0% found this document useful (0 votes)
457 views19 pages

Groups Rings and Fields

Yes

Uploaded by

Bravine Mwaba
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
457 views19 pages

Groups Rings and Fields

Yes

Uploaded by

Bravine Mwaba
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 19

THE COPPERBELT UNIVERSITY

In Association With

TECHNICAL AND VOCATIONAL TEACHERS’ COLLEGE

DEPARTMENT OF APPLIED SCIENCES

BACHELOR OF SCIENCES IN MATHEMATICS AND SCIENCE WITH


EDUCATION

(Second Year)

Course: Linear and Abstract Algebra

Course Code: MAT 240

TOPIC 1: Groups, Rings and Fields

HANDOUT TWO

@ TVTC, 2023

1
Topic 1: Groups

The term group was used by Galois around 1830 to describe sets of one-to-one functions on
finite sets that could be grouped together to form a set closed under composition.

1.1.Binary Operation (*)

Let G be a set. A binary operation on G is a function that assigns each ordered pair of
elements of G an element of G

A binary operation on a set G, then, is simply a method (or formula) by which the members
of an ordered pair from G combine to yield a new member of G.

The most familiar binary operations are ordinary addition, subtraction, and multiplication of
integers.

Division of integers is not a binary operation on the integers because an integer divided by an
integer need not be an integer.

1.2.Group

Let G be a set together with a binary operation (usually called multiplication) that assigns to
each ordered pair (a, b) of elements of G an element in G denoted by a*b. We say G is a
group under this operation if the following four properties are satisfied.

1. Closure
For any two elements a and b, then a*b G for all a, b, c in G.
2. Associativity.
The operation is associative; that is, (a*b)*c = a*(b*c) for all a, b, c in G.
3. Identity.

There is an element e (called the identity) in G such that a*e = e*a = a for all a in G.

4. Inverses.
For each element a in G, there is an element a‘ in G (inverse of a) such that a*b = b*a = e
1.2.1. Abelian group

This is a commutative group where ab = ba for all a,b in G.

A group is non-Abelian if there is some pair of elements a and b for which ab ≠ ba for all a,b
in G.

The following are examples groups

The set of integers Z (so denoted because the German word for numbers is Zahlen) is a group
under ordinary addition. In each case, the identity is 0 and the inverse of a is – a.

1. Closure
For any two elements a and b, then a*b Z for all a, b in Z.

2
3 + 7 = 10

2. Associativity.
The operation is associative; that is, (a*b)*c = a*(b*c) for all a, b, c in Z.
4 + ( 6 + 7) = (4 + 6 ) + 7 = 17 Z
3. Identity.

There is an element e (the identity) in G such that a*e = e*a = a for all a in Z.

2+0=2 Z
4. Inverses.
For each element a in Z, there is an element b in Z (called an inverse of a) such
that a*b = b*a = e
6 + (– 6 ) = 0 Z
5. Commutativity
This is a commutative group where a*b = b*a for all a,b in Z.
4 + 8 = 8 + 4 = 12

TASK

Justify whether the following statements are groups or not with a reason.

i. The set of rational numbers Q (for quotient) ordinary addition.


ii. The set of integers under subtraction
iii. The set of real numbers R is under ordinary addition.
iv. The set of integers under ordinary multiplication.
v. The subset {1, – 1, i, – i} of the complex numbers under complex multiplication.
vi. A rectangular array of a 2 by 2 matrix under component wise addition
vii. The set R* of nonzero real numbers under ordinary multiplication.

ANSWERS TO TASK

i. A group
ii. Not a group, since the operation is not associative.
iii. A group
iv. Not a group, Since the number 1 is the identity, property 3 fails. For example,
there is no integer b such that 5b = 1.
v. A group, Note that – i) is its own inverse, ((– i)(– i) = 1) whereas the inverse of i
is – i,
vi. A group
vii. A group, the identity is 1. The inverse of a is 1/a.

3
Theorem 1.1: In a group G, there is only one identity element (Uniqueness of the Identity)

PROOF

Suppose both e and e‘ are identities of G.

Then,

1. ae = a for all a in G, . . . . . (1)


2. e‘a = a for all a in G. . . . . .(2)

The choices of a = e‘ in (part 1) and a = e in (part 2) yield e‘e = e‘ and e‘e = e. Thus, e
and e‘ are both equal to e‘e and so are equal to each other.

We use e because of German word for identity is Einheit.

Theorem 1.2: In a group G, the right and left cancellation laws hold; that is, b*a = c*a
implies b = c, and a*b = a*c implies b = c.

PROOF

Suppose b*a = c*a. Let a‘ be an inverse of a. Then multiplying on the right by a‘ yields
(b*a)a‘ = (c*a)a‘. Associativity yields b(a*a‘) = c(aa‘). Then b*e = c*e and, therefore, b
= c as desired. Similarly, one can prove that a*b = a*c implies b = c by multiplying by a‘
on the left.

Theorem1.3: For each element a in a group G, there is a unique element b in G such that

a*b = b*a = e.

PROOF

Suppose b and c are both inverses of a. Then a*b = e and a*c = e, so that a*b = a*c.
Cancelling the a on both sides gives b = c, as desired.

Theorem: For group elements a and b, (a*b)– 1 = b– 1 a – 1

Since (a*b)(a*b) – 1 = e and (a*b)(b – 1 *a– 1 ) = a(bb – 1)a – 1 = aea – 1 = aa– 1 = e

1.3.Finite Groups; Subgroups


1.3.1. Order of a Group
- The number of elements of a group (finite or infinite) is called its order. We will
use |G| to denote the order of G.
1.3.2 Subgroup
- If a subset H of a group G is itself a group under the operation of G, we say that
H is a subgroup of G

4
We use the notation H G to mean that H is a subgroup of G. If we want to indicate
that H is a subgroup of G but is not equ al to G itself, we write H < G. Such a
subgroup is called a proper subgroup

Theorem 1.4: H is a subgroup of G. Prove that the identity element of H is equal to the
identity element in G.

Proof:

Given that H is a subgroup of G.

Let us assume that e and e‘ be the two identity elements in H and G, respectively.

Let a ∈ H ⇒ a ∈ G [since H is a subset of G]

Identity element in group H = e

Thus, a ⋆ e = e ⋆ a = a…..(1)

Identity element in group G = e‘

Therefore, a ⋆ e‘ = e‘ ⋆ a = a…..(2)

From (1) and (2),

a ⋆ e = a ⋆ e‘

⇒ e = e‘

That means, the identity element in H is equal to the identity element in G.

Hence proved.

Theorem 1.5 : H is a subgroup of G. The inverse of any element in H is equal to the inverse
of the same element in G.

Proof:

Given that H is a subgroup of G.

Consider a ∈ H ⇒ a ∈ G

Let us assume that b and c are two inverse elements of a in H and G respectively.

Let b be the inverse element of a in H.

Then, a ⋆ b = b ⋆ a = e….(1)

Let c be the inverse element of a in G.

5
Then, a ⋆ c = c ⋆ a = e….(2)

From (1) and (2),

a⋆b=a⋆c

⇒b=c

That means the inverse element of a in H is equal to the inverse element of a in G.

Hence proved.

1.4.Cyclic Groups

A group G is called cyclic if there is an element a in G such that G = {an | n ∈ Z}. Such an
element a is called a generator of G.

e.g The set of integers Z under ordinary addition is cyclic. Both 1 and – 1 are generators.

Theorem 1.6: All subgroups of a cyclic group are cyclic. If G = ⟨a⟩ is cyclic, then for every
divisor d of |G| there exists exactly one subgroup of order d which may be generated by a|G|/d.

Proof: Let |G|=dn. Then 1,an,a2n,...,a(d−1)n are distinct and form a cyclic subgroup ⟨an⟩ of order
d. Conversely, let H={1,a1,...,ad−1 be a subgroup of G for some d dividing G. Then for all i,
ai=ak for some k, and since every element has order dividing |H|, adi=akd=1. Thus
kd=|G|m=ndm for some m, and we have ai=anm so each ai is in fact a power of an. From above
this means it must be one of the d subgroups already described.

Theorem 1.7

Consider the following subset of Z: H = {30x + 42y + 70z | x, y, z ∈ Z}.

Prove that H is a subgroup of Z.

Proof

0 = 30 · 0 + 42 · 0 + 70 · 0 ∈ H.

If 30x + 42y + 70z ∈ H, then

(30x + 42y + 70z) = 30(−x) + 42(−y) + 70(−z) ∈ H.

If 30a + 42b + 70c, 30d + 42e + 70f ∈ H, then

(30a + 42b + 70c) + (30d + 42e + 70f ) = 30(a + d) + 42(b + e) + 70(c + f ) ∈ H.

6
1.5.Permutation Groups
1.5.1. Permutation of A
A permutation of a set A is a function from A to A that is both one-to-one and onto.
1.5.2. Permutation Group of A
A permutation group of a set A is a set of permutations of A that forms a group under
function composition.
Example
1. If
α=[ ]
then
α(1) = 2 α(2) = 3 α(3) = 1 α(4) = 4
2.

β=[ ] and α=[ ]

Then
i. Composition of permutations expressed in array notation is carried out from
right to left by going from top to bottom, then again from top to bottom. For
example to find βα, we say since α is on the right – hand side, we get 1 from α
which yields 2 from same α, then this 2 yields from better β, which is on the
left, and the 2 from β yields a 3 from β

βα = [ ] [ ]=[ ]

ii.

αβ = [ ][ ]=[ ]

iii. β2 = β β = [ ][ ]=[ ]

1.6.Cycle Notation

Given a permutation below

α= [ ]

it can be expressed in cycle notation as follows

7
Theorem 1.8: Every permutation of a finite set can be written as a cycle or as a product of
disjoint cycles. (Products of Disjoint Cycles)

PROOF

Let α be a permutation on A = {1, 2, . . . , n}. To write α in disjoint cycle form, we start by


choosing any member of A, say a1 and let

a2 = α (a1), a3 = α (α (a1)) = α 2(a1),

and so on, until we arrive at a1 = αm(a1) for some m. We know that such an m exists because
the sequence a1, α (a1), α2(a1), . . . must be finite;

So there must eventually be a repetition, say α i(a1) = αj(a1) for some i and j with i < j.

Then a1 = α m(a1), where m = j – i. We express this relationship among a1, a2, . . . , am as

α = (a1, a2, . . . , am) . . .

The three dots at the end indicate the possibility that we may not have exhausted the set A in
this process. In such a case, we merely choose any element b1 of A not appearing in the first
cycle and proceed to create a new cycle as before.

That is, we let b2 = α (b1), b3 = α2(b1), and so on, until we reach b1 = αk (b1) for some k.

This new cycle will have no elements in common with the previously constructed cycle. For,
if so, then αi (a1) = αj (b1) for some i and j. But then α i – j (a1) = b1, and therefore b1 = at for
some t.

This contradicts the way b1 was chosen. Continuing this process until we run out of elements
of A, our permutation will appear as α = (a1, a2, . . . , am)(b1, b2, . . . , bk ) . . . (c1, c2, . . . , cs ).
In this way, we see that every permutation can be written as a product of disjoint cycles.

8
Theorem 1.9: If the pair of cycles α as α = (a1, a2, . . . , am) and β = (b1, b2, . . . , bk ) have no
entries in common, then α β = β α. (Disjoint Cycles Commute)

PROOF

For definiteness, let us say that α and β are permutations of the set

S = { a1, a2, . . . , am, b1, b2, . . . , bn, c1, c2, . . . , ck },

where the c‘s are the members of S left fixed by both α and β (there may not be any c‘s). To
prove that α β = β α., we must show that (α β)(x) = (β α)(x). for all x in S. If x is one of the a
elements, say ai , then

(α β)(ai) = α (β(ai)) = α (ai) = ai+1

since β fixes all a elements. (We interpret ai+1 as a1 if i = m.) For the same reason,

(β α)(ai) = β (α (ai)) = β (ai) = ai+1

. Hence, the functions of α β and β α agree on the a elements. A similar argument shows that
α β and β α agree on the b elements as well. Finally, suppose that x is a c element, say ci .
Then, since both α and β fix c elements, we have

(α β)(ci) = α (β(ci)) = α (ci) = ci+1

and

(β α)(ci) = β (α (ci)) = β (ci) = ci+1

. This completes the proof.

1.7.Even and Odd Permutations

A permutation that can be expressed as a product of an even number of 2-cycles (α β) is


called an even permutation. A permutation that can be expressed as a product of an odd
number of 2-cycles (α β) is called an odd permutation.

If we have = β1, β2, . . . β r – 2, and therefore, by the Second Principle of Mathematical


Induction, r – 2 is even.

Theorem 1.10: If a permutation can be expressed as a product of an even (odd) number of


2-cycles, then every decomposition of into a product of 2-cycles must have an even (odd)
number of 2-cycles. In symbols, if

= β1, β2, . . . β r and = 1, 2, ... s ,

where the ‘s and the ‘s are 2-cycles, then r and s are both even or both odd

9
PROOF

Observe that = β1 , β 2 , . . . β = 1, 2, ... s

implies

( 1 2 ... s)( β1, β2, . . . βr) – 1 = ( 1 2 ... s β1β2, . . . βr

since a 2-cycle is its own inverse, it guarantees that s + r is even. It also follows that r and s
are both even or both odd.

1.8.Group Isomorphism

An isomorphism from a group G to a group ̅ is a one-to-one mapping (or function) from


G onto ̅ that preserves the group operation.

That is,

(ab) = (a) (b) for all a, b in G.

If there is an isomorphism from G onto ̅ ,

we say that G and ̅ are isomorphic and write G ≈ ̅

There are four separate steps involved in proving that a group G is isomorphic to a group ̅ ,.

Step 1 ―Mapping.‖ Define a candidate for the isomorphism; that is, define a function from
G to ̅ ,.

Step 2 Prove that is one-to-one; that is, assume that (a) = (b) and prove that a = b

Step 3 ―Onto.‖ Prove that is onto; that is, for any element ̅ in ̅ , find an element g in G
such that (g) = ̅ .

Step 4 ―O.P.‖ Prove that is operation-preserving; that is, show that

(ab) = (a) (b) for all a and b in G.

10
Take for instance; Let G be the real numbers under addition and let ̅ be the positive real
numbers under multiplication. Then G and ̅ , are isomorphic under the mapping f(x) = 2x .
Certainly, f is a function from G to ̅ , .

Step 2: To prove that it is one-to-one, suppose that 2x = 2y . Then log2 2x = log2 2y , and
therefore x = y.

Step 3: For ―onto,‖ we must find for any positive real number y some real number x such that
f(x) = y; that is, 2x = y. Well, solving for x gives log2 y.

Step 4 : Finally,

f(x + y) = 2x+y = 2x (2y ) = f(x)f(y)

for all x and y in G, so that f is operation-preserving as well.

1.9.Automorphism

An isomorphism from a group G onto itself is called an automorphism of G.

An example is, a function like when you add 2 complex numbers the answer is a complex
number

1.10. Coset of H in G

Let G be a group and let H be a nonempty subset of G. For any a G, the set {ah | h H} is
denoted by aH. Analogously, Ha = {ha | h [ H} and aHa – 1 = { aha – | h [ H}. When H is a
subgroup of G, the set aH is called the left coset of H in G containing a, whereas Ha is called
the right coset of H in G containing a. In this case, the element a is called the coset
representative of aH (or Ha).

We use |aH| to denote the number of elements in the set aH, and |Ha| to denote the number of
elements in Ha

Lagrange’s Theorem1.11: |H| Divides |G|:

If G is a finite group and H is a subgroup of G, then |H| divides |G|.

PROOF

Let a1H, a2H, . . . , ar H denote the distinct left cosets of H in G. Then, for each a in G, we
have aH = ai H for some i. Also, a aH. Thus, each member of G belongs to one of the
cosets ai H.

In symbols,

G = a1H ... arH.

Now, this union is disjoint,

11
so that

|G| = |a1H| + |a2H| + . . . + |arH|.

Finally, since |ai H| = |H| for each i, we have |G| = r|H|.

We pause to emphasize that Lagrange‘s Theorem is a subgroup candidate criterion; that is, it
provides a list of candidates for the orders of the subgroups of a group. Thus, a group of order
12 may have subgroups of order 12, 6, 4, 3, 2, 1, but no others. Warning! The converse of
Lagrange‘s Theorem is false.

For example, a group of order 12 need not have a subgroup of order 6.. A special name and
notation have been adopted for the number of left (or right) cosets of a subgroup in a group.
The index of a subgroup H in G is the number of distinct left cosets of H in G. This number is
denoted by |G:H|.

1.11. Normal Subgroup

A subgroup H of a group G is called a normal subgroup of G if aH = Ha for all a in G

For example every subgroup of an Abelian group is normal.

(In this case, ah = ha for a in the group and h in the subgroup.)

1.12. Factor Groups

Let G be a group and let H be a normal subgroup of G. The set G/H = {aH | a G} is a group
under the operation (aH)(bH) = abH.

1.13. Group Homomorphism

A Homomorphism from a group G to a group ̅ is a mapping (or function) from G into ̅


that preserves the group operation.

That is,

(ab) = (a) (b) for all a, b in G.

1.14. Kernel of a Homomorphism

The kernel of a homomorphism from a group G to a group with identity e is the set {x G
| (x) = e}. The kernel of is denoted by Ker .

12
TOPIC 2: RING

A ring R is a set with two binary operations, addition (denoted by a + b) and multiplication
(denoted by ab), such that for all a, b, c in R:

1. a + b = b + a (Commutative under +)

2. (a + b) + c = a + (b + c). (Associative under +)

3. There is an additive identity 0. That is, there is an element 0 in R such that a + 0 = a for all
a in R.

4. There is an element ‗– a‘ in R such that a + ( – a) = 0.

5. a(bc) = (ab)c. (Associative under ×)

6. a(b + c) = ab + ac and (b + c) a = ba + ca

The set Z of integers under ordinary addition and multiplication is a commutative ring with
unity 1. The units of Z are 1 and – 1.

Theorem 2.1

Let A be a ring such that whenever xy = zx for some x,y A, then z = y . Then A is
commutative

PROOF

Let a, b A. set x = a, y = ba, z = ab. since a(ba) = (ab)a, we have xy = zx. Thus, by
hypothesis we have z = y. hence ab= ba.

Theorem 2.2

Let A be a ring. Suppose that ab = 1 for some a, b A. Then anbn = 1 for each positive integer n.

PROOF

Using mathematical induction. For n = 1, then a1b1 = ab = 1 and we are done.

Then take n + 1, for some n in integers, then an+1 bn+1 = 1, now a( anbn)b = (ab)anbn = 1 and

ab = 1, hence anbn = 1 hence shown.

2.1.Subring

A subset S of a ring R is a subring of R if S is itself a ring with the operations of R.

2.2.Subring Test

A nonempty subset S of a ring R is a subring if S is closed under subtraction and


multiplication—that is, if a – b and ab are in S whenever a and b are in S.

13
The set of Gaussian integers Z[i] = {a + bi | a, b Z} is a subring of the complex numbers C.

2.3.Zero-Divisors

A zero-divisor is a nonzero element a of a commutative ring R such that there is a nonzero


element b R with ab = 0.

2.4.Integral Domain

An integral domain is a commutative ring with unity and no zero- divisors. The ring of
integers is an integral domain.

2.5.Field

A field is a commutative ring with unity in which every nonzero element is a unit.

Theorem 2.3

Let A be a finite commutative ring with no zero divisors. Then A is a field.

PROOF

We only need to show that A has an identity.

Let a A such that a ≠ 0. Since A has no zero divisors, we conclude that if x,y A and x ≠ y,
then ax ≠ ay. Thus, since A is finite, we conclude that that az = a for some z A.

Now, let b A. Since az = a, we have ba = baz = bza. Since ba = bza, we have (b – bz)a = 0.
Since a ≠ 0 and A has no Zero divisors, we conclude that b – bz = 0. Thus bz = b = zb.

Hence, z is the identity of A. Hence, A is a field.

2.6.Ideal

A subring A of a ring R is called a (two-sided) ideal of R if for every r R and every a A


both ra and ar are in A.

2.7.Ideal Test

A nonempty subset A of a ring R is an ideal of R if

1. a – b A whenever a, b A.

2. ra and ar are in A whenever a A and r R.

For any positive integer n, the set nZ = {0, n, 2n, . . .} is an ideal of Z.

14
Theorem 2.4

Let A = R[x] be, and set I = {f(x) A: f(1) = 0}. Then I is a prime ideal of A.

PROOF

It is easy to see that I is Ideal of A. Now suppose f(x) g(x) I for some f(x), g(x) A.

Then f (1) g(1) = 0. Since f(x) R and g(x) R and f(1)g(1) = 0, we conclude that f(x) I
and g(x) I

Theorem 2.5

Let A be a ring with 1, and let I be an Ideal of A such that I contains a unit of A. Then I = A.
In particular, if I contains 1, then I = A.

PROOF

Suppose that I contains a unit u of A. since I is an ideal of A, u – 1 = 1 I. Now, let a A.


then a (1) = a A. Hence, A I, thus I = A.

2.8.Factor Rings

Let R be a ring and let A be an ideal of R. Since R is a group under addition and A is a
normal subgroup of R, we may form the factor group R/A = {r + A | r R}

For example if we have Z/4Z = {0 + 4Z, 1 + 4Z, 2 + 4Z, 3 + 4Z}.

Suppose we take 2 + 4Z and 3 + 4Z.

i. (2 + 4Z) + (3 + 4Z) = 2 + 3 + 4Z + 4Z = 5 + 4Z = 1 + 4 + 4Z = 1 + 4Z,


ii. (2 + 4Z)(3 + 4Z) = 6 + 8Z + 12Z +16Z2 = 6 + (20 + 4)Z = 2 + 4 + 4Z = 2 + 4Z.

2.9.Ring Homomorphism

A ring homomorphism f from a ring R to a ring S is a mapping from R to S that preserves the
two ring operations; that is, for all a, b in R, f(a + b) = f(a) + f(b) and f(ab) = f(a)f(b).

2.8.Ring Isomorphism

A ring homomorphism that is both one-to-one and onto.

2.9.First Isomorphism Theorem for Rings

Let be a ring homomorphism from R to S. Then the mapping from R/Ker to (R), given
by r + Ker (r), is an isomorphism. In symbols, R/Ker ≈ (R).

15
Theorem 2.6:

Let R be a ring with unity 1. The mapping f: Z R given by n n 1 is a ring


homomorphism.

PROOF

Since the multiplicative group property

am+n = aman

log(a)m+n = log (aman)

log(a)m+n = log am + log an

(m + n) = m + n

f(m + n) = f(m) + f(n)

Hence, when the operation is addition the mapping preserves addition

2.10. Ring of Polynomials over R

Let R be a commutative ring. a Ring of Polynomials over R is a set of formal symbols

R[x] = {anxn + an – 1xn – 1 . . .+ a1x 1+ a0|ai R, n is a non-negative integer}

Two elements

anxn + an – 1xn – 1 . . .+ a1x 1+ a0

and

bnxn + bn – 1xn – 1 . . .+ b1x 1+ b0

of R[x] are considered equal iff ai = bi for all non-negative integers i.

2.11. Addition and Multiplication in R[x]

Let R be a commutative ring and let

f(x) = anxn + an – 1xn – 1 . . .+ a1x 1+ a0

and

g(x) = bnxn + bn – 1xn – 1 . . .+ b1x 1+ b0

belong to R[x]. Then

f(x) + g(x) = (as + bs) xs + (as – 1+ bs – 1)xs – 1 . . .+ a0+ b0

while

16
f(x)g(x) = (anxn + an – 1xn – 1 . . .+ a1x 1+ a0) (bnxn + bn – 1xn – 1 . . .+ b1x 1+ b0)

cnxn+m + cn – 1xn+m – 1 . . .+ c1x 1+ c0

Resulting into a polynomial of higher order

Task

Given f(x) = 2x3 + x2 + 2x + 2 and g(x) = x2 + 2x + 1 in Z3[x], that is Z mod 3, x = {0, 1, 2}

Show that

i. f(x) + g(x) = 2x3 + x


ii. f(x)g(x) = x5 + 2x3 + 2
2.12. Remainder Theorem

Let F be a field, a ∈ F, and f(x) ∈ F[x]. Then f(a) is the remainder in the division of f(x) by x – a.

2.13. Factor Theorem:


Let F be a field, a ∈ F, and f(x) ∈ F[x]. Then a is a zero (root or answer) of f(x) if
and only if x – a is a factor of f(x).

Theorem 2.7:

A polynomial of degree n over a field has at most n zeros, counting multiplicity.

PROOF

We proceed by induction on n. Clearly, a polynomial of degree 0 over a field has no zeros.


Now suppose that f(x) is a polynomial of degree n over a field and a is a zero of f(x) of
multiplicity k.

Then, f(x) = (x – a)k q(x) and q(a) ≠ 0; since n = deg f(x) = deg(x - a)k q(x) = k + deg q(x),
we have k n . If f(x) has no zeros other than a, we are done.

On the other hand, if b ≠ a and b is a zero of f(x), then 0 = f(b) = (b – a)k q(b), so that b is also
a zero of q(x) with the same multiplicity as it has for f(x).

By the Second Principle of Mathematical Induction, we know that q(x) has at most deg q(x)
= n – k zeros, counting multiplicity. Thus, f(x) has at most k + n – k = n zeros, counting
multiplicity.

Hence shown

2.14. Factorization of Polynomials


i. Irreducible Polynomial
A non-constant f(x) ∈ F[x] to be irreducible if f(x) cannot be expressed as a
product of two polynomials of lower degree.
e.g. The polynomial f(x) = 2x2 + 4 is irreducible over Q

17
ii. Reducible Polynomial

Let F be a field. If f(x) ∈ F[x] and deg f(x) is 2 or 3, then f(x) is reducible over F if
and only if f(x) has a zero in F

e.g. The polynomial f(x) = x2 – 2 is reducible over R

Gauss’s Lemma 2.8

State that the product of two primitive polynomials is primitive.

Note: A primitive polynomial is a polynomial with the greatest common divisor of


coefficients, constant inclusive is 1.

e.g. f(x) = 2x3 + 7x2 + 5x + 13

TASK

State any two primitive polynomials and show that the product of any two primitive
polynomials is a primitive.

2.15. Extension Field

A field E is an extension field of a field F if F E and the operations of F are those of E


restricted to F.

Fundamental Theorem of Field Theory 2.9

Let F be a field and let f(x) be a non-constant polynomial in F[x]. Then there is an extension
field E of F in which f(x) has a zero.

Even if all the coefficients in a polynomial are real numbers, some solutions may be found in
complex numbers.


R

f(x)

18
References

1. J. A. Gallian and D. Rusin, ―Factoring Groups of Integers Modulo n,‖ Mathematics


Magazine

2. D. Shanks (1978), Solved and Unsolved Problems in Number Theory, 2nd ed., New York:
Chelsea

3. S. Washburn, T. Marlowe, and C. Ryan, (1999) Discrete Mathematics, Reading, MA:


Addison-Wesley,

19

You might also like