, a2 and b so that |G| > 4. Thus, a3 e
and hence G 5 {e, a, a2, a3}, which is abelian. Next, let |G| 5 5. We
can prove that, for any e a G, G 5 {e, a, a2, a3, a4} which is
abelian.
EXERCISE 3(D)
2.
11 1 2 3 4 5 6 7 8 9 10
1 1 2 3 4 5 6 7 8 9 10
2 2 4 6 8 10 1 3 5 7 9
3 3 6 9 1 4 7 10 2 5 8
4 4 8 1 5 9 2 6 10 3 7
5 5 10 4 9 3 8 2 7 1 6
6 6 1 7 2 8 3 9 4 10 5
7 7 3 10 6 2 9 5 1 8 4
8 8 5 2 10 7 4 1 9 6 3
9 9 7 5 3 1 10 8 6 4 2
10 10 9 8 7 6 5 4 3 2 1
4.
14 0 1 2 3
12 0 1 2 0 0 1 2 3
12 0 1 0 0 1 2 1 1 2 3 0
0 0 1 1 1 2 0 2 2 3 0 1
1 1 0 2 2 0 1 3 3 0 1 2
15 0 1 2 3 4
0 0 1 2 3 4
* e a b c
1 1 2 3 4 0
e e a b c
2 2 3 4 0 1
a a e c b
3 3 4 0 1 2
b b c e a
c c b a e 4 4 0 1 2 3
CHAPTER 4
EXERCISE 4(A)
(v) Yes
(vi) Yes
(vii) No
(viii) Yes.
EXERCISE 4(B)
Then,
20. {0}, <2>, <3>, <4>, <6>, <8>, <12> and Z24 are all the subgroups of Z24.
{0}, <2>, <3>, <5>, <6>, <10>, <15> and Z30 are all the subgroups
of Z30.
22. {0}, <5>, <25>, <125> and Z625 are all the subgroups of Z625.
24. Any infinite cyclic group has infinitely many subgroups. Therefore, if G
has finitely many subgroups, then <a> is finite for any a G and, since
G ∪ < a >, G is finite.
a∈G
26. O(a) 5 O(a21) for any a G. Also, if O(a) and O(b) divide n, then
O(ab) divides l.c.m. {O(a), O(b)} and hence O(ab) divides n.
EXERCISE 4(C)
EXERCISE 4(D)
EXERCISE 4(E)
EXERCISE 4(F )
|G|
2. A 5 {e, a3, a6, a9, a12}. G 3. Let A1 5 a5A and A2 5 a7A. Then,
G { A, A , A }. A | A|
A 1 2
A A1 A2
A A A1 A2
A1 A1 A2 A
A2 A2 A A1
CHAPTER 5
EXERCISE 5(A)
2. If f (ab21) 5 f (a)f (b)21 for all a, b G, then f (e) 5 f (aa21) 5 f (a)f (a)21
5 e9, f (b21) 5 f (eb21) 5 f (e)f (b)21 5 e9f (b)21 5 f (b)21 and f (ab) 5
f (a(b21)21) 5 f (a)(f (b)21)21 5 f (a)f (b) for all a and b G. The converse
is clear.
4. G is abelian ⇔ ab 5 ba for all a and b G
⇔ (ab)21 5 b21a21 5 a21b21 for all a, b G
⇔ f (ab) 5 f (a)f (b)
6. For any homomorphism f : Z → Z2, f (n) 5 nf (1) for all n Z and
hence
f (1) 5 0 ⇒ f (n) 5 0 for all n Z
and f (1) 5 1⇒ f (n) 5 0 for all even n and
f (n) 5 1 for all odd n
(since 1 1 1 5 0 in Z2)
8. Let f : Z → Z be a nontrivial endomorphism. Then, f (n) 5 nf (1) for all
n Z and hence f (1) 0 so that
f (n) 5 f (m) ⇒ nf (1) 5 mf (1) ⇒ n 5 m.
14. If |G| 5 p, a prime, then |ker f | divides |G| 5 p and hence |ker f | 5 1 or
p, so that ker f 5 {e} or G.
16. f is a homomorphism ⇒ aabb 5 f (1 1 1, 1 1 1) 5 f (1, 1)f (1, 1) 5 abab
⇒ ab 5 ba
f is a homomorphism ⇔ ab 5 ba.
18. Use induction on n to prove that f (n 1 m) 5 f (n)f (m) (that is, an1m 5
anbm) for all n, m Z. ker f 5 {0} if a is of infinite order and ker f 5 nZ
if O(a) 5 n.
20. If f : Q8 → Z2 is a homomorphism such that f (i) 5 0 and f (j) 5 1, then
f (a) can be evaluated for any a Q8.
22. Follows from the fact that ab > 0 ⇔ (a > 0 and b > 0) or (a < 0 and b <
0) for any a, b R 2 {0}. Also, for any x, y G, xy 5 1 ⇔ x 5 1 5 y
or x 5 21 5 y.
24. If f : Z8 → Z24 is defined by f (a) 5 3a for all a Z8, then f is a mono-
morphism.
26. If f is an isomorphism of (Z, 1) onto (Z, 1), then f (n) 5 nf (1) for all
n Z and f (1) 5 1 or 21 (otherwise f is not surjective) and hence f is
either identity map or the inverse map.
28. (f 1 g)(x) 5 f (x) 1 g(x) for all x. The trivial homomorphism is the
identity in Hom(G, H). Also, if f Hom(G, H), then 2f Hom(G, H)
and 2f is the inverse of f.
30. g o f is an injection ⇒ f is an injection.
g o f is a surjection ⇒ g is a surjection.
EXERCISE 5(B)
EXERCISE 5(C)
4. Define a(A) 5 f21(A) for any subgroup A of G9 and b(B) 5 f (B) for any
subgroup B of G containing ker f. Then, (a o b)(B) 5 B and (b o a)(A) 5
A and hence a is a bijection of the set of subgroup of G9 onto the set of
subgroups of G containing ker f.
6. Refer Example 3.2.8. Let a be the clock-wise rotation about the centre
of the square through an angle p/2 and b be the reflection about the
diagonal D1. Then, O(a) 5 4, O(b) 5 2 and aba 5 b. The group G8 is
of order 8 and G8 5 {e, a, a2, a3, b, ab, a2b, a3b}. It can be verified that
N1 5 {e}, N2 5 {e, a, a2, a3}, N3 5 {e, a2, b, a2b}, N4 5 {e, a2, ab, ab,
a3b}, N5 5 {e, a2} and N6 5 G8 are all the normal subgroups of G8 and
hence G8/Ni, 1 i 6, are all (up to isomorphism) the homomorphic
images of G8. Also, Z(G8) 5 {e, a2} 5 N5 and N2, N3, N4, N5 and G8 are
all subgroups of G8 containing Z(G8) and hence N2/N5, N3/N5, N4/N5, N5/
N5, and G8/N5 are all the subgroups of G8/Z(G8).
EXERCISE 5(D)
CHAPTER 6
EXERCISE 6(A)
1 2 3 4 5 6 7 8
8. a b a1
7 4 6 5 3 1 8 2
1 2 3 4 5 6 7 8
b c b1
5 6 7 3 2 1 8 4
1 2 3 4 5 6 7 8
c a c1
4 1 8 7 2 5 3 6
1 2 3 4 1 2 3 4 1 2 3 4 1 2 3 4
10. e, , , , ,
2 1 4 3 4 3 2 1 3 4 1 2 2 3 1 4
1 2 3 4 1 2 3 4 1 2 3 4 1 2 3 4
S4 , , , ,
3 1 2 4 2 4 3 1 4 1 3 2 3 2 4 1
1 2 3 4 1 2 3 4 1 2 3 4
,
,
4 2 1 3 1 3 4 2 1 4 2 3
EXERCISE 6(C)
2. (i) Yes, (1 4 2 6 5 3 8 7)
(ii) No
(iii) Yes, (1 2 3 4 5 6 7 8)
(iv) No
4. (i) 12
(ii) 2
(iii) 6
(iv) 7.
6. Let |f | 5 Supp f and f and g be disjoint.
12. Follows from the observation that i Supp f if and only if f (i) Supp
f and that f is a 3-cycle if and only if Supp f 5 {i, f (i), f2(i)} for some
i. This cannot be extended for any r-cycle. For example, if f 5 (1 2) o
(3 4), then Supp f 5 {1, 2, 3, 4}, which is a 4-element set, but f is not a
4-cycle.
14. (g f g21)n 5 g f n g21 and hence f n 5 e ⇔ (g f g21)n 5 e for any n Z1.
16. If a 5 (i1 i2 i3 … i2m11), then
22. Partitions of 4 are {1, 1, 1, 1}, {1, 1, 2}, {2, 2}, {1, 3}, {4} and of 5 are
{1, 1, 1, 1, 1}, {1, 1, 1, 2}, {1, 2, 2} {1, 1, 3} {2, 3}, {1, 4}, {5}.
30. Let |G| 5 6. If G is abelian, then G Z6. Suppose that G is not abelian.
Then, O(a) 5 3 for some a G (O(a) 5 2 for all a ⇒ G is abelian and
O(a) 5 6 for some a ⇒ G Z6). Choose b G 2 {e, a, a2}. Then,
we can check that b2 a and b2 a2. Also, b2 {b, ab, a2b}. Since e,
a, a2, b, ab, a2b are all distinct elements, we have G 5 {e, a, a2, b, ab,
a2b}. Therefore, b2 5 e and O(b) 5 2. Since <a> is of index 2, <a> is
normal in G and hence b a b21 5 e, a or a2. If b a b21 5 e, then a 5 e
and, if b a b21 5 a, then ba 5 ab and G is abelian. Therefore, b a b21 5
a2. Thus, G 5 <a, b>, a3 5 e 5 b2, b a b 5 a2. Also, if f 5 (1 2 3) and
g 5 (1 2), then f3 5 e 5 g2 and g f g 5 f2 in S3. If a : S3 → G is defined
by, a(f ) 5 a, a(g) 5 b, a(f2) 5 a2, a(e) 5 e, a(fg) 5 ab and a(f2g) 5
a2b, then a is an isomorphism.
EXERCISE 6(D)
CHAPTER 7
EXERCISE 7(A)
2. Yes, u is effective.
4. ker u 5 {a G : axa21H 5 xH for all x G}, which is not necessarily
{e} and hence u is not necessarily effective.
EXERCISE 7(B)
6. Let (G) be the set of all subgroups of G and let G act on (G) by
conjugation (that is, u(a, K) 5 aKa21). Then, the orbit of H is the set of
conjugates of H in G and the stabilizer of H is the normalizer of H in G.
Now use Corollary 7.2.1.
8. We can assume that the action is nontrivial. By Theorem 7.2.7, St(x) is a
maximal subgroup of G for each x X. If N ⊆ St(x) for all x X, then the
induced action of N on X is trivial. Suppose that N St(x) for some x X.
Then, since N is normal NSt(x) is a subgroup containing St(x) properly and
hence NSt(x) 5 G. Now, for any y X, there exists g G such that gx 5
y (since the action of G on X is primitive, it is transitive by Corollary 7.2.2).
EXERCISE 7(C)
EXERCISE 7(D)
2. (i) <3> 5 {0, 3, 6, 9, 12, 15, 18, 21} is the only Sylow 2-subgroups.
<8> 5 {0, 8, 16} is the only Sylow 3-subgroup {0} is the only
Sylow p-subgroup for prime p 2, 3.
(ii) Refer Worked Exercise 6.2.1. {e, a, b} is the only Sylow 3-sub-
group and {e, c}, {e, d} and {e, s} are all the Sylow 2-subgroups
{e} is the Sylow p-subgroup for any prime p 2, 3.
(iii) Refer Worked Exercise 6.4.2. <a1>, <b1>, <c1> and <d1> are the
only Sylow 3-subgroups. {e, p, q, r} is the only Sylow 2-subgroup.
4. Let |G| 5 56. Then, n7 5 1 1 7k should divide |G| 5 23 ? 7 and hence n7
should divide 23, so that n7 5 1 or 8. Any subgroup of order 7 is cyclic
and hence generated by each of the six nonidentity elements. If n7 5
8, then there are 6 ? 8 5 48 elements of order 7 and the remaining 8
CHAPTER 8
EXERCISE 8(A)
EXERCISE 8(B)
EXERCISE 8(C)
CHAPTER 9
EXERCISE 9(A)
2. (i) 14
(ii) 3
4 4
(iii)
4 4
(iv) 78
(v) 0
(vi) {2}.
EXERCISE 9(B)
EXERCISE 9(C)
2. (a 1 b)3r 5 a3r 1 b3r (since 3 divides 3rCs for any 0 < s < 3r).
4. If R is a finite ring, then O(a) is finite in the group (R, 1) and, if m is the
l.c.m. {O(a) : a R} then ma 5 0 for all a R and hence char(R) is
finite.
6. Z5.
8. (a 1 b)p 5 ap 1 bp 5 a 1 b and (a b)p 5 apbp 5 ab. Also, a A ⇒ a
5 ap ⇒ a(1 2 ap21) 5 0 ⇒ a 5 0 or ap21 5 1 ⇒ a 5 0 or a is a unit.
10. Direct verification.
12. Let ab 5 1 and R 5 {a1, a2, …, an}. Consider Ra 5 {a1a, a2a, …, ana}
⊆ R. For any i and j,
aia 5 aja ⇒ (ai 2 aj)a 5 0 ⇒ ai 2 aj 5 (ai 2 aj)ab 5 0
Therefore, Ra and R have the same number of elements and hence R 5
Ra. In particular, 1 5 ca for some c R. Now, b 5 1b 5 (ca)b 5 c(ab) 5
c and hence ba 5 1.
EXERCISE 9(D)
2. Direct verification.
4. S is a subring of (Zn, 1n, n) ⇔ S is a subgroup of (Zn, 1n) ⇔ S 5 <m>
for some divisor m of n.
6. Let A 5 {x1x2 … xn : n > 0 and x1, x2, …, xn X}.
Then, (X) 5 {a1 1 a2 1 … 1 an : ai or 2ai A}.
8. Let X be an infinite set and R be the set of all finite subsets of X. Then,
(R, 1, ∩) is a ring without unity. For any finite subset A of X, P(A) is a
subring of R and A is the unity in P(A). (Here, 1 is the symmetric differ-
ence of sets.)
10. S ∪ T is a subring ⇒ S ∪ T is a subgroup of (R, 1) ⇒ S ⊆ T or T ⊆ S.
EXERCISE 9(E)
EXERCISE 9(F )
12. The set of all scalar matrices; that is, matrices A 5 (aij) such that aij 5 0
for all i j and aii 5 ajj for all i and j.
14. R
16. If A 5 (aij) is such that BA 5 B 5 AB for all B Mn(R), then prove that
aij 5 1 or 0 according as i 5 j or i j.
EXERCISE 9(G)
2. (i) R 3 R is not integral domain and hence not a field, since (1, 0)
(0, 1) 5 (0, 0)
(ii) Qn is not an integral domain, not a field.
CHAPTER 10
EXERCISE 10(A)
(viii) Z4 has exactly three ideals, namely {0}, {0, 2} and Z4.
4. If R is a division ring, I is a nonzero ideal of R and 0 a I, then 1 5
aa21 I and hence I 5 R, so that R has only two ideals. The converse is
not true; for, M2(R), the set of 2 3 2 matrices forms a ring with exactly
two ideals, but it is not a division ring.
22. Let I be an ideal in Mn(R) and J be the set of all 11th (first row and first
column) entries in members of I. Then prove that I 5 Mn(J), by estab-
lishing that, if A 5 (aij) and Eij’s are the matrix units, then
EXERCISE 10(B)
16. The ideals of Z/nZ are of the form dZ/nZ, where d is a divisor of n and
the homomorphic images of Z/nZ are of the form Z/dZ, where d is a
divisor of n.
EXERCISE 10(C)
2. Direct verification.
EXERCISE 10(D)
16. If m > 1 and m2 divides n, then m2r 5 n for some r $ 1. Now, for any a
R, (mra)2 5 m2r2a 5 n(ra) 5 0 and hence mra N(R) 5 {0}, so that
mra 5 0, which is a contradiction to the least property of n 5 char(R).
EXERCISE 10(E)
2. 9Z
4. Use Ex. 3 and the fact that a nonzero ideal of Z is prime if and only if it
is maximal.
6. Let X 5 {x1, x2, …, xn} and M be a maximal ideal in (P(x), 1, ∩). Then,
n
{xi } X M and hence {xi} M for some i and, since {xi} ∩ {xj} 5
i1
EXERCISE 10(F)
CHAPTER 11
EXERCISE 11(A)
EXERCISE 11(B)
2. (i) 2
(ii) 1
(iii) 0
(iv) 4
(v) 4
4. The Kernel of fa is an ideal of F[x] and any nonzero ideal of F[x] is infinite.
6. (1 1 4x)(1 1 4x) 5 1 in Z8[x].
8. f (a) 5 g(a) ⇔ a is a root of the polynomial f (x) 2 g(x).
10. Use Fermat’s Theorem 4.4.7.
EXERCISE 11(C)
EXERCISE 11(D)
CHAPTER 12
EXERCISE 12(A)
2. (i) 1, 21
(ii) 1, 21, i, 2i
(iii) 61, 6 (1 2), 6 (1 2)
(iv) 61
(v) R 2 {0}
(vi) 1, 2, 3, 4
4. ( 2 3 )( 2 3 ) 1 and ( 2 3 ) 3 3 2 3
6. 1 1 2x 1 3x2, 2 1 4x 2 x2, 3 2 x 1 2x2, 23 1 x 1 5x2, 22 1 3x 1 x2,
21 22x 2 3x2
8. Use induction on n.
10. Let a 5 bu, where u is a unit. Then, a 5 bc ⇒ bu 5 bc ⇒ u 5 c.
EXERCISE 12(B)
EXERCISE 12(C)
i1
n
and [a, b ] ∏ pi max{r ,s }
i i
i1
EXERCISE 12(D)
p ( p 1) p ( p 1)
p 2 p .
2 2
EXERCISE 12(E)
EXERCISE 12(F)
CHAPTER 13
EXERCISE 13(A)
2. (i), (ii), (iii), (iv) are R-submodules of R4 and others are not.
4. Z is a faithful Z-module and, for any nonzero proper ideal I of a
ring R with unity, R/I is an R-module, which is not faithful, since
Ann(R/I) 5 I.
6. Follows from P ⊆ P 1 Q and P ∩ Q ⊆Q.
8. Let M 5 Z2 3 Z2. Then, M is a Z-module. Let N 5 {(0, 0), (1, 1)},
P 5 Z2 3 {0} and Q 5 {0} 3 Z2. Then, N, P, Q are Z-submodules of M.
EXERCISE 13(B)
EXERCISE 13(C)
�
�i
g
gi
EXERCISE 13(D)
EXERCISE 13(E)
EXERCISE 13(F )
2. Let e4 5 (0, 0, 0, 1, 0) and e5 5 (0, 0, 0, 0, 1). Then, {e1, e2, e3, e4, e5} is
a basis of F5.
0 0
0 0 0 0
0 0 0 0
1 0 0 0 0 1 0 0 0 0
0 2 0 0 0 and 1 2 0 0 0.
0 0 3 0 0 1 1 3 0 0
0 0 0 4 0 1 1 1 4 0
1 0 0 0 0 0
1 0 0 0
1 1 0 0 0 1 1 0 0 0
0 1 1 0 0 and 1 1 1 0 0.
0 0 1 1 0 1 1 1 1 0
0 0 0 1 1 1 1 1 1 1
CHAPTER 14
EXERCISE 14(A)
2. [K : Q] 5 4, since [K : Q( p ) ] 5 2 and [ Q( p ) : Q] 5 2
EXERCISE 14(B)
8. F ⊆ K ⊆ E ⇒ [E : K][K : F] 5 (E : F ) 5 p, a prime
⇒ [E : K] 5 1 or [K : F] 5 1
⇒ K 5 E or K 5 F.
EXERCISE 14(C)
∞
2. Let F ∪ Fn . a and b [ F ⇒ a [ Fn and b [ Fm for some n and m ⇒
n1
a and b [ Fn or Fm, according as m # n or n # m ⇒ a 6 b and a ? b [
Fn ⇒ a 6 b and a ? b [ F. Thus, F is a field and Fn is a subfield of F for
each n.
6. Use the argument given in Ex. 2 above and the fact that, for any a and b
[ , Fa ⊆ Fb or Fb ⊆ Fa.
EXERCISE 14(D)
2. Use induction on n.
4. Use induction on m.
6. Use Theorem 14.4.4(2).
EXERCISE 14(E)
CHAPTER 15
EXERCISE 15(A)
EXERCISE 15(B)
EXERCISE 15(C)
E
G m
E Q
G m1 = 2m1
Q Em
G
E
m1
Now, choose a subgroup Hm22 of order 2m22 in G(Em21/Q) and let Em22
be the fixed field of Hm22. This process can be continued to construct the
required fields Q 5 E0 ⊂ E1 ⊂ … ⊂ Em21 ⊂ Em 5 Q(a).
6. Note that the only nonidentity automorphism in G(Q(a)/Q) is the f
for which f (a) 5 a2 (and hence f (a2) 5 a, since a3 5 1). Therefore,
G(Q(a)/Q) is the two element group, which has only two subgroups,
namely {Id} and the whole group and the corresponding fixed fields are
Q(a) and Q, respectively.
CHAPTER 16
EXERCISE 16
A D
algebraic extensions, 14-8–14-19 Dedikind theorem, 15-11–15-12
algebraically closed fields, determinants, 2-43–2-54
14-20–14-26 dihedral group, 6-23–6-36
alternating group, 6-23–6-36 direct products, 8-1–8-10
automorphism groups, 15-10–15-18 disjoint set, 1-7
fixed fields and, 15-10–15-18 division algorithm, 4-13–4-14,
automorphisms, 5-29–5-36 11-15–11-23
B E
binary systems, 3-3–3-16 endomorphisms, 5-29
Burnside’s theorem, 7-23–7-24 equivalence relations, 1-21-1-25
partitions and, 1-21–1-25
C Euclidean division algorithm, 12-36
Euclidean domain, 12-36–12-43
canonical homomorphism, 5-5,
Euler’s theorem, 4-34
10-23
Euler–Totient function, 4-19
cardinality of sets, 1-27–1-35
even permutation, 2-44
Cartesian product, 1-11
extension fields, 14-3–14-38
Cauchy’s Theorem, 7-28–7-29
Cayley’s theorem, 6-1–6-6
Chinese remainder theorem, F
10-29–10-33 Fermat’s theorem, 4-34–4-35
choice function, 2-33 fields, 2-25
class equation, 7-21–7-22 integral domains and, 9-43–9-50
class of sets, 1-6 extensions of, 14-3–14-8
compass, 16-20 polynomials and, 11-25–11-29
completely reducible modules, finite fields, 14-33–14-38
13-31–13-41 finite groups, 3-45–3-50
congruence modulo, 2-14–2-21 finite set, 1-28
coordinate-wise ordering, 2-31 finitely generated abelian groups,
cycles, 6-11–6-20 8-10–8-27
cyclic extensions, 16-5–16-7 invariants of, 8-29–8-32
cyclic groups, 4-12–4-22 fundamental structure theorem for,
cyclotomic polynomial, 12-33 8-16–8-17