Rings N Field
Rings N Field
3YK3
Part 2: ALGEBRA.
5 Polynomial Rings 45
5.1 Polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.2 Polynomials with coefficients in a field . . . . . . . . . . . . . . . . . . 46
5.3 Long division and the euclidean algorithm . . . . . . . . . . . . . . . . 47
1
2 CONTENTS
6 Field Extensions 55
6.1 Extending a given field . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.2 Algebraic number fields . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
6.3 Finite fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
Chapter 1
1.1 Introduction
The Algebra section of this course is about certain types of algebraic structure that
generalise – and include as examples – many such structures with which we are already
familiar.
For example, given two natural numbers a, b, we can add and multiply them to get
new natural numbers a + b and ab. We can also subtract one from the other, but the
result a − b is not always a natural number. (It may be a negative integer.)
If we allow a, b to be arbitrary integers, we can add, multiply and subtract them
and the result will also be an integer. We can also divide a by b (provided b 6= 0), but
the result will not always be an integer.
If a, b are arbitrary rational numbers (or real numbers, or complex numbers) then
we can add, multiply and subtract to get a new number of the same type. We can also
divide a by b if b 6= 0.
We will be interested in properties of Z, Q, R and C with respect to the algebraic
operations of addition, subtraction, multiplication and division, but we will also be
interested in similar algebraic operations on other objects.
For example, we know that we can add or subtract two vectors in Rn to get a new
vector in Rn . If A, B are n × n matrices, we can add, subtract and multiply to get new
n × n matrices A + B, A − B and AB. These operations share many of the familiar
properties of arithmetic of numbers – but not all of them.
For example, if a, b are numbers then we know that ab = ba. But there are examples
of 2×2 matrices A, B such that AB 6= BA. We are interested in developing an abstract
theory that will apply to a wide variety of different algebraic situations.
All the above examples have a common feature: they are abelian groups with respect
to addition. The purpose of this chapter is to revise the features of group theory that
are relevant to our later studies.
3
4 CHAPTER 1. REVISION OF GROUP THEORY
Lemma 1.1 If a binary operation on a set has an identity, then this identity is unique.
Proof. Suppose that e and f are both identities for a binary operation ∗ on a set A.
Then e = e ∗ f = f . The first equality holds because f is an identity. The second holds
because e is an identity.
In the last result, the associativity of ∗ is definitely used in the proof. In fact the
result is not in general true for nonassociative binary operations.
1.3 Groups
1. ∗ is associative;
1. Z, Q, R and C are all abelian groups with respect to addition. In each case 0 is
the identity and the inverse of x is −x.
2. Any vector space V is a group with respect to vector addition. The identity is
the zero vector, and the inverse of v ∈ V is −v.
5. The set of invertible n × n matrices forms a group with respect to matrix multi-
plication. The identity element is the n × n identity matrix In .
6. Let X be a set. Then the set S(X) of all permutations of X, that is, bijective
maps X → X, forms a group with respect to composition of maps. The identity
map X → X is the identity element. This group is called the symmetric group on
X. In the particular case where X is the set {1, 2, . . . , n}, this group is denoted
Sn , and called the symmetric group of degree n.
1
http://www-history.mcs.st-andrews.ac.uk/history/Mathematicians/Abel.html
6 CHAPTER 1. REVISION OF GROUP THEORY
7. Let n > 0 be an integer, and let Zn denote the set {0, 1, . . . , n − 1}. Define a
binary operation ∗ on Zn by a ∗ b = a + b if a + b < n, and a ∗ b = a + b − n
otherwise. Then Zn is an abelian group with respect to ∗, with identity 0. The
inverse of a > 0 in Zn is n−a (the inverse of 0 is 0). This group is called the cyclic
group of order n. The binary operation ∗ is usually denoted +, and referred to
as addition modulo n.
∗ e g
e e g
g g e
1.5 Subgroups
A subgroup of a group G is a subset H ⊆ G that is also a group with respect to the
same binary operation as G. Examples include Z as a subgroup of R (with respect to
addition), R∗ and S 1 as subgroups of C∗ with respect to multiplication.
It is important to recognise when a subset of a group G is actually a subgroup of
G. The following result gives a useful criterion.
Theorem 1.3 (The subgroup test) Let G be a group with respect to a binary operation
∗, and let H be a subset of G. Then H is a subgroup of G if and only if the following
three conditions are satisfied:
1. Closure: x ∗ y ∈ H ∀x, y ∈ H.
eG = eH ∗ e−1 −1 −1
H = (eH ∗ eH ) ∗ eH = eH ∗ (eH ∗ eH ) = eH ∗ eG = eH ∈ H
(where e−1
H denotes the inverse of eH in G, and all the calculations are carried out in
G.
8 CHAPTER 1. REVISION OF GROUP THEORY
(x ∗ y) ∗ z = x ∗ (y ∗ z) ∀ x, y, z ∈ G ⇒ (x ∗ y) ∗ z = x ∗ (y ∗ z) ∀ x, y, z ∈ H.
1. Let n > 0 be an integer and let nZ = {nx : x ∈ Z}, the set of integers divisible
by n. Then nZ is a subgroup of Z with respect to addition. For the closure
property, note that nx + ny = n(x + y). The identity element is 0 = n.0 ∈ nZ.
The inverse in Z of nx ∈ nZ is −(nx) = n(−x) ∈ nZ.
Proof. The set hgi is clearly nonempty and closed under ∗. It contains eG = g 0 , and the
inverse g −k of each of its elements g k . Hence hgi is a subgroup of G, by the subgroup
test.
Suppose that there are integers j < k with g j = g k . Then g k−j = eG , so g has finite
order n ≤ k − j. It follows that the n elements g 0 = eG , g 1 = g, g 2 , . . . , g n−1 of G are
pairwise distinct elements of hgi, so |hgi| ≥ n.
On the other hand, every integer k can be expressed in the form k = an + b for
some k ∈ Z and some b = 0, 1, . . . , n − 1, so
g k = (g n )a g b = (eG )a g b = g b ∈ {g 0 , g 1 , . . . , g n−1 },
and |hgi| ≤ n.
Hence |hgi| = n.
If g j 6= g k whenever j 6= k, then in particular g k 6= eG whenever k > 0, so g has
infinite order. In this case hgi also has infinite order, since the elements g k , k ∈ Z, are
pairwise distinct.
Corollary 1.6 Let G be a group and g ∈ G. Then the order of g divides that of G.
1.6 Homomorphisms
Definition Let (G, ∗) and (H, †) be groups. A map f : G → H is a homomorphism if
f (x ∗ y) = f (x)†f (y) ∀ x, y ∈ G.
Im(f ) := {f (x), x ∈ G} ⊂ H,
Ker(f ) := {x ∈ G, f (x) = eH } ⊂ G,
is a normal subgroup of G.
Proof. I will prove the first part, and leave the second as an exercise – see the example
sheet at the end of this chapter.
We use the subgroup test to check that I := Im(f ) is a subgroup of H.
Firstly, the closure property. If x, y ∈ I, then there are elements a, b ∈ G such that
f (a) = x and f (b) = y (by definition of Im(f )). Then, since f is a homomorphism, we
have
x†y = f (a)†f (b) = f (a ∗ b) ∈ Im(f ) = I.
Next, the identity property. Let eH be the identity element of H, and eG the identity
element of G. Then
f (eG )†f (eG ) = f (eG ∗ eG ) = f (eG ),
so
f (eG ) = f (eG )−1 †f (eG ) †f (eG ) = f (eG )−1 † (f (eG )†f (eG )) = f (eG )−1 †f (eG ) = eH .
(x0 y 0 )N = xn1 yn2 N = xn1 (yN ) = xn1 (N y) = x(n1 N )y = x(N y) = x(yN ) = (xy)N.
Example The set 2Z of even integers is a normal subgroup of the group (Z, +). The
quotient group Z/2Z has two elements: 0 + 2Z = 2Z (the set of all even integers and
1 + 2Z (the set of all odd integers). The Cayley table of this group is:
+ 0 + 2Z 1 + 2Z
0 + 2Z 0 + 2Z 1 + 2Z
1 + 2Z 1 + 2Z 0 + 2Z
Proof. Let K = Ker(f ) and I = Im(f ). We must define a map θ : G/K → I and prove
that it is an isomorphism.
There is only one natural way to define θ: namely θ(gK) := f (g) ∈ I.
12 CHAPTER 1. REVISION OF GROUP THEORY
First, we should check that this is well-defined. In other words, given a different
choice of coset representative g 0 ∈ gK, the definition gives the same element of I for
θ(g 0 K) = θ(gK). We can write g 0 = gk for some k ∈ K. Then
θ(g1 K.g2 K) = θ((g1 g2 )K) = f (g1 g2 ) = f (g1 )f (g2 ) = θ(g1 K)θ(g2 K).
f (g −1 g 0 ) = f (g −1 )f (g 0 ) = f (g)−1 f (g) = eH ,
so g −1 g 0 ∈ Ker(f ) = K, so g 0 K = gK.
A∼
= Zr × Zm(1) × · · · × Zm(s) .
Corollary 1.10 If A is a finite abelian group, then there are integers s ≥ 0 and
m(1), . . . , m(s) ≥ 2, which are uniquely determined by A, such that m(i) divides m(i +
1) for 1 ≤ i ≤ s − 1, and
A∼= Zm(1) × · · · × Zm(s) .
1.8. FINITELY GENERATED ABELIAN GROUPS 13
2. Let M = Mn (R) be the set of 3 × 3 matrices with real coefficients, and let S ⊂ M
be the set of symmetric matrices (that is, matrices A such that A = AT ). Use the
subgroup test to show that S is a subgroup of (M, +) (where + denotes matrix
addition).
(a ∗ H) ⊕ (b ∗ H) = (a ∗ b) ∗ H
Assuming this is well defined (as shown in the notes) show that this gives (G/H, ⊕)
the structure of a group.
6. Describe the elements in each of the following quotient groups (i.e. describe the
appropriate cosets).
(i) C/R;
(ii) Z/3Z;
Can these quotient groups be described in more familiar terms i.e. are they
isomorphic to other groups you know about?
8. Let n ∈ N. Use the First Isomorphism Theorem for Groups to show that Z/nZ ∼
=
Zn . [Hint: Define an appropriate function φ: Z → Zn ].
14 CHAPTER 1. REVISION OF GROUP THEORY
Chapter 2
2.1 Rings
Definition A ring (R, +, ·) is a set R together with two binary operations + (called
addition) and · (called multiplication) that satisfy the following axioms.
Remarks
15
16 CHAPTER 2. RINGS, FIELDS AND INTEGRAL DOMAINS
Examples
3. For similar reasons, the set Z[i] := {m + ni; m, n ∈ Z} ⊂ C is a ring with respect
to addition and multiplication of complex numbers, where i ∈ C denotes a square
root of −1.
The elements of Z[i] are called Gaussian integers.
5. Let Mn (R) denote the set of all n × n matrices with real entries. Then Mn (R) is
a ring with respect to addition and multiplication of matrices. It has an identity,
namely the n × n identity matrix I = In . It is not commutative when n ≥ 2,
since, for example,
1 1 1 0 2 1 1 1 1 0 1 1
= 6= = .
0 1 1 1 1 1 1 2 1 1 0 1
Similarly, the sets Mn (Z), Mn (Q), Mn (C) of n × n matrices with integer, rational
and complex entries, respectively, are noncommutative rings with identity, under
addition and multiplication of matrices.
6. Let X be any set, and let RX denote the set of all functions X → R. Define
addition and multiplication pointwise on RX , that is:
Then RX is a commutative ring with identity. The additive identity is the con-
stant function 0, and the multiplicative identity is the constant function 1.
7. Define R[x] to be the set of all polynomials with real coefficients in the variable
x. Elements of R[x] are formal sums
m
X
p(x) = ak xk = am xm + am−1 xm−1 + · · · + a1 x + a0 ,
k=0
where the operations in the first coordinate position are taking place in R and those
in the second coordinate position are taking place in S.
It is easy to check that this makes R × S into a ring. The additive identity is
(0R , 0S ).
If (a, b) ∈ R × S is a multiplicative identity, than for any (r, s) ∈ R × S we have
(ar, bs) = (a, b)(r, s) = (r, s) = (r, s)(a, b) = (ra, sb),
so ar = r = ra in R and bs = s = sb in S. In other words, a is a multiplicative
identity in R and b is a multiplicative identity in S. Conversely, if each ring R, S has
an identity, then R × S has an identity (1R , 1S ).
2. 0 = 0 + 0, so
0 + 0a = 0a = (0 + 0)a = 0a + 0a.
By the previous property, 0 = 0a. Similarly 0 = a0.
3. ab + (−a)b = (a − a)b = 0b = 0, so (−a)b = −(ab). Similarly a(−b) = −(ab).
4. By the previous property, ab = −(−ab) = −(a(−b)) = (−a)(−b).
Remark The first property in the above lemma says that addition is cancellative –
we can cancel an equal term from each side of an equation involving addition. The
corresponding property is not true in general for multiplication. In other words, it is
possible that ab = ac but b 6= c. Easy examples of this come from the second property
by taking a = 0: 0b = 0 = 0c for all b, c ∈ R. But there are also more subtle examples
of this phenomenon, such as a = 2, b = 1, c = 7 in Z12 : ab = 2 = ac in Z12 , but b 6= c
in Z12 .
2.4. SUBRINGS 19
2.4 Subrings
Just as we can find many examples of groups as subgroups of other groups, many rings
naturally exist as subrings of other rings.
Definition A subring of a ring R is a subset S of R which is itself a ring with respect
to the same addition and multiplication as R.
Examples
There is an easy test for a subset of a ring to be a subring, analogous to the subgroup
test in group theory.
Lemma 2.2 (The Subring Test) A subset S of a ring R is a subring if and only if
Proof. Suppose that S is a subring of R. Then it is a ring with respect to the addition
and multiplication of R. In particular it is a group with respect to the addition of
R, in other words a subgroup of (R, +). It must also be closed with respect to the
multiplication of R, since multiplication on R gives a binary operation on S.
Conversely, suppose that the two conditions of the lemma hold. In particular, the
addition and multiplication of R give binary operations on S, and (S, +) is a group,
being a subgroup of (R, +). Indeed, since (R, +) is an abelian group, so is any of its
subgroups, so (S, +) is an abelian group. Multiplication on S is associative, since it is
associative on the larger set R. For the same reason, multiplication on S is both left
and right distributive over +. Hence (S, +, ·) is a ring, so S is a subring of R.
Remark If R is a commutative ring, then every subring of R is also commutative. If
R has identity, then a subring of R may have identity, but need not.
Examples
2.5 Fields
A ring is an algebraic object in which we can add, subtract, and multiply, but not
necessarily divide (since elements do not need to have inverses).
In many of our nicest rings (such as Q or C, we can divide by any element except
0, since every nonzero element has an inverse. Rings like this are called fields.
Definition A field is a commutative ring with identity element 1 6= 0, in which every
nonzero element x 6= 0 has an inverse x−1 .
Remark It is important to allow 0 to be an exceptional element with no inverse.
Suppose that R is a ring with identity, in which 0 has an inverse. Then 1 = 0.0−1 = 0,
and so, for each r ∈ R, we have r = 1.r = 0.r = 0. In other words, R = {0}, which is
a ring but not a very interesting one.
Examples
1. Q, R and C are fields with respect to the usual addition and multiplication.
To check that it is a subring, we use the subring test: Q[i] is closed under addition
and multiplication ((a + bi) + (c + di) = (a + c) + (b + d)i ∈ Q[i], (a + bi)(c + di) =
(ac − bd) + (ad + bc)i ∈ Q[i]); contains the zero element 0 + 0i, and the additive
inverse −a − bi of each of its elements a + bi.
Q[i] is certainly commutative, being a subring of C, and it contains the identity
1 = 1 + 0i of C. Finally, if a + bi 6= 0 in Q[i], then its multiplicative inverse is
a − bi
∈ Q[i].
a2 + b 2
3. Zp is a field if p is a prime number.
We already know that Zp is a commutative ring with identity, so we need only
check that every nonzero element has an inverse. Suppose x 6= 0 in Zp . Consider
the map f : Zp → Zp given by f (k) = xk mod p. If f (j) = f (k) for some j, k
with 0 ≤ j < k < p, then x(k − j) = xk − xj = f (k) − f (j) = 0 in Zp , so the
integer x(k − j) is divisible by the prime number p, and hence one of x, k − j is
divisible by p (since p is a prime). But this is a contradiction, since 0 < x < p
and 0 < k − j < p by hypothesis.
Hence f is injective, and so by the pigeonhole principle it is also surjective. Hence
there is a (unique) y ∈ Zp with xy = f (y) = 1.
4. Let F denote the set {0, 1, x, y} of four elements, and let addition and multipli-
cation on F be defined by the Cayley tables:
+ 0 1 x y × 0 1 x y
0 0 1 x y 0 0 0 0 0
1 1 0 y x 1 0 1 x y
x x y 0 1 x 0 x y 1
y y x 1 0 y 0 y 1 x
p(x)q(x) = am bn xm+n + . . .
Proof. Let R be a finite integral domain. Then R is a commutative ring with identity.
To prove that R is a field, we need only check that every nonzero element has an
inverse.
Suppose that x ∈ R with x 6= 0. Then, by the Lemma above, xy 6= xz whenever
y 6= z in R. Thus the map f : R → R defined by f (y) = xy is injective. Since R is
finite, the pigeonhole principle implies that f is also surjective. In particular, ∃ y ∈ R
with xy = f (y) = 1R . Hence x has an inverse, as required.
24 CHAPTER 2. RINGS, FIELDS AND INTEGRAL DOMAINS
(a) (S, +, ·), where S = {2k + 1 : k ∈ Z} ∪ {0} and + and · denote the usual
addition and multiplication of real numbers.
(b) (S, +, ·), where S denotes the family of all functions from R to R, and +
and · denote the usual (pointwise) addition and multiplication of functions:
(f + g)(x) = f (x) + g(x), (f g)(x) = f (x)g(x).
(c) (S, +, ◦), where S denotes the family of all functions from R to R, and +
denotes pointwise addition and ◦ denotes composition of functions.
(d) (S, +, ·), where S is the family of all subsets of a given set E and + and ·
are defined by
A + B = (A \ B) ∪ (B \ A) ; A·B =A∩B.
3. In the ring Z48 , find all elements x satisfying x2 = 0, and all elements y satisfying
y 3 = 0.
4. In the ring Z7 × Z7 , find all elements x satisfying x2 = (−1, 1), and all elements
y satisfying y 3 = (1, 0).
5. Let (R, +, ·) be a ring such that x2 = x for all x ∈ R. Show that for all x, y ∈ R,
xy = −yx. [Hint: consider (x + y)2 .] Deduce that R is commutative.
6. Let S denote the set of all rational numbers of the form a/2n , where a, n ∈ Z
and n ≥ 0. Show that S is a subring of Q.
7. Show that the set R of complex numbers of the form a + 2bi, with a, b ∈ Z, is a
subring of C.
a ⊕ b = a + b + 1; a ⊗ b = a + b + ab .
2.6. INTEGRAL DOMAINS 25
10. Let (F, +, ·) be a field and let x, y ∈ F . Is it necessarily true that x3 = y 3 implies
x = y?
11. Let p be a prime number. Find all elements in the field (Zp , +, ·) which are their
own multiplicative inverses.
12. Let R and S be two rings, each with more than one element. Show that there are
nonzero elements x and y in R × S such that xy = 0. In the case where R = Z2
and S = Z3 , find all such pairs x, y ∈ R × S.
13. Determine the units and the divisors of zero of (Z12 , +, ·). Write down the group
table of the group of units of Z12 .
15. Show that a subring R 6= {0} of a field F is an integral domain if and only if
1F ∈ R.
16. Let (R, +, ·) be a commutative ring with identity. Prove that R is an integral
domain if for a, b, c ∈ R with a 6= 0 the relation ab = ac implies that b = c.
26 CHAPTER 2. RINGS, FIELDS AND INTEGRAL DOMAINS
Chapter 3
In group theory, we construct the quotient group G/N of a group G over a normal
subgroup N . The normal subgroups are precisely the kernels of homomorphisms be-
tween groups. In this chapter we follow the entirely analogous story in ring theory. As
in group theory, a ring homomorphism is a map which respects the binary operations.
The analogue of a normal subgroup is something called an ideal in a ring. Given an
ideal, we can construct a quotient ring.
3.1 Homomorphisms
Definition A homomorphism from a ring R to a ring S is a map f : R → S such that
(In each case the operation of + or · on the left side of the equation takes place in R,
while that on the right side takes place in S.)
If the homomorphism f : R → S is bijective, then it is called an isomorphism, and
we say that the rings R and S are isomorphic (denoted R ∼ = S). The relation ∼= is an
equivalence relation between rings. We regard isomorphic rings as being ‘the same’.
If f : R → S is a homomorphism of rings, then in particular f is a homomorphism
of groups from (R, +) to (S, +). From this we have an immediate list of properties of
f:
1. f (0R ) = 0S ;
2. f (−x) = −f (x) ∀ x ∈ R;
27
28 CHAPTER 3. HOMOMORPHISMS, IDEALS, AND QUOTIENT RINGS
4. Ker(f ) is a subgroup of (R, +). (Indeed, Ker(f ) is a normal subgroup of (R, +).
But in any case, since (R, +) is an abelian group, all its subgroups are normal.)
It is not true in general that f (1R ) = 1S , even when both rings R, S have identities.
For example, the map f : Z → Z × Z given by f (r) = (r, 0) is a homomorphism, but
f (1) 6= (1, 1). On the other hand, suppose that f : R → S is an isomorphism Then S
has an identity if and only if R has an identity, and in this case f (1R ) = 1S . To see
this, suppose that R has an identity, and let s ∈ S. Then
and similarly sf (1R ) = s, so f (1R ) is an identity for S. Applying the same argument
to the isomorphism f −1 : S → R gives the converse.
Example The rings R and C are not isomorphic. The reason is that C contains a square
root of −1, but R does not. In detail, suppose that f : C → R is an isomorphism.
Then f (0) = 0, so f (1) 6= 0, since f is injective. But f (1) = f (12 ) = f (1)2 , so
f (1) ∈ R is a nonzero solution of the equation x2 = x. There is only one such
solution, namely x = 1, so f (1) = 1, and hence also f (−1) = −f (1) = −1. Finally,
f (i)2 = f (i2 ) = f (−1) = −1, so f (i) is a square root of −1 in R, a contradiction.
Remark This last example depends in a crucial way on the difference between the mul-
tiplicative structures of R and C. Indeed, as groups, (R, +) and (C, +) are isomorphic.
(They are vector spaces of equal infinite dimension over Q.)
Proof. We have already noted that Im(f ) is a subgroup of (S, +) (since f is a group
homomorphism from (R, +) to (S, +)).
It only remains to show that Im(f ) is closed under multiplication. So suppose that
x, y ∈ Im(f ). Then there are elements a, b ∈ R such that f (a) = x and f (b) = y (by
definition of Im(f )). Since f is a homomorphism, we have
Examples
3.2 Ideals
In this section we study ideals, which are the analogues of normal subgroups in group
theory. Thus ideals should be the objects that occur as kernels of ring homomorphisms.
Suppose then that R, S are rings, and that f : R → S is a ring homomorphism. We
know that Ker(f ) is a subgroup of (R, +). It is not difficult to see that in fact Ker(f )
is closed with respect to multiplication, so is a subring of R:
However, a stronger property holds. In the above equation, in order for f (x)f (y)
to be 0 in S, we do not need both f (x), f (y) to be 0. It is sufficient for any one of
them to be 0 in S.
This suggests the following definition.
Definition An ideal in a ring R is a subset I ⊂ R such that
1. I is a subgroup of (R, +)
2. (∀ x ∈ I) (∀ r ∈ R) xr ∈ I and rx ∈ I.
Thus to check that a given subset I ⊂ R is an ideal of R, we check the above two
properties. For the first, we can use the subgroup test, or we may already know for
other reasons that I is a subgroup of (R, +). The key property to check is usually the
second property rx, xr ∈ I, which I will refer to as the ideal property. In practice, most
30 CHAPTER 3. HOMOMORPHISMS, IDEALS, AND QUOTIENT RINGS
of the rings that we consider will be commutative, in which case the two statements
xr ∈ I and rx ∈ I are equivalent (since rx = xr), so we need only check one of them.
Examples
Proof. We already know that (R/I, +) is a group. Indeed, it is an abelian group since
+ is commutative in R:
(r + I) + (s + I) = (r + s) + I = (s + r) + I = (s + I) + (r + I).
Examples
1. The quotient ring Z/nZ has n elements 0 + nZ, 1 + nZ, . . . (n − 1) + nZ. The sum
or product of two cosets a + nZ and b + nZ is the coset containing a + b or ab,
respectively. Thus addition and multiplication in Z/nZ are the same as in the
ring Zn , if we identify k ∈ Zn with the coset k + nZ in Z/nZ. In other words the
quotient ring Z/nZ is isomorphic to Zn .
2. Let I be the principal ideal (x2 + 1)R[x] in R[x]. Then for any m ≥ 2 the coset
xm +I is the same as the coset −xm−2 +I, since xm −(−xm−2 ) = (x2 +1)xm−2 ∈ I.
If p(x) = am xm + . . . + a1 x + a0 , then
p(x) + I = (am xm + I) + . . . + (a1 x + I) + (a0 + I) = (a + bx) + I,
where a = a0 − a2 + a4 − . . . and b = a1 − a3 + a5 − . . .. Hence every element of
R[x]/I can be (uniquely) expressed in the form (a + bx) + I with a, b ∈ R.
Addition and multiplication in R[x]/I are defined by
(a + bx + I) + (c + dx + I) = (a + c) + (b + d)x + I,
and
(a + bx + I)(c + dx + I) = ac + (ad + bc)x + bdx2 + I = (ac − bd) + (ad + bc)x + I.
These are similar to the rules for adding and multiplying complex numbers, and
indeed the quotient ring R[x]/I is isomorphic to C via the map (a+bx)+I 7→ a+bi.
3. Let I ⊂ R[x] be the principal ideal x2 R[x]. As in the previous example, each
coset in R[x]/I can be uniquely expressed as a + bx + I with a, b ∈ R. As
an additive group, R[x]/I ∼ = R2 ∼ = C, but the multiplication rule in R[x]/I
is different from that in C. We can regard R[x]/I as R2 with multiplication
given by (a, b)(c, d) = (ac, ad + bc). Note that (0, 1)2 = (0, 0) – or, equivalently,
(x + I)2 = x2 + I = I, so that R[x]/I has zero-divisors, and so cannot be a field.
as required.
Examples
S ∼ S+I
= .
S∩I I
34 CHAPTER 3. HOMOMORPHISMS, IDEALS, AND QUOTIENT RINGS
Z4 ∼ Z/4Z ∼ Z ∼
= = = Z2 .
I 2Z/4Z 2Z
3.4. MORE ISOMORPHISM THEOREMS 35
Throughout this chapter, and indeed for the rest of the course, we will restirct our
attention to rings which are commutative and contain an identity.
In this chapter we study three special kinds of ideals in such rings.
3. Let I = {p(x) ∈ Z[x]; p(0) ∈ 2Z}. Then I is an ideal in Z[x] – for example
since it is the kernel of the evaluation homomorphism f : Z[x] → Z2 , f (p(x)) =
p(0) mod 2. But I is not principal. To see this, we must show that I 6= p(x)Z[x]
for any p(x) ∈ Z[x]. Suppose first that p(x) is the constant polynomial a0 for
some a0 ∈ Z. Then a0 = p(0) ∈ 2Z, so I = a0 Z[x] ⊂ 2Z[x], which is impossible
since for example x ∈ I but x ∈ / 2Z[x]. On the other hand, if p(x) has degree
m > 0, then p(x)q(x) has degree m or greater for any nonzero polynomial q(x), so
there are no nonzero constant polynomials in I = p(x)Z[x]. But this contradicts
the fact that 2 ∈ Z[x].
Lemma 4.1 Let R be a commutative ring with identity, and let x ∈ R. Then the
principal ideal xR is the smallest ideal in R that contains x.
37
38 CHAPTER 4. SPECIAL TYPES OF IDEALS
Proof. We have already seen that xR is an ideal. Since R has an identity, we see that
x = x1R ∈ xR.
Conversely, suppose that I is an ideal in R and that x ∈ R. Then the ideal property
says that xr ∈ I for all r ∈ R, and so xR ⊂ I.
Definition A principal ideal domain (or PID) is an integral domain in which every
ideal is principal.
Examples
1. The polynomial ideal Z[x] is an integral domain, but is not a principal ideal
domain, since it contains an ideal which is not principal. (See the example above.)
3. Every field is a principal ideal domain. The only two ideals in a field F are
{0} = 0F and F = 1F . Each of these is principal.
4. If F is a field, then the polynomial ring F [x] is a principal ideal domain. The
proof is similar to the case of Z. We already know that F [x] is an integral domain.
The ideal {0} = 0F [x] is principal, so we suppose I 6= {0} is an ideal and show
that it must be principal.
Choose a nonzero polynomial p(x) ∈ I of least degree (m, say). Then p(x)F [x] ⊂
I. We will show that I = p(x)F [x].
If I 6= p(x)F [x], let q(x) be a polynomial in I \ p(x)F [x] of least degree (n, say).
Then n ≥ m, by choice of p(x). Suppose that the leading coefficient of p(x) is
am , while that of q(x) is bn . Let r(x) = am q(x) − bn xn−m p(x). Then r(x) ∈ I
by the ideal properties, since p(x), q(x) ∈ I. Moreover, r(x) has degree at most
4.2. MAXIMAL IDEALS 39
5. The ring Z[i] of Gaussian integers is a principal ideal domain. Certainly Z[i] is
an integral domain, since it is a subring of the field C that contains the identity
of C. The ideal {0} = 0Z[i] is principal. Suppose that I 6= {0} is an ideal. We
must show that I is principal.
Choose a nonzero element z = a + ib ∈ I such that |z|2 = a2 + b2 ∈ N is
least possible. Clearly zZ[i] ⊂ I. We will show that I = zZ[i]. Suppose that
w = c + id ∈ I. Then v = (w/z) = x + iy ∈ C is a complex number. If we let e, f
be the integers closest to x, y respectively, then u = e + if ∈ Z[i], and |u − v|2 =
|x − e|2 + |y − f |2 ≤ 12 . Now zu − w ∈ I, and |zu − w|2 = |z|2 |u − v|2 ≤ 12 |z|2 . By
choice of z, we must have zu − w = 0, so w = zu ∈ zZ[i], as required.
2. If F is a field, then {0} is a maximal ideal, since the only other ideal of F is F
itself.
ad xd +· · ·+a1 x+a0 . If d > 1 then put q(x) = p(x)−ad xd−2 (x2 +1) ∈ I. Then q(x)
has degree less than d, so q(x) ∈ M . But then p(x) = q(x) + ad xd−2 (x2 + 1) ∈ M ,
contradicting the choice of p(x). We must therefore have d ≤ 1, say p(x) = ax+b,
with (a, b) 6= (0, 0) since p(x) 6= 0 in R[x].
Then a2 + b2 = a2 (x2 + 1) − (ax − b)p(x) ∈ I. Since a, b ∈ R with (a, b) 6= (0, 0),
c = a2 + b2 > 0, and c is a unit in R[x]. Since I contains a unit, it must be the
whole ring. (In other words, p(x) = c(c−1 p(x)) ∈ I for all p(x) ∈ R[x].)
The most important property of maximal ideals relates to the corresponding quo-
tient rings.
Theorem 4.2 Let R be a commutative ring with identity, and let M be an ideal in R.
Then M is a maximal ideal if and only if the quotient ring R/M is a field.
Proof. Suppose first that M is maximal. The ring R/M is commutative, since R is
commutative. ((x + M )(y + M ) = xy + M = yx + M = (y + M )(x + M ).) It also has
an identity 1R/M = 1R + M . Moreover, 1R/M 6= 0R/M since 1R ∈ / M . (If 1R ∈ M , then
M = R, contadicting the definition of maximal ideal.)
To show that R/M is a field, it remains to prove that every nonzero element x + M
in R/M has an inverse. Now x ∈ / M , since x + M 6= 0 + M . We will show that
I = M + xR = {m + xr; m ∈ M, r ∈ R} is an ideal in R. Clearly M ⊂ I, and M 6= I
since x ∈ I and x ∈/ M . Since M is maximal, it follows that I = R, and in particular
1R ∈ I – say 1R = m + xy with m ∈ M and y ∈ R. Then (x + M )(y + M ) = xy + M =
(1R − m) + M = 1R + M , so x + M has an inverse in R/M , as required.
We still have to check that I is an ideal. Let m1 , m2 ∈ M and r1 , r2 , s ∈ R.
Then (m1 + xr1 ) + (m2 + xr2 ) = (m1 + m2 ) + x(r1 + r2 ) ∈ I, 0 = 0 + x0 ∈ I,
−(m1 +xr1 ) = (−m1 )+x(−r1 ) ∈ I, and s(m1 +xr1 ) = (m1 +xr1 )s = m1 s+x(r1 s) ∈ I.
Hence I is an ideal, as claimed.
Conversely, suppose that R/M is a field. Then M 6= R, since otherwise 0 = 1 in
R/M , contrary to the definition of a field. Suppose I is an ideal of R with M ⊂ I ⊂ R.
By the third isomorphism theorem, I/M is an ideal in the field F = R/M . But a
field F has only two ideals F and {0}. If I/M = F = R/M then I = R, while if
I/M = {0} = M/M then I = M . Hence M is maximal, as required.
Theorem 4.3 Let R be a commutative ring with identity, and let P be an ideal in R.
Then P is a prime ideal if and only if the quotient ring R/P is a non-zero integral
domain.
7. Let R be a finite commutative ring with identity. Show that every prime ideal of
R is a maximal ideal.
44 CHAPTER 4. SPECIAL TYPES OF IDEALS
Chapter 5
Polynomial Rings
In this chapter we will study rings of the form R[x], the ring of polynomials in one
variable x with coefficients from the ring R.
In practice, we will mainly be interested in the case where R is a field, or at least
an integral domain. But the definition makes sense for any commutative ring R.
5.1 Polynomials
Let R be a commutative ring, and let n ≥ 0 be an integer. A polynomial in x of degree
n with coefficients from R is a formal expression
Lemma 5.1 1. If p(x), q(x) ∈ R[x] have degrees m, n respectively, and m 6= n, then
p(x) + q(x) has degree max(m, n). (If m = n, then the degree of p(x) + q(x) is at
most n.)
3. If p(x), q(x) ∈ R[x] have degrees m, n respectively, then the degree of p(x)q(x) is
at most m + n. If R is an integral domain and p(x) 6= 0 6= q(x), then the degree
of p(x)q(x) is exactly m + n.
45
46 CHAPTER 5. POLYNOMIAL RINGS
In this course, we will however concentrate on polynomial rings in one variable only.
Lemma 5.3 Let F be a field and p(x), q(x) ∈ F [x]. Then p(x)F [x] = q(x)F [x] if and
only if q(x) = ap(x) for some nonzero constant a ∈ F .
5.3. LONG DIVISION AND THE EUCLIDEAN ALGORITHM 47
Proof. Now p(x)F [x] = q(x)F [x] if and only if there are polynomials a(x) and b(x)
such that q(x) = p(x)a(x) and p(x) = q(x)b(x). This is true if and only if p(x)(1 −
a(x)b(x)) = 0 = q(x)(1 − b(x)a(x)). Since F [x] is an integral domain, this means either
p(x) = 0 = q(x) or a(x)b(x) = 1.
In the second case, the degrees of a(x) and b(x) are both 0, since the product
a(x)b(x) has degree 0, so a is a nonzero constant in F , with inverse b.
Conversely, if q(x) = ap(x) ∈ p(x)F [x] with a a nonzero constant, then a is a
unit in F , so has an inverse, b ∈ F say, and then p(x) = bq(x) ∈ q(x)F [x], and so
p(x)F [x] = q(x)F [x].
Corollary 5.4 Every nonzero ideal I in F [x] has the form p(x) for a monic polynomial
p(x). Moreover, this choice of monic polynomial is uniquely determined by I.
Proof. Certainly I is principal, so I = q(x)F [x] for some nonzero polynomial q(x). Since
q(x) 6= 0, the leading coefficient an (say) of q(x) is nonzero. Define p(x) = a−1 n q(x).
−1
Then p(x) has leading coefficient an an = 1, so is monic. Also, by the lemma, we have
p(x)F [x] = q(x)F [x] = I.
If r(x) is another monic polynomial that generates I, then the lemma says that
r(x) = ap(x) for some constant a 6= 0. Comparing leading coefficients, we see that
1 = a.1, so a = 1 and r(x) = p(x), as claimed.
96 = 1 × 63 + 33.
63 = 1 × 33 + 30.
33 = 1 × 30 + 3.
30 = 10 × 3 + 0.
The last remainder is zero, so the previous remainder, 3 is the highest common factor.
It turns out that the same algorithm works for polynomials with coefficients from a
field F , where the measurement we use for the size of a polynomial is its degree. The
individual division steps of the algorithm work because of the following result.
Lemma 5.5 Let F be a field and a(x), b(x) ∈ F [x] \ {0}. Then there are unique
polynomials q(x), r(x) ∈ F [x] such that
Proof. Let I be the principal ideal a(x)F [x]. If b(x) ∈ I, then we have b(x) = a(x)q(x)
for some q(x) ∈ F [x] (which is unique, since F [x] is an integral domain), and the result
is true with r(x) = 0.
Otherwise, choose r(x) to be a polynomial of least possible degree in the coset
b(x) + I. This degree is less than that of a(x), as the following argument shows.
Suppose that a(x) has degree m and leading coefficient α, whereas r(x) has degree
n ≥ m and leading coefficient ρ. Then r(x) − α−1 ρxn−m a(x) ∈ r(x) + I = b(x) + I has
degree less than n, contrary to the choice of r(x).
It follows that the choice of r(x) is unique – if r0 (x) ∈ r(x) + I also has smaller
degree than a(x), then r(x) − r0 (x) ∈ I = a(x)F [x] is a multiple of a(x) but has degree
less than that of a(x), so must be zero.
Finally, b(x) − r(x) ∈ I = a(x)F [x], so b(x) − r(x) = a(x)q(x) for a (unique)
polynomial q(x).
To perform a division of polynomials – that is, to find q(x) and r(x) in the notation
of the lemma – we can proceed in stages corresponding to the terms of b(x). Suppose
that b(x) has degree k and leading coefficient β, while a(x) has degree m ≤ k and
leading coefficient α, as in the lemma. Then b1 (x) = b(x) − α−1 βxk−m a(x) has degree
less than k. Iterate this process. If b1 (x) = a(x)q1 (x)+r(x), then b(x) = a(x)q(x)+r(x),
where q(x) = α−1 βxk−m + q1 (x).
We can lay this calculation out as a long division. For example:
5.4. REDUCIBLE AND IRREDUCIBLE POLYNOMIALS 49
x2 + 2x + 11
x2 + x + 1)x4 + 3x3 + 14x2 − 7x − 5
x4 + x3 + x2
2x3 + 13x2
2x3 + 2x2 + 2x
11x2 − 9x
11x2 + 11x + 11
− 20x − 16
Theorem 5.6 Let F be a field, and let p(x) be a non-constant polynomial in F [x]. Let
I denote the principal ideal I = p(x)F [x]. Then the following are equivalent:
(i) I is maximal;
(ii) I is prime;
(ii)⇒(iii). If p(x) is reducible, say p(x) = a(x)b(x) where a(x) and b(x) each has
degree less than that of p(x), then a(x), b(x) ∈ / p(x)F [x] = I, but a(x)b(x) = p(x) ∈ I,
so I is not prime.
(iii)⇒(i). It remains to show that, if p(x) is irreducible, then I is maximal. Suppose
than that I ⊂ J ⊂ F [x] for some ideal J of F [x]. Since F [x] is a PID, J(x) = a(x)F [x]
for some a(x). Now p(x) ∈ I ⊂ J = a(x)F [x], so p(x) = a(x)b(x) for some b(x).
Since p(x) is irreducible, it is not possible that both a(x) and b(x) have degrees less
than p(x). Hence one of a(x), b(x) has degree equal to that of p(x), and the other is a
nonzero constant (and hence a unit in F [x]).
There are two cases to consider. If a(x) has the same degree as p(x), then b = b(x)
is a unit. Then a(x) = b−1 p(x) ∈ I, so J ⊂ I and hence in fact I = J. If, on the other
hand, a = a(x) is a unit, then J = aF [x] = F [x]. We have therefore shown that I is
maximal.
Remark Whether or not a given polynomial is irreducible may depend on the field of
coefficients in which we are working, as the following examples show.
Examples
so a = b = 0 and 1 = ac = 0c = 0, a contradiction.
The ring of polynomials over a field resembles in many ways the ring of integers,
with irreducible polynomials playing the part of prime numbers. In particular, there
is the following anaologue of the Fundamental Theorem of Arithmetic, whose proof we
will omit.
Theorem 5.7 Let F be a field, and p(x) ∈ F [x] a nonconstant polynomial. Then
there is a unique constant c and a list α1 (x), . . . , αk (x) (unique up to order) of monic
irreducible polynomials, such that
x 0 1 2 3 4 5 6
x3 + 2 2 3 3 1 3 1 1
2. x4 + x + 1 has no roots in Z2 :
x 0 1
4
x +x+1 1 1
m3 m
+ + 1 = 0.
n3 n
Multiplying both sides by n3 6= 0, we get m3 + mn2 + n2 = 0. Form this equation, we
see that, if m has a prime factor p, then n3 = 0 mod p, so p is also a prime factor of
n. Similarly, if q is a prime factor of n, then m3 = 0 mod q, so q is also a prime factor
of m. But we can choose m, n to have no common factors, so the only possibility is
m = ±n = ±1. By direct evaluation, neither 1 nor −1 is a root of x3 + x = 1, so
x3 + x = 1 has no rational roots, and so is irreducible.
There are two other results which help us decide questions of reducibility in Q[x]:
Theorem 5.12 (Gauss’ Lemma) Let p(x) ∈ Z[x]. Then p(x) is reducible in Q[x] if
and only if it is reducible in Z[x].
Examples
1. x300 − 17 is irreducible in Q[x] (by Eisenstein’s criterion with p = 17).
5.5. TESTING FOR IRRECUCIBILITY 53
(a) x2 + x + 1 in Z2 [x];
(b) x3 + x2 + 3x + 5 in Z7 [x];
(c) x2 + 5x − 3 in R[x];
(d) x3 − 17x2 + 24x − 1 in R[x];
(e) x345 − 53x77 + 1234567x22 − 2x + 1 in C[x].
(f) x2 + x − 5 in Q[x];
(g) x5 + x2 + x in Z2 [x].
in Z5 [x].
3. Let F be a field and I 6= F [x] an ideal in F [x]. Let g(x) ∈ F [x] be an irreducible
polynomial such that g(x) ∈ I. Show that I = g(x)F [x].
7. We have seen in lectures that, if F is a field, then F [x] is a principal ideal domain.
The converse is also true: if R is a commutative ring such that R[x] is a principal
ideal domain, then R is a field. Prove this in steps as follows:
Field Extensions
2. Let F = Q, and let p(x) = x2 + 1, which is irreducible in Q[x]. Then the quotient
field Q[x]/(x2 + 1)Q[x] is isomorphic to the field Q[i] of Gaussian rationals.
3. Let F = Z2 , and let p(x) = x2 +x+1 (the only irreducible quadratic polynomial in
Z2 [x]. Then the resulting field Z2 [x]/(x2 +x+1)Z2 [x] has four elements 0 = 0+I,
1 = 1 + I, a = x + I, b = (x + 1) + I, where I = (x2 + x + 1)Z − 2[x]. The addition
and multiplication tables can be deduced from the rule that (x2 +x+1)+I = 0+I:
for example, a2 = x2 + I = x + 1 + I = b in this field. In fact, it is easy to check
that this field is isomorphic to the field of four elements we saw in an earlier
chapter.
In each of these examples, the first field F is (isomorphic to) a subfield of the
resulting field K. Indeed, this is a general feature of our construction. To see this, recall
that the units in the polynomial ring F [x] are just the nonzero constant polynomials,
in other words the units of F . Since the maximal ideal I = p(x)F [x] is not the whole
ring, it cannot contain any units, so the homomorphism f : F → K, f (r) = r + I, is
injective. (Here we are regarding r ∈ F as a constant polynomial in F [x].)
55
56 CHAPTER 6. FIELD EXTENSIONS
Theorem 6.1 (Kronecker’s Theorem) Let F be a field, and p(x) a non-constant poly-
nomial in F [x]. Then there exists a field K, containing F as a subfield, such that p(x)
has a root in K.
Proof. If p(x) is irreducible, then the construction described above produces the desired
field K. If p(x) is reducible, then it has at least one irreducible factor, q(x) say. Our
construction produces a field K ⊃ F such that K contains a root α of q(x). Since p(x)
is a multiple of q(x), α is also a root of p(x).
Let us take a closer look at the field K = F [x]/I that we have constructed, where
I = p(x)F [x] and p(x) is a monic irreducible polynomial in F [x]. We know that F ⊂ K
and that K contains a root α = x + I of p(x). What are the other elements of K? Of
course, each element of K is a coset a(x) + I = a(α) for some a(x) ∈ F [x]. The element
a(x) is not unique, but if two elements a(x), b(x) ∈ F [x] define the same element of K,
then a(x) − b(x) ∈ I = p(x)F [x]. Given a(x) ∈ F [x], there is a unique representation
a(x) = q(x)p(x) + r(x) with r(x) = 0 or deg(r(x)) < deg(p(x)). In other words, there
is a unique representative r(x) ∈ a(x) + I with r(x) = 0 or deg(r(x)) < deg(p(x)).
If deg(p(x)) = 1, then this says that every coset of I is represented by an element
of F , so the map f : F → F/I = K is an isomorphism.
If deg(p(x)) = n > 1, then p(x) = xn − c(x) for some c(x) 6= 0 with deg(c(x)) < n.
The elements of K correspond to the polynomials of degree less than n in F [x]. These
form an n-dimensional vector space over F , with basis B = {1, α, α2 , . . . , αn−1 }. A
typical element has the form
Examples
√
2. Consider the complex number ω = − 21 + 23 i. This is a cube root of√unity in
C. It is easy to show that ω 2 is the complex conjugate ω = − 12 − 23 i of ω,
and ω 3 = 1. Indeed, ω 2 + ω + 1 = 0 in C, so ω is a root of the irreducible
polynomial x2 + x + 1 ∈ Q[x]. Hence the set Q[ω] of all complex numbers of
the form a + bω, a, b ∈ Q forms a subfield of C isomorphic to the quotient field
Q[x]/(x2 + x + 1)Q[x].
The elements of Q[x] have the form a + bω and addition rule
The multiplication rule in Q[ω] is derived in the same way as that for C, but
using the rule that ω 2 + ω + 1 = 0, or alternatively, ω 2 = −1 − ω. Hence
(1 + x2 )(x + x2 ) = x + x2 + x3 + x4 = x + x2 + (1 + x) + (x + x2 ) = 1 + x.
58 CHAPTER 6. FIELD EXTENSIONS
Not every complex number is algebraic. Indeed, most complex numbers are not
algebraic, in the following sense. The set of algebraic numbers can be shown to be
countable, that is, there is a bijection between that set and the set N of natural numbers.
On the other hand, the set C of complex numbers can be shown to be uncountable,
which means that it is strictly bigger than any countable set. Complex numbers that are
not algebraic are called transcendental. Familiar examples of transcendental numbers
are π and the base e of the natural logarithms.
Examples
is chosen for α). However, all these powers belong to Q[α], and they can be
computed in the form a + bα using the rule α2 = 1 + α:
α2 = 1+α, α3 = α +α2 = 1+2α, α4 = α +2α2 = 2+3α, α5 = 2α +3α2 = 3+5α,
and so on. Can you spot a pattern?
Similarly, we can compute the negative powers of α. To do this, divide the
equation α2 = 1 + α by α and rearrange to get α−1 = −1 + α. Then iterate:
α−2 = −α−1 + 1 = 2 − α, α−3 = 2α−1 − 1 = −3 + 2α, α−4 = −3α−1 + 2 = 5 − 3α,
and so on. Can you spot another pattern?
a0 + a1 x + · · · + ad−1 xd−1 ,
where d = deg(m(x)) and each ai ∈ Zp . There are p possible values for each ai , and
hence pd elements in K.
Indeed, the additive group (K, +) is isomorphic to Zdp , the d-dimensional vector
space over the field Zp .
Examples
+ 0 1 x 1+x × 0 1 x 1+x
0 0 1 x 1+x 0 0 0 0 0
1 1 0 1+x x 1 0 1 x 1+x
x x 1+x 0 1 x 0 x 1+x 1
1+x 1+x x 1 0 1+x 0 1+x 1 x
x 0 1 2 3 4
m(x) 1 3 2 3 1
Hence m(x) is irreducible in Z5 [x], and Z5 [x]/m(x)Z5 [x] is a field of order 52 = 25.
Its elements have the form a = bx, a, b ∈ Z5 , with addition defined modulo 5,
and multiplication defined modulo 5 using the rule x2 = −1 − x = 4 + 4x. Thus,
for example, we could compute
Theorem 6.3 Let F be a finite field. Then there exist a prime number p and a positive
integer d such that F has a subfield isomorphic to Zp , and F has order pd .
K = Im(f ) ∼
= Z/pZ ∼
= Zp .
62 CHAPTER 6. FIELD EXTENSIONS
Note also that the additive group of F is a vector space of dimension 1 over F ,
where we define scalar multiplication to be the multiplication of F . The rules for
scalar multiplication in a vector space are satisfied because of the associativity and
distributivity of multiplication in F :
Theorem 6.4 Let p be a prime number, and d a positive integer. Then there exists a
field of order pd . Moreover, this field is unique up to isomorphism: if F1 and F2 are
fields of order pd , then F1 ∼
= F2 .
I will omit the proof of this theorem, but you can probably imagine how it goes. To
prove existence, we do a counting argument to show that there is at least one monic
irreducible polynomial in Zp of degree d. (For example, in the case d = 2 there are p2
monic quadratics x2 + ax + b, of which p(p + 1)/2 are products (x + c)(x + d) of two
linears.)
To prove uniqueness, we check that (i) every field of order pd contains an element
whose minimal polynomial in Zp [x] has degree d; and (ii) if m(x), n(x) ∈ Zp [x] are
irreducibles of degree d, then Zp [x]/m(x)Zp [x] ∼
= Zp [x]/n(x)Zp [x].
Example Consider the field Z3 . Of the three elements in Z3 , only two of them, 0 and
1, are squares. Thus x2 + 1 has no root in Z3 , so is irreducible. Let I = (x2 + 1)Z3 [x]
and let K = Z3 [x]/I be the field of order 9. Let us denote the element x + I of K by
α. Then the elements of K are a + bα for a, b ∈ Z3 . The addition and multiplication
tables in K are given by
+ 0 1 2 α 1+α 2+α 2α 1 + 2α 2 + 2α
0 0 1 2 α 1+α 2+α 2α 1 + 2α 2 + 2α
1 1 2 0 1+α 2+α α 1 + 2α 2 + 2α 2α
2 2 0 1 2+α α 1+α 2 + 2α 2α 1 + 2α
α α 1+α 2+α 2α 1 + 2α 2 + 2α 0 1 2
1+α 1+α 2+α α 1 + 2α 2 + 2α 2α 1 2 0
2+α 2+α α 1+α 2 + 2α 2α 1 + 2α 2 0 1
2α 2α 1 + 2α 2 + 2α 0 1 2 α 1+α 2+α
1 + 2α 1 + 2α 2 + 2α 2α 1 2 0 1+α 2+α α
2 + 2α 2 + 2α 2α 1 + 2α 2 0 1 2+α α 1+α
6.3. FINITE FIELDS 63
and
× 0 1 2 α 1+α 2+α 2α 1 + 2α 2 + 2α
0 0 0 0 0 0 0 0 0 0
1 0 1 2 α 1+α 2+α 2α 1 + 2α 2 + 2α
2 0 2 1 2α 2 + 2α 1 + 2α α 2+α 1+α
α 0 α 2α 2 2+α 2 + 2α 1 1+α 1 + 2α
1+α 0 1+α 2 + 2α 2+α 2α 1 1 + 2α 2 α
2+α 0 2+α 1 + 2α 2 + 2α 1 α 1+α 2α 2
2α 0 2α α 1 1 + 2α 1+α 2 2 + 2α 2+α
1 + 2α 0 1 + 2α 2+α 1+α 2 2α 2 + 2α α 1
2 + 2α 0 2 + 2α 1+α 1 + 2α α 2 2+α 1 2α
then Ker(φ1+α ) is precisely the principal ideal (x2 + x + 2)Z3 [x], so φ1+α induces an
isomorphism from F = Z3 [x]/(x2 + x + 2)Z3 [x] to K, defined by a + bβ 7→ (a + b) + bα.
In a similar way, 2 + β is a square root of 2 in F : (2 + β)2 = 1 + β + β 2 = 2, so the
inverse isomorphism K → F is defined by a + bα 7→ (a + 2b) + bβ.
In F ∼ = K, the group U (F ) = F \ {0} has order 9 − 1 = 8. The multiplication
table for K tells us that the element β (= 1 + α) has order 8, so that the group U (F )
consists precisely of the powers of β: F = {1, β, β 2 , β 3 , β 4 , β 5 , β 6 , β 7 }, with β 8 = 1, so
U (F ) is isomorphic to the cyclic group of order 8: (U (F ), ×) ∼ = (Z8 , +). This is not an
accident!
Proof. The group U (F ) is a finite abelian group. There is a classification theorem for
finite abelian groups which says that any such group is isomorphic to a direct product
of cylic groups Zm(1) × Zm(2) × · · · × Zm(k) for some k ≥ 1 and positive integers m(i)
such that m(i + 1) is a muplitple of m(i) for each i.
In particular, if q is a prime number dividing m(1), and k > 1, then there are at
least q 2 elements of order dividing q in this group: (am(1)/q, bm(2)/q, 0, . . . , 0), where
0 ≤ a ≤ q − 1, 0 ≤ b ≤ q − 1.
But if α1 , . . . , αq2 are q-th roots of 1 in F , then xq − 1 has q 2 distinct linear factors
x − αi in F [x], and so is divisible by the degree q 2 polynomial
(x − α1 )(x − α2 ) · · · (x − αq2 ).
But this contradicts the rules for the degree of a product in F [x], so is impossible.
Hence k = 1, and U (F ) ∼ = Zm(1) is a cylic group. Its order is N − 1, since every
element of F except for 0 is a unit in F .
Corollary 6.6 Let F be a finite field of order N = pd . Then there exists an element
α ∈ U (F ) = F \ {0} which has order N − 1 in U (F ).
Definition Let F be a finite field, and α ∈ F \{0}. The order of α in the multiplicative
group U (F ) is called the multiplicative order of α. If the multiplicative order of α is
the order N − 1 of U (F ), then we say that α is a primitive element of F . In this
case, the units of F are all powers of α. Indeed, we can list all the elements of F as
F = {0, 1, α, α2 , . . . , αN −2 }.
Examples
Remark The unit groups of finite fields are widely used in cryptography - particularly
in the construction of error-correcting codes. The most commonly used fields are those
of characteristic 2, that is, fields Z2 [x]/p(x)Z2 [x] for some irreducible p(x) of degree d.
This enables the elements of the field to be stored efficiently on a computer, as binary
strings of length d, and the Hamming metric gives a natural distance function between
field elements. Messages are encoded using a subset of the field, and messages with
errors are corrected to the nearest element of this subset.
All this requires efficient computation in the finite field under consideration. To
achieve this, one needs to find a primitive element y of the field, and set up a ‘discrete
logarithm table’
a 0 1 · · · pd − 2
ya 1 y ··· y −1
66 CHAPTER 6. FIELD EXTENSIONS
To multiply two field elements α, β quickly, one locates them on the second row
of the table, and identifies their logarithms a, b from the first row of the table. (This
means that α = y a and β = y b .) One then adds a + b (modulo pd − 1) and uses the
table to find the antilogarithm y a+b , which is the desired product y a+b = y a y b = αβ.
Similarly, the discrete logarithm table can be used to carry out fast exponentiation
in the finite field: (y a )b = y ab , so to raise an element to its b-th power one finds its
logarithm, multiplies by b modulo pd −1, and then finds the antilogarithm of the result.
6.3. FINITE FIELDS 67
3. Let F be the field Z5 [x]/(x2 + x + 1)Z5 [x]. Express each of the following elements
of F as Z5 -linear combinations of 1, x:
5. In the field F = Z2 [x]/(x4 +x+1)Z2 [x] of order 16, show that x has multiplicative
order 15, and hence find all elements of multiplicative order 3 (expressed as Z2 -
linear combinations of 1, x, x2 , x3 ).
6. In the field F = Z7 [x]/(x2 + x + 3)Z7 [x] of order 49, find the multiplicative order
of x. Find an element of multiplicative order 8 in F .
8. Use the Euclidean Algorithm to find a greatest common divisor of the given
elements in the integral domains indicated
9. Find the minimal polynomials (in Q[x]) of the following complex numbers:
(a) 1 + i
p √
(b) 1 + 2
1 i
(c) √ − √
2 2