You are on page 1of 318
A First Course in RINGS AND IDEALS DAVID M. BURTON University of New Hampshire rN vv ADDISON-WESLEY PUBLISHING COMPANY Reading, Massachusetts - Menlo Park, California - London - Don Mills, Ontario This book is in the ADDISON-WESLEY SERIES IN MATHEMATICS Consulting Editor: Lynn H. Loomis Standard Book Number 201.00731.2 AMS 1968 Subject Classifications 1610, 1620. Copyright © 1970 by Addison-Wesley Publishing Company, Inc. Philippines copyright 1970 by Addison-Wesley Publishing Company, Ine. All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise, without the prior written permission of the publisher. Printed in the United States of America. Published simultaneously in Canada. Library of Congress Catalog Card No. 73-100855. To my Father Frank Howard Burton PREFACE As the title suggests, this volume is designed to serve as an introduction to the basic ideas and techniques of ring theory; it is intended to be an expository textbook, rather than a treatise on the subject. The mathe- matical background required for a proper understanding of the contents is not extensive. We assume that the average reader has had some prior contact with abstract algebra, but is still relatively inexperienced in this respect. In consequence, nearly everything herein can be read by a person familiar with basic group-theoretic concepts and having a nodding acquain- tance with linear algebra. The level of material should prove suitable for advanced undergraduates and beginning graduate students. Indeed, a built-in flexibility permits the book to be used, either as the basic text or to be read independently by interested students, in a variety of situations. The reader whose main interest is in ideal theory, for instance, could chart a course through Chapters 2, 3, 5,8, 11, 12, 13. Taken as a whole, the present work is more nearly a begin- ning than an end. Our hope is that it may serve as a natural point of departure for the study of the advanced treatises on ring theory and, in some aspects of the subject, the periodical literature. As regards treatment, our guiding principle is the strong conviction that intelligibility should be given priority over coverage; that a deeper under- standing of a few important topics is preferable to a superficial knowledge of many. This calls for a presentation in which the pace is unhurried and which is complete in the details of proof, particularly of basic results. By adhering to the “theorem-proof” style of writing, we hope to achieve greater clarity (perhaps at the sacrifice of elegance). Apart from the general know- ledge presupposed, an attempt has been made to keep the text technically self-contained, even to the extent of including some material which is undoubtedly familiar. The mathematically sophisticated reader may prefer to skip the earlier chapters and refer to them only if the need arises. At the end of each chapter, there will be found a collection of problems of varying degrees of difficulty. These constitute an integral part of the book and require the reader’s active participation. They introduce a variety v vi PREFACE of topics not treated in the body of the text, as well as impart additional information about material covered earlier; some, especially in the later chapters, provide substantial extensions of the theory. We have, on the whole, resisted the temptation to use the exercises to develop results that will subsequently be needed (although this is not hard and fast). Those problems whose solutions do not appear straightforward are often accom- panied by hints. The text is not intended to be encyclopedic in nature; many fascinating aspects of this subject vie for inclusion and some choice is imperative. To this end, we merely followed our own tastes, condensing or omitting entirely a number of topics that might have been encompassed by a book of the same title. Despite some notable omissions, the coverage should provide a firm foundation on which to build. A great deal of valuable criticism was received in the preparation of this work and our moments of complacence have admitted many improvements, Of those students who helped, consciously or otherwise, we should like particularly to mention Francis Chevarley, Delmon Grapes, Cynthia Kennett, Kenneth Lidman, Roy Morell, Brenda Phalen, David Smith, and John Sundstrom; we valued their critical reading of sections of the manu- script and incorporated a number of their suggestions into the text. It is a pleasure, likewise, to record our indebtedness to Professor James Clay of the University of Arizona, who reviewed the final draft and offered helpful comments leading to its correction and improvement. We also profited from many conversations with our colleagues at the University of New Hampshire, especially Professors Edward Batho, Homer Bechtell, Robb Jacoby, and Richard Johnson. In this regard, special thanks are due Pro- fessor William Witthoft, who was kind enough to read portions of the galleys; his cagle-eyed attention saved us from embarrassment more than ‘once. We enjoyed the luxury of unusually good secretarial help and take this occasion to express our appreciation to Nancy Buchanan and Solange Larochelle for their joint labors on the typescript. To my wife must go the largest debt of gratitude, not only for generous assistance with the text at all stages of development, but for her patience and understanding on those occasions when nothing would go as we wished. Finally, we should like to acknowledge the fine cooperation of the staff. of Addison-Wesley and the usual high quality of their work. The author, needless to say, must accept the full responsibility for any shortcomings and errors which remain. Durham, New Hampshire D.M.B. January 1970 Chapter Chapter Chapter Chapter Chapter Chapter Chapter Chapter Chapter 1 2 3 4 5 6 7 8 ka Chapter 10 Chapter 11 Chapter 12 Chapter 13 Introductory Concepts . Ideals and Their Operations ‘The Classical Isomorphism Theorems Integral Domains and Fields Maximal, Prime, and Primary Ideals . Divisibility Theory in Integral Domains. Polynomial Rings Certain Radicals of a Ring . Two Classic Theorems Direct Sums of Rings Rings with Chain Conditions Further Results on Noetherian Rings . Some Noncommutative Theory Appendix A. Relations . Appendix B, Zorn’s Lemma Bi jiography . Index of Special Symbols Index vii CONTENTS 16 39 52 1 90 M2 157 180 204 217 234 262 287 296 300 303 305 CONVENTIONS Here we shall set forth certain conventions in notation and terminology used throughout the text: the standard symbols of set theory will be employed, namely, €, U, 4, —, and @ for the empty set. In particular, A — B= {x\xe A and x¢B}. As regards inclusion, the symbols S and = mean ordinary inclusion between sets (they do not exclude the possibility of equality), whereas c and = indicate proper inclusion. When we deal with an indexed collection of sets, say {A;|i€ 1}, the cumbersome notations vu {AjjieT} and 9 {A,|ie 1} will generally be abbreviated to U A, and 1 A;; it being understood that the operations are always over the full domain on which the index is defined. Following custom, {a} denotes the set whose only member is a. Provided that there is no risk of confusion, a one-element set wil] be identified with the element itself. A function f (synonymous with mapping) is indicated by a straight arrow going from domain to range, as in the case f: X > Y, and the notation always signifies that f has domain X. Under these circumstances, f is said to be a function on X, or from X, into ¥. In representing functional values, we adopt the convention of writing the function on the left, so that f(x), or occasionally fx, denotes the image of an element x ¢ X. The restriction of f to a subset A of X is the function f|A from A into Y defined by (s|A)@) = fl for all x in A. For the composition of two functions fX > Yand g: Y= Z, we will write g of; that is, g of: X > Z satisfies (9° f)(x) = 9( f(x) for each x eX, (It is important to bear in mind that our policy is to apply the functions from right to left.) Some knowledge of elementary number theory is assumed. We simply remark that the term “prime number” is taken to mean a positive prime; in other words, an integer n > 1 whose only divisors are +1 and +n. Finally, let us reserve the symbol Z for the set of all integers, Z, for the set of positive integers, Q for the set of rational numbers, and R* for the set of real numbers. viii ONE 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 nonempty 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 0. The unique additive inverse of an element ae 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) + ¢ and not for a-(b + c). Because of the general associative law, parentheses 2 FIRST COURSE IN RINGS AND IDEALS 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 : Dat+b=b+a, 2jatb+ecsat+bt+o, 3) there exists an element 0 in R such that a + 0 = a for everyae R, 4) for each a € R, there exists an element —a e R such that a + (—a) = 0, 5) (a-b):¢ = a-(b-c), and 6) a(b + c) = ab + arc and (b + ca = bea + 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, be R. (In case a:b = b-a 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-1 = 1-a = aforallaeR. Given a ring (R, +, -) with identity 1, an element ae 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~', so that a-a~' = a~'-a = 1. Inthe 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+.% (+. (R*, 45°) 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 Example1-2 Let X bea given set and P(X) be the collection of all subsets of X. The symmetric difference of two subsets A, B < X is the set A A B, where AAB=(A— B)U(B- A). If we define addition and multiplication in P(X) by A+B=AAB, a eee ee then the system (P(X), +, -) forms a commutative ring with identity. The empty set & 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, 9) nor (P(X), -, V) constitutes a ring. Example 1-3. Given a ring (R, +, ), we may consider the set M,(R) of n x n matrices over R. If I, = {1,2,...,n}, a typical member of M,(R) is a function f:1, x I,» R. In practice, one identifies such a function with its values a, = f(i,), which are displayed as the n x n rectangular array yy +++ Gig : (a, € R). 4, 4, Into Gn For the sake of simplicity, let us hereafter abbreviate the n x n matrix whose (i, /) entry is ay to (a,). The operations required to make (M,(R), +, °) a ring are provided by the familiar formulas (a,j) + (by) = @j + by) and —(a;)-(6y) = (Cys where Cig = YD Andy. we (We shall often indulge in this harmless duplication of symbols whereby + and - are used with two different meanings.) The zero element of the resulting ring is the n x n matrix all of whose entries are 0; and ~(a;;) = (—a,;). The ring (M,(R), +, *) fails to be commutative for n > 1. It is equally easy to show that if (R, +,-) has an identity element 1, then the matrix with 1’s down the main diagonal (that is, a, = 1) and 0's elsewhere will act as identity for matrix multiplication. In terms of the Kronecker delta symbol 6,,, which is defined by iti) | Aorta the identity matrix can be written concisely as (6,,) i = 1,2,...,0, 4 FIRST COURSE IN RINGS AND IDEALS Example 1-4. To develop our next example, let X be an arbitrary (non- empty) set and (R, +,:) be a ring. We adopt the notation map(X, R) for the set consisting of all mappings from X into R; in symbols, map(X, R) = {f|f: X > R}. (For ease of notation, let us also agree to write map R in place of map(R, R).) Now, the elements of map(X, R) can be combined by performing algebraic operations on their functional values. More specifically, the pointwise sum and product of f and g, denoted by f + g and f-g, respectively, are the functions which satisfy (fF + ) = S%) + 9%), (Fax) = f)-), (EX). It is readily verified that the above definitions provide map(X, R) with the structure of a ring. We simply point out that the zero element of this ring is the constant function whose sole value is 0, and the additive inverse —f of fis characterized by the rule (—f)(x) = —f(x). Notice that the algebraic properties of map(X, R) are determined by what happens in the ring (R, +, -) (the set X furnishes only the points for the pointwise operations). For instance, if (R, +,+) has a multiplicative identity 1, then the ring (map(X, R), +, -) likewise possesses an identity element; namely, the constant function defined by I(x) = 1 for all x € X. Example 1-5. Our final example is that of the ring of integers modulo n, where n is a fixed positive integer. In order to describe this system, we first introduce the notion of congruence: two integers a and b are said to be congruent modulo n, written a = b (mod n), if and only if the difference a — b is divisible by n; in other words, a = b (mod n) if and only if a — b = kn for some ke Z. We leave the reader to convince himself that the relation “congruent modulo n”’ defines an equivalence relation on the set Z of integers. As such, it partitions Z into disjoint classes of congruent elements, called congruence classes. For each integer a, let the congruence class to which a belongs be denoted by [a]: [a] = {xe Z|x = a (mod n)} {a + knlke Z}. Of course, the same congruence class may very well arise from another integer; any integer a’ for which [a’] = [a] is said to be a representative of [a]. One final, purely notational, remark : the collection of all congruence classes of integers modulo n will be designated by Z,. It can be shown that the congruence classes [0], [1], ...,[” — 1] exhaust the elements of Z,. Given an arbitrary integer a, the division algorithm asserts that there exist unique q,r¢Z, with 0 < r 1 is composite; if n = nynz in Z 0 < n,n, 0. With the definition a° = 1, the symbol a" now has a well-defined meaning for every integer n (at least when attention is restricted to invertible elements). Paralleling the exponent notation for powers, there is the notation of integral multiples of an element a ¢ R. For each positive integer n, we define the nth natural multiple na recursively as follows: la=a and na=(n—1)a+a, when n> 1. If it is also agreed to let Oa = 0 and (—n)a = —(na), then the definition of na can be extended to all integers. Integral multiples satisfy several identities which are easy to establish: (n + m)a = na + ma, (nm)a = n(ma), na + b) = na + nb, for a, b € R and arbitrary integers n and m. In addition to these rules, there are two further properties resulting from the distributive law, namely, n(ab) = (na)b = a(n), and -—(na)(mb) = (nm)(ab). Experience impels us to emphasize that the expression na should not be regarded as a ring product (indeed, the integer n may not-even be a member of R); the entire symbol na is just a convenient way of indicating INTRODUCTORY CONCEPTS 11 a certain sum of elements of R. However, when there is an identity for multiplication, it is possible to represent na as a product of two ring elements, namely, na = (nl)a. To proceed further with our limited program, we must first frame a definition. Definition 1-5. Let R be an arbitrary ring. If there exists a positive integer n such that na = O for all ae R, then the smallest positive integer with this property is called the characteristic of the ring. If no such positive integer exists (that is,n = 0 is the only integer for which na = 0 for all a in R), then R is said to be of characteristic zero. We shall write char R for the characteristic of R. The rings of integers, rational numbers, and real numbers are all standard examples of systems having characteristic zero (some writers prefer the expression “characteristic infinity”). On the other hand, the ring P(X) of subsets of a fixed set X is of characteristic 2, since 2A=AAA=(A-A)U(A-A=G for every subset A © X. Although the definition of characteristic makes an assertion about every element of the ring, in rings with identity the characteristic is completely determined by the identity element. We reach this conclusion below. Theorem 1-5. If R is any ring with identity 1, then R has characteristic n > Of and only if n is the least positive integer for which nl = 0. Proof. Ichar R = n > 0,thenna = Ofor every a ¢ Rand so, in particular, nl = 0. Were ml = 0, where 0 < m 0 and infinite when char R = 0. Proof. To verify this assertion, suppose first that char R = n > 0. Accord- ing to the definition of characteristic, each element ‘0 # ae R will then possess a finite additive order m, with m 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 € R, verify that a) ab = 1 if and only if ba = 1, b) if a? = 1, then either a = 1 ora = 1 7. Let a, b be two elements of the ring R. Ifne Z, and a and b commute, derive the binomial expansion (+ b= Far HE Es R)abPE + Bt, r+ (arb where (i) iw = 7 is the usual binomial coefficient. 14 8. % 10. i. 12. 13. 14, ae 16. 17. 18. FIRST COURSE IN RINGS AND IDEALS An element a of a ring R is said to be idempotent if a? = a and nilpotent if a” = 0 for some ne Z,. Show that a) a nonzero idempotent element cannot be nilpotent, b) every nonzero nilpotent element is a zero divisor in R. Given that R is an integral domain, prove that a) the only nilpotent element is the zero element of R, b) the multiplicative identity is the only nonzero idempotent element. If a is a nilpotent clement of R, a ring with identity, establish that 1+ a is invertible in R. [Hint: (1 + a)" =1—a +a? +--+ (-1) 1a", where a" =0] A Boolean ring is a ring with identity every element of which is idempotent. Prove that any Boolean ring R is commutative. [Hint: First show that a = —a for every ae R] 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. 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. Assume that R is a ring and ae R. If C(a) denotes the set of all elements which commute with a, Cla) = {reRlar = ra}, show that C(a) is a subring of R. Also, verify the equality cent R = (ser C(a). Given a ring R, prove that a) if S, is an arbitrary (indexed) collection of subrings of R, then their intersection 17S, is also a subring of R; b) for a nonempty subset T of R, the set 0}, determine whether S, is a subring of R. Establish the following assertions concerning the characteristic of a ring R: 19, 20. 2. PROBLEMS aa a) if there exists an integer k such that ka = 0 for all ae R, then k is divisible by char R; b) ifchar R > 0, then char S < char R for any subring S of R; ©) if R is an integral domain and S is a subdomain of R, then char $ = char R. Let R be a ring with a finite number of elements, say ay, a2, .-., dy, and let n; be the order of a, regarded as a member of the additive group of R. Prove that the characteristic of R is the least common multiple of the integers n,(i = 1,2, ...n). Suppose that R is a ring with identity such that char R = n > 0. Ifmis not prime, show that R has divisors of zero. If R is a ring which has no nonzero nilpotent elements, deduce that all the idem- potent elements of R belong to cent R. [Hint: If a? = a, then (ara — ar)? = (ara — ra)? = 0 for allreR] . Assume that R is a ring with the property that a? + ae cent R for every element ain R. Prove that Ris necessarily a commutative ring, [Hint: Utilize the expression (@ + bP + (a + B) to show first that ab + ba lies in the center for all a, be R.] . Let (G, +) be a commutative group and R be the set of all (group) homomorphisms of G into itself. The pointwise sum f + g and composition f» g of two functions ‘fg € Rare defined by the usual rules (F + MX) =f) + a, (Fo MX) = So) EB). Show that the resulting system (R, +, °) forms a ring. At the same time determine the invertible elements of R. |. Let (G, -) bea finite group (written multiplicatively), say with elements x1, X35 ---. Xp and let R be an arbitrary ring. Consider the set R(G) of all formal sums Sn GER) Two such expressions are regarded as equal if they have the same coefficients Addition and multiplication can be defined in R(G) by taking rat Same Let ss and ¥ tx ot where suet (The meaning of the last-written sum is that the summation is to be extended over all subscripts j and k for which x,x, = x.) Prove that, with respect to these operations, R(G) constitutes a ring, the so-called group ring of G over R. TWO IDEALS AND THEIR OPERATIONS Although it is possible to obtain some interesting conclusions concerning subrings, this concept, if unrestricted, is too general for most purposes. To derive certain highly desirable results (for instance, the fundamental iso- morphism theorems), additional assumptions that go beyond Definition 1-4 must be imposed. Thus, in the present chapter we narrow the field and focus attention on a class of subrings with a stronger type of multiplicative closure, namely, closure under multiplication by an arbitrary ring element. Definition 2-1. A subring / of the ring R is said to be a two-sided ideal of R if and only if re R and aeT imply both rae I and are I. Viewed otherwise, Definition 2-1 asserts that whenever one of the factors in a product belongs to I, then the product itself must be in I. (This may be roughly summarized by saying that the set I “captures” products.) Taking stock of Theorem 1-3, which gives a minimal set of conditions to be a subring, our current definition of a two-sided ideal may be reformu- lated as follows. Definition 2-2, Let J be a nonempty subset of a ring R. Then I isa two-sided ideal of R if and only if 1) a,beT imply a — bel, and 2) re Rand ae/ imply both products ra, ar € I. If condition (2) of the above definition is weakened so as to require only that the product ra belongs to I for every choice of re R and ae I, we are led to the notion of a left ideal; right ideals are defined in a sym- metric way. Needless to say, if the ring R happens to be commutative (the most important case so far as we shall be concerned), then there is no distinction between left, right, and two-sided ideals. CONVENTION In what follows, let us agree that the term “ideal”, un- modified, will always mean two-sided ideal. Before proceeding further, we pause to examine this concept by means of several specific examples. 16 IDEALS AND THEIR OPERATIONS 17 Example 2-1. For each integer a€ Z, let (a) represent the set consisting of all integral multiples of a; that is, (a) = {na|ne Z}. The following relations confirm (a) to be an ideal of the ring of integers: na — ma = (n — ma, m(na) = (mn)a, nneZ. In particular, since (2) = Z,, the ring of even integers forms an ideal of Z. Notice, too, that (0) = {0} 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 € X, we denote by J, the set of all mappings which take on the value 0 at x: I, = {femap(X, R)| f(x) = 0}. Now, choose f, g¢ I, and he map(X, R). From the definition of the ring operations in map(X, R), (f—9)@) = F(%)—gxX) = 0-0 = 0, while (LAX) = fDh(x) = Oh{x) = 0, and, in a similar manner, (hf)(x) = 0. Thus, f — g, fh and hf all belong to I,, which implies that I, is an ideal. More generally, if $ is any nonempty subset of X, then 1 = {fe map(X, R)|f(x) = 0 for all x € S} comprises an ideal of map(X, R). Since I = (),2s[,, we have a situation where the intersection of ideals is once again an ideal. (Theorem 2-2 shows that this is no accident.) Before presenting our next example, we derive a fact which, despite its apparent simplicity, will be frequently applied in the sequel. Theorem 2-1. If / is a proper (right, left, two-sided) ideal of a ring R with identity, then no element of J possesses a multiplicative inverse; that is, I 0 R* = 2. Proof. Let I be an ideal of R and suppose that there is some member a # 0 of I such that a’ exists in R. (The theorem is trivial when I = {0}.) Since J is closed under multiplication by arbitrary ring elements, it follows that 1 = a-'ae/. By the same reasoning, I contains r = rl for every r in R; 18 FIRST COURSE IN RINGS AND IDEALS that is, R < I, whence the equality J = R. This contradicts the hypothesis that I is a proper subset of R. Notice that, en route, we have also established Corollary. In a ring with identity, no proper (right, left, two-sided) ideal contains the identity element. Example 2-3. This example is given to show that the ring M,(R*) of n x n matrices over the real numbers has no nontrivial ideals. As a nota- tional device, let us define E,, to be the n x n matrix having 1 as its ijth entry and zeroes elsewhere. Now, suppose that I + {0} is any ideal of the ring M,(R*). Then J must contain some nonzero matrix (a,,), with, say, rsth entry a,, # 0. Since J is a two-sided ideal, the product Exel ai Ess is a member of J, where the matrix (b,,) is chosen to have the element a;,* down its main diagonal and zeroes everywhere else. As a result of all the zero entries in the various factors, it is easy to verify that this product is equal to E,,. Knowing this, the relation Ey= EE Ey (bf = 1,2,..,0) implies that all n? of the matrices E,, ate contained in J. The clinching point is that the identity matrix (6,)) can be written as (6) = Eis + Baa + + + Ens which leads to the conclusion that (6,)¢J and, appealing to the above corollary, that I = M,(R”). In other words, M,(R*) possesses no nonzero proper ideals, as asserted. As a matter of definition, let us call a ring R # {0} simple if R has no two-sided ideals other than {0} and R. In the light of Example 2-4, the matrix ring M,(R”) is a simple ring. We now take up some of the standard methods for constructing new ideals from given ones. To begin with simpler things: Theorem 2-2. Let {J;} be an arbitrary collection of (right, left, two- sided) ideals of the ring R, where i ranges over some index set. Then 2 [is also a (right, left, two-sided) ideal of R. Proof. We give the proof for the case in which the ideals are two-sided. First, observe that the intersection > J, is nonempty, for each of the ideals I; must contain the zero element of the ring. Suppose that the elements a,be€l,andreR. Then a and b are members of I,, where i varies over the indexing set. INasmuch as J; is assumed to be an ideal of R, it follows that a — b, ar and ra all lie in the set J;. But this is true for every value of IDEALS AND THEIR OPERATIONS 19 i, whence the elements a — b, ar and ra belong to 4 J,, making 9 I, an ideal of R. Consider, for the moment, an arbitrary ring R and a nonempty subset Sof R. By the symbol (S) we shall mean the set (8) = 9 {| SS 1; Lis an ideal of R}. The collection of all ideals which contain S is not empty, since the entire ring itself is an ideal containing any subset of R; thus, the set (S) exists and satisfies the inclusion S ¢ (S), By virtue of Theorem 2-2, (S) forms an ideal of R, known as the ideal generated by the set S. It is noteworthy that when- ever I is any ideal of R with S © I, then necessarily (S) ¢ I. For this reason, ‘one often speaks of (S) as being the smallest ideal of R to contain the set S. It should be apparent that corresponding remarks apply to the one-sided ideals generated by S. If S consists of a finite number of elements, say a,, a3, ...,4,, then the ideal which they generate is customarily denoted by (a,, a2, ...,@,). Such an ideal is said to be finitely generated with the given elements a, as its generators. An ideal (a) generated by just one ring element is termed a principal ideal. A natural undertaking is to determine the precise form of the members of the various ideals (right, left, two-sided) generated by a single element, say a, of an arbitrary ring R. The right ideal generated by a is called a principal right ideal and is denoted by (a),. Being closed with respect to multiplication on the right, (a), necessarily contains all products ar (r € R), as well as the elements na (n an integer), and, hence, includes their sum ar + na. (As usual, the notation na represents the n-fold sum of a.) It is a fairly simple matter to check that the set of elements of the form ar + na constitutes a right ideal of R. Observe, too, that the element a is a member of the ideal, since a = a0 + Ia. These remarks make it clear that (@), = {ar + nalre R: ne Z}. When there is an identity element present, the term na becomes superfluous, for, in this setting, we may write the expression ar + na more simply as ar + na = ar + a(nl) = afr + nl) = ar’, where r’ = r + nl is some ring element. Thus, the set (a), consists of all right multiples of a by elements of R. If R is a ring with identity, we shall frequently employ the more suggestive notation aR in place of (a),; that is, (a), = aR = {arlre R}. Similar remarks apply, of course, to the principal left ideal (a), generated bya. 20 FIRST COURSE IN RINGS AND IDEALS ‘Asa general comment, observe that the products ar (r ¢ R) comprise the set of elements of a right ideal of R even when the ring does not possess an identity. The difficulty, however, is that this ideal need not contain a itself. With regard to the two-sided ideal (a) generated by a, the situation is more complicated. Certainly the elements ras, ra, as and na must all belong to the ideal (a) for every choice of r, s€ R and ne Z. In general, the sum of two elements ras and r’as’ is no longer of the same form, so that, in order to have closure under addition, any finite sum Y'r,as;, where r;, s,€ R, is also required to be in (a). ‘The reader will experience no difficulty in showing that the principal ideal generated by a is given by (a) = {na + ra + as + ¥ rasjlr, s,1r;,5,€ R3 ne Z}. finite In case R happens to have an identity, this description of (a) reduces to the set of all finite sums )' r,as;. A particularly important type of ring is a principal ideal ring, which we now define. Definition 2-3. A ring R is said to be a principal ideal ring if every ideal I of Ris of the form I = (a) for some ae R. The following theorem furnishes an example of such rings. Theorem 2-3. The ring Z of integers is a principal ideal ring; in fact, if I is an ideal of Z, then I = (n) for some nonnegative integer n. Proof. If I = {0}, the theorem is trivially true, since the zero ideal {0} is the principal ideal generated by 0. Suppose then that I does not consist of the zero element alone. Now, if me I, —m also lies in I, so that the set contains positive integers. Let n designate the least positive integer in I. As I forms an ideal of Z, each integral multiple of n must belong to I, whence acl. To establish the inclusion J < (n), let k be an arbitrary element of I. By the division algorithm there exist integers q and r for which k = gn + r, with 0 0, we would have a contradiction to the assumption that n is the smallest positive integer in I. Accordingly, r = 0 and k = qne(n). Thus, only multiples of n belong to I, implying that I < (n). The two inclusions show that J = (n) and the argument is complete. Let us now describe certain binary operations on the set of all ideals of R. (Similar considerations apply to the sets of right and left ideals, but for economy of effort we concentrate on two-sided ideals.) Given a finite number of ideals I,, I, ..., I, of the ring R, one defines their sum in the natural way: T+, +--+ 0, = {a + a, + + a,|a;€ 1}. IDEALS AND THEIR OPERATIONS 21 Then I, + I, + --- + I, is likewise an ideal of R and is the smallest ideal of R which contains every I; phrased in another way, I, + I, + --- + I, is the ideal generated by the union I, U I, U -- U I,. In the special case of two ideals J and J, our definition reduces to I+ J = {a+ blael; be J}. More generally, let {1,} be an arbitrary indexed collection of ideals of R. The sum of this collection may be denoted by ¥ J, and is the ideal of R whose members are all possible finite sums of elements from the various ideals 1: Lhe {Zadar e1,}. The reader will take care to remember “that, although {J,} may be an infinite family of ideals, only finite sums of elements of R are involved in the definition above. An alternative description of 5 I, could be given by YA, = (Lajla,e 1,; all but a finite number of the a, are 0}, where it is understood that )° represents an arbitrary sum with one or more terms. Just as 4 J; can be interpreted as the largest ideal of R contained in every I,, the sum YI, supplies the dual notion of the smallest ideal containing every I. If R = 1, + 1, + -~ + 1,, then each element x € R can be expressed in the form x = a, + a, + -+ + @,, where a; lies in J;, There is no guarantee, however, that this representation of x is unique. To ensure that every member of R is uniquely expressible as a sum of elements from the ideals J, an auxiliary definition is required. Definition 2-4, Let I,, I, ..., 1, be ideals of the ring R. We call R the internal direct sum of Iy, Ip, ..., I, and write R = I, ®@1, @ -- © I, provided that a)R=1,+1,+--+1,,and by + + eg + Tian + > + 1,) = {0} for each i. As was heralded by our remarks, we are now in a position to prove Theorem 2-4. Let I,, I;, ..., I, be ideals of the ring R. Then the following statements are equivalent: 1) Ris the internal direct sum of I), I, ---5 Ip: 2) Each element x of R is uniquely expressible in the form x= a, +a, + + +4, where a;€I,. Proof. There is no ioss in confining ourselves to the case n = 2; the general argument proceeds along similar lines. We begin by assuming that R = 1, © 1,. Suppose further that an element x € R has two representations X=a, +b) =a, +b, (a1, bey). 22 FIRST COURSE IN RINGS AND IDEALS Then a, ~ a, = b, — b,. But the left-hand side of this last equation lies in J,, while the right-hand side is in I,, so that both sides belong to 1,01; = {0}. Itfollowsthata, — a, = b, ~ by = 0,oray = az,b, = bp. In other words, x is uniquely representable as a sum a + b,aeI,,be I>. Conversely, assume that assertion (2) holds and that the element xeéI,1,. We may then express x in two different ways as the sum of an element in I, and an clement in I,; namely, x = x + 0 (here xe J, and 0eJ,) and x = 0 + x (here OI, and xe/,). The uniqueness assumption of (2) implies that x = 0, in consequence of which I, 0 I, = {0}; hence, R = I, @I,. This completes the proof of the theorem. We now come to a less elementary, but extremely useful, notion ; namely, the product of ideals. Once again, assume that I and J are two ideals of the ring R. To be consistent with our earlier definition of the sum I + J, we should define the product IJ to be the collection of all simple products ab, where ae I and be J. Unfortunately, the resulting set fails to form an ideal. (Why?) To counter this difficulty, we instead take the elements of IJ to be all possible finite sums of simple products; stated explicitly, IJ = {Y.abjla,€ 1; be J}. te With this definition IJ indeed becomes an ideal of R. For, suppose that x,y € JJ and re R; then, X = ayby + ayby + + aby, y= aby + ab, + + abn, where the a, and aj are in I, and the b, and bj are in J. From this we obtain x y= ayy to + ayby + (~ aby + + (GBs 1x = (ra,)b, + (ra,)b, + + (ra,)b,. Now, the elements —a; and ra; necessarily lie in I, so that x — y and rx € LJ; likewise, xr ¢ IJ, making IJ an ideal of R. In point of fact, IJ is just the ideal generated by the set of all products ab, ae I, be J. There is no difficulty in extending the above remarks to any finite number of ideals I, Iz, ..., 1, of the ring R. A moment’s thought shows that the product I,J, -- J, is the ideal consisting of finite sums of terms of the form a,a, --- a,, with a; in I. (It is perhaps appropriate to point out that, because of the associative law for multiplication in R, the notation 1,1, +I, 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 I” is the set of finite sums of products of n elements from I: = (9) aay ig [Oi ET} trite IDEALS AND THEIR OPERATIONS 23 In this connection, it is important to observe that T2P2P2-2r2e forms a decreasing chain of ideals. Remark, If J is a right ideal and S a nonempty subset of the ring R, then ST = {Yara S;r,e1} tite forms a right ideal of R. In particular, if $ = {a}, then al (a notation we prefer to {a}I) is given by al = {are 1}. Analogous statements can be made when J is a left ideal of R, but not, of course, a two-sided ideal. The last ideal-theoretic operation which we wish to consider is that of the quotient (or residual), defined below. Definition 2-5. Let J and J be two ideals of the ring R. The right (left) quotient of I by J, denoted by the symbol I ;,J(I J), consists of all elements ae R such that aJ R’ such that S(a + b) = fla) + fb, flab) = ff) for every pair of elements a,b R. A homomorphism which is also one-to-one as a map on the underlying sets is called an isomorphism. We emphasize that the + and - occurring on the left-hand sides of the equations in Definition 2-7 are those of R, whereas the + and - occurring on the right-hand sides are those of R’. This use of the same symbols for the operations of addition and multiplication in two different rings should cause no ambiguity if the reader attends closely to the context in which the notation is employed. If fis a homomorphism of R into R’, then the image f(R) of R under f will be called the homomorphic image of R. When R = R’, so that the two rings are the same, we say that f is a homomorphism of R into itself. In this connection, a homomorphism of R into itself is frequently referred to as an endomorphism of the ring R or, if an isomorphism onto R, an auto- morphism of R. 26 FIRST COURSE IN RINGS AND IDEALS 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 R and 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=04+0= f(a) + f(b), Slab) = 0 = 00 = fa f(b) so that fis 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 Z, of integers modulo n. Define f: Z > Z, by taking f(a) = [a]; that is, map each integer into the congruence class containing it. That fis a homomorphism follows directly from the definition of the operations in Z,,: f(a + b) = [a + b) = [a] +, [6] = £@ +,.£0), S(ab) = [ab] = [4]-,[6] = f@-,f0). Example 2-6. In the ring map(X, R), define 1, to be the function which assigns to each fe map(X, R) its value at a fixed element ae X; in other words, t,(f) = f(a). Then z, is a homomorphism from map(X, R) into R, known as the evaluation homomorphism at a. We need only observe that taf + 9) = (Ff + aha) = f(@) + gla) = t.(f) + 229), tJ) = (fava) = fla)g(a) = 2,(P)ta(g)- We now list some of the structural features preserved under homo- morphisms. (a beR), Theorem 2-7, Let f be a homomorphism from the ring R into the ring R’. Then the following hold: 1) f) = 0, 2) f(-a@) = —f@ forall ae R. If, in addition, R and R’ are both rings with identity and f(R) = R’, then 3) f(Q) = ‘f(a)~* for each invertible element ae R. 4) fla IDEALS AND THEIR OPERATIONS 27 Proof. From f(0) = (0 + 0) = f(0) + f(0), we obtain f(0) = 0. The fact that f(a) + f(—a) = f(a + (—a)) = f(0) = 0 yields f(—a) = —f(a). As regards (3), let the element a € R satisfy f(a) = 1; then,f(1) = f(@f() = flat) = f(a) = 1. Finally, the equation f(a) f(a~") = f(aa~') = f(1) = 1 shows that f(a)! = f(a"), whenever a € R has a multiplicative inverse. Two comments regarding part (3) of the above theorem are in order. First, it is evident that S@1 = f(a) = flal) = fas) for any a in R. Knowing this, one might be tempted to appeal (incorrectly) to the cancellation law to conclude that f(1) = 1; what is actually required is the fact that multiplicative identities are unique. Second, if the hypothesis that f map onto the set R’ is omitted, then it can only be inferred that f(1) is the identity for the homomorphic image f(R). The element f(1) need not serve as an identity for the entire ring R’ and, indeed, it may very well happen that f(1) # 1. We also observe, in passing, that, by virtue of statement (2), f(a — b) = f(a) + f(—b) = f(a) — Fe). In short, any ring homomorphism preserves differences as well as sums and products, The next theorem indicates the algebraic nature of direct and inverse images of subrings under homomorphisms. Among other things, we shall see that if fis a homomorphism from the ring R into the ring R’, then f(R) forms a subring of R’. The complete story is told below. Theorem 2-8, Let f be a homomorphism from the ring R into the ring R.. Then, 1) for each subring S of R, f(S) is a subring of R’; and 2) for each subring S’ of R’, f~1(S’) is a subring of R. Proof. To obtain the first part of the theorem, recall that, by definition, the image f(S) = { f(a)|a € S}. Now, suppose that f(a) and f(b) are arbitrary elements of f(S). Then both a and b belong to the set S, as do a — b and ab (S being a subring of R). Hence, f(a) — f(b) = fla — be f(S) F(a f(b) = flab) f(S). According to Theorem 1-3, these are sufficient conditions for f(S) to be a subring of R’. The proof of the second assertion proceeds similarly. First, remember that f-(S') = {ae R|f(a) eS}. Thus, if a, be f~4(S'), the images f(a) and and 28 FIRST COURSE IN RINGS AND IDEALS f(b) must be members of S’. Since S’ is a subring of R’, it follows at once that Sa — b) = fla) — fibjes' flab) = f@fibyeS’. This means that a — b and ab lie in f~"(S’), from which we conclude that f~ 4S) 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 I’ is an ideal of R’, then the subring f~'(I’) is an ideal of R. For instance, suppose that aef~ (I), so that f(a) eI’, and let r be an arbitrary element of R. Then, fra) = f() f(@ € I’; in other words, the product ra is in f~ (I). Likewise, ar € f~3(I’), which helps to make f~*(’) an ideal of R Without further restriction, it cannot be inferred that the image f(I) will be an ideal of R’, whenever J is an ideal of R. One would need to know that r'f(@)ef(D for all re R’ and ae J. 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: and Corollary. 1) For each ideal J’ of R’, the subring f~*(I') is an ideal of R. 2) If f(R) = R’, then for each ideal J of R, the subring f() is an ideal of 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 of f, denoted by ker f, consists of those elements in R which are mapped by f onto the zero element of the ring R’: kerf = {ae R|f(a) = 0}. Theorem 2-7 indicates that ker fis a nonempty subset of R, since, if nothing else, Oc 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 kerf 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 {0} forms an ideal of R’. Since ker f = f~ (0), the conclusion follows from the last corollary. IDEALS AND THEIR OPERATIONS 29 The kernel of a homomorphism may be viewed as a measure of the extent to which the homomorphism fails to be one-to-one (hence, fails to be an isomorphism). In more concrete terms, we have Theorem 2-10. A homomorphism f from a ring R into a ring R’ is an isomorphism if and only if ker f = {0}. Proof. First, if fis a one-to-one function and f(a) = 0 = f(0), then a = 0, whence ker f = {0}. On the other hand, suppose that the kernel consists exactly of 0. Iff(a) = f(b), then f(a — b) = f(a) — fb) = 0, which means thata — be kerf. Sinceker f = {0},wemusthavea — b = 0, or a = b, making f a one-to-one function. Two rings R and R’ are said to be isomorphic, denoted by R ~ R’, if there exists an isomorphism from the ring R onto the ring R’. Although this definition is unsymmetric in that it makes mention of a function from one particular ring to another, let us remark that iff: R + R’ isa one-to-one, onto, homomorphic mapping, the function f~ ~ R also enjoys these properties. We may therefore ignore the apparent lack of symmetry and merely speak of two rings R and R’ as being isomorphic without specifying ‘one ring as isomorphic to the other ; notationally, this situation is recognized by writing either R ~ R’ or R' =~ R. Isomorphic rings are indistinguishable from the structural point of view, even though they may differ in the notation for and nature of their elements and operations, Two such rings, although not in general formally identical, are the same for all purposes; the underlying feature is the existence of a mapping which transports the algebraic structure of one ring to the other. In practice, we shall often identify isomorphic rings without explicit mention. This seems to be a natural place to insert an example. Example 2-7. Consider an arbitrary ring R with identity and the mapping fi Z + R given by f(n) = nl. (At the risk of being repetitious, let us again emphasize that nl means the n-fold sum of 1.) A simple computation shows that f, so defined, is a homomorphism from the ring Z of integers into the ring R: f(a + m) = (n + mL = nl + mi = fin) + fm) f(nm) = (nm)1 = n(ml) = (nl) (nl) = fr) f(r). Since ker f constitutes an ideal of Z, a principal ideal ring, it follows that kerf = {neZ|nl = 0} = (p) for some nonnegative integer p. A moment's reflection should convince the reader that the integer p is just the characteristic of R. In particular, and 30 FIRST COURSE IN RINGS AND IDEALS any ring R with identity which is of characteristic zero will contain a subring isomorphic to the integers; more specifically, Z ~ Z1, 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 I of the ring R determines an ideal f(J) 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 J and J are two ideals of R with Ic J © I + kerf, then f(I) = f(J). The quickest way to see this is to notice that I) Sf) S fl + kerf) =f) + fkerf) = f(D, 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 = {0} or else narrowing our view to consider only ideals I with kerf ¢ I. In either event, it follows that 1 S J ¢ I + kerf = J and, in consequence, I = J. The first of the restrictions just cited has the effect of making the function f one-to-one, in which case R and 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. Let f be a homomorphism from the ring R onto the ring R’. If J is any ideal of R such that ker f © I, then I = f~*(f(I)). Proof. Suppose that the element a ¢ f-'({I)), so that fa) € fl). Then Sa) = f(r) for some choice ofr in I. Asa result, we will have f(a — r) = 0, or, what amounts to the same thing, a — reker f < I. This implies that ael, yielding the inclusion f~*(f(I)) < 1. Since the reverse inclusion always holds, the desired equality follows. Here now is one of the main results of this section. Theorem 2-11. (Correspondence Theorem). Let f be a homomorphism from the ring R onto the ring R’.. Then there is a one-to-one correspon- dence between those ideals J of R such that ker f © J and the set of all ideals 1’ of R’; specifically, I’ is given by I’ = f(I). Proof. Our first concern is to show that the indicated correspondence actually maps onto the set of all ideals of R’. In other words, starting with an ideal J’ of R’, we must produce some ideal I of the ring R, with ker f SI, IDEALS AND THEIR OPERATIONS 31 such that f(1) = I’. To accomplish this, it is sufficient to take I = f~1(I'. By the corollary to Theorem 2-8, f}(I’) certainly forms an ideal of R and, since Oe I’, kerf =f) Sf") Inasmuch as the function f is assumed to be an onto map, it also follows that f() = f(f-)) = I Next, we argue that this correspondence is one-to-one. To make things more specific, let ideals I and J of R be given, where ker f < I, kerf ¢ J, and satisfying f() = f(J). From the elementary lemma just established, we see that T=f" (f(D) =f (FW) = J. One finds in this way that the correspondence I + f(J), where kerf < I, is indeed one-to-one, completing the proof. Before announcing our next result, another definition is necessary. Definition 2-9. A ring R is said to be imbedded in a ring R’ if there exists some subring S’ of R’ such that R ~ S', In general, if a ring R is imbedded in a ring R’, then R’ is referred to as. an extension of R and we say that R can be extended to R'’. The most important cases are those in which one passes from a given ring R to an extension possessing some property not present in R. Asa simple applica- tion, let us prove that an arbitrary ring can be imbedded in an extension ring with identity. Theorem 2-12. (Dorroh Extension Theorem). Any ring R can be im- bedded in a ring with identity. Proof. Consider the Cartesian product R x Z, where RXZ = {(r, mre Rj ne Z}. If addition and multiplication are defined by (a, n) + (6, m) (a, n)(b, m) then it is a simple matter to verify that Rx Z forms a ring; we leave the actual details as an exercise. Notice that this system has a multiplicative identity, namely, the pair (0, 1); for (a, n)(0, 1) = (a0 + 1a + n0, nl) = (a,n), (a + bn + m), (ab + ma + nb, nm), and, similarly, (0, D(@, n) = (a, n). 32 FIRST COURSE IN RINGS AND IDEALS Next, consider the subset R x {0} of R x Z consisting of all pairs of the form (a, 0). Since (a, 0) — (60) (a, 0)(b, 0} (a — 6,0), (ab, 0), it is evident that R x {0} constitutes a subring of Rx Z. A straightforward calculation, which we omit, shows that Rx {0} is isomorphic to the given ring R under the mapping f: R > Rx {0} defined by f(a) = (a,0). This process of extension therefore imbeds R in R x Z, a ring with identity. A point to be made in connection with the preceding theorem is that the imbedding process may be carried out even if the given ring has an identity to start with. Of course, in this case the construction has no particular merit ; indeed, the original identity element only serves to introduce divisors of zero into the extended ring. Although Theorem 2-12 shows that we could confine our study to rings with identity, it is nonetheless desirable to develop as much of the theory as possible without the assumption of such an element. Thus, unless an explicit statement is made to the contrary, the subsequent discussions will not presuppose the existence of a multiplicative identity. We now take a brief look at a different problem, namely, the problem of extending a function from a subring to the entire ring. In practice, one is usually concerned with extensions which retain the characteristic features of the given function. The theorem below, for instance, presents a situation in which it is possible to extend a homomorphism in such a way that the extended function also preserves both ring operations. Theorem 2-13. Let J be an ideal of the ring R and f a homomorphism from I onto R’, a ring with identity. If I cent R, then there is a unique homomorphic extension of f to all of R. Proof. As a start, we choose the element ue J so that f(u) = 1. Since I constitutes an ideal of R, the product au will lie in the set J for each choice of a€R. It is therefore possible to define a new function g: R R’ by setting g(a) = f(au) for all a in R. If the element a happens to belong to J, then g(a) = flav) = fa fl) = f@! = f@, showing that g actually extends the original function f. The next thing to confirm is that both ring operations are preserved by g. The case of addition is fairly obvious: if a, b ¢ R, then g(a + b) = f(a + bu) = f(au + bu) = Sau) + (bu) = gfa) + g(b). IDEALS AND THEIR OPERATIONS 33 Asa preliminary step to demonstrating that g also preserves multiplication, notice that Sf ((abju?) = flabu) fu) = f(abu). From this we are able to conclude that gab) = flabu) = flabu2) = f (au)(bu)) Flau)f(bu) = g(ag(6). The crucial third equality is justified by the fact that ue cent R, hence, commutes with b. As regards the uniqueness assertion, let us assume that there is another homomorphic extension of f to the set R; call it h. Since fand h must agree on J and, more specifically, at the element u, h(u) = f(u) = 1. With this in mind, it follows that Ha) = h(ayh(u) = h(au) = f(au) = g(a) for all ac R and so hand g are the same function. Hence, there is one and only one way of extending f homomorphically from the ideal J 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,,R,...,R, be a finite number of rings (not necessarily subrings of a common ring) and consider their Cartesian product R = x R, consisting of all ordered n-tuples (@,, 43, ...,4,), with a,¢R,. One can easily convert R into a ring by performing the ring operations componentwise; in other words, if (a,, ay, ...,,) and (by, by, ..., b,) are two elements of R, simply define (Ag, gy sey dy) + Bs Bay 225 Bg) = (Qs + yy dy + Bay e254, + b,) and (yy ay ++ 5 Oq)(B ys Bay +o» Bn) = (@yD1, Gaba, --+ > Gnby). The ring so obtained is called the external direct sum of Ry, Ry, ..-, Ry and is conveniently written R = R, + R, +--+ R,. (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 R, (i = 1,2,..., n), then the individual R, need not be subrings, or even subsets, of R. However, there is an ideal of R which is the isomorphic image of R;. A straightforward calculation will convince the reader that the set T, = {(, ...,0,4, 0, .., Oa; Ry} 34 FIRST COURSE IN RINGS AND IDEALS (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 R, under the mapping which sends (0, ..., 0, a;, 0, ..., 0) to the element a;. Since (4, @y, «+25 G,) = (ay, 0,0, ...,0) + 0, a2, 0, ...,0) + + + 0,0,..., 0, a,), it should also be clear that every member of R is uniquely representable as a sum of elements from the ideals I,. Taking note of Theorem 2-4, this means that R is the internal direct sum of the ideals I, and so R,¢R,¢--4R,=1,0Lh0-Ol, R~ 1). In summary, the external direct sum R of the rings R,, Rz, -.., R, is also the internal direct sum of the ideals I,, I, ..:,1, and, for cach i, R, and I, 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 @-notation. The term “Snternal” 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 ting R is said to be a direct summand of R if there exists another ideal J of R such that R = I ® J. For future use, let us note that should the ideal I happen to have an identity element, say the element ¢¢ J, then it will automatically be a direct summand of R.. The argument proceeds as follows. For any choice of r€ R, the product ree J. The assumption that e serves as an identity for J then ensures that e(re) = re. At the same time (and for the same reasons), (erJe = er. Combining these pieces, we get re = ere = er, which makes it plain that the element ¢ lies in the center of R. This is the key point in showing that the set J = {r — re|r ¢ 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 J and J. Certainly, each element r of R may be written as r= re + (r — re), where reel andr — ree J. Since I 1 J = {0}, this is the only way r can be expressed as a sum of elements of J and J. (A moment’s thought shows that ifaeI 7 J,saya = r — re,thena = ae = (r — ree = r(e — e”) = 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 sct this result in evidence, we first imbed R in the extension ring R’ = RxZ in the standard way (see Theorem 2-12). Then, R = Rx {0}, where, as is easily verified, R x {0} 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 @ 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, 0) + (e,n), for some re R. The last-written equation tells us that e = —r and n = —1; what is important is the resulting conclusion that (e, —1) J. For arbitrary r € R, the product (r, 0)(e, —1) = (re — r, 0) will consequently be in both R and J (each being an ideal of R’). The fact that Rn J = {0} forces (re — r,0) = (0,0); hence, re =r. Ina like fashion, we obtain er = r, proving that R admits the element e as an identity. PROBLEMS 1. If Fis a right ideal and J a left ideal of the ring R such that I 9 J = {0}, prove that ab = O for all ae I, be J. 2, Given an ideal J of the ring R, define the set C(J) by C(l) = {re Rlra ~ are for all ae R}. Verify that C(J) 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{U;} @ = 1,2, ...)isa collection of ideals of the ring R such that I, < I, ¢ j and strictly upper triangular if a, = 0 for i > j. Let T,(R) and T3(R) denote the sets of all upper triangular and strictly upper triangular matrices in M,(R), respectively. Prove each of the following: a) T,(R) and T3(R) are both subrings of M,(R). b) 73(R) is an ideal of the ring 7,(R). ©) A matrix (a;,) € 7,(R) is invertible in 7,(R) if and only if a, is invertible in R for i = 1,2,...,n. [Hint: Induct on the order n.] d) Any matrix (a,,) € T3(R) is nilpotent; in particular, (a,,)" = 5. Let I be an ideal of R, a commutative ring with identity. For an element ae R, the ideal generated by the set I u {a} is denoted by (I, a). Assuming that a ¢ I, show that (a) = {i + raliel,reR). 6. In the ring Z of integers consider the principal ideals (n) and (m) generated by the integers n and m. Using the notation of the previous problem, verify that {(@), m) = ((), n) = (r) + Gm) = (1, m) = @), where d is the greatest common divisor of n and m. 36 FIRST COURSE IN RINGS AND IDEALS 7. Suppose that I is a left ideal and J a right ideal of the ring R. Consider the set I = {Za,b,a,€1; be J}, where = represents a finite sum of one or more terms. Establish that 1J is a two- sided ideal of R and, whenever J and J are themselves two-sided, that IJ < Iq J. 8. If S is any given nonempty subset of the ring R, then ann,$ = {re Rlar = 0 for all ac S} is called the right annihilator of $ (in R); similarly, ann,S = {re Rira = 0 for all ae S} is the left annihilator of S. When R is a commutative ring, we simply speak of the annihilator of S and use the notation ann S. Prove the assertions below: a) ann,$ (ann, 5) is a right (left) ideal of R. b) If S isa right (left) ideal of R, then ann, (ann S) is an ideal of R. ©) If Sis an ideal of R, then ann,S and ann, are both ideals of R. 4d) When R has an identity element, ann,R = ann,R = {0}. 9. Let 1,,13,...,1, be ideals of the ring R with R that this sum is direct if and only if a, + a, + that each a; = 0. I+ lt+-- +1, Show + a, = 0, with a, €J,, implies 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 © X, P(Y) and P(X — Y) are both principal ideals of P(X), with P(X) = P(Y) ® P(X — ¥). 11. Suppose that R is a commutative ring with identity and that the element ae R isan idempotent different from 0 or 1. Prove that R is the direct sum of the principal ideals (a) and (1 — a). 12, Let I, J and K be ideals of the ring R. Prove that a) WJ + K)= 1 + IK, (1 + JK = IK + JK; b) fl 2 J, then IN + K)=J + (oR). 13, Establish that in the ring Z, if ! = (12) and J = (21), then T+J=@0, [oJ=@) U=Q5, EJ=@, J1=(. [Hint: In general, (a):(b) = (c), where = a/ged (a, B).] 14, Given ideals I and J of the ring R, verify that a) 0:1 = ann,J, and 0:,1 = ann, I (notation as in Problem 8); b) I,J (I%,J) is the largest ideal of R with the property that (I;,J)J R defined by f,(x) = axa~' is an automorphism of R. Let f be a homomorphism from the ring R into itself and S be the set of elements that are left fixed by f; in symbols, S = {ae R|f(@ = a}. Establish that S forms a subring of R. If fis a homomorphism from the ring R into the ring R’, where R has positive characteristic, verify that char f(R) < char R. Let fbe 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+ J) =f) +f); b) SU) = SIS); o) flo J) = fl) 0 f(), with equality if either I 2 ker for J 2 kerf; d) fd) = f(D:f(), with equality if 1 > kerf Show that the relation R ~ R’ is an equivalence relation on any set of rings. Let R be an arbitrary ring. For each fixed element a R, define the left-multiplica- tion function T,: R + R by taking T,(x) = ax. If Tz 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) Ty forms a ring, where multiplication is taken to be functional composition; ©) the mapping f(a) = T, determines 2 homomorphism of R onto the ring Ty; d) the kernel of fis the ideal ann, R; ©) if for each 0 # ae R, there exists some b € R such that ab # 0, then R ~ Ty. (In particular, part(e) holds whenever R has an identity element.) Let R be an arbitrary ring and Rx Z be the extension ring constructed in Theorem 2-12, Establish that a) Rx {0} isan ideal of Rx Z; b) Z~ (O}xZ; ©) if. is an idempotent element of R, then the pair (—a, 1) is idempotent in Rx Z, while (a, 0) is a zero divisor.

You might also like