Rings and Ideals A First Course in
Rings and Ideals A First Course in
Vll
CONVENTIONS
INTRODUCTORY CONCEPTS
The present chapter sets the stage for much that follows, by reviewing some
of the basic elements of ring theory. It also serves as an appropriate vehicle
for codifying certain notation and technical vocabulary used throughout
the text. With an eye to the beginning student (as well as to minimize a
sense of vagueness), we have also included a number of pertinent examples
of rings. The mathematically mature reader who finds the pace somewhat
tedious may prefer to bypass this section, referring to it for terminology
when necessary.
As a starting point, it would seem appropriate formally to define the
principal object of interest in this book, the notion of a ring.
Definition 1-1. A ring is an ordered triple (R, +,.) consisting of a
non empty set R and two binary operations + and . defined on R such
that
1) (R, +) is a commutative group,
2) (R,·) is a semigroup, and
3) the operation· is distributive (on both sides) over the operation +.
The reader should understand clearly that + and· represent abstract,
unspecified, operations and not ordinary addition and multiplication. For
convenience, however, one invariably refers to the operation + as addition
and to the operation· as multiplication. In the light of this terminology, it
is natural. then to speak of the commutative group (R, +) as the additive
group of the ring and of (R, .) as the multiplicative semigroup of the ring.
By analogy with the integers, the unique identity element for addition
is called the zero element of the ring and is denoted by the usual symbol O.
The unique additive inverse of an element a E R will hereafter be written
as - a. (See Problem 1 for justification of the adjective "unique".)
In order to minimize the use of parentheses in expressions involving
both operations, we shall stipulate that multiplication is to be performed
before addition. Accordingly, the expression a·b + c stands for (a·b) + c
and not for a·(b + c). Because of the general associative law, parentheses
2
can also be omitted when writing out sums and products of more than two
elements_
With these remarks in mind, we can now give a more elaborate definition
of a ring_ A ring (R, +, -) consists of a nonempty set R together with two
binary operations + and - of addition and multiplication on R for which
the following conditions are satisfied:
1) a + b = b + a,
2) (a + b) + c = a + (b + c),
3)
4)
° °
there exists an element in R such that a + = a for every a E R,
for each a E R, there exists an element - a E R such that a + (- a) = 0,
5) (a-b)-c = a-(b-c), and
6) a:(b + c) = a-b + a-c and (b + c)-a = boa + c-a,
where it is understood that a, b, c represent arbitrary elements of R.
A ring (R, +, -) is said to be a finite ring if, naturally enough, the set R
of its elements is a finite set. By placing restrictions on the multiplication
operation, several other specialized types of rings are obtained_
Definition 1-2. 1) A commutative ring is a ring (R, +, -) in which
multiplication is a commutative operation, a- b = b -a for all a, b E R.
(In case a-b = boa for a particular pair a, b, we express this fact by saying
that a and b commute_)
2) A ring with identity is a ring (R, +, -) in which there exists an
identity element for the operation of multiplication, normally represented
by the symbol 1, so that a-I = I-a = a for all a E R.
Given a ring (R, +, -) with identity 1, an element a E R is said to be
invertible, or to be a unit, whenever a possesses a (two-sided) inverse with
respect to multiplication_ The multiplicative inverse of a is unique, when-
ever it exists, and will be denoted by a- l , so that a-a- l = a-loa = 1_ In the
future, the set of all invertible elements of the ring will be designated by the
symbol R*_ It follows easily that the system (R*, -) forms a group, known
as the group of invertible elements_ In this connection, notice that R* is
certainly nonempty, for, if nothing else, 1 and -1 belong to R*_ (One must
not assume, however, that 1 and -1 are necessarily distinct.)
A consideration of several concrete examples will serve to bring these
ideas into focus_
Example 1-1. If Z, Q, R# denote the sets of integers, rational, and real
numbers, respectively, then the systems
(Z, +,-), (Q, +,-), (R#, +,-)
are all examples of rings (here, + and - are taken to be ordinary addition
and multiplication)_ In each of these cases, the ring is commutative and
has the integer 1 for an identity element.
INTRODUCTORY CONCEPTS 3
Example 1-2 Let X be a given set and P(X) be the collection of all subsets
of X. The symmetric difference of two subsets A, B s; X is the set A Ll B,
where
A Ll B = (A - B) u (B - A).
If we define addition and multiplication in P(X) by
A + B = All B, A·B = A II B,
then the system (P(X), +, .) forms a commutative ring with identity. The
empty set 0 serves as the zero element, whereas the multiplicative identity
is X. Furthermore, each set in P(X) is its own additive inverse. It is
interesting to note that if X is nonempty, then neither (P(X), u, II) nor
(P(X), II, u) constitutes a ring.
Example 1-3. Given a ring (R, +, .), we may consider the set Mn(R) of
n x n matrices over R. If In = {1,2, ... , n}, a typical member of Mn(R)
is a function f: In X In --'f R. In practice, one identifies such a function
with its values aij = f(i, j), which are displayed as the n x n rectangular
array
! ~. 11 •••
a n1 ••• ann
~. 1n ) (a ij E R).
Jij =
f if i
1ifi =j
LO =1= j (i, j = 1, 2, ... , n),
the identity matrix can be written concisely as (Jij).
4
Of course, the same congruence class may very well arise from another
integer; any integer at for which [at] = [a] is said to be a representative
of [a J. One final, purely notational, remark: the collection of all congruence
classes of integers modulo n will be designated by Zn.
It can be shown that the congruence classes [0], [1], ... , [n - 1]
exhaust the elements of Zn. Given an arbitrary integer a, the division
algorithm asserts that there exist unique q, r E Z, with 0 ::::;; r < n, such
that a = qn + r. By the definition of congruence, a == r (mod n), or
INTRODUCTOR Y CONCEPTS 5
Hence, a'b' == ab (mod n) and so [a'b '] = [ab], as desired. The proof that
addition is unambiguously defined proceeds similarly.
We omit the detailed verification of the fact that (Zn' +n' ·n) is a com-
mutative ring with identity (traditionally known as the ring of integers
modulo n), remarking only that the various ring axioms hold in Zn simply
because they hold in Z. The distributive law, for instance, follows in Zn
from its validity in Z:
[aL([b] +n [e]) = [aL[b + e] = [a(b + e)]
= [ab + ae] = [ab] +n [ae]
= [aL[b] +n [aln[e].
Notice, too, that the congruence classes [0] and [1] serve as the zero element
and multiplicative identity, respectively, whereas [- a] is the additive
inverse of [a] in Zn. When no confusion is likely, we shall leave off the
brackets from the elements of Zn' thereby making no genuine distinction
6
in this case the term "integral domain" would merely indicate a commutative
ring without zero divisors.
We change direction somewhat to deal with the situation where a subset
of a ring again constitutes a ring. Formally speaking,
Definition 1-4. Let (R, +, .) be a ring and S ~ R be a nonempty subset
of R. If the system (S, +, .) is itself a ring (using the induced operations),
then (S, +, .) is said to be a subring of (R, +, .).
This definition is adequate, but unwieldy, since all the aspects of the
definition of a ring must be checked in determining whether a given subset
is a subring. In seeking a simpler criterion, notice that (S, +, .) is a subring
of (R, +, .) provided that (S, +) is a subgroup of (R, +), (S, .) is a subsemi-
group of (R, .), and the two distributive laws are satisfied in S. But the
distributive and associative laws hold automatically for elements of S as a
consequence of their validity in R. Since these laws are inherited from R,
there is no necessity of requiring them in the definition of a subring.
Taking our cue from these remarks, a subring could just as well be
defined as follows. The system (S, +, .) forms a subring of the ring (R, +, .)
if and only if
1) S is a nonempty subset of R,
2) (S, +) is a subgroup of (R, +), and
3) the set S is closed under multiplication.
To add the final touch, even this definition can be improved upon; for
the reader versed in group theory will recall that (S, +) is a subgroup of
the group (R, +) provided that a - b E S whenever a, b E S. By these
observations we are led to a set of closure conditions which make it some-
what easier to verify that a particular subset is actually a subring.
Theorem 1-3. Let R be a ring and 0 =1= S ~ R. Then, S is a subring
of R if and only if
1) a, b E S imply a - bE S (closure under differences),
2) a, b E Simply ab E S (closure under multiplication).
If S is a subring of the ring R, then the zero element of S is that of R
and, moreover, the additive inverse of an element of the subring S is the
same as its inverse as a member of R. Verification of these assertions is left
as an exercise.
Example 1-6. Every ring R has two obvious subrings, namely, the set {O},
consisting only of the zero element,· and R itself. These two subrings are
usually referred to as the trivial subrings of R; all other subrings (if any
exist) are called nontrivial. We shall use the term proper subring to mean
a subring which is different from R.
INTRODUCTORY CONCEPTS 9
Example 1-7. The set Ze of even integers forms a subring of the ring Z
of integers, for
2n - 2m = 2(n - m) E Ze'
(2n)(2m) = 2(2nm) E Ze'
This example also illustrates a fact worth bearing in mind: in a ring with
identity, a subring need not contain the identity element.
Prior to stating our next theorem, let us define the center of a ring R,
denoted by cent R, to be the set
cent R = {a E Rlar = ra for all r E R}.
Phrased otherwise, cent R consists of those elements which commute with
every member of R. It should be apparent that a ring R is commutative if
and only if cent R = R.
Theorem 1-4. For any ring R, cent R is a subring of R.
Proof To be conscientious about details, first observe that cent R is non-
empty; for, at the very least, the zero element 0 E R. Now pick any two
elements a, b in cent R. By the definition of center, we know that ar = ra
and br = rb for every choice of r E R. Thus, for arbitrary r E R,
(a - b)r = ar - br = ra - rb = r(a - b),
which implies that a - b E cent R. A similar argument affirms that the
product ab also lies in cent R. In the light of Theorem 1-3, these are
sufficient conditions for the center to be a subring of R.
It has already been remarked that, when a ring has an identity, this
need not be true of its subrings. Other interesting situations may arise.
1) Some subring has a multiplicative identity, but the entire ring does not.
2) Both the ring and one of its subrings possess identity elements, but they
are distinct.
In each of the cited cases the identity for the subring is necessarily a divisor
of zero in the larger ring. To justify this claim, let l' =1= 0 denote the
identity element of the subring S; we assume further that l' does not act as
an identity for the whole ring R. Accordingly, there exists some element
a E R for which aI' =1= a. It is clear that
(al')l' = a(l'l') = aI',
or (aI' - all' = O. Since neither aI' - a nor l' is zero, the ring R has
zero divisors, and in particular l' is a zero divisor.
Example 1-8. To present a simple illustration of a ring in which the second
of the aforementioned possibilities occurs, consider the set R = Z x Z,
10
PROBLEMS
1. Verify that the zero element of a ring R is unique, as is the additive inverse of each
element a E R.
2. Let R be an additive commutative group. If the product of every pair of elements
is defined to be zero, show that the resulting system forms a commutative ring
(this is sometimes called the zero ring).
3. Prove that any ring R in which the two operations are equal (that is, a +b = ab
for all a, b E R) must be the trivial ring R = {O}.
4. In a ring R with identity, establish each of the following:
a) the identity element for multiplication is unique,
b) if a E R has a multiplicative inverse, then a-I is unique,
c) if the element a is invertible, then so also is -a,
d) no divisor of zero can possess a multiplicative inverse in R.
5. a) If the set X contains more than one element, prove that every nonempty proper
subset of X is a zero divisor in the riIig P(X).
b) Show that, if n > 1, the matrix ring M.(R) has zero divisors even though the
ring R may not.
6. Suppose that R is a ring with identity 1 and having no divisors of zero. For a, b E R,
verify that
a) ab = 1 if and only if ba = 1,
b) if a2 = 1, then either a = 1 or a = -1.
7. Let a, b be two elements of the ring R. If n E Z + and a and b commute, derive the
binomial expansion
where
(kn) = k!(n n!
- k)!
is the usual binomial coefficient.
14
11. A Boolean ring is a ring with identity every element of which is idempotent. Prove
that any Boolean ring R is commutative. [H int: First show that a = - a for
every a E R.]
12. Suppose the ring R contains an element a such that (1) a is idempotent and (2) a
is not a zero divisor of R. Deduce that a serves as a multiplicative identity for R.
13. Let S be a nonempty subset of the finite ring R. Prove that S is a subring of R
if and only if S is closed under both the operations of addition and multiplication.
14. Assume that R is a ring and a E R. If C(a) denotes the set of all elements which
commute with a,
C(a) = {r E Rlar = raJ,
show that C(a) is a subring of R. Also, verify the equality cent R = naeR C(a).
15. Given a ring R, prove that
a) if S; is an arbitrary (indexed) collection of subrings of R, then their intersection
n S; is also a subring of R;
b) for a nonempty subset T of R, the set
I
17. Let R be an arbitrary ring and n E Z+. If the set Sn is defined by
Sn = {a E Rlnka = 0 for some k > OJ,
determine whether Sn is a subring of R.
Show that the resulting system (R, +,0) forms a ring. At the same time determine
the invertible elements of R.
24. Let (G, .) be a finite group (written multiplicatively), say with elements Xl' x 2' ••• , xn ,
and let R be an arbitrary ring. Consider the set R(G) of all formal sums
(ri E R).
Two such expressions are regarded as equal if they have the same coefficients.
Addition and multiplication can be defined in R(G) by taking
n n n
I rixi + i=l
i=l
I SiXi = I (ri + S;)Xi
i=l
and
(The meaning of the last-written sum is that the summation is to be extended over
all subscripts j and k for which XjXk = Xi') Prove that, with respect to these
operations, R(G) constitutes a ring, the so-called group ring of Gover R.
TWO
Example 2-1. For each integer a E Z, let (a) represent the set consisting
of all integral multiples of a; that is,
(a) = {naln E Z}.
The following relations confirm (a) to be an ideal of the ring of integers:
na - rna = (n - rn)a,
rn(na) = (rnn)a, "n, nEZ.
In particular, since (2) = Ze' the ring of even integers forms an ideal of Z.
Notice, too, that (0) = {O} and (1) = Z •."
Example 2-2. Another illustration is furnished by map (X, R), the ring of
mappings from the set X into the ring R (see Example 1---4). For a fixed
element x E X, we denote by Ix the set of all mappings which take on the
value 0 at x:
Ix = {f E map (X, R)lf(x) = O}.
Now, choose f, g E Ix and hE map (X, R). From the definition of the ring
operations in map(X, R),
(f -g)(x) = f(x)-g(x) = 0-0 = 0,
while
(fh)(x) = f(x)h(x) = Oh(x) = 0,
Proof. We give the proof for the case in which the ideals are two-sided.
First, observe that the intersection n Ii is nonempty, for each of the ideals
Ii must contain the zero element of the ring. Suppose that the elements
a, bEn Ii and r E R. Then a and b are members of Ii' where i varies over
the indexing set. Inasmuch as Ii is assumed to be an ideal of R, it follows
that a - b, ar and ra all lie in the set Ii. But this is true for every value of
IDEALS AND THEIR OPERATIONS 19
IJ = {L aibilai E I; bi E J}.
finite
where the ai and a; are in I, and the bi and b; are in J. From this we obtain
x - Y = a 1b 1 + ... + anbn + (-a'l)b'l + ... + (-a~)b~,
Now, the elements -a; and rai necessarily lie in I, so that x - y and
rx E IJ; likewise, xr E IJ, making IJ an ideal of R. In point of fact, IJ is
just the ideal generated by the set of all products ab, a E I, b E J.
There is no difficulty in extending the above remarks to any finite
number of ideals 11 , 12 , ••• , In of the ring R. A moment's thought shows
that the product 11 I 2 .•• In is the ideal consisting of finite sums of terms of
the form a 1 a 2 ••• an' with a i in Ii. (It is perhaps appropriate to point out
that, because of the associative law for multiplication in R, the notation
11 I 2 ••• In is unambiguous.) A special case immediately presents itself:
namely, the situation where all the ideals are alike, say equal to the ideal
I. Here, we see that JR is the set of finite sums of products of n elements
from I:
IDEALS AND THEIR OPERATIONS 23
Theorem 2-5. The following relations hold for ideals in a ring R (capital
letters indicate ideals of R):
1) (n Ii) :rJ = n (Ii :rJ),
2) I :r L J i = n (I :r J;),
3) I:r (JK) = (I :r K) :r J.
24
For future use, we shall label the set of all homomorphisms from the
ring R into the ring R' by the symbol hom(R, R'). In the event that R = R',
the simpler notation hom R will be used in place of hom(R, R). (Some
authors prefer to write end R, for endomorphism, in place of hom R; both
notations have a certain suggestive power and it reduces to a matter of
personal preference.)
A knowledge of a few simple-minded examples will help to fix ideas.
Example 2-4. Let Rand R' be arbitrary rings and f: R ~ R' be the function
which sends each element of R to the zero element of R'. Then,
f(a + b) = 0 = 0 + 0 = f(a) + f(b),
(a, bE R),
f(ab) = 0 = 00 = f(a) f(b)
so thatfis a homomorphic mapping. This particular mapping, the so-called
trivial homomorphism, is the only constant function which satisfies Definition
2-7.
Example 2-5. Consider the ring Z of integers and the ring Zn of integers
modulo n. Definef: Z ~ Zn by takingf(a) = [a]; that is, map each integer
into the congruence class containing it. That f is a homomorphism follows
directly from the definition of the operations in Zn:
f(a + b) = [a + b] = [a] +n [b] = f(a) +nf(b),
f(ab) = Cab] = CaLeb] = f(a)·nf(b).
Example 2-6. In the ring map(X, R), define 'a to be the function which
assigns to each f E map (X, R) its value at a fixed element a EX; in other
words, 'a(f) = f(a). Then 'a is a homomorphism from map (X, R) into R,
known as the evaluation homomorphism at a. We need only observe that
'a(f+ g) = (f + g)(a) = f(a) + g(a) = 'a(f) + 'a (g),
'a(fg) = (fg)(a) = f(a)g(a) = 'a (f)'a(g)·
We now list some of the structural features preserved under homo-
morphisms.
Theorem 2-7. Letfbe a homomorphism from the ring R into the ring
R'. Then the following hold:
1) f(O) = 0,
2) f( -a) = -f(a) for all a E R.
If, in addition, Rand R' are both rings with identity and f(R) = R',
then
3) f(1) = 1,
4) f(a -1) = f(a) - 1 for each invertible element a E R.
IDEALS AND THEIR OPERA TIONS 27
f(b) must be members of S'. Since S' is a subring of R', it follows at once
that
f(a - b) = f(a) - f(b) E S'
and
f(ab) = f(a)f(b) E S'.
This means that a - band ab lie inf- 1(S'), from which we conclude that
f- 1(S') forms a subring of R.
Left unresolved is the matter of replacing the term "subring" in Theorem
2-8 by "ideal". It is not difficult to show that part (2) of the theorem
remains true under such a substitution. More precisely: if l' is an ideal of
R', then the subring f- 1(1') is an ideal of R. For instance, suppose that
a Ef-1(1'), so thatf(a) E 1', and let r be an arbitrary element of R. Then,
f(ra) = f(r)f(a) E 1'; in other words, the product ra is inf-1(1'). Likewise,
ar E f- 1(1'), which helps to make f-1(1') an ideal of R.
Without further restriction, it cannot be inferred that the image f(l)
will be an ideal of R', whenever I is an ideal of R. One would need to know
that r'f(a) Ef(l) for all r' E R' and a E I. In general, there is no way of
replacing r' by some f(r) in order to exploit the fact that I is an ideal. The
answer is obvious: just take f to be an onto mapping.
Summarizing these remarks, we may now state:
Corollary. 1) For each ideal l' of R', the subring f- 1(1') is an ideal
of R.
2) If f(R) = R', then for each ideal I of R, the subring f(I) is an ideal
ofR'.
To go still further, we need to introduce a new idea.
Definition 2-8. Let f be a homomorphism from the ring R into the
ring R'. The kernel off, denoted by ker f, consists of those elements in
R which are mapped by f onto the zero element of the ring R':
ker f = {a E R!f(a) = O}.
Theorem 2-7 indicates that ker f is a nonempty subset of R, since, if
nothing else, 0 E ker f. Except for the case of the trivial homomorphism,
the kernel will always turn out to be a proper subset of R.
As one might suspect, the kernel of a ring homomorphism forms an
ideal.
Theorem 2-9. The kernel ker f of a homomorphism f from a ring R
into a ring R' is an ideal of R.
"
Proof. We already know that the trivial subring {O} forms an ideal of R'.
I'
"I Since ker f = f- 1(0), the conclusion follows from the last corollary.
IDEALS AND THEIR OPERATIONS 29
any ring R with identity which is of characteristic zero will contain a subring
isomorphic to the integers; more specifically, Z ~ Zl, where 1 is the
identity of R.
Suppose that f is a homomorphism from the ring R onto the ring R'.
We have already observed that each ideal 1 of the ring R determines an
ideal f(1) of the ring R'. It goes without saying that ring theory would be
considerably simplified if the ideals of R were in a one-to-one correspondence
with those of R' in this manner. Unfortunately, this need not be the case.
The difficulty is reflected in the fact that if 1 and J are two ideals of R
with 1 ~ J ~ 1 + ker f, then f(1) = f(J). The quickest way to see this is
to notice that
from which we conclude that all the inclusions are actually equalities. In
brief, distinct ideals of R may have the same image in R'.
This disconcerting situation could be remedied by either demanding
that ker f = {o} or else narrowing our view to consider only ideals 1 with
ker f ~ 1. In either event, it follows that 1 ~ J ~ 1 + kerf = 1 and, in
consequence, 1 = J. The first of the restrictions just cited has the effect of
making the function f one-to-one, in which case Rand R' are isomorphic
rings (and it then comes as no surprise to find their ideals in one-to-one
correspondence). The second possibility is the subject of our next theorem.
We turn aside briefly to establish a preliminary lemma which will
provide the key to later success.
Lemma. Letfbe a homomorphism from the ring R onto the ring R'.
If 1 is any ideal of R such that kerf ~ 1, then 1 = f-l(J(1)).
Proof Suppose that the element a E jl(t{1)), so that j{a) E j{1). Then
f(a) = f(r) for some choice of r in 1. As a result, we will have f(a - r) = 0,
or, what amounts to the same thing, a - r E ker f ~ 1. This implies that
a E 1, yielding the inclusion f-l(J(l}) ~ 1. Since the reverse inclusion
always holds, the desired equality follows.
Here now is one of the main results of this section.
The crucial third equality is justified by the fact that u E cent R, hence, -
commutes with b.
As regards the uniqueness assertion, let us assume that there is another
homomorphic extension off to the set R; call it h. Sincefand h must agree
on I and, more specifically, at the element u, h(u) = f(u) = 1. With this
in mind, it follows that
h(a) = h(a)h(u) = h(au) = f(au) = g(a)
for all a E R and so hand g are the same function. Hence, there is one and
only one way of extendingfhomomorphically from the ideal I to the whole
ring R.
Before closing the present chapter, there is another type of direct sum
which deserves mention. To this purpose, let R I , R 2 , ... ,Rn be a finite
number of rings (not necessarily subrings of a common ring) and consider
their Cartesian product R = X Ri consisting of all ordered n-tuples
(aI' a2 , ... ,an)' with ai E R i . One can easily convert R into a ring by
performing the ring operations componentwise; in other words, if
(aI' a2 , •.. , an) and (b l , b2 , ••• , bn) are two elements of R, simply define
(aI' a 2 , •.• ,an) + (b l , b2 , .•• , bn) = (a l + b l , a2 + b2 , .•• ,an + bn)
and
(aI' a2 , ••• , an)(b l , b2 , .•• , bn) = (alb l , a2 b2 , .•• , anbn)·
The ring so obtained is called the external direct sum of R I , R 2 , ••. , Rn
and is conveniently written R = RI R2 +
Rn. (Let us caution + ... +
that the notation is not standard in this matter.) In brief, the situation is
this: An external direct sum is a new ring constructed from a given set of
rings, and an internal direct sum is a representation of a given ring as a sum
of certain of its ideals. The connection between these two types of direct
sums will be made clear in the next paragraph.
If R is the external direct sum of the rings Ri (i = 1,2, ... , n), then the
individual Ri need not be subrings, or even subsets, of R. However, there is
an ideal of R which is the isomorphic image of R i • A straightforward
calculation will convince the reader that the set
Ii = {(O, ... , 0, ai' 0, ... , O)iai E R i }
34
(that is, the set consisting of all n-tuples with zeroes in all places but the
ith) forms an ideal of R naturally isomorphic to Ri under the mapping
which sends (0, ... , 0, ai' 0, ... , 0) to the element ai. Since
(ai' a2 , .•• ,an) = (ai' 0, 0, ... ,0) + (0, a 2 , 0, ... ,0) + ... + (0, 0, ... , 0, an),
it should also be clear that every member of R is uniquely representable as
a sum of elements from the ideals Ii. Taking note of Theorem 2-4, this
means that R is the internal direct sum of the ideals Ii and so
Rl + R2 + ... + Rn = Ii EEl 12 EEl ... EEl In (Ri ~ Ii)·
In summary, the external direct sum R of the rings R 1 , R 2 , ••• , Rn is also
the internal direct sum of the ideals Ii' 12 , ••• , In and, for each i, Ri and
Ii are isomorphic.
In view of the isomorphism just explained, we shall henceforth refer to
the ring R as being a direct sum, not qualifying it with the adjective
"internal" or "external", and rely exclusively on the EEl-notation. The term
"internal" merely reflects the fact that the individual summands, and not
isomorphic copies of them, lie in R.
We take this opportunity to introduce the simple, but nonetheless useful,
notion of a direct summand of a ring. In formal terms, an ideal I of the
ring R is said to be a direct summand of R if there exists another ideal J
of R such that R = I EEl J. For future use, let us note that should the ideal
I happen to have an identity element, say the element e E I, then it will
automatically be a direct summand of R. The argument proceeds as follows.
For any choice of r E R, the product re E I. The assumption that e serves
as an identity for I then ensures that e(re) = reo At the same time (and for
thesamereasons),(er)e = er. Combining these pieces, wegetre = ere = er,
which makes it plain that the element e lies in the center of R. This is the
key point in showing that the set J = {r - relr E R} forms an ideal of R;
the details are left to the reader. We contend that the ring R is actually the
direct sum of I and J. Certainly, each element r of R may be written as
r = re + (r - re), where re E I and r - re E J. Since I II J = {O}, this is
the only way r can be expressed as a sum of elements of I and J. (A moment's
thought shows that if a E I II J, say a = r - re, then a = ae = (r - re)e =
r(e - e2 ) = 0.) It is also true that the ideal I = eR = Re, but we did not
need this fact here.
As a further application of the idea of a direct summand, let us record
Theorem 2-14. If the ring R is a direct summand in every extension
ring containing it as an ideal, then R has an identity.
Proof To set this result in evidence, we first imbed R in the extension ring
R' = R x Z in the standard way (see Theorem 2-12). Then, R ~ R x {O},
where, as is easily verified, R x {O} constitutes an ideal of R'. We may
PROBLEMS 35
therefore regard R as being an ideal of the ring R'. Our hypothesis now
comes into play and asserts that R' = R EB J for a suitable ideal J of R'.
It is thus possible to choose an element (e, n) in J so that (0, -1) = (r,O) +
(e, n), for some r E R. The last-written equation tells us that e = - rand
n = -1; what is important is the resulting conclusion that (e, -1) E J.
for arbitrary r E R, the product (r, O)(e, -1) = (re - r, 0) will consequently
be in both Rand J (each being an ideal of R'). The fact that R n J = {O}
forces (re - r, 0) = (0, 0); hence, re = r. In a like fashion, we obtain
er = r, proving that R admits the element e as an identity.
PROBLEMS
1. If I is a right ideal and J a left ideal of the ring R such that I n J = {O}, prove
that ab = 0 for all a E I, b E J.
2. Given an ideal I of the ring R, define the set C(I) by
C(I) = {r E Rlra - ar E I for all a E R}.
Verify that C(I) forms a subring of R.
3. a) Show by example that if I and J are both ideals of the ring R, then I u J need
not be an ideal of R.
b) If {Ii} (i = 1,2, ... ) is a collection of ideals of the ring R such that! 1 ~ I 2 ~ ...
~ In ~ "', prove that u Ii is also an ideal of R.
4. Consider the ring Mn(R) ofn x n matrices over R, a ring with identity. A square
matrix (au) is said to be upper triangular if au = 0 for i > j and strictly upper
triangular if aij = 0 for i ~ j. Let T,,(R) and T!(R) denote the sets of all upper
triangular and strictly upper triangular matrices in Mn(R), respectively. Prove
each of the following:
a) T,,(R) and T!(R) are both subrings of Mn(R).
b) T!(R) is an ideal of the ring T,,(R).
c) A matrix (au) E T,,(R) is invertible in T,,(R) if and only if aii is invertible in R
for i = 1,2, ... , n. [Hint: Induct on the order n.]
d) Any matrix (a ij ) E T!(R) is nilpotent; in particular, (ai)n = O.
5. Let I be an ideal of R, a commutative ring with identity. For an element a E R,
the ideal generated by the set I u {a} is denoted by (I, a). Assuming that a ¢ I,
show that
(I, a) = {i + rali E I, r E R}.
6. In the ring Z of integers consider the principal ideals (n) and (m) generated by
the integers nand m. Using the notation of the previous problem, verify that
7. Suppose that I is a left ideal and J a right ideal of the ring R. Consider the set
where 1: represents a finite sum of one or more terms. Establish that lJ is a two-
sided ideal of R and, whenever I and J are themselves two-sided, that lJ s;;; I II J.
9. Let 11,1 2, ... , In be ideals of the ring R with R = 11 + 12 + ... + In. Show
that this sum is direct if and only if a 1 + a2 + ... + an = 0, with ai E Ii' implies
that each ai = O.
10. If P(X) is the ring of all subsets of a given set X, prove that
a) the collection of all finite subsets of X forms an ideal of P(X);
b) for each subset Y S;;; X, P(Y) and P(X - Y) are both principal ideals of P(X),
with P(X) = P(Y) Et> P(X - Y).
11. Suppose that R is a commutative ring with identity and that the element a E R
is an idempotent different from 0 or 1. Prove that R is the direct sum of the principal
ideals (a) and (1 - a).
15. Let I, J and K be ideals of R, a commutative ring with identity. Prove the following
assertions:
PR08LEMS 37
16. If I is a right ideal of R, a ring with identity, show that I:/R = {a E RIRa ~ I}
is the largest two-sided ideal of R contained in I.
17. Given that f is a homomorphism from the ring R onto the ring R', prove that
a) f(cent R) ~ cent R'.
b) If R is a principal ideal ring, then the same is true of R'. [Hint: For any a E R,
f(a)) = (J(a)).]
c) If the element a E R is nilpotent, then its image f(a) is nilpotent in R'.
18. Let R be a ring with identity. For each invertible element a E R, show that the
function!,.: R --+ R defined by fa (x) = axa- 1 is an automorphism of R.
19. Letfbe a homomorphism from the ring R into itself and S be the set of elements
that are left fixed by f; in symbols,
S = {a E Rlf(a) = a}.
Establish that S forms a subring of R.
20. If f is a homomorphism from the ring R into the ring R', where R has positive
characteristic, verify that char f(R) :::;; char R.
21. Letfbe a homomorphism from the commutative ring R onto the ring R'. If I and
J are ideals of R, verify each of the following:
a) f(I + J) = f(1) + f(1);
b) f(I1) = f(I)f(1);
c) f(I n 1) ~ f(I) n f(1), with equality if either I ;2 kerf or J ;2 kerf;
d) f(I: 1) ~ f(1) :f(1), with equality if I ;2 kerf
22. Show that the relation R ~ R' is an equivalence relation on any set of rings.
23. Let R be an arbitrary ring. For each fixed element a E R, define the left-multiplica-
tion function T.: R --+ R by taking T.(x) = ax. If TR denotes the set of all such
functions, prove the following:
a) T. is a (group) homomorphism of the additive group of R into itself;
b) TR forms a ring, where multiplication is taken to be functional composition;
c) the mappingf(a) = T. determines a homomorphism of R onto the ring TR ;
d) the kernel offis the ideal ann,R;
e) if for each 0 =1= a E R, there exists some bE R such that ab =1= 0, then R ~ TR •
(In particular, part(e) holds whenever R has an identity element.)
24. Let R be an arbitrary ring and R x Z be the extension ring constructed in Theorem
2-12. Establish that
a) Rx {O} is an ideal of RxZ;
b) Z ~ {O}xZ;
c) if a is an idempotent element of R, then the pair ( - a, 1) is idempotent in R x Z,
while (a, 0) is a zero divisor.
38
Given an ideal I of the ring R, let us employ the symbol R/I to denote
the collection of all cosets of I in R; that is,
R/I = {a + Iia E R}.
The set R/I can be endowed with the structure of a ring in a natural way;
all we need do is define addition and multiplication as follows:
(a + I) + (b + I) = (a + b) + I,
(a + I)(b + I) = ab + 1.
One is faced with the usual problem of showing that these operations are
actually well-defined, so that the sum and product of the two co sets a + I
and b + I do not depend on their particular representatives a and b. To
this end, suppose that
a+I=a'+I and b +I= b' + I.
Then a - a' = il and b - b' = iz for some ii' iz E 1. From this we conclude
that
(a + b) - (a' + b') = (a - a') + (b - b')
= il + iz E I,
which, by Theorem 3-2, indicates that (a + b) + I = (a' + b') + I. The
net result is that (a + I) + (b + I) = (a' + I) + (b' + I). With regard
to the multiplication of cosets, we observe that
ab - a'b' = a(b - b') + (a - a')b'
= aiz + i1b' E I,
since both the products aiz and i l b' must be in 1. The implication, of course,
is that ab + 1 = a'b' + I; hence, our definition of multiplication in RjI is
meaningful.
The verification that R/I, under the operations defined above, forms a
ring is easy and the details are left to the reader. To assure completeness,
we simply state
Theorem 3-3. If I is an ideal ofthe ring R, then R/I is also a ring, known
as the quotient ring (or factor ring) of R by 1.
In Theorem 2-9 we saw that certain ideals occur as kernels of homo-
morphisms. Let us now demonstrate that every ideal does indeed arise in
this manner.
Theorem 3-4. Let I be an ideal of the ring R. Then the mapping
nat I : R -+ R/I defined by natI(a) = a +I
is a homomorphism of R onto the quotient ring R/I; the kernel of natI
is precisely the set I.
THE CLASSICAL ISOMORPHISM THEOREMS 41
Proof The fact that nat[ is a homomorphism follows directly from the
manner in which the operations are defined in the quotient ring:
nat[(a + b) = a + b + J = (a + J) + (b + J)
= nat[(a) + nat[(b);
nat[(ab) = ab + J = (a + J)(b + J)
= nat[(a) nat[(b).
That nat[ carries R onto R/J is all but obvious; indeed, every element of
R/J is a coset a + J, with a E R, and so by definition nat[(a) = a + 1.
Inasmuch as the coset J = 0 + J serves as the zero element for the
ring R/J, we necessarily have
Proof. To start, we define a functionJ; R/1 -+ R', called the induced mapping,
by taking
J(a + 1) = f(a) (a E R).
The first question to be raised is whether or not J is actually well-defined.
That is to say, we must establish that this function has values which depend
only upon the co sets of 1 and in no way on their particular representatives.
In order to see this, let us assume a + 1 = b + 1. Then a - b E 1 £ ker f.
This means that
f(a) = f(a - b + b) = f(a - b) + f(b) = f(b)
and, by the manner in which J was defined, that J(a + 1) = J(b + 1).
Hence, the functionJis constant on the cosets of 1, as we wished to demon-
strate.
THE CLASSICAL ISOMORPHISM THEOREMS 43
whence the equality f = J natI . It only remains to show that this fac-
0
torization is unique. Suppose also thatf = go natI for some other function
g: R/f --+ R'. But then
na + f) = f(a) = (g 0 natI )(a) = g(a + f)
for all a in R, and so g = J. The induced mappingJis thus the only function
from the quotient ring R/f into R' satisfying the equation f = J natI . 0
another, the composition of mappings along these paths produces the same
function.)
A rather simple observation, with far-reaching implications, is that
whenever J = ker f, so that both the Factorization Theorem and its Corol-
lary are applicable, f induces a mapping J under which RIJ and R' are
isomorphic rings. We summarize all this in the following theorem, a result
which will be invoked on many occasions in the sequel.
Theorem 3-7. (Fundamental Homomorphism Theorem). If f is a
homomorphism from the ring R onto the ring R', then Riker f ~ R'.
Theorem 3-7 states that the images of R under homomorphisms can
be duplicated (up to isomorphism) by quotient rings of R ; to put it another
way, every homomorphism of R is "essentially" a natural mapping. Thus,
the problem of determination of all homomorphic images of a ring has
been reduced to the determination of its quotient rings.
Let us use Theorem 3-7 to prove that any homomorphism onto the
ring of integers is uniquely determined by its kernel. As the starting point
of our endeavor, we establish a lemma which is of independent interest.
Lemma. The only nontrivial homomorphism from the ring Z of
integers into itself is the identity map iz .
Proof Because each positive integer n may be written as n = 1 + 1 + ...
+ 1 (n summands), the operation-preserving nature of f implies that
f(n) = nf(l). On the other hand, if n is an arbitrary negative integer, then
-n E Z+ and so
f(n) = f( -( -n)) = -f( -n) = -( -n)f(l) = nf(l).
Plainly,f(O) = 0 = Of(l). The upshot is thatf(n) = nf(l) for every n in Z.
Because f is not identically zero, we must have f(l) = 1; to see that this
is so, simply apply the cancellation law to the relation f(m) = f(m1) =
f(m)f(l), where f(m) =1= o. One finds in this way that f(n) = n = iz(n) for
all nEZ, making f the identity map on Z.
Corollary. There is at most one homomorphism under which an
arbitrary ring R is isomorphic to the ring z.
Proof Suppose that the rings Rand Z are isomorphic under two functions
f, g: R -4 Z. Then the composition .r
g - 1 is a homomorphic mapping
0
from the ring Z onto itself. Knowing this, the lemma just proved implies
thatfo g-l = i z , or f = g.
We now have the necessary information to prove the following result.
Theorem 3-8. Any homomorphism from an arbitrary ring R onto the
ring Z of integers is uniquely determined by its kernel.
THE CLASSICAL ISOMORPHISM THEOREMS 45
Proof Let f and g be two homomorphisms from the ring R onto Z with
the property that ker f = ker g. Our aim, of course, is to show that f and g
must be the same function. Now, by Theorem 3-7, the quotient rings
Riker f and Riker g are both isomorphic to the ring of integers via the
induced mappings J and g, respectively. The assumption that f and g have
a common kernel, when combined with the preceding corollary, forces
J = g. It follows at once from the factorizations
f = Jo nat kerf , g = go nat kerg
that the functionsfand g are themselves identical.
The next two theorems are somewhat deeper results than usual and
require the full force of our accumulated machinery. They comprise what
are often called the First and Second Isomorphism Theorems, and have
important applications in the sequel. (The reader is cautioned that there
seems to be no universally accepted numbering for these theorems.)
Theorem 3-9. Let f be a homomorphism of the ring R onto the ring
R' and let I be an ideal of Rlf ker f ~ I, then RII ~ R' I f(1).
Proof Before becoming involved in the details of the proof, let us remark
that 'the corollary to Theorem 2-8 implies that f(1) is an ideal of the ring
R'; thus, it is meaningful to speak of the quotient ring R'I f(1).
Let us now define the function g: R -+ R'I f(l) by g = natf(l) f, where
0
nat1(1): R' -+ R'I f(1) is the usual natural mapping. Thus, g merely assigns
to each element a E R the cosetf(a) + f(1) in R'If(1). Since the functions
f and natf(l) are both onto homomorphisms, their composition carries R
homomorphically onto the quotient ring R' I f(1).
The crux of the argument is to show that ker g = I, for then the desired
conclusion would be an immediate consequence of the Fundamental
Homomorphism Theorem. Since the zero element of R' I f(l) is just the
coset f(1), the kernel of g consists of those members of R which are mapped
by g onto f(1):
ker g = {a E Rlg(a) = f(1)}
= {a E R f(a) + f(l) = f(1)}
= {a E Rlf(a) Ef(l)} = f-1(1(I)).
RjJ
1 ----+
1~
(R/J)/(1jJ)
natll }
By virtue of our assumptions, there exists a (necessarily unique) isomorphism
g: R/I ~ (RjJ)/(IjJ) such that
g natl 0 = nat llJ 0 natJ .
Let us now take up the second of our general isomorphism theorems.
Theorem 3-10. If I and J are ideals of the ring R, then
1/(1 n J) ~ (1 + J)/J.
Proof Reasoning as in Theorem 3-9, we seek a homomorphism f from 1
(regarded a,<; a ring) onto the quotient ring (1 + J)jJ such that ker f = 1 n J.
Our candidate for this functionfis defined by declaring thatf(a) = a + J,
a E 1. A trivial, but useful, observation is that 1 S; 1 + J, whence f can
be obtained by composing the injection map il : 1 ~ I + J with the natural
mapping nat J : 1 + J ~ (1 + J)jJ. To be quite explicit, f = natJ i l or, 0
in diagrammatic language,
THE CLASSICAL ISOMORPHISM THEOREMS 47
Since (N 1 + N 2)/N land N 1 are both nil, it follows from the previous lemma
that N 1 + N 2 is necessarily a nil ideal. Similar reasoning applies to the
nilpotent case.
Corollary. The sum of any finite number of nil (nilpotent) ideals of the
ring R is again nil (nilpotent).
Having completed the necessary preliminaries, let us now establish
Theorem 3-11. The sum L Ni of all the nil ideals Ni of the ring R is a
nil ideal.
Proof If the element a E L N i , then, by definition, a lies in some finite sum
of nil ideals of R; say, aEN l + N2 + ... + N n, where each Nk is nil.
By virtue of the last corollary, the sum Nl + N2 + ... + N n must be a
nil ideal; hence, the element a is nilpotent. This argument shows that L Ni
is a nil ideal.
It is possible to deduce somewhat more, namely,
Corollary. The sum of all the nilpotent ideals of the ring R is a nil
ideal.
Proof Since each nilpotent ideal is a nil ideal, the sum N of all nilpotent
ideals of R is contained in L N i , the sum of all nil ideals. But L Ni is itself
a nil ideal, making N nil.
Example 3-2. For examples of nilpotent ideals, let us turn to the rings
Zpn, where p is a fixed prime and n > 1. By virtue of the remarks on page
42, Zpn has exactly one ideal for each positive divisor of pn and no other
ideals; these are simply the principal ideals (pk) = pkZpn (0 ~ k ~ n). For
o < k ~ n, we have
The reader will find that the zero element of this ring is just the sequence
formed by the zero elements of the various Zpn and the negative of {an}
is {-an}. Now, consider the set R of all sequences in S which become zero
PROBLEMS 49
after a certain, but not fixed, point. One may easily check that R constitutes
a subring of the ring S (in fact, R is not only a subring, but actually an ideal
of S). It is in the ring R that we propose to construct our example of a
non-nilpotent nil ideal.
Let us denote by I the set of sequences in R whose nth term belongs to
the principal ideal in Z pn generated by p; in other words, the sequence
a E I if and only if it is of the form
a = (prl' pr2' ... , prn' 0, 0, .... )
A routine calculation confirms that I is an ideal of R. Since each term of
a is nilpotent in the appropriate ring, it follows that
an = °= (0, 0, 0, ...)
for n large enough, making I a nil ideal. (This also depends on the fact
that a has only a finite number of nonzero terms.)
At the present stage, it is still conceivable that I might be a nilpotent
ideal of R. However, we can show that for each positive integer n there
exist elements (sequences) a E I for which an =1= 0. For instance, define
°
a = {a k} by taking ak = P if k = 1, 2, ... ,n + 1 and ak = if k > n + 1;
that is,
a = (p, ... , p, p, 0, ... ) with n + 1 p's.
One then obtains
an = (0, ... , 0, pn, 0, ... ),
where all the terms are zero except the (n + 1)st, which is pn. Since pn is a
nonzero element of the ring Z pn + 1, the sequence an =1= 0, implying that
1" =1= {OJ. As this argument holds for any n E Z+, the ideal I cannot be
nilpotent.
We shall return to these ideas at the appropriate place in the sequel,
at which time their importance will become clear.
PROBLEMS
a) [R, R] £; kerf;
b) f = Jo nat[R RI' whereJis the induced mapping;
c) if ker f £; [R, R], then R/[R, R] ~ R'/[R', R'J.
12. a) Suppose that 11 and 12 are ideals of the ring R for which R = 11 EB 12. Prove
that R/I1 ~ 12, and R/I2 ~ 1 1.
b) Let R be the direct sum of the rings Ri (i = 1,2, ... , n). If Ii is an ideal of Ri
and I = 11 EB 12 EB ... EB In' show that
[Hint: Find the kernel of the homomorphismJ: R -> 1: EB (RjIJ that sends
a = (a 1,a2, ... ,an) tof(a) = (a 1 + I 1,a2 + 12, ... ,an + In)·]
13. For a proof of Theorem 3-9 that does not depend on the Fundamental Homo-
morphism Theorem, define the function h: R/I -> R'/f(I) by taking h(a + I) =
f(a) + f(I).
a) Show that h is a well-defined isomorphism onto R'/f(I); hence, R/I ~ R'/f(I).
b) Establish that h is the unique mapping that makes the diagram below
commutative:
R L R ' =f(R)
natil 1natf(Il
R/I h R'/ f(I)
verify that
a) I forms an ideal of R'. [Hint: R is an ideal of R'.]
b) R'/I is a ring with identity which has no divisors of zero.
c) R'/I contains a subring isomorphic to R. [Hint: Utilize Problem 9.]
FOUR
Example 4-1. Here are some of the more standard illustrations of fields:
the set Q of all rational numbers, the set F = {a + bJ2la, b E Q}, and
the set R # of all real numbers. In each case the operations are ordinary
addition and multiplication.
52
INTEGRAL DOMAINS AND FIELDS 53
This shows that each nonzero member of C has an inverse under multi-
plication, thereby proving the system C to be a field.
It is worth pointing out that the field C contains a subring isomorphic
to the field of real numbers. For, if
R# x {O} = {(a, O)la E R#},
it follows that R# ~ R# X {O} via the mapping f defined by f(a) = (a, 0).
Inasmuch as the distinction between these systems is only one of notation,
we customarily identify the real number a with the corresponding ordered
pair (a, 0); in this sense, R # may be regarded as a subring of C.
Now, the definition of the operations in C enables us to express an
arbitrary element (a, b) E C as
(a, b) = (a, 0) '+ (b,O)(O, 1),
where the pair (0, 1) is such that (0, 1)2 = (0, 1)(0, 1) = (-1,0). Introducing
the symbol i as an abbreviation for (0, 1), we have
(a, b) = (a, 0) + (b,O)i.
Finally, if it is agreed to replace pairs of the form (a, 0) by the first com-
ponent a (this is justified by the preceding paragraph), the displayed
representation becomes
(a, b) = a + bi, with
In other words, the field C as defined initially is nothing more than a
disguised version of the familiar complex number system.
54
The collection of all equivalence classes a/s relative to '" will be denoted
by Qc)(R):
Qc)(R) = {a/sla E R; s E S}.
From Theorem A-1, we know that the elements of Qc)(R) constitute a
partition of the set R x S. That is, the ordered pairs in R x S fall into
disjoint classes (calledformalfractions), with each class consisting ofequiva-
lent pairs, and nonequivalent pairs belonging to different classes. Further,
two such classes a/s and b/r are identical if and only if ar = sb; in particular,
all fractions of the form as/s, with s E S, are equal.
With these remarks in mind, let us introduce the operations of addition
and multiplication required to make Qc)(R) into a ring. We do this by
means of the formulas
a/s + b/r = (ar + sb)/sr,
(a/s)(b/r) = ab/sr.
Notice, incidentally, that since the set S is closed under multiplication,
the right-hand sides of the defining equations are meaningful.
As usual, our first task is to justify that these operations are well-defined;
62
that is to say, it is necessary to show that the sum and product are independent
of the particular elements of R used in their definition. Let us present the
argument for addition in detail. Suppose, then, that a/s = a'/s' and
b/r = b'/r'; we must show that
(ar + sb)/sr = (a'r' + s'b')/s'r'.
From what is given, it follows at once that
as' = sa', br' = rb'.
These equations imply
(ar + sb)(s'r') - (a'r' + s'b')(sr) = (as' - sa')(rr') + (br' - rb')(ss')
= O(rr') + O(ss') = O.
By the definition of equality of equivalence classes, this amounts to saying
that
(ar + sb)/sr = (a'r' + s'b')/s'r',
which proves addition to be well-defined. In much the same way, one can
establish that
ab/sr = a'b'/s'r'.
The next lemma reveals the algebraic nature of Qc1(R) under these
operations.
,
Lemma. The system Qcl(R) forms a commutative ring with identity.
Proof It is an entirely straightforward matter to confirm that Qcl(R) is a
commutative ring. We leave the reader to make the necessary verifications
at his leisure, and merely point out that O/s serves as the zero element, while
-a/s is the negative of a/so
That the equivalence class s'/s', where s' is any fixed non-zero-divisor
of R, constitutes the multiplicative identity is evidenced by the following
computation:
(a/s)(s'/s') = as'/ss' = a/s
for arbitrary a/s in Qcl(R), since (as')s = (ss')a. Loosely speaking, common
factors belonging to S may be cancelled in a fraction as'/ss'.
This proves part of the theorem below.
Theorem 4-9. Any commutative ring R with at least one non-zero-
divisor possesses a classical ring of quotients.
Proof We begin by establishing that the ring Qcl(R) contains a subring
isomorphic to R. For this, consider the subset K of Qcl(R) consisting of all
INTEGRAL DOMAINS AND FIELDS 63
The reader can easily check that K is a subring of Qcl(R). An obvious (onto)
mappingf: R ~ K is defined by takingf(a) = aso/s o. Since the condition
aso/s o = bso/so implies that as~ = bs~ or, after cancelling, that a = b, f
will be a one-to-one function. Furthermore, it has the property of pre-
serving both addition and multiplication:
members of S but for all elements of Qcl(R) which can be represented in the
form r/s, where r, s are both non-zero-divisors; in fact,
=
rs/sr = so/so.
(r/s)(s/r)
When the ring R is an integral domain, we may take the set S of non-zero-
divisors as consisting of all the elements of R which are not zero. The last
remark of the preceding paragraph then leads to the following important
theorem.
Theorem 4-10. For any integral domain R, the system Qcl(R) forms a
field, customarily known as the field of quotients of R.
Since an integral domain is (isomorphic to) a subring of its field of
quotients, we also obtain
Corollary. A ring is an integral domain if and only if it is a subring of
a field.
It should be pointed out that the hypothesis of commutativity is essential
to this last theorem; indeed, there exist noncommutative rings without
divisors of zero that cannot be imbedded in any division ring.
The field of quotients constructed from the integral domain Z is, of
course, the rational number field Q. Another fact of interest is that the field
of quotients is the smallest field in which an integral domain R can be
imbedded, in the sense that any field in which R is imbeddable contains a
subfield isomorphic to Qc](R) (Problem 20).
The existence theorem for the classical ring of quotients can be supple-
mented by the following result, which shows that it is essentially unique.
Theorem 4-11. Let Rand R' be two commutative rings, each containing
at least one non-zero-devisor. Then, any isomorphism of R onto R' has a
unique extension to an isomorphism of Qc](R) onto Qc1(R').
Proof To begin with, each member of Qc](R) may be written in the form
ab -1, where a, b E Rand b is a non-zero-divisor in R. Given an isomorphism
¢: R -+ R', the element ¢(b) will be a non-zero-divisor of R', so that ¢(b)-1
is present in Qc](R'). Suppose that ¢ admits an extension to an isomorphism
«1>: Qcl(R) -+ Qc1(R'). Since a = (ab- 1)b, we would then have
¢(a) = «I>(a) = «I>(ab- 1)«I>(b) = «I>(ab- 1)¢(b),
which, as a result, yields «I>(ab- 1) = ¢(a)¢(b)-1. Thus, «I>
is completely
determined by the effect of ¢ on R and so determined uniquely, if it exists
at all.
These remarks suggest that, in attempting to extend ¢, we should consider
the assignment:
for all
INTEGRAL DOMAINS AND FIELDS 65
Since every field contains a unique prime subfield, the following sub-
sidiary result is of interest.
Corollary 1. Every field contains a subfield which is isomorphic either
to the field Q or to one of the fields Zp.
Theorem 4-12 also provides some information regarding field auto-
morphisms.
Corollary 2. If f is an automorphism of the field F, then f(a) = a for
each element a in the prime subfield .of F (hence, a prime field has no
automorphism except the identity).
Proof The prime subfield of F is either
Fl = {(n1)(m1)-1In,mEZ; m =1= O}
or
F2 = {n1ln = 0,1, ... ,p - 1},
according as the characteristic of F is 0 or a prime p. Since any automor-
phism of a field carries the identity 1 onto itself, the result should be clear.
PROBLEMS
1. a) Assuming that R is a division ring, show that cent R forms a field.
b) Prove that every subring, with identity, of a field is an integral domain.
2. Let R be an integral domain and consider the set Z1 of all integral multiples of
the identity element:
Z1 = {n11 n E Z}.
(-7Ja P)
a = (a + bi
- c + di
c + di)
a - bi
(a, b, c, dE R#).
Prove that R is a division ring isomorphic to the division ring of real quatemions.
6. By the quaternions over a field F is meant the set of all q = a + bi + cj + dk,
where a, b, c, d E F and where addition and mUltiplication are carried out as with
real quatemions. Given that F is a field in which a 2 + b2 + c2 + d2 = 0 if and
only if a = b = c = d = 0, establish that the quatemions over F form a division
ring.
68
8. Let -r(n) denote the number of (distinct) positive divisors of an integer n > 1.
Prove that
a) If n has the prime factorization n = p~1P22 ... pZ", where the p; are distinct
primes and n; E Z+, then -r(n) = (nl + l)(n2 + 1) ... (nk + 1).
b) The number of ideals of Zn is -r(n).
c) -r(n)</>(n) ~ n. [Hint: II(n; + I)II(1 - lip;) ~ 2kn(1/2t.]
9. Given that the set H n = {[a] E Zn I[a] is not a zero divisor of Zn}, prove that
(Hn' 'n) forms a finite group of order </>(n).
"*
10. a) Derive Fermat's Little Theorem: If p is a prime number and a 0 (mod p),
then aP-1 == 1 (mod p).
b) If gcd (a, n) = 1, show that the equation ax == b (mod n) has a unique solution
modulo n. [Hint: All solutions are given by x = ba<!>(m-I)+ kn.]
12. Let f be a homomorphism from the ring R into the ring R' and suppose that R
has a subring F which is a field. Establish that either F ~ ker f or else R' contains
a subring isomorphic to F.
15. Prove that if the field F is of characteristic p > 0, then every subfield of F has
characteristic p.
PROBLEMS 69
to prove that R/I is a field, it suffices to show that each nonzero element of
R/I has a multiplicative inverse. Now, if the coset a + I =I- I, then a ¢ 1.
By virtue of the fact that I is a maximal ideal, the ideal (I, a) generated by
I and a must be the whole ring R:
R = (l,a) = {i + raliEI,rER}.
That is to say, every element of R is expressible in the form i + ra, where
i E I and r E R. The identity element 1, in particular, may be written as
1 = T + ra for suitable choice of TEl, r E R. But then, the difference
1 - ra E I. This obviously implies that
1 + I = ra + I = (1' + I)(a + I),
which asserts that r + I = (a + 1) - 1. Hence, R/I is a field.
For the opposite direction, we suppose that R/I is a field and J is any
ideal of R for which I c J 5; R. The argument consists of showing that
J = R, for then I will be a maximal ideal. Since I is a proper subset of J,
there exists an element a E J with a ¢ I. Consequently, the coset a + I =l-
I, the zero element of R/1. Since R/I is assumed to be a field, a + I must
have an inverse under multiplication,
(a + I)(b + I) = ab + I = 1 + I,
for some coset b + IE R/1. It then follows that 1 - ab E I c J. But the
product ab also lies in J (recall that a is an element of the ideal J), imply~ng
that the identity 1 = (1 - ab) + ab E J. This in turn yields J = R, as
desired.
Example 5-3. Consider the ring Ze of even integers, a commutative ring
without identity. In this ring, the principal ideal (4) generated by the integer
4 is a maximal ideal, where
(4) = {4(2j) + 4kjj, k E Z} = 4Z.
The argument might be expressed as follows. If n is any element not in (4),
then n is an even integer not divisible by 4; consequently, n can be expressed
in the form n = 4m + 2 for some integer m. We then have
2 = 4(-m) + nE((4),n),
so that Ze = (2) = ((4), n). By virtue of Theorem 5-1, this is sufficient to
demonstrate the maximality of the ideal (4).
Now, note that in the quotient ring Ze/(4),
(2 + (4))(2 + (4)) = 4 + (4) = (4).
The ring Ze/(4) therefore possesses divisors of zero and cannot be a field.
The point which we wish to make is that the assumption of a multiplicative
identity is essential to Theorem 5-5.
76
We now shift our attention from maximal ideals to prime ideals. Before
formally defining this notion, let us turn to the ring Z of integers for
motivation. Specifically, consider the principal ideal (p) generated by a
prime number p. If the product ab E (P), where a, b E Z, then p divides abo
But if a prime divides a product, it necessarily divides one of the factors.
This being the case, either a E (p) or bE (P). The ideal (P) thus has the
interesting property that, whenever (P) contains a product, at least one of
the factors must belong to (P). This observation serves to suggest and
partly to illustrate the next definition.
Definition 5-2. An ideal I of the ring R is a prime ideal if, for all a, b in
R, ab E I implies that either a E I or b E 1.
By induction, Definition 5-2 can easily be extended to finitely many
elements: an ideal I of R is prime if, whenever a product a 1 a 2 ••• an of
elements of R belongs to I, then at least one of the a i E I. In this connection,
we should caution the reader that many authors insist that the term "prime
ideal" always means a proper ideal.
Example 5-4. A commutative ring R with identity is an integral domain if
and only if the zero ideal {O} is a prime ideal of R.
Example 5-5. The prime ideals of the ring Z are precisely the ideals (n),
where n is a prime number, together with the two trivial ideals {O} and Z.
From above, we already know that if n is a prime, then the principal ideal
(n) is a prime ideal of Z. On the other hand, consider any ideal (n) with n
composite (n =F 0, 1); say, n = n 1 n 2 , where 1 < n 1, n 2 < n. Certainly the
product n 1 n 2 = n E (n). However, since neither n 1 nor n 2 is an integral
multiple of n, n1 ¢: (n) and n 2 ¢: (n). Hence, when n is composite, the ideal
(n) cannot be prime. Notice also that although {O} is prime, it is not a
maximal ideal of Z.
Example 5-6. For an illustration of a ring possessing a nontrivial prime
ideal which is not maximal, take R = Z x Z, where the operations are per-
formed componentwise. One may readily verify that Z x {O} is a prime
ideal of R. Since
Zx {O} C ZxZe c R,
with Z x Ze an ideal of R, Z x {O} fails to be maximal.
By analogy with Theorem 5-5, the prime ideals of a ring may be charac-
terized in the following manner.
Theorem 5-6. Let I be a proper ideal of the ring R. Then I is a prime
ideal if and only if the quotient ring RjI is an integral domain.
Proof First, take I to be a prime ideal of R. Since R is a commutative
MAXIMAL, PRIME, AND PRIMARY IDEALS 77
ring with identity, so is the quotient ring Rjl. It remains therefore only
to verify that Rjl is free of zero divisors. For this, assume that
(a + I)(b + I) = I.
In other words, the product of these two cosets is the zero element of the
ring Rjl. The foregoing equation is plainly equivalent to requiring that
ab + I = I, or what amounts to the same thing, ab E I. Since I is assumed
to be a prime ideal; one of the factors a or b must be in 1. But this means
that either the coset a + I = I or else b + I = I; hence, Rjl is without
zero divisors.
To prove the converse, we simply reverse the argument. Accordingly,
suppose that Rjl is an integral domain and the product ab E I. In terms of
cosets, this means that
(a + I)(b + I) = ab + I = I.
By hypothesis Rjl contains no divisors of zero, so that a + I = I or
b + I = 1. In any event, one of a or b belongs to I, forcing I to be a prime
ideal of R.
There is an important class of ideals which are always prime, namely,
the maximal ideals. From the several ways of proving this result, we choose
the argument given below; another approach involves the use of Theorems
5-5 and 5-6.
Theorem 5-7. In a commutative ring with identity, every maximal ideal
is a prime ideal.
Proof Assume that I is a maximal ideal of the ring R, a commutative ring
with identity, and the product ab E I with a ¢ I. We propose to show that
bEl. The maximality of I implies that the ideal generated by I and a must
be the whole ring: R = (I, a). Hence, there exist elements i E I, r E R such
that 1 = i + ra. Since both ab and i belong to I, we conclude that
b = Ib = (i + ra)b = ib + r(ab) E I,
from which it follows that I is a prime ideal of R.
We should point out that without the assumption of an identity element
this last result does not remain valid; a specific illustration is the ring Ze
of even integers, where (4) forms a maximal ideal which is not prime. More
generally, one can prove the following: if R is a commutative ring without
identity, but having a single generator, then R contains a nonprime maximal
ideal. To establish this, suppose that R = (a). First observe that the
principal ideal (a 2 ) is a proper ideal of R, since the generator a ¢ (a 2 ).
Indeed, were a in (a 2 ), we could write a = ra 2 + na 2 for some r E Rand
n E Z; it is a simple matter to check that the element e = ra + na would
78
4) !JI /i. =
80
Proof. Since property (1) is the only fact that will be explicitly required
in the body of the text, we shall content ourselves with its derivation; the
proofs of the remaining assertions are quite elementary and are left as an
exercise.
Now, if an E IJ, then an E I n J, and so an E I, an E J. We thus conclude
that JIJ s;; .jI n J s;; JT n.jJ. On the other hand, if it happens that
a E JI n .jJ, there must exist positive integers n, m, for which an E I and
am E J. This implies that the element an+ m = anamE IJ; hence, a E .jIJ.
Accordingly, JI n .jJ S;; JfJ and the desired equality follows.
P is any prime ideal containing Q and let a E JQ. Then there exists a suitable
positive integer n such that an E Q C;; P. Being prime, the ideal P must
contain the element a itself, which yields the inclusion JQ C;; P.
The primary ideals of R may be characterized in the following way.
Theorem 5-13. Let I be an ideal of the ring R. Then I is a primary
ideal if and only if every zero divisor of the quotient ring R/I is nilpotent.
Proof First, suppose that I is a primary ideal of R and take a + I to be
a zero divisor of R/I. Then there exists some coset b + I =F I, the zero
element of R/I, for which (a + I)(b + I) = I; that is, ab + I = I. There-
fore ab E I and, since b + I =F I, we also have b ¢ I. Now, I is assumed to
be primary, so that an E I for some positive integer n. This being the case,
+ I)n = an + I =
(a I,
which shows that the coset a + I is nilpotent.
Going in the other direction, we assume that any zero divisor of R/I
is nilpotent and let ab E I, with b ¢ 1. It then follows that (a + I)(b + I) = I,
while b + I =F I; if a + I =F I, this amounts to saying that a + I is a zero
divisor in R/I. By hypothesis, there must exist some n E Z + such that
(a + I)" = I, which forces the element an to be in 1. Thus, I is a primary
ideal of R.
Theorem 5-13 serves to emphasize the point that primary ideals are a
modification of the notion of a prime ideal; for, in the quotient ring of a
prime ideal, there are no zero divisors (hence, in a vacuous sense, every zero
divisor is nilpotent).
The following somewhat special result will be needed later, so we pause
to establish it before proceeding.
Corollary. If Ql' Q2' ... , Qn are a finite set of primary ideals of the
ring R, all of them having the same associated prime ideal P, then
Q = n7=1 Qi is also primary, with JQ = P.
Proof Before we delve into the details of the proof, observe that, by
Theorem 5-10,
.JQ = J n Qi = n JQ; = n P = P.
Now, suppose that a + Q is a zero divisor of the quotient ring R/Q. In
this event, we can find a coset b + Q =F Q such that
ab + Q = (a + Q)(b + Q) = Q.
Since b ¢ Q = n Qi' there exists some index i for which b ¢ Qi' Further-
more, ab E Qi with Qi primary, so that the element a E JQi = P = JQ.
This implies an E Q for some integer n; in consequence,
(a + Q)n = an + Q = Q,
MAXIMAL, PRIME, AND PRIMARY IDEALS 83
PROBLEMS
In the following set of problems, all rings are assumed to be commutative with identity.
1. a) Prove that Z EB Ze is a maximal ideal of the external direct sum Z EB z.
b) Show that the ring R is a field if and only if {O} is a maximal ideal of R.
2. Prove that a proper ideal M of the ring R is maximal if and only if, for every
element r ¢ M, there exists some a E R such that 1 + ra E M.
3. Letfbe a homomorphism from the ring R onto the ring R'. Prove that
a) if M is a maximal ideal of R with M ;2 ker J, thenf(M) is a maximal ideal of R',
b) if M' is a maximal ideal of R', thenf-l(M') is a maximal ideal of R,
c) the mapping M ..... f(M) defines a one-to-one correspondence between the set
of maximal ideals of R which contain kerf and the set of all maximal ideals
of R'.
4. If M 1 and M 2 are distinct maximal ideals of the ring R, establish the equality
M1M2 = Ml n M 2 •
5. Let M be a proper ideal of the ring R. Prove that M is a maximal ideal if and
only if, for each ideal I of R, either I ~ M or else I + M = R.
6. An ideal I of the ring R is said to be minimal if I +- {O} and there exists no ideal
J of R such that {O} c J c I.
a) Prove that a nonzero ideal I of R is a minimal ideal if and only if (a) = I for
each nonzero element a E I.
b) Verify that the ring Z of integers has no minimal ideals.
7. Let I be a proper ideal of the ring R. Show that I is a prime ideal if and only
if the complement of I is a multiplicatively closed subset of R.
8. In the ring R = map R #, define the set I by
9. a) With the aid of Theorem 5-5 and Example 5-1, obtain another proof of the fact
that Zp is a field if and only if p is a prime number.
b) Prove that in Zn the maximal ideals are the principal ideals (p) = pZn' where
p is a prime dividing n.
10. Given thatfis a homomorphism from the ring R onto the ring R', verify that
a) R' is a field if and only if ker f is a maximal ideal of R,
b) R' is an integral domain if and only if ker f is a prime ideal of R.
11. a) Show that if P I and P 2 are two ideals of the ring R such that PI '*
P 2 and
P2 '*P l' then the ideal Pin P 2 is not prime.
b) Let {Pi} be a chain of prime ideals of the ring R. Prove that v Pi and n Pi are
both prime ideals of R.
12. Prove that if I is an ideal of the ring Rand P is a prime ideal of I, then P is an
ideal of the whole ring R.
13. Let R denote the set of all infinite sequences {an} of rational numbers (that is,
an E Q for every n). R becomes a commutative ring with identity if the ring
operations are defined termwise:
b) if P' is a prime (primary) ideal of R', thenf -I(P') is a prime (primary) ideal of R;
c) the mapping P ---+ f(P) defines a one-to-one correspondence between the set
of prime (primary) ideals of R which contain ker f and the set of all prime
(primary) ideals of R'.
18. If M is a maximal ideal of the ring Rand n E Z+, show that the quotient ring R/Mft
has exactly one proper prime ideal. [Hint: Problem 17(c).]
19. a) For any ideal I of R, prove that I and .J1 are contained in precisely the same
prime ideals of R.
b) Using part (a), deduce that whenever I is a prime ideal of R, then I = .J1.
20. Letfbe a homomorphism from the ring R onto the ring R'. Prove that
a) if I is an ideal of R with I ;2 kerf, then .Jl(i) = f(.J1),
b) if l' is an ideal of R', then .Jr 1(1') = r I(.J r).
21. Verify that the intersection of semiprime ideals of the ring R is again a semiprime
ideal of R.
22. If I is an ideal ofthe ring R, prove that.J1 is the smallest (in the set-theoretic sense)
semiprime ideal of R which contains 1.
23. Establish that every divisor of zero in the ring Zpn (p a prime, n > 0) is nilpotent.
24. Let" J, and Q be ideals of the ring R, with Q primary. Prove the following
statements:
a) if I <t .jQ, then the q~otient Q:I = Q;
b) ifIJ ~ QandI <t .JQ,thenJ ~ Q;
c) if IJ ~ Q and the ideal J is finitely generated, then either I ~ Q or else r ~ Q
for some n E Z+. [Hint: If I $ Q, each generator of J is in.jQ.]
25. Assume that I is an ideal of the ring R. If.J1 is a maximal ideal of R, show that
I is primary. [Hint: Mimic the argument of the corollary to Theorem 5-14.]
26. Let R be an integral domain and P be a prime ideal of R. Consider Rp , the ring
of quotients of R relative to the complement of P:
Rp = {ab- I E Qc)(R)la E R; b ¢ Pl.
Prove that the ring Rp (which is known as the localization of R at the prime ideal P)
has exactly one maximal ideal, namely, I = {ab- I E Qc)(R)la E P; b ¢ Pl.
27. A ring R is said to be a local ring if it has a unique maximal ideal. If R is a local
ring with M as its maximal ideal, show that any element a ¢ M is invertible in R.
28. A subring R of a field F is said to be a valuation ring of F if for each nonzero
element a E F at least one of a or a-I belongs to R. Assuming that R is a valuation
ring of F, prove the following:
a) R contains all the idempotent elements of F.
b) R is a local ring, with unique maximal ideal M = {a E Rla- I ¢ R}.
[Remark. a-I denotes the inverse of a in F.]
c) For any two elements a, bE R, either aR ~ bR or bR ;2 aR. [Hint: Either
ab- I E R or ba- I E R.]
PROBLEMS 89
As the title suggests, this chapter is concerned with the problem of factoring
elements of an integral domain as products of irreducible elements. The
particular impetus is furnished by the ring of integers, where the Funda-
mental Theorem of Arithmetic states that every integer n > 1 can be written,
in an essentially unique way, as a product of prime numbers; for example,
the integer 360 = 2·2·2·3·3· 5. We are interested here in the possibility
of extending the factorization theory of the ring Z and, in particular, the
aforementioned Fundamental Theorem of Arithmetic to a more general
setting. Needless to say, any reasonable abstraction of these number-
theoretic ideas depends on a suitable interpretation of prime elements (the
building blocks for the study of divisibility questions in Z) in integral
domains. Except for certain definitions, which we prefer to have available
for arbitrary rings, our hypothesis will, for the most part, restrict us to
integral domains. The plan is to proceed from the most general results
about divisibility, prime elements, and uniqueness offactorization to stronger
results concerning specific classes of integral domains.
Throughout this chapter, the rings considered are assumed to be com-
mutative; and it is supposed that each possesses an identity element.
°
Definition 6-1. If a =1= and b are elements of the ring R, then a is
said to divide b, in symbols alb, provided that there exists some c E R
such that b = ac. In case a does not divide b, we shall write a { b.
Other language for the divisibility property a Ib is that a is a factor of
b, that b is divisible by a, and that b is a multiple of a. Whenever the notation
alb is employed, it is to be understood (even if not explicitly mentioned)
that the element a =1= 0; on the other hand, not only may b = 0, but in
such instances we always have divisibility.
Some immediate consequences of this definition are listed below; the
reader should convince himself of each of them.
Theorem 6-1. Let the elements a, b, c E R. Then,
1) alO, 11a, ala;
90
DIVISIBILITY THEORY IN INTEGRAL DOMAINS 91
Proof. To prove the equivalence of (1) and (2), suppose that a = bu, where
u is an invertible element; then, also, b = au-I, so that both alb and bla.
Going in the opposite direction, if alb, we can write b = ax for some x E R;
while, from bla, it follows that a = by with y E R. Therefore, b = (by)x =
b(yx). Since b =1= 0, the cancellation law implies that 1 = yx. Hence, y is
an invertible element of R, with a = by, proving that a and b must be
associates. The equivalence of (2) and (3) stems from our earlier remarks
relating division of ring elements to ideal inclusion.
We next examine the notion of a greatest common divisor.
Definition 6-3. Let aI' a 2 , ... ,an be nonzero elements of the ring R.
An element dE R is a greatest common divisor of aI' a 2 , •• , ,an if it
possesses the properties
1) dla; for i = 1, 2, ... , n (d is a common divisor),
2) cia; for i = 1, 2, ... , n implies that cld.
The use of the superlative adjective "greatest" in this definition does
not imply that d has greater magnitude than any other common divisor c,
but only that d is a multiple of any such c.
A natural question to ask is whether the elements aI' a 2 , ••• , an E R
can possess two different greatest common divisors. For an answer, suppose
that there are two elements d and d' in R satisfying the conditions of
Definition 6-3. Then, by (2), we must have did' as well as d'ld; according
to Theorem 6-2, this implies that d and d' are associates. Thus, the greatest
common divisor of aI' a2 , ••• , an is unique, whenever it exists, up to arbitrary
invertible factors. We shall find it convenient to denote any greatest com-
mon divisor of aI' a 2 , ... , an by gcd (aI' a 2 , ... , an)'
The next theorem will prove, at least for principal ideal rings, that any
finite set of nonzero elements actually does have a greatest common divisor.
DIVISIBILITY THEORY IN INTEGRAL DOMAINS 93
Theorem 6-3. Let a l , a2, ... ,an be nonzero elements of the ring R.
Then a l , a2, ... , an have a greatest common divisor d, expressible in the
form
(ri E R),
if and only if the ideal (a l , a2, ... , an) is principal.
Proof Suppose that d = gcd (a l , a2, ... , an) exists and can be written in
the form d = rla l + r 2a2 + ... + rnan, with ri E R. Then the element d
lies in the ideal (a l , a2, ... ,an)' which implies that (d) £ (a l , a2, ... , an).
To obtain the reverse inclusion, observe that since d = gcd (a l , a2, ... , an),
each ai is a multiple of d; say, ai = xid, where Xi E R. Thus, for an arbitrary
member yla l + Y2a2 + ... + Ynan of the ideal (a l , a2, ... , an), we must have
yla l + Y2 a 2 + ... + ynan = (YIX l + Y2 X2 + ... + Ynxn)d E (d).
This fact shows that (a l , a2, ... ,an) £ (d), and equality follows.
For the converse, let (a l , a2, ... , an) be a principal ideal of R:
(d E R).
Our aim, of course, is to prove that d = gcd (a l , a2, ... ,an)' Since each
ai E (d), there exist elements bi in R for which ai = b;d, whence dla i for
i = 1, 2, ... ,n. It remains only to establish that any common divisor C of
the ai also divides d. Now, ai = SiC for suitable Si E R. As an element of
(a l , a2, ... , an), d must have the form d = rla l + r2a2 + ... + rnan, with
r i in R. This means that
d = (rls l + r 2s2 + ... + rnsn)c,
which is to say that cld. Thus, d is a greatest common divisor of a l , a2, ... , an
and has the desired representation.
Corollary. Any finite set of nonzero elements a l , a2 , ... , anofaprincipal
ideal ring R has a greatest common divisor; in fact,
gcd (a l , a2, ... , an) = rla l + r2a2 + ... + rnan
for suitable choice of r l , r2, ... , rn E R.
When (at, a 2, ... , an) = R, the elements a l , a2, ... ,an must have a
common divisor which is an invertible element of R; in this case, we say
that a l , a2, ... , an are relatively prime and shall write gcd (a l , a2, ... ,an) = 1.
If a l , a2, ... , an are nonzero elements of a principal ideal ring R, then
the corollary to Theorem 6-3 tells us that a l , a2, ... , an are relatively prime
if and only if there exist r l' r2' ... , rn E R such that
(Bezout's Identity).
One ofthe most useful applications of Bezout's Identity is the following
(it also serves to motivate our coming definition of a prime element).
94
b = 1b = rab + scb.
As clab and clc, Theorem 6-1(5) guarantees that cl(rab + scb), or rather,
clb.
Dual to the notion of greatest common divisor there is the idea of a
least common multiple, defined below.
Proof We begin by assuming that d = lcm (ai' a z, ... ,an) exists. Then
the element d lies in each of the principal ideals (a;), for i = 1,2, ... , n,
whence in the intersection n (a i ). This means that (d) S; n (aJ On the
other hand, any element r En (a;) is a common multiple of each of the ai .
But d is a least common multiple, so that dlr, or, equivalently, r E (d). This
leads to the inclusion n (a;) S; (d) and the subsequent equality.
Going in the opposite direction, suppose that the intersection n (a i ) is
a principal ideal of R, say n (a;) = (a). Since (a) S; (a;), it follows that
aila for every i, making a a common multiple of ai' a2 , ... ,an' Given any
other common multiple b of ai' a2 , ... , an' the condition ailb implies that
(b) S; (a;) for each value of i. As a result, (b) S; n (a i ) = (a) and so alb.
Our argument establishes that a = lcm (ai' a2 , ... ,an)' completing the
proof.
DIVISIBILITY THEORY IN INTEGRAL DOMAINS 95
2} If gcd (ra l , ra 2 , ••• , ran) exists, then gcd (at, a2 , ..• , an) also exists and
Proof First, assume that d = km (a l , a2 , ••• ,an) exists. Then a;ld for
each value of i, whence radrd. Now, let d' be any common multiple of
ra 2 , ra 2 , ... ,ran. Then rid', say d' = rs, where s E R. It follows that ads
for every i and so dis. As a result, rdlrs or rdld'. But this means that
km (ra l , ra 2 , •.• , ran) exists and equals rd = r km (a l , a2 , ... , an).
As regards the second assertion, suppose that e = gcd (ra l , ra 2 , ••• , ran)
exists. Then rle; hence, e = rt for suitable t E R. Since elra;, we have tla; for
every i, signifying that t is a common divisor of the a;. Now, consider an
arbitrary common divisor t' of al' a2 , •.• , an. Then rt'lra; for i = 1,2, ... , n
and therefore rt'l e. But e = rt, so that rt'l rt or t' Jt. The implication is
that gcd (a l , a2 , ... , an) exists and equals t. This proves what we wanted:
inverse, then a = r-1p E (P), from which it follows that (a) £ (p), an obvious
contradiction. Accordingly, the elemerit a is invertible, whence (a) = R.
This argument shows that no principal ideal lies between (p) and the whole
ring R, so that (p) is a maximal principal ideal.
On the other hand, let (p) be a maximal principal ideal of R. For a
proof by contradiction, assume that p is not an irreducible element. Then
p admits a factorization p = ab where a, b E R and neither a nor b is in-
vertible (the alterpative possibility that p has an inverse implies (p) = R,
so may be ruled but). Now, if the element a were in (p), then a = rp for
some choice of r E R; hence, p = ab = (rp)b. Using the cancellation law,
we could deduce that 1 = rb; but this results in the contradiction that b
is invertible. Therefore, a ¢ (p), yielding the proper inclusion (p) c (a).
Next, observe that if (a) = R, then a will possess an inverse, contrary to
assumption. We thus conclude that (p) c (a) c R, which denies that (p)
is a maximal principal ideal. Our original supposition is false and must /V
be an irreducible element of R.
With regard to the second assertion of the lemma, suppose that p is any
prime element of R. To see that the principal ideal (p) is in fact a prime
ideal, we let the product ab E (p). Then there exists an element r E R for
which ab = rp; hence, plab. By hypothesis, p is a prime element, so that
either p Ia or pi b. Translating this into ideals, either a E (p) or b E (p); in
consequence, (p) is a prime ideal of R.
The converse is proved in much the same way. Let (p) be a prime ideal
and plab. Then ab E (p). Using the fact that (p) is a prime ideal, it follows
that one of a or b lies in (p). This means that either pia or else plb, and makes
p a prime element of R.
For principal ideal domains, all of this may be summarized by the
following theorem.
Theorem 6-10. Let R be a principal ideal domain. The nontrivial ideal
(p) is a maximal (prime) ideal of R if and only if p is an irreducible (prime)
element of R.
An immediate consequence of this theorem is that every nonzero non-
invertible element of R is divisible by some prime.
Corollary. Let a =1= 0 be a noninvertible element of the principal ideal
domain R. Then there exists a prime pER such that pia.
Proof Since a is not invertible, the principal ideal (a) =1= R. Thus, by
Theorem 5-2, there exists a maximal ideal M of R such that (a) £ M. But
the preceding result tells us that every maximal ideal is of the form M = (p),
where p is a prime element of R (in this setting, there is no distinction
between prime and irreducible elements). Thus, (a) £ (p), which is to say
that pia.
100
We conclude from this that the element ais expressible as the finite product
of primes )
)_ I
a - PlP2 ... Pn-lPn'
where p~ = Pnan, being an associate of a prime, is itself prime.
Corollary. In a principal ideal domain R; every nontrivial ideal is the
product of a finite number of prime (maximal) ideals.
Proof If the element 0 =1= a E R is not invertible, then a has a representation
as a finite product of primes; say a = PlP2 ... Pn' where each Pi is a prime
element of R. It then follows that
(a) = (PlP2 ... Pn) = (Pl)(P2) ... (Pn)'
with (Pi) a prime ideal of R.
Specializing to the ring of integers, we obtain a celebrated result.
Theorem 6-12. (Euclid). There are an infinite number of primes in z.
Proof Assume that the assertion is false; that is, suppose that there are
only a finite number of primes, say Pl' P2' ... ,Pn. Consider the positive
integer
a = (PlP2 ... Pn) + 1.
None of the listed primes Pi divides a. If a were divisible by Pi' for instance,
we would then have Pil(a - PlP2 ... Pm), by Theorem 6-1 (5), or Pill; but
this is impossible by part (2) of the same theorem. Since a > 1, Theoren:t
6-11 asserts that it must have a prime factor. Accordingly, a is divisible
by a prime which is not among those enumerated. This arguments shows
that there is no finite listing of the prime numbers.
Having proved the existence of a prime factorization in principal ideal
domains, one is naturally led to the question of uniqueness. Our next
theorem is the focal point of this chapter.
Theorem 6-13. Every principal ideal domain R is a unique factorization
domain.
Proof Theorem 6-11 shows that each noninvertible element 0 =1= a E R,
has a prime factorization. To establish uniqueness, let us suppose that a
can be represented as a product of primes in two ways, say
a = PlP2 ... Pn = qlq2 ... qm (n.::; m),
102
where the Pi and qi are all primes. Since Pll(qlq2 ... qm), it follows that
PI divides some qi (1 .::; i .::; m); renumbering, if necessary, we may suppose
that Pllql. Now, PI and ql are both prime elements of R, with Pllql' so
they must be associates: ql = PI u l for some invertible element u l E R.
Cancelling the common factor PI' we are left with
P2 ... Pn = u l q2 ... qm·
Continuing this argument, we arrive (after n steps) at
1 = U I U 2 ••• unqn+l ... qm.
Since the qi are not invertible, this forces m = n. It has also been shown
that every Pi has some qj as an associate and conversely. Thus, the two
prime factorizations are identical, apart from the order in which the factors
appear and from replacement of factors by associates.
Attention is called to the fact that the converse of Theorem 6-13 is not
true; in the next chapter, we shall give an example of a unique factorization
domain which is not a principal ideal domain.
A useful fact to bear in mind is that in a unique factorization domain
R, any irreducible element pER is necessarily prime. For, suppose that P
divides the product ab, say pc = abo Let
a = PI ... Pn' b = ql ... qm' and c = tl ... ts
with both <5( -a) = <5(a) < <5(a + b) and <5(b) < <5(a + b). This exhibits the
lack of uniqueness of quotient and remainder in condition (3).
Conversely, assume that the indicated inequality holds and that the
element a E R has two representations;
a = qb + r (r = 0 or c5(r) < <5(b»,
a = q'b + r' (r' = 0 or <5(r') < c5(b»)
with r =1= r' and q =1= qt. Then we have
c5(b) ~ <5((q - q')b) = <5(r - r') < max {<5(r), <5( - r')} < c5(b).
This is only possible if one of r - r' or q - q' is zero. Since each of these
conditions implies the other, uniqueness follows.
Corollary. (Division Algorithm for Z). If a, bE Z, with b =1= 0, then there
exist unique integers q and r such that
a = qb + r, o~ r < lal.
Proof Utilize the valuation <5 given by <5(a) = 14 for all nonzero a E Z.
Unique factorization in Z follows ultimately from the Division Algo-
rithm. It is not surprising that in rings where there is an analog of division
with remainder, we can also prove uniqueness of factorization. The main
line of argument consists of showing that every Euclidean domain is a
principal ideal domain. (One need only consider the ring Ze to see that the
converse of this does not hold.)
Theorem 6-15. Every Euclidean domain is a principal ideal domain.
Proof Let R be a Euclidean domain with valuation <5 and I be an ideal of
R; ignoring trivial cases, we may suppose that I =1= {O}. Consider the set
S defined by
S = {<5(a)la E I; a =1= O}.
Since S is a nonempty subset of nonnegative integers, it has a least element
by the Well-Ordering Principal. Pick bEl, so that <5(b) is minimal in S.
Our contention is that I = (b).
Let a be an arbitrary element of I. By the definition of Euclidean
domain, there exist elements q, r E R for which a = qb + r, where either
r = 0 or <5(r) < <5(b). Now, r = a - qb E I, since I is an ideal containing
both a and b. The alternative <5(r) < <5(b) would therefore contradict the
minimality of c5(b). Consequently, we must have r = 0, and a = qb E (b);
this implies that I £; (b). The reverse inclusion clearly holds, since bEl,
thereby completing the proof.
Corollary. Every Euclidean domain is a unique factorization domain.
DIVISIBILITY THEORY IN INTEGRAL DOMAINS 105
Theorem 6-16. For each square-free integer n, the system Q(.jn) forms
a field; in fact, Q(J"ii) is a subfield of C.
Proof The reader may easily verify that Q(-jn) is a commutative ring with
identity. It remains only to establish that each nonzero element of Q(.jn)
°
has a multiplicative inverse in Q(.jn). Now, if =1= a E Q(.jn), then the
element P = a/N(a) evidently lies in Q(.jn); furthermore, the product
ap = a(a/N(a) ) = N(a)jN(a) = 1,
so that Pserves as the inverse of a.
Contained in each quadratic field Q(.jn) is the integral domain
from which it follows that one of N(P) or N(y) must have the value ± 1.
From the first part of the lemma, we may thus conclude that either P or y
is invertible in Z(.jn), while the other is an associate of a. Accordingly, a
is an irreducible element of Z(.jn).
Example 6-3. Let us find all invertible elements in Z(i) = Z(R), the
domain of Gausian integers, by finding those members a of Z(i) for which
N(a) = 1 (in this setting, the norm assumes nonnegative values). If
a = a + bi E Z(i) and N(a) = 1, then a2 + b2 = 1, with a, b E Z. This
DIVISIBILITY THEORY IN INTEGRAL DOMAINS 107
Definition 6-8 is therefore satisfied in its entirety when n = -1, - 2,2, 3 and
the corresponding quadratic domains Z(Jn) are Euclidean.
Corollary. The domain Z(i) of Gaussian integers is a unique factoriza-
tion domain.
Remark. By investigating further the divisibility properties of Z(i), one can
prove the classic "two squares theorem" of Fermat: every prime number
ofthe form 4n + 1 is the sum oftwo squares; the interested reader is advised
to consult [13] for the details.
Example 6-4. The various integral domains studied in this chapter might
suggest that unique factorization of elements always holds. To round out
the picture with an example of the failure of unique factorization, let us
consider the quadratic domain Z(.j - 5). Observe that the element 9 has
two factorizations in Z(.j - 5), namely,
a1 = 3, a2 = 2 + J - 5, a3 = 2 - J - 5
reducible, we could write a i = py, where neither of the elements p,
y E Z(.j - 5) is invertible. Taking norms, it follows that
9 = N(ai) = N(P)N(y), N(P), N(y) E Z +,
which in turn yields N(P) = N(y) = 3. Hence, if p = a + we find bR,
that we must solve the equation a2 + 5b 2 = 3 for integers a and b; but
°
this equation obviously has no solutions in Z (b 1= implies that a 2 + 5b 2 ~ 5,
and if b = 0, then a 2 = 3). Thus, we have exhibited two genuinely different
factorizations of the element 9 into irreducibles, so that unique factorization
does not hold in Z(.j - 5).
Notice further that the common divisors of 9 and 3(2 + j=5) are
1, 3, and 2 + .j - 5. N one of these latter elements is divisible by the others,
so that gcd (9, 3(2 + J - 5)) fails to exist (in particular, Z(j=5) does not
have the gcd-property). On the other hand, the greatest common divisor
of 3 and 2 + j=5 does exist and, in fact, gcd (3, 2 + j=5) = 1. It follows
that only the right-hand side of the formula
PROBLEMS
b) Show that in the quadratic domain Z(.)6), the relation 6 = (.)6)2 = 3·2 does
not violate unique factorization.
18. Prove that the domain Z(0) is not a unique factorization domain by discovering
two distinct factorizations of the element 10. Do the same for element 9 in the
domain Z(.J - 7). ~
19. Show that the quadratic domain Z(.J -5) is not a principal ideal domain. [Hint:
Consider the ideal (3, 2 + .J - 5).]
20. Describe the field of quotients of the quadratic domain Z(.jYt) where n is a square-
free integer.
SEVEN
POLYNOMIAL RINGS
The next step in our program is to apply some of the previously developed
theory to a particular class of rings, the so-called polynomial rings. For
the moment, we shall merely remark that these are rings whose elements
consist of "polynomials" with coefficients from a fixed, but otherwise
arbitrary, ring. (The most interesting results occur when the coefficients
are specialized to a field.)
As a first order of business, we seek to formalize the intuitive idea of
what is meant by a polynomial. This involves an excursion around the
fringes of the more general question of rings of formal power series. Out
of the veritable multitude of results concerning polynomials, we have
attempted to assemble those facets of the theory whose discussion reinforces
the concepts and theorems expounded earlier; it is hoped thereby to convey
a rough idea of how the classical arithmetic of polynomials fits into ideal
theory. Our investigation concludes with a brief survey of some of the
rudimentary facts relating roots of polynomials to field extensions.
To begin with simpler things, given an arbitrary ring R, let seq R denote
the totality of all infinite sequences
f = (aD, aI' a 2, ... , ak, ... )
of elements ak E R. Such sequences are called formal power series, or merely
power series, over R. (Our choice of terminology will be justified shortly.)
We intend to introduce suitable operations in the set seq R so that the
resulting system forms a ring containing R as a subring. At the outset, it
should be made perfectly clear that two power series
f = (aD, aI' a2, ... ) and g = (b o, bl , b2, ... )
are considered to be equal if and only if they are equal term by term:
f = g if and only if ak = bk for all k ;:::.: O.
Now, power series may themselves be added and multiplied as follows:
f + g = (aD + bo, a l + bl , ... ),
112
POLYNOMIAL RINGS 113
(It is understood that the above summation runs over all integers i, j ~ 0
subject to the restriction that i + j = k.)
A routine check establishes that with these two definitions seq R becomes
a ring. To verify a distributive law, for instance, take
where
dk =
i+j=k
L ai(bj + c) =
i+j=k
L (aib j + aicj )
That is, ax is the specific member of seq R which has the element a for its
°
second term and for all other terms. More generally, the symbol ax n ,
n ~ 1, will denote the sequence
(0, ... , 0, a, 0, ... ),
where the element a appears as the (n + l)st term in this sequence; for
example, we have
ax 2 = (0,0, a, 0, ... )
and
ax 3 = (0, 0, 0, a, 0, ... ).
By use of these definitions, each power series
Next, take M' to be any maximal ideal of R[[x]] and define the set M
to consist of the constant terms of power series in M' :
M = {aoERILakxkEM'}.
The reader can painlessly supply a proof that M forms a maximal ideal of
the ring R. Notice incidentally that M must be a proper ideal. Were M = R,
then there would exist a power series L bnxn in M' with constant term
bo = 1. By the last lemma, L bnxn would then be an invertible element, so
that M' = R[[ x]], which is impossible. Owing to the inclusion M' ~ (M, x)
and the fact that M' is maximal in R[[ x]], it now follows that M' = (M, x).
To verify that the correspondence in question is indeed one-to-one,
suppose that (M, x) = (M, x), where M, M are both maximal ideals of the
ring R; what we want to prove is that M = M. Let rEM be arbitrary.
Given f(x) E R[[x]], the sum r + f(x)x E (M, x) = (M, x), so that
r + f(x)x = r + g(x)x
for appropriate r E M and g(x) E R[[x]J. Hence, r - r= (g(x) - f(x) )x.
If g(x) - f(x) =1= 0, then, upon taking orders,
° = ord(r - r) = ord (g(x) - f(x») + ord x ~ 1,
an absurdity. In consequence, we must have g(x) - f(x) = which, in its
turn, forces r = rEM. The implication is that M ~ M and, since M is
°
maximal, we end up with M = M. This completes the proof of the theorem.
Power series have so far received all the attention, but our primary
concern is with polynomials.
Definition 7-2. Let R[x] denote the set of all power series in R[[x]]
whose coefficients are zero from some index onward (the particular index
varies from series to series) :
R[x] = {a o + a 1 x + ... + anXnlak E R; n ~ o}.
An element of R[ x] is called a polynomial (in x) over the ring R.
In essence, we are defining a polynomial to be a finitely nonzero sequence
of elements of R. Thus, the sequence (1, 1, 1, 0, 0, ... ) would be a polynomial
over Z2' but (1, 0, 1,0, ... , 1,0, ... ) would not.
H is easily verified that R[x] constitutes a subring of R[[x]], the so-
called ring of polynomials over R (in an indeterminant x); indeed, if
f(x) = L akxk, g(x) = L bkxk are in R[x], with ak = °
for all k ~ nand
bk = °for all k ~ m, then
ak + bk = ° for k ~ max {m, n},
L
i+j=k
aib j = ° for k ~ m +-n,
POLYNOMIAL RINGS 119
so that both the sum f(x) + g(x) and product f(x)g(x) belong to R[ x].
Running parallel to the idea of the order of a power series is that of the
degree of a polynomial, which we introduce at this time.
Definition 7-3. Given a nonzero polynomial
in R[ x], we call an the leading coefficient of f(x); and the integer n, the
degree of the polynomial.
The degree of any nonzero polynomial is therefore a nonnegative integer;
no degree is assigned to the zero polynomial. Notice that the polynomials
of degree 0 are precisely the nonzero constant polynomials. If R is a ring
with identity, a polynomial whose leading coefficient is 1 is said to be a
monic polynomial.
As a matter of notation, we shall hereafter write degf(x) for the degree
of any nonzero polynomialf(x) E R[ x].
The result below is similar to that given for power series and its proof
is left for the reader to provide; the only change of consequence is that we
now use the notion of degree rather than order.
Theorem 7-5. If f(x) and g(x) are nonzero polynomials in R[ x], then
1) either f(x)g(x) = 0 or deg (J(x)g(x)) ::;; degf(x) + deg g(x), with
equality whenever R is an integral domain;
2) either f(x) + g(x) = 0 or
multiplicative inverse. For, suppose that f(x) E R[ x], with degf(x) > 0;
if f(x)g(x) = 1 for some g(x) in R[ x], we could obtain the contradiction
o= deg 1 = deg (J(x)g(x)) = degf(x) + deg g(x) 1- o.
The degree of a polynomial is used in the factorization theory of R[ x]
in much the same way as the absolute value is employed in Z. For, it is
through the degree concept that induction can be utilized in R[ x] to develop
a polynomial counterpart of the familiar division algorithm. One can
subsequently establish that the ring F[x] with coefficients in a field forms a
Euclidean domain in which the degree function is taken to be the Euclidean
valuation.
Before embarking on this program, we wish to introduce several new
ideas. To this purpose, let R be a ring with identity; assume further that
R' is any ring containing R as a subring (that is, R' is an extension of R)
and let r be an arbitrary element of R'. For each polynomial
a contradiction; the last inequality relies on the fact that the degrees of
r(x) and r'(x) are both less than the degree of g(x). Thus, q'(x) = q(x), which
in turn implies that r'(x) = r(x).
The polynomials q(x) and r(x) appearing in the division algorithm are
called, respectively, the quotient and remainder on dividing f(x) by g(x).
In this connection, it is important to observe that if g(x) is a monic poly-
nomial, or if R is taken to be a field, one need not assume that the leading
coefficient of g(x) is invertible.
POL YNOMIAL RINGS 123
Proof The polynomial h(x) = f(x) - g(x) is such that deg h(x) ~ nand,
by supposition, has at least n + 1 distinct roots in R. This is imposs~ble
unless h(x) = 0, whence f(x) = g(x).
Corollary 2. Let f(x) E R[ x], where R is an integral domain, and let S
be any infinite subset of R. Iff(a) = 0 for all a E S, thenf(x) is the zero
polynomial.
Example 7-2. Consider the polynomial x P - x E Zp[x], where p is a prime
number. Now, the nonzero elements of Zp form a commutative group under
mUltiplication of order p - 1. Hence, we have aP-l = 1, or aP = a for
every a =1= O. This is equally true if a = O. Our example shows that it may
very well happen that every element of the underlying ring is a root of a
polynomial, yet the polynomial is not zero.
With the Division Algorithm at our disposal, we can prove that the
ring F[ x] is rich in structure.
Theorem 7-10. The polynomial ring F[x], where F is a field, forms a
Euclidean domain.
Proof As has been noted, F[ x] is an integral domain. Moreover, the
function t5 defined by t5(j(x») = degf(x) for any nonzero f(x) E F[x] is a
suitable Euclidean valuation. Only condition (2) of Definition 6-6 fails to
be immediate. But iff(x) and g(x) are two nonzero polynomials in F[x],
Theorem 7-5 implies that
since deg g(x) ;:::: O. Thus, the function t5 satisfies the requisite properties of
a Euclidean valuation.
The reader is no doubt anticipating the corollary below.
Corollary. F[ x] is a principal ideal domain; hence, a unique factoriza-
tion domain.
For a less existential proof of the fact that F[x] is a principal ideal
domain and a considerably more precise description of its ideals, one can
repeat (with appropriate modifications) the pedestrian argument used in
Theorem 2-3. It will appear that any nontrivial ideal I of F[ x] is of the
form I = (j(x»), where f(x) is a nonzero polynomial of minimal degree in 1.
Since a field is trivially a unique factorization domain, part of the last
corollary could be regarded as a special case of the coming theorem.
Theorem 7-11. If R is a unique factoiization domain, then so is R[x].
POL YNOMIAL RINGS 125
Proof Suppose that R[x] is not a unique factorization domain and let S
be the set of all nonconstant polynomials in R[xJ which do not have a
unique factorization into irreducible elements. Select f(x) E S to be of
minimal degree. We may assume that
f(x) = Pl(X)P2(X) ... Pr(x) = ql(X)q2(X) ... qs(x),
where the Pi(X) and qlx) are all irreducible and
Now, either g(x) = 0, which forces aql(x) = bpl(X)Xn - m , or else deg g(x) <
degf(x). In the latter event, g(x) must possess a unique factorization into
irreducibles, some of which are q2(X), ... , qs(x) and Pl(X). The net result of
this is that Pl(x)lg(x), but Pl(X) t qJx) for i > 1, so that
Pl(x)l(aql(x) - bpl(X)X n - m ),
now to the real field, we can obtain the form of the prime factorization in
R # [x] (bear in mind that polynomials with coefficients from R # are poly-
nomials in C[x] and therefore have roots in C).
Corollary 2. If f(x) E R # [ x] is of positive degree, then f(x) can be
factored into linear and irreducible quadratic factors.
Proof Since f(x) also belongs to C[ x ],j(x) factors in C[ x] into a product
of linear polynomials x - Ck, CkEC. If CkER#, then x - CkER#[xJ.
Otherwise, Ck = a + bi, where a, b E R # and b =f O. But the complex roots
of real polynomials occur in conjugate pairs (Problem 7-11), so that
ck = a - bi is also a root off(x). Thus,
(x - ck) (x - ck) = x2 - 2ax + (a 2 + b2 ) E R#[x]
is a factor of f(x). The quadratic polynomial x 2 - 2ax + (a 2 + b2 ) is
irreducible in R # [x], since any factorization in R # [x] is also valid in C[ x}
and (x - ck ) (x - ck ) is its unique factorization in C[xJ.
An interesting remark, to be recorded without proof, is that if F is a
finite field, the polynomial ring F[x] contains irreducible polynomials of
every degree (see Theorem 9-10).
This may be a convenient place to introduce the notion of a primitive
polynomial.
Definition 7-5. Let R be a unique factorization domain. The content
of a nonconstant polynomial f(x) = ao + alx + ... + anxn E R[x],
denoted by the symbol contf(x), is defined to be the greatest common
divisor of its coefficients:
contf(x) = gcd (a o, aI' ... , an)·
We callf(x) a primitive polynomial if contf(x) = 1.
Viewed otherwise, Definition 7-5 asserts that a polynomialf(x) E R[x]
is primitive if and only if there is no irreducible element of R which divides
all of its coefficients. In this connection, it may be noted that in the domain
F[ x] of polynomials with coefficients from a field F, every nonconstant
polynomial is primitive (indeed, there are no primes in F). The reader should
also take care to remember that the notion of greatest common divisor and,
in consequence, the content of a polynomial is not determined uniquely,
but determined only to within associates.
Given a polynomialf(x) E R[x] of positive degree, it is possible to write
f(x) = Cfl(X), where C E Rand fl(X) is primitive; simply let C = contf(x).
To a certain extent this reduces the question of factorization in R[x] (at
least, when R is a unique factorization domain) to that of primitive poly-
nomials. By way of specific illustrations, we observe that f(x) = 3x 3 -
130
Thus, to say that v(J(x)) = 0 for some prime pER is equivalent to asserting
that ak E (p), or rather, plak for all k. But the latter condition signifies that
contf(x) +- 1; hence,f(x) is not primitive.
One of the most crucial facts concerning primitive polynomials is
Gauss's Lemma, which we prove next.
Now consider the reduction off(x) modulo the ideal P. Invoking hypothesis
(2), it can be inferred that
v(g(x))v(h(x)) = v(J(x)) = (an + p)xn.
Since the polynomial ring (R/P)[x] comprises an integral domain, the only
possible factorizations of (an + P)xn are into linear factors. This being so,
a moment's reflection shows that
v(g(x)) = (b r + P)xr ,
v(h(x)) = (c s + P)xs•
This means that the constant terms of these reductions are zero; that is,
bo + P = Co + P = P.
Altogether we have proved that both bo, Co E P, revealing at the same time
that a o = boc o E p2, which is untenable by (3). Accordingly, no such
factorization of f(x) can occur, and f(x) is indeed irreducible in R[ x J.
Theorem 7-18 leads almost immediately to the Eisenstein test for
irreducibility.
)
Corollary. (The Eisenstein Criterion). Let R be a unique factorization
domain and K be its field of quotients. Letf(x) = a o + a1x + ... +
an xn be a nonconstant polynomial in R[x J. Suppose further that for
some prime pER, P fan' plak for 0 :::;; k < n, and p2 {a o. Then,f(x) is
irreducible in K[xJ.
Proof We already know that (p) is a prime ideal of R. Taking stock of
the theorem, f(x) is an irreducible polynomial of R[ x] ; hence, of K[ x] (at
this point a direct appeal is made to Theorem 7-17).
This is probably a good time at which to examine some examples.
Example 7-5. Consider the monic polynomial
f(x) = xn + aEZ[x] (n > 1),
where a =/= ± 1 is a nonzero square-free integer. For any prime p dividing
a, p is certainly a factor of all the coefficients except the leading one, and
our hypothesis ensures that p2 {a. Thus, f(x) fulfils Eisenstein's criterion,
and so is irreducible over Q. Incidentally, this example shows that there
are irreducible polynomials in Q[x] of every degree.
Ontheotherhand,noticethatx4 + 4 = (x 2 + 2x + 2)(x 2 - 2x + 2);
one should not expect Theorem 7-18 to lead to a decision in this case, since,
of course, 4 fails to be a square-free integer.
Example 7-6. Eisenstein's test is not directly applicable to the cyclotonic
polynomial
134
Starting with a ring R we can first form the polynomial ring R[x], with
indeterminant x, and then the polynomial ring (R[ x ])[y] in another in-
determinant y. As the notation indicates, the elements of (R[ x ])[y] are
simply polynomials .
g = fo(x) + fl(X)Y + ... + J,,(x)yn,
where each coefficientf,.(x) E R[x], so that
f,.(x) = a Ok + a 1kx + ... + amkkXmk (aij E R).
Consequently, g can be rewritten as a polynomial in x and y,
m n
g = g(x, y) = L L aijxiy,
i=O j=O
with m, n nonnegative integers and aij elements of R (one makes the obvious
conventions that aooxOyO = aoo, aiOxiyo = aiOxi and aOjxOy = aOjY). In
accordance with tradition, we shall hereafter denote (R[ x ])[y] by R[ x, y]
and refer to the members of this set as polynomials over R in two indeter-
minants x and y.
Two such polynomials with coefficients aij and bij are by definition
equal if aij = bij for all i and j. Addition of polynomials is performed
termwise, while multiplication is given by the rule:
where
C ij = L
k+k'=i
ak1bk'I'·
1+1'=j
POLYNOMIAL RINGS 135
With these operations, the set R[x, y] becomes a ring containing R (or rather
an isomorphic copy of R) as a subring.
The (total) degree of a nonzero polynomial
m n
f(x, y) = L L aijxiyj
i=O j=O
is the largest of the integers i + j for which the coefficient aij +- and is
denoted, as before, by degf(x, y). Without going into details here, let us
°
simply state that it is possible to obtain inequalities involving degrees
analogous to those of Theorem 7-5; in particular, if R is an integral domain,
we still have
deg (j(x, y)g(x, y») = degf(x, y) + deg g(x, y).
From this rule, one can subsequently establish that whenever R forms an
integral domain, then so does the polynomial ring R[ x, y].
Rather than get involved in an elaborate discussion of these matters,
we content ourselves with looking at two examples. :>
Example 7-7. To illustrate that the ideal structure of the ring F[x, y]
(F a field) is more complicated than that of F[ x], let us show that F[ x, y]
is not a principal ideal domain. This is accomplished by establishing that
the ideal I = (x, y) is not principal, where
I = U(x, y)x + g(x, y)ylf(x, y), g(x, y) E F[x, y]}.
Notice that the elements of this ideal are just the p01ynomials in F[x, y]
having zero constant term.
Suppose that I was actually principal, say L; = (h(x, y»), where
deg h(x, y) 2 1. Since both x, y E I, there would exist polynomials f(x, y),
g(x, y) ip F[ x, y] satisfying
x = f(x, y)h(x, y), y = g(x, y)h(x, y).
Now, deg x = deg y = 1, which implies that deg h(x, y) = 1, and degf(x, y) =
deg g(x, y) = 0; what amounts to the same thing,
x = ah(x, y), y = bh(x, y) (a, b E F).
Moreover, h(x, y) must be a linear polynomial; for instance,
(C i E F).
But if the coefficient C2 +- 0, then x cannot be a multiple of h(x, y), and if
C 1 +- 0, y is not a multiple of h(x, y). This being the case, we conclude
that C 1 = c2 = 0, a contradiction to the linearity of h(x, y), and so I does
not form a principal ideal.
136
has degree less than that off(x) or else is the zero polynomial. In fact, the
cosets of I are uniquely determined by remainders on division by f(x) in
the sense that g(x) + I = h(x) + I if and only if g(x) and h(x) leave the
same remainder when divided by f(x).
Thus, if degf(x) = n > 1 (for instance,f(x) = ao + a1x + ... + anx n),
then the extension field F' may be described by
F' = {b o + blx + ... + bn_lxn- 1 + IlbkEF}.
Identifying bk + I with the element bk , we see as before that a typical coset
can be uniquely represented in the form
bo + bl(x + + ... + bn-1(x + I)n-l.
I)
As a final simplification, let us replace x + I by some new symbol A., so that
the elements of F' become polynomials in A:
F' = {b o + blA + ... + bn_lAn-llbkEF}.
Observe that since A = x + I is a root off(x) in F', calculations are carried
out with the aid of the relation ao + alA + ... + anAn = o.
The last paragraph serves to bring out the point that F' is a finite7\
extension of F with basis {l, A, A2 , ••• , An - l }; in particular, we infer that
[F':F] = n = degf(x).
To recapitulate: if f(x) E F[ x] is an irreducible polynomial over F, then
there exists a finite extension F' of F, such that [F' :F] = degf(x), in which
f(x) has a root. Moreover, F' is a simple algebraic extension generated by
a root of f(x). (Admittedly, some work could be saved by an appeal to
Theorems 7-21 and 7-25, but our object here is to present an alternative
approach to the subject.)
We pause now to examine two concrete examples of the ideas just
presented.
Example 7-12. Consider Z2' the field of integers modulo 2, and the poly-
nomial f(x) = x 3 + X + 1 E Z2[X]. Since neither of the elements 0 and 1
is a root of x 3 + x + 1,f(x) must be irreducible in Z2[X].
Theorem 7-27 thus guarantees the existence of an extension of Z2'
specifically, the field Z2[ x ]/(j(x)), in which the given polynomial has a root.
Denoting this root by A, the discussion above tells us that
Z2[X]/(j(x)) = {a + bA + cA2la, b, CE Z2}
= {O, 1, A, 1 + A, A2, 1 + A2, A + A2, 1 + A + A2},
where, of course, A3 + A + 1 = o.
As an example of operating in this field, let us calculate the inverse of
1 + A + A2 • Before starting, observe that by using the relations
A3 = -(A + 1) = A + 1, A4 = A2 + A
146
(our coefficients come from Z2> where -1 = 1), the degree of any product
can be kept less than 3. Now, the problem is to determine elements a, b,
e E Z2 for which
Carrying out the multiplication and substituting for ,13, ,14 in terms of 1,
A, and ,12, we obtain
(a + b + e) + aA + (a + b)A2 = l.
This yields the system of linear equations
a + b + e = 1, a = 0, a + b = 0,
with solution a = b = 0, e = 1; therefore, (1 + A + ,12)-1 = ,12.
It is worth noting that x 3 + x + 1 factors completely into linear
factors in Z2[X]/(j(x)) and has the three roots A, ,12, and A + ,12:
x3 + x + 1 = (x - A)(x - A2)(x - (A + ,P)).
Example 7-13. The quadratic polynomial x 2 + 1 is irreducible in R#[x].
For, if x 2 + 1 were reducible, it would be of the form
x2 + 1 = (ax + b)(ex + d)
= aex2 + (ad + be)x + bd,
where a, b, e, d E R #. It follows at once that ae = bd = 1 and ad + be = 0.
Therefore, be = -(ad), and
1 = (ae)(bd) = (ad)(be) = -(adf,
or rather, (ad)2 = -1, which is impossible.
In this instance, the extension field R#[x]/(x2 + 1) is described by
aj P-' j
F[ x ]/(j(x)) ~ F,[y]/(j'(y))
Certainly, <I> is an isomorphism of F(r) onto F'(r'), for the individual mappings
a, r, p-l are themselves isomorphisms. If a is an arbitrary element of F, then
<I>(a) = (P-l r) (a(a)) = (P-l r)(a + (f(x)))
0 0
PROBLEMS
b) The ideal I" consists of all power series having order 2.n, together with O.
c) nnez+1" = {O}.
2. For any field F, consider the set F<x) consisting of all expressions of the form
3. Let R be a commutative ring with identity. If R is a local ring, prove that the
power series ring R[[x]] is also local.
8. Let P be a prime ideal of R, a commutative ring with identity. Prove that P[x]
is a prime ideal of the polynomial ring R[x]. If M is a maximal ideal of R, is
M[x] a maximal ideal of R[x]?
PROBLEMS 153
9. Consider the polynomial domain F[ x], where F is a field, and a fixed element
rEF. Show that the set of all polynomials having r as a root,
Mr = {f(x) E F[x]lf(r) = O},
then
For any f(x), g(x) E R[x] and any r E R, establish that
a) o(j(x) + g(x») = of(x) + og(x).
b) o(rf(x» = rof(x).
c) o(j(x)g(x») = of(x)'g(x) + f(x)·og(x). [Hint: Induct on the number of terms
off(x).]
15. Suppose that R is a commutative ring with identity and let r E R be a root of the
nonzero polynomial f(x) E R[ x]. We call r a multiple root of f(x) provided that
f(x) = (x - r)·g(x) (n > 1),
154
where g(x) E R[x] is a polynomial such that g(r) 0/= O. Prove that an element
r E R is a multiple root of f(x) if and only if r is a root of both f(x) and t5f(x).
16. Let F be a field and f(x) E F[x] be a polynomial of degree 2 or 3. Deduce that
f(x) is irreducible in F[x] if and only ifit has no root in F. Give an example which
shows that this result need not hold if degf(x) ~ 4.
17. Prove that if a polynomial f(x) E F[x] (F a field of characteristic 0) is irreducible,
then all of its roots in any field containing F must be distinct. [Hint: First show
that gcd (J(x), t5f(x)) = 1.]
18. Given that I is a proper ideal of R, a commutative ring with identity, establish the
assertions below:
a) If v: R[x] ---> (R/I)[x] is the reduction homomorphism modulo the ideal I,
then ker v = I[x]; hence, R[x]/I[x] ~ (R/I)[x].
b) If the polynomial f(x) E R[x] is such that v(J(x)) is irreducible in (R/I)[x],
then f(x) is irreducible in R[ x].
c) The polynomial f(x) = x 3 - x 2 + 1 is irreducible in Z[x]. [Hint: Reduce
the coefficients modulo 2.]
19. Let R be a unique factorization domain. Show that any nonconstant divisor of a
primitive polynomial in R[x] is again primitive.
20. Utilize Gauss's Lemma to give an alternative proof of the fact that if R is a unique
factorization domain, then so is the polynomial ring R[x]. [Hint: For
00/= f(x) E R[x], induct on degf(x); if degf(x) > 0, write f(x) = Cfl(X), where
c E Rand fl(X) is primitive; if fl(X) is reducible, apply induction to its factors.]
21. Apply the Eisenstein Criterion to establish that the following polynomials are
irreducibleinQ[x] :f(x) = x 2 + l,g(x) = x 2 - X + l,andh(x) = 2x 5 - 6x 3 +
9x2 - 15. [Hint: Considerf(x + 1), g(-x).]
22. Let R be a unique factorization domain and K its field of quotients. Assume that
ab- 1 E K (where a and b are relatively prime) is a nonzero root of the polynomial
f(x) = ao + a1x + ... + anx" E R[x]. Verify that ala o and blan.
23. Prove the following assertions concerning the polynomial ring Z[x, y]:
a) The ideals (x), (x, y) and (2, x, y) are all prime in Z[x, y], but only the last is
maximal.
b) (x, y) = .J(x2, y) = .J(x2, xy, y2).
c) The ideal (x\ xy, y2) is primary in Z[x, y] for any integer k E Z +.
d) If I = (x 2, xy), then .Jlis a prime ideal, but I is not primary. [Hint:.JI = (x).]
24. Consider the polynomial domain F[ x, y], where F forms a field.
a) Show that (x 2, xy, y2) is not a principal ideal of F[ x, y].
b) Establish the isomorphism F[x, y]/(x + y) ~ F[x].
25. Let the element r be algebraic over the field F and let f(x) E F[ x] be a monic
polynomial such thatf(r) = O. Prove thatf(x) is the minimum polynomial of r
over F if and only if f(x) is irreducible in F[ x].
26. Assuming that F' is a finite extension of the field F, verify each of the statements
below:
PROBLEMS 155
a) When [F: F] is prime, F is a simple extension of F; in fact, F' = F(r) for every
element rEF' - F.
b) If f(x) E F[ x] is an irreducible polynomial whose degree is relatively prime to
[F':F], thenf(x) has no roots in F.
c) If rEF is algebraic of degree n, then each element of F(r) has as its degree an
integer dividing n.
d) Given fields K; (i = 1,2) such that F:2 K;:2 F, with [KI:F] and [K2:F]
relatively prime integers, necessarily KI ( l K2 = F.
27. Show that the following extension fields of Q are simple extensions and determine
their respective degrees: Q(.J3, .j7), Q(.J3, i), Q(.J2, ~5).
28. a) Prove that the extension field F = F(rl' r2, ... , r.), where each element r; is
algebraic over F, forms a finite extension of F. [Hint: If F; = F;_l(r;), then
F. = F' and [F':F] = TI;[F;+1: FJ]
b) If F" is an algebraic extension of F' and F' is an algebraic extension of F, show
that F" is an algebraic extension of F. [Hint: Each rEF" is a root of some
polynomial f(x) = ao + a1x + ... + a.x" E FT x]; consider the extension
fields K = F(a o, aI' ... , a.) and K' = K(r); r is algebraic over K'.]
29. Let F be an extension of the field F. Prove that the set of all elements in F' which
are algebraic over F constitute a subfield of F' ; applied to the case where F' = C
and F = Q, this yields the field of algebraic numbers. [Hint: If r, s are algebraic
over F, [F(r, s):F] is finite; hence, F(r, s) is an algebraic extension of F.]
30. a) Granting that f(x) = x 2 + X + 2 is an irreducible polynomial in Z3[X],
construct the multiplication table for the field Z3[X]/(j(x»).
b) Show that the polynomial f(x) = x 3 + x 2 + 1 E Z2[X] factors into linear
factors in Z2[ x ]/(j(x») by actually finding the factorization.
31. If n+- 1 is a (nonzero) square-free integer, verify that Q[x]/(x2 - n) forms a field
isomorphic to the quadratic field
35. Suppose that F' is the splitting field for the polynomialf(x) E F[x]; say
Prove that F' = F(rl' r 2 , ••• , r.). As a particular illustration, establish that
Q(J2, .J3)is the splitting field of (x 2 - 2)(x 2 - 3) E Q[x].
36. Letf(x) E Zp[x] be an irreducible polynomial of degree n, p a prime. Verify that
the field F = Zp[x]/(J(x») contains p. elements.
37. If F' is a splitting field of a polynomial of degree n over F, show that [F':F] ::::; n!
38. A field F' is said to be algebraically closed if F' has no proper algebraic extensions.
Assuming that F' is an algebraic extension of F, prove the equivalence of the
following statements:
a) F' is algebraically closed.
b) Every irreducible polynomial in F'[x] is linear.
c) Every polynomial in F[x] splits in F'.
(For a proof that every field has an algebraic extension which is algebraically
closed, the reader is referred to [23].)
EIGHT
The Jacobson radical always exists, since we know by Theorem 5-2 that
any commutative ring with identity contains at least one maximal ideal.
It is also immediately obvious from the definition that rad R forms an
ideal of R which is contained in each maximal ideal.
To fix ideas, let us determine the Jacobson radical in several concrete
rings.
to Example 5-1, the maximal ideals of Z are precisely the principal ideals
generated by the prime numbers; thus,
rad R = n {(p)lp is a prime number}.
that, if every ideal of R is finitely generated, then rad R is not only nil but
nilpotent.
This is a convenient place to also point out that a homomorphic image
of a semisimple ring need not be semisimple. An explicit example of this
situation can easily be obtained from the ring Z of integers. While Z forms
a ring without a Jacobson radical, its homomorphic image Zp" (p a prime;
n > 1) contains the nil ideal (p); appealing to Corollary 3 above, we see
that Z pn cannot be semisimple.
Example 8-4. Consider F[[ x]], the ring of formal power series over a field
F. From the lemma on page 116, it is known that an element f(x) =
ao + a1x + ... + anxn + ... has an inverse in F[[x]] if and only if the
constant term ao =1= O. This observation (in conjunction with Theorem 8-2)
implies that if g(x) = b o + b1x + ... + bnxn + ... , then
rad F[[ x]] = {f(x)ll - f(x)g(x) is invertible for all g(x) E F[[ x]]}
= {f(x) 11 - aob o =1= 0 for all bo E F}
= {f(x)la o = O} = (x).
Thus, we have a second proof of the fact that the Jacobson radical of F[[x]]
is the principal ideal generated by x.
We next prove several results bearing on the Jacobson radical of quotient
rings. The first of these provides a convenient method for manufacturing
semisimple rings; its proof utilizes both implications of the last theorem.
Theorem 8-3. For any ring R, the quotient ring Rlrad R is semisimple;
that is, rad(R/rad R) = {O}.
Proof Before becoming involved in details, let us remark that since rad R
constitutes an ideal of R, we may certainly form the quotient ring Rlrad R.
To simplify notation somewhat, we will temporarily denote rad R by 1.
Suppose that the coset a + J E rad (RIJ). Our strategy is to show that
the element a E J, for then a + J = J, which would imply that rad (RIJ)
consists of only the zero element of RI1. Since a + J is a member of rad (RI J),
Theorem 8-2 asserts that
(1 + J) - (r + J)(a + J) = 1 - ra + J
is invertible in RIJ for each choice of r E R. Accordingly, there exists a
coset b + J (depending, of course, on both r and a) such that
(1 - ra + J)(b + J) = 1 + 1.
This is plainly equivalent to requiring
1 - (b - rab) E J = rad R.
CERTAIN RADICALS OF A RING 161
1) rad (Rjl) 2
rad R
I
+ I
' and,
which is the first part of our theorem (the crucial step requires the inclusion
nI!:;MM 2 I + radR).
With an eye to proving (2), notice that whenever I ~ rad R, then
rad (Rjl) 2
rad R
I
+ I
2 (rad R)jI.
Thus, we need only to show the inclusion (rad R)jl 2 rad (Rjl). To this
purpose, choose the coset a + IE rad (Rjl) and let M be an arbitrary
maximal ideal of R. Since I ~ rad R ~ M, the image natIM = Mjl must
be a maximal ideal of the quotient ring Rjl (Problem 3, Chapter 5). But
then,
a + IE rad (Rjl) ~ Mjl,
forcing the element a to lie in M. As this holds for every maximal ideal of
R, it follows that a E rad R and so a + IE (rad R)jI. All in all, we have
proved that rad (Rjl) ~ (rad R)jl, which, combined with our earlier
inclusion, leads to (2).
Armed with Theorem 8-4, we are in a position to establish:
162
Theorem 8-5. For any ring R, rad R is the smallest ideal I of R such
that the quotient ring R/I is semisimple (in other words, if R/I is a
semisimple ring, then rad R ~ I). .
Proof From Theorem 8-3, it is already known that R/rad R is without
Jacobson radical. Now, assume that I is any ideal of R for which the
associated quotient ring R/I is semisimple. Using part (1) of the preceding
theorem, we can then deduce the equality (I + rad R)/I = I. This in turn
leads to the inclusion rad R ~ I, which is what we sought to prove.
This may be a good place to mention two theorems concerning the
number of maximal ideals in a ring; these are of a rather special character,
but typify the results that can be obtained.
Theorem 8-6. Let R be a principal ideal domain. Then, R is semi-
simple if and only if R is either a field or has an infinite number of
maximal ideals.
Proof Let {Pi} be the set of prime elements of R. According to Theorem
6--10, the maximal ideals of R are simply the principal ideals (Pi). It follows
that an element a E rad R if and only if a is divisible by each prime Pi. If
R has an infinite set of maximal ideals, then a = 0, since every nonzero
noninvertible element of R is uniquely representable as a finite product of
primes. On the other hand, if R contains only a finite number of primes
Ph P2' ... , Pn' we have
rad R = n (Pi) = (P1P2 ... Pn) +- {O},
n
i= 1
so that R cannot be semisimple. Finally, observe that if the set {Pi} is empty,
then each nonzero element of R is invertible and R is a field (in which case
rad R = {O}).
Corollary. The ring Z of integers is semisimple.
Theorem 8-7. Let {Mi }, i E oF, be the set of maximal ideals of the ring
R. If, for each i, there exists an element a i E Mi such that 1 - a i E
rad R - M i , then {Mi} is a finite set.
Proof Suppose that the index set oF is infinite. Then there exists a well-
ordering ::5: of oF under which oF has no last el~ent. (See Appendix A for
terminology.) For each i E oF, we define Ii ~. . ( )i <j Mj . Then {I;} forms
a chain of proper ideals of R. By hypothesis, we can select an element
ai E Mi such that 1 - ai E Ii - Mi. Now the ideal I = u Ii is also a proper
ideal of R, since 1 ¢ I. By our choice of the Ii' I is not contained in any
maximal ideal of R. Indeed, suppose that there does exist an index i for
which I ~ M i ; then,
CERTAIN RADICALS OF A RING 163
yielding the contradiction 1 E Mi' But it is known that every proper ideal
of R is contained in a maximal ideal of R (Theorem 5-2). From this
contradiction we conclude that § must be finite.
Let us now turn to a consideration of another radical which plays an
essential role in ring theory, to wit, the prime radical. Its definition may
also be framed in terms of the intersection of certain ideals.
Definition 8-2. The prime radical of a ring R, denoted by Rad R (in
contrast with rad R), is the set
Rad R = n {pip is a prime ideal of R}.
If Rad R = {O}, we say that the ring R is without prime radical or has
zero prime radical.
Theorem 5-7, together with Definition 8-1, shows that the prime radical
exists, forms an ideal of R, and satisfies the inclusion Rad R ~ rad R. It
is useful to keep in mind that, for any integral domain, the zero ideal is a
prime ideal; for these rings, Rad R = {O}. In particular, the ring F[[ x]]
of formal power series over a field F has zero prime radical but, as we already
know, a nontrivial Jacobson radical.
Perhaps the most striking result of the present chapter is that th~ prime
radical, although seemingly quite different, is actually equal to the nil
radical of a ring. The lemma below provides the key to establishing this
assertion.
Lemma. Let I be an ideal of the ring R. Further, assume that the subset
S ~ R is closed under multiplication and disjoint from 1. Then there
exists an ideal P which is maximal in the set of ideals which contain I
and do not meet S; any such ideal is necessarily prime.
Proof. Consider the family ffi of all ideals J of R such that I ~ J and
J n S = cp. This family is not empty since I itself satisfies the indicated
conditions. Our immediate aim is to show that for any chain of ideals {Ji}
in ffi, their union u J i also belongs to ffi. It has already been established
in Theorem 5-2 that the union of a chain of ideals is again an ideal; more-
over, since I ~ J i for each i, we certainly have I ~ u J i • Finally, observe
that
(u J i ) n S = u (J i n S) = u cp = cp.
The crux of the matter is that Zorn's Lemma can now be applied to infer
that ffi has a maximal element P; this is the ideal that we want.
By definition, P is maximal in the set of ideals which contain I but do
not meet S. To settle the whole affair there remains simply to show that
P is a prime ideal. For this purpose, assume that the product ab E P but
that a ¢ P and b ¢ P. Since it is strictly larger than P, the ideal (P, a) must
164
Proof If the element a ¢.ji, then the set S = {anln E Z+} does not intersect
I. Since S is closed under multiplication, the preceding lemma insures the
existence of some prime ideal P which contains I, but not a; that is, a does
not belong to the intersection of prime ideals containing I. This establishes
the inclusion
II {pip :2 I; P is a prime ideal} ~ .JT.
The reverse inclusion follows readily upon noting that if there exists a prime
ideal which contains I but not a, then a ¢ .JI,
since no power of a belongs
to P.
CERTAIN RADICALS OF A RING 165
As with the case of the Jacobson radical, the prime radical may be
characterized by its elements; this is brought out by a result promised
earlier.
Corollary. The prime radical of a ring R coincides with the nil radical
of R; that is, Rad R is simply the ideal of all nilpotent elements of R.
Proof The assertion is all but obvious upon taking I'= {O} in Theorem
8-8.
An immediate consequence of this last corollary is the potentially
powerful statement: every nil ideal of R is contained in the prime radical,
not simply contained in the larger Jacobson radical (Corollary 3 to Theorem
8-2).
Example 8-5. For an illustration of Theorem 8-8, let us fall back on the
ring Z of integers. In this setting, the nontrivial prime ideals are the principal
ideals (p), where P is a prime number. Given n > 1, the ideal (n) ~ (p) if
and only if P divides n; this being so,
.J(nj = n (pJ
pdn
Thus, if we assume that n has the prime power factorization
n = p~'p~2 ... p~r (k; E Z+),
it follows that
.J(nj = (Pi) n (P2) n ... n (Pr) = (PiP2 ... Pr)·
Let us go back to Theorem 8-8 for a moment. Another of its advantages
is that it permits a rather simple characterization of semiprime ideals.
(The reader is reminded that we defined an ideal I to be semiprime provided
that I = J1).
Theorem 8-9. An ideal I of the ring R is a semiprime ideal if and only
if I is an intersection of prime ideals of R.
Proof The proof is left to the reader; it should offer no difficulties.
Corollary. The prime radical Rad R is a semiprime ideal which is
contained in every semi prime ideal of R.
Before pressing on, we should also prove the prime radical counterpart
of Theorem 8-3.
Theorem 8-10. For any ring R, the quotient ring R/Rad R is without
prime radical.
Proof For clarity of exposition, set I = Rad R. Suppose that a + I is
any nilpotent element of R/I. Then, for some positive integer n,
(a + I)n = an + I = I,
166
Lemma. If e and e' are two idempotent elements of the ring R such that
e - e' E Rad R, then e = e'.
Proof Inasmuch as the product (e - e')(1 - (e + e'») = 0, it is enough
to show that 1 - (e + e') is an invertible element of R. Now, one may
write 1 - (e + e') in the form
1 - (e + e') = (1 - 2e) + (e - e'),
where(1 - 2e)2 = 1 - 4e + 4e 2 = 1. The implication is that 1
- (e + e'),
being the sum of a nilpotent element and an invertible element, is necessarily
invertible in R (Problem 10, Chapter 1).
Corollary. Let I, J be ideals of the ring R with I ~ J ~ Rad R. If the
idempotents of RjJ can be lifted into R, then so can the idempotents of
RjI.
Proof Suppose that u + I is any idempotent of RjI. Since I ~ J, it follows
that u + J is an idempotent element of RjJ; u 2 - U E I ~ J. By assump-
tion, there must exist some e2 = e in R such that e + J = u + J, whence
e - u E J. But then
(e + I) - (u + I) = e - u + IE JjI ~ (Rad R)jI = Rad (RjI).
Applying the lemma to the quotient ring RjI, we conclude that the coset
u + I = e + I and so the idempotents of RjI can be lifted.
The key to showing that the idempotents of RjRad R are liftable is the
circumstance that certain quadratic equations have a solution in the prime
radical of R.
Theorem 8-14. For any ring R, the idempotents of RjRad R can be
lifted into R.
Proof Let u + Rad R be an idempotent element of RjRad R, so that
u2 - u = r E Rad R. The problem is to find an idempotent e E R with
e - u E Rad R or, putting it another way, to obtain a solution a of the
equation (u + a)2 = u + a, with a E Rad R.
We first set a = x(1 - 2u), where x is yet to be determined. Now, the
requirement that the element u + a = u + x(1 - 2u) be idempotent is
equivalent to the equation
(x 2 - x)(1 + 4r) + r = O.
By the quadratic formula, this has a formal solution
X=Hl-~)
= Ij2(2r - (i)r 2 + (~)r3 - ... ).
CER TAIN RADICALS OF A RING 169
Since r is nilpotent (being a member of Rad R), the displayed series will
terminate in a finite number of steps; the result is a perfectly meaningful
polynomial in r with integral coefficients. Thus, the desired idempotent is
e = u + x(1 - 2u), where x E Rad R
Corollary. For any nil ideal I of R, the idempotents of RII can be lifted.
Proof Because I ~ Rad R, an appeal to the last lemma (with J = Rad R)
is legitimate.
Let us define a ring R to be primary whenever the zero ideal is a primary
ideal of R. This readily translates into a statement involving the elements
of R: R is a primary ring if and only if every zero divisor of R is nilpotent.
Integral domains are examples of primary rings. In general, primary rings
can be obtained by constructing quotient rings RIQ, where Q is a primary
ideal of R.
As an application of the preceding ideas, we shall characterize such
rings in terms of minimal prime ideals. (A prime ideal is said to be a minimal
prime ideal if it is minimal in the set of prime ideals; in a commutative ring
with identity, such ideals are necessarily proper.) The crucial step in the
proof is the corollary on page 164.
Theorem 8-15. A ring R is a primary ring if and only if R has a minimal
prime ideal which contains all zero divisors.
Proof For the first half of the proof, let R be a primary ring. Then the set
of zero divisors of R, along with zero, coincides with the ideal N of nilpotent
elements, and N will be prime. Being equal to the prime radical of R, N is
necessarily contained in every prime ideal of R; that is to say, N is a minimal
prime ideal.
The converse is less obvious; in fact, it is easiest to prove the contra-
positive form of the converse. Suppose, then, that R has a minimal prime
ideal P which contains all zero divisors and let a E R be any nonnilpotent
element. We define the set S by
S = {ranlr¢P; n ~ O}.
S is easily seen to be closed under multiplication and 1 E S. Notice, particu-
larly, that the zero element does not lie in S, for, otherwise, we would have
ran = 0 with an =1= 0; this implies that r is a zero divisor and therefore a
member of P. We now appeal to the corollary on page 164 to infer that
the complement of S contains a prime ideal P'. Since pi ~ R - S ~ P,
with P being a minimal prime ideal, it follows that R - S = P. But a E S,
whence a ¢ P, so that a cannot be a zero divisor of R. In other words, every
zero divisor of R is nilpotent, which completes the proof.
For the sake of refinement, let us temporarily drop the assumption that
all rings must have a multiplicative identity (commutativity could also be
170
abandoned, but this seems unnecessarily elaborate for the purposes in mind).
To make things more specific, we pose the problem of constructing a radical
which will agree with the Jacobson radical when an identity element is
available. Of course, it is always possible to imbed a ring in a ring with
identity, but the imbedding is often unnatural and distorts essential features
of the given ring.
The direct approach of considering the intersection of all maximal ideals
is not very effective, because one no longer knows that such ideals exist
(Theorem 5-2, our basic existence theorem for maximal ideals, clearly
requires the presence of an identity). A more useful clue is provided by
Theorem 8-2, which asserts that an element a E rad R if and only if 1 - ra
is invertible for every choice of r in R, or, to put it somewhat differently,
the principal ideal (1 - ra) = R for all r E R. This latter condition can be
written as {x - raxlx E R} = R for each r in R, and is meaningful in the
absence of a multiplicative identity. It thus would appear that Theorem 8-2
constitutes a hopeful starting point to the solution of the problem before
us. Needless to say, it will be necessary to introduce concepts capable of
replacing the notions of an invertible element and maximal ideal which
were so essential to our earlier work.
One begins by associating with each element a E R the set
la = {ax - XIXER}.
A moment's thought shows la to be an ideal of R. Now, it may very well
happen that la = R; in this event, we shall say that a is a quasi-regular
element. There is another way of looking at quasi-regularity:
Theorem 8-16. An element a of the ring R is quasi-regular if and only
if there exists some b E R such that
a + b - ab = 0.
The element b satisfying this equation is called a quasi-inverse of a.
Proof Suppose that a is quasi-regular, so that la = R. Since a E la, we
must have a = ab - b for suitable b in R, whence a + b - ab = o.
On the other hand, if there exists some element b E R satisfying
a + b - ab = 0, then a E la. Thus, for any r in R, ar E la. By virtue of the
definition of la' we also have ar - r E la' which implies that
r = ar - (ar - r) E la.
This means R ~ la' or rather R = la' and a is quasi-regular.
Corollary. An element a E R is quasi-regular if and only if a E la.
Here are some consequences: Every nilpotent element of R is quasi-
regular. Indeed, if an = 0, a straightforward calculation will establish that
CER T AIN RADICALS OF A RING 171
a b = a + b - abo
0
which lar = R for every r E R; recasting this in terms of the notion of quasi-
regularity:
Definition 8-4. The J-radical J(R) of a ring R, with or without an
identity, is the set
J(R) = {a E Rlar is quasi-regular for all r E R}.
If J(R) = {O}, then R is said to be a J-semisimple ring.
To reinforce these ideas, let us consider several examples.
Example 8-7. The ring Ze of even integers is J-semisimple. For, suppose
that the integer n E J(Ze). Then, in particular, n2 is quasi-regular. But the
equation
n2 + x - n2 x = 0,
or, equivalently, n2 = (n 2 - l)x, has no solution among the even integers
unless n = O. This implies that J(Ze) = {O}.
Example 8-8. Any commutative regular ring R is J-semisimple. Indeed,
given a E J(R), there is some a' in R such that a2 a' = a. Now, aa' must be
quasi-regular, so we can find an element x E R satisfying
aa' +x- aa'x = O.
Multiplying this equation by a and using the fact that a2 a' = 0, we deduce
that a = 0, whence J(R) = {O}.
Example 8-9. Consider the ring F[[xJ] of formal power series over the
field F. As we know, F[[ xJ] is a commutative ring with identity in which
an element f(x) = L akxk is invertible if and only if ao =1= o. If f(x) belongs
to the principal ideal (x), thenf(x) has zero constantterm; hence, (1 - f(X))-l
exists in F[[xJJ. Takingf(x), = (1 - f(x)t 1, we see thatf(x) f(x)' = O.
0
Thus, every member of(x) is quasi-regular, which implies that (x) s;; J(F[[ xJ]).
On the other hand, any element not in (x) is invertible and therefore cannot
be in the J-radical. (In general, if a E J(R) has a multiplicative inverse, then
1 = aa -1 is quasi-regular; but zero is the only quasi-regular idempotent,
so that 1 = 0, a contradiction.) The implication is that J(F[[ xJ]) s;; (x)
and equality follows:
J(F[[ xJ]) = (x) = rad F[[ xJJ.
Turning once again to generalities, let us show that any element a E J(R)
is itself quasi-regular. Since ar is quasi-regular for each choice of r in R,
it follows that a2 in particular will be quasi-regular. Therefore, we can
obtain an element b E R for which a2 b = O. But a simple computation
0
shows that
ao((-a)ob) = (ao(-a))ob = a2 0b = 0
CERTAIN RADICALS OF A RING 173
One finds in this way that the element a is quasi-regular with quasi-inverse
(-a)ob.
It is by no means apparent from Definition 8-3 that J(R) forms an ideal
of R; our next concern is to establish that this is actually the case.
Theorem 8-17. For any ring R, the J-radical J(R) is an ideal of R.
Proof Suppose that the element a E J(R), so that for any choice of x E R,
ax is quasi-regular. If r E R, then certainly (ar)y = a(ry) must be quasi-
regular for all y in R, and therefore ar E J(R).
It remains to show that whenever a, b E J(R), then the difference a - b
lies in J(R). Given x, u, v E R, a fairly routine calculation establishes the
identity
(a - b)x 0 (u 0 v) = a(x - ux) 0 v + (-bx 0 u) - (-bx 0 u)v.
Taking stock of the fact that a, b belong to J(R), we can select an element u
such that -(bx) u = b( -x) u = 0 and a second element v for which
0 0
Proof If the element a ¢ J(R), then there is some x E R such that ax is not
quasi-regular. Corollary 1 asserts the existence of a modular maximal ideal
of R which excludes ax and, in consequence, does not contain a.
Corollary 3. An element a E R is qlJasi-regular if and only if, for each
modular maximal ideal M of R, there exists an element b such that
a 0 bE M.
Proof The indicated condition is clearly necessary, for it suffices to take
the quasi-inverse of a as the element b. Suppose now that the condition is
satisfied, but that a does not possess a quasi-inverse. Then the modular
idealla = {ar - rlr E R} will be contained in some modular maximal ideal
M of R. By assumption, we can find an element b in R for which a 0 b E M.
But ab - bEla' whence
a = a0 b + (ab - b) E M.
Now, let e be an identity element for R modulo M. Then there exist suitable
i E M, r E R and an integer n for which
e = i + ra + na.
As J(R) forms an ideal of R, the sum ra + na E J(R), so that e - i E J(R).
176
PROBLEMS
Unless indicated to the contrary, all rings are assumed to be commutative with identity .
1. Describe the Jacobson radical of the ring Z. of integers modulo n. [Hint: Consider
the prime factorization of n.] In particular, show that Z. is semisimple if and
only if n is a square-free integer.
2. Prove that F[ x], the ring of polynomials in x over a field F, is semisimple.
PROBLEMS 177
3. Prove that rad R is the largest (in the set-theoretic sense) ideal 1 of R such that
1 + a is invertible for all a E 1.
4. a) Let the ring R have the property that all zero divisors lie in fad R. If(a) = (b),
show that the elements a and b must be associates.
b) Verify that if the element a E rad R and ax = x for some x E R, then x = o.
5. Provethatapowerseriesf(x) = ao + a1x + ... +a"xn + ... belongstoradR[[x]]
if and only if its constant term ao belongs to rad R.
6. Prove the following assertions concerning semisimple rings:
a) A ring R is semisimple if and only if a =1= 0 implies that there exists some element
r E R for which 1 - ra is not invertible.
b) Every commutative regular ring is semisimple.
c) Suppose that {Ii} is a family of ideals of R such that R/li is semisimple for
each i and Illi = {a}. Then R itself is a semisimplering. [Hint: Theorem 8-5.]
7. If R = Rl (9 R2 (9 ... (9 Rn is the direct sum of a finite number of rings Ri
(i = 1, 2, ... , n), prove that
12. Establish that an ideal 1 of R contains a prime ideal if and only if for each n,
a1a2 ... an = 0 implies that ak E 1 for some k. [Hint: The set
S = {b 1 b2 ••• bnlbk¢J; n ~ 1}
is closed under multiplication and 0 ¢ S.]
13. Show that the prime radical of a ring R contains the sum of all nilpotent ideals of R.
14. Establish the equivalence of the statements below:
a) {O} is the only nilpotent ideal of R;
b) R is without prime radical; that is, Rad R = {O};
c) for any ideals 1 and J of R, lJ = {OJ implies that 1 Il J = {OJ.
178
16. a) Prove that Rad R is the maximal nil ideal of R (maximal among the set of nil
ideals); this property is often taken as the definition of the prime radical of R.
b) If R has no nonzero nil ideals, deduce that the polynomial ring R[x] is semi-
simple.
c) Let char R = n > 0. Prove that if R is without prime radical, then n is a
square-free integer. [H int: Assume that n = p2 q for some prime p; then there
exists an element a E R such that pqa =I 0, but (pqa)2 = 0.]
17. Prove that the following statements are equivalent:
a) R has a unique proper prime ideal;
b) R is a local ring with rad R = Rad R;
c) every noninvertible element of R is nilpotent;
d) R is a primary ring and every noninvertible element of R is either a zero divisor
or zero.
18. Supply a proof of Theorems 8-11 and 8-12.
19. A ring R is termed a Hilbert ring if each proper prime ideal of R is an intersection
of maximal ideals. Prove that:
a) R is a Hilbert ring if and only if for every proper ideal I of R,
rad (R/I) = Rad (R/I).
b) Any homomorphic image of a Hilbert ring is again a Hilbert ring.
c) If the polynomial ring R[x] is a Hilbert ring, then R is one also. [Hint: Utilize
(b) and the fact that R[ x ]/(x) ~ R.]
20. If I is an ideal of the ring R, prove each of the following statements:
a) The nil radical of I is the intersection of all the minimal prime ideals of I.
b) Rad R is the intersection of all the minimal prime ideals of R.
c) The union of all the minimal prime ideals of R is the set
RELATIONS
for every a, b, c E S, (1) a '" a, (2) a '" b implies b '" a, (3) a '" band b '" c
together imply a '" c. We say that a is equivalent to b if and only if a '" b
holds.
In the following set of miscellaneous examples, we leave to the reader
the task of verifying that each relation described actually is an equivalence
relation.
Example A-I. Let S be a nonempty set and define, for a, b E S, a '" b if
and only if a = b; that is, a and b are the same element. (Technically
speaking, '" is the subset {(a, a)la E S} of S x S.) Then '" satisfies the
requirements of Definition A-2, showing that the equivalence concept is a
generalization of equality.
Example A-2. In the Cartesian product R# x R#, let (a, b) '" (c, d) signify
that a - c and b - d are both integers. A simple calculation reveals that
"', so defined, is an equivalence relation in R # X R # .
Example A-3. As another illustration, consider the set L of all lines in a
plane. Then the phrase "is parallel to" is meaningful when applied to the
elements of L and may be used to define a relation in L. If we agree that
any line is parallel to itself, this yields an equivalence relation.
Example A-4. Let f: X --. Y be a given mapping. Take for '" the relation
a '" b if and only if f(a) = f(b); then '" is an equivalence relation in X,
called the equivalence relation associated with the mapping/. More generally,
if fF is an arbitrary family offunctions from X into Y, an equivalence relation
can be introduced in X by interpreting a '" b to meanf(a) = f(b) for every
f E fF. (The underlying feature in the latter case is that any intersection
of equivalence relations in X is again an equivalence relation, for
'" = nfe, {(a, b)lf(a) = f(b)}.)
One is often led to conclude, incorrectly, that the reflexive property is
redundant in Definition A-2. The argument proceeds like this: if a '" b,
then the symmetric property implies that b '" a; from a '" band b '" a,
together with the transitivity of "', it follows that a '" a. Thus, there appears
to be no necessity for the reflexive condition at all. The flaw in this reasoning
lies, of course, in the fact that, for some element a E S, there may not exist
any b in S such that a '" b. As a result, we would not have a '" a for every
member of S, as the reflexive property requires.
Any equivalence relation '" determines a separation of the set S into
a collection of subsets of a kind which we now describe. For each a E S,
let [a] denote the subset of S consisting of all elements which are equivalent
to a:
[a] = {b E Sib", a}.
This set [a] is referred to as the equivalence class determined by a. (The
reader should realize that, in general, the equivalence class [a] is the same
RELA TIONS 289
as the class [a'] for many elements a' E S.) As a notational device, let us
henceforth use the symbol Sj'" to represent the set of all equivalence classes
of the relation '" ; that is,
Sj", = {[a]ja E S}.
Some of the basic properties of equivalence classes are listed in the
theorem below.
Theorem A-I. Let", be an equivalence relation in the set S. Then,
1) for each a E S, the class [a] =1= 0;
2) if b E [a], then [a] = [b]; in other words, any element of an
equivalence class determines that class;
3) for all a, bE S, [a] = [b] if and only if a '" b;
4) for all a, bE S, either [a] n [b] = 0 or [a] = [b];
5) U [a] = S.
aeS
Proof. Clearly, the element a E [a], for a '" a. To prove the second
assertion, let bE [a], so that b '" a. Now, suppose that x E [a], which
means x '" a. Using the symmetric and transitive properties of "', we thus
obtain x '" b, whence x E [b]. Since x is an arbitrary member of [a], this
establishes the inclusion [a] £; [b]. A similar argument yields the reverse
inclusion and equality follows. As regards (3), first assume that [a] = [b];
then a E [a] = [b] and so a '" b. Conversley, if we let a '" b, then the
element a E [b]; hence, [a] = [b] from (2).
To derive (3), suppose that [a] and [b] have an element in common,
say, C E [a] n [b]. Statement (2) then informs us that [a] = [c] = [b].
In brief, if [a] n [b] =1= 0, then we must have [a] = [b]. Finally, since
each class [a] £; S, the inclusion u {[a]la E S} £; S certainly holds. For
the opposite inclusion, it is enough to show that each element a E S belongs
to some equivalence class; but this is no problem, for a E [a].
As evidenced by Example A-4, any mapping determines an equivalence
relation in its domain. The following corollary indicates that every
equivalence relation arises in this manner; that is to say, each equivalence
relation is the associated equivalence relation of some function.
Corollary. Let", be an equivalence relation in the set S. Then there
exists a set T and a mapping g: S ~ T such that a '" b if and only if
g(a) = g(b).
Proof. Simply take T = Sj'" and g : S ~ T to be the mapping defined by
g(a) = [a]; in other words, send each element of S onto the (necessarily
unique) equivalence class to which it belongs. By the foregoing theorem,
a '" b if and only if [a] = [b], or equivalently, g(a) = g(b).
290
Example A-5. This example is given to illustrate that any mapping can be
written as the composition of a one-to-one function and an onto function.
Letf: X - Ybe an arbitrary mapping and consider the equivalence relation
- associated withf If the element a e X, then we have
[a] = {b e Xif(b) = f(a)} = f-l(J(a».
In effect, the equivalence classes for the relation - are just the inverse
images off-l(y), where yef(X) £; Y.
Now, define the function!: Xj- - Yby the ruleJ([a]) = f(a). Since
[a] = [b] ifand only iff(a) = f(b),Jis well-defined. Observe that whereas
the original function f may not have been one-to-one, J happens to be
one-to-one; indeed, J([ a]) = J([b]) implies that f(a) = feb), whence
[a] = [b]. At this point, we introduce the onto function g: X - Xj- by
setting g(a) = [a]. Then f(a) = 1([a]) = J(g(a») = (J 0 g)(a) for all a in
X, in consequence of which f = log. This achieves our stated aim.
We next connect the notion of an equivalence relation in S with that
of a partition of S.
Definition A-3. By a partition of the set S is meant a family r!J of subsets
of S with the properties
1) 0 ¢ r!J,
2) for any A, B e r!J, either A = B or A n B = 0 (pairwise disjoint),
3) u r!J = S.
Expressed otherwise, a partition of S is a collection r!J of nonempty
subsets of S such that every element of S belongs to one and only one member
of r!J. The set Z of integers, for instance, can be partitioned into the subsets
of odd and even integers; another partition of Z might consist of the sets
of positive integers, negative integers and {O}.
Theorem A-I may be viewed as asserting that each equivalence relation
- on a set S yields a partition of S, namely, the partition Sj - into the
equivalence classes for -. (In this connection, notice that, for the equiva-
lence relation of equality, the corresponding classes contain only one element
each; hence, the resulting partition is the finest possible.) We now reverse
the situation and show that a given partition of S induces an equivalence
relation in S. But first a preliminary lemma is required.
Lemma. Two equivalence relations - and -' in the set S are the same
if and only if Sj - and Sj -' are the same.
Proof If - and -' are the same, then surely Sj- = Sj-'. So, suppose
that - and -' are distinct equivalence relations in S. Then there exists
a pair of elements a, b e S which are equivalent under one of the relations,
RELA TIONS 291
but not under the other; say a '" b, but not a ",' b. By Theorem A-I,
there is an equivalence class in S/ '" containing both a and b, while no such
class appears in S/"". Accordingly, S/'" and S/",' differ.
Theorem A-2. If &J is a partition of the set S, then there is a unique
equivalence relation in S whose equivalence classes are precisely the
members of &J.
Proof Given a, b E S, we write a '" b if and only if a and b both belong to
the same subset in &J. (The fact that &J partitions S guarantees that each
element of S lies in exactly one member of &J.) The reader may easily check
that the relation "', defined in this way, is indeed an equivalence relation
in S.
Let us prove that the partition &J has the form S/ "'. If the subset P E &J,
then a E P for some a in S. Now, the element b E P if and only if b '" a,
or, what amounts to the same thing, if and only if b E [a]. This demonstrates
the equality P = [a] E S/ "'. Since this holds for each P in &J, it follows
that &J s;;; S/ "'. On the other hand, let [a] be an arbitrary equivalence
class and P be the partition set in &J to which the element a belongs. By
similar reasoning, we conclude that [a] = P; hence, S/'" s;;; &J. Thus, the
set of equivalence classes for '" coincides with the partition &J. The
uniqueness assertion is an immediate consequence of the lemma.
To summarize, there is a natural one-to-one correspondence between
the equivalence relations in a set and the partitions of that set; every
equivalence relation gives rise to a partition and vice versa. We have a
single idea, which has been considered from two different points of view.
Another type of relation which occurs in various branches of mathe-
matics is the so-called partial order relation. Just as equivalence generalizes
equality, this relation (as we define it below) generalizes the idea of "less
than or equal to" on the real line.
Definition A-4. A relation R in a nonempty set S is called a partial
order in S if the following three conditions are satisfied:
1) aRa (reflexive property),
2) if aRb and bRa, then a = b (antisymmetric property),
3) if aRb and bRc, then aRc (transitive property),
where a, b, c denote arbitrary elements of S.
From now on, we shall follow custom and adopt the symbol ::;; to
represent a partial order relation, writing a ::;; b in place of aRb; the fore-
going axioms then read: (1) a ::;; a, (2) if a ::;; band b ::;; a, then a = b,
and (3) if a ::;; b and b ::;; c, then a ::;; c. As a linguistic convention, let us
also agree to say (depending on the circumstance) that "a precedes b" or
292
ZORN'S LEMMA
In this Appendix, we give a brief account of some ofthe axioms of set theory,
with the primary purpose of introducing Zorn's Lemma. Our presentation
is descriptive and most of the facts are merely stated. The reader who is
not content with this bird's-eye view should consult [12J for the details.
As we know, a given partially ordered set need not have a first element
and, if it does, some subset could very well fail to possess one. This prompts
the following definition: a partially ordered set (S, :::;;) is said to be well-
ordered if every nonempty subset A !;;; S has a first element ("with respect
to :::;; .. being understood). The set Z + is well-ordered by the usual :::;;; each
nonempty subset has a first element, namely, the integer of smallest
magnitude in the set.
Notice that any well-ordered set (S, :::;; ) is in fact totally ordered. For,
each subset {a, b} !;;; S must have a first element. According as the first
element is a or b, we see that a :::;; b or b :::;; a, whence the two elements
a and b are comparable. Going in the other direction, any total ordering
of a finite set is a well-ordering of that set. Let it also be remarked that a
subset of a well-ordered set is again well-ordered (by the restriction of the
ordering).
asserting the existence of a set S with the property that A (') S contains
exactly one element, for each A in ~.
Granting Zermelo's Theorem, it is clear that a choice function f can
be defined for any collection ~ of nonempty sets: having well-ordered u ~,
simply take f to be the function which assigns to each set A in ~ its first
element. As indicated earlier, Zermelo's Theorem was originally derived
from the axiom of choice, so that these are in reality two equivalent principles
(although seemingly quite different).
Although the axiom of choice may strike the reader as being intuitively
obvious, the sou:ldness of this principle has aroused more philosophical
discussion than any other single question in the foundations of mathematics.
At the heart of the controversy is the ancient problem of existence. Some
mathematicians believe that a set exists only if each of its elements can be
designated specifically, or at the very least if there is a rule by which each
of its members can be constructed. A more liberal school of thought is that
an axiom about existence of sets may be used if it does not lead to a contra-
diction. In 1938 Godel demonstrated that the axiom of choice is not in
contradiction with the other generally accepted axioms of set theory
(assuming that the latter are consistent with one another). It was sub-
sequently established by Cohen (1963) that the denial of this axiom is also
consistent with the rest of set theory. Thus, the axiom of choice is in fact
an independent axiom, whose use or rejection is a matter of personal
inclination. The feeling among most mathematicians today is that the axiom
of choice is harmless in principle and indispensable in practice (provided
that one calls attention to the occasions of its use). It is also valuable as
an heuristic tool, since every proof by means of this assumption represents
a result for which we can then seek proofs along other lines.
A non-constructive criterion for the existence of maximal elements is
given by the so-called "maximality principle", which generally is cited in
the literature under the name Zorn's Lemma. (From the point of view of
priority, this principle goes back to Hausdorff and Kuratowski, but Zorn
gave a formulation of it which is particularly suitable to algebra; he was also
the first to state, without proof, that a maximality principle implies the
axiom of choice.)
Zorn's Lemma. Let S be a nonempty set partially ordered by :::;; .
Suppose that every subset A S;; S which is totally ordered by :::;; has an
upper bound (in S). Then S possesses at least one maximal element.
Zorn's Lemma is a particularly handy tool when the underlying set is
partially ordered and the required object of interest is characterized by
maximality. To demonstrate how it is used in practit:e, let us prove what is
sometimes known as Hausdorff's Theorem (recall that by a chain is meant
a totally ordered set):
ZORN'S LEMMA 299
GENERAL REFERENCES
Our purpose here is to present a list of suggestions for collateral reading and
further study. The specialized sources will carry the reader considerably
beyond the point attained in the final pages of this work.
18. JACOBSON, N., Structure of Rings, Rev. Ed. Providence: American Mathematical
Society, 1964.
19. JANS, J., Rings and Homology. New York: Holt, 1964.
20. KUROSH, A., General Algebra. New York: Chelsea, 1963.
21. LANG, S., Algebra. Reading, Mass.: Addison-Wesley, 1965.
22. LAMBEK, J., Lectures on Rings and Modules. Waltham, Mass.: Blaisdell, 1966.
23. MCCARTIlY, P., Algebraic Extensions of Fields. Waltham, Mass.: Blaisdell, 1966.
24. McCoy, N., Rings and Ideals. (Carus Monographs). Menascha, Wis.: Mathematical
Association of America, 1948.
25. McCoy, N., Theory of Rings. New York: Macmillan, 1964.
26. NAGATA, M., Local Rings. New York: Interscience, 1962.
27. NORTIlCOTT, D. G., Ideal Theory. Cambridge, England: Cambridge University Press,
1953.
28. NORTIlCOTT, D. G., Lessons on Rings, ModulesandMultiplicities. Cambridge, England:
Cambridge University Press, 1968.
29. REDEl, L., Algebra, Vo!.1. Oxford, England: Pergamon, 1967.
30. SAH, C.-H., Abstract Algebra. New York: Academic Press, 1967.
31. WARNER, S., Modern Algebra, 2 Vols. Englewood Cliffs, New Jersey: Prentice-Hall,
1965.
32. WEISS, E., Algebraic Number Theory. New York: McGraw-Hill, 1963.
33. ZARISKI, O. and P. SAMUEL, Commutative Algebra, Vo!' I. Princeton: Van Nostrand,
1958.
JOURNAL ARTICLES
34. BROWN, B. and N. McCoY, "Radicals and Subdirect Sums," Am. J. Math. 69, 46-58
(1947).
35. BUCK, R. C., "Extensions of Homorphisms and Regular Ideals," J. Indian Math. Soc.
14, 156-158 (1950).
36. COHEN, I. S., "Commutative Rings with Restricted Minimum Condition," Duke Math.
J. 17,27-42 (1950).
37. DIVINSKY, N., "Commutative Subdirectly Irreducible Rings," Proc. Am. Math Soc.
8,642-648 (1957).
38. FELLER, E., "A Type of Quasi-Frobenius Rings," Canad. Math. Bull. 10, 19-27 (1967).
39. GIFFEN, c., "Unique Factorization of Polynomials," Proc. Am. Math. Soc. 14, 366
(1963).
40. HENDERSON, D., "A Short Proof of Wedderburn's Theorem," Am. Math. Monthly 72,
385-386 (1965).
41. HERSTEIN, I. N., "A Generalization of a Theorem of Jacobson, I," Am. J. Math. 73,
756-762 (1951).
42. HERSTEIN, I. N., "An Elementary Proof of a Theorem of Jacobson," Duke Math. J.
21,45-48 (1954).
43. HERSTEIN, I. N., "Wedderburn's Theorem and a Theorem of Jacobson," Am. Math.
Monthly 68, 249-251 (1961).
44. JACOBSON, N., "The Radical and Semi-Simplicity for Arbitrary Rings," Am. J. Math.
67, 300-320 (1945).
45. KOHLS, C., "The Space of Prime Ideals of a Ring," Fund. Math. 45,17-27 (1957).
302 BIBLIOGRAPHY
46. KOVACS, L., "A Note on Regular Rings," Publ. Math. Debrecen 4, 465-468 (l956).
47. LUH, J., "On the Commutativity of J-Rings," Canad. J. Math. 19, 1289-1292 (l967).
48. McCoy, N., "Subdirectly Irreducible Commutative Rings," Duke Math. J.12, 381-387
(1945).
49. McCoy, N., "Subdirect Sums of Rings," Bull. Am. Math. Soc. 53, 856-877 (1947).
50. McCoy, N., "A Note on Finite Unions of Ideals and Subgroups," Proc. Am. Math.
Soc. 8, 633--637 (1957).
51. NAGATA, M., "On the Theory of Radicals in a Ring," J. Math. Soc. Japan 3,330-344
(l951).
52. VON, NEUMANN, J., "On Regular Rings," Proc. Natl. Acad. Sci. U.S. 22, 707-713 (1936).
53. NORTHCOTT, D., "A Note on the Intersection Theorem for Ideals," Proc. Cambridge
Phil. Soc. 48,.366-367 (1952).
54. PERLIS, S. "A Characterization of the Radical of an Algebra," Bull. Am. Math. Soc.
48, 128-132 (1942).
55. SAMUEL, P., "On Unique Factorization Domains," Illinois J. Math. 5, 1-17 (1961).
56. SATYANARAYANA, M., "Rings with Primary Ideals as Maximal Ideals," Math. Scand.
20, 52-54 (1967).
57. SATYANARAYANA, M., "Characterization of Local Rings," Tohoku Math. J.19, 411-416
(l967).
58. SNAPPER, E., "Completely Primary Rings, I," Ann. Math. 52, 666-693 (1950).
59. STONE, M. R., "The Theory of Representations of Boolean Algebras," Trans. Am.
Math. Soc. 40, 37-111 (l936).
INDEX OF SPECIAL SYMBOLS
The following is by no means a complete list of all the symbols used in the
text, but is rather a listing of certain symbols which occur frequently.
Numbers refer to the page where the symbol in question is first found.
ABCDE79876S43210