0% found this document useful (0 votes)
16 views86 pages

Lecture note_Polynomial Theory

The document is a comprehensive study on polynomial theory, covering topics such as rings of polynomials, factoring, substitution, and field extensions. It includes detailed explanations of polynomial operations, properties, and theorems, particularly in the context of commutative rings and integral domains. The content is structured into chapters that progressively build on the concepts of polynomials and their applications in algebra.

Uploaded by

Hiếu Lê
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)
16 views86 pages

Lecture note_Polynomial Theory

The document is a comprehensive study on polynomial theory, covering topics such as rings of polynomials, factoring, substitution, and field extensions. It includes detailed explanations of polynomial operations, properties, and theorems, particularly in the context of commutative rings and integral domains. The content is structured into chapters that progressively build on the concepts of polynomials and their applications in algebra.

Uploaded by

Hiếu Lê
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/ 86

VINH UNIVERSITY

COLLEGE OF EDUCATION

NGUYEN THI HONG LOAN

POLYNOMIAL THEORY

VINH 2024
CONTENTS

Contents 1

Introduction 2

1 Rings of Polynomials 3
1.1. Polynomials in One Variable . . . . . . . . . . . . . . . . . . 3
1.2. Division Algorithm . . . . . . . . . . . . . . . . . . . . . . . 9

2 Factoring Polynomials 18
2.1. Ideals of F[x] . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2. Properties of the GCD . . . . . . . . . . . . . . . . . . . . . 19
2.3. Irreducible Polynomials . . . . . . . . . . . . . . . . . . . . . 21

3 Substitution in Polynomials 28
3.1. Polynomial Functions . . . . . . . . . . . . . . . . . . . . . . 28
3.2. Polynomials over Z and Q . . . . . . . . . . . . . . . . . . . . 31
3.3. Eisenstein’s Irreducibility Criterion . . . . . . . . . . . . . . . 34
3.4. Polynomials over R and C . . . . . . . . . . . . . . . . . . . . 35
3.5. Solution of Cubic and Quartic Equations by Formulas . . . . . 36

4 Extensions of Fields 60
4.1. Algebraic and Transcendental Elements . . . . . . . . . . . . . 60
4.2. The Minimum Polynomial . . . . . . . . . . . . . . . . . . . . 62
4.3. Basic Theorem on Field Extensions . . . . . . . . . . . . . . . 64
4.4. Degrees of field extensions . . . . . . . . . . . . . . . . . . . . 66

1
INTRODUCTION

2
CHAPTER 1

RINGS OF POLYNOMIALS

1.1 Polynomials in One Variable

In elementary algebra an important role is played by polynomials in an


unknown x. These are expressions such as
1
2x3 − x2 + 3
2
whose terms are grouped in powers of x. The exponents, of course, are
positive integers and the coefficients are real or complex numbers.
Polynomials are involved in countless applications-applications of every
kind and description. For example, polynomial functions are the easiest
functions to compute, and therefore one commonly attempts to approxi-
mate arbitrary functions by polynomial functions. A great deal of effort
has been expended by mathematicians to find ways of achieving this. Aside
from their uses in science and computation, polynomials come up very nat-
urally in the general study of rings, as the following example will show:
Suppose we wish to enlarge the ring Z by adding to it the number π. It
is easy to see that we will have to adjoin to Z other new numbers besides
just π; for the enlarged ring (containing π as well as all the integers) will
also contain such things as −π, π + 7, 6π 2 − 11, and so on.
As a matter of fact, any ring which contains Z as a subring and which
also contains the number π will have to contain every number of the form

an π n + an−1 π n−1 + . . . + a1 π + a0

where a0 , a1 , . . . , an are integers. In other words, it will contain all the

3
polynomial expressions in π with integer coefficients.
But the set of all the polynomial expressions in π with integer coefficients
is a ring (it is a ring of because it is obvious that the sum and product of
any two polynomials in π is again a polynomial in π). This ring contains
Z because every integer a is a polynomial with a constant term only, and
it also contains π.
Thus, if we wish to enlarge the ring Z by adjoining to it the new number
π, it turns out that the “next largest” ring after Z which contains as a
subring and includes π, is exactly the ring of all the polynomials in π with
coefficients in Z.
As this example shows, aside from their practical applications, polyno-
mials play an important role in the scheme of ring theory because they
are precisely what we need when we wish to enlarge a ring by adding new
elements to it.
In elementary algebra one considers polynomials whose coefficients are
real numbers, or in some cases, complex numbers. As a matter of fact,
the properties of polynomials are pretty much independent of the exact
nature of their coefficients. All we need to know is that the coefficients
are contained in some ring. For convenience, we will assume this ring is a
commutative ring with unity.
Let A be a commutative ring with unity. Up to now we have used letters
to denote elements or sets, but now we will use the letter x in a different
way. In a polynomial expression such as ax2 + bx + c, where a, b, c ∈ A, we
do not consider x to be an element of A, but rather x is a symbol which we
use in an entirely formal way. Later we will allow the substitution of other
things for x, but at present x is simply a placeholder.
Notationally, the terms of a polynomial may be listed in either ascending
or descending order. For example, 4x3 − 3x2 + x + 1 and 1 + x − 3x2 + 4x3
denote the same polynomial. In elementary algebra descending order is
preferred, but for our purposes ascending order is more convenient.

4
Let A be a commutative ring with unity, and x an arbitrary symbol.
Every expression of the form

a0 + a1 x + . . . + an xn

is called a polynomial in x with coefficients in A, or more simply, a polyno-


mial in x over A. The expressions ak xk , for k ∈ {1, . . . , n}, are called the
terms of the polynomial.
Polynomials in x are designated by symbols such as a(x), b(x), q(x), and
so on. If a(x) = a0 + a1 x + . . . + an xn is any polynomial and ak xk is any one
of its terms, ak is called the coefficient of xk . By the degree of a polynomial
a(x) we mean the greatest n such that the coefficient of xn is not zero. In
other words, if a(x) has degree n, this means that an ̸= 0 but am = 0 for
every m > n. The degree of a(x) is symbolized by deg a(x). For example,
1 + 2x − 3x2 + x3 is a polynomial degree 3.
The polynomial 0 + 0x + 0x2 + . . . all of whose coefficients are equal
to zero is called the zero polynomial, and is symbolized by 0. It is the
only polynomial whose degree is not defined (because it has no nonzero
coefficient).
If a nonzero polynomial a(x) = a0 + a1 x + . . . + an xn has degree n, then
an is called its leading coefficient: it is the last nonzero coefficient of a(x).
The term an xn is then called its leading term, while a0 is called its constant
term.
If a polynomial a(x) has degree zero, this means that its constant term
a0 is its only nonzero term: a(x) is a constant polynomial. Beware of
confusing a polynomial of degree zero with the zero polynomial.
Two polynomials a(x) and b(x) are equal if they have the same degree
and corresponding coefficients are equal. Thus, if a(x) = a0 + . . . + an xn is
of degree n, and b(x) = b0 + . . . + bm xm is of degreem, then a(x) = b(x) iff
n = m and ak = bk for each k from 0 to n.

5
The familiar sigma notation for sums is useful for polynomials. Thus,
n
X
n
a(x) = a0 + . . . + an x = ak xk .
k=0

with the understanding that x0 = 1. Addition and multiplication of poly-


nomials is familiar from elementary algebra. We will now define these op-
erations formally. Throughout these definitions we let a(x) and b(x) stand
for the following polynomials:

a(x) = a0 + . . . + an xn ,

b(x) = b0 + . . . + bn xn .
Here we do not assume that a(x) and b(x) have the same degree, but allow
ourselves to insert zero coefficients if necessary to achieve uniformity of
appearance.
We add polynomials by adding corresponding coefficients. Thus,

a(x) + b(x) = (a0 + b0 ) + (a1 + b1 )x + . . . + (an + bn )xn .

Note that the degree of a(x) + b(x) is less than or equal to the higher of
the two degrees, deg a(x) and deg b(x). Multiplication is more difficult, but
quite familiar:

a(x)b(x) = a0 b0 + (a0 b1 + b0 a1 )x + (a0 b2 + a1 b1 + a2 b0 )x2 + . . . + an bn x2n .

In other words, the product of a(x) and b(x) is the polynomial

c(x) = c0 + c1 x + . . . + c2n x2n

whose k th coefficient (for any k from 0 to 2n) is


X
ck = ai bj .
i+j=k

This is the sum of all the ai bj for which i + j = k. Note that

deg[a(x)b(x)] ≤ deg a(x) + deg b(x).

6
If A is any ring, the symbol A[x] designates the set of all the polyno-
mials in x whose coefficients are in A, with addition and multiplication of
polynomials as we have just defined them.
1.1.1 Theorem. Let A be a commutative ring with unity. Then A[x] is a
commutative ring with unity.
Proof. To prove this theorem, we must show systematically that A[x] satis-
fies all the axioms of acommutative ring with unity. Throughout the proof,
let a(x), b(x), and c(x) stand for the following polynomials:

a(x) = a0 + . . . + an xn ,

b(x) = b0 + . . . + bn xn ,
c(x) = c0 + . . . + cn xn .
The axioms which involve only addition are easy to check: for example,
addition is commutative because
a(x) + b(x) = (a0 + b0 ) + (a1 + b1 )x + . . . + (an + bn )xn
= (b0 + a0 ) + (b1 + a1 )x + . . . + (bn + an )xn
= b(x) + a(x).
The associative law of addition is proved similarly, and is left as an exercise.
The zero polynomial has already been described, and the negative of a(x)
is
−a(x) = (−a0 ) + (−a1 )x + . . . + (−an )xn ,
To prove that multiplication is associative requires some care. Let
b(x)c(x) = d(x), where d(x) = d0 + d1 x + . . . + d2n x2n . By the definition of
polynomial multiplication, the k th coefficient of b(x)c(x) is
X
dk = b i cj .
i+j=k

Then a(x)[b(x)c(x)] = a(x)d(x) = e(x), where e(x) = e0 +e1 x+. . .+e3n x3n .
Now, the ℓth coefficient of a(x)d(x) is
X X X
ek = ah dk = ah ( bi cj ).
h+k=ℓ h+k=ℓ i+j=k

7
It is easy to see that the sum on the right consists of all the terms ah bi cj
such that h + i + j = ℓ. Thus,
X
ek = ah bi cj .
h+i+j=ℓ

For each ℓ from 0 to 3n, eℓ is the ℓth coefficient of a(x)[b(x)c(x)].


If we repeat this process to find the ℓth coefficient of [a(x)b(x)]c(x), we
discover that it, too, is eℓ . Thus,

a(x)[b(x)c(x)] = [a(x)b(x)]c(x).

To prove the distributive law, let a(x)[b(x) + c(x)] = d(x) where d(x) =
d0 + d1 x + . . . + d2n x2n . By the definitions of polynomial addition and mul-
tiplication, the k th coefficient a(x)[b(x) + c(x)] is
X X X X
dk = ai (bj + cj ) = ai bj + ai cj = ai b j + ai cj .
i+j=k i+j=k i+j=k i+j=k

ai bj is exactly the k th coefficient of a(x)b(x), and ai cj is the k th


P P
But
i+j=k i+j=k
th
coefficient of a(x)c(x), hence dk is equal to the k coefficient of a(x)b(x) +
a(x)c(x). This proves that

a(x)[b(x) + c(x)] = a(x)b(x) + a(x)c(x).

The commutative law of multiplication is simple to verify and is left to the


student. Finally, the unity polynomial is the constant polynomial 1.

1.1.2 Theorem. If A is an integral domain, then A[x] is an integral do-


main.

Proof. If a(x) and b(x) are nonzero polynomials, we must show that their
product a(x)b(x) is not zero. Let an be the leading coefficient of a(x), and
bm the leading coefficient of b(x). By definition, an ̸= 0, and bm ̸= 0. Thus
an bm ̸= 0 because A is an integral domain. It follows that a(x)b(x) has a
nonzero coefficient (namely, an bm ), so it is not the zero polynomial.

8
If A is an integral domain, we refer to A[x] as a domain of polynomials,
because A[x] is an integral domain. Note that by the preceding proof, if
an and bm are the leading coefficients of a(x) and b(x), then an bm is the
leading coefficient of a(x)b(x). Thus, deg a(x)b(x) = n + m: In a domain
of polynomials A[x], where A is an integral domain,

deg[a(x)b(x)] = deg a(x) + deg b(x).

In the remainder of this chapter we will look at a property of polynomials


which is of special interest when all the coefficients lie in a field. Thus,
from this point forward, let F be a field, and let us consider polynomials
belonging to F[x].

1.2 Division Algorithm

It would be tempting to believe that if F is a field then F[x] also is a


field. However, this is not so, for one can easily see that the multiplicative
inverse of a polynomial is not generally a polynomial. Nevertheless, by
Theorem 1.1.2, F[x] is an integral domain. Domains of polynomials over a
field do, however, have a very special property: any polynomial a(x) may
be divided by any nonzero polynomial b(x) to yield a quotient q(x) and a
remainder r(x). The remainder is either 0, or if not, its degree is less than
the degree of the divisor b(x). For example, x2 may be divided by x − 2 to
give a quotient of x + 2 and a remainder of 4:

x2 = (x − 2) (x + 2) + |{z}
|{z} 4 .
| {z } | {z }
a(x) b(x) c(x) r(x)

This kind of polynomial division is familiar to every student of elemen-


tary algebra. It is customarily set up as follows:
The process of polynomial division is formalized in the next theorem.

1.2.1 Theorem (Division algorithm for polynomials). If a(x) and b(x) are
polynomials over a field F,and b(x) ̸= 0, there exist polynomials q(x) and

9
r(x) over F such that

a(x) = b(x)q(x) + r(x)

and r(x) = 0 or deg r(x) < deg b(x).

Proof. Let b(x) remain fixed, and let us show that every polynomial a(x)
satisfies the following condition: There exist polynomials q(x) and r(x) over
F such that
a(x) = b(x)q(x) + r(x),

and r(x) = 0 or deg r(x) < deg b(x).


We will assume there are polynomials a(x) which do not fulfill the con-
dition, and from this assumption we will derive a contradiction. Let a(x)
be a polynomial of lowest degree which fails to satisfy the conditions. Note
that a(x) cannot be zero, because we can express 0 as 0 = b(x).0 + 0, where
by a(x) would satisfy the conditions. Furthermore, deg a(x) ≥ deg b(x), for
if deg a(x) < deg b(x) then we could write a(x) = b(x).0 + a(x), so again
a(x) would satisfy the given conditions.
Let a(x) = a0 + . . . + an xn and b(x) = b0 + . . . + bm xm . Define a new
polynomial
an n−m
A(x) = a(x) + x b(x)
bm
an an an
= a(x) − (b0 xn−m + b1 xn−m+1 + . . . + bm xn−m+m )
am am | am {z }
an xn

This expression is the difference of two polynomials both of degree n and


both having the same leading term an xn . Because an xn cancels in the sub-
traction, A(x) has degree less than n.
Remember that a(x) is a polynomial of least degree which fails to sat-
isfy the given condition; hence A(x) does satisfy it. This means there are
polynomials p(x) and r(x) such that

A(x) = b(x)p(x) + r(x)

10
where r(x) = 0 or deg r(x) < deg b(x). But then
an n−m
A(x) = a(x) + x b(x)
bm
an
= b(x)p(x) + r(x) + xn−m b(x)
bm
an n−m
= b(x)(p(x) + x ) + r(x).
bm
If we let p(x) + (an /bm )xn − m be renamed q(x), then

a(x) = b(x)q(x) + r(x),

so a(x) fulfills the given condition. This is a contradiction, as required.

11
EXERCISES

A. Elementary Computation in Domains of Polynomials

REMARK ON NOTATION: In some of the problems which follow, we con-


sider polynomials with coefficients in Zn for various n. To simplify notation,
we denote the elements of Zn by 1, 2, . . . , n − 1 rather than the more correct
1, 2, . . . , n − 1.
1. Let a(x) = 2x2 +3x+1 and b(x) = x3 +5x2 +x. Compute a(x)+b(x),
a(x)–b(x) and a(x)b(x) in Z[x], Z5 [x], Z6 [x], and Z7 [x].
2. Find the quotient and remainder when x3 + x2 + x + 1 is divided by
x2 + 3x + 2 in Z[x] and in Z5 [x].
3. Find the quotient and remainder when x3 +2 is divided by 2x2 +3x+4
in Z[x], in Z3 [x], and in Z5 [x].
We call b(x) a factor of a(x) if a(x) = b(x)q(x) for some q(x), that is, if
the remainder when a(x) is divided by b(x) is equal to zero.
4. Show that the following is true in A[x] for any ring A: For any odd
n,
(a) x + 1 is a factor of xn + 1,
(b) x + 1 is a factor of xn + xn−1 + . . . + x + 1.
5. Prove the following statements:
(a) In Z3 [x], x + 2 is a factor of xm + 2, for all m,
(b) In Zn [x], x + (x − 1) is a factor of xm + (n − 1), for all m and n.
6. Prove that there is no integer m such that 3x2 + 4x + m is a factor of
6x4 + 50 in Z[x].
7. For what values of n is x2 + 1 a factor of x5 + 5x + 6 in Zn [x]?

B. Problems Involving Concepts and Definitions

1. Is x8 + 1 = x3 + 1 in Z5 [x]? Explain your answer.


2. Is there any ring A such that in A[x], some polynomial of degree 2 is

12
equal to a polynomial of degree 4? Explain.
3. Write all the quadratic polynomials in Z5 [x]. How many are there?
How many cubic polynomials are there in Z5 [x]? More generally, how many
polynomials of degree m are there in Zn [x]?
4. Let A be an integral domain, prove the following statements:
(a) If (x + 1)2 = x2 + 1 in A[x], then A must have characteristic 2,
(b) If (x + 1)4 = x4 + 1 in A[x], then A must have characteristic 2,
(c) If (x + 1)6 = x6 + 2x3 + 1 in A[x], then A must have characteristic 3.
5. Find an example of each of the following in Z8 [x]: a divisor of zero,
an invertible element (find nonconstant examples).
6. Explain why x cannot be invertible in any A[x], hence no domain of
polynomials can ever be a field.
7. There are rings in which every element ̸= 0, 1 is a divisor of zero.
Explain why this cannot happen in any ring of polynomials A[x], even
when A is not an integral domain.
8. Show that in every A[x], there are elements ̸= 0, 1 which are not
idempotent, and elements ̸= 0, 1 which are not nilpotent.

C. Rings A[x] Where A Is Not an Integral Domain

1. Prove: If A is not an integral domain, neither is A[x].


2. Give examples of divisors of zero, of degrees 0, 1, and 2, in Z4 [x]. 3
In Z10 [x], (2x + 2)(2x + 2) = (2x + 2)(5x3 + 2x + 2), yet (2x + 2) cannot be
canceled in this equation. Explain why this is possible in Z10 [x], but not in
Z5 [x].
4. Give examples in Z4 [x], in Z6 [x], and in Z9 [x] of polynomials a(x) and
b(x) such that
deg a(x)b(x) < deg a(x) + deg b(x).

5. If A is an integral domain, we have seen that in A[x], deg a(x)b(x) =


deg a(x)+deg b(x). Show that if A is not an integral domain, we can always

13
find polynomials a(x) and b(x) such that deg a(x)b(x) < deg a(x)+deg b(x).
6. Show that if A is an integral domain, the only invertible elements in
A[x] are the constant polynomials with inverses in A. Then show that in
Z4 [x] there are invertible polynomials of all degrees.
7. Give all the ways of factoring x2 into polynomials of degree 1 in Z9 [x];
in Z5 [x]. Explain the difference in behavior.
8. Find all the square roots of x2 + x + 4 in Z5 [x]. Show that in Z8 [x],
there are infinitely many square roots of 1.

D. Domains A[x] Where A Has Finite Characteristic

In each of the following, let A be an integral domain:


1. Prove that if A has characteristic p, then A[x] has characteristic p.
2. Use part 1 to give an example of an infinite integral domain with
finite characteristic.
3. Prove: If A has characteristic 3, then x + 2 is a factor of xm + 2 for
all m. More generally, if A has characteristic p, then x + (p–1) is a factor
of xm + (p − 1) for all m.
4. Prove that if A has characteristic p, then in A[x], (x + c)p = xp + cp .
5. Explain why the following “proof of part 4 is not valid: (x + c)p =
xp + cp in A[x] because (a + c)p = ap + cp for all a, c ∈ A. (Note the following
example: in Z2 , a2 + 1 = a4 + 1 for every a, yet x2 + 1 ̸= x4 + 1 in Z2 [x]).
6. Use the same argument as in part 4 to prove that if A has character-
istic p, then [a(x) + b(x)]p = a(x)p + b(x)p for any a(x), b(x) ∈ A[x]. Use
this to prove:

(a0 + a1 x + . . . + an xn )p = ap0 + ap1 xp + . . . + apn xnp .

E. Subrings and Ideals in A[x]

1. Show that if B is a subring of A, then B[x] is a subring of A[x].


2. If B is an ideal of A, B[x] is an ideal of A[x].

14
3. Let S be the set of all the polynomials a(x) in A[x] for which every
coefficient ai for odd i is equal to zero. Show that S is a subring of A[x].
Why is the same not true when “odd” is replaced by “even”?
4. Let J consist of all the elements in A[x] whose constant coefficient is
equal to zero. Prove that J is an ideal of A[x].
5. Let J consist of all the polynomials a0 + a1 x + . . . + an xn in A[x] such
that a0 + a1 + . . . + an = 0. Prove that J is an ideal of A[x].
6 Prove that the ideals in both parts 4 and 5 are prime ideals (assume
A is an integral domain).

F. Homomorphisms of Domains of Polynomials

Let A be an integral domain.


1. Let h : A[x] → A map every polynomial to its constant coefficient;
that is, h(a0 + a1 x + . . . + an xn ) = a0 . Prove that:
(a) h is a homomorphism from A[x] onto A,
(b) Kernel of h is a principal ideal (x) in A[x].
2. Using parts 1, explain why A[x]/(x) ∼ = A.
3. Let g : A[x] → A send every polynomial to the sum of its coefficients.
Prove that g is a surjective homomorphism, and describe its kernel.
4. If c ∈ A, let h : A[x] → A[x] be defined by h(a(x)) = a(cx), that is,

h(a0 + a1 x + . . . + an xn ) = a0 + a1 cx + a2 c2 x2 + . . . + an cn xn .

Prove that h is a homomorphism and describe its kernel.


5. If h is the homomorphism of part 4, prove that h is an automorphism
(isomorphism from A[x] to itself) iff c is invertible.

G. Homomorphisms of Polynomial Domains Induced by a Homo-


morphism of the Ring of Coefficients

Let A and B be rings and let h : A → B be a homomorphism with

15
kernel K. Define h : A[x] → B[x] by

h(a0 + a1 x + . . . + an xn ) = h(a0 ) + h(a1 )x + . . . + h(an )xn

(we say that h is induced by h).


1. Prove that h is a homomorphism from A[x] to B[x].
2. Describe K the kernel of h.
3. Prove that h is surjective iff h is surjective.
4 Prove that h is injective iff h is injective.
5 Prove that if a(x) is a factor of b(x), then h(a(x)) is a factor of h(b(x)).
6 If h : Z → Zn is the natural homomorphism, let h : Z[x] → Zn [x]
be the homomorphism induced by h. Prove that h(a(x)) = 0 iff n divides
every coefficient of a(x).
7. Let h be as in part 6, and let n be a prime. Prove that if a(x)b(x) ∈
ker h, then either a(x) or b(x) is in ker h. (HINT: Use Exercise F2 of
Chapter 19).

H. Polynomials in Several Variables

A[x1 , x2 ] denotes the ring of all the polynomials in two letters x1 and x2
with coefficients in A. For example, x2 − 2xy + y 2 + x − 5 is a quadratic
polynomial in Q[x, y]. More generally, A[x1 , . . . , xn ] is the ring of the poly-
nomials in n letters x1 , . . . , xn with coefficients in A. Formally it is defined
as follows: Let A[x1 ] be denoted by A1 ; then A1 [x2 ] is A[x1 , x2 ]. Contin-
uing in this fashion, we may adjoin one new letter xi at a time, to get
A[x1 , . . . , xn ].
1. Prove that if A is an integral domain, then A[x1 , . . . , xn ] is an integral
domain.
2. Give a reasonable definition of the degree of any polynomial p(x, y)
in A[x, y] and then list all the polynomials of degree ≤ 3 in Z3 [x, y].
aij xi y j
P
Let us denote an arbitrary polynomial p(x, y) in A[x, y] by
P
where ranges over some pairs i, j of nonnegative integers.

16
3. Imitating the definitions of sum and product of polynomials in A[x],
give a definition of sum and product of polynomials in A[x, y].
4. Prove that deg a(x, y)b(x, y) = deg a(x, y) + deg b(x, y) if A is an
integral domain.

I. Fields of Polynomial Quotients

Let A be an integer domain. Then A[x] is also an integer domain.


Let A(x) be the field of quotients of A[x]. Note that A(x) consists of all
a(x)
fractions with a(x) and b(x) ̸= 0 in A[x], and these fractions can be
b(x)
added, subtracted, multiplied, and divided in the customary way.
1. Show that A(x) has the same characteristic as A.
2. Using part 1, explain why there is an infinite field of characteristic p,
for every prime p.
3. If A and B are integral domains and h : A → B is an isomorphism,
prove that h determines an isomorphism h : A(x) → B(x).

J. Division Algorithm: Uniqueness of Quotient and Remainder

In the division algorithm, prove that q(x) and r(x) are uniquely deter-
mined. [HINT: Suppose

a(x) = b(x)q1 (x) + r1 (x) = b(x)q2 (x) + r2 (x),

and subtract these two expressions, which are both equal to a(x).]

17
CHAPTER 2

FACTORING POLYNOMIALS

2.1 Ideals of F[x]

Just as every integer can be factored into primes, so every polynomial can
be factored into “irreducible” polynomials which cannot be factored further.
As a matter of fact, polynomials behave very much like integers when it
comes to factoring them. This is especially true when the polynomials have
all their coefficients in a field.
Throughout this chapter, we let F represent some field and we consider
polynomials over F. It will be found that F[x] has a considerable number
of properties in common with Z. To begin with, all the ideals of F[x] are
principal ideals, which was also the case for the ideals of Z. Note carefully
that in F[x], the principal ideal generated by a polynomial a(x) consists of
all the products a(x)s(x) as a(x) remains fixed and s(x) ranges over all the
members of F[x].

2.1.1 Theorem. Every ideal of F[x] is principal.

Proof. Let J be any ideal of F[x]. If J contains nothing but the zero poly-
nomial, J is the principal ideal generated by 0.
If there are nonzero polynomials in J, let b(x) be any polynomial of
lowest degree in J. We will show that J = ⟨b(x)⟩, which is to say that
every element of J is a polynomial multiple b(x)q(x) of b(x).
Indeed, if a(x) is any element of J, we may use the division algorithm to
write a(x) = b(x)q(x) + r(x), where r(x) = 0 or deg r(x) < deg b(x). Now,
r(x) = a(x) − b(x)q(x); but a(x) was chosen in J, and b(x) ∈ J; hence

18
b(x)q(x) ∈ J. It follows that r(x) is in J.
If r(x) ̸= 0, its degree is less than the degree of b(x). But this is im-
possible because b(x) is a polynomial of lowest degree in J. Therefore, of
necessity, r(x) = 0. Thus, finally, a(x) = b(x)q(x); so every member of J
is a multiple of b(x), as claimed.

It follows that every ideal J of F[x] is principal. In fact, as the proof


above indicates, J is generated by any one of its members of lowest degree.

2.2 Properties of the GCD

Throughout the discussion which follows, remember that we are consid-


ering polynomials in a fixed domain F[x] where F is a field.
Let a(x) and b(x) be in F[x]. We say that b(x) is a multiple of a(x) if
b(x) = a(x)s(x) for some polynomial s(x) in F[x]. If b(x) is a multiple of
a(x), we also say that a(x) is a factor of b(x), or that a(x) divides b(x). In
symbols, we write a(x) | b(x).
Every nonzero constant polynomial divides every polynomial. For if
c ̸= 0 is constant and

a(x) = a0 + a1 x + . . . + an xn ,

then
a0 a1 an
a(x) = c( + x + . . . + xn ),
c c c
hence c | a(x). A polynomial a(x) is invertible iff it is a divisor of the unity
polynomial 1. But if a(x)b(x) = 1, this means that a(x) and b(x) both have
degree 0, that is, are constant polynomials: a(x) = a, b(x) = b, and ab = 1.
Thus, we have the following statement.

2.2.1 Proposition. The invertible elements of F[x] are all the nonzero
constant polynomials.

A pair of nonzero polynomials a(x) and b(x) are called associates if


they divide one another: a(x) | b(x) and b(x) | a(x). That is to say,

19
a(x) = b(x)c(x) and b(x) = a(x)d(x) for some c(x) and d(x). If this
happens to be the case, then

a(x) = b(x)c(x) = a(x)d(x)c(x)

hence d(x)c(x) = 1 because F[x] is an integral domain. But then c(x) and
d(x) are constant polynomials, and therefore a(x) and b(x) are constant
multiples of each other. Thus, we have the following statement.

2.2.2 Proposition. In F[x], a(x) and b(x) are associates iff they are con-
stant multiples of each other.

If a(x) = a0 + a1 x + . . . + an xn , the associates of a(x) are all its nonzero


constant multiples. Among these multiples is the polynomial
a0 a1 an
+ x + . . . + xn
an an an
which is equal to (1/an )a(x), and which has 1 as its leading coefficient. Any
polynomial whose leading coefficient is equal to 1 is called monk. Thus,
every nonzero polynomial a(x) has a unique monic associate. For example,
3
the monic associate of 3 + 4x + 2x3 is + 2x + x3 .
2
2.2.3 Definition. A polynomial d(x) is called a greatest common divisor
of a(x) and b(x) if d(x) divides a(x) and b(x), and is a multiple of any other
common divisor of a(x) and b(x); in other words,

(1) d(x) | a(x) and d(x) | b(x), and

(2) For any u(x) in F[x], if u(x) | a(x) and u(x) | b(x), then u(x) | d(x).

According to this definition, two different gcd’s of a(x) and b(x) divide
each other, that is, are associates. Of all the possible gcd’s of a(x) and
b(x), we select the monic one, call it the gcd of a(x) and b(x), and denote
it by gcd[a(x), b(x)]. It is important to know that any pair of polynomials
always has a greatest common divisor.

20
2.2.4 Theorem. Any two nonzero polynomials a(x) and b(x) in F[x] have
a gcd d(x). Furthermore, d(x) can be expressed as a “linear combination”

d(x) = r(x)a(x) + s(x)b(x)

where r(x) and s(x) are in F[x].

Proof. The proof is analogous to the proof of the corresponding theorem


for integers. If J is the set of all the linear combinations

u(x)a(x) + v(x)b(x)

as u(x) and v(x) range over F[x], then J is an ideal of F[x], say the ideal
⟨d(x)⟩ generated by d(x). Now a(x) = la(x) + 0b(x) and b(x) = 0a(x) +
1b(x), so a(x) and b(x) are in J. But every element of J is a multiple of
d(x), so d(x) | a(x) and d(x) | b(x).
If k(x) is any common divisor of a(x) and b(x), this means there are
polynomials F(x) and g(x) such that a(x) = k(x)f (x) and b(x) = k(x)g(x).
Now, d(x) ∈ J, so d(x) can be written as a linear combination
d(x) = r(x)a(x) + s(x)b(x)
= r(x)k(x)f (x) + s(x)g(x)g(x)
= k(x)[r(x)f (x) + s(x)g(x)].
hence k(x) | d(x). This confirms that d(x) is the gcd of a(x) and b(x).

2.2.5 Definition. Polynomials a(x) and b(x) in F[x] are said to be rela-
tively prime if their gcd is equal to 1.

This is equivalent to saying that their only common factors are constants
in F.

2.3 Irreducible Polynomials

2.3.1 Definition. A polynomial a(x) of positive degree is said to be re-


ducible over F if there are polynomials b(x) and c(x) in F[x], both of positive
degree, such that a(x) = b(x)c(x).

21
Because b(x) and c(x) both have positive degrees, and the sum of their
degrees is deg a(x), each has degree less than deg a(x).

2.3.2 Definition. A polynomial p(x) of positive degree in F[x] is said


to be irreducible over F if it cannot be expressed as the product of two
polynomials of positive degree in F[x].

Thus, p(x) is irreducible iff it is not reducible.


When we say that a polynomial p(x) is irreducible, it is important that
we specify irreducible over the field F. A polynomial may be irreducible
over F, yet reducible over a larger field E. For example, p(x) = x2 + 1 is
irreducible over R; but over C it has factors (x + i)(x − i).
We next state the analogs for polynomials of Euclid’s lemma and its
corollaries. The proofs are almost identical to their counterparts in Z;
therefore they are left as exercises.

2.3.3 Proposition (Euclid’s lemma for polynomials). Let p(x) be irre-


ducible. If p(x) | a(x)b(x), then p(x) | a(x) or p(x) | b(x).

2.3.4 Corollary. Let p(x) be irreducible. If p(x) | a1 (x)a2 (x) . . . an (x),


then p(x) | ai (x) for one of the factors ai (x) among a1 (x), . . . , an (x).

2.3.5 Corollary. Let q1 (x), . . . , qr (x) and p(x) be monic irreducible poly-
nomials. If p(x) | q1 (x), . . . qr (x), then p(x) is equal to one of the factors
q1 (x), . . . , qr (x).

2.3.6 Theorem (Factorization into irreducible polynomials). Every poly-


nomial a(x) of positive degree in F[x] can be written as a product

a(x) = kp1 (x)p2 (x) . . . pr (x),

where k is a constant in F and p1 (x), . . . , pr (x) are monic irreducible poly-


nomials of F[x].

If this were not true, we could choose a polynomial a(x) of lowest de-
gree among those which cannot be factored into irreducibles. If a(x) is

22
reducible, so a(x) = b(x)c(x) where b(x) and c(x) have lower degree than
a(x). But this means that b(x) and c(x) can be factored into irreducibles,
and therefore a(x) can also. This is contradictory.

2.3.7 Theorem (Unique factorization). If a(x) can be written in two ways


as a product of monic irreducibles, say

a(x) = kp1 (x)p2 (x) . . . pr (x) = lq1 (x)q2 (x) . . . qs (x)

then k = l, r = s, and each pi (x) is equal to a qj (x).

The proof is the same, in all major respects, as the corresponding proof
for Z ; it is left as an exercise.
In the next chapter we will be able to improve somewhat on the last two
results in the special cases of R[x] and C[x]. Also, we will learn more about
factoring polynomials into irreducibles.

23
EXERCISES

A. Examples of Factoring into Irreducible Factors


1. Factor x4 − 4 into irreducible factors over Q, over R, and over C.
2. Factor x6 − 16 into irreducible factors over Q, over R, and over C.
3. Find all the irreducible polynomials of degree ≤ 4 in Z2 [x].
4. Show that x2 + 2 is irreducible in Z5 [x]. Then factor x4 − 4 into
irreducible factors in Z5 [x] (by Theorem 2.3.6, it is sufficient to search for
monic factors).
5. Factor 2x3 + 4x + 1 in Z5 [x] (factor it as in Theorem 2.3.6).
6. In Z6 [x], factor each of the following into two polynomials of degree
1: x, x + 2, x + 3. Why is this possible?

B. Short Questions Relating to Irreducible Polynomials


Let F be a field. Explain why each of the following is true in F[x]:
1. Every polynomial of degree 1 is irreducible.
2. If a(x) and b(x) are distinct monic polynomials, they cannot be
associates.
3. Any two distinct irreducible polynomials are relatively prime.
4. If a(x) is irreducible, any associate of a(x) is irreducible.
5. If a(x) ̸= 0, a(x) cannot be an associate of 0.
6. In Zp [x], every nonzero polynomial has exactly p − 1 associates.
7. x2 + 1 is reducible in Zp [x] iff p = a + b where ab ≡ 1 (mod p).

C. Number of Irreducible Quadratics over a Finite Field


1. Without finding them, determine how many reducible monic quadrat-
ics there are in Z5 [x] [HINT: Every reducible monic quadratic can be
uniquely factored as (x + a)(x + b)].
2. How many reducible quadratics are there in Z5 [x]? How many irre-
ducible quadratics?
3. Generalize: How many irreducible quadratics are there over a finite

24
field of n elements?
4 How many irreducible cubics are there over a field of n elements?

D. Ideals in Domains of Polynomials


1. Let F be a field, and let J designate any ideal of F[x]. Prove that:
(a) Any two generators of J are associates.
(b) J has a unique monic generator m(x). An arbitrary polynomial
a(x) ∈ F [x] is in J iff m(x) | a(x).
(c) J is a prime ideal iff it has an irreducible generator.
(d) If p(x) is irreducible, then ⟨p(x)⟩ is a maximal ideal of F[x].
2. Let S be the set of all polynomials a0 + a1 x + . . . + an xn in F[x] which
satisfy a0 + a1 + . . . + an = 0. It has been shown (Chapter 1, Exercise E5)
that S is an ideal of F[x]. Prove that x − 1 ∈ S, and explain why it follows
that S = ⟨−1⟩.
3. Conclude from part 2 that F[x]/⟨x−1⟩ ∼
= F (see Chapter 24, Exercise
F4).
aij xi y j in two
P
4. Let F[x, y] denote the domain of all the polynomials
letters x and y, with coefficients in F. Let J be the ideal of F[x, y] which
contains all the polynomials whose constant coefficient in zero. Prove that
J is not a principal ideal. Conclude that Theorem 2.1.1 is not true in F[x, y]

D. Proof of the Unique Factorization Theorem


1 Prove Euclid’s lemma for polynomials.
2 Prove the two corollaries of Euclid’s lemma.
3 Prove the unique factorization theorem for polynomials.

F. A Method for Computing the gcd


Let a(x) and b(x) be polynomials of positive degree. By the division
algorithm, we may divide a(x) by b(x):

a(x) = b(x)q1 (x) + r1 (x).

25
1. Prove that every common divisor of a(x) and b(x) is a common divisor
of b(x) and r1 (x).
It follows from part 1 that the gcd of a(x) and b(x) is the same as the
gcd of b(x) and r1 (x). This procedure can now be repeated on b(x) and
r1 (x); divide b(x) by r1 (x):

b(x) = r1 (x)q2 (x)r2 (x),

next
r1 (x) = r2 (x)q3 (x) + r3 (x),
..
.
Finally,
rn−1 (x) = rn (x)qn+1 (x) + 0.
In other words, we continue to divide each remainder by the succeeding
remainder. Since the remainders continually decrease in degree, there must
ultimately be a zero remainder. But we have seen that

gcd[a(x), b(x)] = gcd[b(x), r1 (x)] = . . . = gcd[rn−1 (x), rn (x)].

Since rn (x) is a divisor of rn−1 (x), it must be the gcd of rn (x) and rn−1 .
Thus,
rn (x) = gcd[a(x), b(x)].
This method is called the euclidean algorithm for finding the gcd.
2. Find the gcd of x3 + 1 and x4 + x3 + 2x2 + x − 1. Express this gcd as
a linear combination of the two polynomials.
3. Do the same for x24 − 1 and x15 − 1.
4. Find the gcd of x3 + x2 + x + 1 and x4 + x3 + 2x2 + 2x in Z3 [x].

G. A Transformation of F[x]
Let G be the subset of F[x] consisting of all polynomials whose constant
term is nonzero. Let h : G → G be defined by

h(a0 + a1 x + . . . + an xn ) = an + an−1 x + . . . + a0 xn .

26
1. Prove that:
(a) h preserves multiplication, that is, h[a(x)b(x)] = h[a(x)]h[b(x)].
(b) h is injective and surjective and h.h = 1G .
(c) a0 + a1 x + . . . + an xn is irreducible iff an + an−1 x + . . . + a0 xn is
irreducible.
2. Let a0 + a1 x + . . . + an xn = (b0 + . . . + bm xm )(c0 + . . . + cq xq ). Factor
an + an−1 x + . . . + a0 xn .
3. Let a(x) = a0 + a1 x + . . . + an xn and b
a(x) = an + an−1 x + . . . + a0 xn .
If c ∈ F , prove that a(c) = 0 iff b
a(1/c) = 0.

27
CHAPTER 3

SUBSTITUTION IN POLYNOMIALS

3.1 Polynomial Functions

Up to now we have treated polynomials as formal expressions. If a(x) is


a polynomial over a field F, say a(x) = a0 +a1 x+. . .+an xn , this means that
the coefficients a0 , a1 , . . . , an are elements of the field F, while the letter x
is a placeholder which plays no other role than to occupy a given position.
When we dealt with polynomials in elementary algebra, it was quite
different. The letter x was called an unknown and was allowed to assume
numerical values. This made a(x) into a function having x as its indepen-
dent variable. Such a function is called a polynomial function.
This chapter is devoted to the study of polynomial functions. We begin
with a few careful definitions.
Let a(x) = a0 + a1 x + . . . + an xn be a polynomial over F. If c is any
element of F, then a0 + a1 c + . . . + an cn is also an element of F, obtained
by substituting c for x in the polynomial a(x). This element is denoted by
a(c). Thus, a(c) = a0 + a1 c + . . . + an cn .
Since we may substitute any element of F for x, we may regard a(x) as
a function from F to F. As such, it is called a polynomial function on F.
The difference between a polynomial and a polynomial function is mainly
a difference of viewpoint. Given a(x) with coefficients in F: if x is regarded
merely as a placeholder, then a(x) is a polynomial; if x is allowed to assume
values in F, then a(x) is a polynomial function. The difference is a small
one, and we will not make an issue of it.

28
If a(x) is a polynomial with coefficients in F, and c is an element of F
such that a(c) = 0 then we call c a root of a(x). For example, 2 is a root
of the polynomial 3x2 + x − 14 ∈ R[x], because 3.22 + 2 − 14 = 0.
There is an absolutely fundamental connection between roots of a poly-
nomial and factors of that polynomial. This connection is explored in the
following pages, beginning with the next theorem:

3.1.1 Theorem. Let a(x) be a polynomial over a field F. c is a root of


a(x) iff x − c is a factor of a(x).

Proof. If x − c is a factor of a(x), this means that a(x) = (x − c)q(x) for


some q(x). Thus, a(c) = (c − c)q(c) = 0, so c is a root of a(x).
Conversely, if c is a root of a(x), we may use the division algorithm to
divide a(x) by x − c:

a(x) = (x − c)q(x) + r(x).

The remainder r(x) is either 0 or a polynomial of lower degree than x − c;


but lower degree than x − c means that r(x) is a constant polynomial:
r(x) = r ∈ F. Then

0 = a(c) = (c − c)q(c) + r = 0 + r = r

Thus, r = 0, and therefore x − c is a factor of a(x).

Theorem 3.1.1 tells us that if c is a root of a(x), then x − c is a factor


of a(x) (and vice versa). This is easily extended: if c1 and c2 are two roots
of a(x), then x − c1 and x − c2 are two factors of a(x). Similarly, three
roots give rise to three factors, four roots to four factors, and so on. This
is stated concisely in the next theorem.

3.1.2 Theorem. If a(x) has distinct roots cl , . . . , cm in F, then

(x − c1 )(x − c2 ) . . . (x − cm )

is a factor of a(x).

29
Proof. To prove this, let us first make a simple observation: if a polynomial
a(x) can be factored, any root of a(x) must be a root of one of its factors.
Indeed, if a(x) = s(x)t(x) and a(c) = 0, then s(c)t(c) = 0, and therefore
either s(c) = 0 or t(c) = 0.
Let cl , . . . , cm be distinct roots of a(x). By Theorem 3.1.1,

a(x) = (x − c1 )q1 (x).

By our observation in the preceding paragraph, c2 must be a root of x − c1


or of q1 (x). It cannot be a root of x − c1 because c2 − c1 ̸= 0; so c2 is a root
of q1 (x). Thus, q1 (x) = (x − c2 )q2 (x), and therefore

a(x) = (x − c1 )(x − c2 )q2 (x).

Repeating this argument for each of the remaining roots gives us our result.

An immediate consequence is the following important fact:

3.1.3 Theorem. If a(x) has degree n, it has at most n roots.

Proof. If a(x) had n + 1 roots c1 , . . . , cn+1 , then by Theorem 3.1.2, (x −


c1 ) . . . (x − cn+1 ) would be a factor of a(x), and the degree of a(x) would
therefore be at least n + 1.

It was stated earlier in this chapter that the difference between polyno-
mials and polynomial functionsis mainly a difference of viewpoint. Mainly,
but not entirely! Remember that two polynomials a(x) and b(x) are equal
iff corresponding coefficients are equal, whereas two functions a(x) and b(x)
are equal iff a(x) = b(x) for every x in their domain. These two notions of
equality do not always coincide!
For example, consider the following two polynomials in Z5 [x]:

a(x) = x5 + 1

b(x) = x − 4.

30
You may check that a(0) = b(0), a(1) = b(1), . . . , a(4) = b(4); hence a(x)
and b(x) are equal functions from Z5 to Z5 . But as polynomials, a(x) and
b(x) are quite distinct (they do not even have the same degree)!
It is reassuring to know that this cannot happen when the field F is
infinite. Suppose a(x) and b(x) are polynomials over a field F which has
infinitely many elements. If a(x) and b(x) are equal as functions, this
means that a(c) = b(c) for every c ∈ F . Define the polynomial d(x) to
be the difference of a(x) and b(x): d(x) = a(x) − b(x). Then d(c) = 0
for every c ∈ F. Now, if d(x) were not the zero polynomial, it would be a
polynomial (with some finite degree n) having infinitely many roots, and
by Theorem 3.1.3 this is impossible! Thus, d(x) is the zero polynomial (all
its coefficients are equal to zero), and therefore a(x) is the same polynomial
as b(x) (they have the same coefficients).
This tells us that if F is a field with infinitely many elements (such as
Q, R, or C), there is no need to distinguish between polynomials and poly-
nomial functions. The difference is, indeed, just a difference of viewpoint.

3.2 Polynomials over Z and Q

In scientific computation a great many functions can be approximated by


polynomials, usually polynomials whose coefficients are integers or rational
numbers. Such polynomials are therefore of great practical interest. It is
easy to find the rational roots of such polynomials, and to determine if a
polynomial over Q is irreducible over Q. We will do these things next.
First, let us make an important observation: Let a(x) be a polynomial
with rational coefficients, say
k0 k1 kn
a(x) = + x + . . . + xn .
l0 l1 ln
We may now factor out s from all but the first term to get
1
a(x) = (k0 l0 . . . ln + . . . + kn l0 . . . ln−1 xn ).
l0 . . . ln | {z }
b(x)

31
The polynomial b(x) has integer coefficients; and since it differs from a(x)
only by a constant factor, it has the same roots as a(x). Thus, for every
polynomial with rational coefficients, there is a polynomial with integer
coefficients having the same roots. Therefore, for the present we will confine
our attention to polynomials with integer coefficients. The next theorem
makes it easy to find all the rational roots of such polynomials:

3.2.1 Theorem. Let s/t be a rational number in simplest form (that is,
the integers s and t do not have a commonfactor greater than 1). Let
a(x) = a0 + . . . + an xn be a polynomial with integer coefficients. If s/t is a
root of a(x), then s | a0 and t | an .

Proof. If s/t is a root of a(x), this means that


s s
a0 + a1 . . . + an ( )n = 0.
t t
Multiplying both sides of this equation by tn we get

a0 tn + a1 stn−1 + . . . + an sn = 0 (3.2.1.1)

We may now factor out s from all but the first term to get

−a0 tn = s(al tn−1 + . . . + an sn−1 ).

Thus, s | a0 tn ; and since s and t have no common factors, s | a0 . Similarly,


in Equation 3.2.1.1, we may factor out t from all but the last term to get

t(a0 tn−1 + . . . + an−1 sn−1 ) = an sn .

Thus, t | an sn ; and since s and t have no common factors, t | an .

As an example of the way Theorem 3.2.1 may be used, let us find the
rational roots of

a(x) = 2x4 + 7x3 + 5x2 + 7x + 3.

Any rational root must be a fraction s/t where s is a factor of 3 and t is a


1 3
factor of 2. The possible roots are therefore ±1, ±3, ± and ± . Testing
2 2
32
each of these numbers by direct substitution into the equation a(x) = 0,
1
we find that − and −3 are roots.
2
Before going to the next step in our discussion we note a simple but
fairly surprising fact.

3.2.2 Lemma. Let a(x) = b(x)c(x), where a(x), b(x), and c(x) have integer
coefficients. If a prime number p divides every coefficient of a(x), it either
divides every coefficient of b(x) or every coefficient of c(x).

Proof. If this is not the case, let br be the first coefficient of b(x) not divisible
by p, and let ct be the first coefficient of c(x) not divisible by p. Now,
a(x) = b(x)c(x), so

ar+t = b0 cr+t + . . . + br ct + . . . + br+t c0 .

Each term on the right, except br ct is a product bi cj where either i < r or


j < t. By our choice of br and ct , if i < r then p | bi , and j < t then p | cj .
Thus, p is a factor of every term on the right with the possible exception
of br ct but p is also a factor of ar+t . Thus, p must be a factor of br ct hence
of either br or ct and this is impossible.

We saw (in the discussion immediately preceding Theorem 3.2.1) that


any polynomial a(x) with rational coefficients has a constant multiple ka(x),
with integer coefficients, which has the same roots as a(x). We can go one
better.

3.2.3 Theorem. Let a(x) ∈ Z[x]. Suppose a(x) can be factored as

a(x) = b(x)c(x),

where b(x) and c(x) have rational coefficients. Then there are polynomials
B(x) and C(x) with integer coefficients, which are constant multiples of
b(x) and c(x), respectively, such that

a(x) = B(x)C(x).

33
Proof. Let k and l be integers such that kb(x) and lc(x) have integer coef-
ficients. Then kla(x) = [kb(x)][lc(x)]. By Lemma 3.2.2, each prime factor
of kl may now be canceled with a factor of either kb(x) or lc(x).

3.3 Eisenstein’s Irreducibility Criterion

Remember that a polynomial a(x) of positive degree is said to be re-


ducible over F if there are polynomials b(x) and c(x) in F[x], both of pos-
itive degree, such that a(x) = b(x)c(x). If there are no such polynomials,
then a(x) is irreducible over F.
If we use this terminology, Theorem 3.2.3 states that any polynomial
with integer coefficients which is reducible over Q is reducible already over
Z.
In Chapter 2 we saw that every polynomial can be factored into ir-
reducible polynomials. In order to factor a polynomial completely (that
is, into irreducibles), we must be able to recognize an irreducible polyno-
mial when we see one! This is not always an easy matter. But there is a
method which works remarkably well for recognizing when a polynomial is
irreducible over Q:

3.3.1 Theorem (Eisenstein’s irreducibility criterion). Let

a(x) = a0 + a1 x + . . . + an xn

be a polynomial with integer coefficients. Suppose there is a prime num-


ber p which divides every coefficient of a(x) except the leading cofficient
an ; suppose p does not divide an and p2 does not divide a0 . Then a(x) is
irreducible over Q.

Proof. If a(x) can be factored over as a(x) = b(x)c(x), then by Theorem


3.2.3 we may assume b(x) and c(x) have integer coefficients: say b(x) =
b0 + . . . + bk xk and c(x) = c0 + . . . + cm xm . Now, a0 = b0 c0 ; p divides a0
but p2 does not, so only one of b0 , c0 is divisible by p. Say p | c0 and p ∤ b0 .
Next, an = bk cm and p ∤ an , so p ∤ cm .

34
Let s be the smallest integer such that p ∤ cs . We have

as = b0 cs + b1 cs−1 + . . . + bs c0

and by our choice of cs , every term on the right except b0 cs is divisible by p.


But as also is divisible by p, and therefore b0 cs must be divisible by p. This
is impossible because p ∤ b0 and p ∤ cs . Thus, a(x) cannot be factored.

For example, x3 + 2x2 + 4x + 2 is irreducible over because p = 2 satisfies


the conditions of Eisenstein’s criterion.

3.4 Polynomials over R and C

One of the most far-reaching theorems of classical mathematics concerns


polynomials with complex coefficients. It is so important in the frame
work of traditional algebra that it is called the fundamental theorem of
algebra. It states the following: Every nonconstant polynomial with complex
coefficients has a complex root.
It follows immediately that the irreducible polynomials in C[x] are ex-
actly the polynomials of degree 1. For if a(x) is a polynomial of degree
greater than 1 in C[x], then by the fundamental theorem of algebra it has
a root c and therefore a factor x − c.
Now, every polynomial in C[x] can be factored into irreducibles. Since
the irreducible polynomials are all of degree 1, it follows that if a(x) is a
polynomial of degree n over C, it can be factored into

a(x) = k(x − c1 )(x − c2 ) . . . (x − cn ).

In particular, if a(x) has degree n it has n (not necessarily distinct) complex


roots c1 . . . cn .
Since every real number a is a complex number (a = a + 0i), what
has just been stated applies equally to polynomials with real coefficients.
Specifically, if a(x) is a polynomial of degree n with real coefficients, it can

35
be factored into a(x) = k(x − c1 ) . . . (x − cn ), where cl , . . . , cn are complex
numbers (some of which may be real).
For our closing comments, we need the following lemma:

3.4.1 Lemma. Suppose a(x) ∈ R[x]. If a + bi is a root of a(x), so is a − bi.

Proof. Remember that a − bi is called the conjugate of a + bi. If r is


any complex number, we write r for its conjugate. It is easy to see that
the function F(r) = r is a homomorphism from C to C (in fact, it is an
isomorphism). For every real number a, F(a) = a. Thus, if a(x) has real
coefficients, then

F(a0 + a1 r + . . . + an rn ) = a0 + a1 r + . . . + an rn .

Since F(0) = 0, it follows that if r is a root of a(x), so is r.

Now let a(x) be any polynomial with real coefficients, and let r = a + bi
be a complex root of a(x). Then r is also a root of a(x), so

(x − r)(x − r) = x2 − 2ax + (a2 + b2 )

and this is a quadratic polynomial with real coefficients! We have thus


shown that any polynomial with real coefficients can be factored into poly-
nomials of degree 1 or 2 in R[x]. In particular, the irreducible polyno-
mials over R are linear polynomials or irreducible quadratics (that is, the
ax2 + bx + c where b2 − 4ac < 0).

3.5 Solution of Cubic and Quartic Equations by For-


mulas

In this section we focus on polynomials that have their coefficients in


the field R of real numbers. Up to this point, results have been stated
with emphasis on the zeros of polynomials or on the related property of
irreducibility.

36
We now place emphasis on a different point of view. Finding the zeros
of a polynomial

f (x) = a0 + a1 x + . . . + an−1 xn−1 + an xn

is equivalent to finding the solutions of the equation

a0 + a1 x + . . . + an−1 xn−1 + an xn = 0.

Historically, mathematics developed with emphasis on the solution of equa-


tions.
The solution of linear equations

ax + b = 0

by the formula
b
x=−
a
and the solution of quadratic equations

ax2 + bx + c = 0

by the formula √
b2 − 4ac
−b ±
x=
2a
long ago prompted mathematicians to seek similar formulas for equations
of higher degree with real coefficients.
In the 16th century, Italian mathematicians named Ferro, Tartaglia, Fer-
rari, and Cardano developed methods for solving third- and fourth-degree
equations with real coefficients by the use of formulas that involved only
the operations of addition, subtraction, multiplication, division, and the
extraction of roots. For more than two hundred years afterward, mathe-
maticians struggled to obtain similar formulas for equations with degree
higher than 4 or to prove that such formulas did not exist. It was in the
early 19th century that the Norwegian mathematician Abel„ proved that

37
it was impossible to obtain such formulas for equations with degree greater
than 4.
The proof of Abel’s result is beyond the level of this text, but the for-
mulas for cubic and quartic (third- and fourth-degree) equations with real
coefficients are within our reach.
We consider first the solution of the general cubic equation

a0 + a1 x + a2 x2 + a3 x3 = 0

where the coefficients are real numbers and a3 ̸= 0. There is no loss of


generality in assuming that the cubic polynomial is monic since division of
both sides of the equation by a3 yields an equivalent equation. Thus we
assume an equation of the form

x3 + ax2 + bx + c = 0.

As would be expected, cube roots of complex numbers play a major role


in the development. For this reason, some remarks on cube roots are in
order. We have the cube roots of 1 are given by

cos 0 + i sin 0 = 1,

2π 2π −1 + i 3
cos + i sin = ,
3 3 2

4π 4π −1 − i 3
cos + i sin = .
3 3 2

−1 + i 3
If we let ω = the distinct cube roots of 1 are ω, ω 2 , and
2 √
ω 3 = 1. For an arbitrary nonzero complex number z, let 3 z denote any
√ √ √
fixed cube root of z. Then each of the numbers 3 z, ω 3 z, and ω 2 3 z and is
a cube root of z, and they are clearly distinct. Thus the three cube roots
of z are given by
√ √ √
3
z, ω 3 z, ω 2 3 z,

−1 + i 3
where ω = . This result is used in solving the cubic equation in
2
Theorem 3.5.2.

38
The following two theorems lead to formulas for the solutions of the
general cubic equation

x3 + ax2 + bx + c = 0.

3.5.1 Theorem (Change of Variable in the Cubic). The change of variable


a
x=y−
3
in x3 + ax2 + bx + c = 0 yields the equation

y 3 + py + q = 0,
a2 ab 2a3
where p = b − , q = c − + .
3 3 27
Proof. The theorem can be proved by direct substitution, but the details
are neater if we first consider a substitution of the form x = y + h, where
h is unspecified at this point. This substitution yields

(y + h)3 + a(y + h)2 + b(y + h) + c = 0.

When this equation is simplified, it appears as

y 3 + (3h + a)y 2 + (3h2 + 2ah + b)y + (h3 + ah2 + bh + c) = 0.


a
If we let h = − the coefficients then simplify as follows:
3
a
3h + a = 3(− ) + a = 0,
3
2 a 2 a a2
3h + 2ah + b = 3(− ) + 2a(− ) + b = b − ,
3 3 3
a a a ab 2a3
h3 + ah2 + bh + c = (− )3 + a(− )2 + b(− ) + c = c − + .
3 3 3 3 27
This establishes the theorem.

3.5.2 Theorem (Solutions to the Cubic Equation). Consider the equation

y 3 + py + q = 0,

39
√ r  r 
−1 + i 3 q q 2  p 3 q q 2  p 3
and let ω = ,A=− + + ,B=− − + .
2 2 2 3 2 2 3
The solutions to y 3 + py + q = 0 are given by
√ √ √ √ √ √
A + B, ω A + ω 2 B, ω 2 A + ω B
3 3 3 3 3 3

√ √
where 3 A and 3 B denote (real or complex) cube roots of A and B chosen
√ √ p
so that 3 A 3 B = − .
3
Proof. For an efficient proof, we resort to a “trick” substitution: We let
p
y=z−
3z
in y 3 + py + q = 0. This substitution yields
 p 3  p
z− +p z− + q = 0.
3z 3z
This equation then simplifies to
3 p3
z − +q =0
27z 3
and then to
6 3 p3
z + qz − = 0.
27
This is a quadratic equation in z 3 and we can use the quadratic formula to
obtain r
4p3
−q ± q2 + q
r 
q 2  p 3
z3 = 27 = − ± + .
2 2 2 3
With A and B as given in the statement of the theorem, we have

z 3 = A or z 3 = B.

Noting that
r  ! r  !
q q 2  p  3 q q 2  p  3
AB = − + + − − +
2 2 3 2 2 3
 q 2   q 2  p 3 
= − +
2 2 3
p3
=− ,
27
40

3

3
we see that A and B need to be chosen so that

3

3 2

3

3

3 2

3
A, ω A, ω A, B, ω B, ω B.

Substituting these values in



3

p A3B
y=z− =z+ ,
3z z
1 1
and using = ω 2 or 2 = ω we obtain the following three solutions for y :
ω ω
√3
√3

3

2 3

2 3
√3
A + B, ω A + ω B, ω A + ω B.

3.5.3 Example. We shall use the formulas in Theorem 3.5.2 to solve the
equation
y 3 − 9y − 12 = 0.

We have p = −9 and q = −12. Thus


12
q √
A= + (−6)2 + (−3)3 = 6 + 9 = 9,
2

B =6− 9=3
√ √ √ √ p
and the real cube roots 3 9 and 3 3 satisfy 3 9 3 3 = − . The solutions are
3
given by
√3

9 + 3 3,
√ ! √ !2
√ √ −1 + i 3 3 √ −1 + i 3 √
ω 9 + ω2 3 =
3 3 3
9+ 3
2 2
√ ! √ !
−1 + i 3 √ 3 −1 − i 3 √
3
= 9+ 3
2 2

1 √3
√3 i 3 √
3
√ 3
= − ( 9 + 3) + ( 9 − 3),
2 2

41
√ !2 √ !

2 3

3 −1 + i 3 √ 3 −1 + i 3 √ 3
ω 9+ω 3= 9+ 3
2 2
√ ! √ !
−1 − i 3 √ 3 −1 + i 3 √3
= 9+ 3
2 2

1 √ 3

3 i 3 √
3
√ 3
= − ( 9 + 3) − ( 9 − 3),
2 2
The results of Theorem 3.5.1 and Theorem 3.5.2 combine to yield the
following theorem. The formulas in the theorem are known as Cardano’s
Formulas.

3.5.4 Theorem (Cardano’s Formulas). The solutions to the cubic equation

x3 + ax2 + bx + c = 0,

are given by
√ √ a √ √ a √ √ a
B − , ω A + ω2 B − , ω2 A + ω B − ,
3 3 3 3 3 3
A+
3 3 3
where √
−1 + i 3 a2 ab a3
ω= ,p = b − ,q = c − + ,
2 3 3 27
r  r
q q 2  p 3 q  q 2  p 3
A=− + + ,B = − − + ,
2 2 3 2 2 3
√ √ √ √ p
with 3 A and 3 B chosen so that 3 A 3 B = − .
3
The use of Theorems 3.5.4 is demonstrated in the following example.

3.5.5 Example. For the equation

x3 − 3x2 − 6x − 4 = 0.

we have a = −3, b = −6 and c = −4. The formulas in Theorem 3.5.4 yield


9
p = −6 − = −9,
3
18 54
q = −4 − − = −12,
3 27
42
q
A=6+ (−6)2 + (−3)3 = 9,
q
B = 6 − (−6)2 + (−3)3 = 3.
√ √ √ √ p
The real cube roots 3 9 and 3 3 satisfy 3 A 3 B = − , and the solutions
3
are given by
√3

9 + 3 3 + 1,

√ √ 1 √ √ i 3 √ √
ω 9 + ω 2 3 + 1 = − (ω 9 + ω 2 3 − 2) + (ω 9 − ω 2 3),
3 3 3 3 3 3

2 2

√3

2 3 1 √ 3

2 3 i 3 √ 3

2 3
ω 9 + ω 3 + 1 = − (ω 9 + ω 3 − 2) − (ω 9 − ω 3).
2 2
We turn our attention now to the solution of quartic equations. As in
the case of the cubic equation, there is no loss of generality in assuming
that the equation is monic. Thus we assume an equation of the form

x4 + ax3 + bx2 + cx + d = 0.

We find again that an appropriate substitution will remove the term of


second-highest degree.

3.5.6 Theorem (Change of Variable in the Quartic). The change of vari-


able
a
x=y−
4
in x4 + ax3 + bx2 + cx + d = 0 yields an equation of the form

y 4 + py 2 + qy + r = 0.

Theorem 3.5.6 can be proved by direct substitution, and this proof is


left as an exercise. In contrast to Theorem 3.5.1, we are not interested in
formulas for p, q, and r at this time.
Consider now an equation of the form

y 4 + py 2 + qy + r = 0

43
which can be written as

y 4 = −py 2 − qy − r.

The basic idea of our method, which was devised by Ferrari, is to add an
expression to each side of the last equation that will make both sides perfect
squares (squares of binomials). With this idea in mind, we add

2 t2
ty +
4
to both sides, where t is yet to be determined. This gives
t2 t2
y 4 + ty 2 + = −py 2 − qy − r + ty 2 + ,
4 4
or 2
t2
  
2 t 2
y + = (t − p)y − qy + −r .
2 4
We recall that a quadratic polynomial Ay 2 + By + C is the square of a
binomial
Ay 2 + By + C = (Dy + E)2

if and only if B 2 − 4AC = 0. Thus


 2 
t
(t − p)y 2 − qy + − r = (Dy + E)2
4
if and only if
t2
 
(−q 2 ) − 4(t − p) −r = 0.
4
This equation simplifies to the equation

t3 − pt2 − 4rt + 4rp − q 2 = 0,

which is known as the resolvent equation for y 4 + py 2 + qy + r = 0.


The resolvent equation can be solved for t by Cardano’s method. Any
one of the three solutions for t may be used in
t 2
   2 
2 2 t
y + = (t − p)y − qy + −r
2 4
44
to obtain an equation of the form
 2  2 
t t
y2 + = (t − p)y 2 − qy + − r = (Dy + E)2 .
2 4
The solutions to the original equation can then be found by solving the two
quadratic equations
   
2 t 2 t
y + = Dy + E and y + = −Dy − E.
2 2
The method is illustrated in the following example.

3.5.7 Example. We illustrate the preceding discussion by solving the equa-


tion
y 4 + y 2 − 2y + 6 = 0.

We have p = 1, q = −2 and r = 6. The resolvent equation is given by

t3 − t2 − 24t + 20 = 0.

We find that t = 5 is a solution to the resolvent equation, and the equation

t 2
   2 
2 2 t
y + = (t − p)y − qy + −r
2 4
becomes
5 2 1 2
   
2 2 1
y + = 4y + 2y + = 2y + .
2 4 2
Equating square roots, we obtain
 
5 1 5 1
y 2 + = 2y + or y 2 + = − 2y +
2 2 2 2
and then
y 2 − 2y + 2 = 0 or y 2 + 2y + 3 = 0.

The quadratic formula then yields



y = 1 ± i and y = −1 ± i 2

as the solutions of the original equation.

45
We can now describe a method of solution for an arbitrary quartic equa-
tion
x4 + ax3 + bx2 + cx + d = 0.

We first make the substitution


a
x=y−
4
and obtain an equation for the form

y 4 + py 2 + qy + r = 0.

We next use the method of Example 3.5.7 to find the four solutions y1 , y2 , y3
and y4 of the equation in y. Then the solutions to the original equation are
given by
a
xj = yj − , for j = 1, 2, 3, 4.
4
This is illustrated in Example 3.5.8.

3.5.8 Example. Consider the equation

x4 + 4x3 + 7x2 + 4x + 6 = 0.
a
The substitution formula x = y − yields x = y − 1 and the resulting
4
equation

(y − 1)4 + 4(y − 1)3 + 7(y − 1)2 + 4(y − 1) + 6 = 0.

This equation simplifies to

y 4 + y 2 − 2y + 6 = 0.

From Example 3.5.7, the solutions to the last equation are


√ √
y1 = 1 + i, y2 = 1 − i, y3 = −1 + i 2, y4 = −1 − i 2.

Hence xj = yj − 1 the solutions are given by


√ √
x1 = i, x2 = −i, x3 = −2 + i 2, x4 = −2 − i 2.

46
Just as the discriminant b2 −4ac can be used to characterize the solutions
of the quadratic equation ax2 +bx+c = 0, the discriminant of a polynomial
equation can be used to characterize its solutions. In particular, we will see
that a cubic equation will have either exactly one real solution or exactly
three real solutions. We begin with the next definition.

3.5.9 Definition (Discriminant of a Cubic Polynomial). Let

f (y) = y 3 + py + q

have zeros c1 , c2 , and c3 . The discriminant of f (y) is D2 where


Y
D= (ci − cj ) = (c1 − c2 )(c1 − c3 )(c2 − c3 ).
i<j

The reason for defining the discriminant as D2 , rather than as D, is


because the sign of D depends on the order of the zeros. However, the sign
of D2 is independent on the order of the zeros.

3.5.10 Theorem (Discriminant of a Cubic Polynomial). The discriminant


of f (y) = y 3 + py + q is D2 = −27q 2 − 4p3 .

Proof. Let c1 , c2 , and c3 be zeros of f (y) = y 3 + py + q. Then we can write

f (y) = (y − c1 )(y − c2 )(y − c3 )

where

3

3
c1 = A+
B,
√ √
c2 = ω A + ω 2 B,
3 3

√ √
c3 = ω 2 A + ω B
3 3

and √
−1 + i 3
ω= ,
2
r 
q q 2  p 3
A=− + + ,
2 2 3

47
r 
q q 2  p 3
B=− − + .
2 2 3
The discriminant is

D2 = (c1 − c2 )2 (c1 − c3 )2 (c2 − c3 )2

and using ω 3 = 1, we have


√ √ √ √
c1 − c2 = ( A + B) − (ω A + ω 2 B)
3 3 3 3

√ √ √ √
= A + B − ω A − ω2 B
3 3 3 3

√ √
= (1 − ω)( A − ω 2 B),
3 3

√ √ √ √
c1 − c3 = ( A + B) − (ω 2 A + ω B)
3 3 3 3

√ √ √ √
= A + B − ω2 A − ω B
3 3 3 3

√ √
= −ω 2 (1 − ω)( A − ω B),
3 3

√ √ √ √
c2 − c3 = (ω A + ω 2 B) − (ω 2 A + ω B)
3 3 3 3

√ √ √ √
= ω A + ω2 B − ω2 A − ω B
3 3 3 3

√ 3

3
= ω(1 − ω)( A − B).
Then

√ √ √ √ √ √
D = −(1 − ω)( A − ω 2 B)ω 2 (1 − ω)( A − ω B)ω(1 − ω)( A − B)
3 3 3 3 3 3

√ √ √ √ √ √
= −ω 3 (1 − ω)3 ( A − ω 2 B)( A − ω B)( A − B)
3 3 3 3 3 3

√ √ √ √ √ √ √
= 3i 3( A − ω 2 B)( A − ω B)( A − B)
3 3 3 3 3 3

√ √ 3

3 3
√ √
3 3 3
√3
√3

3

3
= 3i 3( A − ω B + ω A B − A B − ω A2 B
3 3 2 2
√ √3

3
√ √ √3
+ ω A B 2 − ω 2 A2 B + ω 2 A B 2 )
3 3 3

√ √ 3

3
√ 3
√3
= 3i 3(A − B + A B − A B 2 2
√3
√ √ √3

3
√ √ √
3
− ω( A2 B − A B 2 ) − ω 2 ( A2 B − A B 2 ))
3 3 3 3

√ √3
√ √ √
3
= 3i 3(A − B + (−1 − ω − ω 2 )( A2 B − A B 2 ))
3 3


= 3i 3(A − B)

r
4
= 3i 3 q 2 + p3
27
48
since −1 − ω − ω 2 = 0 and
r  r  !
q q 2  p  3 q q 2  p  3
A−B =− + + − − − +
2 2 3 2 2 3
r 
q 2  p 3
=2 +
r 2 3
4
= q 2 + p3 .
27
Thus !2

r
4 3
D2 = 3i 3 q 2 + p
27
= −27q 2 − 4p3 .

The result of Theorem 3.5.10 can be used to characterize the solutions


to the polynomial equation

y 3 + py + q = 0.

3.5.11 Theorem (Real Solutions of a Cubic Equation). The equation y 3 +


py + q = 0 has exactly three real solutions if and only if D2 ≥ 0; that is, if
and only if −27q 2 − 4p3 ≥ 0.

Proof. Let c1 , c2 , and c3 be real solutions to y 3 + py + q = 0. Then

D = (c1 − c2 )(c1 − c3 )(c2 − c3 )

is real, and the discriminant D2 ≥ 0.


Now assume that there are exactly one real solution c1 and two nonreal
solutions c2 and c3 . We know that the nonreal solutions must be conjugates,
so let c2 = z = a + bi and c3 = z = a − bi. Then
D = (c1 − z)(c1 − z)(z − z)
= (c1 − (a + bi))(c1 − (a − bi))((a + bi) − (a − bi))
= 2bi((a − c1 )2 + b2 ).

49
and
D2 = (2bi((a − c1 )2 + b2 ))2
= −4b2 ((a − c1 )2 + b2 )2 < 0
since b ̸= 0. Thus, if there is a nonreal solution, then the discriminant
is negative. It follows that if the discriminant is nonnegative, then the
solutions must all be real.

We note that the discriminant for the polynomial y 3 −9y−12 in Example


3.5.3 with two nonreal zeros is

D2 = 227q 2 − 4p3 = −27(−12)2 − 4(−9)3 = −972 < 0.

50
EXERCISES

A. Finding Roots of Polynomials over Finite Fields

In order to find a root of a(x) in a finite field F, the simplest method (if
F is small) is to test every element of F by substitution into the equation
a(x) = 0.
1. Find all the roots of the following polynomials in Z5 [x], and factor
the polynomials:

x3 + x2 + x + 1; 3x4 + x2 + 1; x5 + 1; x4 + 1; x4 + 4.

2. Use Fermat’s theorem to find all the roots of the following polynomials
in Z7 [x]:
x100 − 1; 3x98 + x19 + 3; 2x74 − x55 + 2x + 6.
3. Using Fermat ’s theorem, find polynomials of degree ≤ 6 which
determine the same functions as the following polynomials in Z7 [x]:

3x75 − 5x54 + 2x13 − x2 ; 4x108 + 6x101 − 2x81 ; 3x103 − x73 + 3x55 − x25 .

4. Explain why every polynomial in Zp [x] has the same roots as a poly-
nomial of degree < p.

B. Finding Roots of Polynomials over Q


1. Find all the rational roots of the following polynomials, and factor
them into irreducible polynomials in Q[x]:

9x3 + 18x2 − 4x − 8; 4x3 − 3x2 − 8x + 6;

2x4 + 3x3 − 8x − 12; 6x4 − 7x3 + 8x2 − 7x + 2.


2. Factor each of the preceding polynomials in R[x] and in C[x].
3. Find associates with integer coefficients for each of the following
polynomials:
3 4 2 1 1 1 1 √ 3 1 √ 1
x3 + x2 − x − ; x3 − x2 − x + ; 3x + √ x2 − 3x − √ .
2 9 3 2 4 2 4 3 3
51
4. Find all the rational roots of the polynomials in part 3 and factor
them over R.
5. Does 2x4 + 3x2 − 2 have any rational roots? Can it be factored into
two polynomials of lower degree in Q[x]? Explain.

C. Short Questions Relating to Roots

Let F be a field.
1. Prove that the following statements are true in F[x]:
(a) The remainder of p(x), when divided by x − c, is p(c).
(b) (x − c) | (p(x) − p(c)).
(c) Every polynomial has the same roots as any of its associates.
2. If a(x) and b(x) have the same roots in F, are they necessarily asso-
ciates? Explain.
3. Prove: If a(x) is a monic polynomial of degree n, and a(x) has n
roots c1 , . . . , cn ∈ F , then a(x) = (x − c1 ) . . . (x − cn ).
4. Suppose a(x) and b(x) have degree < n. If a(c) = b(c) for n values of
c, prove that a(x) = b(x).
7. There are infinitely many irreducible polynomials in Z5 [x].
8. How many roots does x2 − x have in Z10 ? In Z11 ? Explain the
difference.

D. Irreducible Polynomials in Q[x] by Eisenstein’s Criterion (and


Variations on the Theme)

1. Show that each of the following polynomials is irreducible over Q:


2 5 1 4 1
3x4 − 8x3 + 6x2 − 4x + 6; x + x − 2x2 + ;
3 2 2
1 4 1 3 2 1 4 4 3 2 2
x − x − x + 1; x + x − x + 1.
5 3 3 2 3 3
2. It often happens that a polynomial a(y), as it stands, does not satisfy
the conditions of Eisenstein’s criterion, but with a simple change of variable

52
y = x + c, it does. It is important to note that if a(x) can be factored into
p(x)q(x), then certainly a(x+c) can be factored into p(x+c)q(x+c). Thus,
the irreducibility of a(x + c) implies the irreducibility of a(x).
(a) Use the change of variable y = x + 1 to show that x4 + 4x + 1 is
irreducible in Q[x] (In other words, test (x+1)4 +4(x+1)+1 by Eisenstein’s
criterion).
(b) Find an appropriate change of variable to prove that the following
are irreducible in Q[x]:

x4 + 2x2 − 1; x3 − 3x + 1; x4 + 1; x4 − 10x2 + 1.

3. Prove that for any prime p, xp − 1 + xp − 2 + . . . + x + 1 is irreducible


in Q[x].
HINT: By elementary algebra,

(x − 1)(xp − 1 + xp − 2 + . . . + x + 1) = xp − 1

hence
p−1 p−2 xp − 1
x + x + ... + x + 2 = .
x−1
Use the change of variable y = x + 1, and expand by the binomial theorem.
4. By Exercise
G1(c)

of Chapter 2, the function

h(a0 + a1 x + . . . + an xn ) = an + an−1 x + . . . + a0 xn

restricted to polynomials with nonzero constant term matches irreducible


polynomials with irreducible polynomials. Use this fact to state a dual
version of Eisenstein’s irreducibility criterion.
5. Use part 4 to show that each of the following polynomials is irreducible
in Q[x] :
1 3 4 1 3 6
6x4 + 4x3 − 6x2 − 8x + 5; x4 − x2 + x − ; x3 + x2 − x + .
2 2 3 2 2 5
53
E. Irreducibility of Polynomials of Degree ≤ 4

1. Let F be any field. Explain why, if a(x) is a quadratic or cubic


polynomial in F[x], a(x) is irreducible in F[x] iff a(x) has no roots in F.
2. Prove that the following polynomials are irreducible in Q[x]:
1 3 3 3 1 1 5 3
x + 2x − ; 3x2 − 2x − 4; x3 + x2 + x + ; x3 + ; x2 − x + .
2 2 2 2 2 2 2
3. Suppose a monic polynomial a(x) of degree 4 in F[x] has no roots in
F. Then a(x) is reducible iff it is aproduct of two quadratics x2 + ax + b
and x2 + cx + d, that is, iff

a(x) = x4 + (a + c)x3 + (ac + b + d)x2 + (bc + ad)x + bd.

If the coefficients of a(x) cannot be so expressed (in terms of any a, b, c, d ∈


F ) then a(x) must be irreducible.
Example a(x) = x4 + 2x3 + x + 1; then bd = 1, so b = d = ±1; thus,
bc + ad = ±(a + c), but a + c = 2 and bc + ad = 1; which is impossible.
Prove that the following polynomials are irreducible in Q[x] (use Theo-
rem 3.2.3, searching only for integer values of a, b, c, d):

x4 − 5x2 + 1; 3x4 − x2 − 2; x4 + x3 + 3x + 1.

4. Prove that the following polynomials are irreducible in Z5 [x]:

2x3 + x2 + 4x + 1; x4 + 2; x4 + 4x2 + 2; x4 + 1.

F. Mapping onto Zn to Determine Irreducibility over Q

If h : Z → Zn is the natural homomorphism, let h : Z[x] → Zn [x] be


defined by

h(a0 + a1 x + . . . + an xn ) = h(a0 ) + h(al )x + . . . + h(an )xn .

In Chapter 1, Exercise G, it is proved that is a homomorphism. Assume


this fact and prove:

54
1. If h(a(x)) is irreducible in Zn [x] and a(x) is monic, then a(x) is
irreducible in Z[x];
2. x4 +10x3 +7 is irreducible in Z[x] by using the natural homomorphism
from Z to Z5 ;
3. The following are irreducible in Z[x] (find the right value of n and
use the natural homomorphism from Z to Zn ):

x4 − 10x2 + 1; x4 + 7x3 + 14x2 + 3; x5 + 1.

G. Roots and Factors in A[x] When A Is an Integral Domain

It is a useful fact that Theorems 3.1.1, 3.1.2, and 3.1.3 are still true in
A[x] when A is not a field, but merely an integral domain. The proof of
Theorem 3.1.1 must be altered a bit to avoid using the division algorithm.
We proceed as follows: If a(x) = a0 + a1 x + . . . + an xn and c is a root of
a(x), consider

a(x) − a(c) = a1 (x − c) + a2 (x2 − c2 ) + . . . + an (xn − cn ).

1. Prove that for k = 1, . . . , n :

ak (xk − ck ) = ak (x − c)(xk−1 + xk−2 c + . . . + ck−1 ).

2. Conclude from part 1 that a(x) − a(c) = (x − c)q(x) for some q(x).
3. Complete the proof of Theorem 3.1.1, explaining why this particular
proof is valid when A is an integral domain, not necessarily a field.
4 Check that Theorems 3.1.2 and 3.1.3 are true in A[x] when A is an
integral domain.

H. Polynomial Interpolation

One of the most important applications of polynomials is to problems


where we are given several values of x (say, x = a0 , a1 , . . . , an ) and corre-
sponding values of y (say, y = b0 , b1 , . . . , bn ), and we need to find a function

55
y = f (x) such that f (a0 ) = b0 , f (a1 ) = b1 , . . . , f (an ) = bn . The simplest
and most useful kind of function for this purpose is a polynomial function
of the lowest possible degree.
We now consider a commonly used technique for constructing a poly-
nomial p(x) of degree n which assumes given values b0 , b1 , . . . , bn are given
points a0 , a1 , . . . , an . That is,

p(a0 ) = b0 , p(a1 ) = b1 , . . . , p(an ) = bn .

First, for each i = 0, 1, . . . , n, let

qi (x) = (x − a0 ) . . . (x − ai−1 )(x − ai+1 ) . . . (x − an ).

1. Show that qi (aj ) = 0 for j ̸= i, and qi (ai ) ̸= 0.


Let qi (ai ) = ci , and define p(x) as follows:
n
X bi b0 b1 bn
p(x) = qi (x) = q0 (x) + q1 (x) + . . . + qn (x)
c
i=0 i
c0 c1 cn
(this is called the Lagrange interpolation formula).
2. Explain why p(a0 ) = b0 , p(a1 ) = b1 , . . . , p(an ) = bn .
3. Prove that there is one and only one polynomial p(x) of degree ≤ n
such that p(a0 ) = b0 , . . . , p(an ) = bn .
4. Use the Lagrange interpolation formula to prove that if F is a finite
field, every function from F to F is equal to a polynomial function (in fact,
the degree of this polynomial is less than the number of elements in F).
5. If t(x) is any polynomial in F[x], and a0 , . . . , an ∈ F , the unique
polynomial p(x) of degree ≤ n such that p(a0 ) = t(a0 ), . . . , p(an ) = t(an )
is called the Lagrange interpolator for t(x) and a0 , . . . , an . Prove that the
remainder, when t(x) is divided by (x − a0 )(x − a1 ) . . . (x − an ), is the
Lagrange interpolator.

I. Polynomial Functions over a Finite Field

1. Find three polynomials in Z5 [x] which determine the same function


as x2 − x + 1.

56
2. Prove that xp − x has p roots in Zp [x], for any prime p. Draw the
conclusion that in Zp [x], xp − x can be factored as

xp − x = x(x − 1)(x − 2) . . . [x − (p − 1)]3 .

3. Prove that if a(x) and b(x) determine the same function in p[x], then
(xp − x) | (a(x) − b(x)).
In the next four parts, let F be any finite field.
4. Let a(x) and b(x) be in F[x]. Prove that if a(x) and b(x) determine
the same function, and if the number of elements in F exceeds the degree
of a(x) as well as the degree of b(x), then a(x) = b(x).
5. Prove: The set of all a(x) which determine the zero function is an
ideal of F[x]. What its generator?
6. Let F(F ) be the ring of all functions from F to F, defined in the
same way as F(R). Let h : F [x] → F(F ) send every polynomial a(x) to the
polynomial function which it determines. Show that h is a homomorphism
from F[x] onto F(F ).
(NOTE: To show that h is onto, use Exercise H4).
7. Let F = {c1 , . . . , cn } and p(x) = (x − c1 ) . . . (x − cn ). Prove that

F[x]/⟨p(x)⟩ ∼
= F(F).

J. Solution of Cubic and Quartic Equations by Formulas

1. Label each of the following statements as either true or false.


(1) Every cubic equation over the reals has at least one real solution.
(2) Every quartic equation over the reals has at least one real solution.
(3) If the discriminant is positive for a quadratic or cubic polynomial
over the reals, then all the zeros must be real.
(4) If the discriminant is negative for a quadratic or cubic polynomial
over the reals, then all the zeros must be nonreal.
2. Use the techniques presented in this section to find all solutions of
the given equation.

57
(1) x3 − 15x − 30 = 0,
(2) x3 − 9x + 12 = 0,
(3) x3 − 12x − 20 = 0,
(4) x3 + 15x − 20 = 0,
(5) x3 − 6x − 6 = 0,
(6) x3 + 6x − 2 = 0,
(7) x3 + 9x + 6 = 0,
(8) x3 + 9x − 6 = 0,
(9) 2x3 + 6x − 3 = 0,
(10) 2x3 − 6x − 5 = 0,
(11) x3 − 6x2 + 33x − 92 = 0,
(12) x3 + 3x2 + 21x + 13 = 0,
(13) 8x3 + 12x2 + 150x + 25 = 0,
(14) 8x3 − 12x2 + 54x − 9 = 0,
(15) x4 + x2 − 2x + 6 = 0,
(16) x4 − 2x2 + 8x − 3 = 0,
(17) x4 + 4x3 + 3x2 + 4x + 2 = 0,
(18) x4 − 4x3 + 4x2 − 8x + 4 = 0.
3. Characterize the solutions to the following equations by evaluating
the discriminant D2 .
(1) x3 − 91x + 90 = 0,
(2) x3 − 32x + 24 = 0,
(3) x3 − 55x − 72 = 0,
(4) x3 − 124x − 240 = 0,
(5) x3 − 47x − 136 = 0,
(6) x3 − 3x + 52 = 0.
a
4. Prove Theorem 3.5.6: The change of variable x = y − in
4
x4 + ax3 + bx2 + cx + d = 0

58
yields an equation of the form

y 4 + py 2 + qy + r = 0.
1
5. Show that the change of variable x = y − an−1 in
n
xn + an−1 xn−1 + an−2 xn−2 + . . . + a1 x + a0 = 0

yields an equation of the form

y n + 0y n−1 + bn−2 y n−2 + . . . + b1 x + b0 = 0

or
y n + bn−2 y n−2 + . . . + b1 x + b0 = 0.

6. Derive the quadratic formula by using the change in variable


 
1 b
x=y−
2 a
to transform the quadratic equation
b c
x2 + x + = 0
a a
into one involving the difference of two squares and solve the resulting
equation.
7. Use the definition of the discriminant
Y
D2 = (ci − cj )2
i<j
 2
b c b c
to show that the discriminant of x2 + x + is −4 .
a a a a

59
CHAPTER 4

EXTENSIONS OF FIELDS

4.1 Algebraic and Transcendental Elements

If F is a field, then a subfield of F is any nonempty subset of F which


is closed with respect to addition and subtraction, multiplication and di-
vision (it would be equivalent to say: closed with respect to addition and
negatives, multiplication and multiplicative inverses). As we already know,
if K is a subfield of F, then K is a field in its own right.
If K is a subfield of F, we say also that F is an extension field of K.
When it is clear in context that both F and K are fields, we say simply
that F is an extension of K.
Given a field F, we may look inward from F at all the subfields of F.
On the other hand, we may look outward from F at all the extensions
of F. Just as there are relationships between F and its subfields, there
are also interesting relationships between F and its extensions. One of
these relationships, as we shall see later, is highly reminiscent of Lagrange’s
theorem—an inside-out version of it.
Why should we be interested in looking at the extensions of fields? There
are several reasons, but one is very special. If F is an arbitrary field, there
are, in general, polynomials over F which have no roots in F. For example,
x2 + 1 has no roots in R. This situation is unfortunate but, it turns out, not
hopeless. For, as we shall soon see, every polynomial over any field F has
roots. If these roots are not already in F, they are in a suitable extension
of F. For example, x2 + 1 = 0 has solutions in C.

60
In the matter of factoring polynomials and extracting their roots, is
utopia! In every polynomial a(x) of degree n has exactly n roots c1 , . . . , cn
and can therefore be factored as

a(x) = k(x − c1 )(x − c2 ) . . . (x − cn ).

This ideal situation is not enjoyed by all fields - far from it! In an arbitrary
field F, a polynomial of degree n may have any number of roots, from no
roots to n roots, and there may be irreducible polynomials of any degree
whatever. This is a messy situation, which does not hold the promise of an
elegant theory of solutions to polynomial equations. However, it turns out
that F always has a suitable extension E such that any polynomial a(x)
of degree n over F has exactly n solutions in E. Therefore, a(x) can be
factored in E[x] as

a(x) = k(x − c1 )(x − c2 ) . . . (x − cn ).

Thus, paradise is regained by the expedient of enlarging the field F. This is


one of the strongest reasons for our interest in field extensions. They will
give us a trim and elegant theory of solutions to polynomial equations.
Now, let us get to work! Let E be a field, F a subfield of E, and c any
element of E. We define the substitution function σc as follows: For every
polynomial a(x) in F[x],
σc (a(x)) = a(c).
Thus, σc is the function “substitute c for x”. It is a function from F[x] into
E. In fact, σc is a homomorphism. This is true because

σc (a(x) + b(x)) = σc (a(x) + σc (b(x),


| {z } | {z } | {z }
a(c)+b(c) a(c) b(c)

σc (a(x)b(x)) = σc (a(x) σc (b(x) .


| {z } | {z } | {z }
a(c)b(c) a(c) b(c)
The kernel of the homomorphism σc is the set of all the polynomials a(x)
such that a(c) = σc (a(x)) = 0. That is, the kernel of σc consists of all the
polynomials a(x) in F[x] such that c is a root of a(x).

61
Let Jc denote the kernel of σc ; since the kernel of any homomorphism is
an ideal, Jc is an ideal of F[x]. An element c in E is called algebraic over
F if it is the root of some nonzero polynomial a(x) in F[x]. Otherwise, c is
called transcendental over F. Obviously c is algebraic over F iff Jc contains
nonzero polynomials, and transcendental over F iff Jc = {0}.
We will confine our attention now to the case where c is algebraic. The
transcendental case will be examined in Exercise G at the end of this chap-
ter.

4.2 The Minimum Polynomial

Let c be algebraic over F, and let Jc be the kernel of σc (where σc is


the function “substitute c for x”). Remember that in F[x] every ideal is
a principal ideal; hence Jc = ⟨p(x)⟩ = the set of all multiples of p(x), for
some polynomial p(x). Since every polynomial in Jc is a multiple of p(x),
p(x) is a polynomial of lowest degree among all the nonzero polynomials
in Jc . It is easy to see that p(x) is irreducible; otherwise we could factor it
into polynomials of lower degree, say p(x) = f (x)g(x). But then 0 = p(c) =
f (c)g(c), so f (c) = 0 or g(c) = 0, and therefore either f (x) or g(x) is in
Jc . This is impossible, because we have just seen that p(x) has the lowest
degree among all the polynomials in Jc , whereas f (x) and g(x) both have
lower degree than p(x).
Since every constant multiple of p(x) is in Jc , we may take p(x) to be
monic, that is, to have leading coefficient 1. Then p(x) is the unique monic
polynomial of lowest degree in Jc (also, it is the only monic irreducible
polynomial in Jc ). This polynomial p(x) is called the minimum polynomial
of c over F, and will be of considerable importance in our discussions in a
later course.
Let us look at an example: R is an extension field of Q, and R contains
√ √
the irrational number 2. The function σ√2 is the function “substitute 2
for x”; for example σ√2 (x4 − 3x2 + 1) = −3 + 1 = −1. By our discussion

62
above, σ√2 : Q[x] → R is a homomorphism and its kernel consists of all

the polynomials in Q[x] which have 2 as one of their roots. The monic

polynomial of least degree in Q[x] having 2 as a root is p(x) = x2 − 2;

hence x2 − 2 is the minimum polynomial of 2 over Q.
Now, let us turn our attention to the range of σc . Since σc is a homomor-
phism, its range is obviously closed with respect to addition, multiplication,
and negatives, but it is not obviously closed with respect to multiplicative
inverses. Not obviously, but in fact it is closed for multiplicative inverses,
which is far from self-evident, and quite a remarkable fact. In order to prove
this, let f (c) be any nonzero element in the range of σc . Since f (c) ̸= 0, f (x)
is not in the kernel of σc . Thus, f (x) is not a multiple of p(x), and since p(x)
is irreducible, it follows that f (x) and p(x) are relatively prime. Therefore
there are polynomials s(x) and t(x) such that s(x)f (x) + t(x)p(x) = 1. But
then
s(c)f (c) + t(c)p(c) = 1.

Since p(c) = 0, s(c) is the multiplicative inverse of f (c).


We have just shown that the range of σc is a subfield of E. Now, the
range of σc is the set of all the elements a(c), for all a(x) in F[x]:

Range σc = {a(c) | a(x) ∈ F [x]}.

We have just seen that range σc is a field. In fact, it is the smallest field con-
taining F and c: indeed, any other field containing F and c would inevitably
contain every element of the form

a0 + a1 c + . . . + an cn (a0 , . . . , an ∈ F )

in other words, would contain every element in the range of σc .


By the smallest field containing F and c we mean the field which contains
F and c and is contained in any other field containing F and c. It is called
the field generated by F and c, and is denoted by the important symbol
F(c).

63
Now, here is what we have, in a nutshell: σc is a homomorphism with
domain F[x], range F(c), and kernel Jc = ⟨p(x)⟩. Thus, by the fundamental
homomorphism theorem

F(c) ∼
= F [x]/⟨p(x)⟩.

Finally, here is an interesting sidelight: if c and d are both roots of p(x),


where c and d are in E, then, by what we have just proved, F(c) and F(d)
are both isomorphic to F[x]/⟨p(x)⟩, and therefore isomorphic to each other:
If c and d are roots of the same irreducible polynomial p(x) in F[x], then
F(c) ∼
= F (d). In particular, this shows that, given F and c, F(c) is unique
up to isomorphism.

4.3 Basic Theorem on Field Extensions

It is time now to recall our main objective: if a(x) is a polynomial in F[x]


which has no roots in F, we wish to enlarge F to a field E which contains
a root of a(x). How can we manage this?
An observation is in order: finding extensions of F is not as easy as
finding subfields of F. A subfield of F is a subset of an existing set: it is
therel But an extension of F is not yet there. We must somehow build it
around F.
Let p(x) be an irreducible polynomial in F[x]. We have just seen that if F
can be enlarged to a field E containing a root c of p(x), then F(c) is already
what we are looking for: it is an extension of F containinga root of p(x).
Furthermore, F(c) is isomorphic to F[x]/⟨p(x)⟩. Thus, the field extension
we are searching for is precisely F[x]/⟨p(x)⟩. Our result is summarized in
the next theorem.

4.3.1 Theorem (Basic theorem of field extensions). Let F be a field and


a(x) a nonconstant polynomial in F[x]. There exists an extension field E of
F and an element c in E such that c is a root of a(x).

64
Proof. To begin with, a(x) can be factored into irreducible polynomials
in F[x]. If p(x) is any nonconstant irreducible factor of a(x), it is clearly
sufficient to find an extension of F containing a root of p(x), since such a
root will also be a root of a(x).
In Exercise D4 of Chapter 2, the reader was asked to supply the simple
proof that, if p(x) is irreducible in F[x], then ⟨p(x)⟩ is a maximal ideal of
F[x]. Furthermore, if ⟨p(x)⟩ is a maximal ideal of F[x], then the quotient
ring F[x]/⟨p(x)⟩ is a field.
It remains only to prove that F[x]/⟨p(x)⟩ is the desired field extension
of F. When we write J = ⟨p(x)⟩, let us remember that every element of
F[x]/J is a coset of J. We will prove that F[x]/J is an extension of F by
identifying each element a in F with its coset a + J.
To be precise, define h : F → F [x]/J by h(a) = a + J. Note that h is
the function which matches every a in F with its coset a + J in F[x]/J. We
will now show that h is an isomorphism.
By the familiar rules of coset addition and multiplication, h is a homo-
morphism. Now, every homomorphism between fields is injective (this is
true because the kernel of a homomorphism is an ideal, and a field has no
nontrivial ideals). Thus, h is an isomorphism between its domain and its
range.
What is the range of h? It consists of all the cosets a + J where a ∈ F ,
that is, all the cosets of constant polynomials (if a is in F, then a is a
constant polynomial). Thus, F is isomorphic to the subfield of F[x]/J
containing all the cosets of constant polynomials. This subfield is therefore
an isomorphic copy of F, which may be identified with F, so F[x]/J is an
extension of F.
Finally, if p(x) = a0 + a1 x + . . . + an xn , let us show that the coset x + J
is a root of p(x) in F[x]/J. Of course, in F[x]/J, the coefficients are not
actually a0 , a1 , . . . , an , but their cosets a0 + J, a1 + J, . . . , an + J. Writing

a0 = a0 + J, . . . , an = an + J, and x = x + J.

65
We must prove that

a0 + a1 x + . . . + an xn = J (J is the zero coset).

Well
a0 + a1 x + . . . + an xn = (a0 + J) + (a1 + J)(x + J) + . . . + (an + J)(x + J)n
= (a0 + J) + (a1 x + J) + . . . + (an xn + J)
= p(x) + J
= J ( because p(x) ∈ J).
This completes the proof of the basic theorem of field extensions. Observe
that we may use this theorem several times in succession to get the follow-
ing: Let a(x) be a polynomial of degree n in F[x]. There is an extension
field E of F which contains all n roots of a(x).

4.4 Degrees of field extensions

Let F and K be fields. If K is an extension of F , we may regard K as


being a vector space over F . We may treat the elements in K as “vectors”
and the elements in F as “scalars.” That is, when we add elements in K,
we think of it as vector addition; when we add and multiply elements in F ,
we think of this as addition and multiplication of scalars; and finally, when
we multiply an element of F by an element of K, we think of it as scalar
multiplication.
We will be especially interested in the case where the resulting vector
space is of finite dimension. If K, as a vector space over F , is of finite
dimension, we call K a finite extension of F . If the dimension of the vector
space K is n, we say that K is an extension of degree n over F . This is
symbolized by writing [K : F ] = n which should be read, “the degree of K
over F is equal to n.”
Let us recall that F (c) denotes the smallest field which contains F and
c. This means that F (c) contains F and c, and that any other field K

66
containing F and c must contain F (c). We saw in the above that if c is
algebraic over F , then F (c) consists of all the elements of the form a(c),
for all a(x) in F [x]. Since F (c) is an extension of F , we may regard it as a
vector space over F . Is F (c) a finite extension of F ?
Well, let c be algebraic over F , and let p(x) be the minimum polynomial
of c over F (that is, p(x) is the monic polynomial of lowest degree having
c root). Let the degree of the polynomial p(x) be equal to n. It turns out,
then, that the n elements

1, c, c2 , . . . , cn−1

are linearly independent and span F (c). We will prove this fact in a mo-
ment, but meanwhile let us record what it means. It means that the set
of n “vectors” {1, c, c2 , . . . , cn−1 } is a basis of F (c); hence F (c) is a vector
space of dimension n over the field F . This may be summed up concisely
as follows:

4.4.1 Theorem. The degree of F (c) over F is equal to the degree of the
minimum polynomial of c over F .

Proof. It remains only to show that the n elements 1, c, . . . , cn−1 span F (c)
and are linearly independent. Well, if a(c) is any element of F (c), use the
division algorithm to divide a(x) by p(x):

a(x) = p(x)q(x) + r(x) where deg r(x) ≤ n − 1.

Therefore,
a(c) = p(c)q(c) + r(c) = 0 + r(c) = r(c).

This shows that every element of F (c) is of the form r(c) where r(x) has
degree n − 1 or less. Thus, every element of F (c) can be written in the form

a0 + a1 c + . . . + an−1 cn−1

which is a linear combination of 1, c, . . . , cn−1 .

67
Finally, to prove that 1, c, . . . , cn−1 are linearly independent, suppose
that a0 + a1 c + . . . + an−1 cn−1 = 0. If the coefficients a0 , a1 , . . . , an−1 were
not all zero, c would be the root of a nonzero polynomial of degree n − 1
or less, which is impossible because the minimum polynomial of c over F
has degree n. Thus, a0 = a1 = . . . = an−1 = 0.
√ √
4.4.2 Example. Let us look at Q( 2): the number 2 is not a root of any
monic polynomial of degree 1 over Q. For such a polynomial would have to
√ √
be x − 2, and the latter is not in Q[x] because 2 is irrational. However,
√ √
2 is a root of x2 − 2, which is therefore the minimum polynomial of 2
over Q, and which has degree 2. Thus,

[Q( 2) : Q] = 2.

In particular, every element in Q( 2) is therefore a linear combination of
√ √
1 and 2, that is, a number of the form a + b 2 where a, b ∈ Q.

4.4.3 Example. i is a root of the irreducible polynomial of over x2 + 1 in


R[x]. Therefore x2 + 1 is the minimum polynomial of i over R, x2 + 1 has
degree 2, so [R(i) : R] = 2. Thus, R(i) consists of all the linear combinations
of 1 and i with real coefficients, that is, all the a+bi where a, b ∈ R. Clearly
then, R(i) = C, so the degree of C over R is equal to 2.

In the sequel we will often encounter the following situation: E is a finite


extension of K, where K is a finite extension of F . If we know the degree
of E over K and the degree of K over F , can we determine the degree of
E over F . This is a question of major importance! Fortunately, it has an
easy answer, based on the following lemma:

4.4.4 Lemma. Let a1 , a2 , . . . , am be a basis of the vector space K over F ,


and let b1 , b2 , . . . , bn be a basis of the vector space E over K. Then the set
of mn products {ai bj } is a basis of the vector space E over the field F .

Proof. To prove that the set {ai bj } spans E, note that each element c in E

68
can be written as a linear combination

c = k1 b1 + . . . + kn bn

with coefficients ki in K. But each ki because it is in K, is a linear combi-


nation
ki = li1 a1 + . . . + lim am
with coefficients lij in F . Substituting,
c = (l11 a1 + . . . + l1n am )b1 + . . . + (ln1 a1 + . . . + lnm am )bn
X
= lij ai bj

and this is a linear combination of the products ai bj with coefficient lij in


F.
P
To prove that {ai bj } is linearly independent, suppose lij ai bj = 0. This
can be written as

(l11 a1 + . . . + l1n am )b1 + . . . + (ln1 a1 + . . . + lnm am )bn = 0

and since b1 , . . . , bn are independent, li1 a1 + . . . + lim am = 0 for each i. But


a1 , . . . , am are also independent, so every lij = 0.

With this result we can now conclude the following:

4.4.5 Theorem. Suppose F ⊆ K ⊆ E where E is a finite extension of K


and K is a finite extension of F .Then E is a finite extension of F , and

[E : F ] = [E : K][K : F ].

This theorem is a powerful tool in our study of fields. It plays a role in


field theory analogous to the role of Lagrange’s theorem in group theory.
See what it says about any two extensions, K and E of a fixed “base field”
F : If K is a subfield of E9 then the degree of K (over F ) divides the degree
of E (over F ).
If c is algebraic over F , we say that F (c) is obtained by adjoining c to F .
If c and d are algebraicover F , we may find adjoin c to F , thereby obtaining

69
F (c), and then adjoin d to F (c). The resulting field is denoted F (c, d) and
is the smallest field containing F , c and d. [Indeed, any field containing F ,
c and d must contain F (c), hence also F (c, d)]. It does not matter whether
we first adjoin c and then d or vice versa.
If c1 , . . . , cn are algebraic over F , we let F (c1 , . . . , cn ) be the smallest
field containing F and c1 , . . . , cn . We call it the field obtained by adjoining
c1 , . . . , cn to F . We may form F (c1 , . . . , cn ) step by step, adjoining one ci
at a time, and the order of adjoining the ci is irrelevant.
An extension F (c) formed by adjoining a single element to F is called
a simple extension of F . An extension F (c1 , . . . , cn ) formed by adjoining
a finite number of elements c1 , . . . , cn is called an iterated extension. It is
called “iterated” because it can be formed step by step, one simple extension
at a time:

F ⊆ F (c1 ) ⊆ F (c1 , c2 ) ⊆ F (c1 , c2 , c3 ) ⊆ . . . ⊆ F (c1 , . . . , cn−1 ) (4.4.5.1)

If c1 , . . . , cn are algebraic over F , then by Theorem 4.4.1, each extension


in Condition 4.4.5.1 is a finite extension. By Theorem 4.4.5, F (c1 c2 ) is a
finite extension of F ; applying Theorem 4.4.5 again, F (c1 , c2 , c3 ) is a finite
extension of F ; and so on. So finally, if c1 , . . . , cn are algebraic over F , then
F (c1 , . . . , cn ) is a finite extension of F .
Actually, the converse is true too: every finite extension is an iterated
extension. This is obvious: for if K is a finite extension of F , say an exten-
sion of degree n, then K has a basis {a1 , . . . , an } over F . This means that
every element in K is a linear combination of a1 , . . . , an with coefficients
in F ; but any field containing F and a1 , . . . , an obviously contains all the
linear combinations of a1 , . . . , an ; hence K is the smallest field containing
F anda1 , . . . , an . That is, K = F (a1 , . . . , an ).
In fact, if K is a finite extension of F and K = F (a1 , . . . , an ), then
a1 , . . . , an have to be algebraic over F . This is a consequence of a simple
but important little theorem:

70
4.4.6 Theorem. If K is a finite extension of F , every element of K is
algebraic over F .

Proof. Indeed, suppose K is of degree n; over F , and let c be any element


of K. Then the set {c, c2 , . . . , cn } is linearly dependent, because it has n + 1
elements in a vector space K of dimension n.
Consequently, there are scalars a0 , . . . , an ∈ F , not all zero, such that
a0 + a1 c + . . . + an cn = 0. Therefore c is a root of the polynomial a(x) =
a0 + a1 x + . . . + an xn in F [x].

Let us sum up: Every iterated extension F (c1 , . . . , cn ), where c1 , . . . , cn


are algebraic over F , is finite extension of extension of F . Conversely,
every finite extension of F is an iterated extension F (c1 , . . . , cn ), where
c1 , . . . , cn are algebraic over F .
Here is an example of the concepts presented in this section. We have
√ √
already seen that Q( 2) is of degree 2 over Q, and therefore Q( 2) consists
√ √
of all the numbers a + b 2 where a, b ∈ Q. Observe that 3 cannot be
√ √ √
in Q( 2); for if it were, we would have 3 = a + b 2 for rational a and
√ √
b; squaring both sides and solving for 2 would give us 2 = a′ rational
number, which is impossible.
√ √ √
Since 3 is not in Q( 2), 3, cannot be a root of a polynomial of
√ √
degree 1 over Q( 2) (such a polynomial would have to be x − 3). But
√ √
3 is a root of x2 − 3, which is therefore the minimum polynomial of 3
√ √ √ √
over Q( 2).Thus, Q( 2, 3) is of degree 2 over Q( 2), and therefore by
√ √
Theorem 4.4.5, Q( 2, 3) is of degree 4 over Q.
√ √
By the comments preceding Theorem 4.4.1, {1, 2} is a basis of Q( 2)
√ √ √ √
over Q, and {1, 3} is a basis of Q( 2, 3) over Q( 2). Thus, by Lemma
√ √ √ √ √
4.4.4, {1, 2, 3, 6} is a basis of Q( 2, 3) over Q. This means that
√ √ √ √ √
Q( 2, 3) consists of all the numbers a + b 2 + c 3 + d 6 , for all a, b, c,
and d in Q.
For later reference. The technical observation which follows will be

71
needed later.
By the comments immediately preceding Theorem 4.4.1, every element
of F (c1 ) is a linear combination of powers of c1 , with coefficients in F . That
is, every element of F (c1 ) is of the form ki ci1 where the ki are in F . For
P
i
the same reason, every element of F (c1 , c2 ) is of the form lj ci1 where the
P
j
coefficients lj are in F (c1 ). Thus, each coefficient lj is equal to a sum of
the form (2). But then, clearing brackets, it follows that every element of
F (c1 , c2 ) is of the form kij ci1 cj2 where the coefficients kij are in F .
P
i
If we continue this process, it is easy to see that every element of
F (c1 , c2 , . . . , cn ) is a sum of terms of the form kci11 ci22 . . . cinn where the coef-
ficient k of each term is in F .

72
EXERCISES

A. Recognizing Algebraic Elements

1. Prove that each of the following numbers is algebraic over Q:


(a) i,

(b) 2,
(c) 2 + 3i,
p √
(d) 1 + 3 2,
p √
(e) i − 2,
√ √
(f) 2 + 3,
√ √
(g) 3 2 + 3 4.
2. Prove that each of the following numbers is algebraic over the given
field:

(a) π over Q(π),

(b) π over Q(π 2 ),
√ 2
(c) π − 1 over Q(π 3 ).
p √
Example: To show that 1 + 2 is algebraic over Q, one must find a
p √
polynomial p(x) ∈ Q[x] such that 1 + 2 is a root of p(x).
p √ √ √
Let a = 1 + 2; then a2 = 1 + 1 + 2, a2 − 1 = 2, and finally,
(a2 − l)2 = 2. Thus, a satisfies p(x) = x4 − 2x2 − 1 = 0.
NOTE: Recognizing a transcendental element is much more difficult, since
it requires proving that the element cannot be a root of any polynomial
over the given field. In recent times it has been proved, using sophisticated
mathematical machinery, that π and e are transcendental over Q.

B. Finding the Minimum Polynomial

1. Find the minimum polynomial of each of the following numbers over


Q (where appropriate, use the methods of Chapter 3, Exercises D, E, and
F to ensure that your polynomial is irreducible).

73
(a) 1 + 2i,

(b) 1 + 2,

(c) 1 + 2i,
p √
(d) 2 + 3 3,
√ √
(e) 3 + 5,
p √
(f) 1 + 2.

2. Show that the minimum polynomial of 2 + i is

(a) x2 − 2 2x + 3 over R,
(b) x4 − 2x2 + 9 over Q,
(c) x2 − 2ix − 3 over Q(i).
3. Find the minimum polynomial of the following numbers over the
indicated fields:
√ √
(a) 3 + i over R; over Q; over Q(i); over Q( 3),
p√ √
(b) 2 + i over R; over Q; over Q(i); over Q( 2).
4. For each of the following polynomials p(x), find a number a such that
p(x) is the minimum polynomial of a over Q:
(a) x2 + 2x − 1,
(b) x4 + 2x2 − 1,
(c) x4 − 10x2 + 1.
5 Find a monic irreducible polynomial p(x) such that Q[x]/⟨p(x)⟩ is
isomorphic to

(a) Q( 2),

(b) Q(1 + 2),
p √
(c) Q( 1 + 2).

C. The Structure of Fields F[x]/⟨p(x)⟩

Let p(x) be an irreducible polynomial of degree n over F. Let c denote


a root of p(x) in some extension of F (as in the basic theorem on field
extensions).

74
1. Prove: Every element in F(c) can be written as r(c), for some r(x) of
degree < n in F[x].
[HINT: Given any element t(c) ∈ F (c), use the division algorithm to
divide t(x) by p(x)].
2. If s(c) = t(c) in F(c), where s(x) and t(x) have degree < n, prove
that s(x) = t(x).
3. Conclude from parts 1 and 2 that every element in F(c) can be written
uniquely as r(c), with deg r(x) < n.
4. Using part 3, explain why there are exactly four elements in Z2 [x]/⟨x2 +
x + 1⟩. List these four elements, and give their addition and multiplication
tables.
[HINT: Identify Z2 [x]/<x2 + x + 1⟩ with Z2 (c), where c is a root of
x2 + x + 1. Write the elements of Z2 (c) as in part 3. When computing the
multiplication table, use the fact that c2 + c + 1 = 0].
5. Describe Z2 [x]/⟨x3 + x + 1⟩, as in part 4.
6. Describe Z3 [x]/⟨x3 + x2 + 2⟩, as in part 4.

D. Short Questions Relating of Field Extensions

1. Let F be any field. Prove that:


(a) If c is algebraic over F, so are c + 1 and kc (where k ∈ F ),
(b) If c ̸= 0 and c is algebraic over F, so is 1/c,
(c) If cd is algebraic over F, then c is algebraic over F(d); If c + d is
algebraic over F, then c is algebraic over F(d) (assume c ̸= 0 and d ̸= 0),
(d) If the minimum polynomial of a over F is of degree 1, then a ∈ F ,
and conversely.
(e) Suppose F ⊆ K and a ∈ K. If p(x) is a monic irreducible polynomial
in F[x], and p(a) = 0, then p(x) is the minimum polynomial of a over F.
2. Name a field (̸= R or C) which contains a root of x5 + 2x3 + 4x2 + 6.
√ √
= Q(1 − i). However, Q( 2) ∼
3. Prove: Q(1 + i) ∼ = Q( 3).
4. If p(x) is irreducible and has degree 2, prove that F[x]/⟨p(x)⟩ contains

75
both roots of p(x).

E. Simple Extensions

1. Recall the definition of F(a). It is a field such that


(i) F ⊆ F (a);
(ii) a ∈ F (a);
(iii) any field containing F and a contains F(a).
Use this definition to prove the following statements, where F ⊆ K,
c ∈ F , and a ∈ K:
(a) F(a) = F (a + c) and F(a) = F (ca) (assume c ̸= 0);
(b) F(a2 ) ⊆ F (a) and F(a + b) ⊆ F (a, b). [F(a, b) is the field containing
F, a, and b, and contained in any other field containing F, a and b]. Why
are the reverse inclusions not necessarily true?
(c) a + c is a root of p(x) iff a is a root of p(x + c); ca is a root of p(x)
iff a is a root of p(cx);
(d) Let p(x) be irreducible, and let a be a root of p(x + c). Then
F[x]/⟨p(x + c)⟩ ∼
= F (a) and F[x]/⟨p(x)⟩ ∼
= F (a + c). Conclude that
F[x]/⟨p(x + c)⟩ ∼
= F [x]/⟨p(x)⟩;
(e) Let p(x) be irreducible, and let a be a root of p(cx). Then F[x]/⟨p(cx)⟩ ∼
=
F (a) and F[x]/⟨p(cx)⟩ ∼= F (ca). Conclude that
F[x]/⟨p(x)⟩ ∼
= F [x]/⟨p(x)⟩.
2. Use parts (d) and (e) to prove the following:
(a) Z11 [x]/⟨x2 + 1⟩ ∼
= Z11 [x]/⟨x2 + x + 4⟩;
(b) If a is a root of x2 −2 and b is a root of x2 −4x+2, then Q(a) ∼
= Q(b);
1
(c) If a is a root of x2 − 2 and b is a root of x2 − , then Q(a) ∼= Q(b).
2
F. Quadratic Extensions

If the minimum polynomial of a over F has degree 2, we call F(a) a


quadratic extension of F.

76
1. Prove that, if F is a field whose characteristic is ̸= 2, any quadratic

extension of F is of the form F( a), for some a ∈ F (HINT: Complete the
square, and use Exercise E1(d)).
Let F be a finite field, and F∗ the multiplicative group of nonzero ele-
ments of F. Obviously H = {x2 | x ∈ F ∗ } is a subgroup of F∗ ; since every
square x2 in F∗ is the square of only two different elements, namely ±x,
exactly half the elements of F∗ are in H. Thus, H has exactly two cosets:
H itself, containing all the squares, and aH (where a ∈ / H), containing all
a
the nonsquares. If a and b are nonsquares, then ab−1 = ∈ H. Thus: if a
a b
and b are nonsquares, is a square. Use these remarks in the following:
b
2. Let F be a finite field. If a, b ∈ F , let p(x) = x2 − a and q(x) = x2 − b
√ √
be irreducible in F[x], and let a and b denote roots of p(x) and q(x) in
a a
an extension of F. Explain why is a square, say = c2 for some c ∈ F.
b b
Prove that is a root of p(cx).

3. Use part 2 to prove that F[x]/⟨p(cx)⟩ ∼ = F ( b); then use Exercise
√ √
E1(e) to conclude that F( a) ∼ = F ( b).
4. Use part 3 to prove: Any two quadratic extensions of a finite field are
isomorphic.
a
is a square (why?). Use the same
5. If a and b are nonsquares in R,
b
argument as in part 4 to prove that any two simple extensions of R are
isomorphic (hence isomorphic to C).

G. Questions Relating to Transcendental Elements

Let F be a field, and let c be transcendental over F. Prove the following:


1. {a(c) | a(x) ∈ F [x]} is an integral domain isomorphic to F[x].
2. F(c) is the field of quotients of {a(c) | a(x) ∈ F [x]}, and is isomorphic
to F(x), the field of quotients of F[x].
3. If c is transcendental over F, so are c + 1, kc (where k ∈ F and k ̸= 0),
c2 .

77
4. If c is transcendental over F, every element in F(c) but not in F is
transcendental over F.

H. Common Factors of Two Polynomials: Over F and over Ex-


tensions of F

Let F be a field, and let a(x), b(x) ∈ F [x]. Prove the following:
1. If a(x) and b(x) have a common root c in some extension of F,
they have a common factor of positive degree in F[x]. [Use the fact that
a(x), b(x) ∈ ker σc ].
2. If a(x) and b(x) are relatively prime in F[x], they are relatively prime
in K[x], for any extension K of F. Conversely, if they are relatively prime
in K[x], then they are relatively prime in F[x].

I. Derivatives and Their Properties

Let a(x) = a0 + a1 x + . . . + an xn ∈ F [x]. The derivative of a(x) is the


following polynomial a′ (x) ∈ F [x]:

a′ (x) = a1 + 2a2 x + . . . + nan xn−1

(this is the same as the derivative of a polynomial in calculus). We now


prove the analogs of the formal rules of differentiation, familiar from cal-
culus. 1. Let a(x), b(x) ∈ F [x], and let k ∈ F . Prove the following
statements:
(a) [a(x) + b(x)]′ = a′ (x) + b′ (x),
(b) [a(x)b(x)]′ = a′ (x)b(x) + a(x)b′ (x),
(c) [ka(x)]′ = ka′ (x),
(e) If F has characteristic 0 and a′ (x) = 0, then a(x) is a constant poly-
nomial. Why is this conclusion not necessarily true if F has characteristic
p ̸= 0?
2. Find the derivative of the following polynomials in Z5 [x]:

x6 + 2x3 + x + 1, x5 + 3x2 + 1, x15 + 3x10 + 4x5 + 1.

78
3. If F has characteristic p ̸= 0, and a′ (x) = 0, prove that the only
nonzero terms of a(x) are of the form amp xmp for some m [that is, a(x) is
a polynomial in powers of xp ].

J. Multiple Roots

Suppose a(x) ∈ F [x], and K is an extension of F. An element c ∈ K is


called a multiple root of a(x) if (x − c)m | a(x) for some m > 1. It is often
important to know if all the roots of a polynomial are different, or not. We
now consider a method for determining whether an arbitrary polynomial
a(x) ∈ F [x] has multiple roots in any extension of F.
Let K be any field containing all the roots of a(x). Suppose a(x) has a
multiple root c.
1. Prove that a(x) = (x − c)2 q(x) ∈ K[x].
2. Compute a′ (x), using part 1.
3. Show that x − c is a common factor of a(x) and a′ (x). Use Exercise
H 1 to conclude that a(x) and a′ (x) have a common factor of degree > 1
in F[x].
Thus, if a(x) has a multiple root, then a(x) and a′ (x) have a common
factor in F[x]. To prove the converse, suppose a(x) has no multiple roots.
Then a(x) can be factored as

a(x) = (x − c1 ) . . . (x − cn )

where c1 , . . . , cn are all different.


4. Explain why a′ (x) is a sum of terms of the form

(x − c1 ) . . . (x − ci−1 )(x − ci+1 ) . . . (x − cn ).

5. Using part 4, explain why none of the roots c1 , . . . , cn of a(x) are


roots of a′ (x).
6. Conclude that a(x) and a′ (x) have no common factor of degree > 1
in F[x].

79
This important result is stated as follows: A polynomial a(x) in F[x] has
a multiple root iff a(x) and a′ (x) have a common factor of degree > 1 in
F[x].
7. Show that each of the following polynomials has no multiple roots in
any extension of its field of coefficients:

x3 − 7x2 + 8 ∈ Q[x], x2 + x + 1 ∈ Z5 [x], x100 − 1 ∈ Z7 [x].

The preceding example is most interesting: it shows that there are 100
different hundredth roots of 1 over Z7 . (The roots ±1 are in Z7 , while the
remaining 98 roots are in extensions of Z7 ). Corresponding results hold for
most other fields.

EXERCISES OF FIELD EXTENSIONS

A. Examples of Finite Extensions


√ √
1. Find a basis for Q(i 2) over Q, and describe the elements of Q(i 2)
(see the two examples immediately following Theorem 4.4.1).
2. Show that every element of R(2 + 3i) can be written as a + bi, where
a, b ∈ R. Conclude that R(2 + 3i) = C.
p √
3. If a = 1 + 3 2, show that {1, 21/3 , 22/3 , a, 21/3 a, 22/3 a} is a basis of
Q(a) over Q. Describe the elements of Q(a).
√ √
4. Find a basis of Q( 2 + 3 4) over Q, and describe the elements of
√ √
Q( 2 + 3 4).
√ √
5. Find a basis of Q( 5, 7) over Q and describe the elements of
√ √
Q( 5, 7) (see the example at the end of this chapter).
√ √ √
6. Find a basis of Q( 2, 2, 5) over Q, and describe the elements of
√ √ √
Q( 2, 2, 5).
7. Name an extension of Q over which π is algebraic of degree 3.

B. Further Examples of Finite Extensions

80
Let F be a field of characteristic ̸= 2. Let a ̸= b be in F .
√ √ √ √
1. Prove that any field F containing a + b also contains a and b.
√ √ √ √
Conclude that F ( a + b) = F ( a, b).
√ √ √
[HINT: Compute ( a + b)2 and show that ab ∈ F . Then compute
√ √ √
ab( a + b), which is also in F ].
√ √
2. Prove that if b ̸= x2 a for any x ∈ F , then b ∈ / F ( a). Conclude
√ √
that F ( a, b) is of degree 4 over F .
√ √
3. Show that x = a + b satisfies x4 − 2(a + b)x2 + (a − b)2 = 0. Show
p √
that x = a + b + 2 ab also satisfies this equation. Conclude that
q √ √ √
F ( a + b + 2 ab) = F ( a, b).

4. Using parts 1 to 3, find an uncomplicated basis for Q(d) over Q,


4 2
p √
where d is a root of x − 14x + 9. Then find a basis for Q( 7 + 2 10)
over Q.

C. Finite Extensions of Finite Fields

By the proof of the basic theorem of field extensions, if p(x) is an irre-


ducible polynomial of degree n in F [x], then F [x]/⟨p(x)⟩ ∼
= F (c) where c is
a root of p(x). By Theorem 4.4.1 in this chapter, F (c) is of degree n over
F . Using the paragraph preceding Theorem 4.4.1:
1. Prove that every element of F (c) can be written uniquely as a0 +
a1 c + . . . + an−1 cn−1 , for some a0 , . . . , an−1 ∈ F .
2. Construct a field of four elements (it is to be an extension of Z2 ).
Describe its elements, and supply its addition and multiplication tables.
3. Construct a field of eight elements (it is to be an extension of Z2 ).
4. Prove that if F has q elements, and a is algebraic over F of degree n,
then F (a) has q n elements.
5. Prove that for every prime number p, there is an irreducible quadratic
in Zp [x]. Conclude that for every prime p, there is a field with p2 elements.

81
D. Degrees of Extensions (Applications of Theorem 4.4.5)

Let F be a field, and K a field extension of F . Prove the following:


1. [K : F ] = 1 iff K = F .
2. If [K : F ] is a prime number, there is no field properly between F
and K (that is, there is no field L such that F ⊊ L ⊊ K).
3. If [K : F ] is a prime, then K = F (a) for every a ∈ K \ F .
4. Suppose a, b ∈ K are algebraic over F with degrees m and n, where
m and n are relatively prime. Then:
(a) F (a, b) is degree mn over F .
(b) F (a) ∩ F (b) = F.
5. If the degree of F (a) over F is a prime, then F (a) = F (an ) for any n
(on the condition that an ∈
/ F ).
6. If an irreducible polynomial p(x) ∈ F [x] has a root in K, then
deg p(x) | [K : F ].

E. Short Questions Relating to Degrees of Extensions

1. Let F be a field. Prove that


(a). The degree of a over F is the same as the degree of 1/a over F . It
is also the same as the degrees of a + c and ac over F , for any c ∈ F .
(b) a is of degree 1 over F iff a ∈ F .
(c) If a real number c is a root of an irreducible polynomial of degree
> 1 in Q[x], then c is irrational.
2. Use part 1 (c) and Eisentein’s irreducibility criterion to prove that
p
m/n (where mn ∈ Q) is irrational if there is a prime number which
divides m but not n,and whose square does not divide m.
p
3. Show that part 2 remains true for q m/n where q > 1.
4. If a and b are algebraic over F , prove that F (a, b) is a finite extension
of F .

82
F. Further Properties of Degrees of Extensions

Let F be a field, and K a finite extension of F . Prove each of the


following:
1. Any element algebraic over K is algebraic over F , and conversely.
2. If b is algebraic over K, then [F (b) : F ] | [K(b) : F ].
3. If b is algebraic over K, then [K(b) : K] ≤ [F (b) : F ].
(HINT: The minimum polynomial of b over F may factor in K[x], and
b will then be a root of one of its irreducible factors).
4. If b is algebraic over K, then [K(b) : F (b)] ≤ [K : F ].
[HINT: Note that F ⊆ K ⊆ K(b) and F ⊆ F (b) ⊆ K(b). Relate the
degrees of the four extensions involved here, using part 3].
5. Let p(x) be irreducible in F [x]. If [K : F ] and deg p(x) are relatively
prime, then p(x) is irreducible in K[x].

G. Fields of Algebraic Elements: Algebraic Numbers

Let F ⊆ K and a, b ∈ K. We have seen that if a and b are algebraic


over F , then F (a, b) is a finite extension of F .
Use the above to prove parts 1 and 2.
1. If a and b are algebraic over F , then a + b, a − b, ab, and a/b are
algebraic over F (in the last case, assume b ̸= 0).
2. The set {x ∈ K | x is algebraic over F } is a subfield of K, containing
F.
Any complex number which is algebraic over Q is called an algebraic
number. By part 2, the set of all the algebraic numbers is a field, which we
shall designate by A.
Let a(x) = a0 + a1 x + . . . + an xn be in A[x], and let c be any root of
a(x). We will prove that c ∈ A. To begin with, all the coefficients of a(x)
are in Q(a0 , a1 , . . . , an ).
3. Prove: Q(a0 , a1 , . . . , an ) is a finite extension of Q.

83
4. Let Q(a0 , . . . , an ) = Q1 . Since a(x) ∈ Q1 [x], c is algebraic over Q1 .
Prove that:
(a) Q1 (c) is a finite extension of Q1 hence a finite extension of Q. Why?
(b) c ∈ A.
Conclusion: The roots of any polynomial whose coefficients are algebraic
numbers are themselves algebraic numbers.
A field F is called algebraically closed if the roots of every polynomial
in F [x] are in F . We have thus proved that A is algebraically closed.

84
BIBLIOGRAPHY

[1] Nguyen Thi Hong Loan, Pham Hung Quy and Nguyen Thanh Quang
(20023), Dai so dai cuong, Vinh University Publications.

[2] Charles C. Pinter (2010), A Book of Abstract Algebra, Dover Publica-


tions, Inc., Mineola, New York.

85

You might also like