Download as pdf or txt
Download as pdf or txt
You are on page 1of 198

Ring

and
Field
Theory
B1948 Governing Asia

This page intentionally left blank

B1948_1-Aoki.indd 6 9/22/2014 4:24:57 PM


Ring
and
Field
Theory
Kaiming Zhao
Wilfrid Laurier University, Canada

World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TAIPEI • CHENNAI • TOKYO
Published by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE

British Library Cataloguing-in-Publication Data


A catalogue record for this book is available from the British Library.

RING AND FIELD THEORY


Copyright © 2022 by World Scientific Publishing Co. Pte. Ltd.
All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means,
electronic or mechanical, including photocopying, recording or any information storage and retrieval
system now known or to be invented, without written permission from the publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy
is not required from the publisher.

ISBN 978-981-125-577-9 (hardcover)


ISBN 978-981-125-578-6 (ebook for institutions)
ISBN 978-981-125-579-3 (ebook for individuals)

For any available supplementary material, please visit


https://www.worldscientific.com/worldscibooks/10.1142/12819#t=suppl

Printed in Singapore

Yumeng - 12819 - Ring and Field Theory.indd 1 2/3/2022 4:29:44 pm


March 18, 2022 9:45 amsart-9x6 12819-main page v

Preface

Topics included: rings, subrings, quotient rings and ring homo-


morphisms; field of quotients of an integral domain, ideal theory,
isomorphism theorems; unique factorization domains, principal ideal
domains, Euclidean domains and Gaussian integers; polynomial rings
over unique factorization domains, Schönemann-Eisenstein Irreduci-
bility Criterion for unique factorization domains, Perron’s Criterion,
Cohn’s Criterion and Osada’s Criterion; Noetherian rings, modules
over rings, free modules, finitely generated modules over Euclidean
domains, Smith normal form of matrices over Euclidean domains;
fields, field extensions, algebraic closure, finite fields; splitting ex-
tension fields, separable extension fields, perfect fields, finite normal
extension fields; the Fundamental Theorem of Galois Theory and
solvability by radicals.

Overview and approach: In addition to being an important branch


of mathematics in its own right, ring and field theory (Galois theory)
is now an essential tool in number theory, geometry, topology, Lie
groups, algebraic geometry, differential equations.
The main object of study in Galois theory are roots of single vari-
able polynomials. Many ancient mathematics civilizations (Babylo-
nian, Egyptian, Greek, Chinese, Indian, Persian) knew how to solve
quadratic equations. Today, most middle school students memorize
the “quadratic formula” by heart. While various incomplete meth-
ods for solving cubic equations were developed in the ancient world, a
general “cubic formula” (as well as a “quartic formula”) was not known
until the 16th century Italian school. It was proven by Ruffini and
Abel in 1824, that the roots of the general quintic polynomial could
not be solvable in terms of nested roots. Galois theory provides a
satisfactory explanation for this. More generally, Galois theory is all
about symmetries of the roots of polynomials. An essential concept
is the field extension generated by the roots of a polynomial, called
the splitting field of a polynomial. The philosophy of Galois theory
has also impacted other branches of higher mathematics (Lie groups,

v
March 18, 2022 9:45 amsart-9x6 12819-main page vi

vi RING AND FIELD THEORY

topology, number theory, algebraic geometry, differential equations).


This book will provide a rigorous proof-based modern treatment of
the main results of ring and field theory (Galois theory).

About this book: This book was originally intended as a textbook


for a one-term senior undergraduate (or graduate) course in ring and
field theory, or Galois theory. The students are required to have some
knowledge on calculus, linear systems, determinants and matrices,
and to have taken a first course on abstract algebra. This book can
also serve as a reference for professional mathematicians. Earlier
drafts of this book were used several times when the author taught
MA475 (and MA675), the third course in abstract algebra at Wilfrid
Laurier University. When the author prepared these lecture notes he
mainly took [F] as his reference.
The author tries to make this book self-contained. Readers will be
only required to have some knowledge of first and second year uni-
versity math background. The book contains 241 carefully selected
exercise questions of varying difficulty which will allow students to
practice their own computational and proof-writing skills. Sample
solutions to some exercise questions are provided, from which stu-
dents can learn to approach and write their own solutions and proofs.
Besides standard ones, some of the exercises are new and very inter-
esting. Some are rather hard. It is not a surprise if the reader cannot
solve some of the exercises, particularly for first learners.

Feature of this book: The book is written in a way that is easy


to understand, simple and concise with simple historic remarks to
show the beauty of algebraic results and algebraic methods. This
makes the book a small one which can help students build up their
confidence so that they can easily pick up big volumes to learn in
their future academic career. The book provides a lot of interesting
examples that illustrate definitions, theorems, and methods, and help
students to learn how to approach and write correct and good proofs.

A guide for the instructor: This book was originally intended as


a textbook for a one-term (36 lectures of 50-minute, or 24 lectures
of 80-minute) senior undergraduate or graduate algebra course. For
a one-term course an instructor can cover Chapters 1–4 in the order
of the book, or can cover Chapters 1, 2, 4, 5 and 6 in the order of
March 18, 2022 9:45 amsart-9x6 12819-main page vii

PREFACE vii

the book and consider some sections and some theorems as optional
material (for example, Sections 1.8, 2.5, 6.5 and 6.6; the proof of
Theorem 4.3.11; Theorems 4.3.13, 4.4.9, 4.4.10, 5.4.16 and 6.2.1; the
second half of Sections 5.4 and 6.4). For more in-class examples
an instructor can take some exercise questions with solutions from
the book. Of course, these are just guides, and an instructor will
certainly want to customize the materials in the book to fit his/her
own interests and requirements.

A guide for the student:

(1). Always attend to lectures. Class provides informal discus-


sions, and you will profit from comments of your classmates,
as well as gain confidence by providing your insights and in-
terpretations of a topic. Don’t be absent!
(2). Ask and answer questions during the lecture: Is it correct? I
do not get it. Can you explain it?
(3). Take a fresh look at the notes on the topic once the lec-
ture of a particular topic has been given. Well understand
definitions, theorems, corollaries and examples, and try to
memorize them.
(4). Try to solve more exercises questions in the textbook. Write
detailed full solutions to more exercise questions in your own
words.
(5). Don’t fall behind! The sequence of topics is closely interre-
lated, with one topic leading to another.
(6). Work in study groups. Have weekly meeting with your group!
Articulating your ideas to others will ensure you know mate-
rials better and gain confidence.
(7). When midterm exam or assignments are returned, rework
the problems on which you lost points to find out exactly
what you did wrong. Then carefully read and understand all
solutions.

Feel free to contact the author at kzhao@wlu.ca for any questions


involving the book (e.g., comments, suggestions, corrections, etc.).
The author welcomes your input.

Acknowledgments: The author would like to thank Dr. Dongfang


Gao for pointing out some errors in the early draft of the book.
March 18, 2022 9:45 amsart-9x6 12819-main page viii

viii RING AND FIELD THEORY

At last, the author wishes all instructors and students who use
this book a happy mathematical journey they will undertake into
this delightful and beautiful realm of algebra.

Kaiming Zhao
Department of Mathematics
Wilfrid Laurier University
75 University Ave. W., Waterloo
Ontario, Canada, N2L 3C5
June 1, 2021
March 18, 2022 9:45 amsart-9x6 12819-main page ix

Contents

Preface v
Notations xi
1. Basic Theory on Rings 1
1.1. Basic properties of rings 1
1.2. Isomorphism theorems 9
1.3. The field of quotients of an integral domain 13
1.4. Rings of polynomials 17
1.5. Ideal theory 21
1.6. Division algorithm for polynomials over a field 26
1.7. Irreducible polynomials over a field 31
1.8. Other irreducibility criteria 37
1.9. Exercises 41
2. Unique Factorization Domains 47
2.1. Basic definitions 47
2.2. Principal ideal domains 50
2.3. Euclidean domains 52
2.4. Polynomial rings over UFDs 55
2.5. Multiplicative norms 60
2.6. Exercises 63
3. Modules and Noetherian rings 67
3.1. Modules, submodules and isomorphism theorems 67
3.2. Free modules 71
3.3. Finitely generated modules over Euclidean domains 74
3.4. Noetherian rings 82
3.5. Exercises 86
4. Fields and Extension Fields 89
4.1. Prime fields and extension fields 89
4.2. Algebraic and transcendental elements 92
4.3. Algebraic extensions and algebraic closure 98
4.4. Finite fields 103
4.5. Exercises 108

ix
March 18, 2022 9:45 amsart-9x6 12819-main page x

x RING AND FIELD THEORY

5. Automorphisms of Fields 111


5.1. Automorphisms 111
5.2. The isomorphism extension theorem 116
5.3. Splitting fields 119
5.4. Separable extensions 122
5.5. Exercises 127
6. Galois Theory 131
6.1. Galois Theorem 131
6.2. Examples and an application 137
6.3. Cyclotomic extensions 142
6.4. Solvability by radicals 146
6.5. Insolvability of equations of higher degree 155
6.6. Dedekind’s Theorem and discriminants of
polynomials 157
6.7. Exercises 159
7. Sample Solutions 165
Appendix A. Equivalence Relations and
Kuratowski-Zorn Lemma 179
References 181
Index 183
March 18, 2022 9:45 amsart-9x6 12819-main page xi

Notations

The readers are supposed to be familiar with the following conven-


tional notations throughout this book:
• |A|: the cardinality of the set A,
• B A = {all maps A → B} where A and B are sets,
• ∅: the empty set,
• a ∈ A: a is an element of the set A,
• A ⊆ B: the set A is a subset of the set B,
• A ⊂ B: the set A is a proper subset of the set B,
• A ∪ B: the union of the sets A and B,
• A ∩ B: the intersection of the sets A and B,
• A \ B = {a ∈ A : a ∈ / B} where A and B are sets,
• Z: the set of integers,
• N: the set of positive integers,
• Q: the set of rational numbers,
• R: the set of real numbers,
• C: the set of complex numbers,
• Z+ = {0, 1, 2, . . . },
• Q+ = {r ∈ Q|r > 0},
• R+ = {r ∈ R|r > 0},
• Z∗ = Z\ {0},
• Q∗ = Q\ {0},
• R∗ = R\ {0},
• C∗ = C\ {0},
• Mm×n (F ): the set of all m × n matrices with entries in a field
F,
• Mn (F ) = Mn×n (F ),
• In : the identity matrix of size n.

xi
B1948 Governing Asia

This page intentionally left blank

B1948_1-Aoki.indd 6 9/22/2014 4:24:57 PM


March 18, 2022 9:45 amsart-9x6 12819-main page 1

1. Basic Theory on Rings

The conceptualization of rings spanned the 1870s to the 1920s, with


key contributions by Richard Dedekind (1831–1916), David Hilbert
(1862–1943), Abraham Fraenkel (1891–1965), and Emmy Noether
(1882–1935). Rings were first formalized as a generalization of
Dedekind domains that occur in number theory, and of polynomial
rings and rings of invariants that occur in algebraic geometry and
invariant theory. They later proved useful in other branches of math-
ematics such as geometry and analysis.
In this chapter we will first recall some definitions and results
about rings and fields from an elementary course on abstract algebra,
including basic properties of rings and isomorphism theorems. Proofs
of results from an elementary course on abstract algebra will often
be omitted in class although we contain details in this book. Then
we introduce some basic tools for this book, for example, the field
of quotients of an integral domain, basic properties for irreducible
polynomials over a field.

1.1. Basic properties of rings.


In this section we will recall some definitions and results from Ab-
stract Algebra. Most contents of the first two sections and Section 1.5
can be found in [LZ].
Definition 1.1.1. A group G is a set with a binary operation (gen-
erally called product) G × G → G, (a, b) 7→ a · b for a, b ∈ G satisfying
the following three axioms.
(G1) The product is associative:
a · (b · c) = (a · b) · c, ∀a, b, c ∈ G.
(G2) There is an identity element e in G:
a · e = e · a = a, ∀a ∈ G.
(G3) Each element a in G has an inverse a−1 ∈ G :
a · a−1 = a−1 · a = e.
1
March 18, 2022 9:45 amsart-9x6 12819-main page 2

2 RING AND FIELD THEORY

We generally denote the above group G as (G, ·, e), and a · b is


denoted by ab for convenience.
Definition 1.1.2. A group G is called abelian if ab = ba for all
a, b ∈ G.
Example 1.1.1. We have the following familiar examples of groups:
(Z, +, 0), (Q, +, 0), (R, +, 0), (C, +, 0), (Q∗ , ·, 1), (R∗ , ·, 1), (C∗ , ·, 1).
Example 1.1.2. Let X be a nonempty set. Denote SX by the set of
all bijections from X to itself. So any f ∈ SX has an inverse map in
SX denoted by f −1 . Clearly the identity map idX belongs to SX . For
any f, g ∈ SX , define the composition g · f by (g · f )(x) = g(f (x)),
x ∈ X. This composition gives a binary operation
SX × SX → SX , (g, f ) 7→ g · f,
satisfying
(f · g) · h = f · (g · h),
idX · f = f · idX = f,
f −1 · f = f · f −1 = idX ,
for all f, g, h ∈ SX . Thus (SX , ·, idX ) is a group, called the symmet-
ric group on X. We simply write SX as Sn if X = {1, 2, · · · , n}.
Definition 1.1.3. A subset H ⊆ G is a subgroup of a group G if
H is a group with the same operation as G, i.e.:
(1). H is closed under “·”,
(2). e ∈ H,
(3). a−1 ∈ H, for all a ∈ H.
We denote this by writing H ≤ G (or G ≥ H). The trivial sub-
groups of G are {e} , G ≤ (G, ·, e). The meaning for H < G (or
G > H) is clear.
Definition 1.1.4. Let (G, ·, e) be a group, H ≤ G and a ∈ G. The
(left) coset of H containing a is the subset aH = {ah : h ∈ H}
of G. We denote
G
= {aH : a ∈ G},
H
which is also denoted as G/H.
Definition 1.1.5. A subgroup N of a group G is called a normal
subgroup of G if g −1 xg ∈ N for all x ∈ N and all g ∈ G. We write
March 18, 2022 9:45 amsart-9x6 12819-main page 3

BASIC THEORY ON RINGS 3

N  G to indicate that N is a normal subgroup of G. By N  G we


have the obvious meaning.
Theorem 1.1.6. Let G be a group and N be a normal subgroup of
G. Define a binary operation on G/N by
xN · yN = xyN
for x, y ∈ G. With this multiplication, G/N is a group.
Definition 1.1.7. If G is a group and N is a normal subgroup of
G, we call group G/N (with the above multiplication) the quotient
group of G by N .
Definition 1.1.8. A ring R is a set with two binary operations, +
and ·, satisfying:
(R1) (R, +, 0) is an abelian group,
(R2) associativity under multiplication: (ab)c = a(bc) for all a, b,
c ∈ R,
(R3) distributivities: a(b + c) = ab + ac and (a + b)c = ac + bc for
all a, b, c ∈ R.
The ring R is usually denoted as (R, +, ·). From now on in this
book we always assume that R is a ring. A ring R is called commu-
tative if ab = ba for all a, b ∈ R.
Proposition 1.1.9. For any r, s ∈ R, we have
(1). r0 = 0r = 0,
(2). (−r)s = r(−s) = −(rs).
Proof. (1). r0 = r(0 + 0) = r0 + r0. Adding −(r0) to both sides,
we get:
0 = r0 − (r0) = r0 + r0 − r0 = r0.
Similarly, 0r = 0.
(2). 0 = 0s by (1) and
0 = 0s = (−r + r)s = (−r)s + rs.
Add −(rs) to both sides to get −(rs) = (−r)s. Similarly, r(−s) =
−(rs). 
Example 1.1.3. (a). We can easily see that the following number
systems are rings:
(Z, +, ·), (Q, +, ·), (C, +, ·), (R, +, ·), (Zn , +, ·).
March 18, 2022 9:45 amsart-9x6 12819-main page 4

4 RING AND FIELD THEORY

(b). The set (Mn (R), +, ·) with matrix multiplication and addition
is a ring. Note that if we replace R with any number system
this still holds true. For example, (Mn (Z), +, ·) is a ring.
Example 1.1.4. (a). Fix m ∈ N. For any n ∈ Z, write n =
{n + mk : k ∈ Z}. Define
n1 + n2 = n1 + n2 , and n1 · n2 = n1 n2 .
The classes 0, 1, · · · , m − 1 are called residues modulo m.
The set {0, 1, · · · , m − 1} is denoted by Zm or by Z/mZ. Then
(Zm , +, ·) is a commutative ring.
(b). The set of polynomials in x with coefficients in Q (or in R or
C)
Q[x] = {f (x) = a0 + a1 x + · · · + an xn : n ∈ N, ai ∈ Q}
with usual addition and multiplication is a commutative ring.
If an 6= 0 then n is the degree of f (x), denoted by deg(f (x)) =
n and we define deg(0) = −∞.
Definition 1.1.10. A subring of a ring R is a subset S of R which
is a ring under the same addition and multiplication as in R, denoted
by S ≤ R.
The meaning for S < R (or R > S) is clear.
Proposition 1.1.11. A non-empty subset S of a ring R is a subring
of R if and only if a + b, ab, −a ∈ S for any a, b ∈ S.
Proof. (⇒). Clearly, a subring has these properties.
(⇐). If S is a non-empty subset of R such that a + b, ab, −a ∈ S
for any a, b ∈ S, then (S, +) is a subgroup of (R, +) (from group
theory), and S is closed under multiplication. Associativity (R2)
and distributivities (R3) hold for S because they hold for R. 
Definition 1.1.12. Let d be an integer which is not a square. Define
√ √
Z[ d] = {a + b d : a, b ∈ Z}.
√ √
Call Z[ −1] = {a+b −1, a, b ∈ Z} the ring of Gaussian integers.
Proposition
√ 1.1.13. Let d be an integer√which is not a√square. √Then
(Z[ d], ·, +) is a ring. Moreover, m + n d = m0 + n0 d in Z[ d] if
and only if m = m0 and n = n0 .
March 18, 2022 9:45 amsart-9x6 12819-main page 5

BASIC THEORY ON RINGS 5



Proof. We will show that (Z[ d], ·, +) < (C, ·, +). Consider
m, n, a, b ∈ Z. Then we have: √ √
√Closure under addition: (m + n d) + (a + b d) = (m + a) + (n +
b) d. √ √
Closure √under multiplication: (m + n d)(a + b d) = ma + nbd +
(mb + na) d. √ √
Also, −(m√+ n d) = (−m) + (−n) d.
Hence (Z[ d], ·, +)√< (C, ·, +). √
√ Finally, if m + n d = m0 + n0 d,√ then if n 6= n0 we write
m−m0
d = n0 −n which is not possible since d is not a rational number.
Therefore, n = n0 hence m = m0 . 

Definition 1.1.14. An nonzero element a of a ring R is called a left


zero divisor if there exists nonzero b ∈ R such that ab = 0. We can
similarly define a right zero divisor of a ring. The multiplicative
identity element of a ring R, if it exists, is denoted by 1 and is called
the unity. A unital ring R is ring with unity 1 such that 1 6= 0.

If 0 = 1, then x · 1 = x and so x = x · 1 = x · 0 = 0. Hence if 0 = 1


then R = {0}. Note that, in some other books, the rings we defined
are called rng (or pseudo-ring), and unital rings are called rings.

Definition 1.1.15. If there exists a positive integer n such that na =


0 for all a ∈ R, then the least such positive integer is called the
characteristic of R, denoted by char(R) = n. If no such positive
integer exists, then the characteristic of R is 0, denoted by char(R) =
0.

Example 1.1.5. We know that (C, +, ·) is a ring with characteristic


char(C) = 0, and (Zn , +, 0) has characteristic char(Zn ) = n.

Example 1.1.6. Let R be the subring of M2 (Z3 ) consisting of ma-


trices of the form
ï ò
a b
, a, b ∈ Z3 .
0 0
ï ò
1 0
We can see that is a right zero divisor but not a left zero
0 0
divisor.
March 18, 2022 9:45 amsart-9x6 12819-main page 6

6 RING AND FIELD THEORY

Definition 1.1.16. A ring R is called an integral domain if


(1) R is commutative,
(2) R is unital, and
(3) R has no zero divisors.

For example (Z, +, ·), (Q, +, ·), (R, +, ·), (C, +, ·), (Z[ d]+, ·),
(Q[x]+, ·) are integral domains.
If R is an integral domain (or any commutative ring), then R[x]
denotes the set of polynomials in x with coefficients from R with
usual addition and multiplication. Clearly R[x] is a commutative
ring.
Proposition 1.1.17. If R is an integral domain, then R[x] is also.
Proof. We need only to check that there are no zero divisors. For
contradiction, assume that
f (x) = a0 + a1 x + · · · + am xm , g(x) = b0 + · · · + bn xn
are elements of R[x] such that f (x)g(x) is the zero polynomial.
Without loss of generality assume that am 6= 0, bn 6= 0, i.e., m =
degf (x), n = degg(x). Then f (x)g(x) = a0 b0 + · · · + am bn xm+n .
Since R is an integral domain, am bn 6= 0. Therefore we get a
contradiction, hence f (x)g(x) 6= 0. 
From the above proof we see that for any f (x), g(x) ∈ R[x],
deg(f g) = deg(f ) + deg(g),
if R is an integral domain.
Definition 1.1.18. Let R be a unital ring. An element a ∈ R is
a unit if there exists a−1 ∈ R such that aa−1 = a−1 a = 1. The
element a−1 is unique if it exists, and is called the inverse of a. The
unit group of R is the set consisting of all units of R is denoted as
U(R).
Definition 1.1.19. Let R be a ring and let a, b ∈ R. We say that
then a is a factor of b or a divides b if there exists c ∈ R such that
b = ac, denoted by a|b. We read a 6 | b as “a does not divide b”. Notice
that a|0 for any element a ∈ R.
Definition 1.1.20. Let R be a unital ring. A non-unit element a ∈ R
is irreducible if a = bc for some b, c ∈ R implies that b or c is a
unit.
March 18, 2022 9:45 amsart-9x6 12819-main page 7

BASIC THEORY ON RINGS 7

Note that irreducible polynomials in Q[x] are exactly irreducibles


in the polynomial ring Q[x].
Theorem 1.1.21. In (Zn , +, ·), the 0-divisors are the nonzero ele-
ments that are not coprime to n.
Proof. Let m ∈ Zn \ {0} , 1 ≤ m ≤ n − 1.
Case 1: Suppose that gcd(m, n) = d 6= 1 (i.e., m and n are not
m n m n
relatively prime). Then , ∈ Z. Then it follows that 0 < , <
d d d d
m n n m
n, so , 6= 0 in Zn . Then m · = · n = 0 in Zn . Thus m is a
d d d d
0-divisor.
Case 2: Now suppose gcd(m, n) = 1 (i.e., m and n are relatively
prime). Let s ∈ Zn so that ms = 0 in Zn , i.e., n ms. So n s. So
s = 0 in Zn . So m is not a 0-divisor. 
Proposition 1.1.22. Let n be a positive integer. Then (Zn , +, ·) is
an integral domain if and only if n is prime.
Proof. This follows from the previous theorem. 
Proposition 1.1.23. Every integral domain R satisfies the cancel-
lation property: if ax = ay and a 6= 0 then x = y for all x, y, a ∈ R.
Proof. If ax = ay then a(x − y) = 0. Since R has no zero divisors
and a 6= 0, we conclude that x − y = 0, so that x = y. 
Definition 1.1.24. A unital ring D is called a division ring if
U(D) = D \ {0}. A unital ring F is a field if it is a commutative
division ring.
A first step towards the notion of a field was made in 1770 by
Joseph-Louis Lagrange (1736–1813). The first clear definition of an
abstract field (1893) is due to Heinrich Martin Weber (1842–1913).
Example 1.1.7. (1). We have the following familiar examples of
fields: Q, R, C, Z2 , Z3 , where Z2 = {0, 1},√Z3 = {0, 1, 2}
√= {0, 1, −1}.
(2). For a non-square d ∈ Q, then Q[ d] = {x + y d : x, y ∈ Q}
is a field. √
It is easy to see that Q[ d] is a subring of C. Assuming x 6= 0 or
y 6= 0, we compute
√ √
1 x−y d x−y d √
√ = √ √ = 2 2
∈ Q[ d].
x+y d (x − y d)(x + y d) x −y d
Note that x2 − y 2 d 6= 0 since d is not a square of a rational number.
March 18, 2022 9:45 amsart-9x6 12819-main page 8

8 RING AND FIELD THEORY

Definition 1.1.25. A subset S of a field F is a subfield if S is a


field with the same addition and multiplication, denoted by S ≤ F .
The obvious meaning of S < F is clear.
To check that a subset S of a field F is a subfield, it is enough to
check that 0, 1, a − b, ab−1 ∈ S (if b 6= 0) for any a, b ∈ S.
Definition 1.1.26. Let F be a subfield of a field K and α1 , · · · , αn ∈
K. The smallest subfield of K containing F and α1 , · · · , αn is de-
noted by F (α1 , α2 , · · · , αn ).
√ √
Example 1.1.8. This notation agrees with Q( d) = {a + b d :
a, b√∈ Q} where d is a non-square rational number. Let’s check √ that
Q( d) is indeed the smallest subfield of C containing
√ Q and d. The
smallest subfield must contain all numbers like a d, a √ ∈ Q, since it
is closed under ·, and hence also all numbers like a + a 0 d, a, a0 ∈ Q,
√ √
since closed under√ +. So {a + b d : a, b ∈ Q} ⊂ Q( d). We also
know that {a + b d : a, b ∈ Q} is a field.
√ √
Similarly we can consider Q( d1 , d2 ), and more complicated
fields.
Theorem 1.1.27. Every field is an ID.
Proof. Let F be a field. Then 0 6= 1 ∈ F and F is commutative.
Suppose on the contrary that a ∈ F is a 0-divisor. Then a 6= 0 and
there exists b 6= 0 such that ab = 0. Since F is a field, then a−1 exists
and so
0 6= b = a−1 (ab) = a−1 0 = 0
and is clearly a contradiction. So F cannot have any 0-divisors. So
F must be an ID. 
The following theorem was proved by Joseph Wedderburn (1882–
1948) in 1905.
Theorem 1.1.28 (Wedderburn’s Little Theorem). Every finite ID
is a field.
Proof. Let D = {0, 1, a1 , · · · , an } be a finite ID. We need to show
that the unit group U(D) = D\ {0}. Let a ∈ D\ {0}. Since ax = ay
implied x = y we have
{a1, aa1 , · · · , aan } = {1, a1 , · · · , an } .
| {z }
n+1 elements
March 18, 2022 9:45 amsart-9x6 12819-main page 9

BASIC THEORY ON RINGS 9

Since these two sets are equal, there exists b ∈ D \ {0} such that
ab = 1. So b = a−1 and so D must be a field. 
Corollary 1.1.29. The ring Zm is a field if and only if m is prime.
Proof. If m is not prime then we know that Zm has zero divisors,
hence is not a field.
If m is a prime, then Zm is a finite integral domain, hence a field
by the previous theorem. 
1.2. Isomorphism theorems.
Definition 1.2.1. A subring I of a ring R is called an ideal if
ar, ra ∈ I for all a ∈ I, r ∈ R. If I is an ideal of R we denote this
fact by I  R. By I  R (or I  R) we mean I  R and I 6= R.
Proposition 1.2.2. A non-empty subset I of a ring R is an ideal of
R if and only if a − b, ar, ra ∈ I whenever a, b ∈ I and r ∈ R.
Proof. This is easy to see. 
Definition 1.2.3. Let I be an ideal of a ring R and x ∈ R. The
coset of I in R containing x is
x + I = {x + i : i ∈ I}.
When dealing with cosets, it is more important to realize that, in
general, a given coset can be represented in more than one way. The
next lemma shows how the coset representatives are related.
Lemma 1.2.4. Let R be a ring with an ideal I and x, y ∈ R. Then
x + I = y + I if and only if x − y ∈ I.
Proof. We omit the detailed proof since it is easy. 
Theorem 1.2.5. Let I  R. Then R/I = {a + I : a ∈ R} is a ring
with
(a + I) + (b + I) = (a + b) + I, (a + I)(b + I) = ab + I, ∀a, b ∈ R.
The ring (R/I, +, ·) is call the quotient ring of R by I.
Proof. The proof is fairly standard and can be found in any
Abstract Algebra book like [LZ]. 
The zero element of R/I is I = 0 + I = a + I for any a ∈ I. If S is
a subset of R with S ⊇ I we denote by S/I the subset {s + I : s ∈ S}
of R/I.
March 18, 2022 9:45 amsart-9x6 12819-main page 10

10 RING AND FIELD THEORY

Definition 1.2.6. A map θ of a ring R into a ring S is said to be


a (ring) homomorphism if θ(x + y) = θ(x) + θ(y) and θ(xy) =
θ(x)θ(y) for all x, y ∈ R.
The map θ : R → S defined by θ(r) = 0 for all r ∈ R is a
homomorphism. It is called the zero homomorphism.
The map φ : R → R defined by φ(r) = r for all r ∈ R is also a
homomorphism. It is called the identity homomorphism.
Let I  R. Then σ : R → R/I defined by σ(x) = x + I for all
x ∈ R is a homomorphism of R onto R/I. This is called the natural
(or canonical) homomorphism.
Proposition 1.2.7. Let R, S be rings and θ : R → S a homomor-
phism. Then
(1). θ(0R ) = 0S ,
(2). θ(−r) = −θ(r) for all r ∈ R,
(3). K = {x ∈ R : θ(x) = 0S } is an ideal of R,
(4). θR = {θ(r) : r ∈ R} is subring of S,
(5). θ−1 (θ(x)) = x + K for any x ∈ R.
Proof. The proof is standard. 
In the above theorem, K is called the kernel of θ and θR is called
the (homomorphic) image of R. The ideal K is sometimes denoted
by ker(θ).
Definition 1.2.8. Let θ be a homomorphism of a ring R into a
ring S. Then θ is called an isomorphism if θ is a one to one and
onto map. We say that R and S are isomorphic rings if there is an
isomorphism θ : R → S, denote this by R ∼
= S.
Lemma 1.2.9. Let N  R and let
[[R, N ]] = {K ≤ R | K ⊇ N } , [[R/N ]] = K 0 ≤ R/N .


(1). Ψ : [[R, N ]] → [[R/N ]] defined by K 7→ K/N is a bijection.


(2). For K ∈ [[R, N ]] we have that K  R if and only if K/N 
R/N .
Proof. (1). Consider the homomorphism γ : R → R/N, x 7→
x + N , which is onto. For any K ∈ [[R, N ]], we see that Ψ(K) =
γ(K) = K/N .
For K 0 ∈ [[R/N ]], let K = γ −1 (K 0 ). Then N ≤ K ≤ R. We see
that Ψ(K) = γ(γ −1 (K 0 )) = K 0 since γ is onto. So Ψ is onto.
March 18, 2022 9:45 amsart-9x6 12819-main page 11

BASIC THEORY ON RINGS 11

Let K1 , K2 ∈ [[R, N ]] such that Ψ(K1 ) = Ψ(K2 ). Then γ(K1 ) =


γ(K2 ), and
K1 = K1 + N = γ −1 (γ(K1 )) = γ −1 (γ(K2 )) = K2 + N = K2 .
So Ψ is 1-1. Thus Ψ : [[R, N ]] → [[R/N ]] is a bijection.
(2). We see that K  R if and only if rK, Kr ⊆ K for any r ∈ R,
if and only if r(K/N ) = (r + N )(K/N ), (K/N )r = (K/N )(r + N ) ⊆
K/N for any r ∈ R, if and only if K/N  R/N . 
Theorem 1.2.10 (First Isomorphism Theorem). Let θ : R → S be
a ring homomorphism. Then θR ∼
= R/I where I = ker θ.
Proof. Defined σ : R/I → θR by σ(x + I) = θ(x) for all x ∈ R.
The map σ is well defined since for x, y ∈ R,
x+I = y+I ⇐⇒ x−y ∈ I = ker θ ⇐⇒ θ(x−y) = 0 ⇐⇒ θ(x) = θ(y).
Clearly σ is 1-1 and onto. Then σ is the required isomorphism. 
Theorem 1.2.11 (Second Isomorphism Theorem). Let I be an ideal
and L a subring of a ring R. Then L ∩ I  L, I  L + I, and
L/(L ∩ I) ∼
= (L + I)/I.
Proof. It is easy to see that L ∩ I  L, I  L + I.
Let σ be the natural homomorphism R → R/I. Restrict σ to the
ring L. We have σL = (L + I)/I. The kernel of σ restricted to L is
L ∩ I. Now apply previous theorem. 
Theorem 1.2.12 (Third Isomorphism Theorem). Let I, K  R such
that I ⊆ K. Then
(R/I)/(K/I) ∼= R/K.
Proof. Since K/I  R/I, so (R/I)/(K/I) is defined. Define a
map
γ : R/I → R/K, γ(x + I) = x + K, ∀x ∈ R.
The map γ is easily seen to be well defined and a homomorphism
onto R/K. Further,
γ(x + I) = K ⇐⇒ x+K =K
⇐⇒ x∈K
⇐⇒ x + I ∈ K/I.
Therefore ker γ = K/I. Now apply the first isomorphism theorem.

March 18, 2022 9:45 amsart-9x6 12819-main page 12

12 RING AND FIELD THEORY

The isomorphism theorems were formulated in some generality for


homomorphisms of modules by Emmy Noether in her paper
“Abstrakter Aufbau der Idealtheorie in algebraischen Zahlund Funk-
tionenkrpern”, which was published in 1927 in “Mathematische An-
nalen”. (See Theorems 3.1.10–3.1.12.) Three years later, B.L. van
der Waerden (1903–1996) published his influential “Algebra”, the first
abstract algebra textbook that took the groups-rings-fields approach
to the subject. The three isomorphism theorems, called homomor-
phism theorem, and two laws of isomorphism when applied to rings,
appeared explicitly.
Definition 1.2.13. Let R1 , . . . , Rn be rings. We define the external
direct sum S to be the set of all n-tuples {(r1 , . . . , rn ) : ri ∈ Ri }.
On S we define addition and multiplication component wise. This
makes S a ring. We write S = R1 ⊕ · · · ⊕ Rn .
The set (0, . . . , 0, Rj , 0, . . . , 0) is an ideal of S. Clearly S is the
internal direct sum of these ideals. But (0, . . . , Rj , . . . 0) ∼ = Rj . Be-
cause of this S can be considered as a ring in which the Rj are ideals
and S is their internal direct sum. Also in internal direct sum we can
consider I1 ⊕ · · · ⊕ In to be the external direct sum of the rings Ij .
Hence, in practice, we do not need to distinguish between external
and internal direct sums.
Definition 1.2.14. Let {Iλ }λ∈Λ be a collection of ideals of a ring R.
We define their sum to be
X
Iλ = {x1 + · · · + xk : xi ∈ Iλi , λk ∈ Λ}.
λ∈Λ
That is, the sum is the collection of finite sums of elements of the
Iλ ’s.
P We say that the sum of the Iλ ’s is direct if each element of
λ∈Λ Iλ is uniquely expressible as
x1 + · · · + xk , with xi ∈ Iλi .
In this case we denote this sum as ⊕λ∈Λ Iλ or I1 ⊕ · · · ⊕ In if Λ is
finite.
P
Proposition 1.2.15. The sum λ∈Λ Iλ is direct if and only if
Ñ é
X
Iµ ∩ Iλ = 0, for all µ ∈ Λ.
λ∈Λ,λ6=µ
March 18, 2022 9:45 amsart-9x6 12819-main page 13

BASIC THEORY ON RINGS 13

Proof. The proof is standard and we omit the details. 


Example 1.2.1. If φ is a nonzero ring homomorphism from the real
numbers R to R, show that φ is the identity map. (Hint: show x > 0
implies φ(x) > 0.)
Solution. Since φ 6= 0 we see that φ is one-to-one, φ(±1) = ±1
and φ(0) = 0.
Let x > 0 be a positive real number. Then there exists y ∈ R such
that x = y 2 . Hence φ(x) = φ(y)2 > 0.
If a < b, then b − a > 0. Hence φ(b) − φ(a) = φ(b − a) > 0 by the
above. Hence φ(a) < φ(b). This means that for a, b ∈ R,
a < b =⇒ φ(a) < φ(b).
If n ∈ N, using n = 1 + · · · + 1, we have
φ(n) = φ(1 + · · · + 1) = φ(1) + · · · + φ(1) = 1 + · · · + 1 = n, ∀n ∈ N,
and φ(−n) = −n for all n ∈ N. So
φ(n) = n, ∀n ∈ Z.
Now, any rational number is of the form r = ac−1 for a, c ∈ Z with
c 6= 0. It follows that
φ(r) = φ(ac−1 ) = φ(a)φ(c)−1 = ac−1 = r, ∀r ∈ Q.
Let a ∈ R and suppose φ(a) 6= a, say φ(a) > a. We know that
there is q ∈ Q such that a < q < φ(a). Hence, φ(a) < φ(q) = q <
φ(a). This is a contradiction. Thus φ(a) = a for all a ∈ R. 
1.3. The field of quotients of an integral domain.
If every nonzero element in an integral domain D has a multiplicative
inverse, then D is a field. It is the purpose of this section to show
that every integral domain can be regarded as subring of a field, a
field of quotients of the integral domain. This field will be a minimal
field containing the integral domain. For example, the integers are
contained in the field Q, whose elements can all be expressed as
quotients of integers.
We can follow the steps by the way Q can be formed from Z.

Let D be an integral domain that we desire to enlarge to a field of


quotients F . We take four steps to obtain F as follows.
1. Define the elements of F .
2. Define addition and multiplication on F .
March 18, 2022 9:45 amsart-9x6 12819-main page 14

14 RING AND FIELD THEORY

3. Show that F is a field under these operations.


4. Show that D can be considered a subring of F .

Step 1. Consider S = {(a, b)|a, b ∈ D, b 6= 0}.


Definition 1.3.1. Two elements (a, b), (c, d) ∈ S are equivalent,
denoted by (a, b) ∼ (c, d), if ad = bc.
Theorem 1.3.2. The relation ∼ on S is an equivalence relation.
Proof. Reflexive: (a, b) ∼ (a, b) since ab = ba (D is an integral
domain).
Symmetric: (a, b) ⇐⇒ ad = bc ⇐⇒ (c, d) ∼ (a, b).
Transitive: If (a, b) ∼ (c, d) and (c, d) ∼ (r, s), then ad = bc and
cs = dr. We have
asd = sad = sbc = bcs = bdr = brd.
Now d 6= 0, and D is an integral domain, so cancellation is valid.
Hence from asd = brd we obtain as = br, so that (a, b) ∼ (r, s). 

We now know that ∼ gives a partition of S into equivalence classes.


We shall let ab be the equivalence class of (a, b) in S under the relation
∼, i.e.,
a
= {(c, d) ∈ S : (c, d) ∼ (a, b)}.
b
Let na o
F = : (a, b) ∈ S .
b
Step 2. Define addition and multiplication in F . Observe that if
D = Z and ab is viewed as a/b ∈ Q, these definitions applied to Q
give the usual operations.
Theorem 1.3.3. For ab , dc ∈ F ,the operations
a c ac a c ad + bc
· = , + =
b d bd b d bd
are well-defined on F .
Proof. Since ab , dc ∈ F , then (a, b) and (c, d) are in S, and bd 6= 0.
So
(ad + bc, bd), (ac, bd) ∈ S.
ac ad+bc
Thus bd , bd ∈ F.
March 18, 2022 9:45 amsart-9x6 12819-main page 15

BASIC THEORY ON RINGS 15

To see these operations of addition and multiplication are well


defined, suppose that ab11 = ab and dc11 = dc . Then
a1 b = b1 a, c1 d = d1 c. (1.1)
We must show that
a1 d1 + b1 c1 ad + bc a1 c1 ac
= , and = ,
b1 d1 bd b1 d1 bd
i.e.,
(a1 d1 + b1 c1 )bd = b1 d1 (ad + bc), and a1 c1 bd = b1 d1 ac.
These can be verified by using (1.1). Now we complete the proof. 
From this theorem we see that,
ab b
= , ∀a, b, c ∈ D with ac 6= 0.
ac c
Step 3. Check that F is a field under these operations.
Theorem 1.3.4. The above defined (F, +, ·) from D is a field.
Proof.
(1). Addition in F is commutative: Since ab + dc = bc+da bd and
c a ad+bc a c c a
d + b = bd . So b + d = d + b .
(2). Addition is associative. This is easy to verify.
(3). The element 01 is an identity element for addition in F . This
is clear.
(4). The element −a a
b is an additive inverse for b in F . This is
clear.
(5). Multiplication in F is associative. This is easy to verify.
(6). Multiplication in F is commutative. This is easy to verify.
(7). The distributive laws hold in F . This is easy to verify.
(8). The element 11 is a multiplicative identity element in F . This
is clear.
(9). If ab ∈ F is not the additive identity element, then a 6= 0 in
D and ab is a multiplicative inverse for ab :
Let ab ∈ F. If a = 0, then 0b = b0 0 0
b1 = 1 . But 1 is the additive identity
by (3). Thus if ab 6= 01 in F , we have a 6= 0. Now ab ab = ab 1
ba = 1 . Thus
ab 1
= ,
ba 1
1
and 1 is the multiplicative identity by (8). So (F, ·, +) is a field. 
This completes Step 3.
March 18, 2022 9:45 amsart-9x6 12819-main page 16

16 RING AND FIELD THEORY

Step 4. Show that F can be regarded as containing D.


a
Theorem 1.3.5. The map ι : D → F given by ι(a) = 1 is an
isomorphism of D with a subring of F .
Proof. For a, b ∈ D, we have
a b a1 + 1b a+b
ι(a) + ι(b) = + = = = ι(a + b),
1 1 1 1
ab ab
ι(a)ι(b) = = = ι(ab).
11 1
It remains for us to show only that ι is one to one. If ι(a) = ι(b),
then a1 = 1b , so (a, 1) ∼ (b, 1) giving a1 = 1b; that is, a = b. Thus ι is
an isomorphism of D with ι(D), of course, as a subdomain of F . 
Since ab = a1 1b = a1 ( 1b )−1 = ι(a)ι(b)−1 clearly holds in F , we have
now proved the following theorem.
Theorem 1.3.6. Any integral domain D can be enlarged to (or em-
bedded in) a field F such that every element of F can be expressed as
a quotient of two elements of D. (Such a field F is called a field of
quotients of D, or field of fractions of D.)
The next theorem will show that the field of quotients of D is
unique.
Theorem 1.3.7. Let F be a field of quotients of D and let E be any
field containing D. Then there exists a map ψ : F → E that gives an
isomorphism of F with a subfield of E such that ψ(a) = a for a ∈ D.
Proof. For a, b ∈ D, by a/b we mean the quotient regarded as
elements of F , by a/E b we mean the quotient regarded as elements
of E. Define
ψ : F → E, ψ(a/b) = a/E b, ∀a, b ∈ D with b 6= 0.
D ≤ F
|| ↓ψ
D ≤ E
We first show that ψ is well-defined. If a/b = c/d in F , then
ad = bc in D. Thus
a/E b = c/E d,
in E, so ψ is well-defined. The equations
ψ(xy) = ψ(x)ψ(y), and ψ(x + y) = ψ(x) + ψ(y), ∀x, y ∈ F
March 18, 2022 9:45 amsart-9x6 12819-main page 17

BASIC THEORY ON RINGS 17

follow easily from the definition of ψ on F and from the fact that ψ
is the identity on D.
If a/E b = c/E d we have ad = bc. So a/b = c/d. Thus ψ is one to
one.
By definition, ψ(a) = a for a ∈ D. The theorem follows. 
Theorem 1.3.8. Every field E containing an integral domain D
contains a field of quotients of D.
Proof. Let F be a field of quotients of D. In the above Theorem
the subfield ψ[F ] of E is a quotient field of D. 
Theorem 1.3.9. Any two fields of quotients of an integral domain
D are isomorphic.
Proof. This directly follows from Theorem 1.3.7. 

We remark that, in general, not every unital noncommutative ring


without zero divisors can be embedded into a division ring. This
leads to Ore theory.
The right Ore condition for a multiplicative subset S = R \ {0}
of a ring R is that for any a ∈ R and any s ∈ S, the intersection
aS ∩ sR 6= ∅. A (non-commutative) integral domain for which the
set of non-zero elements satisfies the right Ore condition is called a
right Ore domain. Only right Ore domains can be embedded in
some division rings.

1.4. Rings of polynomials.


Definition 1.4.1. Let R be a ring. A polynomial f (x) with coef-
ficients in R is a sum
f (x) = a0 + a1 x + · · · + an xn
where n ∈ Z+ , ai ∈ R. The ai ’s are coefficients of f (x). If an 6= 0
then an is called the leading coefficient of f (x). A polynomial
f (x) ∈ R[x] is called a monic polynomial if its leading coefficient
is 1 (assuming that R is unital).
If R is unital, we will write a term 1xk in such a sum as xk . For
example, in Z[x], we will write the polynomial 2+1x as 2+x. Finally,
we shall agree that we may omit altogether from the formal sum any
term 0xi , or a0 if a0 = 0 but not all ai = 0. Thus 0, 2, x, and 2 + x2
March 18, 2022 9:45 amsart-9x6 12819-main page 18

18 RING AND FIELD THEORY

are polynomials with coefficients in Z. An element of R is a constant


polynomial.
Addition and multiplication of polynomials with coefficients in a
ring R are defined in a way familiar to us. If
f (x) = a0 + a1 x + · · · + an xn ,
and
g(x) = b0 + b1 x + · · · + bn xn ,
then for polynomial addition, we have
f (x) + g(x) = (a0 + b0 ) + (a1 + b1 )x + · · · + (an + bn )xn ,
and for polynomial multiplication, we have
k
X
f (x)g(x) = d0 + d1 x + · · · + d2n x2n where dk = ai bk−i .
i=0
With these definitions of addition and multiplication, we have the
following theorem.
Theorem 1.4.2. The set R[x] of all polynomials in an indeterminate
x with coefficients in a ring R is a ring under polynomial addition
and multiplication. If R is commutative, then so is R[x], and if R
unital then so is R[x].
Proof. Clearly, we have the abelian group (R[x], +, 0). The asso-
ciative law for multiplication and the distributive laws are straight-
forward, but slightly cumbersome, computations. We only prove the
associative law.
Applying ring axioms to ai , bj , ck ∈ R, we obtain

m
!Ñ n é
r
! m+n+rÑ é
X X X X X
 ai xi bj xj  ck xk = ai bj ck xs
i=0 j=0 k=0 s=0 i+j+k=s

m
! Ñ n
é
r
!
X X X
= ai xi  bj xj ck xk  .
i=0 j=0 k=0
The distributive laws are similarly proved.
The comments prior to the statement of the theorem show that
R[x] is a commutative ring if R is commutative, and a unity 1 6= 0
in R is also unity for R[x], in view of the definition of multiplication
in R[x]. 
March 18, 2022 9:45 amsart-9x6 12819-main page 19

BASIC THEORY ON RINGS 19

Thus Z[x] is the ring of polynomials in the indeterminate x with


integral coefficients, Q[x] the ring of polynomials in x with rational
coefficients, and so on.

Example 1.4.1. In Z2 [x], we have

(x + 1)2 = (x + 1)(x + 1) = x2 + (1 + 1)x + 1 = x2 + 1.

Still working in Z2 [x], we obtain

(x + 1) + (x + 1) = (1 + 1)x + (1 + 1) = 0x + 0 = 0.

If R is a ring and x and y are two indeterminates, then we can


form the ring (R[x])[y], that is, the ring of polynomials in y with
coefficients that are polynomials in x. Every polynomial in y with
coefficients that are polynomials in x can be rewritten in a natural
way as a polynomial in x with coefficients that are polynomials in
y. This indicates that (R[x])[y] is naturally isomorphic to (R[y])[x],
although a careful proof is tedious. We shall identify these rings
by means of this natural isomorphism, and shall consider this ring
R[x, y] the ring of polynomials in two indeterminates x and y with
coefficients in R. The ring R[x1 , · · · , xn ] of polynomials in the n
indeterminates xi with coefficients in R is similarly defined.
(a). If D is an integral domain then so is D[x]. In particular, if F
is a field, then F [x] is an integral domain.
(b). We can construct the field of quotients F (x) of F [x]. Any
element in F (x) can be represented as a quotient f (x)/g(x) of two
polynomials in F [x] with g(x) 6= 0. We similarly define F (x1 , · · · , xn )
to be the field of quotients of F [x1 , · · · , xn ]. This field F (x1 , · · · , xn )
is the field of rational functions in n indeterminates over F . These
fields play a very important role in algebraic geometry.

Let E and F be fields, with F a subfield of E, that is, F ≤ E. The


next theorem asserts the existence of very important homomorphisms
of F [x] into E. These homomorphisms will be the fundamental tools
in this book.

Theorem 1.4.3. Let F be a subfield of a field E, let α ∈ E, and let


x be an indeterminate. The map φα : F [x] → E defined by

φα (a0 + a1 x + · · · + an xn ) = a0 + a1 α + · · · + an αn
March 18, 2022 9:45 amsart-9x6 12819-main page 20

20 RING AND FIELD THEORY

is a ring homomorphism of F [x] into E. Also, φα (x) = α, and


φα (a) = a for any a ∈ F . The homomorphism φα is called the
evaluation at α.
Proof. Clearly φα is well defined. If
f (x) = a0 + a1 x + · · · + an xn , g(x) = b0 + b1 x + · · · + bm xm ,
h(x) = f (x) + g(x) = c0 + c1 x + · · · + cr xr ,
then
φα (f (x) + g(x)) = φα (h(x)) = c0 + c1 α + · · · + cr αr ,
while
φα (f (x))+φα (g(x)) = (a0 +a1 α+· · ·+an αn )+(b0 +b1 α+· · ·+bm αm ).
Since by definition of polynomial addition we have ci = ai + bi , we
see that
φα (f (x) + g(x)) = φα (f (x)) + φα (g(x)).
For the multiplication, if f (x)g(x) = d0 + d1 x + · · · + ds xs , then
φα (f (x)g(x)) = d0 + d1 α + · · · + ds αs ,
while
(φα (f (x)))(φα (g(x))) = (a0 +a1 α+· · ·+an αn )(b0 +b1 α+· · ·+bm αm ).
Thus φα is a homomorphism.
The definition of φα applied to a constant polynomial a ∈ F [x],
where a ∈ F, gives φα (a) = a. Clearly,
φα (x) = φα (1x) = 1α = α.

Example 1.4.2. (a). Consider the evaluation homomorphism φ0 :
Q[x] → R. Here
φ0 (a0 + a1 x + · · · + an xn ) = a0 + a1 0 + · · · + an 0n = a0 .
Thus every polynomial is mapped onto its constant term.
(b). We have the evaluation homomorphism φ2 : Q[x] → R. Here
φ2 (a0 + a1 x + · · · + an xn ) = a0 + a1 2 + · · · + an 2n .
Note that φ2 (x2 +x−6) = 22 +2−6 = 0. Thus x2 +x−6 is in
the kernel N of φ2 . Of course, x2 +x−6 = (x−2)(x+3), and
the reason that φ2 (x2 +x−6) = 0 is that φ2 (x−2) = 2−2 = 0.
March 18, 2022 9:45 amsart-9x6 12819-main page 21

BASIC THEORY ON RINGS 21

(c). We have the evaluation homomorphism


φi : Q[x] → C, φi (a0 + a1 x + · · · + an xn ) = a0 + a1 i + · · · + an in
and φi (x) = i. Note that φi (x2 + 1) = i2 + 1 = 0, so x2 + 1 is
in the kernel N of φi .
(d). We have the evaluation homomorphism
φπ : Q[x] → R, φπ (a0 + a1 x + · · · + an xn ) = a0 + a1 π + · · · + an π n .
We know that a0 + a1 π + · · · + an π n = 0 if and only if ai = 0
for i = 0, 1, · · · , n. Thus the kernel of φπ is 0, and φπ is a
one-to-one map. This shows that all formal polynomials in π
with rational coefficients form a ring isomorphic to Q[x] in a
natural way with φπ (x) = π.
Using evaluation homomorphisms, by solving a polynomial equa-
tion, we shall refer to finding a zero of a polynomial.
Definition 1.4.4. Let F be a subfield of a field E, and let α ∈ E.
Let f (x) = a0 + a1 x + · · · + an xn ∈ F [x]. Let f (α) denote
φα (f (x)) = a0 + a1 α + · · · + an αn .
If f (α) = 0, then α is a zero of f (x).
In terms of this definition, we can rephrase the classical problem
of finding all real numbers r such that r2 + r − 6 = 0 by
{α ∈ R : φα (x2 + x − 6) = 0} = {r ∈ R : r2 + r − 6 = 0} = {2, −3}.

1.5. Ideal theory.


Let R be a ring. We know that the two trivial ideals of R are R itself
and 0. If I  R and I 6= R, then I is said to be a proper ideal of
R, denoted I  R. Let’s introduce the following result.
Theorem 1.5.1. Let R be a unital ring and let I  R such that I
contains a unit. Then I = R.
Proof. Let u ∈ I be a unit. There exists u−1 ∈ R such that
1 = uu−1 ∈ I.
Then R = R · 1 ⊆ I ⊆ R, yielding that R = I. 
Let’s define what a maximal ideal is.
March 18, 2022 9:45 amsart-9x6 12819-main page 22

22 RING AND FIELD THEORY

Definition 1.5.2. A maximal ideal of a ring R is a proper ideal


M of R so that there is no proper ideal of R properly containing M .
In other words, a proper ideal M of R is maximal if
M ⊆N  R =⇒ N = M or N = R.
Example 1.5.1. Let R = (Z, +, ·) and p be a prime. Show that
pZ  R, and pZ is a maximal ideal of Z.
Proof. It is easy to see that pZ  R.
Let pZ ⊆ N  Z. If pZ 6= N , there is r ∈ N \ pZ. We see that
p 6 | r, and gcd(p, r) = 1. there exist a, b ∈ Z so that 1 = ap + br ∈ N .
From Theorem 1.5.1 we see that N = Z. So pZ is a maximal ideal.

Lemma 1.5.3. Let R be a unital commutative ring. Then R is a
field if and only if R has exactly the two ideals, i.e., 0 and R.
Proof.
R is a field ⇐⇒ any nonzero element of R is a unit
⇐⇒ any nonzero ideal contains a unit
⇐⇒ any nonzero ideal is R
⇐⇒ R has only the trivial ideals.

Theorem 1.5.4. Let R be a unital commutative ring and M  R.
Then M is maximal if and only if R/M is a field.
Proof.
M is max ⇐⇒ {M, R} = {N  R| N ⊇ M }
⇐⇒ N 0  R/M = {M/M, R/M } = {0, R/M }


⇐⇒ R/M has exactly two ideals


⇐⇒ R/M is a field.

Definition 1.5.5. Let R be a commutative ring, N  R. We say
that N is prime if ab ∈ N implies that a ∈ N or b ∈ N for a, b ∈ R.
The set of all prime ideals of R is call the spectrum of R.
Example 1.5.2. A non-zero ideal nZ  Z is prime if and only if n
is a prime.
March 18, 2022 9:45 amsart-9x6 12819-main page 23

BASIC THEORY ON RINGS 23

Proof. Suppose n = p is a prime, and a · b ∈ pZ. So p | a · b. So


p | a or p | b, i.e., a ∈ pZ or b ∈ pZ.
For the other direction, suppose n = pq is a composite number
(p, q 6= 1). Then n ∈ nZ but p 6∈ nZ and q 6∈ nZ, since 0 < p, q < n.

Definition 1.5.6. Let R be a ring, S ⊆ R. The ideal hSi is the
smallest ideal of R containing S, which is called the ideal generated
by S.
For any A, B ⊆ R, we define
( n )
X
AB = ai bi : ai ∈ A, bi ∈ B .
i=1

If R is unital, then hSi = SR + RS.


Theorem 1.5.7. Let R be a unital commutative ring, N  R. Then
N is prime if and only if R/N is an ID.
Proof. (⇒). Suppose that N is prime. R/N is a unital commu-
tative ring where 1 6= 0. We need to demonstrate that there are no
zero divisors. Let a + N, b + N ∈ R/N with (a + N )(b + N ) = 0 + N
(since either a or b could be zero divisors). Then
ab + N = N ⇒ ab ∈ N.
Since N is prime, it follows that a ∈ N or b ∈ N . So a + N = 0 + N
or b + N = 0 + N . So R/N has no zero divisors. By definition, it
follows that R/N is an ID.
(⇐). Suppose that the quotient ring R/N is an ID. Let ab ∈ N
for a, b ∈ R. Then
(a + N )(b + N ) = ab + N = 0 + N in R/N.
Since R/N is an ID, then either a + N = 0 + N or b + N = 0 + N .
So a ∈ N or b ∈ N . So N is prime. 
Combining this theorem with the fact that any field is an integral
domain, we have the following corollary.
Corollary 1.5.8. Let R be a unital commutative ring and M  R.
Then any maximal ideal of R is prime.
Example 1.5.3. Is the ideal h2, x2 + 5i ⊂ Z[x] prime?
March 18, 2022 9:45 amsart-9x6 12819-main page 24

24 RING AND FIELD THEORY

Solution. Consider Z[x]/h2, x2 + 5i ∼ = Z2 [x]/hx2 − 1i. Then


x + 1, x − 1 6= 0 in Z2 [x]/hx − 1i, but (x + 1)(x − 1) = x2 − 1 = 0.
2

So Z[x]/h2, x2 + 5i is not an ID. Thus the ideal h2, x2 + 5i is not a


prime ideal of Z[x]. 
Lemma 1.5.9. Let R be a finite commutative unital ring. Then any
prime ideal of R is maximal.
Proof. Let N  R and N be prime. We show that N is maximal.
So R/N is an integral domain (by the previous theorem). Since R is
finite, then R/N is finite. Since any finite integral domain is a field,
then R/N is a field and so, N must be maximal. 
Definition 1.5.10. If R is a commutative unital ring and a ∈ R,
the ideal {ra|r ∈ R} of all multiples of a is the principal ideal
generated by a and is denoted by hai. An ideal N of R is a principal
ideal if N = hai for some a ∈ R. An integral domain is a principal
ideal domain (PID, for short) if all its ideals are principal.
Theorem 1.5.11. For any field F , the polynomial ring F [x] is a
PID.
Proof. Let N  F [x]. If N = {0}, then N = h0i. Suppose
that N 6= {0}, and let g(x) be a nonzero element of N of minimal
degree. If degg(x) = 0, then g(x) ∈ F and is a unit. We see that
N = F [x] = h1i, so N is principal. If degg(x) ≥ 1, for any f (x) ∈ N,
then
f (x) = g(x)q(x) + r(x), where degr(x) < degg(x).
Since f (x) ∈ N and g(x) ∈ N then f (x) − g(x)q(x) = r(x) ∈ N by
definition of an ideal. Since g(x) is a nonzero element of minimal
degree in N , so r(x) = 0. Thus f (x) = g(x)q(x) ∈ hg(x)i and N ⊆
hg(x)i ⊆ N . So N = hg(x)i, i.e., F [x] is a PID. 
Similarly, we can easily show the following result.
Corollary 1.5.12. The ring (Z, +, ·) is a principal ideal domain.
Theorem 1.5.13. Let p(x) ∈ F [x]. Then hp(x)i is maximal in F [x]
if and only if p(x) is irreducible over F .
Proof. (⇒). Since hp(x)i is a maximal ideal of F [x], then hp(x)i =
6
{0} and hp(x)i =
6 F [x], so p(x) ∈/ F. Let p(x) = f (x)g(x) where
March 18, 2022 9:45 amsart-9x6 12819-main page 25

BASIC THEORY ON RINGS 25

f (x), g(x) ∈ F [x]. Since hp(x)i is a maximal ideal and hence also a
prime ideal, f (x)g(x) ∈ hp(x)i implies that
f (x) ∈ hp(x)i or g(x) ∈ hp(x)i.
We may assume that f (x) ∈ hp(x)i. Then f (x) = p(x)u(x) for some
u(x) ∈ F [x]. So p(x)u(x)g(x) = p(x), i.e., g(x)u(x) = 1. Thus g(x)
is a unit in F [x]. So p(x) is irreducible over F .
(⇐). Suppose that N  F [x] such that hp(x)i ⊆ N ⊆ F [x]. Since
N is a principal ideal, we may assume that N = hg(x)i for some
g(x) ∈ N. Then p(x) ∈ N implies that
p(x) = g(x)q(x) for some q(x) ∈ F [x].
Since p(x) is irreducible, either g(x) or q(x) is of degree 0. If g(x) is of
degree 0, that is, a nonzero constant in F , then g(x) is a unit in F [x].
Thus hg(x)i = N = F [x]. If q(x) is of degree 0, then q(x) = c ∈ F .
So g(x) = (1/c)p(x) is in hp(x)i, i.e., N = hp(x)i. Hence hp(x)i is
maximal. 
Theorem 1.5.14. Let p(x) be an irreducible polynomial in F [x]. If
p(x) r(x)s(x) for r(x), s(x) ∈ F [x], then either p(x)|r(x) or p(x)|s(x).
Proof. Suppose p(x)|r(x)s(x). Then r(x)s(x) ∈ hp(x)i, which is
maximal. Therefore, hp(x)i is a prime ideal. Hence r(x)s(x) ∈ hp(x)i
implies that either r(x) ∈ hp(x)i, yielding p(x)|r(x), or that s(x) ∈
hp(x)i, yielding p(x)|s(x). 
Example 1.5.4. Suppose that D is an integral domain. If D[x] is a
principal ideal domain, show that D is a field.
Proof. We need to show that any a ∈ D \ {0} has an inverse.
We have to use the given condition that D[x] is a principal ideal
domain. So consider the ideal ha, xi  D[x]. Since it is principal
there is f (x) ∈ D[x] such that ha, xi = hf (x)i. Since f |a we know
that f = d for some d ∈ D. At the same time d|x. The x = d(bx + c)
for some b, c ∈ D. It follows that bd = 1, i.e., d is invertible. Thus
ha, xi = D[x]. So
1 = xg(x) + ah(x)
for some g, h ∈ D[x]. We have 1 = ah(0). So a is invertible. There-
fore D is a field. 
Example 1.5.5. Prove that a prime ideal N in a commutative ring
R contains every nilpotent element. Deduce that the nilradical of R
March 18, 2022 9:45 amsart-9x6 12819-main page 26

26 RING AND FIELD THEORY

(the set of all nilpotent elements in R) is contained in the intersection


of all the prime ideals of R.
Proof. We first show by induction on n that a ∈ N if an ∈ N for
some n ∈ N. This is true for n = 1. If n > 1, since an = aan−1 ∈ N
we see that an−1 ∈ N . By inductive hypothesis we deduce that
a ∈ N.
Suppose b ∈ R is nilpotent, i.e., bn = 0 for some n ∈ N. Then
bn ∈ N . From the above established result we see that b ∈ N . 
Solution. Let x ∈ R with xn = 0 for some n ∈ N. Since N is a
prime ideal of R, we know that R/N in an ID. From (x + N )n = N
(the zero element in R/N ), we see that x + N = N . Thus x ∈ N . 
Example 1.5.6. Find all prime ideals and all maximal ideals of the
finite commutative ring (Z12 , +, ·).
Solution. Note that 12 = 22 3. We know that the only subgroups
of the cyclic group (Z12 , +) are
{0}, 2Z12 , 3Z12 , 4Z12 , 6Z12 , Z12 .
We can easily see that all of them are the ideals of Z12 . So they
are all the ideals of Z12 . Note that the prime and the maximal
ideals coincide for any finite unital commutative ring. The prime
and maximal ideals are
2Z12 = {0, 2, 4, 6, 8, 10} and 3Z12 = {0, 3, 6, 9}
because the factor rings are isomorphic to the fields
Z12 ∼ Z12 ∼
= Z2 and = Z3
2Z12 3Z12
respectively, and any other quotients are not fields. 
Example 1.5.7. Let F and K be fields. If F [x] ∼
= K[x], prove that

F = K.
Proof. Let ϕ : F [x] → K[x] be a ring isomorphism. Since U(K) =
K ∗ and U(F ) = F ∗ , we deduce that ϕ(F ∗ ) = K ∗ . Then ϕ|F : F → K
is an isomorphism of fields. So F ∼= K. 

1.6. Division algorithm for polynomials over a field.


In this section we always assume that F is field. The following the-
orem is the basic tool for our work in this section.
March 18, 2022 9:45 amsart-9x6 12819-main page 27

BASIC THEORY ON RINGS 27

Theorem 1.6.1 (Division Algorithm). Let f (x), g(x) ∈ F [x] with


g(x) 6= 0. Then there are unique polynomials q(x), r(x) ∈ F [x] such
that
f (x) = g(x)q(x) + r(x), with deg(r(x)) < deg(g(x)).
Proof. Let
f (x) = an xn + an−1 xn−1 + · · · + a0 ,
g(x) = bm xm + bm−1 xm−1 + · · · + b0 6= 0,
where bm 6= 0. Let
S = {f (x) − g(x)s(x) : s(x) ∈ F [x]}.
If 0 ∈ S then there exists an s(x) such that f (x) − g(x)s(x) = 0,
so f (x) = g(x)s(x). Taking q(x) = s(x) and r(x) = 0, we are done.
Otherwise, let r(x) be an element of minimal degree in S. Then
f (x) = g(x)q(x) + r(x)
for some q(x) ∈ F [x]. We must show that the degree of r(x) is less
than m. Suppose that
r(x) = ct xt + ct−1 xt−1 + · · · + c0 ,
with cj ∈ F and ct 6= 0. If t ≥ m, then
f (x) − q(x)g(x) − (ct /bm )xt−m g(x) = r(x) − (ct /bm )xt−m g(x),
and the latter is of the form r1 (x) = r(x) − (ct xt +terms of lower
degree), which is a polynomial of degree lower than t, the degree of
r(x). However, the polynomial can be written in the form
r1 (x) = f (x) − g(x)[q(x) + (ct /bm )xt−m ],
so r1 (x) ∈ S, contradicting the fact that r(x) was selected to have
minimal degree in S. Thus the degree of r(x) is less than the degree
m of g(x).
For uniqueness, suppose
f (x) = g(x)q1 (x) + r1 (x) = g(x)q2 (x) + r2 (x).
We see that g(x)[q1 (x) − q2 (x)] = r2 (x) − r1 (x). Because deg(r2 (x) −
r1 (x)) < m, this can hold only if q1 (x) − q2 (x) = 0 so q1 (x) = q2 (x).
Then we must have r2 (x) − r1 (x) = 0 so r1 (x) = r2 (x). 
We can compute the polynomials q(x) and r(x) of the above the-
orem by long division just as we divided polynomials in R[x] in high
school.
March 18, 2022 9:45 amsart-9x6 12819-main page 28

28 RING AND FIELD THEORY

Example 1.6.1. Let us work with polynomials in Z5 [x] and divide

f (x) = x4 − 3x3 + 2x2 + 4x − 1

by g(x) = x2 − 2x + 3 to find q(x) and r(x) of the theorem.

Solution. The long division should be easy to follow, but remem-


ber that we are working in Z5 [x]. For example, 4x − (−3x) = 2x.

x2 − x − 3
x2 − 2x + 3 x4 − 3x3 + 2x2 + 4x − 1

− x4 + 2x3 − 3x2
− x3 − x2 + 4x
x3 − 2x2 + 3x
− 3x2 + 7x − 1
3x2 − 6x + 9
x+8

Thus q(x) = x2 − x − 3, and r(x) = x + 3. 

Theorem 1.6.2 (Factor Theorem). Let f (x) ∈ F [x] and a ∈ F .


Then f (a) = 0 if and only if x − a|f (x).

Proof. From Theorem 1.6.1, we know that there are q(x) ∈ F [x]
and r ∈ F such that

f (x) = q(x)(x − a) + r.

Then f (a) = 0 if and only if r = 0, if and only if f (x) = q(x)(x − a),


if and only if x − a|f (x). 

Example 1.6.2. Again in Z5 [x], factor x4 + 3x3 + 2x + 4.

Solution. Note that 1 is a zero of x4 + 3x3 + 2x + 4. We should


be able to factor x4 + 3x3 + 2x + 4 into (x − 1)q(x) in Z5 [x]. Let us
March 18, 2022 9:45 amsart-9x6 12819-main page 29

BASIC THEORY ON RINGS 29

find the factorization by long division.


x3 + 4x2 + 4x + 6
x4 + 3x3

x−1 + 2x + 4
− x4 + x3
4x3
− 4x3 + 4x2
4x2 + 2x
− 4x2 + 4x
6x + 4
− 6x + 6
10
Thus x4 + 3x3 + 2x + 4 = (x − 1)(x3 − x2 − x + 1) in Z5 [x]. Since 1 is
seen to be a zero of x3 −x2 −x+1 also, we can divide this polynomial
by x − 1 and get
x2 −1
3 2

x−1 x −x −x+1
− x3 + x2
−x+1
x−1
0
Thus x4 + 3x3 + 2x + 4 = (x − 1)2 (x2 − 1) = (x − 1)3 (x + 1) in Z5 [x].

The next corollary should also look familiar.
Corollary 1.6.3. Any f (x) ∈ F [x] with deg(f (x)) = n ≥ 1 has at
most n distinct zeros in F .
Proof. We will prove this by induction on n. For n = 1, the
statement is clear. Now suppose that the statement holds for n =
k ≥ 1 and we consider n = k + 1. If f (x) does not have any zeros,
the statement holds clearly. If a1 ∈ F is a zero of f (x), then f (x) =
(x−a1 )g(x), where g(x) ∈ F [x] with deg(g(x)) = k. A zero a2 ∈ F of
f (x) with a2 6= a1 satisfies (a2 −a1 )g(a2 ) = 0, yielding that g(a2 ) = 0,
i.e., a2 is a zero of g(x). Since g(x) has at most k zeros, we conclude
that f (x) has at most k + 1 zeros in F . The proof completes. 
Let us recall a result on finitely generated abelian groups [ZTL,
Theorem 4.1.11].
March 18, 2022 9:45 amsart-9x6 12819-main page 30

30 RING AND FIELD THEORY

Theorem 1.6.4. Any finitely generated abelian group is isomorphic


to
Zr1 × Zr2 × · · · × Zrs × Zn
where s, n ∈ Z+ , ri ∈ N and r1 |r2 , r2 |r3 , · · · , rs−1 |rs .

Our next corollary is concerned with the structure of the multi-


plicative group F ∗ of nonzero elements of a field F , rather than with
factorization in F [x].

Corollary 1.6.5. If G is a finite subgroup of the multiplicative group


(F ∗ , ·, 1) of a field F , then G is cyclic. In particular, the multiplica-
tive group of all nonzero elements of a finite field is cyclic.

Proof. As a finite abelian group, G is isomorphic to a direct


product Cd1 × Cd2 × · · · × Cdr , where each di |di+1 , and each of the
Cdi is a cyclic group of order di in multiplicative notation. If ai ∈ Cdi ,
then adi i = 1, so adi r = 1 since di divides dr . Thus for all α ∈ G, we
have αdr = 1, so every element of G is zero of xdr − 1. But G has
d1 d2 · · · dr elements, while xdr − 1 can have at most dr zeros in the
field F , so d1 d2 · · · dr ≤ dr , we deduce that d1 = d2 = · · · = dr−1 = 1.
Therefore G is isomorphic to the cyclic group Cdr . 

Corollary 1.6.6. If p is a prime, then


(a). np−1 ≡ 1 (mod p) for any n ∈ Z with p 6 |n;
(b). np ≡ n (mod p) for any n ∈ Z.

Proof. Applying the above corollary we obtain the statements.




Example 1.6.3. Find all zeros in Z5 for the polynomial 2x219 +


3x74 + 2x57 + 3x44 .

Solution. We can write 2x219 +3x74 +2x57 +3x44 = 2(x219 −x74 +


x57 − x44 ). Clearly 0 is a solution for f (x) = x219 − x74 + x57 − x44 .
If a ∈ Z5 \ {0}, we see that a4 = 1 in Z5 . Then

f (a) = a3 − a2 + a − 1 = (a − 1)(a2 + 1) = (a − 1)(a2 − 22 )


= (a − 1)(a + 2)(a − 2).

Thus all zeros for the polynomial are 0, 1, 2, −2. 


March 18, 2022 9:45 amsart-9x6 12819-main page 31

BASIC THEORY ON RINGS 31

1.7. Irreducible polynomials over a field.


We consider polynomials over a field F in this section.
Definition 1.7.1. A nonconstant polynomial f (x) ∈ F [x] is irre-
ducible over F or is an irreducible polynomial in F [x] if f (x)
cannot be expressed as a product g(x)h(x) of two polynomials g(x)
and h(x) in F [x] both of lower degree than deg(f (x)).
This definition concerns the concept irreducible over F and not
just the concept irreducible. A polynomial f (x) may be irreducible
over F , but may not be irreducible if viewed over a larger field E
containing F.
Example 1.7.1. We know that x2 − 3 viewed in Q[x] has no zeros in
Q. This shows that x2 − 3 is irreducible over Q, for a factorization
x2 − 3 = (ax + b)(cx + d) for a, b, c, d ∈ Q would give rise to zeros of
x2 − 3 in Q. However, x2 − 3 viewed in R[x] √ is not √
irreducible over
R, because x2 − 3 factors in R[x] into (x − 3)(x + 3).
Note that the units in F [x] are precisely the nonzero elements of
F , i.e., U(F [x]) = F ∗ . Thus we could have defined an irreducible
polynomial f (x) as a nonconstant polynomial such that in any fac-
torization f (x) = g(x)h(x) in F [x], either g(x) or h(x) is a unit.
Irreducible polynomials will play a very important role in almost
everywhere. The problem of determining whether a given f (x) ∈
F [x] is irreducible over F may be difficult. We now give some criteria
for irreducibility that are useful in certain cases.
Theorem 1.7.2. Let f (x) ∈ F [x], and let f (x) be of degree 2 or 3.
Then f (x) is reducible over F if and only if it has a zero in F .
Proof. “⇐”. Suppose that f (x) has a root a in F . By Theo-
rem 1.6.2 we know that x − a|f (x), i.e., f (x) = (x − a)g(x) for some
g ∈ F [x] with positive degree since deg(f (x)) = 2 or 3. Thus f (x) is
reducible.
“⇒”. Suppose that f (x) is reducible. Then f (x) = g(x)h(x)
where g(x), h(x) ∈ F [x] of positive degrees. Let deg(g(x)) = r or
deg(h(x)) = s. We see that r ≥ 1, s ≥ 1 and r + s = 2 or 3 since
deg(f (x)) = 2 or 3. We deduce that r = 1 or s = 1. We may assume
that r = 1, i.e., g(x) = ax − b where a ∈ F ∗ , b ∈ F . Thus g(x) has a
root in b/a ∈ F , and hence f (b/a) = 0. 
Example 1.7.2. Show that f (x) = x3 + 3x + 2 ∈ Z5 [x] is irreducible
over Z5 .
March 18, 2022 9:45 amsart-9x6 12819-main page 32

32 RING AND FIELD THEORY

Solution. We compute that


f (0) = 2, f (1) = 1, f (−1) = −2, f (2) = 1, and f (−2) = −2,
showing that f (x) has no zeros in Z5 . From Theorem 1.7.2 we know
that f (x) is irreducible over Z5 . 
We turn to some conditions for irreducibility of polynomials in
Q[x].
Definition 1.7.3. f (x) = n0 ai xi ∈ Z[x] is primitive if degf > 0
P
and the coefficients a0 , . . . , an are relatively prime.
Lemma 1.7.4 (Gauss’ Lemma). If f (x), g(x) ∈ Z[x] are primitive,
then so is f (x)g(x).
Proof. Let f (x) = m ai xi , g(x) = ni=0 bi xi , and f (x)g(x) =
P P
P l P i=0
cl x , where cl = i+j=l ai bj . Let p ∈ Z be prime. It suffices to
show that there exists l such that p - cl . Suppose
s = min{i : p - ai } and t = min{i : p - bi }.
Then X X
cs+t = as bt + ai bj + ai bj ,
i+j=s+t i+j=s+t
i<s j<t
so p - cs+t since p - as bt but p divides all the other terms on the right
hand side. So f (x)g(x) is primitive. 
Lemma 1.7.5. Let f (x) ∈ Q[x] \ {0}. Then
(1). there exists c ∈ Q and primitive g(x) ∈ Z[x] such that f (x) =
cg(x);
(2). if also f (x) = dh(x) with d ∈ Q and primitive h(x) ∈ Z[x],
then c = ±d and g(x) = ±h(x).
Proof. (i) Clearly there is a ∈ Z∗ such that af (x) = n0 ai xi ∈
P
Z[x]. LetPb = gcd(a0 , . . . , an ). For each i put bi = ai /b, and then put
g(x) = n0 bi xi . We have f (x) = ab g(x) by construction, and g(x) is
primitive since the bi are relatively prime.
0
(ii) If also f (x) = ab 0 h(x) with a0 , b0 ∈ Z and primitive h(x) ∈ Z[x],
then a0 bg(x) = ab0 h(x) ∈ Z[x]. Since both |a0 b| and |ab0 | are the
GCD of the coefficients of a0 bg(x) = ab0 h(x), so b/a = ±b0 /a0 , and
h(x) = ±g(x). 
Lemma 1.7.6. Let f (x) ∈ Z[x] be primitive. Then f (x) is reducible
in Z[x] if and only if f (x) is reducible in Q[x].
March 18, 2022 9:45 amsart-9x6 12819-main page 33

BASIC THEORY ON RINGS 33

Proof. “⇐”. Assume that f (x) is reducible in Q[x], and choose


g(x), h(x) ∈ Q[x] such that f (x) = g(x)h(x) and degg(x), degh(x) >
0. From Lemma 1.7.5, there exist a, b ∈ Q∗ that ag(x), bh(x) ∈
Z[x] are primitive. So (ab)−1 (ag(x))(bh(x)) = f (x). Since both
(ag(x))(bh(x)), f (x) are primitive, from Lemma 1.7.5 we know that
f (x) = ±(ag(x))(bh(x)). Thus f (x) is reducible in Z[x].
“⇒”. Assume that f (x) is reducible in Z[x], and choose noncon-
stant g(x), h(x) ∈ Z[x] such that f (x) = g(x)h(x). Since f (x) is
primitive, then g(x), h(x) are primitive, and degg(x), degh(x) > 0.
Clearly f (x) is reducible in Q[x]. 
Theorem 1.7.7. Let f (x) ∈ Z[x]. Then f (x) factors into a product
of two polynomials of lower degrees r and s in Q[x] if and only if it
has such a factorization with polynomials of the same degrees r and
s in Z[x].
Proof. “⇒”. Suppose f (x) = g(x)h(x) where h(x), g(x) ∈ Q[x]
and degg(x), degh(x) < degf (x). From Lemma 1.7.5, there exist
a, b ∈ Q∗ such that ag(x), bh(x) ∈ Z[x] are primitive, and there is
c ∈ Z∗ such that f (x) = cf1 (x) where f1 (x) ∈ Z[x] is primitive.
So (ab)−1 (ag(x))(bh(x)) = cf1 (x). Since both (ag(x))(bh(x)), f1 are
primitive, from Lemma 1.7.5, we know that f1 (x) = ±(ag(x))(bh(x)).
Thus f (x) = ±c(ag(x))(bh(x)) in Z[x].
“⇐”. This is trivial since Z[x] is a subring of Q[x]. 
Corollary 1.7.8. Let f (x) = a0 +a1 x+· · ·+an xn ∈ Z[x] with n ≥ 1
and a0 an 6= 0. If p/q, where p, q ∈ Z∗ with gcd(p, q) = 1, is a zero of
f (x), then p|a0 , q|an .
Proof. From f (p/q) = 0 we see that
a0 q n + a1 pq n−1 + · · · + an−1 pn−1 q + an pn = 0.
So p|a0 q n and q|an pn . Since gcd(p, q) = 1, then p|a0 and q|an . 
Example 1.7.3. (a). Is Q[x]/hx2 − 6x − 6i a field? Why?
(b). Find all c ∈ Z5 so that Z5 [x]/hx2 + x + ci is a field.
Solution. (a). We know that Q[x]/hx2 − 6x − 6i is a field iff
hx2− 6x − 6i is maximal if and only if x2 − 6x − 6 is irreducible.
Possible rational zeros of x2 −6x−6 are ±1, ±2, ±3, ±6. By testing
none of them are. So x2 −6x−6 is irreducible over Q. Thus Q[x]/hx2 −
6x − 6i is a field.
March 18, 2022 9:45 amsart-9x6 12819-main page 34

34 RING AND FIELD THEORY

(b). We know that f (x) = x2 + x + c is irreducible if and only if


f (x) has non zeros in Z5 . Note that
f (0) = f (−1) = c, f (1) = f (−2) = c + 2, f (2) = c + 1.
Thus f (x) with c = 0, −1, −2 is not irreducible. So c = 1, 2. 
Example 1.7.4. Show that f (x) = x4 + 8x3 − 2x2 + 1 is irreducible
over Q.
Solution. If f (x) has a linear factor in Q[x], then it has a zero
in Z, and this zero would have to be a divisor in Z of 1, that is,
either ±1. But f (1) = 8, and f (−1) = −8, so such a factorization is
impossible. If f (x) factors into two quadratic factors in Q[x], then it
has a factorization.
(x2 + ax + b)(x2 + cx + d)
in Z[x]. Equating coefficients of powers of x, we find that we must
have
bd = 1, ad + bc = 0, ac + b + d = −2, and a + c = 8
for integers a, b, c, d ∈ Z. From bd = 1, we see that either b = d = 1
or b = d = −1. In any case, b = d and from ad + bc = 0, we deduce
that a + c = 0 which is impossible. Thus a factorization into two
quadratic polynomials is also impossible and f (x) is irreducible over
Q. (We will have a much shorter proof for this in Example 1.8.1.) 
No we provide the famous Schönemann-Eisenstein criterion for
irreducibility.
Theorem 1.7.9 (Schönemann-Eisenstein Criterion). Let f (x) =
an xn + · · · + a1 x + a0 ∈ Z[x] with n ≥ 1 and an 6= 0. If there is
a prime p such that
(i). p|ai for 0 ≤ i < n,
(ii). p 6 |an ,
(iii). p2 6 |a0 ,
then f (x) is irreducible in Q[x].
Proof. To the contradiction, assume that f (x) is reducible in
Q[x]. Then f (x) = g(x)h(x) for some
g(x) = br xr + · · · + b1 x + b0 ∈ Q[x] \ Q,
h(x) = cs xs + · · · + c1 x + c0 ∈ Q[x] \ Q.
March 18, 2022 9:45 amsart-9x6 12819-main page 35

BASIC THEORY ON RINGS 35

From Theorem 1.7.7, we may assume that bi , cj ∈ Z. Denoting by


f (x), g(x), h(x) the reductions mod p of these polynomials (i.e., we
consider the coefficients as elements in Zp ), we have g(x)h(x) =
f (x) = an xn , which means that g(x) = br xr and h(x) = cs xs . This
shows that p|b0 and p|c0 , yielding that p2 |b0 c0 , i.e., p2 |a0 , which
contradicts the hypothesis of the theorem. 
This criterion is named after Theodor Schönemann (1812–1868)
and Gotthold Eisenstein (1823–1852). It was also known as the
Eisenstein Criterion. Schönemann was the first to publish his proof
in 1846, and then by Eisenstein in 1850.
Note that if we take p = 2, the Schönemann-Eisenstein Criterion
gives us still another proof of the irreducibility of x2 − 2 over Q.
Example 1.7.5. Taking p = 3 in the above theorem, we see that
25x5 − 9x4 − 3x2 − 12 is irreducible over Q.
Corollary 1.7.10. The polynomial
Φp (x) = (xp − 1)/(x − 1) = xp−1 + xp−2 + · · · + x + 1
is irreducible over Q for any prime p.
Proof. It is clear that Φp (x) is irreducible over Q iff Φp (x + 1) is
irreducible over Q. Let
xp + p1 xp−1 + · · · + px

p
g(x) = Φp (x+1) = ((x+1) −1)/((x+1)−1) = .
x
The coefficient of xp−r for 0 < r < p is the binomial coefficient
p!/(r!(p−r)!) which is divisible by p because p|p! but p6 |r! or p6 |(p−r)!
when 0 < r < p. Thus
Ç å
p−1 p p−2
g(x) = x + x + ··· + p
1
satisfies the Schönemann-Eisenstein Criterion for the prime p and is
thus irreducible over Q. Thus Φp (x) must also be irreducible over Q.

The polynomial Φp (x) in the above corollary is called the p-th
cyclotomic polynomial.
The p-th cyclotomic polynomial Φp (x) is generally not irreducible
over a field of finite characteristic.
Example 1.7.6. In Z2 [x],
Φ7 (x) = x6 + x5 + · · · + x + 1 = (x3 + x2 + 1)(x3 + x + 1).
March 18, 2022 9:45 amsart-9x6 12819-main page 36

36 RING AND FIELD THEORY

Using Theorem 1.7.2 one can further show that both polynomials x3 +
x2 + 1, x3 + x + 1 are irreducible over Z2 .
Theorem 1.7.11 (Unique Factorization Theorem). For any f (x) ∈
F [x] of positive degree, there exist a unique constant c; unique dis-
tinct irreducible monic polynomials f1 (x), f2 (x), · · · , fk (x) ∈ F [x];
and unique positive integers n1 , n2 , · · · , nk such that
f (x) = cf1 (x)n1 f2 (x)n2 · · · fk (x)nk .
Proof. Write f (x) as a product of as many as possible positive
polynomials
f (x) = g1 (x)g2 (x) · · · gr (x)
where each gi (x) is of positive degree and r is maximal. Then each
gi (x) is irreducible. We may assume that f (x) = ch1 (x)h2 (x) · · · hr (x),
where each hi (x) is irreducible and monic.
Let us prove the uniqueness by induction on r. Suppose f (x) has
another decomposition of monic irreducible polynomials:
f (x) = cp1 (x)p2 (x) · · · ps (x),
where each pi (x) is irreducible and monic. If r = 1 the uniqueness is
clear. Now assume that r > 1. From Theorem 1.5.14 there is a pj (x)
such that h1 (x) = pj (x). Without loss of generality we may assume
that h1 (x) = p1 (x). The we obtain that
h2 (x) · · · hr (x) = p2 (x) · · · ps (x).
Using inductive hypothesis we deduce that r = s and hi (x) = pi (x)
for all i = 1, 2, . . . , r after renumbering pi (x) if necessary. 
Example 1.7.7. By Schönemann-Eisenstein Criterion, we know that
x4 − 4x3 + 2x − 2 is irreducible over Q. But in Z7 [x],
x4 − 4x3 + 2x − 2 = (x − 2)3 (x + 2) = (x − 1)2 (2x − 4)(4x + 8).
Example 1.7.8. Find primes p so that x + 2|x4 + x3 + x2 − x + 1
in Zp [x].
Solution. Let f (x) = x4 + x3 + x2 − x + 1. From Factor The-
orem 1.6.2 we know that x + 2 f (x) if and only if f (−2) = 0, i.e.,
15 = 0. Thus p = 3, or 5.
You may use Division Algorithm to solve this question. 
It is generally hard to check whether a polynomial is irreducible
over Q. Here we provide a method to help do so.
March 18, 2022 9:45 amsart-9x6 12819-main page 37

BASIC THEORY ON RINGS 37

Proposition 1.7.12. Let f (x) ∈ Z[x]. Suppose that f (x)Zp [x] is the
reduction of f (x) modulo p with deg(f ) = deg(f ). If f (x) is reducible
in Q[x], then f (x) is reducible in Zp [x]. Or by the contrapositive: if
f (x) is irreducible in Zp [x], then f (x) is irreducible in Q[x].
Example 1.7.9. Show that the polynomial f (x) = x5 + (2a + 1)x2 +
(2b + 1) ∈ Z[x] is irreducible for any integers a, b ∈ Z.
Proof. Consider the polynomial in Z2 [x]. We have f (x) =
x5 + x2 + 1. Since f (0)f (1) 6= 0 we know that f (x) does not have
degree one factors. If f (x) is not irreducible in Z2 [x], it must have an
irreducible of degree 2 in Z2 [x]. Since x2 +x+1 is the only irreducible
of degree 2 in Z2 [x], and
f (x) = (x2 + x + 1)(x3 + x2 ) + 1,
we see that x2 +x+1 6 |f (x). Thus f (x) is irreducible for any integers
a, b ∈ Z. 
1.8. Other irreducibility criteria.
We will provide several other important simple-to-use irreducibility
criteria for integer polynomials. The first one is the following Perron’s
irreducibility criterion which was first proved by Oskar Perron (1880–
1975) in 1907 using Complex Analysis.
Theorem 1.8.1 (Perron’s Irreducibility Criterion). Suppose
f (x) = xn + an−1 xn−1 + · · · + a1 x + a0 ∈ Z[x]
where a0 6= 0. If either of the following two conditions applies:
|an−1 | > 1 + |an−2 | + · · · + |a0 |; (1.2)
|an−1 | = 1 + |an−2 | + · · · + |a0 | and f (±1) 6= 0, (1.3)
then f (x) is irreducible over Q.
Let us first recall Rouche’s Theorem from Complex Analysis.
Theorem 1.8.2 (Rouche’s Theorem). Let f (z) and g(z) be analytic
functions on and inside a simple closed curve C. Suppose that |f (z)+
g(z)| < |f (z)| + |g(z)| for all points z on C. Then f (z) and g(z) have
the same number of zeros (counting multiplicities) interior to C.
Lemma 1.8.3. Let f (z) = z n + an−1 z n−1 + · · · + a1 z + a0 be as in
Theorem 1.8.1. Then exactly one zero z of f (x) satisfies |z| > 1, and
the other n − 1 zeros of f (z) satisfy |z| < 1.
March 18, 2022 9:45 amsart-9x6 12819-main page 38

38 RING AND FIELD THEORY

Proof. (a). Suppose that |an−1 | > 1 + |an−2 | + · · · + |a0 |. We need


to apply Rouche’s Theorem to the functions g(z) = z n + an−1 z n−1
and f (z). For |z| = 1 we have
|f (z) − g(z)| = |an−2 z n−2 + · · · + a0 |
(1.4)
≤ |an−2 | + · · · + |a0 | < |an−1 | − 1,
|f (z)| + |g(z)| ≥ |g(z)| = |z + an−1 | ≥ |an−1 | − 1. (1.5)
It follows that f (z) has the same number of zeros as z + an−1 z n−1
n

inside the unit circle. We know that z n + an−1 z n−1 has n − 1


zeros inside the unit circle. It follows that f (z) has exactly n − 1
zeros inside the unit circle. It is easy to see that f (z) has no zeros
on the unit circle.
(b). Suppose that |an−1 | = 1 + |an−2 | + · · · + |a0 | and f (±1) 6= 0.
Equation (1.4) becomes |f (z) − g(z)| ≤ |an−1 | − 1. Now in Equa-
tion (1.5), if the second inequality is not strict, then z = ±1 and
the first inequality in Equation (1.5) is strict. Thus we also have
|f (z) − g(z)| < |f (z)| + |g(z)| on the unit circle. The remaining
argument is the same as (a). 
Proof of Theorem 1.8.1. Suppose that f (x) = g(x)h(x) where
g(x) and h(x) are integer polynomials. Since, by the above Lemma,
f (x) has only one zero with modulus not less than 1, one of the
polynomials g(x), h(x) has all its zeroes strictly inside the unit circle.
Suppose that z1 , . . . , zk are the zeroes of g(x), and |z1 |, . . . , |zk | < 1.
Note that g(0) is a nonzero integer, and |g(0)| = |z1 · · · zk | < 1,
contradiction. Therefore, f (x) is irreducible. 
Example 1.8.1. Let us revisit Example 1.7.4. Using Perron’s Irre-
ducibility Criterion, we easily see that the polynomial f (x) = x4 +
8x3 − 2x2 + 1 is irreducible over Q.
It is interesting to know the following generalization. We will omit
its proof.
Theorem 1.8.4 (Perron’s Irreducibility Criterion). Let F be a field.
Suppose
f (x, y) = an (y)xn + an−1 (y)xn−1 + · · · + a1 (y)x + a0 (y) ∈ F [x, y]
where ai (y) ∈ F [y] with a0 an 6= 0 and an (y) ∈ F . If
deg(an−1 ) > max{deg(a0 ), deg(a1 ), · · · , deg(an−2 )}
then f (x, y) is irreducible over F (y).
March 18, 2022 9:45 amsart-9x6 12819-main page 39

BASIC THEORY ON RINGS 39

The other simple-to-use criterion is Cohn’s irreducibility criterion


which was first proved by Arthur Cohn (1894–1940) in 1925 using
Number Theory. We omit its long proof here.
Theorem 1.8.5 (Cohn’s Irreducibility Criterion). Let
f (x) = an xn + an−1 xn−1 + · · · + a1 x + a0 ∈ Z[x]
where all ai ≥ 0. Suppose that b ∈ N with b ≥ 2 and 0 ≤ ai ≤ b − 1.
If f (b) is a prime, then f (x) is irreducible over Q.
We introduce our next simple-to-use criterion which was proved
by Hiroyuki Osada in 1987.
Theorem 1.8.6 (Osada’s Irreducibility Criterion). Let p be a prime,
f (x) = an xn + an−1 xn−1 + · · · + a1 x ± p ∈ Z[x],
with n ≥ 1 and an 6= 0. If p > |a1 |+· · ·+|an |, then f (x) is irreducible
over Q.
Proof. Let α be any complex zero of f (x). Suppose that |α| ≤ 1,
then p = |a1 α + · · · + an αn | ≤ |a1 | + · · · + |an |, a contradiction.
Therefore, all the zeros of f (x) satisfies |α| > 1. Now, suppose that
f (x) = g(x)h(x), where g(x), h(x) are nonconstant integer polynomi-
als. Then a0 = f (0) = g(0)h(0). Since p is prime, one of |g(0)|, |h(0)|
equals 1. Say |g(0)| = 1, and let b be the leading coefficient of g(x).
If α1 , · · · , αk are the roots of g(x), then |α1 α2 · · · αk | = 1/|b| ≤ 1.
However, α1 , · · · , αk are also zeros of f (x), and so each has an mag-
nitude greater than 1. Contradiction. Therefore, f (x) is irreducible.

The last simple-to-use criterion is named after Alfred Brauer (1894–
1985) who proved it in 1951.
Theorem 1.8.7 (Brauer’s Irreducibility Criterion). Let a1 ≥ a2 ≥
· · · ≥ an be positive integers and n ≥ 2. Then the polynomial f (x) =
xn − a1 xn−1 − · · · − an−1 x − an is irreducible over Q.
Proof. If all zeros of f (x) are negative or non-real, an would be
negative. So f (x) has at least one non-negative zeros, say α. Then
we see that
αn = a1 αn−1 + · · · + an−1 α + an > a1 αn−1 .
So α > a1 ≥ 1. Now we show that all other zeros β of f (x) satisfy
|β| < 1.
March 18, 2022 9:45 amsart-9x6 12819-main page 40

40 RING AND FIELD THEORY

Consider the polynomial g(x) = (x − 1)f (x). Clearly,


g(x) = xn+1 − b1 xn + b2 xn−1 + · · · + bn+1 ,
where b1 = a1 + 1, b2 = a1 − a2 , · · · , bn = an−1 − an , bn+1 = an . The
numbers b1 , · · · , bn+1 are non-negative integers and b1 = 1 + b2 +
· · · + bn+1 with bn+1 ≥ 1. Let
h(z) = b1 z n − b2 z n−1 − · · · − bn+1 .
For all sufficiently small  > 0, first we show that
|h(z)| > |z n+1 | = |g(z) + h(z)|, if |z| = 1 + .
Indeed, if |z| = 1 + , then
|h(z)| − |z n+1 | ≥ b1 (1 + )n − b2 (1 + )n−1 − · · · − bn+1 − (1 + )n+1
= (nb1 − (n − 1)b2 − · · · − 2bn−1 − bn − (n + 1))
+ (higher terms in )
= (b2 + 2b3 + · · · + (n − 1)bn + nbn+1 − 1)
+ (higher terms in ).
The coefficient of  is positive. For sufficiently small  > 0, we have
|h(z)| − |z n+1 | > 0. In this case
|g(z) + h(z)| = |z n+1 | < |h(z)| ≤ |g(z)| + |h(z)|.
Therefore, by Rouche’s theorem, the polynomial g(z) and h(z) have
the same number of roots inside the disk |z| ≤ 1 + .
If |z| ≥ 1, then
|h(z)| ≥ b1 |z|n − b2 |z|n−1 − · · · − bn+1
≥ |z|n (b1 − b2 − · · · − bn+1 ) = |z|n > 0.
So all the n roots of h(z) lie strictly inside the unit disk |z| ≤ 1.
Letting  → 0, we see that inside and on the boundary of the unit
disk there are exactly n roots of the polynomial g(x) = (x − 1)f (x).
If |z| = 1 and z 6= 1, we claim that g(z) 6= 0. Otherwise there is
z0 such that |z0 | = 1, z0 6= 1 and g(z0 ) = 0. Then
|z0n+1 − b1 z0n | = |b2 z0n−1 + · · · + bn z0 + bn+1 |,
|z0n+1 − b1 z0n | > b1 − 1 = b2 + · · · + bn+1
≥ |b2 z0n−1 + · · · + bn z0 + bn+1 |,
March 18, 2022 9:45 amsart-9x6 12819-main page 41

BASIC THEORY ON RINGS 41

which is a contradiction. Hence exactly n − 1 roots of f (x) lie inside


the unit disk.
Using the similar arguments in the proof of Theorem 1.8.1, we see
that f (x) is irreducible. 
We remark that the proofs for Lemma 1.8.3 and Theorem 1.8.7
are modified from that in [P].

Example 1.8.2. For any positive integer n, show that x2n − 2xn − 7
is irreducible in Z[x].

Proof. The constant term 7 is a prime satisfying 7 > 1 + | − 2|.


From Osada’s Irreducibility Criterion we know that x2n − 2xn − 7 is
irreducible in Z[x]. 

1.9. Exercises.
(1) Let R be a ring that contains at least two elements. Suppose
for each a ∈ R there is a unique ϕ(a) ∈ R such that aϕ(a)a =
a. Show that R is a division ring.
Hints: You may follow the steps below.
(a). Show that R has no zero divisors.
(b). Show that ϕ(a)aϕ(a) = ϕ(a).
(c). Show that R has unity.
(d). Show that R is a division ring.
(2) Let R be a unital ring. If (xy)2 = x2 y 2 for all x, y ∈ R, show
that R is commutative.
(3) Let R be a unital ring such that x6 = x for all x ∈ R.
(a). Prove that x2 = x for all x ∈ R.
(b). Prove that R is commutative.
(4) Define the center of a ring R as Z(R) = {a ∈ R : ab =
ba, ∀b ∈ R}. Prove that a ring R is commutative if a2 − a ∈
Z(R), for all a ∈ R.
(5) Let (R, +, ·) be a finite ring with at least two distinct ele-
ments. Suppose that the multiplication satisfies the cancel-
lation rules. Show that R is a division ring.
(6) Consider (S, +, ·), where S is a set and + and · are binary
operations on S such that
(a). (S, +) is a group,
(b). (S ∗ , ·) is a group where S ∗ = S \{0} and 0 is the additive
identity element,
March 18, 2022 9:45 amsart-9x6 12819-main page 42

42 RING AND FIELD THEORY

(c). a(b + c) = (ab) + (ac) and (a + b)c = (ac) + (bc) for all
a, b, c ∈ S.
Show that (S, +, ·) is a division ring.
(7) Show that in any unital ring R the commutative law for addi-
tion is redundant, in the sense that it follows from the other
axioms for a ring.
(8) Determine all irreducible polynomials of degree 3 in Z2 [x].
Justify your answer.
(9) P R be a finite ring without zero-divisors. Show that
Let
2 r∈R r = 0.
(10) Let a, b be elements of the unital ring R. Show that 1 − ab is
a unit if and only if 1 − ba is a unit.
(11) Find all real roots of the polynomial 2x4 +3x3 +3x−2 ∈ R[x].
(12) Factor the polynomial f (x) = x5 + 3x3 + x2 + 2x ∈ Z5 [x] into
a product of irreducible polynomials.
(13) Whether is the rational polynomial x3 + 3x2 − 8 irreducible?
(14) Let R be a commutative ring, let a be a unit of R, and let b
be any element of R. Define a function φ : R[x] → R[x] by
φ(f (x)) = f (ax + b), for all f (x) ∈ R[x]. Show that φ is an
automorphism of R[x].
(15) In the polynomial ring Z[x], show that the ideal hn, xi gen-
erated by n ∈ Z and x is a prime ideal if and only if n is a
prime number.
(16) Show that any principal ideal in the polynomial ring Z[x]
cannot be a maximal ideal.
(17) Let P be a prime ideal of a commutative ring R. Prove that
A ∩ B ⊆ P implies A ⊆ P or B ⊆ P , for all ideals A, B of R.
Give an example to show that the converse is false.
(18) Let R be a commutative unital ring. Prove that if every
proper ideal of R is prime, then R is a field.
(19) Let R be a commutative ring with 1. Show that the intersec-
tion of all the prime ideals of R is precisely the set of nilpotent
elements of R.
(20) Let R be a commutative ring with 1 and let {Pi } be a chain
of prime ideals in R. Prove that ∩Pi is prime. Deduce that
every prime ideal P of R contains a minimal prime ideal of R.
(21) Let R be a commutative ring with 1 and let I be the intersec-
tion of all the maximal ideals of R. Prove that a ∈ I if and
only if 1 + ax is a unit in R for all x ∈ R.
March 18, 2022 9:45 amsart-9x6 12819-main page 43

BASIC THEORY ON RINGS 43

(22) Let R be a finite unital ring, and x ∈ R be nonzero. Show


that x is a left 0-divisor if and only if it is a right 0-divisor.
(23) Let R denote the real numbers. What familiar ring is isomor-
phic to R[x]/hx2 − x + 1i? Prove your assertion. Prove that
Z[x]/hx2 + 1i is an integral domain, and find its cardinality.
(24) Let Qn be the space of all n-tuples of rational numbers, made
into a ring by component-wise addition and multiplication.
Find all ring homomorphisms of Qn onto Q and all ring ho-
momorphisms of Qn onto Qn .
(25) Show that the ideal of Z[x] generated by 2 and x4 + x2 + 1 is
not maximal.
(26) Let R1 = Zp [x]/hx2 − 3i, R2 = Zp [x]/hx2 + 2i. Determine
whether R1 and R2 are isomorphic for p = 2, 5, 11 respec-
tively.
(27) Let F be one of the following fields R, Q, C, Z9 . Let I ⊂ F [x]
be the ideal generated by x4 + 2x − 2. For which choices of
F is the ring F [x]/I a field?
(28) Let R = C([0, 1]) be the ring of continuous real-valued func-
tions on the interval [0, 1] with the usual definition of sum
and product of functions from calculus. Show that f ∈ R
is a zero divisor if and only if f is not identically zero and
{x : f (x) = 0} contains an open interval. What are the
idempotents of this ring? What are the nilpotents? What
are the units?
(29) Let R = (Z2 , +, ·) and let G = (Z2 , +), the group of order 2.
Find a quotient of a polynomial ring which is isomorphic to
the group ring R[G].
(30) Let G be a finite elementary 2-group (i.e., a direct product
of finitely many copies of Z2 , the cyclic group of order 2).
(a). Show that Z[G] has zero divisors.
(b). Show that there is a one-to-one correspondence between
ring homomorphisms ψ : Z[G] → Z and group homo-
morphisms χ : G → {±1}.
(c). The augmentation mapping is the homomorphism Z[G]
→ Z defined by sending all g ∈ G to 1. Let I be the
kernel of this homomorphism. What are all the maximal
ideals containing I?
(d). Let P be a minimal prime ideal of Z[G]. You may assume
that P ∩ Z = 0.
March 18, 2022 9:45 amsart-9x6 12819-main page 44

44 RING AND FIELD THEORY

(e). Let M be the maximal ideal from (d) that contains I


and 2. Show that P ⊂ M.
(31) 4
Is x + 1 irreducible over the field of real numbers? over the
field of rational numbers? over a field with 16 elements?
(32) In R[x], consider the set of polynomials f (x) for which f (2) =
f 0 (2) = f 00 (2) = 0. Prove that this set forms an ideal and find
its monic generator. Do the polynomials such that f (2) =
f 0 (3) = 0 form an ideal? Prove or give a counterexample.
(33) Let R be a commutative ring with identity, and √ let A, B be
ideals of R. The radical of A is defined to be A = {r ∈ R :
rn ∈ A for some n ≥ 0}.
(a). Show A is an√ideal of √ R.
(b). Show that if A + B = R, then A + B = R.
(34) Let R be a ring. Show that the two-sided ideals of the ring
Mn (R) of n × n matrices over R are precisely the subsets of
the form
Mn (I) = {(aij ) : aij ∈ I ∀ i, j},
where I is some ideal of R.
(35) Show that the ring of even integers contains a maximal ideal
M such that E/M is not a field.
(36) If N is the ideal of all nilpotent elements in a commutative
ring R, show that R/N is a ring with no nonzero nilpotent
elements.
(37) Show that the maximal ideals of the polynomial ring C[x] are
principal ideals generated by x − c for some c ∈ C.
(38) Show that the prime ideals of C[x, y] are h0i, hf (x, y)i for
an irreducible f (x, y) ∈ C[x, y] and hx − a, y − bi for some
a, b ∈ C.
(39) Let R be a commutative ring. A minimal prime ideal in
R is a prime ideal I such that if J ⊂ I is a prime ideal of
R then I = J. Show that every prime ideal I contains a
nonzero minimal prime ideal. (Hints: Use Kuratowski-Zorn
lemma.)
(40) Let P1 , P2 , · · · , Pn , n ≥ 2, be ideals in a ring R, with P3 , · · · , Pn
prime (if n ≥ 3). Let P be any ideal of R. If P ⊆ ∪ni=1 Pi ,
show that P ⊆ Pi for some i.
(41) Show that all nonzero prime ideals are maximal in a principal
ideal domain.
March 18, 2022 9:45 amsart-9x6 12819-main page 45

BASIC THEORY ON RINGS 45

(42) Let D be an integral domain for which IJ = I ∩ J for all


ideals I, J of D. Prove that D is a field.
(43) A valuation ring R is a commutative unital ring in which,
for any a and b in R, either a|b or b|a. Show that the non-
units (i.e., non-invertible elements) of a valuation ring R form
an ideal that is necessarily the unique maximal ideal in R.
(44) Show that End ((Z, +)) ∼ = (Z, +, ·) and that End((Zn , +)) ∼ =
(Zn , +, ·).
(45) Show that End((Z2 × Z2 , +)) is not isomorphic to (Z2 ×
Z2 , +, ·).
(46) Find a subring of the ring (Z × Z, +, ·) that is not an ideal of
Z × Z.
(47) Find all ideals of the ring (F [x] × F [x], +, ·) where F is a
field.
(48) Let F be a field, and let S be any subset of F × F × · · · × F
for n factors. Show that the set NS of all f (x1 , · · · , xn ) ∈
F [x1 , · · · , xn ] that have every element (a1 , · · · , an ) of S as a
zero is an ideal in F [x1 , · · · , xn ].
(49) How many maximal ideals of Z[x] contain {42, x2 + 1}?
(50) Is Q[x]/hx3 − 6x − 6i a field? Why?
(51) Find all c ∈ Z7 so that Z7 [x]/hx2 + x + ci is a field.
n
(52) For any n ∈ Z+ , show that x2 + 1 is irreducible over Q.
(53) Let be p a prime. RFor any monic f (x) ∈ Z[x] with deg(f (x)) =
x
p − 1, show that 0 f (t)dt + 1 is irreducible over Q.
(54) Is x4 + x3 + x2 + x + 1 is irreducible in Q[x]? in Z2 [x]?
(55) Factor the polynomial f (x) = x12 + x9 + x6 + x3 + 1 into
product of irreducibles over in Q[x]. (Hint: Consider (x3 −
1)f (x).)
(56) Factor the polynomial x7 +1 into product of irreducibles over
in Z2 [x].
(57) Is x4 + x + 1 is irreducible in Q[x]? in Z2 [x]?
(58) Show that f (x) = (1 + x + · · · + xn )2 − xn ∈ Z[x] is a product
of two polynomials in Z[x]. (Hints: Consider (x − 1)2 f (x).)
(59) Let n ≥ 2 be an integer. Show that the polynomial fn (x) =
xn−1 + xn−2 + · · · + x + 1 ∈ Q[x] is irreducible if and only if
n is a prime number.
(60) For any integer n > 2, show that all roots of f (x) = xn−1 +
2xn−2 + 3xn−3 + · · · + (n − 1)x + n have norm larger than 1.
(Hints: Consider (x − 1)f (x).)
March 18, 2022 9:45 amsart-9x6 12819-main page 46

46 RING AND FIELD THEORY

(61) For any prime p, show that


f (x) = xp−1 + 2xp−2 + 3xp−3 + · · · + (p − 1)x + p
is irreducible in Z[x]. (Hints: Use the previous question.)
(62) Let a be a nonzero integer, and n ≥ 3 be another integer.
Show that the polynomial f (x) = xn + axn−1 + axn−2 + · · · +
ax − 1 is irreducible over the integers. (Hints: Use Brauer’s
Theorem twice.)
(63) In Theorem 1.8.6, if we replace the condition “p > |a1 | + · · · +
|an |” with “p = |a1 | + · · · + |an | and no root of f (x) is a root
of unity”, show that f (x) is also irreducible over Q.
March 18, 2022 9:45 amsart-9x6 12819-main page 47

2. Unique Factorization Domains

In this chapter we will study an important class of integral domains:


unique factorization domains (UFD, for short). We further inves-
tigate some special UFDs: PID and Euclidean domains. We also
provide some applications for them. We remark that many materials
in this chapter and Chapter 4 have appeared in Chapters 5 and 6
in [LZ].

2.1. Basic definitions.


We already knew that the ring Z of integers and the polynomial ring
F [x] over a field F are integral domains in which every element can
be factored into unique product of irreducibles. Now we consider the
general case.
Recall that U(R) is the unit group of a unital ring R.
Definition 2.1.1. Let R be a unital commutative ring. Two elements
a, b ∈ R are associates in R if a = bε for some ε ∈ U(R), denoted
by a ∼ b.
It is easy to show that the relation a ∼ b is an equivalence relation
on R.
Example 2.1.1. We know that U(Z) = {±1}. So the only associates
of 6 in Z are ±6.
The following easy results will be repeatedly used later.
Lemma 2.1.2. For nonzero elements a and b of an integral domain
D, we have
(i). hai ⊆ hbi if and only if b|a, and
(ii). hai = hbi (or equivalently, a|b and b|a) if and only if a ∼ b.
Proof. (i). Note that hai ⊆ hbi = bD if and only if a ∈ hbi, if and
only if a = bd for some d ∈ D, if and only if b|a.
(ii). Using (a), we see that hai = hbi if and only if a = bc and
b = ad for some c, d ∈ D. But then a = adc and by canceling, we
obtain 1 = dc. Thus d and c are units so a ∼ b. 
47
March 18, 2022 9:45 amsart-9x6 12819-main page 48

48 RING AND FIELD THEORY

Definition 2.1.3. Let D be an integral domain and 0 6= a ∈ D. All


units and associates of a are called the trivial factors of a.
Let D be an integral domain. Recall from Definition 1.1.20 that
an element a ∈ D is irreducible if a = bc for some b, c ∈ D implies
that b or c is a unit.
Lemma 2.1.4. Let D be an integral domain. If p ∈ D is irreducible
and p ∼ q, then q is irreducible.
Proof. Since p ∼ q, then p = εq for an ε ∈ U(D). Any factoriza-
tion q = ab, where a, b ∈ D, gives p = εab. Clearly a and b cannot
be both units. We may assume that b is not. We deduce that εa is a
unit since p is irreducible, and further a is a unit. Thus q is also an
irreducible. 
Definition 2.1.5. Let D be an integral domain. We say that D is
a unique factorization domain (UFD for short) if the following
conditions are satisfied.
(i). Every element of D that is neither 0 nor a unit can be factored
into a product of a finite number of irreducibles.
(ii). If p1 · · · ps = q1 · · · qt where pi , qj are irreducibles in D, then
s = t and the qj can be renumbered so that pi ∼ qi .
Example 2.1.2. If F is a field, we know that the ring F [x] is a
UFD. Also we know that the integer Z is a UFD. Consider in Z the
factorizations
30 = (2)(3)(5) = (−2)(−3)(5) = (3)(−2)(−5).
Clearly 2 ∼ −2, 3 ∼ −3 and 5 ∼ −5. Thus except for order and
associates, the irreducible factors in these three factorizations of 30
are the same.
Notice that in a UFD D, any nonzero nonunit element a ∈ D can
be written as
a = εpr11 pr22 · · · prss ,
where ε is a unit, ri ≥ 1, p1 , p2 , · · · , ps are irreducibles and not
associates.
Definition 2.1.6. A nonzero nonunit element a of an integral do-
main D is called a prime if, for all b, c ∈ D, a|bc implies either a|b
or a|c.
March 18, 2022 9:45 amsart-9x6 12819-main page 49

UNIQUE FACTORIZATION DOMAINS 49

Note that, a ∈ D is prime if and only if hai is a prime ideal of D.


Lemma 2.1.7. Let D is an ID. If p ∈ D is prime, then p is an
irreducible.
Proof. Suppose that p = ab for some a, b ∈ D. Since p is prime,
then p|a or p|b. We may assume that p|a, i.e., a = pc for some c ∈ D.
We have p = pbc. So bc = 1, and b ∈ U(D). Thus p is an irreducible.

One can show that in a UFD an irreducible is also a prime (see
Exercise 2.6(2)). Thus the concepts of prime and irreducible coincide
in a UFD. The concepts do not coincide in every domain.
Example 2.1.3. Consider the subdomain D = R[x2 , xy, y 2 ] of R[x, y].
Then x2 , xy, and y 2 are irreducibles in D, but
(x2 )(y 2 ) = (xy)(xy).
Since xy divides x2 y 2 but not x2 or y 2 , we see that xy is not a prime
in D. Similarly, neither x2 nor y 2 is a prime.
Definition 2.1.8. Let D be a UFD and let a1 , a2 , · · · , an be nonzero
elements of D. An element d of D is called a greatest common
divisor (abbreviated gcd) of all of the ai if the following are satisfied.
(i). d|ai for i = 1, · · · , n,
(ii). d0 ∈ D and d0 |ai imply that d0 |d.
When we write d ∼ gcd(a1 , a2 , · · · , an ) we mean that d is one of
the gcds of a1 , a2 , · · · , an .
We can easily see that any two gcd’s are associates. The well-
known technique in the example below shows that gcd’s exist in a
UFD.
Example 2.1.4. We knew how to find gcd(420, −168, 252) in Z.
Factoring, we obtain 420 = 22 · 3 · 5 · 7, −168 = 23 · (−3) · 7, and
252 = 22 · 32 · 7. Then gcd(420, −168, 252) = 4 · 3 · 1 · 7 = 84. The
only other gcd of these numbers in Z is −84, because 1 and −1 are
the only units.
The technique in the above Example depends on being able to
factor an element of a UFD into a product of irreducibles. This
can be a tough job, even in Z. Later we will learn a technique, the
Euclidean Algorithm, that will allow us to find gcd’s in some UFDs.
March 18, 2022 9:45 amsart-9x6 12819-main page 50

50 RING AND FIELD THEORY

2.2. Principal ideal domains.


Let R be a commutative unital ring. Let a ∈ R. Recall that the
principal ideal hai consists of all multiples of the element a, i.e.,
hai = Ra. From Definition 1.5.10 we know that an integral domain
D is a principal ideal domain (abbreviated PID) if every ideal in D
is a principal ideal. We see that Z is a PID because every ideal is of
the form nZ, the ideal generated by some integer n. If F is a field,
then F [x] is a PID (Theorem 1.5.11). Our purpose in this section is
to prove that every PID is a UFD.
Let us start with an easy result.
Lemma 2.2.1. Let I1 ⊆ I2 ⊆ · · · be an ascending chain of ideals Ii
in a ring R. Then N = ∪i Ii  R.
Proof. Let a, b ∈ N, r ∈ R. Then there are ideals Ii and Ij in the
chain, with a ∈ Ii and b ∈ Ij . Now either Ii ⊆ Ij or Ij ⊆ Ii . We may
assume that Ii ⊆ Ij , so both a and b are in Ij . This implies that
a ± b, ar, ra ∈ Ij , so a ± b, ar, ra ∈ N . Hence N  R. 
Lemma 2.2.2 (Ascending Chain Condition). Let D be a PID, and
I1 ⊆ I2 ⊆ · · · be an ascending chain of ideals Ii of D. Then there
exists a positive integer m such that In = Im for all n ≥ m.
Proof. By the above Lemma, we know that N = ∪i Ii  D. Since
D is a PID, there is a c ∈ D such that N = hci. Since N = ∪i Ii , we
must have c ∈ Im , for some m ∈ N. For n ≥ m, we have
hci ⊆ Im ⊆ In ⊆ N = hci.
Thus Im = In for n ≥ m. 
The above lemma tells us that every strictly ascending chain of
ideals (that is, all inclusions proper) in a PID is of finite length. In
this situation we say that the ascending chain condition (ACC)
holds for ideals in the PID.
Theorem 2.2.3. Let D be a PID, and a ∈ D neither 0 nor a unit.
Then a can be factored into a product of irreducibles.
Proof. Claim 1. a has at least one irreducible factor.
If a is an irreducible, we are done. If a is not an irreducible, then
a = a1 b1 , for some nonunit elements a1 , b1 ∈ D. Now hai ⊂ ha1 i.
In this manner, then starting now with a1 , we arrive at a strictly
ascending chain of ideals
hai ⊂ ha1 i ⊂ ha2 i ⊂ · · · .
March 18, 2022 9:45 amsart-9x6 12819-main page 51

UNIQUE FACTORIZATION DOMAINS 51

By Lemma 2.2.2, this chain terminates with some har i, and ar must
then be irreducible. Thus a has an irreducible factor ar . Claim 1
follows.
From Claim 1, either a is irreducible or a = p1 c1 for p1 an ir-
reducible and c1 not a unit. We see that hai ⊂ hc1 i. If c1 is not
irreducible, then c1 = p2 c2 for an irreducible p2 with c2 not a unit.
Continuing in this manner, we get a strictly ascending chain of ideals
hai ⊂ hc1 i ⊂ hc2 i ⊂ · · · .
By Lemma 2.2.2 this chain must terminate, i.e., with some cr = qr
that is an irreducible. Thus a = p1 p2 · · · pr qr . 
We first give the following result which is of some interest in itself.
Lemma 2.2.4. Let D is a PID and p ∈ D \{0}. Then hpi is maximal
in D if and only if p is an irreducible.
Proof. (⇒). Let hpi be a maximal ideal of D. Suppose that
p = ab where a, b ∈ D. It follows that hpi ⊂ hai. Suppose that
hai = hpi. Then a ∼ p, so b must be a unit. If hai = 6 hpi, we see
that hpi ⊂ hai, i.e., hai = h1i = D, since hpi is maximal. Then a is a
unit. Thus, p is an irreducible of D.
(⇐). Suppose that p is an irreducible in D. If hpi ⊆ hai  D for
some a ∈ D, we have p = ab for some b ∈ D. Since p is an irreducible,
we see that a or b is a unit.
Now if a is a unit, then hai = h1i = D.
If b is a unit, then a = b−1 p, so hai ⊆ hpi, and we have hai = hpi.
Thus hpi ⊆ hai implies that either hai = D or hai = hpi. Hence hpi
is a maximal ideal. 
Lemma 2.2.5. Let D be a PID and p ∈ D. If p is an irreducible,
then p is prime.
Proof. Suppose that p|ab where a, b ∈ D. Then ab ∈ hpi. Since p
is an irreducible, we know that hpi is a maximal ideal. From Theo-
rems 1.7.2 and 1.5.4, then hpi is a prime ideal. So either a ∈ hpi or
b ∈ hpi, yielding that either p|a or p|b. Thus, p is prime. 
Corollary 2.2.6. If p is an irreducible in a PID D and p|a1 a2 · · · an
for ai ∈ D, then p|ai for some i.
Proof. This is immediate from using mathematical induction. 
Now we can prove the main result in this section.
March 18, 2022 9:45 amsart-9x6 12819-main page 52

52 RING AND FIELD THEORY

Theorem 2.2.7. Every PID is a UFD.


Proof. Let D be a PID and a ∈ D, where a is neither 0 nor a
unit. Then a has a factorization a = p1 p2 · · · pr into irreducibles. We
need only to show uniqueness. Let
p1 p2 · · · pr = q1 q2 · · · qs
where pi , qj are irreducibles. Then we have p1 |q1 q2 · · · qs , which im-
plies that p1 |qj for some j. By changing the order of the qj if neces-
sary, we can assume that j = 1 so p1 |q1 . Then q1 = p1 u1 , and since
q1 , p1 are irreducibles, u1 is a unit, so p1 ∼ q1 . We have
p1 p2 · · · pr = p1 u1 q2 · · · qs .
Then
p2 · · · pr = u1 q2 · · · qs .
In this manner, we will arrive at pi ∼ qi for i ≤ r and
1 = u1 u2 · · · ur qr+1 · · · qs .
Since each qj is irreducible, we must have r = s. 
The converse to the above theorem is false. That is, a UFD need
not be a PID (see Example 2.4.3).
The following well-known result from elementary number theory
can easily follow from the fact that Z is a PID.
Corollary 2.2.8 (Fundamental Theorem of Arithmetic). The ring
Z of integers is a UFD.
2.3. Euclidean domains.
In this section we shall study an important class of UFDs, the Eu-
clidean domains.
Definition 2.3.1. Let D be an ID. A Euclidean norm on D is a
function ν : D∗ → Z+ such that the following conditions are satisfied
for all a, b ∈ D∗ = D \ {0}.
(i). There exist q, r ∈ D such that a = bq + r, where either r = 0
or ν(r) < ν(b).
(ii). ν(a) ≤ ν(ab).
An integral domain D is a Euclidean domain if it has a Eu-
clidean norm.
March 18, 2022 9:45 amsart-9x6 12819-main page 53

UNIQUE FACTORIZATION DOMAINS 53

Example 2.3.1. The integer domain Z is a Euclidean domain, since


it has a Euclidean norm
ν : Z → Z+ , ν(n) = |n|, ∀n ∈ Z.
Condition 1 holds by the division algorithm for Z. Condition (ii)
follows from |ab| = |a| · |b| and |b| ≥ 1 for b 6= 0 in Z.
Example 2.3.2. For any field F , the polynomial ring F [x] is a Eu-
clidean domain, since from Division Algorithm F [x] has a Euclidean
norm
ν : F [x]∗ → Z+ , ν(f (x)) = deg(f (x)), ∀f (x) ∈ F [x]∗ .
Of course, we shall have some examples of Euclidean domains other
than these familiar ones later.
Theorem 2.3.2. Every Euclidean domain D is a PID.
Proof. Let D have a Euclidean norm ν, and let I  D. If I = 0,
then I = h0i. Now suppose that I 6= 0. Take b ∈ I \ {0} such that
ν(b) is minimal among all ν(g) for g ∈ I \{0}. We claim that I = hbi.
Let a ∈ I. Then by Condition (i) for a Euclidean domain, there exist
q, r ∈ D such that a = bq +r, where either r = 0 or ν(r) < ν(b). Now
r = a − bq and a, b ∈ I, so that r ∈ I since I  D. So ν(r) < ν(b)
is impossible by the choice of b. Thus r = 0, so a = bq, and a ∈ hbi.
Hence, I ⊆ hbi ⊆ I, and further I = hbi. 
Corollary 2.3.3. A Euclidean domain is a UFD.
Finally, we should mention that examples of PIDs that are not
Euclidean are not easily found, however.
Let D be a Euclidean domain with a Euclidean norm ν. We can
use Condition (ii) of a Euclidean norm to characterize the units of D.
Theorem 2.3.4. Let D be a Euclidean domain with a Euclidean
norm ν.
(i). ν(1) ≤ ν(a) for any nonzero a ∈ D.
(ii). u ∈ D is a unit if and only if ν(u) = ν(1).
Proof. (i). For a ∈ D \ {0}, we have ν(1) ≤ ν(1a) = ν(a).
(ii). (⇒). If u ∈ U(D), then
ν(u) ≤ ν(uu−1 ) = ν(1).
Thus ν(u) = ν(1).
March 18, 2022 9:45 amsart-9x6 12819-main page 54

54 RING AND FIELD THEORY

(⇐). Suppose u ∈ D with ν(u) = ν(1). Then by the division


algorithm, there exist q, r ∈ D such that
1 = uq + r,
where either r = 0 or ν(r) < ν(u). By (i), we know that ν(r) < ν(u)
is impossible. Then r = 0 and 1 = uq. Hence u ∈ U(D). 
Example 2.3.3. (a). For Z with ν(n) = |n|, the minimum of
ν(n) for nonzero n ∈ Z is 1. Thus, ±1 are the only ele-
ments of Z with ν(n) = 1. So ±1 are the only units of Z by
Theorem 2.3.4.
(b). The polynomial ring F [x] with ν(f (x)) = deg(f (x)) for f (x) 6=
0 is a Euclidean domain. The minimum value of ν(f (x))
for all nonzero f (x) ∈ F [x] is 0. The nonzero polynomi-
als of degree 0 are exactly the nonzero elements of F , So
F ∗ = U(F [x]) by Theorem 2.3.4.
We know that in any UFD D, gcd(a, b) exists for any a, b ∈ D.
But it is generally very hard to find gcd(a, b). The best property for
a Euclidean domain is that there is a nice algorithm for this, as the
next theorem shows.
Theorem 2.3.5 (Euclidean Algorithm). Let D be a Euclidean do-
main with a Euclidean norm ν, and let a and b be nonzero elements
of D.
(i). There are qi , ri ∈ D such that
a = bq1 + r1 ,
b = r1 q2 + r2 ,
r1 = r2 q3 + r3 ,
(2.1)
···
rs−3 = rs−2 qs−1 + rs−1 ,
rs−2 = rs−1 qs + rs ,
where rs = 0, ν(rs−1 ) < ν(rs−2 ) < · · · < ν(r2 ) < ν(r1 ) <
ν(b). Furthermore gcd(a, b) ∼ rs−1 .
(ii). If gcd(a, b) ∼ d, then there exist λ, µ ∈ D such that d =
λa + µb.
Proof. (i). Since ν(ri ) < ν(ri−1 ) and ν(ri ) is a nonnegative
integer, it follows that after some finite number of steps we must
arrive at some rs = 0. Thus, we have all equations in (2.1).
March 18, 2022 9:45 amsart-9x6 12819-main page 55

UNIQUE FACTORIZATION DOMAINS 55

Suppose d ∼ gcd(a, b). From d|a and d|b, we have d|r1 . From d|b
and d|r1 , we have d|r2 . In this manner we deduce that d|ri for any i.
In particularly d|rs−1 .
On the other hand rs−1 |rs−2 . From (2.1) backward we deduce that
rs−1 |rs−2 , rs−1 |rs−3 , · · · , rs−1 |b
and rs−1 |a. Thus rs−1 |d. Therefore rs−1 ∼ d.
(ii). We may assume that d = rs−1 . We shall prove by induction
on k that rk = λk a + µk b for some λk , µk ∈ D. If s = 1, i.e.,
r1 = 0, then d = b, and d = 0a + 1b and we are done. Suppose that
rj = λj a + µj b for j = 1, 2, · · · , k. Use rk−1 = rk qk+1 + rk+1 we
deduce that
rk+1 = rk−1 − rk qk+1 = (λk−1 a + µk−1 b) − qk+1 (λk a + µk b)
= λk+1 a + µk+1 b.
Thus
d = rs−1 = λs−1 a + µs−1 b
where λs−1 , µs−1 ∈ D. 
Example 2.3.4. Use Euclidean Algorithm in Q[x] to find a gcd(f (x),
g(x)), where
f (x) = x4 + x3 − x2 − 1, g(x) = x3 + x2 − 2x.
Solution. Notice that
x4 + x3 − x2 − 1 = x(x3 + x2 − 2x) + (x − 1),
x3 + x2 − 2x = (x2 + 2x)(x − 1) + 0.
So gcd(f (x), g(x)) = x − 1. 
2.4. Polynomial rings over UFDs.
In this section we shall prove that polynomial rings over UFDs are
UFDs. We always assume that D is a UFD.
Definition 2.4.1. Let D be a UFD, and let
f (x) = a0 + a1 x + · · · + an xn ∈ D[x] \ D.
An element c ∈ D is a content of f (x) if c ∼ gcd(a0 , a1 , · · · , an ).
We say that f (x) is primitive if gcd(a0 , a1 , · · · , an ) ∼ 1.
Example 2.4.1. In Z[x], x2 + 3x + 2 is primitive, but 4x2 + 2x + 8 is
not, since 2, a nonunit in Z, is a common divisor of the coefficients
4, 2, and 8.
March 18, 2022 9:45 amsart-9x6 12819-main page 56

56 RING AND FIELD THEORY

Observe that every irreducible in D[x] of positive degree must be


a primitive polynomial.
Lemma 2.4.2. Let D be a UFD. Then for every nonconstant f (x) ∈
D[x] we have f (x) = cg(x), where c ∈ D, g(x) ∈ D[x], and g(x) is
primitive. Also g(x) is unique up to a unit factor in D.
Proof. Let f (x) = a0 +a1 x+· · ·+an xn ∈ D[x], where a0 , a1 , . . . , an
with an 6= 0 and n ≥ 1. Let c ∼ gcd(a0 , a1 , . . . , an ). Write ai = cqi
for some qi ∈ D. We have f (x) = cg(x), where no irreducible in
D divides all of the coefficients q0 , q1 , . . . , qn of g(x). So g(x) is a
primitive polynomial.
For uniqueness, if f (x) = cg(x) = dh(x) for c, d ∈ D, h(x), g(x) ∈
D[x], and g(x), h(x) primitive. Since both c and d are contents of
f (x), then d = cu for a unit u ∈ D. We see that g(x) = uh(x) for a
unit u ∈ D. From f (x) = cg(x), we see that the primitive polynomial
g(x) is also unique up to a unit factor. 
Example 2.4.2. In Z[x], 4x2 + 6x − 2 = 2(2x2 + 3x − 1), where
2x2 + 3x − 1 is primitive.
Lemma 2.4.3 (Gauss’s Lemma). Let D be a UFD, and f (x), g(x) be
two primitive polynomials in D[x]. Then f (x)g(x) is also primitive.
Proof. Let
f (x) = a0 + a1 x + · · · + an xn ,
g(x) = b0 + b1 x + · · · + bm xm ,
and let h(x) = f (x)g(x). Let p be an arbitrary irreducible in D.
Then p does not divide all ai , and p does not divide all bj , since f (x)
and g(x) are primitive. Let ar be the first coefficient of f (x) not
divisible by p; i.e., p|ai for i < r, and p6 |ar . Similarly, let p|bj for
j < s, and p6 |bs . The coefficient of xr+s in h(x) = f (x)g(x) is
cr+s = (a0 br+s + · · · + ar−1 bs+1 ) + ar bs + (ar+1 bs−1 + · · · + ar+s b0 ).
Since p|ai for i < r, then
p|(a0 br+s + · · · + ar−1 bs+1 ).
Since p|bj for j < s, then
p|(ar+1 bs−1 + · · · + ar+s b0 ).
Since p6 |ar or p6 |bs , so p6 |ar bs , and consequently p6 |cr+s . This shows
that any irreducible p ∈ D does not divide some coefficient of f (x)g(x).
Therefore f (x)g(x) is primitive. 
March 18, 2022 9:45 amsart-9x6 12819-main page 57

UNIQUE FACTORIZATION DOMAINS 57

Corollary 2.4.4. Let D be a UFD. Then a finite product of primitive


polynomials in D[x] is also primitive.
Proof. This follows from the above Lemma by induction. 
Now let D be a UFD and let F be a field of quotients of D. Then
we have known that F [x] is a UFD.
Lemma 2.4.5. Let D be a UFD and let F be a field of quotients of
D. Let f (x) ∈ D[x] with deg(f (x)) > 0.
(i). If f (x) is an irreducible in D[x], then f (x) is also an irre-
ducible in F [x].
(ii). If f (x) is primitive in D[x] and irreducible in F [x], then f (x)
is irreducible in D[x].
Proof. (i). Suppose that f (x) = r(x)s(x) for r(x), s(x) ∈ F [x]
with deg(r(x)) < deg(f (x)) and deg(s(x)) < deg(f (x)). Since F is
a field of quotients of D, each coefficient in r(x) and s(x) is of the
form a/b for some a, b ∈ D. By clearing denominators, we can get
df (x) = r1 (x)s1 (x)
for d ∈ D, and r1 (x), s1 (x) ∈ D[x], where deg(r1 (x)) = deg(r(x)) and
deg(s1 (x)) = deg(s(x)). Write f (x) = ag(x), r1 (x) = a1 r2 (x), and
s1 (x) = a2 s2 (x) for primitive polynomials g(x), r2 (x), and s2 (x),
and a, a1 , a2 ∈ D. Then
(da)g(x) = a1 a2 r2 (x)s2 (x),
and r2 (x)s2 (x) is primitive. By the uniqueness, a1 a2 = dau for some
unit u in D. So
(da)g(x) = daur2 (x)s2 (x),
yielding that
f (x) = ag(x) = aur2 (x)s2 (x).
This is impossible. Thus f (x) ∈ D[x] is irreducible in F [x].
(ii). A nonconstant f (x) ∈ D[x] that is primitive in D[x] and
irreducible in F [x] is also irreducible in D[x], since D[x] ⊂ F [x]. 
The above Lemma shows that if D is a UFD, the irreducibles in
D[x] are precisely the irreducibles in D, together with the noncon-
stant primitive polynomials that are irreducible in F [x], where F is
a field of quotients of D.
Corollary 2.4.6. Let D be a UFD and let F be a field of quotients
of D. Let f (x) ∈ D[x] with deg(f (x)) > 0. Then f (x) factors into a
March 18, 2022 9:45 amsart-9x6 12819-main page 58

58 RING AND FIELD THEORY

product of two polynomials of degrees r and s in F [x] if and only if


f (x) has a factorization into polynomials of the same degrees r and
s in D[x].
Proof. (⇒). This was shown in the proof for (i) of the previous
lemma.
(⇐). This holds trivially since D[x] ⊆ F [x]. 
Now we prove our main theorem in this section.
Theorem 2.4.7. Let D be a UFD. Then D[x] is a UFD.
Proof. Let f (x) ∈ D[x], where f (x) is neither 0 nor a unit.
If f (x) is of degree 0, we are done, since D is a UFD. Suppose that
degf (x) > 0. Let
f (x) = g1 (x)g2 (x) · · · gr (x)
be a factorization of f (x) in D[x] having the greatest number r of
factors of positive degree. Now write each gi (x) = ai hi (x) where ai
is a content of gi (x) and hi (x) is a primitive polynomial. From the
maximality of r, each of the hi (x) is irreducible. Thus we now have
f (x) = a1 a2 · · · ar h1 (x)h2 (x) · · · hr (x)
where the hi (x)’s are irreducibles in D[x]. If we now factor the
a1 a2 · · · ar into irreducibles in D, we obtain a factorization of f (x)
into a product of irreducibles in D[x].
Now we prove the uniqueness. Let
a1 a2 · · · ar g1 (x)g2 (x) · · · gs (x) = b1 b2 · · · br0 h1 (x)h2 (x) · · · hs0 (x)
(2.2)
where the ai , bi , gj (x), hj (x) are irreducibles in D[x]. Then a1 a2 · · · ar
∼ b1 b2 · · · br0 since they are content of the above polynomial, and also
g1 (x)g2 (x) · · · gs (x) ∼ h1 (x)h2 (x) · · · hs (x) in F [x].
Then after renumbering bi ’s and using Theorem 2.4.5 and the fact
that F [x] is a UFD, we have
r = r0 , ai ∼ bi in D,
s = s0 , gj (x) ∼ hj (x) in F [x].
Note that gj (x), hj (x) are irreducibles in F [x]. There are cj , dj ∈ D∗
c
such that gj (x) = djj hj (x). Then dj gj (x) = cj hj (x) and further
cj ∼ dj in D, hence gj (x) ∼ hj (x) in D[x]. The uniqueness follows.

March 18, 2022 9:45 amsart-9x6 12819-main page 59

UNIQUE FACTORIZATION DOMAINS 59

Corollary 2.4.8. Let F be a field and x1 , · · · , xn indeterminates.


Then F [x1 , · · · , xn ] is a UFD.
Proof. By the above theorem, F [x1 ] is a UFD. Again by the above
theorem, so is (F [x1 ])[x2 ] = F [x1 , x2 ]. Continuing in this procedure,
by induction we see that F [x1 , · · · , xn ] is a UFD. 
We have seen that a PID is a UFD. It is easy for us to give an
example that shows that not every UFD is a PID.
Example 2.4.3. Consider the polynomial ring F [x, y] over a field F .
We know that F [x, y] is a UFD. Let I = xF [x, y] + yF [x, y]. Then
I  F [x, y]. If I = aF [x, y] for some a ∈ F [x, y]. Since x, y ∈ I, we
have a|x, a|y. Thus a ∈ F which is impossible. Hence F [x, y] is not
a PID.
Now we have following generalization of Theorem 1.7.9.
Theorem 2.4.9 (Schönemann-Eisenstein Criterion). Let D be a UFD
with quotient field F , and f (x) = an xn + · · · + a1 x + a0 ∈ D[x] with
n ≥ 1 and an 6= 0. If p is prime in D such that
(i). p|ai for 0 ≤ i < n,
(ii). p6 |an ,
(iii). p2 6 |a0 ,
then f (x) is irreducible over F .
Proof. The proof is identical to that of Theorem 1.7.9.
For a contradiction, suppose that f (x) is reducible in F [x], and
f (x) = g(x)h(x) for some positive degree polynomials g(x), h(x) ∈
F [x]. By Corollary 2.4.6 we may assume that g(x), h(x) ∈ D[x] as
well. Then denote by f (x), g(x), h(x) the reductions mod p of these
polynomials, i.e., consider them as polynomials with coefficients in
the ID D/hpi (since hpi is prime). We have gh(x) = f (x) = an xn ,
which means that g(x) = axk and h(x) = bxn−k . This shows that
all the other coefficients of g and h are zero mod p, so they’re all
divisible by p. Since their constant terms are both divisible by p,
the constant term of f (x) is divisible by p2 , which contradicts the
hypothesis of the theorem. 
Example 2.4.4. Show that f (x, y) = x3 + y 3 + 1 is irreducible
in Q[x, y]. You need to take D = Q[y], p(y) = y + 1 and use
Schönemann-Eisenstein Criterion.
March 18, 2022 9:45 amsart-9x6 12819-main page 60

60 RING AND FIELD THEORY

2.5. Multiplicative norms.


In this section we shall give some examples of Euclidean domains
different from the ring Z of integers and the polynomial ring F [x].
Definition 2.5.1. Let Z[i] = {a + bi : a, b ∈ Z} which is a subring of
C. Any number in Z[i] is called a Gaussian integer. The norm of
a+bi ∈ Z[i], where a, b ∈ Z, is define as N (a+bi) = |a+bi|2 = a2 +b2 .
We can easily extend the function N to C, i.e., define N (a +
bi) = a2 + b2 for any a + bi ∈ C where a, b ∈ R. Note that the
Gaussian integers include all the integers. Recall that the norm or
p of a + bi ∈ C, where a, b ∈ R, was defined as |a + bi| =
absolute value

a2 + b2 = N (a + bi). So here we have different meaning for the
word norm.
Lemma 2.5.2. For all α, β ∈ C we have
(i). N (α) ≥ 0,
(ii). N (α) = 0 if and only if α = 0,
(iii). N (αβ) = N (α)N (β).
Proof. These results directly follow from properties of absolute
value of complex numbers. 
Lemma 2.5.3. Z[i] is an integral domain.
Proof. This follows from the fact that Z[i] ⊂ C which is a field.

Theorem 2.5.4. The norm N (α) for nonzero α ∈ Z[i] is a Euclidean
norm on Z[i], i.e., Z[i] is a Euclidean domain.
Proof. For β = b1 + b2 i 6= 0 we know that N (b1 + b2 i) = b21 + b22 .
So N (β) ≥ 1. Then
N (α) ≤ N (α)N (β) = N (αβ), ∀α, β ∈ Z[i] \ {0}.
This proves Condition (ii) in Definition 2.3.1 for a Euclidean norm.
Now we prove Condition (i) in Definition 2.3.1 for N . Let α = a1 +
a2 i, β = b1 + b2 i ∈ Z[i], where β 6= 0. We want to find σ and ρ in Z[i]
such that α = βσ + ρ, where either ρ = 0 or N (ρ) < N (β) = b21 + b22 .
Let αβ = r +si for r, s ∈ Q. Take q1 , q2 ∈ Z such that |r −q1 | ≤ 1/2
and |s − q2 | ≤ 1/2. Let σ = q1 + q2 i and ρ = α − βσ. If ρ = 0, we
are done. Otherwise, we see that
Å ã
α
N − σ = N ((r + si) − (q1 + q2 i))
β
March 18, 2022 9:45 amsart-9x6 12819-main page 61

UNIQUE FACTORIZATION DOMAINS 61

= N ((r − q1 ) + (s − q2 )i) ≤ (1/2)2 + (1/2)2 = 1/2.


Thus we obtain
Å Å ãã Å ã
α α
N (ρ) = N (α − βσ) = N β −σ = N (β)N −σ
β β
≤ N (β)/2 < N (β).

Example 2.5.1. In Z[i], find U(Z[i]) and factor 5 into a product of
irreducibles.
Solution. In Z[i], since N (1) = 1, the units of Z[i] are exactly
the α = a1 + a2 i with N (α) = a21 + a22 = 1. Since a1 , a2 ∈ Z, it
follows that a1 = ±1 with a2 = 0, or a1 = 0 with a2 = ±1. Thus
U(Z[i]) = {±1, ±i}.
We know that 5 is an irreducible in Z. But 5 is no longer an
irreducible in Z[i] since 5 = (1 + 2i)(1 − 2i), where neither 1 + 2i nor
1 − 2i is a unit. 
Example 2.5.2. Use a Euclidean algorithm in Z[i] to find a gcd(8 +
6i, 5 − 15i).
Solution. Since 5−15i 1 3
8+6i = − 2 − 2 i, we have 5 − 15i = −i(8 + 6i) −
8+6i
(1 + 7i). Since 1+7i = 1 − i, we have 8 + 6i = (1 + 7i)(1 − i). We put
them together
5 − 15i = −i(8 + 6i) − (1 + 7i),
8 + 6i = (1 + 7i)(1 − i) + 0.
Thus gcd(8 + 6i, 5 − 15i) ∼ 1 + 7i. 
Let us study integral domains that have a multiplicative norm.
Definition 2.5.5. A multiplicative norm N on an integral do-
main D is a function mapping N : D → Z with the following condi-
tions hold for all α, β ∈ D.
(i). N (α) = 0 if and only if α = 0.
(ii). N (αβ) = N (α)N (β).
Note that the Euclidean norm on C[x] is no longer a multiplicative
norm.
Theorem 2.5.6. Let D be an ID with a multiplicative norm N.
(i). |N (u)| = 1 for every unit u ∈ U(D).
March 18, 2022 9:45 amsart-9x6 12819-main page 62

62 RING AND FIELD THEORY

(ii). If U(D) = {α ∈ D : |N (α)| = 1} and β ∈ D is such that


|N (β)| = p for a prime p ∈ Z, then β is an irreducible of D.
Proof. (i). From
N (1) = N ((1)(1)) = N (1)N (1)
we see that N (1) = 1. If u is a unit in D, then
1 = N (1) = N (uu−1 ) = N (u)N (u−1 ).
Since N (u) is an integer, we deduce t that |N (u)| = 1.
(ii). If β = αγ where α, γ ∈ D, we have
p = |N (β)| = |N (α)N (γ)| = |N (α)| · |N (γ)|.
Then either |N (α)| = 1 or |N (γ)| = 1. By (i) we know that either α
or γ is a unit of D. So β is an irreducible of D. 
Example 2.5.3. It is easy to see that the function N defined by
N (a + bi) = a2 + b2 gives a multiplicative norm on Z[i]. We know
that
U(Z[i]) = {α ∈ Z[i] : |N (α)| = 1} = {±1, ±i}.
We see that 13 is not an irreducible in Z[i] since 13 = (3+2i)(3−2i).
Since N (3 + 2i) = N (3 − 2i) = 32 + 22 = 13 and 13 is a prime in
Z, we see from the above theorem that 3 + 2i and 3 − 2i are both
irreducibles in Z[i].
The next example gives another example of an integral domain
that is not a UFD.
Example 2.5.4. We now consider the ID
√ √
Z[ −5] = {a + b −5 i : a, b ∈ Z} ⊂ C.

Define a multiplicative norm N on Z[ −5] by

N (a + b −5) = a2 + 5b2 , ∀a, b ∈ Z.

Clearly, N (a + b −5) = 0 if and √ only if a2 + 5b2 = 0 if and only
if a = b = 0 if and only if a + b −5 √ = 0. It is easy to see that
N (αβ) = N (α)N√ (β) for any α, β ∈ Z[ −5]. So N is a multiplicative
norm N on Z[ √ −5].
Now N (a + b −5) = 1 if and only √ if a2 + 5b2 = 1 if and
√ only if
a = ±1 and b = 0 if and only if a + b −5 = ±1. So U(Z[ −5]) =
{±1}.
March 18, 2022 9:45 amsart-9x6 12819-main page 63

UNIQUE FACTORIZATION DOMAINS 63



In Z[ −5], we have
√ √
9 = 3 · 3, 9 = (2 + −5)(2 − −5).
√ √
We√will show that√3, 2 + −5, and 2 − −5 are all irreducibles in
Z[ −5]. Then Z[ −5] is not a UFD.
If 3 = αβ, then
9 = N (3) = N (α)N (β).
We see that N (α) = 1, 3, or 9. If N (α) = 1, then α is a unit. If
N (α) = 3 then a2 + 5b2 = 3 which is impossible. If N (α) = √ 9, then
N (β) = 1, so β is a unit. Thus √ 3 is an irreducible
√ in Z[ −5]. A
similar argument
√ shows that 2
√ + −5 and 2 − −5 are also irre-
ducibles in Z[ −5]. Hence Z[ −5] is an integral domain but not a
UFD.
In conclusion, we know that the numbers
√ √
±3, 2 + −5, 2 − −5

are all irreducibles in Z[ −5], but none of them is prime. 
2.6. Exercises.
(1) Show that the ring (Z[x], +, ·) is not a PID.
(2) If p is an irreducible in a UFD D, show that p is a prime.
(3) Factor the polynomial 4x2 − 4x + 8 into a product of irre-
ducibles viewing it as an element of the UFD Z[x]; or Q[x];
or Z11 [x].
(4) Find a gcd of the following polynomials in Q[x] :
x10 − 3x9 + 3x8 − 11x7 + 11x6 − 11x5
+ 19x4 − 13x3 + 8x2 − 9x + 3,
x6 − 3x5 + 3x4 − 9x3 + 5x2 − 5x + 2.
(5) Let α, β ∈ Z[i]. Show that gcd(α, β) ∼ 1 if and only if
gcd(N (α), N (β)) = 1, where N (α) = |α|2 .
(6) Let D be a UFD, F be the field of quotients of D, f (x1 , x2 , · · · ,
xn ) ∈ D[x1 , x2 , · · · , xn ] be primitive in the obvious meaning.
Show that the polynomial f (x1 , x2 , · · · , xn ) is irreducible in
D[x1 , x2 , · · · , xn ] if and only if it is irreducible in F [x1 , x2 , · · · ,
xn ].
(7) Show that f (x) = x4 − 4x2 + 1 is irreducible in Q[x] but it is
reducible in Zp [x] for any prime p.
March 18, 2022 9:45 amsart-9x6 12819-main page 64

64 RING AND FIELD THEORY

(8) Use the Euclidean algorithm and UFD’s property in Z[i] to


find gcd(15 − 12i, 6 − 5i), and gcd(16 + 7i, 10 − 5i).
(9) Find all prime numbers p such that p = a4 + 4b4 , where
a, b ∈ Z. (Hint: Consider the numbers in Z[i].)
(10) Prove that Z[i]/h2 + ii is a field.
(11) Let R be a PID, a, b ∈ R. Prove that gcd(a, b) ∼ 1 if and
only if there exist u, v ∈ R such that au + bv = 1.
(12) Show that in a PID, any proper ideal is contained in a max-
imal ideal.
(13) Show that the integer solutions of x2 + 2 = y 3 are x = ±5,
y = 3. √
Hints: We know that D = Z[ −2] is a Euclidean domain
under the usual complex norm. Let x, y ∈ Z such that x2 +
2 = y3. √
(a). Show that −2 √ √ in D.
is irreducible
(b). Show that (x + −2), (x − −2) are relatively prime.
(c). Deduce that the integer solutions of x2 + 2 = y 3 are
x = ±5, y = 3.
(14) Let D be a PID.
(a). Show that every nonzero prime ideal of D is maximal.
(b). If S is an integral domain and φ : D → S is a surjective
ring homomorphism, show that either φ is an isomor-
phism or S is a field.
(c). If D[x]√is a PID, show that D is a field.
(15) Let R = Z[ −a], where a is an integer ≥ 3. Show that 2R is
not a prime ideal in R, but that 2 is an irreducible element
of R. Is R a PID? a UFD? Why or why not?
(16) Let D be a PID, and let I, J be nonzero principal ideals of
D. Show that IJ = I ∩ J if and only if I + J = D.
(17) Let F be a field, and R = {f (x) ∈ F [x] : f 0 (0) = 0}.
(a). Show that R is a subring of F [x].
(b). Show that x2 and x3 are irreducibles in R.
(c). Show that R is not a UFD.
(d). Explicitly demonstrate an ideal that is not principal.
(18) Show that the following polynomials are irreducible in the
integral domain C[x, y]:

y 5 + xy 4 − y 4 + x2 y 2 − 2xy 2 + y 2 + x3 − 1,
xy 3 + x2 y 2 − x5 y + x2 + 1.
March 18, 2022 9:45 amsart-9x6 12819-main page 65

UNIQUE FACTORIZATION DOMAINS 65

(19) For any integer n > 2, show that the following polynomials
f (x) = xn + 2xn−1 + 2xn−2 + · · · + 2x + 1 + i,
g(x) = xn + 5xn−1 + 5xn−2 + · · · + 5x + 2 + i
are irreducible over Z[i].
(20) Show that x6 + x3 + 1√is irreducible √ over Q.
(21) Show that the rings Z[ 2] and Z[ 3] are Euclidean√domains.
(22) Show that 6 does not factor uniquely in the ID Z[ −5].
(23) Let F be a field. Find all f (x) ∈ F [x] such that f (x2 ) =
f (x)2 .
(24) Let n ∈ N and a1 , a2 , · · · , an ∈ Z be pairwise distinct. Show
that
f (x) = (x − a1 ) (x − a2 ) · · · (x − an ) − 1
is irreducible over Q. (This was proved by Issai Schur (1875–
1941) in 1908. Hints: If f (x) = g(x)h(x), then g(ai )h(ai ) =
−1, and furthermore g(ai ) = −h(ai ).)
(25) Let n ∈ N and a1 , a2 , · · · , an ∈ Z be pairwise distinct. Show
that
f (x) = [(x − a1 ) (x − a2 ) · · · (x − an )]2 + 1
is irreducible over Q.
(26) Let f (x) ∈ Z[x] such that it takes value 1 at four distinct
integers. Show that f (x) does not take value −1 at any inte-
ger.
(27) Let g(x) = ax2 +bx+1 ∈ Z[x] be irreducible over Q of degree
2. Let n ∈ N and a1 , a2 , · · · , an ∈ Z be pairwise distinct and
let
f (x) = (x − a1 ) (x − a2 ) · · · (x − an ) .
Show that g(f (x)) is irreducible over Q if n ≥ 7.
(28) Show that, the result in Perron’s Irreducibility Criterion 1.8.1
still holds if we replace Z and Q with Z[i] and Q[i], respec-
tively.
(29) Let p be a prime in Z[i],
f (x) = an xn + an−1 xn−1 + · · · + a1 x + p ∈ Z[i][x],
with n ≥ 1 and an 6= 0. If |p| > |a1 | + · · · + |an |, show that
f (x) is irreducible over Q[i].
B1948 Governing Asia

This page intentionally left blank

B1948_1-Aoki.indd 6 9/22/2014 4:24:57 PM


March 18, 2022 9:45 amsart-9x6 12819-main page 67

3. Modules and Noetherian rings

In this chapter we will introduce Noetherian rings and modules over


a ring, which are powerful tools to study rings. We will establish very
basic properties for Noetherian rings and modules. In this chapter
we always assume that R is a ring. Unlike many other books, we do
not assume that R is unital at the beginning.

3.1. Modules, submodules and isomorphism theorems.


Definition 3.1.1. A set M is called a left R-module or a module
over R if M is an additive abelian group with a map R × M → M
defined by (r, u) → ru such that for u, v ∈ M and r1 , r2 ∈ R we have
(1). r1 (u + v) = r1 u + r1 v,
(2). (r1 + r2 )u = r1 u + r2 u,
(3). (r1 r2 )u = r1 (r2 u).
A right R-module can be defined analogously. Here the product of
u ∈ M and r ∈ R is denoted by ur.
Example 3.1.1. (1). R and {0} are naturally left (and also right)
R-modules in the similar manner. These R-modules are called
regular left and regular right R-modules.
(2). Any abelian group (A, +) can be considered a left Z-module
as follows. For g ∈ A and k ∈ Z we defined
kg = g + · · · + g if k > 0, 0Z g = 0A ,
| {z }
k times
and kg = −[(−k)g] if k < 0.
of all n × n matrices over a ring R becomes
(3). The set Mn (R)
a left R-module
if we define
â ì
r 0 0 ··· 0
0 r 0 ··· 0
rX = 0 0 r ··· 0 X, ∀r ∈ R and X ∈ Mn (R).
.. .. ..
..
. . . .
0 0 0 r
67
March 18, 2022 9:45 amsart-9x6 12819-main page 68

68 RING AND FIELD THEORY

Clearly, we can also make Mn (R) a right R-module in the


same manner.
By MR we denote a right R-module M , while R M will denote M
as a left R-module. For convenience, we generally work with left R-
modules while dealing with non-commutative rings. We simply say
that M is a module if other details are clear from the context.
When R is a commutative ring, left R-module and right R-modules
are the same, i.e., if M is a R-module we can use ru = ur for any
r ∈ R and u ∈ M .
Proposition 3.1.2. Let M be a left R-module. Then:
(1). r0M = 0M for all r ∈ R,
(2). 0R u = 0M for all u ∈ M ,
(3). r(−u) = (−r)u = −ru for all u ∈ M and r ∈ R.
Proof. This is easy to prove. 
Definition 3.1.3. Let M be a left R-module.
(a). A subset K of M is called a R-submodule of M if K is also
a left R-module under the same action defined on M , denoted
by K ≤ M .
(b). A nonzero R-module M is called to be simple if {0} and M
are the only submodules of M .
(c). Let S ⊆ R. The submodule of M generated by S is the
smallest submodule of M that contains S, denoted by hSi.
Proposition 3.1.4. Let K be a non-empty subset of R M. Then K ≤
M if and only if u − v ∈ K and rx ∈ K for all u, v ∈ K and r ∈ R.
Proof. This is easy to prove. We omit the details. 
Definition 3.1.5. Submodules of RR are called right ideals of R
and submodules of R R are called left ideals of R.
Let K be a submodule of a left R-module M . Consider the factor
group M/K. Elements of M/K are cosets of the form u + K with
m ∈ M . We can make M/K a left R-module by defining
r(u + K) = ru + K ∀ u ∈ M and r ∈ R.
Check that this action is well defined and the module axioms are
satisfied to make M/K a left R-module. We define the following sets
[[M, K]] = {L ≤ M | L ⊇ K} , [[M/K]] = K 0 ≤ M/K .

March 18, 2022 9:45 amsart-9x6 12819-main page 69

MODULES AND NOETHERIAN RINGS 69

Similar to rings and subrings we have the following correspondence


theorem for submodules.
Proposition 3.1.6. Let K be a submodule of R M .
(1). Every submodule of M/K has the form A/K where A is a
submodule of M and A ⊇ K.
(2). The map [[M, K]] → [[M/K]], L 7→ L/K is a one to one and
onto map.
Definition 3.1.7. Let M and M 0 be left R-modules. A map ϕ :
M → M 0 is called an R-module homomorphism if:
ϕ(u + v) = ϕ(u) + ϕ(v), ∀u, v ∈ M,
ϕ(ru) = rϕ(u), ∀u ∈ M, r ∈ R.
If K is a submodule of R M then the map
σ : M → M/K, σ(m) = m + K ∀m ∈ M
is a homomorphism of M onto M/K. It is called the canonical
homomorphism.
Proposition 3.1.8. Let ϕ :R M →R M 0 be an R-module homomor-
phism. Then:
(1). ϕ(0M ) = 0M 0 ;
(2). ker(ϕ) = {u ∈ M : ϕ(u) = 0M 0 } ≤ M ;
(3). ϕ(M ) = {ϕ(u) : u ∈ M } ≤ M 0 .
Proof. This is easy to prove. We omit the details. 
The above ker(ϕ) is called the kernel of ϕ, and ϕ(M ) is called
the image of ϕ.
Definition 3.1.9. Let ϕ : M → M 0 be an R-homomorphism. Then
ϕ is called an R-isomorphism if it is in addition a one to one and
onto map. In this case we write M ∼
= M 0.
There are similar isomorphism theorems to those for rings.
Theorem 3.1.10 (First isomorphism theorem). Let M and M 0 be
left R-modules and ϕ : M → M 0 and R-homomorphism. Then
ϕ(M ) ∼
= M/ ker(ϕ).
Theorem 3.1.11 (Second isomorphism theorem). Let L, K be sub-
modules of R M . Then (L + K)/K ∼
= L/(L ∩ K).
March 18, 2022 9:45 amsart-9x6 12819-main page 70

70 RING AND FIELD THEORY

Theorem 3.1.12 (Third isomorphism theorem). If K, L are sub-


modules of R M and K ⊆ L then L/K ≤ M/K and (M/K)/(L/K) ∼ =
M/L.
The proofs of these theorems are similar to those for rings.
Let M1 , · · · , Mn be left R-modules. The set of n-tuples {(u1 , · · · ,
un ) : ui ∈ Mi } becomes a left R-modules if we define
(u1 , · · · , un ) + (u01 , · · · , u0n ) = (u1 + u01 , · · · , un + u0n )
and r(u1 , · · · , un ) = (ru1 , · · · , run ). This is the external direct
sum of the Mi and is denoted
⊕ni=1 Mi or M1 ⊕ · · · ⊕ Mn .
For simplicity we denote M n = M ⊕ · · · ⊕ M , the direct sum of
n copies of an R-module M . For convenience, sometimes we use
column vectors to denote elements in M n , say (u1 , · · · , un )t ∈ M n .
Let {Mλ }λ∈Λ be a collection of submodules of a left R-modules
M . We define their sum
X
Mλ = {uλ1 + · · · + uλk : uλi ∈ MΛi for all possible subsets
λ∈Λ
{λ1 , · · · , λk } of Λ}.
P
Thus λ∈Λ Mλ is the set of all finite sums of elements of the Mλ ’s.
Peasy to check that this is a submodule of M . P
It is
λ∈Λ Mλ is said to be direct if each element in λ∈Λ Mλ has a
unique expression as uλ1 + · · · + uλk for some uλi ∈ Mλi . As before
we can show that
 
X  X 
Mλ is direct ⇐⇒ Mµ ∩ Mλ = {0} ∀ µ ∈ Λ.
 
λ∈Λ λ∈Λ,λ6=µ

If λ∈Λ MΛ is direct and Λ is a finite set, we denote it by ⊕ni=1 Mi or


P
M1 ⊕· · ·⊕Mn . As explained for rings before, there is no real difference
between (finite) external and internal direct sums of modules.
Definition 3.1.13. Let R be a unital ring. A module R M is said to
be unital if 1u = u for all u ∈ M .
We shall assume that all modules considered are unital whenever
R is a unital ring. A vector space V over a field F is exactly a unital
F -module.
March 18, 2022 9:45 amsart-9x6 12819-main page 71

MODULES AND NOETHERIAN RINGS 71

Definition 3.1.14 (Cyclic submodule). A submodule of an R-module


M generated by a single nonzero element is called a cyclic submod-
ule of M . An R-module M is call to be cyclic if it can be generated
by a single element.
Definition 3.1.15 (Finitely generated module). An R-module M
is finitely generated if there are u1 , · · · , un ∈ M such that M =
hu1 , · · · , un i.
Let R be a ring with an ideal I and M a left R-module. In general,
M need not be a left R/I-module. However, we can give M a left
R/I-module structure if IM = 0. In this case we define
(r + I)u = ru for all u ∈ R and r ∈ R.
It can be checked that this is a well-defined left R/I-module action.
Further, under this action the R-submodules and R/I-submodules
of M coincide.
Example 3.1.2. Let I, J be ideals of the unital ring R. Show that
R/I ∼
= R/J as left R-modules if and only if I = J.
Proof. If I = J we clearly have R/I ∼ = R/J as left R-modules.
Now suppose that R/I ∼ = R/J as left R-modules. Since I(R/I) =
0 we have I(R/J) = 0, i.e., I ⊂ J. Similarly J ⊂ I. Thus I = J. 
3.2. Free modules.
Let R be a unital ring and M a left R-module.
Definition 3.2.1 (Linear independence). Let u1 , · · · , un ∈ M . Then
{u1 , · · · , un } is linearly independent if
n
X
ri ui = 0 for ri ∈ R
i=1
implies r1 = r2 = · · · = rn = 0.
Definition 3.2.2 (Freely generate). A subset S ⊆ M generates M
freely if
(1) S generates M , and
(2) any set map ϕ : S → N to an R-module N can be extended
to an R-module homomorphism ϕ̃ : M → N .
One can show that the R-module homomorphism ϕ̃ : M → N is
uniquely determined by the map ϕ. See Exercise (15).
March 18, 2022 9:45 amsart-9x6 12819-main page 72

72 RING AND FIELD THEORY

Thus, what this definition tells us is that giving an R-module


homomorphism from M to N is exactly the same thing as giving a
map from S to N .
Definition 3.2.3 (Free module and basis). An R-module M is free
if it is freely generated by some subset S ⊆ M , and S is called a
basis of M .
Similar to what we do in linear algebra, we have
Proposition 3.2.4. Let M be a module over a unital ring R. For a
subset S = {u1 , · · · , un } ⊆ M , the following are equivalent:
(a). S is a basis of M ;
(b). S generates M and S is linearly independent;
(c). Every element of u ∈ M is uniquely expressible as
u = r1 u1 + r2 u2 + · · · + rn un
for some ri ∈ R.
Proof. The proof for the equivalence of (b) and (c) is the same as
in Linear Algebra. So we only show that (a) and (b) are equivalent.
(a)⇒(b). If S is not independent, then we can write
r1 u1 + · · · + rn un = 0,
with ri ∈ M and, say, r1 non-zero. We define the set function ϕ :
S → R by sending u1 7→ 1R and ui 7→ 0 for all i 6= 1. As S generates
M freely, this extends to an R-module homomorphism ϕ̃ : M → R.
By definition of a homomorphism, we can compute
0 = ϕ̃(0)
= ϕ̃(r1 u1 + r2 u2 + · · · + rn un )
= r1 ϕ̃(u1 ) + r2 ϕ̃(u2 ) + · · · + rn ϕ̃(un )
= r1 .
This is a contradiction. So (b) follows.
(b)⇒(a). Suppose every element can be uniquely written as r1 u1 +
· · · + rn un . Given an R-module N any set function ϕ : S → N , we
define ϕ̃ : M → N by
ϕ̃(r1 u1 + · · · + rn un ) = r1 ϕ(u1 ) + · · · + rn ϕ(un ).
This is well-defined by uniqueness, and is clearly a homomorphism.
So it follows that S is a basis of M . 
March 18, 2022 9:45 amsart-9x6 12819-main page 73

MODULES AND NOETHERIAN RINGS 73

Proposition 3.2.5. Let M be a module over a unital ring R. Let


S = {u1 , · · · , un } be a basis for the R-module M . Then M ∼
= Rn .
Proof. We can easily prove that the map
ϕ : Rn → M, (r1 , r2 , · · · , rn ) 7→ r1 u1 + · · · + rn un
is an R-module isomorphism. 
Proposition 3.2.6. If I  R is an ideal and M is an R-module, then
(a). IM = { ni=1 ai ui : ai ∈ I, ui ∈ M } ≤ M ;
P
(b). M/IM is an R/I module via (r + I) · (u + IM ) = r · u + IM
for all r ∈ R, u ∈ M.
Proof. (a). It is easy to see that IM ≤ M.
(b). If b ∈ I, then its action on M/IM is
b(u + IM ) = bu + IM = IM, ∀u ∈ M,
i.e., everything in I kills everything in M/IM . We can consider
M/IM as an R/I module by
(r + I) · (u + IM ) = r · u + IM, ∀r ∈ R, u ∈ M.

We next need to use the following general fact:
Proposition 3.2.7. Every unital ring has a maximal ideal.
Proof. We observe that an ideal I  R is proper if and only if
1 6∈ I. So every increasing union of proper ideals is proper. By
Kuratowski-Zorn Lemma, there is a maximal ideal of R. 
Lemma 3.2.8. Let M1 , M2 , · · · , Mn be R-modules with H1 ≤ M1 ,
H2 ≤ M2 , · · · , Hn ≤ Mn . Then
M1 ⊕ M2 ⊕ · · · ⊕ M n ∼ M 1 M 2 Mn
= ⊕ ⊕ ··· ⊕ .
H1 ⊕ H2 ⊕ · · · ⊕ Hn H1 H2 Hn
Proof. Consider the map
M1 M2 Mn
φ :M1 ⊕ M2 ⊕ · · · ⊕ Mn → ⊕ ⊕ ··· ⊕ ,
H1 H2 Hn
(g1 , g2 , · · · , gn ) 7→ (g1 + H1 , g2 + H2 , · · · , gn + Hn ).
It is easy to see that φ is a module homomorphism with ker(φ) =
H1 ⊕ H2 ⊕ · · · ⊕ Hn . Using the first isomorphism theorem we obtain
the result in the lemma. 
March 18, 2022 9:45 amsart-9x6 12819-main page 74

74 RING AND FIELD THEORY

Theorem 3.2.9 (Invariance of rank). Let R be a unital commutative


ring. If Rn ∼
= Rm as R-modules, then n = m.
Proof. Let I be a maximal ideal of R. Suppose we have Rn ∼
= Rm .
Then we must have
Rn /IRn ∼
= Rm /IRm ,
as R/I modules. Using the fact that
IRn = (IR)n , IRm = (IR)m , Rn /(IR)n ∼ n
= (R/I) ,
and Lemma 3.2.8 we see that
(R/I)n ∼ m
= (R/I) ,
are vector spaces over the field R/I. By Linear Algebra we must
have m = n. 
Definition 3.2.10. Let R be a unital commutative ring. If R-module
M∼= Rk , we say that M is of rank k.
Example 3.2.1. Let R be a unital commutative ring. If every sub-
module of a free R-module is free, show that R is a principal ideal
domain.
Proof. We first show that R has no 0-divisors. Otherwise assume
that ab = 0 for a, b ∈ R \ {0}. Then the principal ideal Rb is a free
module. This is impossible since a(Rb) = 0, i.e., Rb cannot have a
basis.
Let I  R be nonzero. Then I is an R-module. From the assump-
tions we know that I is a free R-module with a basis {ui : i ∈ J}.
If |J| > 1, say 1, 2 ∈ J. We know that Ru1 ∩ Ru2 = 0. Since
u1 u2 ∈ Ru1 ∩ Ru2 , we deduce that u1 u2 = 0 which is impossible.
We obtain that |J| = 1, i.e., I = Ru for some u ∈ R. Thus R is a
principal ideal domain. 
3.3. Finitely generated modules over Euclidean domains.
In this section we will study finitely generated modules over a Eu-
clidean domain D.
Theorem 3.3.1. Let X = {x1 , · · · , xr } be a basis for a free D-
module M and t ∈ D. Then for i 6= j, the set
Y = {x1 , · · · , xj−1 , xj + txi , xj+1 , · · · , xr }
is also a basis for M .
March 18, 2022 9:45 amsart-9x6 12819-main page 75

MODULES AND NOETHERIAN RINGS 75

Proof. Since xj = (−t)xi + (1)(xj + txi ), we see that xj ∈ hY i,


i.e., Y generates M . Let
a1 x1 + · · · + aj−1 xj−1 + aj (xj + txi ) + aj+1 xj+1 + · · · + ar xr = 0,
where ai ∈ D. Then
a1 x1 + · · · + (ai + aj t) xi + · · · + aj xj + · · · + ar xr = 0.
Since X is a basis, we deduce that
a1 = · · · = ai + aj t = · · · = aj = · · · = ar = 0.
We see that
a1 = · · · = ai = · · · = aj = · · · = ar = 0.
So Y is a basis. 
Theorem 3.3.2. Let D be a Euclidean domain with Euclidean norm
φ : R \ {0} → Z+ , M be a nonzero free D-module of finite rank n,
and let K be a nonzero submodule of M .
(1). K is a free D-module of rank s ≤ n;
(2). There exists a basis {x1 , x2 , · · · , xn } for M and d1 , d2 , · · · ,
ds ∈ D with di |di+1 , such that {d1 x1 , d2 x2 , · · · , ds xs } is a
basis for K.
Proof. We only prove (2) since (1) follows from (2).
For any basis Y = {y1 , · · · , yn } of M , all nonzero elements in K
can be expressed in the form
a1 y1 + · · · + an yn , ai ∈ D
where some ai is nonzero.
Step 1: Constructing x1 .
Among all bases Y for M , select one Y1 so that φ(ai ) is minimal as
all nonzero elements of K are written in terms of the basis elements
in Y1 . By renumbering the elements of Y1 if necessary, we can assume
there is w1 ∈ K such that
w1 = d1 y1 + a2 y2 + · · · + an yn
where d1 = 6 0 and φ(d1 ) is the minimal as just described. Write
aj = d1 qj + rj where rj = 0 or φ(rj ) < φ(d1 ) for j = 2, · · · , n. Then
w1 = d1 (y1 + q2 y2 + · · · + qn yn ) + r2 y2 + · · · + rn yn .
March 18, 2022 9:45 amsart-9x6 12819-main page 76

76 RING AND FIELD THEORY

Take x1 = y1 + q2 y2 + · · · + qn yn . By Theorem 3.3.1 then X1 =


{x1 , y2 , · · · , yn } is also a basis for M . From our choice of y1 for
minimal coefficient d1 , we see that r2 = · · · = rn = 0. Thus d1 x1 ∈ K.
Claim 1. If w = d1 x1 + a2 y2 + · · · + an yn ∈ K then d1 |aj for any
j = 2, · · · , n.
Write aj = d1 qj +rj where rj = 0 or φ(rj ) < φ(d1 ) for j = 2, · · · , n.
Then
w = d1 (x1 + q2 y2 + · · · + qn yn ) + r2 y2 + · · · + rn yn .
Now let x01 = x1 +q2 y2 +· · ·+qn yn . By Theorem 3.3.1 then {x01 , y2 , · · · ,
yn } is also a basis for M . From our choice of y1 for minimal coefficient
d1 , we see that r2 = · · · = rn = 0. Claim 1 follows.
Step 2: Constructing x2 .
We will use the basis X1 = {x1 , y2 , · · · , yn }. If a1 x1 + a2 y2 + · · · +
an yn ∈ K can imply that a2 = a3 = · · · = an = 0, then K = hd1 xi.
We are done in this case.
Consider elements of the form
a2 y2 + · · · + an yn ∈ K.
Note that d1 |aj . There is an element in K with minimal φ(ai ) and
ai 6= 0 for some i = 2, 3, · · · , n. By renumbering the elements of X1
we can assume that there is w2 ∈ K such that
w2 = d2 y2 + · · · + an yn
where d2 6= 0 and φ(d2 ) is minimal as just described. Exactly
as in Step 1 and Claim 1, we can modify our basis from X1 =
{x1 , y2 , · · · , yn } to a basis X2 = {x1 , x2 , y3 , · · · , yn } for M where
d1 x1 , d2 x2 ∈ K, and w = a1 x1 + a2 y2 + · · · + an yn ∈ K implies d1 |a1
and d2 |ai for any i = 2, · · · , n.
Step 3: Finishing.
We have the basis X2 = {x1 , x2 , y3 , · · · , yn } for M and examine
elements of K of the form a3 y3 +· · ·+an yn . The pattern is clear. The
process continues until we obtain a basis {x1 , x2 , · · · , xs , ys+1 , · · · , yn }
where the only element of K of the form as+1 ys+1 +· · ·+an yn is zero,
that is, all ai are zero. We then let xs+1 = ys+1 , · · · , xn = yn and
obtain a basis for M of the form described in the statement of the
theorem. 
Similar to the case in Linear Algebra, we can define the three types
of elementary row operations on a matrix A ∈ Mm×n (R) where R is
a commutative unital ring:
March 18, 2022 9:45 amsart-9x6 12819-main page 77

MODULES AND NOETHERIAN RINGS 77

1. Multiply a row by a constant c ∈ U (R). (cRi → Ri )


2. Interchange two rows. (Ri ↔ Rj for i 6= j)
3. Add a constant c times one row to another. (cRi + Rj → Rj
for i 6= j)
If B is the matrix that results from A by performing one of the
above operations, then the matrix A can be recovered from B by
performing the corresponding operation in the following list:
1. Multiply the same row by 1/c. (c−1 Ri → Ri )
2. Interchange the same two rows. (Ri ↔ Rj for i 6= j)
3. If B resulted by adding c times row ri of A to row rj , then add
−c times ri to rj . (−cRi + Rj → Rj for i 6= j)
It follows that if B is obtained from A by performing a sequence
of elementary row operations, then there is a second sequence of
elementary row operations, which when applied to B recovers A.

Definition 3.3.3. A matrix E ∈ Mn (R) is called an elementary


matrix over a commutative unital ring R if it can be obtained from
an identity matrix by performing a single elementary row operation.

Definition 3.3.4 (Invertible matrices). Let R be a commutative uni-


tal ring. A matrix A ∈ Mn (R) is invertible if there is A−1 ∈ Mn (R)
such that

AA−1 = A−1 A = In .

We denote the set of all invertible n×n matrices in Mn (R) by GLn (R)
which is called the general linear group of rank n over R.

Using adjoint matrix of A, one can easily show that A ∈ Mn (R)


is invertible if and only if det(A) ∈ U (R). Unlike in Liner Algebra,
now det(A) 6= 0 cannot generally deduce the invertibility of A.

Definition 3.3.5 (Equivalent matrices). Two m × n matrices A, B


over a commutative unital ring R are equivalent if B = P AQ for
some invertible matrices P ∈ GLm (R) and Q ∈ GLn (R).

We will prove the following important result.


March 18, 2022 9:45 amsart-9x6 12819-main page 78

78 RING AND FIELD THEORY

Theorem 3.3.6 (Smith normal form). An m × n matrix A over a


Euclidean domain D is equivalent to a matrix of the form
  
d1
 d2  
O
  
 ..  1

 .  

 dr 
O2 O3
with all the di ∈ D non-zero and d1 | d2 , d2 | d3 , · · · dr−1 | dr , where
O1 , O2 , O3 are zero matrices.
Before proving this theorem we will generalize some results from
Linear algebra. Assume that V, W are free D-modules.
Definition 3.3.7. A map L : V → W is called a linear map from
V to W if, for all u, v ∈ V and c ∈ D, we have
(a). L(u + v) = L(u) + L(v),
(b). L(cu) = cL(u).
We often simply call L linear. Linear maps from V into itself are
also called a linear operators on V . We denote the set of all linear
maps from V to W by L(V, W ), or by HomD (V, W ). In the case that
V = W, we simply write L(V, V ) as L(V ) which is also denoted by
EndD (V ).
Definition 3.3.8. An ordered basis for a free D-module V is a
basis for V with a specific order. If β = {u1 , u2 , · · · , un } is an ordered
basis for V , for convenience we simply write it as β = (u1 , u2 , · · · , un )
and consider it as a 1×n matrix. So an ordered basis is a 1×n matrix
of vectors.
We call (e1 , e2 , · · · , en ) the standard ordered basis for the mod-
ule Dn where ei = (δ1,i , δ2,i , · · · , δn,i )t and we consider
Dn = (a1 , a2 , · · · , an )t : ai ∈ D .


Theorem 3.3.9. Let β = (v1 , v2 , · · · , vn ) be an ordered basis for a


free D-module V . For A, B ∈ Mn×m (D), if βA = βB, then A = B.
Proof. Let A = (aij ) and B = (bij ). From βA = βB we see that
a1i v1 + a2i v2 + · · · + ani vn = b1i v1 + b2i v2 + · · · + bni vn .
Since v1 , v2 , · · · , vn are linearly independent, we deduce that aij = bij
for all i, j, that is A = B. 
March 18, 2022 9:45 amsart-9x6 12819-main page 79

MODULES AND NOETHERIAN RINGS 79

Suppose that β = (v1 , v2 , · · · , vn ) and γ = (w1 , w2 , · · · , wm ) are


ordered bases for V and W , respectively. Let L ∈ L(V, W ). Then
there exist unique scalars aij ∈ D for each 1 ≤ i, j ≤ m, such that
L(v1 ) = a11 w1 + a21 w2 + · · · + am1 wm
L(v2 ) = a12 w1 + a22 w2 + · · · + am2 wm
(3.1)
············
L(vn ) = a1n w1 + a2n w2 + · · · + amn wm .
We define the matrix
 
a1,1 a1,2 ··· a1,n
 a2,1 a2,2 ··· a2,n 
A =  .. ..  .
 
.. ..
 . . . . 
am,1 am,2 · · · am,n
Definition 3.3.10. We call the above m × n matrix A = (aij ) the
matrix of L in the ordered bases β and γ and denote A = [L]γβ .
We can write the formula as matrix product
L(β) = (L(v1 ), L(v2 ), · · · , L(vn ))
(3.2)
= (w1 , w2 , · · · , wm )[L]γβ = γ[L]γβ .
Note that many properties for linear maps in Linear Algebra triv-
ially hold here for D-linear maps also. For example L is invertible if
and only if the matrix [L]γβ is invertible.
Proof of Theorem 3.3.6. Let α, β be the standard ordered basis
for Dn , Dm respectively. Consider the D-linear map
   
a1 a1
 a2   a2 
LA : Dn → Dm ,  ..  7→ A  ..  .
   
. .
an an

Then [LA ]βα = A, LA (α) = βA and LA (Dn ) ≤ Dm . Using Theo-


rem 3.3.2 we have an ordered basis β 0 = (x1 , x2 , . . . , xm ) for Dm and
d1 , d2 , . . . , ds ∈ D with di |di+1 , such that (d1 x1 , d2 x2 , . . . , ds xs ) is a
basis for LA (Dn ).
There are y1 , y2 , . . . , yr ∈ Dn such that LA (yi ) = di xi . Let (yr+1 ,
. . . , ys ) be an ordered basis for ker(LA ). We can easily show that
March 18, 2022 9:45 amsart-9x6 12819-main page 80

80 RING AND FIELD THEORY

α0 = (y1 , y2 , . . . , ys ) is a basis for Dn . So s = n and


  
d1
 d2  
O
0
  
β
[LA ]α0 = 
 . .. 
 1
.

 
 dr 
O2 O3
Using Linear Algebra formulas (say, [Z, Theorem 4.23]) we have
0 0
[LA ]βα0 = [IDn ]ββ [LA ]βα [IDm ]αα0 = P AQ
0
where P = [IDn ]ββ ∈ GLm (D), Q = [IDm ]αα0 ∈ GLm (D). 
We can actually find the Smith normal form of an m × n matrix A
over a Euclidean domain D by simply applying elementary row and
column operations to A, see Example 3.3.1.
Definition 3.3.11 (Invariant factors). The elements dk ∈ D ob-
tained in the Smith normal form of the m × n matrix A are called
the invariant factors of A.
Theorem 3.3.12 (Classification of finitely-generated modules over
a Euclidean domain). Let D be a Euclidean domain, and M be a
finitely generated D-module. Then
M∼ = D/hd1 i ⊕ D/hd2 i ⊕ · · · ⊕ D/hdr i ⊕ Dn
for some nonzero nonunit di ∈ D with d1 | d2 , d2 | d3 , · · · dr−1 | dr .
Proof. Since M is finitely-generated, there is a surjective module
homomorphism φ : Dm → M . So by the first isomorphism, we have
M∼ = Dm /ker φ.
Since ker φ is a submodule of Dm , by the previous theorem, there is a
basis v1 , · · · , vm of Dm such that ker φ is generated by d1 v1 , · · · , dr vr
for some nonzero di ∈ D with d1 | d2 , d2 | d3 , · · · dr−1 | dr . So we
know
Dm
M∼ = .
h(d1 , 0, · · · , 0), (0, d2 , 0, · · · , 0), · · · , (0, · · · , 0, dr , 0, · · · , 0)i
This is just
D D D
⊕ ⊕ ··· ⊕ ⊕ D ⊕ · · · ⊕ D,
hd1 i hd2 i hdr i
March 18, 2022 9:45 amsart-9x6 12819-main page 81

MODULES AND NOETHERIAN RINGS 81

with m − r copies of D. If di ∈ U (D) we see that R/hdi i = 0. So we


can delete this unit di . 
A different form of the classification of finitely-generated modules
over a Euclidean domain can be found in Exercise (20). We point
out that all results so far in this section hold also for principal ideal
domains, see [J, Chapter 3].
We can consider any additive abelian group as a Z-module. Tak-
ing D = Z, and applying the classification of finitely generated
D-modules, we recover the following classification theorem for finitely-
generated abelian groups (Theorem 1.6.4).

Theorem 3.3.13 (Classification of finitely-generated abelian groups).


Every nontrivial finitely generated abelian group G is isomorphic to
a group of direct product of nontrivial cyclic groups

Zm1 × Zm2 × · · · × Zmr × Zn ,

where r, n ∈ Z+ , m1 , · · · , mr ∈ 1 + Z+ with mi |mi+1 for i = 1, 2, · · · ,


r − 1.

Example 3.3.1. Find the Smith normal form of the integer matrix
 
0 1 1
A= 1 0 −3  .
1 −3 0

Proof. We compute that


   
0 1 1 0 1 0
R3 −R2 →R3 
A= 1 0 −3  −− −−−−−→ 1 0 −3 
C3 −C2 →C3
1 −3 0 0 −3 6

   
0 1 0 1 1 0
C +3C1 →C3  R2 +R1 →R1 
−−3−−−− −−→ 1 0 0  −− −−−−−→ 1 0 0 
R3 +3R1 →R3
0 0 6 0 0 6
   
1 0 0 1 0 0
R2 −R1 →R2  −R →R2 
−− −−−−−→ 0 −1 0  −−−2−−−→ 0 1 0 ,
C2 −C1 →C2
0 0 6 0 0 6
which is the Smith normal form of A. 
March 18, 2022 9:45 amsart-9x6 12819-main page 82

82 RING AND FIELD THEORY

3.4. Noetherian rings.


We now introduce Noetherian rings, left Noetherian rings, and right
Noetherian rings in this section and prove Hilbert basis theorem.
Definition 3.4.1 (Noetherian ring). A ring R is Noetherian (or
left Noetherian, or right Noetherian, respectively) if any chain
of ideals (or left ideals, or right ideals, respectively) of R
I1 ⊆ I2 ⊆ I3 ⊆ · · · ,
satisfies ACC, i.e., there is some n ∈ N such that In = In+1 =
In+2 = · · · .
Example 3.4.1. (a). Every finite ring is Noetherian, left Noe-
therian and right Noetherian.
(b). Every principal ideal domain D is Noetherian. This is because
any nonzero and nonunit a ∈ D has only finitely factors up
to associates.
(c). The ring Z[x1 , x2 , x3 , · · · ] is not Noetherian since it has the
chain of strictly increasing ideals
hx1 i ⊂ hx1 , x2 i ⊂ hx1 , x2 , x3 i ⊂ · · · .
We have the following proposition that makes Noetherian rings
much more concrete, and makes it obvious why PIDs are Noetherian.
Definition 3.4.2 (Finitely generated ideal). An ideal I of a ring
R is finitely generated if there are r1 , · · · , rn ∈ R such that I =
hr1 , · · · , rn i.
Similarly, we can define finitely-generated left ideals and finitely-
generated right ideals.
Proposition 3.4.3. A ring R is Noetherian (or left Noetherian, or
right Noetherian, respectively) if and only if every ideal (or left ideal,
or right ideal, respectively) of R is finitely generated.
Proof. (⇒). Suppose every ideal of R is finitely generated. Given
the chain I1 ⊆ I2 ⊆ · · · , we have the ideal
I = I1 ∪ I2 ∪ I3 ∪ · · · .
We know I is finitely generated, say I = hr1 , · · · , rn i, with ri ∈ Iki .
Let
n = max {ki }.
i=1,··· ,n
March 18, 2022 9:45 amsart-9x6 12819-main page 83

MODULES AND NOETHERIAN RINGS 83

Then r1 , · · · , rn ∈ IK . So In = I, and furthermore In = In+1 =


In+2 = · · · .
(⇐). Suppose there is an ideal I  R that is not finitely generated.
We pick r1 ∈ I. Since I is not finitely generated, we know hr1 i 6= I.
So we can find some r2 ∈ I \ (r1 ).
Again hr1 , r2 i 6= I. So we can find r3 ∈ I \ hr1 , r2 i. We continue
on, and then can find an infinite strictly ascending chain
hr1 i ⊆ hr1 , r2 i ⊆ hr1 , r2 , r3 i ⊆ · · · .
So R is not Noetherian.
For left Noetherian, or right Noetherian cases the proof is similar.

If R is Noetherian, not necessarily every subring of R has to be
Noetherian. For example, since Z[x1 , x2 , · · · ] is an integral domain,
we can take its field F of fractions, which is a field, hence Noetherian,
but Z[x1 , x2 , · · · ] is a subring of F . For quotient rings we have the
following result.
Proposition 3.4.4. Let R be a Noetherian ring and I  R. Then
R/I is Noetherian.
Proof. Consider the natural homomorphism
π : R → R/I, x 7→ x + I.
Let J  R/I. We want to show that J is finitely generated. We know
that π −1 (J)  R, and is hence finitely generated, since R is Noe-
therian. So π −1 (J) = hr1 , · · · , rn i for some r1 , · · · , rn ∈ R. Then J
is generated by π(r1 ), · · · , π(rn ). So R/I is Noetherian by Proposi-
tion 3.4.3. 
Now we can prove the following powerful theorem which was, sur-
prisingly, proven by David Hilbert (1862–1943) in 1890:
Theorem 3.4.5 (Hilbert basis theorem). Let R be a Noetherian ring.
Then so is R[x].
Proof. To the contrary, suppose a ⊆ R[x] is a non-finitely-
generated ideal. Then by recursion there is a sequence {f0 , f1 , . . .} ⊂
a such that if bn with n ≥ 1 is the ideal generated by f0 , . . . , fn−1 ,
then fn ∈ a \ bn is of minimal degree. It is clear that {deg(f0 ), deg(f1 ),
. . .} is a non-decreasing sequence of nonnegative integers. Let an be
the leading coefficient of fn and let b be the ideal of R generated by
March 18, 2022 9:45 amsart-9x6 12819-main page 84

84 RING AND FIELD THEORY

a0 , a1 , . . .. Since R is Noetherian the chain of ideals


ha0 i ⊂ ha0 , a1 i ⊂ ha0 , a1 , a2 i ⊂ · · ·
must terminate. Suppose that b = ha0 , . . . , an−1 i for some integer n.
So in particular,
ni
XX
an = ui,j ai vi,j , ui,j , vi,j ∈ R,
i<n j=1

where the sum is finite. Now consider


XX ni
g= ui,j xdeg(fn )−deg(fi ) fi vi,j ∈ bn ,
i<n j=1

whose leading term is equal to that of fn . However, fn ∈


/ bn , which
means that
fn − g ∈ a \ bn
has degree less than fn , contradicting the minimality. 
Modifying the above proof we can have the following result:
Theorem 3.4.6. Let R be a left (or right) Noetherian ring. Then
so is R[x].
A direct consequence of the above results are the following corol-
lary.
Corollary 3.4.7. If R is a Noetherian ring (or left Noetherian ring,
or right Noetherian ring, respectively), then the polynomial ring R[x1 ,
. . . , xn ] in commutative indeterminates x1 , . . . , xn is a Noetherian
ring (or left Noetherian ring, or right Noetherian ring, respectively).
We now explain an application of the above theorem. Let S ⊆
F [x1 , x2 , · · · , xn ] be any set of polynomials where F is an arbitrary
field. We define the zero-locus Z(S) to be the set of points in F n on
which the functions in S simultaneously vanish, that is
Z(S) = {α ∈ F n | f (α) = 0 for all f ∈ S} .
A subset V of F n is called an affine algebraic set if V = Z(S) for
some S ⊆ F [x1 , x2 , · · · , xn ].
We view this as a set of equations f = 0 for each f ∈ S. The
claim is that to solve the potentially infinite set of equations S, we
actually only have to solve finitely many equations.
March 18, 2022 9:45 amsart-9x6 12819-main page 85

MODULES AND NOETHERIAN RINGS 85

Consider the ideal hSi  F [x1 , · · · , xn ]. By the Hilbert basis the-


orem, there is a finite list f1 , · · · , fk such that
hf1 , · · · , fk i = hSi.
We can easily see that
Z(S) = Z(hSi) = Z(f1 , · · · , fk ).
So solving S is the same as solving f1 , · · · , fk . This is extremely
useful.
Next we will only introduce some important concepts in Commu-
tative Algebra.
Definition 3.4.8. The Krull dimension of a unital ring R, denoted
by dim R, is the maximum length n of a chain I0 ⊂ I1 ⊂ · · · ⊂ In of
prime ideals of R. If there is no upper bound on the length of such a
chain, we take n = ∞.
Example 3.4.2. (a). A field F has Krull dimension 0.
(b). The polynomial ring F [x1 , x2 , · · · , xn ] has Krull dimension n.
(c). A principal ideal domain that is not a field has Krull dimen-
sion 1.
(d). The Krull dimension of the non-Noetherian ring F [x1 , x2 , · · · ]
is infinity, where F is a field. We have the infinite chain of
prime ideals
hx1 i ⊂ hx1 , x2 i ⊂ hx1 , x2 , x3 i ⊂ · · · .
Definition 3.4.9. The height of a prime ideal P of a unital ring
R is the maximum length n of a chain of prime ideals I0 ⊂ I1 ⊂
· · · ⊂ In = P of R.
Definition 3.4.10. Let P be a maximal ideal of a unital ring R. The
sequence a1 , · · · , at of nonzero elements in P is a regular sequence
for R, if each ai is not a zero-divisor of R/ha1 , · · · , ai−1 i.
Definition 3.4.11. The depth of a ring R is the maximal number
of elements in some regular sequence, denoted by depth(R). A ring
in which depth(R) = dim(R) is called a Cohen–Macaulay ring.
Example 3.4.3. Let F be a field.
(a). The rings F [x1 , · · · , xn ], F [x, y]/hxyi, F [x, y, z]/hxy, xz, yzi,
F [x, y, z]/hxy −zi, and F [x, y, z, w]/hxy −zwi are all Cohen–
Macaulay.
March 18, 2022 9:45 amsart-9x6 12819-main page 86

86 RING AND FIELD THEORY

(b). None of the rings F [x, y]/hx2 , xyi, F [x, y, z]/hxy, xzi, or
F [x, y, z, w]/hwy, wz, xy, xzi is Cohen–Macaulay.
We leave the proofs as an exercise.
3.5. Exercises.
(1) Classify all simple modules over the ring R = (Z, +, ·).
(2) Classify all simple modules over the ring R = (2Z, +, ·).
(3) Classify all simple modules over the ring R = (Q[x], +, ·).
(4) Classify all simple modules over the ring R = (R[x], +, ·).
(5) Find the Smith normal form of the integer matrix
 
5 −417 129 50
 −6 111 −36 6 
A=  5 −672 210 74  .

−7 255 −81 −10


(6) Find the invariant factors of the integer matrix
 
2 4 4
A = −6 6 12.
10 4 16
(7) Without doing any elementary row or column operations find
the Smith normal form of the integer matrix
   
2 0 0 15 0 0
A =  0 3 0  , B =  0 10 0  .
0 0 4 0 0 6
(8) Find the Smith normal form of the matrix over C[x]:
 
1−x 1 1
 0 1−x 0 .
0 1 2−x
(9) Find the invariant factors of the matrix over Q[x]:
2 2x + 3 2x2 + 3x
 
 1 6x 6x2 + 6x  .
1 3 x
(10) Suppose that V be a finitely generated module over C[x] that
is not a free module over C[x]. Show that x has an eigenvector
on V .
March 18, 2022 9:45 amsart-9x6 12819-main page 87

MODULES AND NOETHERIAN RINGS 87

(11) Find a basis for the Z[x]-submodule of Z[x]-module Z[x]3


generated by u1 = (3x + 1, 3x + 2, (x + 1)2 ), u1 = (x + 1, 2x +
1, x2 ), u1 = (1 + x, 1, 2x + 1).
(12) Let M be a left R-module. Show that M is finitely generated
if there exists a submodule N ⊂ M such that N and M/N
are both finitely generated.
(13) Show that if x2 = 0 implies x = 0, for all x in the ring R,
then all idempotent elements of R are central.
(14) Let R be a commutative ring with a unique maximal ideal I,
and let M be a nonzero finitely generated R-module. Show
that HomR (M, R/I) 6= 0.
(15) In Definition 3.2.2 show that the R-module homomorphism
ϕ̃ : M → N is uniquely determined by the map ϕ.
(16) Let R be a unital ring, and let M be a left R-module that
has a minimal submodule S such that M/S = S. Prove that
either S is a direct summand of M , in which case M = S ⊕ S,
or else S is the only proper nontrivial submodule of M.
(17) Let A and B be finitely generated modules over a Euclidean
domain D. If A ⊕ A ∼ = B ⊕ B, prove that A ∼ = B.
(18) Let R be any ring, A a R-module, and A = B ⊕ C = D ⊕ E
two direct sum decompositions of A. Let f be the projection
on B, restricted to D, let g be the projection on E, restricted
to C.
(a). If f is one-to-one, so is g.
(b). If f is onto, so is g.
(19) Let D be a Euclidean domain, and a, b ∈ D be such that
gcd(a, b) = 1. Show that D/habi ∼ = D/hai ⊕ D/hbi as D-
modules.
(20) Let D be a Euclidean domain, and M be a finitely-generated
R-module. Show that

M∼
= N1 ⊕ N2 ⊕ · · · ⊕ Nr ,

where each Ni is either D or is D/hpn i for some prime p ∈ D


and some n ≥ 1.
(21) The ideal I = h2, xi of the ring R = (Z[x], +, ·) is not a direct
sum of cyclic Z[x]-modules.
(22) Determine all simple modules over the ring R = (Z[x], +, ·).
March 18, 2022 9:45 amsart-9x6 12819-main page 88

88 RING AND FIELD THEORY

(23) In the non-Noetherian ring F [x1 , x2 , · · · ], where F is a field,


show that the ideals hx1 , x2 , · · · , xn i are prime for any posi-
tive integer n.
(24) Show that an R-module homomorphism between two simple
modules over a unital ring R is zero or an isomorphism.
(25) Let M be a nonzero finitely generated module over a ring R.
Show that M has a maximal submodule.
(26) Let R be a unital ring, and n ∈ N. Show that R is Noetherian
if and only if Mn (R) is Noetherian.
(27) Prove Theorem 3.4.6.
(28) Let R be a left-Noetherian ring and M be a finitely-generated
left R-module. Show that any submodule of M is finitely-
generated. (Hints: Use induction on the size of the generating
set.)
(29) A commutative ring R is finitely generated if there are u1 , · · · ,
un ∈ R such that every element of R is a finite sum of ele-
ments of the form
auk11 · · · uknn , where a ∈ Z, ki ∈ N.
Show that every finitely generated commutative ring R is
Noetherian.
(30) Let M be a finitely generated module over a unital commuta-
tive ring R, and I ≤ R such that IM = M . Show that there
exists a ∈ I such that (1 + a)M = 0. (Hints: Use adjoint
matrix.)
(31) (Cohen’s Theorem). Let R be a unital commutative ring.
Show that R is Noetherian if and only if all prime ideals of
R are finitely generated. (Hints: By contradiction.)
(32) Prove Example 3.4.3.
(33) Show that the ring C[x1 , · · · , x5 ]/I where I = hx1 x5 , x1 x2 ,
x2 x3 , x3 x4 , x4 x5 i is Cohen–Macaulay.
March 18, 2022 9:45 amsart-9x6 12819-main page 89

4. Fields and Extension Fields

Historically, three algebraic disciplines led to the concept of a field:


the question of solving polynomial equations, algebraic number the-
ory, and algebraic geometry. A first step towards the notion of a field
was made in 1770 by Joseph-Louis Lagrange. The first clear defini-
tion of an abstract field (1893) is due to Heinrich Martin Weber. In
particular, Weber’s notion included the field Zp .
In this chapter we will study the basic theory on the extension
fields, including algebraic extensions and transcendental extensions,
and establish basic results for finite fields.

4.1. Prime fields and extension fields.


We first introduce some basic definitions on prime fields and exten-
sion fields.
Let E be a field. Recall that a subset F of E is called a subfield of
E if F is a field with the same operations as E, denoted by F ≤ E.
It is easy to see that if F ⊆ E, then F is a subfield of E if and only
if
1, a − b, ab−1 ∈ F, ∀a, b ∈ F with b 6= 0.
It is clear that the intersection of all subfields of E is again a sub-
field. The following theorem describes the structure of the smallest
subfield of E.
Theorem 4.1.1. Let E be a field and P the smallest subfield of E.
(1). If char(E) = 0, then P is isomorphic to Q.
(2). If char(E) = p is a prime, then P is isomorphic to Zp .
Proof. Let F be any subfield of E. Then 1 ∈ F, and Z · 1 ⊆ F.
(1). If char(E) = 0, then the subring Z · 1 of F is isomorphic to
Z. So F must contain a quotient field of this subring and that this
quotient field must be isomorphic to Q. Thus, F contains a subfield
isomorphic to Q. Therefore, P ∼= Q.
(2). If char(E) = p, then the subring Z · 1 of F is isomorphic to
Zp . It follows that P ∼
= Zp . 
89
March 18, 2022 9:45 amsart-9x6 12819-main page 90

90 RING AND FIELD THEORY

Definition 4.1.2. (1). The fields Q and Zp for any prime p are
call prime fields.
(2). If F ≤ E are fields, then E is called an extension field
of F .

We see that every field F is an extension field of the prime subfield


P of F. In particular, R is an extension field of Q, and C is an
extension field of both R and Q. If F ≤ E are fields, we can consider E
as a vector space over F. We denote [E : F ] = dimF E, the dimension
of E as a vector space over F .

Definition 4.1.3. Let E be an extension of a field F. If [E : F ] =


n < ∞, then E is called a finite extension of degree n over F .
Otherwise, we say that E is an infinite extension field over F .

Although a field E is a finite extension of a field F , it does not


mean that E is a finite field. It means that E is a finite-dimensional
vector space over F .
Notice that [E : F ] = 1 if and only if E = F.

Theorem 4.1.4. Suppose F ≤ K ≤ E are fields.


(1). [E : F ] is finite if and only if [E : K] and [K : F ] are finite.
(2). If [E : F ] < ∞, then [E : F ] = [E : K][K : F ].

Proof. (1). (⇒). Assume that [E : F ] < ∞. It follows that [K :


F ] < ∞ since K is a subspace of E over F. Let γ = {γ1 , γ2 , · · · , γn }
be a basis for E as a vector space over F . Then every element of
E is a linear combination of γ with coefficients in F and hence with
coefficients in K. Therefore, by a result from linear algebra, a subset
of γ forms a basis of E over K. Thus [E : K] < ∞.
(⇐). Suppose that [E : K] and [K : F ] are finite. Let α =
{α1 , α2 , · · · , αr } be a basis of K as a vector space over F , and let
β = {β1 , β2 , · · · , βs } be a base of E as a vector space over K. It is
enough to show that the rs elements αi βj form a basis for E as a
vector space over F.
For any γ ∈ E since β is a basis for E over K, we have
s
X
γ= bj βj
j=1
March 18, 2022 9:45 amsart-9x6 12819-main page 91

FIELDS AND EXTENSION FIELDS 91

where bj ∈ K. Since α is a basis for K over F , we have


r
X
bj = aij αi
i=1
where aij ∈ F. Then we have
r s
!
X X X
γ= aij αi βj = aij (αi βj ),
j=1 i=1 i,j

so every element of E is an F -linear combination of the rs vectors


αi βj .
P rs elements αi βj are linearly independent
Next we show that the
over F . Suppose that i,j aij (αi βj ) = 0, with aij ∈ F. Then
s r
!
X X
aij αi βj = 0
j=1 i=1
Pn
and i=1 aij αi ∈ K. Since β is independent over E, we see that
r
X
aij αi = 0
i=1
for all j. Since α is independent over F , so aij = 0 for all i and j.
Thus {αi βj : i = 1, 2, · · · , r; j = 1, 2, · · · , s} forms a basis for E over
F and
[E : F ] = [E : K][K : F ] < ∞,
and (2) follows also. 
A direct consequence of the above theorem is the following result.
Corollary 4.1.5. Let E1 ≤ E2 ≤ · · · ≤ Es be fields and [Ei+1 :
Ei ] < ∞ for i = 1, · · · , s − 1. Then [Es : E1 ] is finite and
[Es : E1 ] = [Es : Es−1 ][Es−1 : Es−2 ] · · · [E2 : E1 ].
A very useful observation is that if [E : F ] is finite, then [K :
F ] [E : F ] for any K ≤ E.
Definition 4.1.6. Let F ≤ E be fields and ∅ =
6 S ⊆ E. The field
\
F (S) = K
S∪F ⊆K≤E

is called the subfield generated by S over F.


March 18, 2022 9:45 amsart-9x6 12819-main page 92

92 RING AND FIELD THEORY

It is not hard to see that the subfield F (S) is the minimum sub-
field of E containing F and S. If S = {a1 , a2 , · · · , an } is a finite
subset of E, we write F (S) as F (a1 , a2 , · · · , an ). Clearly, elements in
F (a1 , a2 , · · · , an ) are of the form:
f (a1 , a2 , · · · , an )
,
g(a1 , a2 , · · · , an )
where f, g ∈ F [x1 , x2 , · · · , xn ] with g(a1 , a2 , · · · , an ) 6= 0.
In particular, for any a ∈ E,
ß ™
f (a)
F (a) = : f, g ∈ F [x], g(a) 6= 0 .
g(a)
Definition 4.1.7. If F ≤ E be fields and E = F (a) for some a ∈ E,
then E is called a simple extension field of F .
Example 4.1.1.√ We regard√R as an extension √field of Q. It is easy
to see that Q( 2) = {a + b 2 : a, b ∈ Q} = Q[ 2].
4.2. Algebraic and transcendental elements.
In this section we always assume that F is a field. We firstly prove an
important result that follows quickly and elegantly. This theorem is
named after Leopold Kronecker (1823–1891) who proved it in 1884.
Theorem 4.2.1 (Kronecker’s Theorem). For any field F and f (x) ∈
F [x] with deg(f (x)) > 0, there exists an extension field E of F and
an α ∈ E such that f (α) = 0.
Proof. We have known that f (x) can be written as a product
of irreducible polynomials in F [x]. Now let p(x) be an irreducible
factor of f (x). Thus it is sufficient to find an extension field E of F
containing an element α such that p(α) = 0.
Since p(x) is irreducible in F [x], from Theorem 1.5.13 we see that
hp(x)i is a maximal ideal of F [x], and further E = F [x]/hp(x)i is a
field by Theorem 1.5.4. We first want to identify F with a subfield
of F [x]/hp(x)i in a natural way. Define the map
ψ : F → E, ψ(a) = a + hp(x)i, ∀a ∈ F.
It is easy to see that ψ is a ring homomorphism. If ψ(a) = ψ(b) for
a, b ∈ F , that is, if a + hp(x)i = b + hp(x)i for some a, b ∈ F , then
a − b ∈ hp(x)i, i.e., p(x)|a − b. Thus we deduce that a − b = 0, so
a = b. Then ψ is one to one.
March 18, 2022 9:45 amsart-9x6 12819-main page 93

FIELDS AND EXTENSION FIELDS 93

So ψ maps F one-to-one onto a subfield of F [x]/hp(x)i. We may


identify F with Im(ψ) = {a + hp(x)i : a ∈ F }. Thus we shall view
E = F [x]/hp(x)i as an extension field of F. Next we show that E
contains a zero of p(x).
Consider α = x + hp(x)i ∈ E. Take the evaluation homomor-
phism
φα : F [x] → E, φα (g(x)) = g(α), ∀g(x) ∈ F [x].
If p(x) = a0 + a1 x + · · · + an xn , where ai ∈ F, then we have
φα (p(x)) = a0 + a1 (x + hp(x)i) + · · · + an (x + hp(x)i)n
in E = F [x]/hp(x)i. Therefore,
p(α) = a0 + a1 x + · · · + an xn + hp(x)i = p(x) + hp(x)i = hp(x)i = 0
in F [x]/hp(x)i. Thus, p(α) = 0, and therefore f (α) = 0. 
We illustrate the construction involved in the proof by an example.
Example 4.2.1. Take F = R and let f (x) = x2 + 1 ∈ R[x]. We
know that f (x) has no zeros in R and thus is irreducible in R[x]. Then
hx2 +1i is a maximal ideal in R[x], and further R[x]/hx2 +1i is a field.
Identifying r ∈ R with r + hx2 + 1i in R[x]/hx2 + 1i, we can view R as
a subfield of E = R[x]/hx2 +1i. Take α = x+hx2 +1i ∈ R[x]/hx2 +1i.
In R[x]/hx2 + 1i, we compute
α2 + 1 = (x + hx2 + 1i)2 + (1 + hx2 + 1i) = (x2 + 1) + hx2 + 1i = 0.
So α is a zero of x2 + 1. We shall identify R[x]/hx2 + 1i with C.
In the following we put an element of an extension field E of a
field F into one of two categories.
Definition 4.2.2. Let F ≤ E be fields and α ∈ E. If f (α) = 0
for some nonzero f (x) ∈ F [x], then α is called algebraic over F.
Otherwise α is called transcendental over F.
Example√ 4.2.2. (1). Regard C as an extension√field of Q. Since
3 is a zero of x2 − 3 ∈ Q[x], we see that 3 is an algebraic
element over Q. Also, i is an algebraic element over Q, since
i is a zero of x2 + 1 ∈ Q[x].
(2). It is well known (but the proof is not easy) that the real num-
bers π and e are transcendental over Q. Here e is the base
for the natural logarithm.
March 18, 2022 9:45 amsart-9x6 12819-main page 94

94 RING AND FIELD THEORY

(3). The real number π is transcendental over Q. However, π is


over R, for it is a zero of x − π ∈ R[x].
algebraic p

(4). Let α = 1 + 2 ∈ R. It√is easy to see that α is algebraic
over Q. Since α2 − 1 = 2 and (α2 − 1)2 = 2. Therefore
α4 − 2α2 − 1 = 0, so α is a zero of the polynomial f (x) =
x4 − 2x2 − 1 ∈ Q[x].
Definition 4.2.3. Regard C as an extension field of Q. The complex
number α is called an algebraic number if it is algebraic over Q,
otherwise, α is called a transcendental number.
The definition above connect these ideas in the theory of field with
those in number theory. There is an extensive and elegant theory of
algebraic numbers in number theory.
The next theorem provides a useful characterization of algebraic
and transcendental elements over F in an extension field E of F .
Theorem 4.2.4. Let F ≤ E be fields and let α ∈ E. Then α is
transcendental over F if and only if the evaluation homomorphism
φα : F [x] → E defined by φα (g(x)) = g(α) for g(x) ∈ F [x] is one-to-
one. Thus, F [x] ∼
= im(φα ) = F [α].
Proof. By definition, α is transcendental over F if and only if
f (α) 6= 0 for all nonzero f (x) ∈ F [x], if and only if φα (f (x)) 6= 0 for
all nonzero f (x) ∈ F [x], if and only if the kernel of φα is 0, if and
only if φα is one-to-one. The isomorphism F [x] ∼ = im(φα ) = F [α]
follows from Theorem 1.2.10. 
The next theorem plays a central role in our later sections.
Theorem 4.2.5. Let F ≤ E be fields and let α ∈ E be algebraic
over F .
(1). There is an unique monic irreducible polynomial p(x) ∈ F [x]
such that p(α) = 0.
(2). If f (α) = 0 for f (x) ∈ F [x], then p(x)|f (x).
Proof. (1). Let φα : F [x] → E be the evaluation homomorphism.
We know that ker(φα )  F [x]. Since F [x] is a PID, then ker(φα ) =
hp(x)i for some monic p(x) ∈ F [x].
Now we prove that p(x) is irreducible. Suppose that p(x) =
r(x)s(x) with deg(r(x)), deg(s(x)) < deg(p(x)). Since p(α) = 0 then
r(α)s(α) = 0, yielding that r(α) = 0 or s(α) = 0, since E is a field.
March 18, 2022 9:45 amsart-9x6 12819-main page 95

FIELDS AND EXTENSION FIELDS 95

Say r(α) = 0. Then r(x) ∈ ker(φα ) = hp(x)i. This is impossible


since 0 < deg(r(x)) < deg(p(x)). Hence p(x) is irreducible over F .
If q(x) ∈ F [x] is monic irreducible such that q(α) = 0. Then
q(x) ∈ ker(φα ) = hp(x)i. Thus p(x)|q(x). So p(x) = q(x). Part (1)
follows.
(2). Now the principal ideal hp(x)i consists precisely of those poly-
nomials of F [x] having α as a zero. If f (α) = 0 for f (x) ∈ F [x], then
f (x) ∈ ker(φα ) = hp(x)i. It follows that p(x)|f (x). 

Definition 4.2.6. Let F ≤ E be fields and α ∈ E be algebraic


over F. The unique monic polynomial p(x) with p(α) = 0 is called
the irreducible polynomial of α over F and will be denoted by
irr(α, F ). The degree of irr(α, F ) is the degree of α over F , denoted
by deg(α, F ).

The irreducible polynomial of α over F is also called minimal poly-


nomial of α over F in some textbooks.

Example √ 4.2.3. By the previous


√ examples, we
√ know that
(1). irr(√2, Q) = x2 − 2; irr( 2, R) = x − 2.
(2). irr( −1, Q) = x2 +p 1.

(3). We see that for α = 1 + 3 in R, α is a zero of x4 −2x2 −2 ∈
Q[x]. Since x4 − 2x2 − 2 is irreducible over Q (bypSchönemann-

Eisenstein Criterionpwith p = 2), we know that irr( 1 + 3, Q) =

x4 − 2x2 − 2. Thus 1 + 3 is algebraic of degree 4 over Q.

Let F ≤ E be fields and α ∈ E. Let φα : F [x] → E be the evalua-


tion homomorphism. By the previous results we should consider the
following two cases.
Case 1: α is algebraic over F . Then ker(φα ) = hirr(α, F )i which
is a maximal ideal of F [x]. Therefore, F [x]/hirr(α, F )i is a field and
is isomorphic to the image φα (F [x]) in E. This subfield φα (F [x]) of
E is then the smallest subfield of E containing F and α. Thus, in
this case we have φα (F [x]) = F [α] = F (α).
Case 2: α is transcendental over F . Then φα gives an iso-
morphism of F [x] with a subdomain F [α] of E. Thus in this case
φα (F [x]) is not a field but an integral domain. We see that E con-
tains a field of quotients of F [α], which is just the smallest subfield
F (α) of E containing F and α.
March 18, 2022 9:45 amsart-9x6 12819-main page 96

96 RING AND FIELD THEORY

Example 4.2.4. Since e is transcendental over Q, the field Q(e)


is isomorphic to the field Q(x) of rational functions over Q in the
indeterminate x.
The next theorem characterize algebraic elements over a field F .
Theorem 4.2.7. Let E = F (α) be an extension of F where α ∈ E is
algebraic over F . Let deg(α, F ) = n ≥ 1. As a vector space over F ,
then E has a basis {1, α, α2 , · · · , αn−1 }. Consequently deg(α, F ) =
[F (α) : F ].
Proof. Since α ∈ E is algebraic over F, we know that E = F (α) =
F [α]. Assume that
irr(α, F ) = p(x) = xn + an−1 xn−1 + · · · + a0
be the irreducible polynomial of α over F. For any element β ∈ E,
we can write β = f (α) for some polynomial f (x) ∈ F [x]. By the Eu-
clidean Algorithm, write f (x) = p(x)q(x) + r(x), where q(x), r(x) ∈
F [x] and r(x) = 0 or deg(r(x)) < deg(p(x)) = n. Thus we have
β = f (α) = p(α)q(α) + r(α) = r(α).
It follows that β can be expressed as
β = b0 + b1 α + · · · + bn−1 αn−1
with coefficients bi ∈ F.
For linear independence, if
c0 + c1 α + · · · + cn−1 αn−1 = 0
for ci ∈ F, then we have
g(x) = c0 + c1 x + · · · + cn−1 xn−1 ∈ F [x]
with g(α) = 0. Noticing that deg(g(x)) ≤ n − 1 < n = deg(p(x))
and p(x) is the irreducible polynomial of α over F , we must have
g(x) = 0. Therefore, ci = 0, so the linear independence of the αi is
established. 
We give an impressive example illustrating the theorem.
Example 4.2.5. Is p(x) = x2 + x + 1 ∈ Z2 [x] irreducible? Find a
field that has a zero of p(x).
Solution. Since p(0) = p(1) 6= 0, p(x) is irreducible over Z2 .
By Kronecker Theorem we know that the field E = Z2 [x]/hp(x)i
contains a zero α of x2 + x + 1. It is clear that |E| = 4. Also we
March 18, 2022 9:45 amsart-9x6 12819-main page 97

FIELDS AND EXTENSION FIELDS 97

know that the extension field Z2 (α) of Z2 contains the following four
elements
0 + 0α, 1 + 0α, 0 + 1α, 1 + 1α,
that is, 0, 1, α and 1 + α. This is a new finite field with four elements!
Moreover, the addition and multiplication tables for this field are
shown below. For example, to compute (1 + α)(1 + α) ∈ Z2 (α),
notice that p(α) = α2 + α + 1 = 0, we have
α2 = −α − 1 = α + 1.
Therefore,
(1 + α)(1 + α) = 1 + α + α + α2 = 1 + α2 = 1 + α + 1 = α.
+ | 0 1 α 1+α · | 0 1 α 1+α

0 | 0 1 α 1+α 0 | 0 0 0 0
1 | 1 0 1+α α 1 | 0 1 α 1+α
α | α 1+α 0 1 α | 0 α 1+α 1
1+α | 1+α α 1 0 1+α | 0 1+α 1 α


Let F be a field and f (x) ∈ F [x]. We know by Kronecker Theorem


that there exists an extension field E of F and an α ∈ E such that
α is a root of f (x). In order to discuss whether f (x) has a multiple
root, we need the following definition.
xi ∈ F [x].
P
Definition 4.2.8. Let F be a field and f (x) = i aiP
Define the derivative polynomial of f (x) to be f 0 (x) = i iai xi−1 .
Note that one must be careful to realize that multiplication by i
denotes multiplication by the image of i under the standard map from
Z to F . In particular, it may be zero. For instance, the derivative
polynomial of xp − 1 ∈ Zp [x] is zero.
The follow lemma is easy to prove or from calculus.
Lemma 4.2.9. Let f (x), g(x) ∈ F [x]. Then we have
[f (x)g(x)]0 = f 0 (x)g(x) + f (x)g 0 (x).
Theorem 4.2.10. Let f (x) ∈ F [x]. Then f (x) has a multiple root
in an extension field E if and only if gcd(f (x), f 0 (x)) 6= 1.
Proof. (⇒). Suppose that α is a multiple root of f (x) ∈ F [x] in
an extension field E. Then, there exists g(x) ∈ E[x] such that f (x) =
(x − α)2 g(x). Using the product rule in Lemma 4.2.9, we see that
f 0 (x) = 2(x − α)g(x) + (x − α)2 g 0 (x) and hence 1 6= gcd(f (x), f 0 (x)).
March 18, 2022 9:45 amsart-9x6 12819-main page 98

98 RING AND FIELD THEORY

(⇐). Suppose that gcd(f (x), f 0 (x)) 6= 1. Let d(x) = gcd(f (x),
f 0 (x)).
Then there exists an extension field E in which d(x) has a
root, say α. Since (x − α)|d(x) and hence f (x), there exists h(x) ∈
E[x] such that f (x) = (x − α)h(x). So
f 0 (x) = h(x) + (x − α)h0 (x).
Since (x − α) divides f 0 (x), it must also divide h(x). But then (x −
α)2 |f (x) in E[x] as required. 
4.3. Algebraic extensions and algebraic closure.
In this section we will prove that every field F has an extension E
such that every nonconstant polynomial in F [x] always has a root in
E. Such a minimal extension field of F is the algebraic closure of F .
We always assume that F is a field in this section.
Definition 4.3.1. Let F ≤ E be fields. Then E is called is an
algebraic extension of F if every element in E is algebraic over F .
We first have the following theorem which says that any finite
extension field is an algebraic extension.
Theorem 4.3.2. Let F ≤ E be fields with [E : F ] < ∞. Then E is
an algebraic extension of F.
Proof. Assume that [E : F ] = n. For any α ∈ E, then {1, α, · · · ,
αn } is linearly dependent over F , so there exist ai ∈ F , 0 ≤ i ≤ n,
such that
an αn + · · · + a1 α + a0 = 0,
and not all ai = 0. Then f (x) = an xn + · · · + a1 x + a0 ∈ F [x] is
nonzero, and f (α) = 0. Thus α is an algebraic element over F . 
Corollary 4.3.3. Let F ≤ E be fields and α ∈ E be algebraic over
F, and β ∈ F (α). Then deg(β, F )|deg(α, F ).
Proof. Note that deg(α, F ) = [F (α) : F ] and deg(β, F ) = [F (β) :
F ]. Since F ≤ F (β) ≤ F (α), from Theorem 4.1.4 we see that [F (β) :
F ] [F (α) : F ]. 
Example 4.3.1. Find a basis for Q(21/2 , 21/3 ) over Q, and show
that Q(21/2 , 21/3 ) = Q(21/6 ).
Solution. Since deg(21/2 , Q) = 2 and 26 | 3 = deg(21/3 , Q), we see
that 21/2 ∈
/ Q(21/3 ). Then x2 − 2 is irreducible over Q(21/3 ) and
[Q(21/3 , 21/2 ) : Q(21/3 )] = 2.
March 18, 2022 9:45 amsart-9x6 12819-main page 99

FIELDS AND EXTENSION FIELDS 99

So {1, 21/3 , 22/3 } is a basis for Q(21/3 ) over Q and {1, 21/2 } is a
basis for Q(21/3 , 21/2 ) over Q(21/3 ). Moreover, by Theorem 4.2.7,
{1, 21/2 , 21/3 , 25/6 , 22/3 , 27/6 } is a basis for Q(21/2 , 21/3 ) over Q.
Since 27/6 = 2(21/6 ), we see that 21/6 ∈ Q(21/2 , 21/3 ). Note that
21/6 is a zero of x6 − 2, which is irreducible over Q by Schönemann-
Eisenstein Criterion. Since
Q ≤ Q(21/6 ) ≤ Q(21/2 , 21/3 ),
we have
6 = [Q(21/2 , 21/3 ) : Q] = [Q(21/2 , 21/3 ) : Q(21/6 )][Q(21/6 ) : Q]
= 6[Q(21/2 , 21/3 ) : Q(21/6 )].
Thus, [Q(21/2 , 21/3 ) : Q(21/6 )] = 1, and Q(21/2 , 21/3 ) = Q(21/6 ). 
The previous example shows that it is possible for an extension
F (α1 , · · · , αn ) for n > 1 of a field F to be a simple extension.
Theorem 4.3.4. Let E be an algebraic extension of a field F . Then
there exist α1 , · · · , αn ∈ E such that E = F (α1 , · · · , αn ) if and only
if E is a finite extension of F .
Proof. (⇒). Suppose that E = F (α1 , · · · , αn ) for some ele-
ments αi ∈ E. Since E is an algebraic extension of F, each αi is
algebraic over F . So α1 is algebraic over F , and moreover, αj is
algebraic over F (α1 , · · · , αj−1 ) for j = 2, · · · , n. So [F (α1 , · · · , αj ) :
F (α1 , · · · , αj−1 )] < ∞. For the sequence of finite extensions
F ≤ F (α1 ) ≤ F (α1 , α2 ), · · · ≤ F (α1 , · · · , αn ) = E,
we know that E is a finite extension of F (Corollary 4.1.5).
(⇐). Suppose that E is a finite algebraic extension of F , i.e.,
[E : F ] = n < ∞. Take a basis {α1 , · · · , αn } of E over F . Clearly
F (α1 , · · · , αn ) = E.

We have observed that if E is an extension of a field F and α, β ∈ E
are algebraic over F, then so are α + β, αβ, α − β, and α/β if β 6= 0.
This follows also from the following theorem.
Theorem 4.3.5. Let F ≤ E be fields. Then
F E = {α ∈ E|α is algebraic over F } ≤ E,
called the algebraic closure of F in E.
March 18, 2022 9:45 amsart-9x6 12819-main page 100

100 RING AND FIELD THEORY

Proof. Let α, β ∈ F E . Then F (α, β) is a finite extension of F , and


every element of F (α, β) is algebraic over F , that is, F (α, β) ⊂ F E .
Thus F E contains α + β, αβ, α − β, and also contains α/β for β 6= 0,
so F E ≤ E. 
Corollary 4.3.6. The set of all algebraic numbers forms a field.
Proof. This corollary follows from the previous Theorem by tak-
ing F = Q and E = C. 
It is well known that the complex numbers have the property that
every nonconstant polynomial in C[x] has a zero in C. This is known
as the Fundamental Theorem of Algebra. We will give a proof for
this theorem in Theorem 6.2.1. We now give a name for such fields
in general.
Definition 4.3.7. Let F be a field. F is called algebraically closed
if every nonconstant polynomial in F [x] has a zero in F .
Theorem 4.3.8. A field F is algebraically closed if and only if every
nonconstant polynomial in F [x] is a product of degree one polynomials
in F [x].
Proof. (⇒). Let F be algebraically closed and f (x) ∈ F [x] with
deg(f (x)) = n ≥ 1. We prove this by induction on n. If n = 1
this is trivial. If n > 1, then f (x) has a zero α ∈ F. So x − α is a
factor of f (x), and f (x) = (x − a)g(x) for some g(x) ∈ F [x]. Note
that deg(g(x)) = n − 1. By inductive hypothesis g(x) is a product of
degree one polynomials in F [x]. Hence f (x) is a product of degree
one polynomials in F [x].
(⇐). Suppose that every nonconstant polynomial f (x) of F [x] can
be written as a product of linear factors, say
f (x) = (a1 x − b1 )(a2 x − b2 ) · · · (an x − bn ).
Then we have bi /ai , 1 ≤ i ≤ n, are all zeros of f (x). Thus F is
algebraically closed. 
Corollary 4.3.9. Let F be an algebraically closed field. Then there
is no algebraic extension E of F such that F < E.
Proof. Let E be an algebraic extension of F and α ∈ E. Since
F is algebraically closed, we have irr(α, F ) = x − α. Thus α ∈ F, it
follows that F = E. 
March 18, 2022 9:45 amsart-9x6 12819-main page 101

FIELDS AND EXTENSION FIELDS 101

If a field F has an algebraic extension F which is algebraically


closed, then F will certainly be a maximal algebraic extension of
F , since F is algebraically closed, it can have no proper algebraic
extensions. Such an extension of F is called an algebraic closure
of F .
The proof of the following lemma is left as Exercise (4).
Lemma 4.3.10. Let K ≤ F ≤ E be fields. If E is algebraic over F
and F is algebraic over K, then E is algebraic over K.
Theorem 4.3.11. Every field F has an algebraic closure F .
Proof. Let
P ={pj (x) ∈ F [x] : j ∈ J,
each pj (x) is monic and irreducible over F },
that is, P is the set of all irreducible monic polynomials in F [x].
Construct a polynomial ring R = F [xj : j ∈ J} in infinitely many
variables xj . Now let I = hpj (xj ) : j ∈ Ji  R.
First I 6= R. Otherwise if I = R then 1 ∈ I, so one can write
n
X
1= aj pj (xj ) (4.1)
j=1

for some ai ∈ R. Using Theorem 4.2.1 we can have an extension field


E of F containing all the roots for the polynomial p1 p2 · · · pn , and
choose a root αj ∈ E for each pj , 1 ≤ i ≤ n. We now consider the
evaluation map
φ :R → E,
ß
αj , if 1 ≤ j ≤ n
xj →
0, otherwise.
The Equation (4.1) becomes 1 = 0, which is nonsense. This proves
that I 6= R.
Consider S = {K  R : I ⊂ K} as a partially ordered set under the
set inclusion. Let C be a chain in S. One can easily show that ∪K∈C K
is an upper bound of C in S. By Kuratowski-Zorn Lemma there’s a
maximal ideal N of R with I ⊂ N ⊂ R. Define F1 = R/N and we
have the canonical map φF : F → F1 . Furthermore every polynomial
f (x) ∈ F [x] of degree 1 or more has a root in F1 ! We identify F with
φF (F ), that is, F is a subfield of F1 . Since pj (xj + N ) = 0 ∈ F1 =
R/N , we know that F1 is an algebraic extension of F (Lemma 4.3.10).
March 18, 2022 9:45 amsart-9x6 12819-main page 102

102 RING AND FIELD THEORY

It would be great if we were now done. Unfortunately we are


not yet there. The problem is that F1 has the property that every
polynomial in F [x] of positive degree has a root. However there may
be polynomials of positive degree in the larger ring F1 [x] that do not
have roots in F1 .
In this manner, if we start with F1 then we build a field F2 con-
taining F1 such that every element of F1 [x] of positive degree has a
root in F2 . Similarly, F2 is an algebraic extension of F1 and hence
of F (Lemma 4.3.10). And so on. We continue, getting an infinite
collection of algebraic extensions
F ⊂ F1 ⊂ · · · ⊂ Fk ⊂ · · · .
Now, we let F = ∪k∈N Fk . Then F is an algebraic extension field of F
(Lemma 4.3.10). Because any f ∈ F [x] of degree 1 or more will have
each coefficient in some Fk for k large enough, so f (x) has a root in
Fk+1 and hence in F . Then F will be an algebraically closed field,
and is an algebraic extension of F . Thus F is an algebraic closure
of F . 
Later in Corollary 5.2.3 we will show that each field F has a unique
algebraic closure F up to isomorphisms.

Definition 4.3.12. Let F ≤ E be fields. A subset S of a field E is al-


gebraically independent over F if the elements of S do not satisfy
any non-trivial polynomial equation with coefficients in F. A maximal
algebraically independent subset of E over F is called a transcen-
dence basis of E over F .

Using Kuratowski-Zorn Lemma one can show that transcendence


bases of E over F always exist and all have the same cardinality
which is called the transcendence degree of E over F . It is in-
teresting to know the following theorem that is named for Ferdinand
von Lindemann (1852–1939) and Karl Weierstrass (1815–1897). Lin-
demann proved in 1882 that ea is transcendental for every non-zero
algebraic number a, thereby establishing that π is transcendental.
Weierstrass proved the above more general statement in 1885. Here
we do not provide a proof which can be found in [J].

Theorem 4.3.13 (Lindemann-Weierstrass Theorem). If α1 , α2 , · · · ,


αn are algebraic numbers that are linearly independent over Q, then
eα1 , eα2 , · · · , eαn are algebraically independent over Q.
March 18, 2022 9:45 amsart-9x6 12819-main page 103

FIELDS AND EXTENSION FIELDS 103

Open Problem. Although both π and e are known to be transcen-


dental, it is not known whether π and e are algebraically independent
over Q. Even it is not known whether e − π or e + π is irrational.

Example 4.3.2. Prove that x2 − 3 is irreducible over Q( 3 2).

Solution. If x2 − 3 were reducible√ over
√ Q( 3 2), then it would
factor
√ into linear factors over √ Q( 3 2), so√ 3 would lie in the field
Q( 2), and we would have Q( 3) ≤ Q( 3 2). But then
3

√3

3
√ √
[Q( 2) : Q] = [Q( 2) : Q( 3][Q( 3) : Q].
√ √
This equation is impossible because [Q( 3 2) : Q] = 3 while √ [Q( 3) :
Q] = 2. This is impossible. So x2 − 3 is irreducible over Q( 3 2). 
√ √ √
Example
√ 4.3.3.
√ Let a, b ∈ Q. If a + b 6= 0, show that Q( a +

b) = Q( a, b).
Solution. If√a = √ √ is√clear; we assume a 6= b. It is
b the result
obvious that Q( a + b) ⊂ Q( a, b).
√ √ √ √
We now show that Q( a, b) ⊂ Q( a + b). Let α = √a−b √ ∈
a+ b
√ √ √ √ √ √
Q( a + b). Now α = a − b. Thus Q( a + b) contains 12 [α +
√ √ √ √ √ √ √
( a + b)] = a and hence also contains ( a + b) − a = b.
√ √ √ √
Thus Q( a, b) ⊂ Q( a + b). 
√ √
Example 4.3.4. Find a basis for Q( 2, 3) over Q.
√ √
Solution. Since 3 ∈ / Q( 2), we see that
√ √ √ √ √ √
[Q( 2, 3) : Q] = [Q( 2, 3) : Q( 3)][Q( 3) : Q] = 2 × 2 = 4.
√ √ √ √ √ √
Since Q( 2 + 3) √ = Q(
√ 2, 3) and 2 + 3 is a zero of x4 −
10x2 + 1, then irr( 2 + 3, Q) = x4 − 10x2 + 1. (This is a method √ to
show that this√polynomial
√ is√ irreducible.)
√ Consequently,
√ {1, 3} is
a basis
√ √ √ 2, 3) = (Q( 2))(
for Q( √ 3) √ over Q( 2). This shows that
{1, 2, 3, 6} is a basis for Q( 2, 3) over Q. 
4.4. Finite fields.
We shall now apply the established results in the extension field the-
ory to determine the structure of all finite fields. Observe that if F
is a finite field, then char(F ) = p is a prime and the prime field of
F can be identified with Zp . In the usual way we may regard F as
a vector space over Zp . Assume that [F : Zp ] = n, then we have a
March 18, 2022 9:45 amsart-9x6 12819-main page 104

104 RING AND FIELD THEORY

basis {α1 , · · · , αn } of F over Zp , every element on F can be written


in one and only one way as a linear combination
a1 α1 + a2 α2 + · · · + an αn ,
where ai ∈ Zp . Since each ai may be any of the p elements of Zp , the
total number of such distinct linear combinations of the αi is pn , that
is, |F | = pn . The same method shows that if F ≤ E, [E : F ] = n and
|F | = q then |E| = q n . Thus we have the basic facts on finite fields.
Theorem 4.4.1. Let F be a finite field with char(F ) = p.
(1). |F | = pn for some positive integer n.
(2). If F ≤ E, [E : F ] = n, and |F | = q, then |E| = q n .
We shall now consider the structure of the multiplicative group of
nonzero elements of a finite field.
Theorem 4.4.2. Let E be a finite field. Then the multiplicative
group (E ∗ , ·) of E is cyclic.
Proof. This follows from Corollary 1.6.5. 
Corollary 4.4.3. Any finite extension E of a finite field F is a
simple extension of F .
Proof. If α is a generator for the cyclic group E ∗ of nonzero
elements of E, we see that E = F (α). 
Lemma 4.4.4. Let F be a field with char(F ) = p. Then for all
α, β ∈ F and all n ∈ Z+ we have
n n n
(α + β)p = αp + β p .
Proof. Applying the binomial theorem to (α + β)p , we have
p(p − 1)
(α + β)p = αp + (p · 1)αp−1 β +
2
· 1α β + · · · + (p · 1)αβ p−1 + β p
p−2 2

= αp + 0αp−1 β + 0αp−2 β 2 + · · · + 0αβ p−1 + β p


= αp + β p .
n−1 n−1 n−1
By induction on n, suppose that we have (α+β)p = αp +β p .
Then
n n−1 n−1 n−1 n n
(α + β)p = [(α + β)p ]p = (αp + βp )p = αp + β p .

March 18, 2022 9:45 amsart-9x6 12819-main page 105

FIELDS AND EXTENSION FIELDS 105

Let Zp be an algebraic closure of the field Zp , n ∈ N, and


n
Fpn = {a ∈ Zp : ap = a}.
Theorem 4.4.5. Let p be a prime and n ∈ N. Then Fpn is a field
of order pn .
Proof. Using the previous lemma we can easily show that Fpn is a
subfield of Zp of order at most pn since elements in Fpn are solutions
n
of the polynomial f (x) = xp − x. From Theorem 4.2.10 and the
fact that gcd(f (x), f 0 (x)) = 1 we see that the number of solutions of
n
xp − x is pn . 
Thus, for each prime p we have the sequence of finite fields
Zp = Fp < Fp2 < Fp3 < · · · < Fpr < · · · < Zp .
This tells us that F p = ∪∞
k=1 Fpk . See Exercise (20).
Definition 4.4.6. Let E be a field and α ∈ E. Then α is called an
n-th root of unity if αn = 1. It is called a primitive n-th root
of unity if αn = 1 and αm 6= 1 for 0 < m < n.
We know that the nonzero elements of a finite field of pn elements
are all (pn − 1)-th roots of unity.
Example 4.4.1. Find the generators for (Z∗11 , ·), i.e., all primitive
10th roots of unity in the field (Z11 , +, ·), and all primitive 5th roots
of unity.
Solution. Consider the finite field Z11 . We know that (Z∗11 , ·) is
cyclic of order 10. Let us find all generators of Z∗11 . Noticing that
|Z∗11 | = 10, we see that ord(2)|10. It follows that ord(2) = 2, 5, or
10. Since
22 = 4, 24 = 42 = 5, and 25 = (2)(5) = 10 = −1,
so ord(2) = 10, and Z∗11 = h2i, that is, 2 is a primitive 10th root of
unity in Z11 .
10
By the theory of cyclic groups, we know that ord(2n ) = gcd(n,10) .
∗ n
Then all the generators of Z11 , are of the form 2 , where gcd(n, 10) =
1. Thus, these elements are
21 = 2, 23 = 8, 27 = 7, 29 = 6.
The primitive 5th roots of unity in Z11 are of the form 2m , where
gcd(m, 10) = 2, that is,
22 = 4, 24 = 5, 26 = 9, 28 = 3.
The primitive square root of unity in Z11 is 25 = 10 = −1. 
March 18, 2022 9:45 amsart-9x6 12819-main page 106

106 RING AND FIELD THEORY

Corollary 4.4.7. Let F be a finite field. Then for every positive


integer n, there is an irreducible polynomial in F [x] of degree n.
Proof. Let char(F ) = p, |F | = pr and F be the algebraic closure
of F . Then there is a field K ≤ F consisting precisely of the prn
rn r
zeros of xp − x. Notice that all elements in F are zeros of xp − x,
by Theorem 4.4.5 we know that F ≤ K and [K : F ] = n. Since K
is simple over F (Corollary 4.4.3), so K = F (β) for some β ∈ K.
Hence, the irreducible polynomial irr(β, F ) is of degree n = [K : F ]
(Theorem 4.2.7). 
Theorem 4.4.8. Let p be a prime and let n ∈ N. If E and E 0 are
fields of order pn , then E ∼
= E0.
Proof. We may regard both E and E 0 as extensions of the prime
field Zp up to isomorphism. Then E is a simple extension of Zp of
degree n, say E = Zp (α). Let f (x) = irr(α, Zp ) which has degree n.
By considering the evaluation homomorphism φα : Zp [x] → E we see
that E ∼
n
= Zp [x]/hf (x)i. Because elements of E are zeros of xp − x,
n
we see that f (x)|xp −x in Zp [x]. Because E 0 also consists of all zeros
n
of xp −x, then E 0 also contains zeros of irreducible f (x) ∈ Zp [x], say
α0 ∈ E 0 with f (α0 ) = 0. Thus E 0 = Zp (α0 ) since they have the same
size. By considering the evaluation homomorphism φα0 : Zp [x] → E 0
we see that E 0 ∼ = Zp [x]/hf (x)i since ker(φα0 ) = hf (x)i. Therefore
0 ∼
E = E. 
As a classical application, we will use Gaussian integers and prop-
erties of finite fields to prove the following result in Number Theory.
Theorem 4.4.9. Let p be an odd prime in Z. Then p = a2 + b2 for
some a, b ∈ Z if and only if p ≡ 1 (mod 4).
Proof. (⇒). Suppose that p = a2 + b2 . Since p is odd, then a and
b cannot be both even or both odd. We may assume that a = 2r and
b = 2s + 1, then a2 + b2 = 4r2 + 4(s2 + s) + 1, so p ≡ 1 (mod 4).
(⇐). Assume that p ≡ 1 (mod 4). From Theorem 4.4.2 we know
that the multiplicative group of nonzero elements of the finite field
Zp is cyclic of order p − 1. Since 4|p − 1, we see that the group (Z∗p , ·)
contains an element n of order 4. Then n2 has multiplicative order
2. So n2 = −1 in Zp . Thus n2 ≡ −1 (mod p), so p|n2 + 1 in Z.
Next we work within in the Euclidean domain Z[i] (Theorem 2.5.4).
We see that p|n2 + 1 = (n + i)(n − i) in Z[i]. If p is irreducible in Z[i],
March 18, 2022 9:45 amsart-9x6 12819-main page 107

FIELDS AND EXTENSION FIELDS 107

then p|n + i or p|n − i, which is impossible. Thus p is not irreducible


in Z[i].
Let p = (a + bi)(c + di) where neither a + bi nor c + di is a
unit. Then a2 + b2 6= 1 nor c2 + d2 6= 1. Taking norms, we have
p2 = (a2 + b2 )(c2 + d2 ). So p = a2 + b2 = c2 + d2 , which completes
our proof. 
To conclude this chapter we prove the following beautiful result in
Number Theory using Gaussian integers Z[i].
Theorem 4.4.10. Let n = pn1 1 pn2 2 · · · pnk k ∈ 1+N where p1 , p2 , · · · , pk
are pairwise distinct primes with each nj ∈ N. Then n = a2 + b2 for
some a, b ∈ Z, if and only if, pj ≡ 3 (mod 4) implies nj is even.
Proof. (⇒). We have n = a2 + b2 = (a + ib)(a − ib) = N (a + ib).
Let z = a + ib ∈ Z[i]. Write z = α1 · · · αq as a product of irreducibles
in Z[i]. By Exercise (34), we see that
(a). αj = ±1 ± i,
(b). αj = p ≡ 3 (mod 4),
(c). N (αj ) = p ≡ 1 (mod 4),
where p is a prime number in Z. We now take the norm for z = a+ib
to obtain
n = a2 + b2 = N (z) = N (α1 )N (α2 ) · · · N (αq ).
If αj = ±1 ± i, we know that N (αj ) = 2; if αj = p ≡ 3 (mod 4), we
know that N (αj ) = p2 ; the remaining case for αj is N (αj ) = p ≡ 1
(mod 4). So pj ≡ 3 (mod 4) implies nj is even.
(⇐). Let n = pn1 1 pn2 2 · · · pnk k be a product of distinct primes such
that pj ≡ 3 (mod 4) implies nj is even. By Theorem 4.4.9, we know
n
that each pj j is a sum of integer squares. From Exercise (32), we see
that n = a2 + b2 for some a, b ∈ Z. 
Example 4.4.2. Can you write the integer 330000 as a sum of two
integer squares?
Solution. Note that 330000 = 3 · 11 · 1002 . Since the power 3 is
odd, by Theorem 4.4.10 we cannot write the integer 330000 as a sum
of two integer squares. 
For n ∈ N with n = a2 + b2 for some a, b ∈ Z, note that n may
have more than one expressions as a sum of two integer squares. For
example,
65 = 82 + 12 = 42 + 72 .
March 18, 2022 9:45 amsart-9x6 12819-main page 108

108 RING AND FIELD THEORY

4.5. Exercises.
√ √ √
(1) Calculate the irreducible polynomial of 3 2 and 1 + 3 2 + 3 4
over Q.
(2) Show that Q[x]/hx2 − 1i is not an integral domain, but Q[x]/
hx2 + 1i is.
(3) Use Kuratowski-Zorn Lemma to show that every proper ideal
of a ring R with unity is contained in some maximal ideal.
(4) Prove Lemma 4.3.10.
(5) Let E be a finite extension field of a field F . Let D be an
integral domain such that F ⊂ D ⊂ E. Show that D is a
field.
(6) Find√ the
√ degree
√ and a basis√ for√ the given field extensions:
Q( 3 2, 3 6, 3 24) over Q, Q( 3 2, 3) over Q.
(7) Let E be an extension field of a field F and [E : F ] be a
prime. For any α ∈ E \ F, show that E = F (α).
(8) Let E be an extension field of F and α ∈ E be algebraic of
odd degree over F . Show that α2 is algebraic of odd degree
over F and F (α) = F (α2 ).
(9) Let f (x) be an irreducible polynomial in Zp [x]. Show that
n
f (x) is a divisor of xp − x for some n.
(10) Let c ∈ F, where F is a field of characteristic p > 0. Prove
that xp − x − c is irreducible in F [x] if and only if xp − x − c
has no root in F. Show this is false if F is of characteristic 0.
(11) Let f (x) = xp − x − c where p is a prime not dividing c ∈ Z.
Show that f (x) is irreducible over Q. (Hint: try the following
steps: Show that if f (x) is irreducible over Zp , then it is
irreducible over Q. Show that f (x) does not have a root in
Zp . Then consider the previous exercise.)
(12) Under what conditions on q is the polynomial x2 + x + 1 irre-
ducible over a finite field F with q elements? (Hint: consider
the multiplicative group of nonzero elements of F .)
(13) Find the conditions on a ∈ C such that x5 − 5x + a = 0 has
multiple roots.
(14) Let F be an algebraically closed field. Find conditions on
a ∈ F such that the equation x5 + 5ax + 4a = 0 has no
multiple roots in F .
(15) Let p be a prime. Show that a finite field of pn elements has
exactly one subfield of pm elements for each divisor m of n.
(16) Find a primitive root of unity of order 6 in F7 .
March 18, 2022 9:45 amsart-9x6 12819-main page 109

FIELDS AND EXTENSION FIELDS 109

(17) Find the number of primitive root of unity of order 48 in F72 .


(18) Find the number of primitive roots of unity of order 24 in
F72 .
(19) Let p be a prime, m ∈ N such that p6 |m. Show that there is
n ∈ N such that m|pn − 1.
(20) Let p be a prime, Fp < Fp2 < Fp3 < · · · < Fpr < · · · . Show
that F p = ∪∞ k=1 Fpk .
(21) Let F be a field consisting of 1311 elements. Find the number
of subfields of F .
(22) Let F be a field consisting of 1311 elements and α ∈ F . Find
deg(α, Z13 ) where Z13 is the prime subfield of F .
(23) Let F be a field consisting of 1311 elements. How many dis-
tinct irreducible polynomials of degree 11 over Z13 .
(24) Let p be a prime, F be the extension field of Zp of degree
n
n, consisting of all zeros of xp − x. For each d|n let fd,1 (x),
fd,2 (x), · · · , fd,rd (x) be all the distinct irreducible monic poly-
nomials of degree d in Z[x]. Show that
rd
n
YY
xp − x = fd,j (x).
d|n j=1

(25) Let F be the field Z2 [x]/hx6 + x + 1i of order 26 = 64.


(a). Calculate x7 , x9 , x14 (= (x7 )2 ) and x21 (= (x7 )3 ) in F as
Z2 -linear combinations of 1, x, · · · , x5 .
(b). Find the order of x in the group U(F ) of units of F .
(26) There are two irreducible cubics x3 + x + 1 and x3 + x2 + 1
in Z2 [x]. Show that Z2 [x]/hx3 + x + 1i ∼ = Z2 [x]/hx3 + x2 + 1i.
(27) Let p be a prime, and let f (x), g(x) ∈ Zp [x] be irreducible
polynomials of order 6. Show that Zp [x]/hf (x)i ∼ = Zp [x]/
hg(x)i.
(28) Find the number of different subfields of F78 .
(29) Let p be a prime. Let E be a field extension of F of degree
p. If a ∈ E \ F , prove that the irreducible polynomial of a
over F has degree p.
(30) Prove that e2 − 2 is algebraic over Q(e3 ). Find irr(e2 −
2, Q(e3 )).
March 18, 2022 9:45 amsart-9x6 12819-main page 110

110 RING AND FIELD THEORY

(31) Let F be a finite field with ch(F ) 6= 2. For any k ∈ Z+ show


that X
rk = 0,
r∈F
where 00 = 1. (Hint: Use Theorem 4.4.2.)
(32) Can you easily show that (a2 + b2 )(c2 + d2 ) = (ac − bd)2 +
(ad + bc)2 for any a, b, c, d ∈ Z without expanding both sides?
(33) If p ∈ N is prime with p ≡ 3 (mod 4), show that p is irre-
ducible in the Gaussian integers Z[i].
(34) Let a+bi ∈ Z[i] where a, b ∈ Z. Show that a+bi is irreducible
in Z[i] if and only if one of the following holds:
(a). a + bi = ±1 ± i,
(b). b = 0 and a ≡ 3 (mod 4) is a prime, or
(c). a2 + b2 ≡ 3 (mod 4) is a prime.
(35) Can you write the integer 690000 as a sum of two integer
squares?
(36) Can you write the integer 1146600 as a sum of two integer
squares?
(37) Show that Z[x2 ] ∩ Z[x2 − x] = Z in Z[x].
(38) Using Theorem 4.3.13 show that π is a transcendental num-
ber.
(39) If α is a non-zero algebraic number, using Theorem 4.3.13
show that sin(α), cos(α), tan(α) are transcendental numbers.
(Note that sin2 (α) + cos2 (α) = 1.)
(40) If α 6= 1 is a positive algebraic number, show that ln(α) is a
transcendental number.
March 18, 2022 9:45 amsart-9x6 12819-main page 111

5. Automorphisms of Fields

This chapter is a preparation for Galois Theorem. We will study


automorphism groups for various different extension fields. In this
chapter we always assume that E, F are fields, F is the algebraic
closure of F .
5.1. Automorphisms.
Definition 5.1.1. Let F ≤ E be fields, and α, β ∈ E be algebraic
over F . We say that α, β ∈ E to be conjugate over F if irr(α, F ) =
irr(β, F ).
Example 5.1.1. If a, b ∈ R with b 6= 0, the conjugate complex num-
bers a + bi and a − bi are both zeros of x2 − 2ax + a2 + b2 , which is
irreducible in R[x].
Theorem 5.1.2. Let α, β ∈ F be algebraic over F with deg(α, F ) =
n. The map ψα,β : F (α) → F (β) defined by
ψα,β (c0 + c1 α + · · · + cn−1 αn−1 ) = c0 + c1 β + · · · + cn−1 β n−1 , ∀ci ∈ F
is an isomorphism (called conjugation isomorphism) if and only
if α and β are conjugate over F .
Proof. (⇒). Let irr(α, F ) = a0 + a1 x + · · · + an xn . Then a0 +
a1 α + · · · + an αn = 0, so
ψα,β (a0 + a1 α + · · · + an αn ) = a0 + a1 β + · · · + an β n = 0.
This implies that irr(β, F ) irr(α, F ). Since both polynomials are monic
and irreducible, then irr(α, F ) = irr(β, F ), so α and β are conjugate
over F .
(⇐). Suppose irr(α, F ) = irr(β, F ) = p(x). Then both the evalu-
ation homomorphisms φα : F [x] → F (α) and φβ : F [x] → F (β) have
the same kernel hp(x)i.
φβ φα
F (β) ←− F [x] −→ F (α)

ψβ ψα
F (β)  F [x]/hp(x)i  F (α)
111
March 18, 2022 9:45 amsart-9x6 12819-main page 112

112 RING AND FIELD THEORY

We have the natural isomorphisms ψα : F [x] → F (α) and ψβ :


F [x] → F (β). Let ψα,β = ψβ ψα−1 . Clearly, ψα,β is an isomorphism
mapping F (α) onto F (β). For c0 + c1 α + · · · + cn−1 αn−1 ∈ F (α), we
have
ψα,β (c0 + c1 α + · · · + cn−1 αn−1 ) = ψβ ψα−1 (c0 + c1 α + · · · + cn−1 αn−1 )
= ψβ ((c0 +c1 x+· · ·+cn−1 xn−1 )+hp(x)i) = c0 +c1 β +· · ·+cn−1 β n−1 .
Thus ψα,β is the isomorphism defined in the statement of the the-
orem. 
Corollary 5.1.3. Let α ∈ F be algebraic over F , and ψ be an iso-
morphism mapping F (α) onto a subfield of F such that ψ(a) = a for
a ∈ F . Then β = φ(α) is a conjugate of α over F .
Proof. Let irr(α, F ) = a0 + a1 x + · · · + an xn . Then
a0 + a1 α + · · · + an αn = 0,
0 = ψ(a0 + a1 α + · · · + an αn ) = a0 + a1 ψ(α) + · · · + an ψ(α)n ,
and β = ψ(α) is a conjugate of α. 
A special case of the above corollary is a familiar result.
Corollary 5.1.4. Let f (x) ∈ R[x]. If f (a + bi) = 0 for a + bi ∈ C,
where a, b ∈ R, then f (a − bi) = 0 also.
Proof. We have seen that C = R(i). Now irr(i, R) = irr(−i, R) =
x2 + 1, so i and −i are conjugate over R. By the previous Theorem,
the conjugation map ψ = ψi,−i : C → C where ψ(a + bi) = a − bi is
an isomorphism. Since f (x) ∈ R[x], applying ψ to f (a + bi) = 0 we
obtain that f (a − bi) = 0. 
√ √
Example 5.1.2.
2
√ Consider Q( 2) over Q. The zeros of irr( 2, Q) =
x − 2 are ± 2, so they √are conjugate√ over Q. The conjugation
isomorphism ψ√2,−√2 : Q( 2) → Q( 2) defined by
√ √
ψ√2,−√2 (a + b 2) = a − b 2

is an automorphism of Q( 2).
As illustrated in the preceding corollary and example, a field may
have a nontrivial isomorphism onto itself.
Definition 5.1.5. An isomorphism of a field F onto itself is an
automorphism of the field of F . We denote the set of all auto-
morphisms of F by Aut(F ).
March 18, 2022 9:45 amsart-9x6 12819-main page 113

AUTOMORPHISMS OF FIELDS 113

Definition 5.1.6. Let E ≤ K be fields, σ an isomorphism of E onto


a subfield of K. Then a ∈ E is left fixed by σ if σ(a) = a. A
collection S of isomorphisms of E leaves a subfield F of E fixed
if
sigma(a) = a, ∀a ∈ F, σ ∈ S.
If {σ} leaves F fixed, then σ leaves F fixed.
√ √
Example 5.1.3. Let E = Q( 2, 3). The map σ : E → E defined
by √ √ √ √ √ √
σ(a + b 2 + c 3 + d 6) = a + b 2 − c 3 − d 6
for a, b, c, d ∈ Q is an automorphism of E; it is the conjugation
√ √ iso-
morphism ψ√3,−√3 of E onto itself if we view E as (Q( 2))( 3). We

see that σ leaves Q( 2) fixed.
Theorem 5.1.7. Let H = {σk |k ∈ I} ⊆ Aut(E) where E is a field.
Then the set EH = {a ∈ E : σk (a) = a ∀k ∈ I} is a subfield of E.
Proof. Let a, b ∈ EH , i.e., σk (a) = a and σk (b) = b for all k ∈ I.
Then
σk (a ± b) = σk (a) ± σk (b) = a ± b,
σk (ab) = σk (a)σk (b) = ab, ∀k ∈ I,
i.e., a ± b, ab ∈ EH . Also, if b 6= 0, then
σk (a/b) = σk (a)/σk (b) = a/b
for all k ∈ I, i.e., a/b ∈ EH . Since the σk are automorphisms,
σk (0) = 0, σk (1) = 1 for all k ∈ I, i.e., 0, 1 ∈ EH . Thus EH ≤ E. 
Definition 5.1.8. The field EH in the above Theorem is the fixed
field of H. For a single automorphism σ, we shall refer to Eσ as the
fixed field of σ.
Example 5.1.4. Consider the conjugation automorphism ψ√2,−√2

of Q( 2) given in the previous example. For a, b ∈ Q, we have
√ √
ψ√2,−√2 (a + b 2) = a − b 2,
√ √
and a − b 2 = a + b 2 if and only if b = 0. Thus the fixed field of
ψ√2,−√2 is Q.
Theorem 5.1.9. The set Aut(E) is a subgroup of the symmetric
group (SE , ◦).
March 18, 2022 9:45 amsart-9x6 12819-main page 114

114 RING AND FIELD THEORY

Proof. We know that (SE , ◦) is a group, where the multiplication


is map composition. The identity map ι : E → E is in Aut(E). For
σ, τ ∈ Aut(E), it is easy to see that στ, σ −1 ∈ Aut(E). Thus Aut(E)
is a subgroup of SE . 
Theorem 5.1.10. Let E be a field, and let F be a subfield of E.
Then the set
Gal(E/F ) = {σ ∈ Aut(E) : σ(a) = a ∀a ∈ F }
is a subgroup of Aut(E). Furthermore, F ≤ EGal(E/F ) .
Proof. For σ, τ ∈ Gal(E/F ) and a ∈ F, we have
(στ )(a) = σ(τ (a)) = σ(a) = a,
so στ ∈ Gal(E/F ). Of course, the identity automorphism ι is in
Gal(E/F ). Also, if σ(a) = a for a ∈ F, then a = σ −1 (a) so
σ ∈ Gal(E/F ) implies that σ −1 ∈ Gal(E/F ). Thus Gal(E/F ) is
a subgroup of Aut(E).
By definition we know that σ(a) = a for a ∈ F and σ ∈ Gal(E/F ).
So F ≤ EGal(E/F ) . 
Definition 5.1.11. The group Gal(E/F ) of the preceding theorem
is called the Galois group of E over F .
√ √
Example 5.1.5. Find Gal(Q( 2, 3), Q).
√ √ √
√ Solution.
√ We√know that [Q( 2, 3) : Q] = 4. If we view Q( 2,
√ √
3) as (Q( 3))( 2), the conjugation isomorphism ψ 2,− 2 defined
by √ √
ψ√2,−√2 (a + b 2) = a − b 2
√ √ √ √
for a, b ∈ Q( 3) is an automorphism of Q( 2, 3) having Q( 3) as
fixed field. Similarly, we have the automorphism ψ√3,−√3
√ √ √
of Q( 2, 3) having Q( 2)√as fixed √ field. The automorphisms
ψ√2,−√2 ψ√3,−√3 moves both 2 and 3, that is, leaves neither num-
ber fixed. Let
ι = the identity automorphism,
σ1 = ψ√2,−√2 ,
σ2 = ψ√3,−√3 ,
σ3 = ψ√2,−√2 ψ√3,−√3 .
March 18, 2022 9:45 amsart-9x6 12819-main page 115

AUTOMORPHISMS OF FIELDS 115


√ √
The group of all automorphisms of Q( 2, 3) has a fixed field.
This fixed field must contain Q, since every automorphism of √a field

leaves 1 and hence √ √ √ the prime subfield fixed.
√ A basis
√ for√ Q( 2, √3)
over Q is √ {1, 2, √ 3, 6}. Since σ 1 ( 2) = − 2, σ1 ( 6) = − 6
and σ2 ( 3) = − 3, we see that Q is exactly the fixed field of
{ι, σ1 , σ2 , σ3 }. It is readily checked that G = {ι, σ1 , σ2 , σ3 } is a group
under automorphism multiplication (function composition). For ex-
ample,

σ1 σ3 = ψ√2,−√2 (ψ√2,−√2 ψ√3,−√3 ) = ψ√3,−√3 = σ2 .

The group G is actually


√ isomorphic
√ to (Z2 , +) × (Z2 , +). We can
show √that√G = Gal(Q( √ 2, 3)/Q), because
√ every automorphism
√ τ
of Q( 2,√ 3) maps 2 onto either
√ √ √ √ ± 2. Similarly, τ maps
√ √ 3 onto
either ± 3. But since 1, 2, √ √ 3, 2 3 is a basis for Q( 2, 3) over
Q, an automorphism
√ of√Q( 2, 3) leaving Q fixed is determined by
its values on 2 and 3. √ Now, ι, σ√1 , σ2 , and σ3 give all possible
combinations of values √ √ 2 and 3, and hence are all possible
on
automorphisms of Q(√ 2,√ 3). √ √
Note that Gal(Q( 2, 3)/Q) has order 4, and [Q( 2, 3) : Q] =
4. This holds for a general situation, as we shall prove later. 
For any finite field F we shall show later that the group Aut(F )
is cyclic. Actually the group Aut(F ) has a canonical generator, the
Frobenius automorphism given by the next theorem.

Theorem 5.1.12. Let F be a finite field of characteristic p. Then


the map σp : F → F defined by σp (a) = ap for a ∈ F is an automor-
phism, the Frobenius automorphism, called of F . Also, Fσp ' Zp .

Proof. Let a, b ∈ F. We see that (a + b)p = ap + bp . So

σp (a + b) = (a + b)p = ap + bp = σp (a) + σp (b).

σp (ab) = (ab)p = ap bp = σp (a)σp (b),


i.e., σp is at least a homomorphism. If σp (a) = 0, then ap = 0, and
a = 0, i.e., ker(σp ) = {0}. Then σp is a one-to-one map. Since F is
finite, σp is also onto. Thus σp is an automorphism of F .
From Theorem 4.4.5 we know that Zp = {a ∈ F : ap = a}, and
Fσp = {a ∈ F : σp (a) = a}. We see that Zp = Fσp . 
March 18, 2022 9:45 amsart-9x6 12819-main page 116

116 RING AND FIELD THEORY

5.2. The isomorphism extension theorem.


Remember that we always assume that E, F are fields, and F is an
algebraic closure of F .
Theorem 5.2.1 (Isomorphism Extension Theorem). Let E be an
algebraic extension of a field F . Let σ be an isomorphism of F onto
a field F 0 . Then σ can be extended to an isomorphism τ of E onto a
subfield of F 0 , i.e., τ |F = σ.
σ
F  F0
∩ ∩
∃τ
E −→ F 0
Proof. Let S be the set of all pairs (L, λ), where L is a field such
that F ≤ L ≤ E and λ is an isomorphism of L onto a subfield of F 0
such that λ(a) = σ(a) for a ∈ F. Clearly, (F, σ) ∈ S. Define a partial
ordering on S by (L1 , λ1 ) ≤ (L2 , λ2 ), if L1 ≤ L2 and λ1 (a) = λ2 (a)
for a ∈ L1 . This relation ≤ gives a partial ordering of S.
Let T = {(Hi , λi ) : i ∈ I} be a chain in S. We claim that H =
∪i∈I Hi is a subfield of E. Let a, b ∈ H, where a ∈ H1 and b ∈ H2 ;
then either H1 ≤ H2 or H2 ≤ H1 , since T is a chain. Say, H1 ≤ H2 .
Then a, b ∈ H2 , so a ± b, ab, and a/b for b 6= 0 are all in H2 and hence
in H. Since for each i ∈ I, F ⊂ Hi ⊂ E, we have F ⊂ H ⊂ E. Thus
H ≤ E.
Define λ : H → F 0 as follows. Let c ∈ H. Then c ∈ Hi for some
i ∈ I, and let λ(c) = λi (c). The map λ is well defined because if
c ∈ H1 and c ∈ H2 , then either (H1 , λ1 ) ≤ (H2 , λ2 ) or (H2 , λ2 ) ≤
(H1 , λ1 ), since T is a chain. In either case, λ1 (c) = λ2 (c). We claim
that λ is an isomorphism of H onto a subfield of F 0 . If a, b ∈ H then
there is an Hi such that a, b ∈ Hi , and
λ(a + b) = λi (a + b) = λi (a) + λi (b) = λ(a) + λ(b).
Similarly,
λ(ab) = λi (ab) = λi (a)λi (b) = λ(a)λ(b).
If λ(a) = 0, then a ∈ Hi for some i implies that λi (a) = 0, so a = 0.
Therefore, λ is an isomorphism. Thus (H, λ) ∈ S, and it is clear
from our definitions of H and λ that (H, λ) is an upper bound for T .
We have shown that every chain of S has an upper bound in S.
So the hypotheses of Kuratowski-Zorn lemma are satisfied. Hence
there exists a maximal element (K, τ ) ∈ S. Let τ (K) = K 0 , where
March 18, 2022 9:45 amsart-9x6 12819-main page 117

AUTOMORPHISMS OF FIELDS 117

K 0 ≤ F 0 . If K 6= E, take α ∈ E \ K. Now α is algebraic over F ,


so α is algebraic over K. Also, let p(x) = irr(α, K). Extending the
isomorphism τ we have the ring isomorphism τx : K[x] → K 0 [x] by
τx (a0 + a1 x + · · · + an xn ) = τ (a0 ) + τ (a1 )x + · · · + τ (an )xn , ∀ai ∈ K.
Let q(x) = τx (p(x)). Since τx is an isomorphism, q(x) is irreducible
in K 0 [x], and we have the induced isomorphism
τ : K[x]/hp(x)i  K 0 [x]/hq(x)i
from τx . Let ψα be the canonical isomorphism
ψα : K[x]/hp(x)i  K(α),
corresponding to the evaluation homomorphism φα : K[x] → K(α).
Since K 0 ≤ F 0 , there is a zero α0 of q(x) in F 0 . Let
ψα0 : K 0 [x]/hq(x)i  K 0 (α0 )
be the isomorphism analogous to ψα .
φα τx φ 0
K(α) ←− K[x]  K 0 [x] α
−→ K 0 (α0 )

ψα τ ψα0
K(α)  K[x]/hp(x)i  K 0 [x]/hq(x)i  K 0 (α0 )
Then the composition of maps
ψα0 τ ψα−1 : K(α) → K 0 (α0 )
is an isomorphism of K(α) onto a subfield of F 0 . Since
ψα0 τ ψα−1 (k) = ψα0 τ (k) = ψα0 τ (k) = τ (k), ∀k ∈ K,
so, (K, τ ) < (K(α), ψα0 τ ψα−1 ) which contradicts that (K, τ ) is maxi-
mal. Therefore we must have had K = E. 

Example 5.2.1. We have the automorphism ψ√2,−√2 : Q( 2) →

Q( 2). Can we extend ψ√2,−√2 to an automorphism of R? (No.
See Example 1.2.1.) Can we extend ψ√2,−√2 to an automorphism of
C? (Yes, a lot. But at this moment we cannot prove this.)
We give as a corollary the existence of an extension of one of our
conjugation isomorphisms ψα,β , as discussed at the beginning of this
section.
March 18, 2022 9:45 amsart-9x6 12819-main page 118

118 RING AND FIELD THEORY

Corollary 5.2.2. If F ≤ E ≤ F where E is algebraic over F , and


α, β ∈ E are conjugate over F , then the conjugation isomorphism
ψα,β : F (α) → F (β) can be extended to an isomorphism of E onto a
subfield of F .
Proof. Proof of this corollary is immediate from Theorem 5.2.1 if
in the statement of the theorem we replace F by F (α), F 0 by F (β),
and F 0 by F . 
Corollary 5.2.3. Let σ : F → F 0 be an isomorphism of the fields.
Then σ can be extended to an isomorphism σ : F → F 0 .
Proof. By Theorem 5.2.1, σ can be extended to an isomorphism
τ : F onto a subfield of F 0 . We need only show that τ is onto F 0 .
But the map τ −1 : τ (F ) → F can be extended to an isomorphism of
F 0 onto a subfield of F . Since τ −1 is already onto F , we must have
τ (F ) = F 0 . 
As a direct consequence of this corollary, we see that an algebraic
closure of F is unique, up to an isomorphism over F , i.e.,
0
Corollary 5.2.4. Let F and F be two algebraic closures of F . Then
there is an isomorphism τ : F → F 0 such that τ (a) = a for any
a ∈ F.
Theorem 5.2.5. Let E be a finite extension of a field F , and σ :
F → F 0 be an isomorphism of fields. Then the number of extensions
of σ to an isomorphism τ of E onto a subfield of F 0 is finite, and
independent of F 0 , F 0 , and σ. That is, the number of extensions is
completely determined by the two fields E and F .
Proof. By Corollary 5.2.3 we can extend σ −1 : F 0 → F to an
isomorphism σ : F 0 → F . Each isomorphism σ̃ of E onto a subfield
of F 0 that extends σ, one-to-one corresponds an isomorphism σσ̃ :
E → F . Indeed, if isomorphisms σ̃, σ̃ 0 of E onto a subfield of F 0
extends σ, then σσ̃ = σσ̃ 0 iff σ̃ = σ̃ 0 .
σ σ −1
F  F0  F F = F
∩ ∩ ∩ ∩ ∩
σ ?∃τ
?∃σ̃
E −−→ F 0  F E −−→ F
In order to prove the theorem we may assume that F 0 = F and σ =
idF . Since E is a finite extension of F , we take a basis of E over F :
β = {a1 , a2 , · · · , an }. Let pi (x) = irr(ai , F ). Let τ : E → F be any
March 18, 2022 9:45 amsart-9x6 12819-main page 119

AUTOMORPHISMS OF FIELDS 119

isomorphism from E onto a subfield of F with τ (a) = a for any a ∈ F .


We know that τ is uniquely determined by τ (a1 ), τ (a2 ), · · · , τ (an ) ∈
F.
Applying τ to pi (ai ) = 0, we see that pi (τ (ai )) = 0, that is ai and
τ (ai ) are zeros of pi (x). Each pi (x) has only finitely many zeros. We
conclude that we have only finitely may τ . This completes the proof.

Definition 5.2.6. Let E be a finite extension of a field F . Let
S(E/F ) be the set of all isomorphisms of E onto a subfield of F
over F . Then we call |S(E/F )| the index of E over F , denoted by
{E : F }.
You may compare the two sets Gal(E/F ) and S(E/F ).
Corollary 5.2.7. If F ≤ E ≤ K, where K is a finite extension field
of F , then {K : F } = {K : E}{E : F }.
Proof.
F ≤ E ≤ K
|| σ↓ τ↓
F ≤ F = F
By Theorem 5.2.1, each of the {E : F } isomorphisms σ of E onto a
subfield of F over F has {K : E} extensions to an isomorphism τ of
K onto a subfield of F . 
√ √
Example 5.2.2. Consider E = Q( √ 2, 3) over Q. We√know that
{E : Q} = [E : Q] = 4. Also, {E : Q( 2)} = 2, and {Q( 2) : Q} =
2, so
√ √
4 = {E : Q} = {E : Q( 2)}{Q( 2) : Q} = (2)(2).
5.3. Splitting fields.
We will determine for what extension field F ≤ E ≤ F , every iso-
morphic mapping of E onto a subfield of F over F is actually an
automorphism of E.
Definition 5.3.1. Let F be a field with algebraic closure F . Let
P = {fi (x) : i ∈ I} be a collection of polynomials in F [x]. A field
E ≤ F is the splitting field of P over F if E is the smallest subfield
of F containing F and all the zeros in F of each of fi (x) ∈ P. A field
K ≤ F is a splitting field over F if it is the splitting field of some
set of polynomials in F [x].
March 18, 2022 9:45 amsart-9x6 12819-main page 120

120 RING AND FIELD THEORY


√ √
Example 5.3.1. We see that Q[ 2, 3] is a splitting field of {x2 −
2, x2 − 3} and also of x4 − 5x2 + 6.
For one polynomial f (x) ∈ F [x], we shall often refer to the splitting
field of {f (x)} over F as the splitting field of f (x) over F , denoted
by Ff (x) .
Let P = {fi (x) : i ∈ I} ⊂ F [x], and let
RP = {a ∈ F : fi (a) = 0 for some i}.
Then we can see that F (RP ) is the splitting field of {fi (x) : i ∈ I}
over F . If P = {f (x)} we will simple denote Rf (x) = R{f (x)} .

Theorem 5.3.2. A field E, where F ≤ E ≤ F , is a splitting field


over F if and only if every τ ∈ Gal(F /F ) maps E onto itself, i.e.,
τ |E ∈ Gal(E/F ).
Proof. (⇒). Let E be a splitting field over F in F of P = {fi (x) :
i ∈ I}. Let RP be defined above. Then E = F (RP ). For any α ∈ F ,
fi (α) = 0 if and only if fi (τ (α)) = 0. So τ (RP ) = RP . Thus τ (E) =
τ (F (RP )) = F (τ (RP )) = F (RP ) = E, i.e., τ |E ∈ Gal(E/F ).
(⇐). Suppose that τ |E ∈ Gal(E/F ) for every τ ∈ Gal(F /F ).
Take a basis β for E over F . For each a ∈ β let pa (x) = irr(a, F ).
Let P = {pa (x) : a ∈ β}. We claim that E = F (Rp ), i.e., E is the
splitting field over F in F of P . Since E ⊆ F (Rp ), it is enough to
show that F (Rp ) ⊆ E.
Take any g(x) = pa (x) ∈ P . If b is any zero of g(x) in F , then
there is a conjugation isomorphism ψa,b of F (a) onto F (b) over F .
By Theorem 4.2.1, ψa,b can be extended to an automorphism τ of F .
Since τ (E) = E we see that τ (a) = b ∈ E. Consequently, RP ∈ E.
Thus F (Rp ) ⊆ E. 
Definition 5.3.3. Let E be an extension field of a field F . A poly-
nomial f (x) ∈ F [x] splits in E if it factors into a product of linear
factors in E[x].
Corollary 5.3.4. If E ≤ F is a splitting field over F , then every
irreducible polynomial in F [x] having a zero in E splits in E.
Proof. Since E is a splitting field over F in F , then τ |E ∈
Gal(E/F ) for every τ ∈ Gal(F /F ). The last paragraph of the proof
of above Theorem showed precisely that any irreducible polynomial
March 18, 2022 9:45 amsart-9x6 12819-main page 121

AUTOMORPHISMS OF FIELDS 121

g(x) ∈ F [x] having a zero in E have all zeros in E, i.e., its factoriza-
tion into linear factors in F [x], actually takes place in E[x], so g(x)
splits in E. 
Corollary 5.3.5. If E ≤ F is a splitting field over F , then S(E/F ) =
Gal(E/F ). If E is further of finite degree over F , then {E : F } =
|Gal(E/F )|.
Proof. It is clear that Gal(E/F ) ⊆ S(E/F ). Each σ ∈ S(E/F )
can be extended to an automorphism τ of F .
F ≤ E ≤ F
|| σ↓ τ↓
F ≤ F = F
Since E is a splitting field over F , then, σ = τ |E ∈ Gal(E/F ). Thus
S(E/F ) = Gal(E/F ).
The equation {E : F } = |Gal(E/F )| then follows clearly. 
Corollary 5.3.6. Let α ∈ F be algebraic over F . Then {F (α) : F }
is the number of different zeros of irr(α, F ).
Proof. By Theorem 5.1.2 we know that α, σ(α) are conjugate
over F for each σ ∈ S(F (α)/F ), and each conjugate β of α can have
a conjugation isomorphism ψα,β . 
If F ≤ E ≤ K is a chain of field extensions such that F ≤ E is
splitting and E ≤ K is splitting, it is falsepto conclude that F ≤ K
√ √
is splitting. For example Q ≤ Q[ 2] ≤ Q[ 2 − 1].
In the next section we will determine conditions under which
|Gal(E/F )| = {E : F } = [E : F ] for finite extensions E of F .
Example 5.3.2. Find the splitting field E of x3 − 2 over Q.

Solution.
√ We know that x3 − 2 does not split in Q( 3 2), for
Q( 3√2) < R and only one zero of x3 − √ 2 is real. Thus x3 − 2 factors in
(Q( 3 2))[x] into a √
linear factor x − 3 2 and an irreducible quadratic
3
factor. So [E : Q( 2)] = 2. Then

3
√3
[E : Q] = [E : Q( 2)][Q( 2) : Q] = (2)(3) = 6.
We can easily see that
√ √
−1 + i 3 √ 3 −1 − i 3 √3
2 and 2
2 2
√ √
are the other zeros of x3√− 2 in √ C. Thus E = Q( 3 2, i 3). (This is
not the same field as Q( 3 2, i, 3), which is of degree 12 over Q.) 
March 18, 2022 9:45 amsart-9x6 12819-main page 122

122 RING AND FIELD THEORY

5.4. Separable extensions.


We now assume that all algebraic extensions of a field F under con-
sideration are contained in one fixed algebraic closure F of F .
Our aim in this section is to determine, for a finite extension E of
F , what conditions ensure {E : F } = [E : F ]. The key to answering
this question is to consider the multiplicity of zeros of polynomials.
Definition 5.4.1. Let f (x) ∈ F [x]. An element α ∈ F such that
f (α) = 0 is a zero of f (x) of multiplicity r if (x − α)r |f (x) but
(x − α)r+1 6 |f (x) in F [x].
Theorem 5.4.2. Let f (x) be irreducible in F [x]. Then all zeros of
f (x) in F have the same multiplicity.
Proof. Let α, β ∈ Rf (x) . We have a conjugation isomorphism
ψα,β : F (α) → F (β). Then ψα,β can be extended to an isomorphism
τ : F → F . Thus τ induces a natural isomorphism τx : F [x] → F [x],
with τx (x) = x. Now τx leaves f (x) fixed, since f (x) ∈ F [x] and
ψα,β leaves F fixed. However,
τx ((x − α)r ) = (x − β)r ,
which shows that the multiplicity of β in f (x) is greater than or equal
to the multiplicity of α. A symmetric argument gives the reverse
inequality, so the multiplicity of α equals that of β. 
Corollary 5.4.3. If f (x) ∈ F [x] is irreducible, then f (x) has a
factorization in F [x] of the form
Yn
a (x − αk )r ,
k=1

where the αk are the distinct zeros of f (x) in F and a ∈ F.


Proof. The corollary is immediate from the previous Theorem.

To illustrate the above result let us work out some examples next.
Example 5.4.1. Let E = Zp (y), where y is an indeterminate. Let
α = y p , and let F = Zp (α) ≤ E. Now E = F (y) is algebraic over F ,
since y is a zero of (xp − α) ∈ F [x]. Then irr(y, F ) xp − α in F [x].
Since F (y) 6= F , we must have the degree of irr(y, F ) ≥ 2. Note
that xp − α = xp − y p = (x − y)p , since E has characteristic p. Thus
irr(y, F ) (x−y)p , and deg(y, F ) = k > 1. Then irr(y, F ) = (x−y)k ∈
March 18, 2022 9:45 amsart-9x6 12819-main page 123

AUTOMORPHISMS OF FIELDS 123

F [x]. We deduce that k = p and irr(y, F ) = xp −α, so the multiplicity


of y is p.
Theorem 5.4.4. If E is a finite extension of F , then {E : F } [E :
F ].
Proof. If E is finite over F , then E = F (α1 , . . . , αn ), for some
αk ∈ F . Let irr(αk , F (α1 , . . . , αk−1 )) have αk as one of nk distinct
zeros that are all of a common multiplicity rk . From Corollary 5.3.6
we know that
[F (α1 , . . . , αk ) : F (α1 , . . . , αk−1 )] = nk rk
= {F (α1 , . . . , αk ) : F (α1 , . . . , αk−1 )}rk .
From Corollaries 4.1.5 and 5.2.7 we see that
Y Y
[E : F ] = nk rk , and {E : F } = nk .
k k

Therefore, {E : F } [E : F ]. 
Definition 5.4.5. A finite extension E of F is a separable exten-
sion of F if {E : F } = [E : F ]. An algebraic element α of F is
separable over F if F (α) is a separable extension of F . An irre-
ducible polynomial f (x) ∈ F [x] is separable over F if every zero of
f (x) in F is separable over F .
√ √
Example 5.4.2. (a). The field E = Q[ 2, 3] is separable over
Q since we saw in previous examples that {E : Q} = 4 = [E :
Q].
(b). The extension E is not a separable extension of F in Exam-
ple 5.4.1.
By Corollary 5.1.3 we know that {F (α) : F } is the number of
distinct zeros of irr(α, F ). Also, the multiplicity of α in irr(α, F )
is the same as the multiplicity of each conjugate of α over F , by
Theorem 5.4.2. Thus α is separable over F if and only if irr(α, F )
has all zeros of multiplicity 1. This tells us at once that an irreducible
polynomial f (x) ∈ F [x] is separable over F if and only if f (x) has
all zeros of multiplicity 1.
Theorem 5.4.6. Let K be a finite extension of E and E a finite
extension of F . Then K is separable over F if and only if K is
separable over E and E is separable over F .
March 18, 2022 9:45 amsart-9x6 12819-main page 124

124 RING AND FIELD THEORY

Proof. Note that


[K : F ] = [K : E][E : F ], {K : F } = {K : E}{E : F },
{K : E} [K : E], {E : F } [E : F ].
Then K is separable over F , if and only if [K : F ] = {K : F },
if and only if [K : E][E : F ] = {K : E}{E : F }, if and only if
[K : E] = {K : E} and [E : F ] = {E : F } (Theorem 5.4.4). 
Theorem 5.4.6 can be extended in the obvious way, by induction,
to any finite tower of finite extensions. The top field is a separable
extension of the bottom one if and only if each field is a separable
extension of the one immediately under it.
Corollary 5.4.7. Let E be a finite extension of F . Then E is sep-
arable over F if and only if each α ∈ E is separable over F .
Proof. (⇒). Suppose that E is separable over F , and let α ∈ E.
Then
F ≤ F (α) ≤ E,
and Theorem 5.4.6 shows that F (α) is separable over F , i.e., α in E
is separable over F .
(⇐). Suppose that every α ∈ E is separable over F . Since E is a
finite extension of F , there exist α1 , · · · , αn such that
F < F (α1 ) < F (α1 , α2 ) < · · · < E = F (α1 , · · · , αn ).
Now since αk is separable over F , αk is separable over F (α1 , · · · , αk−1 ),
because q(x) = irr(αk , F (α1 , · · · , αk−1 )) divides irr(αk , F ), so that
αk is a zero of q(x) of multiplicity 1. Thus F (α1 , · · · , αk ) is separa-
ble over F (α1 , · · · , αk−1 ), so E is separable over F by Theorem 5.4.6.
So E is separable over F . 

We next prove that α can fail to be separable over F only if F is


an infinite field of characteristic p 6= 0.
Theorem 5.4.8. Let f (x) ∈ F [x] be an irreducible polynomial and
let E be a splitting field for f (x).
(1). If F has characteristic zero, then f (x) does not have multiple
roots in E.
(2). If F has characteristic p and f (x) has multiple roots in E,
then f (x) = g(xp ) for some g(x) ∈ F [x].
March 18, 2022 9:45 amsart-9x6 12819-main page 125

AUTOMORPHISMS OF FIELDS 125

Proof. (1) Since f (x) is irreducible, degf (x) ≥ 1. So since F has


characteristic zero, f 0 (x) 6= 0. Because f (x) is irreducible, we must
then have that gcd(f (x), f 0 (x)) = 1. So f (x) cannot have multiple
roots.
(2) By Theorem 4.2.10 we know that f (x) has multiple roots in E if
and only if gcd(f (x), f 0 (x)) 6= 1, if and only if gcd(f (x), f 0 (x)) ∼ f (x)
since f (x) is irreducible, if and only if f 0 (x) = 0, if and only if
f (x) = g(xp ) for some g(x) ∈ F [x]. 
Definition 5.4.9. A field F is perfect if every finite extension of
F is separable.
Theorem 5.4.10. Every field of characteristic zero is perfect.
Proof. Let E be a finite extension of a field F of characteristic
zero, and let α ∈ E. From Theorem 5.4.8 then f (x) = irr(α, F ) does
not have multiple roots in F . Therefore, α is separable over F for
all α ∈ E. By Corollary 5.4.7, this means that E is a separable
extension of F . 
We will find the answer for a field F of characteristic p > 0 to be
perfect.
Theorem 5.4.11. Let F be a field of characteristic p such that every
element of F is a pth power, and f (x) ∈ F [x] \ F. Then f (xp ) is not
irreducible in F [x].
Proof. We know that F = {ap : a ∈ F }.
Let f (x) = an xn + an−1 xn−1 + · · · + a1 x + a0 where ai ∈ F . Then
ai = bpi for some bi ∈ F . We see that
f (xp ) = bpn xpn + bpn−1 xp(n−1) + · · · + bp1 xp + bp0
= (bn xn + bn−1 xn−1 + · · · + b1 x + b0 )p .
Thus f (xp ) is not irreducible in F [x]. 
Theorem 5.4.12. A field F of characteristic p > 0 is perfect if and
only if every element of F is a pth power.
Proof. (⇐). Suppose that F = {ap : a ∈ F }. Let E be a
finite extension of F , and let α ∈ E. Since f (x) = irr(α, F ) is irre-
ducible, from Theorems 5.4.8 and 5.4.11 we know that it does not
have multiple roots. Therefore, α is separable over F for all α ∈ E.
By Corollary 5.4.7, this means that E is a separable extension of F .
Consequently, F is perfect.
March 18, 2022 9:45 amsart-9x6 12819-main page 126

126 RING AND FIELD THEORY

(⇒). Now suppose that F 6= {ap : a ∈ F }. Take a ∈ F which is


not a prime of any element in F . Let E be the splitting field of xp −a
over F . Take a root α ∈ E of xp − a. We know that a = αp and
xp − a = (x − α)p . Since α ∈
/ F , deg(α, F ) > 1 and irr(α, F )|(x − α)p .
Thus irr(α, F ) must have multiple root α, i.e., α is not separable over
F . Consequently, F is not perfect. 
As a consequence we have the following result.
Theorem 5.4.13. Every finite field F is perfect.
Proof. This follows from the fact that F = {ap : a ∈ F } (Theo-
rem 5.1.12) and Theorem 5.4.12. 

We have completed our aim: for finite extensions E of such perfect


fields F , [E : F ] = {E : F } if and only if E is a separable extension.

The following theorem is very useful.


Theorem 5.4.14. If E is a finite separable extension of a field F ,
then E = F (α) for some α ∈ E.
Proof. If F is a finite field, then E is also finite. Let α be a
generator for the cyclic group E ∗ of nonzero elements of E under
multiplication. Clearly, E = F (α), so α is a primitive element in this
case.
We now assume that F is infinite, and prove our theorem in the
case that E = F (β, γ). The induction argument from this to the gen-
eral case is straightforward. Let f (x) = irr(β, F ) have distinct zeros
β = β1 , · · · , βn , and let irr(γ, F ) have distinct zeros γ = γ1 , · · · , γm
in F , where all zeros are of multiplicity 1, since E is a separable
extension of F . Since F is infinite, we can find a ∈ F such that
a 6= (βi − β)/(γ − γj )
for all i and j, with j 6= 1. That is, a(γ − γj ) 6= βi − β. Letting
α = β + aγ, we have
α = β + aγ 6= βi + aγj ,
so α − aγj 6= βi for all i and all j 6= 1. Consider
h(x) = f (α − ax) ∈ (F (α))[x].
Now h(γ) = f (β) = 0. However, h(γj ) 6= 0 for j = 6 1 by con-
struction, since the βi were the only zeros of f (x). Hence h(x)
March 18, 2022 9:45 amsart-9x6 12819-main page 127

AUTOMORPHISMS OF FIELDS 127

and g(x) = irr(γ, F ) have a common factor in (F (α))[x], namely


irr(γ, F (α)), which must be linear, since γ is the only common zero
of g(x) and h(x). Thus γ ∈ F (α), and therefore β = α − aγ is in
F (α). Hence F (β, γ) = F (α). 
Corollary 5.4.15. Any finite extension of a field of characteristic
zero is a simple extension.
Proof. This corollary follows at once from Theorems 5.4.10 and
5.4.14. 
We see that the only possible “bad case” where a finite extension
may not be simple is a finite extension of an infinite field of charac-
teristic p 6= 0.
Example 5.4.3. Let p be a prime, L = Fp (x1 , x2 ), the fraction field
of the polynomial ring Fp [x1 , x2 ]. Let K = Fp (xp1 , xp2 ). Show that L
is not a simple extension over K.
Proof. If α ∈ L, there are f (x1 , x2 ), g(x1 , x2 ) ∈ Fp [x1 , x2 ] such
that α = fg(x
(x1 ,x2 )
1 ,x2 )
. Then

p f (x1 , x2 )p f (xp1 , xp2 )


α = = ∈K
g(x1 , x2 )p g(xp1 , xp2 )
since ap = a for any a ∈ Fp . So α is a root of xp − αp ∈ K[x].
Thus the irreducible polynomial irr(α, K) has degree at most p. So
[K(α) : K] = deg(irr(α, K)) ≤ p. On the other hand, we have
[L : K] = p2 , since {xk1 xj2 : 0 ≤ k, j < p} is a basis. For any α ∈ L,
we have K(α) 6= L. So K ≤ L is not a simple extension. This then
implies K ≤ L is not separable. 

It is interesting to know the following useful result which is named


after Jacob Lüroth (1844–1910), who proved it in 1876. We omit the
proof while a detailed proof can be found in [M, Theorem 8.19].
Theorem 5.4.16 (Lüroth’s Theorem). Let K ≤ E ≤ K(x) be fields
where x is an indeterminate. Then there is α ∈ K(x) such that
E = K(α).
5.5. Exercises.
(1) Find the centralizer of complex conjugation in Aut(C/Q).
(2) (a). Show that x4 + 4 is reducible over Q.
(b). Find the splitting field over Q for the polynomial x4 + 4.
March 18, 2022 9:45 amsart-9x6 12819-main page 128

128 RING AND FIELD THEORY

(c). Find the Galois group over Q of the polynomial x4 + 4.


(3) Let F be a field generated over the field K by u and v of
relatively prime degrees m and n, respectively, over K. Prove
that [F : K] = mn.
(4) Let F ⊇ K be an extension field, with u ∈ F . If [K(u) : K]
is an odd number, show that K(u2 ) = K(u).
(5) Find the degree [F : Q], where F is the splitting field of the
polynomial x3 − 11 over the field Q of rational numbers.
(6) Let f (x) ∈ Q[x] be irreducible over Q, and let F be the
splitting field for f (x) over Q. If [F : Q] is odd, prove that
all of the roots of f (x) are real.
(7) A field F is called formally real if −1 is not expressible as
a sum of squares in F . Let f (x) ∈ F [x] be an irreducible
polynomial of odd degree, and let α be a root of f (x). Prove
that if F is formally real, then F [α] is also formally real.
(8) Let K1 , K2 be finite extensions of a field F contained in the
field K, and assume both are splitting field over F . Prove
that K1 K2 and K1 ∩ K2 are splitting field over F .
(9) Calculate the splitting field E of f (x) = x3 + x + 1 over Z2 ,
and find |E|.
(10) Find the degree of the splitting field over Z2 for the polyno-
mial (x3 + x + 1)(x2 + x + 1).
(11) Determine the splitting field over Q for x4 + x2 + 1.
(12) Determine the splitting field over Q for x4 + 2.
(13) Find the splitting field for x3 + x + 1 over Z2 .
(14) Let p be a prime number. Find the splitting fields for xp − 1
over Q and over R.
(15) Find the splitting field of x6 − 1 ∈ Z5 [x] over Z5 .
(16) Find the splitting field of x5 − 1 ∈ Z2 [x] over Z2 .
(17) Find the degree of the splitting field over Z2 for the polyno-
mial (x3 + x + 1)(x2 + x + 1).
(18) Let F be the splitting field in Q of x4 + 1.
(a). Show that [F : Q] = 4. √
(b). Find automorphisms
√ of F that have fixed fields Q( 2),
Q(i), and Q( 2i), respectively.
(19) Let F be an algebraic closure of F , and let
f (x) = xn + an−1 xn−1 + · · · + a1 x + a0 ∈ F [x].
March 18, 2022 9:45 amsart-9x6 12819-main page 129

AUTOMORPHISMS OF FIELDS 129

If (f (x))m ∈ F [x] and m · 1 6= 0 in F , show that f (x) ∈ F [x],


that is, all ai ∈ F.
(20) Let F be a field, u = f (x)/g(x) ∈ F [x] \ F where gcd(f (x),
g(x)) = 1. Show u is transcendental over F , x is algebraic
over F [u] and [F (x) : F (u)] = max(deg(f (x), deg(f (x)).
(21) Let I be a maximal ideal in F [x1 , · · · , xn ], where F is a field.
Show that the field F [x1 , · · · , xn ]/I is a finite extension of F .
(22) Show that σ ∈ Gal(F (x)/F ) if and only if
ax + b
σ(x) = , a, b, c, d ∈ F
cx + d
with ad − bc 6= 0.
(Remark: The group Gal(F (x1 , x2 , · · · , xn )/F ) is quite com-
plicated for n = 2 and is unknown for n ≥ 3.)
B1948 Governing Asia

This page intentionally left blank

B1948_1-Aoki.indd 6 9/22/2014 4:24:57 PM


March 18, 2022 9:45 amsart-9x6 12819-main page 131

6. Galois Theory

In 1830, Evariste Galois (1811–1832), at the age of 18, submitted to


the Paris Academy of Sciences a memoir on his theory of solvability
by radicals; Galois’ paper was ultimately rejected in 1831 as being
too sketchy and for giving a condition in terms of the roots of the
equation instead of its coefficients. Galois then died in a duel in 1832,
and his paper, “Mémoire sur les conditions de résolubilité des équa-
tions par radicaux”, remained unpublished until 1846 when it was
published by Joseph Liouville (1809–1882) accompanied by some of
his own explanations. Prior to this publication, Liouville announced
Galois’ result to the Academy in a speech he gave on July 4, 1843.
Galois’s characterization “dramatically supersedes the work of Abel
and Ruffini.” (Cited from Wikipedia.)
In this last chapter we will prove the elegant Galois Theorem that
explains the relationship between subfields of a field and its auto-
morphism group. As an application we provide the necessary and
sufficient conditions for polynomial equations over a field of charac-
teristic 0 in one variable to be solvable in terms of nested roots.
6.1. Galois Theorem.
The Galois theory gives a beautiful interplay of group theory and
field theory. We shall start by recalling the main results we have
developed and should keep well in mind. We always assume that F
is the algebraic closure of the field F .
1. Let F ≤ E ≤ F , α ∈ E, and let β be a conjugate of α over
F , that is, irr(α, F ) = irr(β, F ). Then we have the conjugation
isomorphism ψα,β : F (α) → F (β) that leaves F fixed and maps α
onto β.
2. If F ≤ E ≤ F and α ∈ E, then any σ ∈ Gal(F /F ) maps α
onto some conjugate of α.
3. If F ≤ E, we have the Galois group Gal(E/F ). For any S ⊆
Gal(E/F ), ES ≤ E. In particular, EGal(E/F ) ≤ E.
4. A field E, F ≤ E ≤ F , is a splitting field over F if and only if
S(E/F ) ≤ Gal(E/F ). If E is a finite extension and a splitting field
over F , then |Gal(E/F )| = {E : F }.
131
March 18, 2022 9:45 amsart-9x6 12819-main page 132

132 RING AND FIELD THEORY

5. If E is a finite extension of F, then {E : F }|[E : F ]. If E is


also separable over F, then {E : F } = [E : F ]. Also, E is separable
over F if and only if irr(α, F ) has all zeros of multiplicity 1 for every
α ∈ E.
6. If E is a finite extension of F and is a separable splitting field
over F , then |Gal(E/F )| = {E : F } = [E : F ].
Definition 6.1.1. A finite extension K of F is a finite normal
extension of F if K is a separable splitting field over F.
A finite normal extension of a field F is also called a Galois ex-
tension of F .
Theorem 6.1.2. Let K be a finite normal extension of F, and let E
be an extension of F such that F ≤ E ≤ K ≤ F .
(a). K is a finite normal extension of E, and
Gal(K/E) = {σ ∈ Gal(K/F ) : σ(a) = a ∀a ∈ E}.
(b). σ, τ ∈ Gal(K/F ) induce the same isomorphism of E onto a
subfield of F if and only if they are in the same left coset of
Gal(K/E) in Gal(K/F ).
Proof. (a). If K is the splitting field of a set P = {fi (x)|i ∈ I} of
polynomials in F [x], then K is the splitting field over E of P ⊂ E[x].
Since K is separable over F then K is separable over E. Thus K is
a normal extension of E. The first statement follows.
For any σ ∈ Gal(K/E) we have σ(a) = a for any a ∈ F . Thus
Gal(K/E) ⊆ Gal(K/F ). Since Gal(K/E) is a group under function
composition also, we see that Gal(K/E) ≤ Gal(K/F ).
(b). Let σ, τ ∈ Gal(K/F ). Then σ and τ are in the same left coset
of Gal(K/E) if and only if µ = τ −1 σ ∈ Gal(K/E), if and only if
σ(α) = (τ µ)(α) = τ (µ(α)) = τ (α),
since µ(α) = α for α ∈ E. 
If F ≤ E ≤ K is a chain of field extensions such that F ≤ E is
normal and E ≤ K is normal, it is false to conclude that F ≤ K is
√ p√
normal. For example Q ≤ Q[ 2] ≤ Q[ 2 − 1].
√ √
Example 6.1.1. Let E = Q( 2, 3). Find all subgroups of
Gal(E/Q) and their corresponding fixed subfields of E.
March 18, 2022 9:45 amsart-9x6 12819-main page 133

GALOIS THEORY 133

Solution. Now E is a finite normal extension of Q, and a previous


example showed that |Gal(E/Q)|
√ √ √ = 4. We recall them by giving their
values on the basis {1, 2, 3, 6} for E over Q.
ι : The
√ identity
√ map,
√ √
σ1 : 2 → − 2, 6 → − 6, and leaves the other basis elements
fixed, √ √ √ √
σ2 : 3 → − 3, 6 → − 6, and leaves the other basis elements
fixed, √ √ √ √
σ3 : 2 → − 2, 3 → − 3, and leaves the other basis elements
fixed.
We saw that {ι, σ1 , σ2 , σ3 } is isomorphic to Z2 × Z2 . The complete
list of subgroups, with each subgroup paired off with the correspond-
ing intermediate field that it leaves fixed, is as follows:

{ι, σ1 , σ2 , σ3 } ⇐⇒ Q,

{ι, σ1 } ⇐⇒ Q( 3),

{ι, σ2 } ⇐⇒ Q( 2),

{ι, σ3 } ⇐⇒ Q( 6),
√ √
{ι} ⇐⇒ Q( 2, 3).

The diagram of subgroups and the diagram of subfields are as follows.


√ √
Q( 2, 3)
√  |
√  √
Q( 2) Q( 6) Q( 3)
 | 
Q

{ι}
 | 
{ι, σ2 } {ι, σ3 } {ι, σ1 }
 | 
{ι, σ1 , σ2 , σ3 }
All subgroups of the abelian group {ι, σ1 , σ2 , σ3 } are normal sub-
groups, and all the intermediate fields are normal extensions of Q.

March 18, 2022 9:45 amsart-9x6 12819-main page 134

134 RING AND FIELD THEORY

When K is a finite normal extension of F, and G a group, we


define the following sets
[[K, F ]] = {E : F ≤ E ≤ K}, [[G]] = {H : H ≤ G}.
Them we have the following maps:
λ :[[K, F ]] → [[Gal(K/E)]], E 7→ Gal(K/E), ∀E ∈ [[K, F ]];
(6.1)
µ :[[Gal(K/E)]] → [[K, F ]], H 7→ EH , ∀H ∈ [[Gal(K/E)]].
We will prove our main theorem in a few steps.
Theorem 6.1.3. Let K be a finite normal extension of a field F , let
E ∈ [[K, F ]].
(1). λ : [[K, F ]] → [[Gal(K/E)]] is a one-to-one map, and E =
Kλ(E) .
(2). [K : E] = |Gal(K/E)| and [E : F ] = (Gal(K/F ) : Gal(K/E)),
the number of left cosets of Gal(K/E) in Gal(K/F ).
Proof. (1). By definition we know that E ≤ KGal(K/E) . Next we
prove that KGal(K/E) ≤ E.
Let α ∈ K \ E. Since K is a normal extension of E, by using a
conjugation isomorphism and the Isomorphism Extension Theorem,
we can find an automorphism of K leaving E fixed and mapping α
onto a different zero of irr(α, E). So α ∈ / KGal(K/E) . This implies
that KGal(K/E) ≤ E, so E = KGal(K/E) .
For any E1 , E2 ∈ [[K, F ]], if λ(E1 ) = λ(E2 ), then
E1 = Kλ(E1 ) = Kλ(E2 ) = E2 .
So λ is one to one.
(2). Since K is a finite normal extension over F and E, from
Corollary 5.3.5 we see that [K : E] = {K : E} = |S(K/E)| =
|Gal(K/E)|. Again from Corollary 5.3.5 we obtain that
[E : F ] = [K : F ]/[K : E] = {K : F }/{K : E} = |S(K/F )|/|S(K/E)|
= |Gal(K/F )|/|Gal(K/E)| = (Gal(K/F ) : Gal(K/E)).

Theorem 6.1.4. Let K be a finite normal extension of a field F .
For any G ≤ Gal(K/F ), we have Gal(K/KG ) = G.
Proof. Let G ≤ Gal(K/F ). Then KG = {a ∈ K : σ(a) = a, ∀σ ∈
G} and
G ≤ Gal(K/KG ) ≤ Gal(K/F ).
March 18, 2022 9:45 amsart-9x6 12819-main page 135

GALOIS THEORY 135

We need to show that it is impossible to have G a proper subgroup


of Gal(K/KG ). We shall suppose that G < Gal(K/KG ) and shall
derive a contradiction. As a finite separable extension, K = KG (α)
for some α ∈ K. Let
n = [K : KG ] = {K : KG } = |Gal(K/KG )|.
Then G < Gal(K/KG ) implies that r = |G| < |Gal(K/KG )| = n.
Let G = {σ1 , · · · , σr }, and consider the polynomial
r
Y
f (x) = (x − σi (α)).
i=1
Then f (x) is of degree r < n. Now the coefficients of each power of
x in f (x) are symmetric expressions in the σi (α). For example, the
coefficient of xr−1 is
−σ1 (α) − σ2 (α) − · · · − σr (α).
Thus these coefficients are invariant under each isomorphism σi ∈ G,
since G = {σσ1 , · · · , σσr } for any σ ∈ G. Hence f (x) ∈ KG [x].
Since some σi is ι, we see that some σi (α) is α, so f (α) = 0. So
irr(α, KG )|f (x). Therefore, we would have
deg(α, KG ) ≤ r < n = [K : KG ] = [KG (α) : KG ] = deg(α, KG ).
This is impossible. Thus we have proved our result. 
Theorem 6.1.5. Let K be a finite normal extension of a field F , let
E ∈ [[K, F ]].
(1). E is a normal extension of F if and only if Gal(K/E) 
Gal(K/F ).
(2). If E is a normal extension of F , then Gal(E/F ) ' Gal(K/F )/
Gal(K/E).
Proof. (1). Since E ∈ [[K, F ]], it is separable over F . Thus E
is normal over F if and only if E is a splitting field over F. By the
Isomorphism Extension Theorem, every isomorphism of E onto a
subfield of F leaving F fixed can be extended to an automorphism of
K, since K is normal over F . Thus the automorphisms of Gal(K/F )
induce all possible isomorphisms of E onto a subfield of F over F .
We know that E is a splitting field over F , (hence is normal over F ),
if and only if
σ(α) ∈ E, ∀σ ∈ Gal(K/F ), α ∈ E,
March 18, 2022 9:45 amsart-9x6 12819-main page 136

136 RING AND FIELD THEORY

if and only if
τ (σ(α)) = σ(α), ∀σ ∈ Gal(K/F ), α ∈ E, τ ∈ Gal(K/E),
if and only if
(σ −1 τ σ)(α) = α, ∀α ∈ E, σ ∈ Gal(K/F ), and τ ∈ Gal(K/E),
if and only if
(σ −1 τ σ) ∈ Gal(K/E), ∀σ ∈ Gal(K/F ), and τ ∈ Gal(K/E),
if and only if Gal(K/E)  Gal(K/F ).
(2). For σ ∈ Gal(K/F ), let σE be the automorphism of E induced
by σ (we are assuming that E is a normal extension of F ). Thus
σE ∈ Gal(E/F ). The map
φ : Gal(K/F ) → Gal(E/F ), σ 7→ σE
is a homomorphism. By the Isomorphism Extension Theorem, every
µ ∈ Gal(E/F ) can be extended to some automorphism of K; that
is, it is µ = τE for some τ ∈ Gal(K/F ). Thus φ is onto Gal(E/F ).
The kernel of φ is Gal(K/E). Therefore, by the First Isomorphism
Theorem, Gal(E/F ) ' Gal(K/F )/Gal(K/E). 

We have proved the Galois Theorem.


Theorem 6.1.6 (Galois Theorem). Let K be a finite normal exten-
sion of a field F , let E ∈ [[K, F ]].
(1). The maps λ and µ defined in (6.1) are bijections and λ = µ−1 .
(2). [K : E] = |Gal(K/E)| and [E : F ] = (Gal(K/F ) : Gal(K/E)).
(3). E is a normal extension of F if and only if Gal(K/E) 
Gal(K/F ).
(4). If E is a normal extension of F , then Gal(E/F ) ' Gal(K/F )/
Gal(K/E).
The Main Theorem of Galois Theory is a strong tool in the study
of zeros of polynomials. If f (x) ∈ F [x] is such that every irreducible
factor of f (x) is separable over F , then the splitting field K of f (x)
over F is a normal extension of F . The Galois group Gal(K/F ) is
the (Galois) group of the polynomial f (x) over F which is also
denoted by Gal(f (x)/F ), even for any f (x) ∈ F [x].
The following lemma is obvious.
March 18, 2022 9:45 amsart-9x6 12819-main page 137

GALOIS THEORY 137

Lemma 6.1.7. Let K be a splitting field for f (x) ∈ F [x] over F ,


let Rf (x) = {α1 , · · · , αn }. Then K = F (α1 , · · · , αn ), and any σ ∈
Gal(K/F ) permutes the roots α1 , · · · , αn so gives an injective group
homomorphism Gal(K/F ) → Sn , i.e., Gal(K/F ) ≤ Sn .
We can determine the Galois groups of finite fields.
Theorem 6.1.8. Let K be a finite extension of degree n of a finite
field F of pr elements. Then Gal(K/F ) is cyclic of order n, and is
r
generated by σpr , where for α ∈ K, σpr (α) = αp .
Proof. From Theorem 5.4.13 (any finite field is perfect) we have
seen that K is a separable extension of F . Since |F | = pr and
[K : F ] = n, so |K| = prn . Then we have seen that K is the splitting
rn
field of xp − x over F . Hence K is a normal extension of F .
r
From Theorem 4.4.5 we know that Kσpr = {a ∈ K : ap = a} = F ,
and (σpr )k = ι implies that n|k. So |hσpr i| = n. Since |Gal(K/F )| =
[K : F ] = n, then Gal(K/F ) = hσpr i is cyclic and generated by σpr .

We use this theorem to give another illustration of the Main The-
orem of Galois Theory.
Example 6.1.2. Let F = Zp , and let E = Fp12 , so [E : F ] = 12.
Then Gal(E/F ) is isomorphic to the cyclic group hZ12 , +i. All the
subgroups of Gal(E/F ) = hσp i are the following:
hσp i, hσp2 i, hσp3 i, hσp4 i, hσp6 i, hσp12 i.
For example,
hσp4 i = {ι, σp4 , σp8 }.
6.2. Examples and an application.
In this section we will give some examples to compute the Galois
groups of some polynomials and use the Galois Theorem to prove
the fundamental theorem of algebra.
Example 6.2.1. Let E be the splitting field of x3 − 2 over Q. Find
all subgroups of Gal(E/Q) and their corresponding fixed subfields.
√ √
Solution. From Example 5.3.2 we know that E = Q( 3 2, 3 2ω,
√ √ 2πi
3
2ω 2 ) = Q( 3 2, ω), where ω = e 3 , which is the splitting field of
x3 − 2. So√E is√finite√normal over Q. Any element of G = Gal(E/Q)
permutes 3 2, 3 2ω, 3 2ω 2 . From Lemma 6.1.7 we know that G ⊆ S3 .
March 18, 2022 9:45 amsart-9x6 12819-main page 138

138 RING AND FIELD THEORY

Since |G| = [E : Q] = 6, we must have G = S3 . All the subgroups of


G are as follows.
{id}
2
2
2
3
h(12)i h(13)i h(23)i

3
h(123)i 3
3

2
S3
It is easy to compute all the corresponding subfields of G as follows.
E
2
2
2
√ √ √
3 Q( 3 2ω 2 ) Q( 3 2ω) Q( 3 2)

3
Q(ω) 3
3

2
Q

Example 6.2.2. Let E be the splitting field of x4
− 2 over Q. Find
all subgroups of Gal(K/Q) and their corresponding fixed subfields.
Solution. Now x4 − 2 is irreducible
√ over Q by Schönemann-
Eisenstein Criterion. Let α = 2. Then the four zeros of x4 − 2
4

in C are α, −α, iα, and −iα. We see that (iα)/α = i ∈ K. Since α is


a real number, Q(α) < R, so Q(α) 6= K. However, since Q(α, i) con-
tains all zeros of x4 − 2, we see that Q(α, i) = K. Letting E = Q(α).
Now {1, α, α2 , α3 } is a basis for E over Q, and {1, i} is a basis for
K over E. Thus {1, α, α2 , α3 , i, iα, iα2 , iα3 } is a basis for K over Q.
From Lemma 6.1.7 we know that Gal(K/Q) ⊆ S4 . Since [K : Q] = 8,
we must have |Gal(K/Q)| = 8. We will find eight automorphisms
σ ∈ Gal(K/Q).
March 18, 2022 9:45 amsart-9x6 12819-main page 139

GALOIS THEORY 139

Since K = Q(α, i), each Gal(K/Q) is completely determined by


σ(i) and σ(α). But σ(α) ∈ {1, α, α2 , α3 }. Likewise, σ(i) = ±i, zeros
of irr(i, Q) = x2 + 1. Thus the four possibilities for σ(α), combined
with the two possibilities for σ(i), must give all eight automorphisms.
ρ0 ρ1 ρ2 ρ3 µ1 δ1 µ2 δ2
α 7→ α iα −α −iα α iα −α −iα
i 7→ i i i i −i −i −i −i
For example, ρ3 (α) = −iα and ρ3 (i) = i, while ρ0 is the identity
automorphism.
Now
(µ1 ρ1 )(α) = µ1 (ρ1 (α)) = µ1 (iα) = µ1 (i)µ1 (α) = −iα,
and, similarly,
(µ1 ρ1 )(i) = −i,
so µ1 ρ1 = δ2 . A similar computation shows that
(ρ1 µ1 )(α) = iα and (ρ1 µ1 )(i) = −i.
Thus ρ1 µ1 = δ1 , so ρ1 µ1 6= µ1 ρ1 and Gal(K/Q) is not abelian.
There are two noisomorphic nonabelian groups of order 8. Since
ρ1 is of order 4, µ1 is of order 2, {ρ1 , µ1 } generates Gal(K/Q), and
µ1 ρ1 µ1 = ρ−1
1 . Thus Gal(K/Q) is isomorphic to the dihedral group
D4 (also called octic group). Actually all the subgroups of Gal(K/Q)
are
H0 = Gal(K/Q), H1 = {ρ0 , ρ2 , µ1 , µ2 },
H2 = {ρ0 , ρ1 , ρ2 , ρ3 }, H3 = {ρ0 , ρ2 , δ1 , δ2 },
H4 = {ρ0 , µ1 }, H5 = {ρ0 , µ2 },
H6 = {ρ0 , ρ2 }, H7 = {ρ0 , δ1 },
H8 = {ρ0 , δ2 }, H9 = {ρ0 }.
The determination of the fixed fields KHi sometimes requires a bit
of ingenuity. Let’s illustrate.
To find KH2 , we merely have to find an extension of Q of degree
2 left fixed by {ρ0 , ρ1 , ρ2 , ρ3 }. Since all ρj leave i fixed, then KH2 =
Q(i).
To find KH4 , we have to find an extension of Q of degree 4 left fixed
by ρ0 and µ1 . Since µ1 leaves α fixed and α is a zero of irr(α, Q) =
x4 − 2, we see that Q(α) is of degree 4 over Q and is left fixed by
{ρ0 , µ1 }. By Galois Theorem 6.1.6, it is the only such field. Here we
March 18, 2022 9:45 amsart-9x6 12819-main page 140

140 RING AND FIELD THEORY

are using strongly the one-to-one correspondence given by the Galois


theory. Then KH4 = Q(α).
Let us find KH7 . Since H7 = {ρ0 , δ1 } is a group, for any β ∈ K
we see that ρ0 (β) + δ1 (β) is left fixed by ρ0 and δ1 . Taking β = α,
we see that ρ0 (α) + δ1 (α) = α + iα is left fixed by H7 . We can check
and see that ρ0 and δ1 are the only automorphisms leaving α + iα
fixed. Thus, by the one-to-one correspondence, we must have
√4
√4
Q(α + iα) = Q( 2 + i 2) = KH7 .
Next we find irr(α + iα, Q). If γ = α + iα, then for every conjugate
of γ over Q, there exists an automorphism of K mapping γ into
that conjugate. Thus we need only compute the various different
values σ(γ) for σ ∈ Gal(K/Q) to find the other zeros of irr(γ, Q).
Elements σ of Gal(K/Q) giving these different values can be found
by taking a set of representatives of the left cosets of Gal(K/Q(γ)) =
{ρ0 , δ1 } in Gal(K/Q). A set of representatives for these left cosets is
{ρ0 , ρ1 , ρ2 , ρ3 }.
The conjugates of γ = α + iα are thus α + iα, iα − α, −α − iα, and
−iα + α. Hence
irr(γ, Q) = [(x−(α+iα))(x−(iα−α))]·[(x−(−α−iα))(x−(−iα+α))]
= (x2 − 2iαx − 2α2 )(x2 + 2iαx − 2α2 ) = x4 + 4α4 = x4 + 8.
You may try to use a simpler way to compute irr(γ, Q).
We leave all other subgroups as exercises for you to compute their
fixed subfields. 
Our next example will give an extension of degree 4 for the splitting
field of a quartic.
Example 6.2.3. Let K be the splitting field of x4 + 1 over Q. Find
all subgroups of Gal(K/Q) and their corresponding fixed subfields.
Solution. Since x4 + 1 = (x2 − i)(x2 + i) is the irreducible fac-
torization over the field Q[i], which is not a factorization in Q[x],
√ so
x4 + 1 is irreducible
√ over Q. The zeros of x 4 + 1 are (1 ± i)/ 2 and

(−1 ± i)/ 2. A computation shows that if


1+i
α= √ ,
2
then
−1 + i −1 − i 1−i
α2 = i, α3 = √ , α5 = √ , and α7 = √ .
2 2 2
March 18, 2022 9:45 amsart-9x6 12819-main page 141

GALOIS THEORY 141

Thus K = Q(α), and [K : Q] = 4. Let us compute Gal(K/Q) and


give the group. Since there exist automorphisms of K mapping α
onto each conjugate of α, and since an automorphism σ of Q(α)
is completely determined by σ(α), we see that the four elements of
Gal(K/Q) are defined by
σ1 σ3 σ5 σ7
α 7→ α α3 α5 α7
Since
(σj σk )(α) = σj (αk ) = (αj )k = αjk
and α8 = 1, we see that Gal(K/Q) is isomorphic to the group
{1, 3, 5, 7} under multiplication modulo 8. We know that σj2 = σ1 ,
the identity, for all j. We can see that Gal(K/Q) is the Klein 4-group
which is isomorphic to Z2 × Z2 .
To find K{σ1 ,σ3 } , it is only necessary to find an element of K not in
Q left fixed by {σ1 , σ3 }, since [K{σ1 ,σ3 } : Q] = 2. Clearly σ1 (α)+σ3 (α)
is left fixed by both σ1 and σ3 , since {σ1 , σ3 } is a group. We have

σ1 (α) + σ3 (α) = α + α3 = i 2.

So K{σ1 ,σ3 } = Q[i 2]. Similarly,

σ1 (α) + σ7 (α) = α + α7 = 2

is left fixed by {σ1 , σ7 }. So K{σ1 ,σ7 } = Q[ 2].
This technique is of no use in finding K{σ1 ,σ5 } , for
σ1 (α) + σ5 (α) = α + α5 = 0.
But by a similar argument, σ1 (α)σ5 (α) is left fixed by both σ1 and
σ5 , and
σ1 (α)σ5 (α) = αα5 = −i.
Thus K{σ1 ,σ5 } = Q(−i) = Q(i). 
There are many proofs for the fundamental theorem of algebra.
Next we will use the Galois theorem to give a very short algebraic
proof for the fundamental theorem of algebra.
Theorem 6.2.1 (The Fundamental Theorem of Algebra). The field
of complex numbers C is algebraically closed and R = C.
Proof. We know that C is an algebraic extension field of R of de-
gree 2. Let p(x) ∈ C[x] be irreducible. Any root u of p(x) in the alge-
braic closure C is algebraic over R, so in C[x] we have p(x) | irr(u, R).
March 18, 2022 9:45 amsart-9x6 12819-main page 142

142 RING AND FIELD THEORY

The splitting field of p(x) over C is contained in the splitting field E
of irr(u, R) x2 + 1 over R. The extension E over R is finite normal.
Since C 6 E, we have 2 | [E : R] and so 2 | Gal(E/R) |.
Now consider a 2-Sylow subgroup P 6 Gal(E/R) ([ZTL, Theo-
rem 3.4.1]). Then | Gal(E/R)|/|P | is odd. From Theorem 6.1.6, we
have P = Gal(E/EP ) and
| Gal(E/R)| | Gal(E/R)|
[EP : R] = =
| Gal(E/EP )| |P |
which shows that [EP : R] is odd. Theorem 5.4.14 allows us to write
EP = R(v) for some v whose minimal polynomial over R must also
have odd degree. Since every real polynomial of odd degree has
a real root, irreducibility implies that v has degree 1 over R, and
furthermore EP = R. Thus Gal(E/R) = P , i.e., Gal(E/R) is a
2-group.
Since Gal(E/C) ≤ Gal(E/R), we know that Gal(E/C) is also a
2-group. There is a normal subgroup N /Gal(E/C) of index 2 ([ZTL,
Theorem 3.3.8]). We have the Galois extension EN over C of degree 2.
It is well-known that every quadratic ax2 +bx+c ∈ C[x] has complex
roots. So we cannot have an irreducible quadratic polynomial in
C[x]. We deduce that | Gal(E/C)| = 1 and E = C. Therefore any
irreducible polynomial over C is of degree 1. Hence C is algebraically
closed and R = C. 
6.3. Cyclotomic extensions.
This section deals with the splitting extension fields of xn − 1 over a
field F .
Definition 6.3.1. The splitting field of xn − 1 over F is called the
nth cyclotomic extension of F .
Lemma 6.3.2. Let p be a prime and n ∈ N with p6 |n, F a field of
characteristic p. Then xn − 1 has no multiple roots in F .
Proof. We see that g(x) = (xn −1)0 = nxn−1 6= 0. Then gcd(xn −
1, g(x)) = 1. By Theorem 4.2.10, we know that xn −1 has no multiple
roots in F . 
Let us recall the Euler φ-function.
Definition 6.3.3. The Euler φ-function φ : N → N is defined by

φ(n) = k ∈ {1, · · · , n} : gcd(k, n) = 1 .
March 18, 2022 9:45 amsart-9x6 12819-main page 143

GALOIS THEORY 143

Let n = pr11 pr22 · · · prss , where p1 , · · · , ps are distinct primes, and


r1 , · · · , rs ∈ N. Then
Å ãÅ ã Å ã
1 1 1
φ(n) = n 1 − 1− ··· 1 − .
p1 p2 ps
For example φ(20) = φ(22 5) = 20(1 − 1/2)(1 − 1/5) = 8.
Under the conditions in Lemma 6.3.2, let K be the splitting field
of xn − 1 over F . Then xn − 1 has n distinct zeros in K, and these
form a cyclic group of order n under the field multiplication. We saw
that a cyclic group of order n has φ(n) generators. It is clear that
these φ(n) generators are exactly the primitive nth roots of unity.
Definition 6.3.4. Under the conditions in Lemma 6.3.2, the poly-
Qφ(n)
nomial Φn (x) = i=1 (x − αi ), where the αi are all the primitive
nth roots of unity in F , is called the nth cyclotomic polynomial
over F .
If char(F ) = p is a prime and p|n, then the nth cyclotomic poly-
nomial over F is not well-defined at this moment since there is no
primitive nth roots of unity. From Example 1.7.6 we know that Φn (x)
can be reducible over a field of finite characteristic.
Since any automorphism of the Galois group Gal(K/F ) must per-
mute the primitive nth roots of unity, we see that Φn (x) is left fixed
under every element of Gal(K/F ) regarded as extended in the nat-
ural way to K[x]. Thus Φn (x) ∈ F [x]. In particular, if F = Q, then
Φn (x) ∈ Q[x], and Φn (x)|xn − 1. Thus over Q, we must actually
have Φn (x) ∈ Z[x]. Certainly we can consider this integer polyno-
mial Φn (x) as a polynomial in Zp [x] for any prime p. We have seen
that, for any prime p, Φp (x) is irreducible over Q in Corollary 1.7.10.
Actually all Φn (x) are irreducible over Q.
Theorem 6.3.5. For any n ∈ N, Φn (x) is irreducible over Q.
Proof. Let ω be a primitive nth root of unity and let f (x) be its
irreducible polynomial over Q. Since ω is also a zero of xn − 1, it
follows that f (x)|xn − 1 and f (x) ∈ Z[x] by Gauss Lemma.
Claim 1: If p6 |n is a prime then ω p is a zero of f (x).
Suppose this claim is not true, i.e., ω p is not a zero of f (x). Let
g(x) = irr(ω p , Q). Then ω is a zero of g(xp ). So f (x)|g(xp ). Note
that f (x)g(x)|Φn (x).
Now, considering f (x)|g(xp ) in Zp [x] we see that f (x)|g(x)p since
g(xp ) = g(x)p in Zp [x]. Let h(x) be an irreducible common factor of
March 18, 2022 9:45 amsart-9x6 12819-main page 144

144 RING AND FIELD THEORY

f (x) and g(x) in Zp [x]. Then in Zp [x], we have


h2 (x)|f (x)g(x), f (x)g(x)|xn − 1,
i.e., h2 (x)|xn − 1. This contradicts Lemma 6.3.2. The claim is now
proved.
Using Claim 1 we see that f (ω k ) = 0 for any integer k with
gcd(k, n) = 1. These ω k are exactly all the primitive nth roots
of unity. Thus Φn (x) f (x). Hence Φn (x) = f (x). Thus Φn (x) is
irreducible over Q. 
Let us now limit our discussion to characteristic 0, in particular
to subfields of the complex numbers. We know that cos(2π/n) +
i sin(2π/n) is a primitive nth root of unity, a zero of Φn (x).
Example 6.3.1. Find the cyclotomic polynomial Φ8 (x) over Q.
Solution. A primitive 8th root of unity in C is
√ √ 1+i
ω = cos(2π/8) + i sin(2π/8) = 1/ 2 + i/ 2 = √ .
2
All the primitive 8th roots of unity in Q are ω, ω 3 , ω 5 , and ω 7 , so
Φ8 (x) = (x − ω)(x − ω 3 )(x − ω 5 )(x − ω 7 ).
We can compute, directly from this expression, Φ8 (x) = x4 + 1.
We can also find Φ8 (x) from deg(Φ8 (x)) = φ(8) = 4 and
x8 − 1 = (x4 + 1)(x2 + 1)(x2 − 1).

Theorem 6.3.6. The Galois group Gal(K/Q) of the nth cyclotomic
extension K of Q has φ(n) elements and is isomorphic to the abelian
unit group U(Zn ) of the commutative ring (Zn , +, ·).
Proof. Let
ω = cos(2π/n) + i sin(2π/n).
So ω is a generator of the cyclic multiplicative group of order n
consisting of all nth roots of unity. All the primitive nth roots of
unity, that is, all the generators of this group, are of the form ω m for
1 ≤ m < n with gcd(m, n) = 1. The field Q(ω) is the whole splitting
field of xn − 1 over Q. Then K = Q(ω). If ω m is another primitive
nth root of unity, then since ω and ω m are conjugate over Q, there
is an automorphism τm in Gal(K/Q) mapping ω onto ω m . Let τr be
March 18, 2022 9:45 amsart-9x6 12819-main page 145

GALOIS THEORY 145

the similar automorphism in Gal(K/Q) corresponding to a primitive


nth root of unity ω r . Then
(τm τr )(ω) = τm (ω r ) = (τm (ω))r = (ω m )r = ω rm .
This shows that the Galois group Gal(K/Q) is isomorphic to the
group U(Zn ) consisting of elements of Zn relatively prime to n un-
der multiplication modulo n. This group has φ(n) elements and is
abelian. 
Using similar arguments as in the proof of Theorem 6.3.6 and the
fact that (Rxn −1 , ·) is an abelian group (not necessarily of order φ(n))
we can similarly deduce the following results.
Theorem 6.3.7. Let E be the splitting field of xn −1 over F (not nec-
essarily a cyclotomic extension). Then the Galois group Gal(E/F )
is abelian.
Example 6.3.2. For any prime p if p|n, prove that Φnp (x) = Φn (xp )
over Q.
Proof. We denote an nth primitive root of unity by ωn . Since
p|n we see that φ(pn) = pφ(n). Then Φnp (x) and Φn (xp ) have the
p
same degree φ(pn) = pφ(n). Note that ωnp is an nth primitive root
of unity. We see that ωnp is a root of Φn (xp ). Since Φnp (x) =
irr(ωnp , Q), then Φnp (x)|Φn (xp ). Since both of them are monic and
of the same degree therefore Φnp (x) = Φn (xp ). 
For some other properties of cyclotomic polynomials, see Sec. 6.7
(34–40).
Next we list the first few cyclotomic polynomials over Q of small
composite degree.
Φ2 (x) = x + 1
Φ4 (x) = x2 + 1
Φ6 (x) = x2 − x + 1
Φ8 (x) = x4 + 1
Φ9 (x) = x6 + x3 + 1
Φ10 (x) = x4 − x3 + x2 − x + 1
Φ12 (x) = x4 − x2 + 1
Φ14 (x) = x6 − x5 + x4 − x3 + x2 − x + 1
March 18, 2022 9:45 amsart-9x6 12819-main page 146

146 RING AND FIELD THEORY

Φ15 (x) = x8 − x7 + x5 − x4 + x3 − x + 1
Φ16 (x) = x8 + 1
Φ18 (x) = x6 − x3 + 1
Φ20 (x) = x8 − x6 + x4 − x2 + 1
Φ21 (x) = x12 − x11 + x9 − x8 + x6 − x4 + x3 − x + 1
Φ22 (x) = x10 − x9 + x8 − x7 + x6 − x5 + x4 − x3 + x2 − x + 1
Φ24 (x) = x8 − x4 + 1
Φ25 (x) = x20 + x15 + x10 + x5 + 1
Φ26 (x) = x12 − x11 + x10 − x9 + x8 − x7
+ x6 − x5 + x4 − x3 + x2 − x + 1
Φ27 (x) = x18 + x9 + 1
Φ28 (x) = x12 − x10 + x8 − x6 + x4 − x2 + 1
Φ30 (x) = x8 + x7 − x5 − x4 − x3 + x + 1.
6.4. Solvability by radicals.
We knew that a quadratic polynomial f (x) = ax2 + bx + c, a 6= 0,
with real coefficients, has the following zeros in C:

−b ± b2 − 4ac
.
2a
Actually, this formula holds for f (x) ∈ F [x], where F is any field of
characteristic 6= 2 and the zeros are in F . For example, x2 + 2x + 3 ∈
Q[x] has its zeros

−2 ± −8 √ √
= −1 ± −2 ∈ Q( −2).
2
You may wonder whether the zeros of a cubic polynomial over Q
can also always be expressed in terms of radicals. The answer is yes
indeed. Also the zeros of a polynomial of degree 4 (called quartic)
over Q can be expressed in terms of radicals. But we will see that
some 5th degree polynomial (called quintic) do not have the “radical
formula” for their zeros. We will describe precisely what this means.
Definition 6.4.1. An extension K of a field F is an extension
of F by radicals (or a radical extension of F ) if there are el-
ements α1 , · · · , αr ∈ K and positive integers n1 , · · · , nr such that
March 18, 2022 9:45 amsart-9x6 12819-main page 147

GALOIS THEORY 147

K = F (α1 , · · · , αr ), α1n1 ∈ F and αini ∈ F (α1 , · · · , αi−1 ) for 1 < i ≤


r. We say an element α in an extension of F is expressible by rad-
icals if α is contained in some radical extension of F. A polynomial
f (x) ∈ F [x] is solvable by radicals over F if the splitting field
Ff (x) of f (x) over F is contained in a radical extension of F .
Lemma 6.4.2. Let K be a radical finite extension of a field F . Then
there is a finite splitting radical extension E of F containing K.
Proof. We need only to show the case where K = F (α) where
αn ∈ F . Let a = αn ∈ F . Then p(x) = irr(α, F )|xn − a ∈ F [x].
For any β ∈ Rp(x) = {α1 , · · · , αr } we also have p(β) = 0 = β n − a
So β n = a ∈ F . We see that E = F (α1 , · · · , αr ) is a finite splitting
radical extension of F containing K since α1n , · · · , αrn ∈ F . 
If further char(F ) = 0, in the above lemma E is a finite normal
radical extension of F containing K.
Example 6.4.1. Let ω be a primitive 5th root of unity. The splitting
field of x5 − 1 is K = Q(ω). Thus the polynomial x5 − 1 is solvable
by radicals over Q.
Similarly, x5 −√ 2 is solvable√ by radicals over K, for its splitting
field over Q is K( 2), where 5 2 is the real zero of x5 − 2. Thus the
5

polynomial
√ x5 − 2 is solvable by radicals over Q since Q ≤ Q(ω) ≤
5
Q(ω, 2).
In this section we shall show that a polynomial f (x) ∈ F [x] with
char(F ) = 0 is solvable by radicals over a field F if and only if its
splitting field E over F has a solvable Galois group Gal(E/F ). Then
we will find some quintic polynomials f (x) ∈ Q[x] with a splitting
field E over Q such that Gal(E/Q) ' S5 , the symmetric group on 5
letters. Since S5 is not solvable, the f (x) is not solvable by radicals.
We shall start with an arbitrary field F in this section and later
we shall restrict ourselves to fields of characteristic 0.
We first recall the definition for a group to be solvable and several
related results from [ZTL].
Definition 6.4.3. A group G is called solvable if it has a subnormal
series whose factor groups (quotient groups) are all abelian, that is,
if there are subgroups
1 = G0 < G1 < · · · < Gk = G
such that each Gj−1 is normal in Gj , and each Gj /Gj−1 is an abelian
group, for j = 1, 2, · · · , k.
March 18, 2022 9:45 amsart-9x6 12819-main page 148

148 RING AND FIELD THEORY

Definition 6.4.4. A subnormal series


{e} = G0  G1  · · ·  Gn = G
of a group G is a composition series if all the factor groups Gi+1 /Gi
are simple.
Theorem 6.4.5. Let G be a group and N be a normal subgroup of G.
(1). G is solvable iff both N and G/N are solvable.
(2). If G is solvable, and H is a subgroup of G, then H is solvable.
(3). If G and H are solvable, the direct product G × H is solvable.
(4). G is finite solvable if and only if G has a composition se-
ries with all composition factors of prime order (which are
certainly cyclic).
(5). The groups Sn and An are solvable if and only if n ≤ 4.
Lemma 6.4.6. Let a ∈ F. If K is the splitting field of xn − a over
F , then Gal(K/F ) is a solvable group.
Proof. From Corollary 1.6.5 we know that the nth roots of unity

form a cyclic subgroup (Un , ·) of (F , ·) with generator ω of order r.
Then the nth roots of unity are
1, ω, ω 2 , · · · , ω r−1 .
Let β ∈ E be a zero of xn − a ∈ F [x]. Then all zeros of xn − a are
β, ωβ, ω 2 β, · · · , ω r−1 β.
Case 1: ω ∈ F .
In this case we have K = F (β). Then an automorphism σ ∈
Gal(K/F ) is determined by the value σ(β). Now if σ(β) = ω i β and
τ (β) = ω j β, where τ ∈ Gal(K/F ), then
(τ σ)(β) = τ (σ(β)) = τ (ω i β) = ω i τ (β) = ω i ω j β,
since ω i ∈ F. Similarly,
(στ )(β) = ω j ω i β.
Thus στ = τ σ, and Gal(K/F ) is abelian and therefore solvable.
Case 2: ω ∈ / F , i.e., F does not contain any generator of (Un , ·).
Since β, ωβ are both zeros of xn − a, then ω = (ωβ)/β ∈ K. Let
F 0 = F (ω), so we have F < F 0 ≤ K. Now F 0 is the splitting field
of xn − 1. Since F 0 = F (ω), an automorphism µ ∈ Gal(F 0 /F ) is
determined by µ(ω). We must have µ(ω) = ω i for some i, since all
March 18, 2022 9:45 amsart-9x6 12819-main page 149

GALOIS THEORY 149

zeros of xn − 1 are powers of ω. If η(ω) = ω j for η ∈ Gal(F 0 /F ),


then
(µη)(ω) = ω ij .
and, similarly,
(ηµ)(ω) = ω ij .
Thus Gal(F 0 /F ) is abelian. By Theorem 6.1.6,
{ι} ≤ Gal(K/F 0 ) ≤ Gal(K/F )
is a normal series and hence a subnormal series of groups. The first
part of the proof shows that Gal(K/F 0 ) is abelian, and Galois The-
orem 6.1.6 tells us that
Gal(K/F )/Gal(K/F 0 ) ' Gal(F 0 /F ),
which is abelian. So Gal(K/F ) is solvable by Theorem 6.4.5 (1). 
Corollary 6.4.7. Let F be a field with char(F ) = 0, E be a finite
normal radical extension of F. Then Gal(E/F ) is solvable.
Proof. There exists a sequence of field extensions
F = F0 ⊂ F1 ⊂ · · · ⊂ Fr = E
for which there exist βi ∈ Fi and positive integers ni such that Fi =
Fi−1 (βi ) and βini ∈ Fi−1 . Let n = lcm(n1 , · · · , nr ) and let E 0 be the
splitting field for xn − 1 over E. Let ω be a generator of (Rxn −1 , ·),
the nth root of unity in E 0 . Then E 0 = E(ω) which is also a finite
normal extension of F . Let F 0 = F (ω) and let Fi0 = Fi (ω). Again we
have a sequence of field extensions
F ⊂ F 0 = F00 ⊂ F10 ⊂ · · · ⊂ Fr0 = E 0
0 (β ) and β ni ∈ F 0 . Moreover F 0 is a normal
such that Fi0 = Fi−1 i i i−1 i
0 0 = F . Let G = Gal(E 0 /F 0 ). Then these
extension of Fi−1 with F−1 i i
form a chain of subgroups, all normal in G−1 = Gal(E 0 /F ):
Gr = {e}  Gr−1  · · ·  G−1 .
We see that
Gj /Gj+1 = Gal(E 0 /Fj0 )/Gal(E 0 /Fj+1
0 0
) ' Gal(Fj+1 /Fj0 ),
G−1 /G0 = Gal(E 0 /F )/Gal(E 0 /F00 ) ' Gal(F00 /F ).
Now Gal(Fj+10 /F 0 ) is solvable by Lemma 6.4.6. So each G is solv-
j i
able for i = r, r − 1, · · · , −1 by induction and using Theorem 6.4.5.
March 18, 2022 9:45 amsart-9x6 12819-main page 150

150 RING AND FIELD THEORY

In particular, G−1 = Gal(E 0 /F ) is solvable. Finally Gal(E/F ) =


Gal(E 0 /F )/Gal(E 0 /E) must also be solvable by Theorem 6.4.5 (3).

Let us recall the following definition from Sec. 6.1.
Definition 6.4.8. Let f (x) ∈ F [x]. We define the Galois group of
f (x) to be the Galois group Gal(E/F ) of a splitting field E of f (x),
denoted by Gal(f (x)/F ).
Let Gal(E/F ) be the Galois group of a splitting field E of f (x) ∈
F [x], which has been considered as a subgroup of the symmetric
group on Rf (x) = {α1 , α2 , · · · , αn }, all the distinct zeros of f (x). See
Lemma 6.1.7.
Theorem 6.4.9. Let F be a field with char(F ) = 0 and f (x) ∈
F [x]. If f (x) is solvable by radicals over F , then the Galois group
Gal(f (x)/F ) is solvable.
Proof. Let E be the splitting field of f (x) over F . Then E is
normal over F since char(F ) = 0. Since f (x) is solvable by radicals,
by Lemma 6.4.2 we know that E is contained in a finite normal radical
extension, say E 0 . Then Gal(E 0 /F ) is solvable by Corollary 6.4.7.
Therefore, Gal(E/F ) = Gal(E 0 /F )/Gal(E 0 /E) is also solvable by
Theorem 6.4.5 (3). 

Next we will produce a concrete irreducible quintic polynomial


whose Galois group is S5 .
Theorem 6.4.10. Let p be a prime. Suppose that H is a subgroup
of Sp that contains a transposition and a p-cycle. Then H = Sp .
Proof. We may assume that σ = (1, 2), τ1 = (1, k1 , · · · , kp−1 ) ∈
H. If ki = 2 we see that τ1i = (1, 2, · · · ) ∈ H. Without loss of
generality we may assume that σ = (1, 2), τ = (1, 2, · · · , p) ∈ H.
Computing τ i στ −i we see that
(1, 2), τ (1, 2)τ −1 = (2, 3), τ 2 (1, 2)τ −2 = (3, 4), · · · , τ p−1 (1, 2)τ 1−p
= (p − 1, p) ∈ H.
Further
(1, 2), (1, 2)(2, 3)(1, 2) = (1, 3), (1, 3)(3, 4)(1, 3) = (1, 4),
· · · , (1, p − 1)(p − 1, p)(1, p − 1) = (1, p) ∈ H.
Thus H = Sp . 
March 18, 2022 9:45 amsart-9x6 12819-main page 151

GALOIS THEORY 151

Question: Can we change p to any n ∈ N in the previous theo-


rem?
Theorem 6.4.11. Let p be a prime, let f (x) ∈ Q[x] be an irreducible
polynomial of degree p with exactly p − 2 real roots. Then the Galois
group of f (x) over Q is Sp . Consequently f (x) is not solvable by
radicals over Q.
Proof. Let α be a root of f (x) and let E be its splitting field over
Q. Then E is finite normal over F , [Q(α) : Q] = p and Q(α) ⊂ E.
So the order of G = Gal(E/Q) is divisible by p since
|G| = [E : Q] = [E : Q[α)][Q[α) : Q].
By the Sylow Theorem, we certainly know that G must contain an
element τ of order p, which must be a p-cycle.
On the other hand the set of roots is invariant under complex
conjugation, so complex conjugation permutes the roots of f (x) and
therefore restricts to an automorphism σ of E of order two. Note
that we have identified G as a group of permutations of the roots.
Then σ identifies with the transposition (α1 , α2 ), where α1 , α2 are
the non-real roots. Hence G contains a transposition and p-cycle, so
by the previous theorem G = Sp . 
Question: Can we have the following result? “Let f (x) ∈ Q[x] be
an irreducible polynomial of degree n with exactly n − 2 real roots.
Then the Galois group of f (x) over Q is Sn .”
Example 6.4.2. Show that f (x) = 2x5 − 10x + 5 is irreducible over
Q, and find the number of real roots. Find the Galois group of f(x)
over Q, and explain why f (x) is not solvable by radicals.
Solution. The polynomial f (x) is irreducible over Q since it sat-
isfies Schönemann-Eisenstein Criterion for p = 5. Consider y = f (x)
as a continuous real valued function. The derivative f 0 (x) = 10x4 −10
has two real roots ±1. We see that
ß 0
f (x) > 0 if |x| > 1,
f 0 (x) < 0 if |x| < 1.
That is, f (x) is strictly increasing for |x| > 1, and f (x) is strictly
decreasing for |x| < 1. Since
f (−∞) = −∞, f (−1) = 13 > 0, f (1) = −3 < 0, f (+∞) = +∞,
March 18, 2022 9:45 amsart-9x6 12819-main page 152

152 RING AND FIELD THEORY

y
20

10

x
−1.5 −1 −0.5 0.5 1 1.5

−10

Figure 1. y = 2x5 − 10x + 5

x −∞ % −1 % 1 % ∞
f 0 (x) ∞ > 0 0 < 0 0 > 0 ∞
f (x) −∞ % 13 & −3 % ∞
we see that f (x) must have exactly three real roots. It follows from
Theorem 6.4.11 that the Galois group of f (x) over Q is S5 , and so it
is not solvable. Thus f (x) is not solvable by radicals. See Figure 1.

We have finally proved the famous Abel-Ruffini Theorem which
is named after Paolo Ruffini (1765–1822), who made an incomplete
proof in 1799, and Niels Henrik Abel (1802–1829), who provided
a proof in 1824. This is one of the greatest achievements of 19th
century mathematics.
Theorem 6.4.12. There exist quintic equations which are not solv-
able by radicals.
As we mentioned earlier this immediately implies that no formula
analogous to those for the cubic and quartic equations can exist for
quintic equations.
To have perfect Galois theory let us prove the converse of Theo-
rem 6.4.9 next.
Lemma 6.4.13. Let E be a finite normal extension of F with cyclic
Gal(E/F ) = hσi of order n > 1, where F contains a primitive nth
March 18, 2022 9:45 amsart-9x6 12819-main page 153

GALOIS THEORY 153

root of unity ω. Then β + ωσ(β) + · · · + ω n−1 σ n−1 (β) 6= 0 for some


β ∈ E.
Proof. Since E be a finite normal extension of F then E =
F (α) for some α ∈ E. Then the numbers {α, σ(α), . . . , σ n−1 (α)} are
pairwise distinct. Let
g = 1 + ωσ + ω 2 σ 2 + · · · + ω n−1 σ n−1
which is an F -linear multiplicative function on E. To the contrary,
we suppose g(E) = 0. Considering g(1) = g(α) = g(α2 ) = · · · =
g(αn−1 ) = 0 we have
1 + ω + · · · + ω n−1 = 0,
α + σ(α)ω + · · · + σ n−1 (α)ω n−1 = 0,
α2 + σ(α)2 ω + · · · + σ n−1 (α)2 ω n−1 = 0,
·········
αn−1 + σ(α)n−1 ω + · · · + σ n−1 (α)n−1 ω n−1 = 0.
This is impossible since the coefficient matrix is the invertible Vander-
monde matrix of α, σ(α), . . . , σ n−1 (α). So g(E) 6= 0, the statement
follows. 
Theorem 6.4.14. Let E be a finite normal extension of F with cyclic
Gal(E/F ) of order n, where F contains a primitive nth root of unity
ω. Then there exists a ∈ F such that f (x) = xn − a is irreducible
over F and E is a splitting field for f over F . Consequently, E is a
radical normal extension of F .
Proof. Let Gal(E/F ) = hσi. From the previous lemma there is
β ∈ E such that θ = β + ωσ(β) + · · · + ω n−1 σ n−1 (β) 6= 0. Now
σ(θ) = σ(β) + ωσ 2 (β) + · · · + ω n−2 σ n−1 (β) + ω n−1 σ n (β) = ω −1 θ
since σ n (β) = β. We take a = θn . Since Gal(E/F ) = hσi and
σ(θn ) = (σ(θ))n = (ω −1 θ)n = θn , then θn ∈ F . Now by defini-
tion of a, θ is a root of f (x) = xn − a, so the roots of xn − a are
θ, ωθ, · · · , ω n−1 θ. Therefore F (θ) is a splitting field for f (x) over
F . Since σ(θ) = ω −1 θ, the distinct automorphisms 1, σ, · · · , σ n−1
can be restricted to distinct automorphisms of F (θ). Consequently,
n ≤ |Gal(F (θ)/F )| = [F (θ) : F ] ≤ degf = n so [F (θ) : F ] = n. It
follows that E = F (θ) and (since f must be the irreducible polyno-
mial of θ over F ) f is irreducible over F . 
March 18, 2022 9:45 amsart-9x6 12819-main page 154

154 RING AND FIELD THEORY

In Theorem 6.4.14 we assumed that F contains a primitive nth


root of unity ω. This cannot be satisfied if char(F ) = p and p|n. In
this case we have the following result. One can find a proof in [J].
Theorem 6.4.15. Let char(F ) = p, E be a finite normal extension
of F with Gal(E/F ) of order p. Then there exists a ∈ E such that
E = F (a) where ap − a ∈ F .
Theorem 6.4.16. Let F be a field with char(F ) = 0. If E is a finite
normal extension of F with solvable Gal(E/F ), then E is contained
in a finite radical normal extension of F .
Proof. Since Gal(E/F ) is solvable, by Theorem 6.4.5 there is a
subnormal series
{0} = Gr  · · ·  G1  G0 = Gal(E/F )
such that each Gk /Gk+1 is cyclic. Let Ek = EGk . By Galois Theo-
rem 6.1.6, we get a sequence of field extension
F = E0 ⊆ · · · ⊆ Er = E.
We see that Gk = Gal(E/Ek ). Moreover, we know that Ek ⊆ Ek+1
is a normal extension with Galois group Gal(Ek+1 /Ek ) ∼ = Gk /Gk+1 .
So Gal(Ek+1 /Ek ) is cyclic. Let |Gk−1 /Gk | = nk , and n = [E : F ].
Then n = n1 n2 · · · nr . Let ω be an nth primitive root of unity.
From Theorem 6.4.14 we know that if Ek contains a primitive
nk th root of unity, then Ek+1 is a radical normal extension of Ek .
However, we may not have this in our hands. Now we add ω to each
field in the sequence:
0 (ω) ⊆ ES
ES 1 (ω) ⊆ · · · ⊆ ES
r (ω) = E(ω).
| | |
F = E0 ⊆ E1 ⊆ ··· ⊆ Er = E.
From Theorem 6.4.14 we know that E0 (ω) is a normal radical exten-
sion of E0 = F . We also know that Ek+1 (ω) is a normal extension
of Ek (ω) for all i. Next we show that Ek+1 (ω) is a radical extension
of Ek (ω) for all k.
We define a group homomorphism
φ : Gal(Ek+1 (ω)/Ek (ω)) → Gal(Ek+1 /Ek ), σ 7→ σ|Ek+1 .
This is well-defined because Ek+1 is normal extension of Ek , and
hence σ(Ek+1 ) = Ek+1 for any σ ∈ Gal(Ek+1 (ω)/Ek (ω)).
March 18, 2022 9:45 amsart-9x6 12819-main page 155

GALOIS THEORY 155

Next we show that φ is injective. Let φ(σ) = id. Since σ ∈


Gal(Ek+1 (ω)/Ek (ω)), it fixes Ek (ω). In particular, it fixes ω. So σ
must fix the whole of Ek+1 (ω). So σ = id.
By injectivity, we know that Gal(Ek+1 (ω)/Ek (ω)) is isomorphic to
a subgroup of Gal(Ek+1 /Ek ). So Gal(Ek+1 (ω)/Ek (ω)) is cyclic. By
Theorem 6.4.14, we know that Ek+1 (ω) is a radical normal extension
of Ek (ω). Then Er (ω) = E(ω) is a finite radical normal extension of
F containing E. 

Corollary 6.4.17. Let F be a field with char(F ) = 0 and f (x) ∈


F [x]. Then f (x) can be solved by radicals if and only if the Galois
group Gal(f (x)/F ) is solvable.

Proof. (⇒) This is Theorem 6.4.9.


(⇐) Let E be the splitting of f (x) over F . Then E is a finite nor-
mal extension of F with solvable Gal(E/F ). From Theorem 6.4.16
we know that E is contained in a finite radical normal extension of
F . So f (x) is solvable by radicals. 
Remark. When F has characteristic p, Corollary 6.4.17 fails.
Actually Theorem 6.4.16 fails.
Now we have the explanations why a polynomial equation of degree
≤ 4 has radical solutions.

Corollary 6.4.18. Let F be a field with charF = 0, and let f (x) ∈


F [x] with degf (x) = n ≤ 4. Then f (x) is solvable by radicals.

Proof. From Lemma 6.1.7 we know that Gal(f (x)/F ) is isomor-


phic to a subgroup of Sn which is solvable when n ≤ 4 by Theo-
rem 6.4.5. Applying Corollary 6.4.17 we know that f (x) is solvable
by radicals. 

6.5. Insolvability of equations of higher degree.


In this section we will construct some polynomials whose Galois
group is Sn for any n ∈ N.
Let F be a field, E = F (x1 , · · · , xn ), the field of rational func-
tions over F in the variables x1 , · · · , xn . Then there is an injective
homomorphism Sn → AutF (E) given by permutations of xi .
We define the field of symmetric rational functions K = ESn
to be the fixed subfield of Sn in E. We first prove a few important
results on symmetric rational functions.
March 18, 2022 9:45 amsart-9x6 12819-main page 156

156 RING AND FIELD THEORY

Definition 6.5.1. The elementary symmetric polynomials on


n-variables x1 , · · · , xn are s1 , s2 , · · · , sn defined by
X
si = xk1 · · · xki .
1≤k1 <k2 <···<ki ≤n

It is easy to see that


s1 = x1 + x2 + · · · + xn ,
s2 = x1 x2 + x1 x3 + · · · + xn−1 xn ,
(6.2)
·········
s n = x1 · · · xn .
Obviously, s1 , · · · , sn ∈ K.
Theorem 6.5.2. Let F be a field, E = F (x1 , · · · , xn ) and K = ESn .
(1) E is the splitting field of f (x) = xn − s1 xn−1 + · · · + (−1)n sn
over K.
(2) K = ESn ⊆ E is a finite normal extension with Gal(E/K) '
Sn .
(3) K = F (s1 , · · · , sn ).
(4) f (x) is irreducible over K.
Proof. (1) In E[x], we have
f (x) = (x − x1 ) · · · (x − xn ).
So E is the splitting field of f over K.
(2) Since f (x) is separable and E is the splitting field of f (x), then
K = ESn ⊆ E is a finite normal extension. By Galois Theorem, we
see that Gal(E/K) ' Sn .
(3) Let K1 = F (s1 , · · · , sn ). Clearly, K1 ⊆ K. Now K ⊆ E is a
finite normal extension, since E is the splitting field of f over K and
f has no repeated roots.
By Galois Theorem, since we have the finite normal extensions
K1 ⊆ K ⊆ E, we have Sn = Gal(E/K) ≤ Gal(E/K1 ). We also know
that Gal(E/K1 ) is a subgroup of Sn , we must have Gal(E/K1 ) =
Gal(E/K) = Sn . So we must have K = K1 .
(4) If f (x) = g(x)h(x) where g(x), h(x) ∈ K[x] are of positive
degree. For any σ ∈ Gal(E/K) = Gal(f (x)/K) we would have
σ(Rg(x) ) = Rg(x) , σ(Rh(x) ) = Rh(x) ,
which contradicts the fact that Gal(E/K) = Sn . So f (x) is irre-
ducible over K. 
March 18, 2022 9:45 amsart-9x6 12819-main page 157

GALOIS THEORY 157

Theorem 6.5.3. Let F be a field with charF = 0, x1 , x2 , · · · , xn be


variables over F , K = F (s1 , s2 , · · · , sn ) ≤ F (x1 , · · · , xn ). Then the
polynomial
f (x) = xn − s1 xn−1 + · · · + (−1)n sn ∈ K[x]
is solvable by radicals over K if and only if n < 5.
Proof. We know that K(x1 , x2 , · · · , xn ) is the splitting field of the
separable irreducible polynomial f (x) over K. From Theorem 6.5.2
we know that Gal(E/K) = Sn which is not solvable if n ≥ 5 by
Theorem 6.4.5. From Theorem 6.4.9 we see that f (x) is solvable by
radicals over K if and only if n < 5. 

6.6. Dedekind’s Theorem and discriminants of


polynomials.
In general it is difficult to compute the Galois groups of polynomials
over a field, particularly for higher degree polynomials. We include
the following theorem to help computing the Galois groups of poly-
nomials over Q. This theorem is named after Julius Wilhelm Richard
Dedekind. We omit its proof here. For the detailed proof see Sec. 61
in [J].
Theorem 6.6.1 (Dedekind’s Theorem). Let f (x) ∈ Z[x] be monic
with no repeated roots. Let f (x) ∈ Zp [x] be the obvious polynomial
obtained by reducing the coefficients of f mod p for a prime p. As-
sume that f (x) has no repeated roots. If f (x) factors as a product of
irreducibles of degree n1 , n2 , · · · , nr , then Gal(f (x)/Q) contains an
element of cycle type [n1 , · · · , nr ] (on Rf (x) ).

Example 6.6.1. Show that f (x) = x5 + 3x2 + 2x + 3 is not solvable


by radicals over Q.
Solution. In Z2 [x], f (x) = x5 + x2 + 1 is irreducible.
In Z3 [x], f (x) = x5 − x = (x2 + 1)x(x − 1)(x + 1) where x2 + 1 is
irreducible.
From Dedekind’s Theorem we know that Gal(f (x)/Q) ≤ S5 con-
tains a transposition and a 5-cycle. By Theorems 6.4.11 and 6.6.1
we see that Gal(f (x)/Q) = S5 over Q, and f (x) is not solvable by
radicals over Q. 
Discriminant of a polynomial is another tool to help computing
the Galois group of a polynomial.
March 18, 2022 9:45 amsart-9x6 12819-main page 158

158 RING AND FIELD THEORY

Definition 6.6.2 (Discriminant). Let F be a field, f (x) ∈ F [x], and


E the splitting field of f (x) over F with
f (x) = a(x − α1 ) · · · (x − αn ), a ∈ F, α1 , · · · , αn ∈ E.
We define
Y Y
∆f (x) = (αi − αj ), Df (x) = ∆2f (x) = (−1)n(n−1)/2 (αi − αj ),
i<j i6=j
and call Df (x) the discriminant of f .
Clearly, Df (x) 6= 0 if and only if f (x) has no repeated roots.
Theorem 6.6.3. Let F be a field, f (x) ∈ F [x], and E the splitting
field of f (x) over F . Suppose Df (x) 6= 0 and char(F ) 6= 2. Then
(1) Df (x) ∈ F ;
(2) Gal(E/F ) ⊆ An if and only if ∆f (x) ∈ F (if and only if Df (x)
is a square in F ).
Proof. (1). Note that E is a finite normal extension of F . It
is clear that Df (x) is fixed by Gal(E/F ) since it only permutes the
roots.
(2). For any transposition σ ∈ Sn switching αi , αj with i 6= j, we
see that
σ(∆f (x) ) = −∆f (x) .
So ∆f (x) ∈ F if and only if ∆f (x) is fixed by Gal(E/F ) if and only
if every element of Gal(E/F ) is even, if and only if Gal(E/F ) ≤ An .

Lemma 6.6.4. Let F be a field with char(F ) 6= 2, 3. Let f (x) =
x3 + bx + c ∈ F [x]. Then Df (x) = −4b3 − 27c2 .
Proof. Let f (x) = (x − α1 )(x − α2 )(x − α2 ). Then
α1 + α2 + α3 = 0, α1 α2 + α1 α3 + α2 α3 = b, α1 α2 α3 = −c,
i.e.,
b = −α1 α2 − α12 − α22 , c = α12 α2 + α22 α1 .
We compute
Df (x) + 4b3 + 27c2 = (α1 − α2 )2 (α1 − α3 )2 (α3 − α2 )2 + 4b3 + 27c2
= (α1 − α2 )2 (2α1 + α2 )2 (2α2 + α1 )2 + 4(−α1 α2 − α12 − α22 )3
+ 27(α12 α2 + α22 α1 )2 = 0
Thus Df (x) = −4b3 − 27c2 . 
March 18, 2022 9:45 amsart-9x6 12819-main page 159

GALOIS THEORY 159

Corollary 6.6.5. Let F be a field with char(F ) = 0. Let f (x) ∈ F [x]


be irreducible of degree 3. If Df (x) is a square of an element in F ,
then Gal(E/F ) = A3 , otherwise Gal(E/F ) = S3 .
Proof. (1). If Df (x) is a square of an element in F , then ∆f (x) ∈
F , Gal(E/F ) ⊆ A3 from Theorem 6.6.3. Note that Gal(E/F ) has
more that one element. So Gal(E/F ) = A3 .
(2). If Df (x) is not a square of an element in F , then ∆f (x) ∈
/ F,
Gal(E/F ) 6⊆ A3 from Theorem 6.6.3. The only nontrivial subgroups
of S3 are S3 , A3 and {(1), (i, j)} where 1 ≤ i ≤ j ≤ 3. Since
Gal(E/F ) acts on the three roots of f (x) transitively, we must have
Gal(E/F ) = S3 . 
By Theorem 6.1.8, the Galois group Gal(E/F ) is always cyclic for
any finite fields F ≤ E.
Example 6.6.2. Find Gal(f (x)/Q) for
(a). f (x) = x3 − 2x + 2,
(b). f (x) = x3 − 9x + 3.
Solution.
(a). We know that f (x) is irreducible over Q by Schönemann-
Eisenstein criterion. Since Df (x) = −4(−2)3 − 27(2)2 < 0,
using Corollary 6.6.5 we know that Gal(f (x)/Q) = S3 .
(b). We know that f (x) is irreducible over Q by Schönemann-
Eisenstein Criterion. Since Df (x) = −4(−9)3 − 27(3)2 > 0,
using Corollary 6.6.5 we know that Gal(f (x)/Q) = A3 .

Similar to Lemma 6.6.4 we can obtain the following formula.
Lemma 6.6.6. Let F be a field with char(F ) 6= 2, 3. Let f (x) =
x4 + cx2 + dx + e ∈ F [x]. Then Df (x) = 256e3 − 128c2 e2 + 144cd2 e −
27d4 + 16c4 e − 4c3 d2 . In particular, Dx4 +ax2 +b = 16b(a2 − 4b)2 .
6.7. Exercises.
(1) Find the Galois groups of x4 − 2 over the fields (a) (Z3 , +, ·),
(b) (Z7 , +, ·).
(2) Find the Galois group of x4 + 2 over the field (a) (Z3 , +, ·),
(b) (Z5 , +, ·).
(3) Let E ≤ F be a finite normal extension. Let
G = Gal(E/F ), K = {u ∈ E|στ (u) = τ σ(u)∀σ, τ ∈ G}.
March 18, 2022 9:45 amsart-9x6 12819-main page 160

160 RING AND FIELD THEORY

(a). Show that K is an intermediate subfield.


(b). Show that F ≤ K is a normal extension with abelian
Galois group.
(4) Let ω be a primitive 20th root of unity in C, and let E =
Q(ω).
(a). Identify the Galois group Gal(E/Q), explaining how the
individual automorphisms act on E.
(b). How many subfields of E are there which are quadratic
extensions of Q?
(c). Determine the irreducible polynomial of ω over Q.
(5) Let K be the splitting field of x6 − 25 over Q. Determine
Gal(K/Q). Explicitly determine all subfields of K, giving
generators over Q. Indicate which are normal over Q.
(6) Let F be a field of characteristic 0, F ⊂ E ⊆ F (x) be fields.
Show that Gal(F (x)/E) is finite.
(7) Let F be a field of characteristic 0. Show that F (x2 )∩F (x2 −
x) = F .
(8) Let F be a field of characteristic 0. Show that F (x2 − x) ∩
F (x−1 − x) = F .
(9) Let F = Q(α), where α3 = 5. Determine the irreducible
polynomial of α + α2 over Q.
(10) Consider f (x) = x4 + x3 + x2 + x + 1 as a polynomial over
Q. How many subfields are there of the splitting field E of
f (x) over Q? Justify your answer.
(11) Let E be a field containing exactly 64 elements. Find all the
subfields of E. How many elements α ∈ E satisfy E = Z2 (α)?
How many irreducible polynomials of degree 6 are there in
Z2 [x]?
(12) Let E be the splitting field of the function f (x) = x4 − 5 over
Q. Find the Galois group G of f (x) over Q, and list all the
subgroups of G and the corresponding fixed subfields.
(13) Let E ≤ L be a normal extension of fields, with Galois group
Gal(L/E) = {σ1 , · · · , σn }, and let α ∈ L. Show that L =
E(α) if and only if σ1 (α), · · · , σn (α) are distinct.
(14) Let p be a prime. Demonstrate the existence of a normal
extension of Q whose Galois group is cyclic with p elements.
(15) Let f (x) ∈ F [x]. Show that f (x) is irreducible over F if and
only if the action of Gal(f (x)/F ) on Rf (x) is transitive.
March 18, 2022 9:45 amsart-9x6 12819-main page 161

GALOIS THEORY 161

(16) Let f (x) be irreducible over Q, and let F be its splitting field
over Q. Show that if Gal(F/Q) is abelian, then F = Q(u)
for all roots u of f (x).
(17) Let K be a field of characteristic 0 in which every cubic poly-
nomial has a root. Let f (x) be an irreducible quartic poly-
nomial with coefficients in K whose discriminant is a square
in K. What is the Galois group of f (x) over K?
(18) Find the order of the Galois group of x5 − 2 over Q.
(19) Show that f (x) = x5 − 4x + 2 is irreducible over Q, and find
the number of real roots. Find the Galois group of f (x) over
Q, and explain why f (x) is not solvable by radicals
(20) Calculate the Galois group of x3 − 3x + 1 over Q.
(21) Let f (x) = x3 − 3x − 1 ∈ Q[x]. Show that Gal(f (x)/Q) =
(Z3 , +).
(22) Find infinitely many examples of polynomials of the form
x3 + 2ax + a over Q with Galois group S3 .
(23) Calculate the Galois group of x4 + 5x + 5 over Q.
(24) Calculate the Galois group of x4 + px + p over Q, where p is
a prime greater than 5.
(25) Show that the Galois group of f (x) = x4 + 3x2 + 2x + 1 over
Q is S4 . (Hint: Use Theorem 6.6.1.)
(26) Find infinitely many polynomials over Q with Galois group
S4 . (Hint: Use Theorem 6.6.1 and the fact that S4 is gener-
ated by a 4-cycle
p and a 3-cycle.)

(27) Show that Q( 2 + 2) is normal over Q with Galois group
isomorphic to the cyclic group (Z4 , +).
(28) Determine the Galois group of the polynomial x4 − 14x2 + 9
over Q.
(29) Find the Galois groups of x3 − 2 over the fields Z5 , Z7 and
Z11 .
(30) Find the Galois group of x4 − 1 over the field Z7 .
(31) Let F be a finite, normal extension of Q for which |Gal(F/Q)|
= 8 and each element of Gal(F/Q) has order 2. Find the
√ √ of F that have degree 4 over Q.
number of subfields
(32) Let F = Q( 2, 3 2). Find [F : Q] and prove that F is not
normal over Q.
(33) Find the Galois group of x9 − 1 over Q.
March 18, 2022 9:45 amsart-9x6 12819-main page 162

162 RING AND FIELD THEORY

(34) For any prime p relatively prime to n, prove that Φnp (x) =
Φn (xp )
Φn (x) over Q.
(35) Prove that Φ12 (x3 ) = Φ18 (x2 ) over Q.
(36) For any prime p, show that Φ12 (x) is reducible in Zp [x].
(37) Show that Φ10 (x) is irreducible in Z2 [x], but reducible in
Z5 [x].
(38) Over Q, show that
Y
Φd (x) = xn − 1.
d|n

n
(39) For any n ∈ Z+ , show that Φ2n+1 (x) = x2 + 1 over Q.
(40) For any m, n ∈ N, show that Φn (xm ) over Q is irreducible
over Q if and only if each prime factor of m is a factor of n.
2n
(41) Let n ∈ Z+ . Show that f (x) = x2 + x + 1 is irreducible
2
over Q. (Hints: Show that x + x − ω is irreducible over Q(ω)
where ω is a 2n+1 -th primitive root of unity.)
(42) Show that x4 − x3 + x2 − x + 1 is irreducible over Q, and use
it to find the Galois group of x10 − 1 over Q.
(43) Let f (x) = x5 − 5x2 + 1. Show f (x) has precisely three real
roots and is irreducible over Q. Let G = Gal(f (x)/Q), the
Galois group of f over Q. Show G contains a 5-cycle and a
2-cycle. What is the Galois group of f (x)? Is f (x) solvable
by radicals? explain.
(44) For any prime p, show that f (x) = x5 − p2 x + p is irreducible
over Q, and find the number of real roots of f (x). Find the
Galois group of f (x) over Q, and explain why the group is
not solvable. Consequently, f (x) is not solvable by radicals.
(45) For any prime p, show that f (x) = x5 −5p4 x+p is irreducible
over Q, and find the number of real roots of f (x). Find the
Galois group of f (x) over Q, and explain why the group is
not solvable. Consequently, f (x) is not solvable by radicals.
(46) For any primes q ≥ p with q ≥ 5, show that f (x) = xq −px+p
is not solvable by radicals over Q.
(47) Show that f (x) = 15x7 −84x5 −35x3 +420x+105 is irreducible
over Q, and find the number of real roots of f (x). Find the
Galois group of f (x) over Q, and explain why the group is
not solvable. Consequently, f (x) is not solvable by radicals.
March 18, 2022 9:45 amsart-9x6 12819-main page 163

GALOIS THEORY 163

(48) (Kaplansky’s Theorem). Let f (x) = x4 + ax2 + b ∈ Q[x] be


irreducible.
(a). If b is a square in Q then Gal(f (x)/Q) ' Z2 × Z2 .
(b). If b(a2 − 4b) is a square in Q then Gal(f (x/Q) ' Z4 .
(c). If neither b nor b(a2 −4b) is a square in Q then Gal(f (x/Q)
' D8 .
(Hints: You may assume that the roots of f (x) are ±α, ±β.)
(49) Show that Z[x1 , · · · , xn ]∩Q(s1 , · · · , sn ) = Z[s1 , · · · , sn ] where
si are defined in (6.2).
(50) Let F be a field, E = F (x1 , · · · , xn ), K = F (s1 , · · · , sn )
where si are defined in (6.2).
(a). Find [E : K], i.e., the dimension of the vector space E
over the field K.
(b). Find a basis for the vector space E over the field K.
(51) Let G be an arbitrary finite group. Show that there is a field
F and a polynomial f (x) ∈ F [x] such that the Galois group
of f (x) is isomorphic to G. (Hint: Use Theorem 6.5.2.)
(52) Let F ≤ E be a finite normal extension. Let Gal(E/F ) =
{ϕ1 , · · · , ϕn }. Define the trace and norm of α ∈ E as
Xn Yn
trE/F (α) = ϕi (α), NE/F (α) = ϕi (α).
i=1 i=1
Show that trE/F (α), NE/F (α) ∈ F .
(53) Let F ≤ E be a finite normal extension. Show that there is
α ∈ E such that trE/F (α) 6= 0.
(54) Let F ≤ E ≤ K be finite normal extensions. Show that
trK/F = trE/F ◦ trK/E , NK/F = NE/F ◦ NK/E .
B1948 Governing Asia

This page intentionally left blank

B1948_1-Aoki.indd 6 9/22/2014 4:24:57 PM


March 18, 2022 9:45 amsart-9x6 12819-main page 165

7. Sample Solutions

Chapter 1.

(1) Proof.
(a). Let a 6= 0. We want to show that a is not a 0-divisor. Note
that ϕ(a) is the unique element such that aϕ(a)a = a. Sup-
pose ac = 0 or ca = 0. Then a(ϕ(a)+c)a = aϕ(a)a+aca = a.
By uniqueness ϕ(a) + c = ϕ(a) so c = 0.
(b). From aϕ(a)a = a, we know that ϕ(a) 6= 0 also. Multiplying
on the left by ϕ(a) we obtain ϕ(a)aϕ(a)a = ϕ(a)a. Because R
has no divisors of zero by Part (a), multiplicative cancellation
is valid and we see that ϕ(a)aϕ(a) = ϕ(a).
(c). We claim that aϕ(a) is unity for nonzero a and ϕ(a) given
in the statement of the exercise. Let c ∈ R. From aϕ(a)a =
a, we see that ca = caϕ(a)a. Canceling a, we obtain c =
c(aϕ(a)). From Part (b), we have ϕ(a)c = ϕ(a)aϕ(a)c, and
cancelling ϕ(a) yields c = (aϕ(a))c. Thus aϕ(a) satisfies
(aϕ(a))c = c(aϕ(a)) = c for all c ∈ R, so aϕ(a) is unity.
(d). Let a be a nonzero element of the ring. By Part (a), aϕ(a)a =
a. Using cancellation we obtain that aϕ(a) = 1 and ϕ(a)a =
1. So ϕ(a) is the inverse of a, and a is a unit. This shows
that R is a division ring.

(3) Solution.
(a). We have y = y 6 = (−y)6 = −y, hence 2y = 0 for any y ∈ R.
Now let x be an arbitrary element in R. Using the binomial
formula, we obtain

x + 1 = (x + 1)6
= x6 + 6x5 + 15x4 + 20x3 + 15x2 + 6x + 1
= x4 + x2 + x + 1,
165
March 18, 2022 9:45 amsart-9x6 12819-main page 166

166 RING AND FIELD THEORY

where we canceled the terms that had even coefficients. Hence


x4 + x2 = 0, or x4 = −x2 = x2 . We then have
x = x6 = x2 x4 = x2 x2 = x4 = x2 ,
and so x2 = x, as desired.
(b). Expanding the equality (x + y)2 = x + y we deduce xy + yx =
0, so xy = −yx = yx for any x, y ∈ R. This shows that the
ring is commutative, as desired.

(5) Proof. Let G = R \ {0}. We need to prove that (G, ·) is a
group, i.e., the identity axiom and the inverse element axiom hold
for (G, ·).
Let a, b, c ∈ G. Because of the cancellation rules we see that,
ab = ac iff b = c. Thus we see that aG = G. Since a ∈ G = aG,
there exists ea ∈ G so that aea = a.
From this we also have aea a = aa. By the cancellation rules we
have ea a = a.
From ab = aea b and the cancellation rules we have b = ea b. Thus
e = ea is the identity element in G.
Since e ∈ G = aG = Ga, there exists x, y ∈ G so that ax = ya = e.
Then x = ex = yax = ye = y is the inverse of a. Therefore G is a
group. 
(6) Proof. We need only to show that “+” is commutative.
Denote by 1 the identity element of hS ∗ , ·i. For any x ∈ S, using
(iii) we have 0x = x0 = 0 and x1 = 1x = x.
For any x, y ∈ S, using (iii) we have
(1 + x)(1 + y) = 1 + y + x + xy, and (1 + x)(1 + y) = 1 + y + x + xy.
Thus x + y = y + x, i.e., “+” is commutative.
Therefore, hS, +, ·i is a division ring. 
(8) Solution. The Polynomials of degree 3 in Z2 [x] are
x3 : not irreducible because 0 is a zero,
x3 + 1: not irreducible because 1 is a zero,
x3 + x: not irreducible because 0 is a zero,
x3 + x2 : not irreducible because 0 is a zero,
x3 + x + 1: irreducible, neither 0 nor 1 is a zero,
x3 + x2 + 1: irreducible, neither 0 nor 1 is a zero,
x3 + x2 + x: not irreducible because 0 is a zero,
x3 + x2 + x + 1: not irreducible, 1 is a zero.
Thus the irreducible cubics are x3 + x + 1 and x3 + x2 + 1. 
March 18, 2022 9:45 amsart-9x6 12819-main page 167

SAMPLE SOLUTIONS 167

Solution 2. An irreducible polynomials of degree 3 in Z2 [x] has


to be of the form f (x) = x3 + ax2 + bx + 1 for some a, b ∈ Z2 . We
know that f is irreducible iff f (0)f (1) 6= 0, iff a + b 6= 0, iff a = 1 and
b = 0, or a = 0 and b = 1. Thus the irreducible cubics are x3 + x + 1
and x3 + x2 + 1. 
(10) Solution. First assume that 1 − ab is invertible, and let
x = (1 − ab)−1 . Then x(1 − ab) = 1, so bx(1 − ab)a = ba, and
therefore (1 − ba) + bx(1 − ab)a = 1. Now

1 = (1−ba)+bx(a−aba) = (1−ba)+bxa(1−ba) = (1+bxa)(1−ba).

It can be checked easily that (1 − ba)(1 + bxa) = 1, so (1 − ba)−1 =


(1 + bxa). A similar argument shows that if 1 − ba is invertible, then
so is 1 − ab. 
(11) Solution. We know that all possible rational zeros of the
polynomial are ±1, ±2, ±1/2. By direct verification we see that the
only rational zeros of the polynomial are −2, 1/2. By simple compu-
tations we see that 2x4 + 3x3 + 3x − 2 = (x + 2)(2x − 1)(x2 + 1).
Thus all real roots of the polynomial are −2, 1/2. 
(12) Solution. f (x) = x(x4 −2x2 +x+2) = x(x+1)(x3 −x2 −x+2).
It is easy to check that x3 − x2 − x + 2 does not have any solutions in
Z5 . Thus it is irreducible. So the irreducible factorization is f (x) =
x(x + 1)(x3 − x2 − x + 2). 
(13) Solution. It is irreducible. If x3 +3x2 −8 is reducible over Q,
then it factors in Z[x], and must therefore have a linear factor of the
form x−a in Z[x]. Then a must be a zero of the polynomial and must
divide −8, so the possibilities are a = ±1, ±2, ±4, ±8. Computing
the polynomial at these eight values, we find none of them is a zero
of the polynomial, which is therefore irreducible over Q. 
(14) Proof. It is clear that φ is a ring homomorphism. Similarly
ψ : R → R defined by ψ(f (x)) = f (a−1 x − a−1 b) is also a ring
homomorphism. It is easy to verify that ψ = φ−1 . Thus φ is an
automorphism of R[x]. 
(15) Proof. Since hn, xi = Z[x]x + nZ and hxi = Z[x]x, we see
that

Z[x] Z[x]x + nZ
' Z, ' nZ.
Z[x]x Z[x]x
March 18, 2022 9:45 amsart-9x6 12819-main page 168

168 RING AND FIELD THEORY

Then
Z[x]
Z[x]x Z
Z[x]/hn, xi ' ' ' Zn .
Z[x]x+nZ nZ
Z[x]x

So hn, xi is a prime ideal iff Zn is an integral domain, which occurs


iff n is a prime number. 
(18) Proof. We first note that R is an integral domain since the
zero ideal is prime. Let a be a nonzero element of R. If a2 R = R,
we see that a ∈ U(R). Otherwise, by assumption, the ideal a2 R is
prime, and so a2 ∈ a2 R implies a ∈ a2 R. Thus a = a2 r for some
r ∈ R, and since R is an integral domain, we can cancel a to obtain
1 = ar, showing that a is invertible, which is impossible. So R is a
field. 
(26) Solution. In Z5 [x] we have x2 − 3 = x2 + 2. Then R1 = R2 .
In Z2 [x] we have R1 = Z2 [x]/((x + 1)2 ), R2 = Z2 [x]/(x2 ). Then
R1 ' R2 by the map f : R1 → R2 , f (ax + b) = a(x − 1) + b.
In Z11 [x] since (2x)2 − 3 = 4(x2 + 2) then R1 ' R2 by the map
f : R1 → R2 , f (ax + b) = 2ax + b. 
(33) Solution.

(a). Let a, b ∈ A. Then am ∈ A and bn ∈ N for some positive
integers m and n. In a commutative ring, the binomial expan-
sion is valid. Consider (a+b)m+n . In the binomial expansion,
each summand contains a term ai bm+n−i . Now either i > m
so that ai ∈ A or m + n − i > n so that bm+n−i ∈ A. Thus
each summand of (a + b)√m+n is in A, so (a + b)m+n ∈ A and

a + b ∈ A. Also sa ∈ A since (as)m ,√ (sa)m ∈ A for any
s ∈ R. Because 01 ∈ A, we see that 0 ∈ A. Also (−a)m is
either √am or −am , and both√am and −(am ) are in A. Thus
−a ∈ √A. This √ shows that A is an ideal √ of R. √
(b). Since A + B = R, there exist a ∈ A and b ∈ B such
that a + b = 1. There exists an integer n such that an ∈ A
and bn ∈ B. Then
n Ç å n Ç å
2n
X 2n 2n−i i X 2n
1 = (a + b) = a b + an−i bn+i
i n+i
i=0 i=1
n Ç å n Ç å
n
X 2n n−i i n
X 2n
=a a b +b an−i bi .
i n+i
i=0 i=1
March 18, 2022 9:45 amsart-9x6 12819-main page 169

SAMPLE SOLUTIONS 169

Clearly x = an ni=0 2n b ∈ A and y = bn ni=1 2n 


P  n−i i P
i a n+i
an−i bi ∈ B. From x + y = 1 we see that A + B = R.

(35) Solution. We know that E = 2Z. Take M = 4Z. Then
M  E, and E/M = {0 + 4Z, 2 + 4Z}. It is easy to see that M is a
maximal ideal. Since (2 + 4Z)2 = 0 therefore E/M is not a field. 
(36) Solution. Let I be the set of all prime ideals of R, P =
∩I∈I I, N is the set of all nilpotent elements of R. We will show that
N = P.
In an example in class we know that N ⊂ I for any I ∈ I, yielding
that N ⊂ P .
It is enough to show that for any a ∈ R \ N there is I ∈ I with
a∈ / I. Let A = {ak : k ∈ Z+ }. Using Kuratowski-Zorn Lemma there
is an ideal M  R such that M is maximal A ∩ M = {0}. We shall
show that M is prime.
Let x, y ∈ R \ M such that xy ∈ M . We will try to deduce some
contradictions. Consider the ideals M + Rx and M + Ry which
properly contain M . Then (M + Rx) ∩ A 6= {0} and (M + Ry) ∩ A 6=
{0}. Thus there are k1 , k2 ∈ Z+ , m1 , m2 ∈ M, r1 , r2 ∈ R such that
ak1 = m1 + r1 x, ak2 = m2 + r2 y.
We deduce that
ak1 +k2 = (m1 + r1 x)(m2 + r2 y) ∈ M
a contradiction. Thus M is prime. 
(37) Solution. Let I be a maximal ideal of C[x]. Then we know
that I = hf (x)i for some irreducible f (x) ∈ C[x]. We know that all
irreducible polynomials are of degree 1. Then f (x) = ax − b for some
a, b ∈ C with a 6= 0. So I = hax + bi = hx − ci for some c ∈ C. 
4 3 2 8 7 5 4 3
(55) Answer. (x + x + x + x + 1)(x − x + x − x + x − x + 1).

Chapter 2.

(2) Proof. Let a, b ∈ D∗ that p|ab. Then there is c ∈ D such that


ab = pc. We may assume that
a = p1 p2 · · · ps1 , b = q1 q2 · · · qs2 , c = r1 r2 · · · rs3 ,
where pi , qj , rk are irreducibles in D. Then
p1 p2 · · · ps1 q1 q2 · · · qs2 = pr1 r2 · · · rs3 .
March 18, 2022 9:45 amsart-9x6 12819-main page 170

170 RING AND FIELD THEORY

Since D is an UFD, we see that p ∼ pi or p ∼ qj . Thus p|a or p|b,


i.e., p is prime. 
3
(4) Answer. x + 2x − 1.
(8) Solution. Since 15−12i 150 3i
6−5i = 61 + 61 , we have 15 − 12i = 2(6 −
5i) − (3 − 2i). Since 6−5i 28 3i
3−2i = 13 − 13 , we have 6 − 5i = 2(3 − 2i) − i.
We put them together
15 − 12i = 2(6 − 5i) − (3 − 2i)
6 − 5i = 2(3 − 2i) − i
3 − 2i = i(−2 − 3i) + 0.
Thus gcd(15 − 12i, 6 − 5i) ∼ 1.
Since 16+7i 6
10−5i = 1 + 5 i, we have 16 + 7i = (10 − 5i)(1 + i) + (1 + 2i).
Since 10−5i
1+2i = −5i, we have 10 − 5i = (1 + 2i)(−5i). We put them
together
16 + 7i = (10 − 5i)(1 + i) + (1 + 2i)
10 − 5i = (1 + 2i)(−5i) + 0.
Thus we have gcd(16 + 7i, 10 − 5i) ∼ 1 + 2i. √ 
√(13) Solution. We work on the UFD Z[ −2][x, y]. Let √ v(a +
b −2) = a2 + 2b2 be the standard
√ multiplicative
√ norm on Z[ −2].
Write the equation as (x + −2)(x − −2) = y 3 .
√ √
(a). Since v( −2)√ = 2, by Theorem 2.5.6 we know that −2 is
prime in Z[ −2]. √ √ √
(b). Let√p be a prime
√ in Z[ −2], p|x+ −2 and √ p|x− −2. √ Then
p|2 −2 = ( −2)3 . We know that p = ± −2. Then −2|x,
and 2|x2 , and furthermore 4|x2 , which is impossible√from
x2 + √
2 = y 3 (we would have 23 |y 3 = x2 + 2). So (x + −2),
(x − −2) √ are relatively
√ prime.
(c). Since
√ (x + √ −2), (x − −2) are relatively
√ prime, 3from (x +
−2)(x√− −2) = y 3 we know √ that x + −2 = y1 for some
y1 ∈ Z[ −2]. Let y1 = a + b −2. We have
√ √ √
x + −2 = (a + b −2)3 = a3 − 6ab2 + (3a2 b − 2b3 ) −2,
yielding that
1 = 3a2 b − 2b3 = b(3a2 − 2), x = a3 − 6ab2 .
We obtain that b = 1 and a = ±1, hence x = ±5 and y = 3.

March 18, 2022 9:45 amsart-9x6 12819-main page 171

SAMPLE SOLUTIONS 171

(16) Solution. Since R is a PID, we may suppose that I = hai,


J = hbi for some nonzero a, b ∈ D. Let d = gcd(a, b). Then IJ =
habi, I + J = {ax + by|a, b ∈ R} = hdi.
It is clear that hab/di ⊂ I ∩ J.
Let z ∈ D. Then z ∈ I ∩ J iff z = ax = by for some x, y ∈ R, iff
a|by, ad | db y, iff ad |y, iff b ad |by iff b ad |z. Thus z ∈ hab/di, i.e., I ∩ J ⊆
hab/di, and further I ∩ J = hab/di.
We see that IJ = I ∩ J iff hab/di = habi iff d is invertile, iff
I + J = R. 
(17) Solution.
(a). We see that R = {a0 +a2 x2 +a3 x3 +· · · an xn |n ∈ Z+ , ai ∈ F }.
Clearly R is a subring of F [x].
(b). Suppose that x2 = f g for some f, g ∈ R. By comparing the
degree of f and g we deduce that deg(f ) = 0 or deg(g) = 0,
that is, f or g is in U(R) = F ∗ . So x2 is irreducible in R.
Similarly, x3 is irreducible in R as well.
(c). From (x2 )3 = (x3 )2 , we deduce that x2 |(x3 )2 , but x2 6 |x3 ,
and that x2 |(x3 )2 , but x2 6 |x3 in R. We know that both x2
and x3 are irreducibles but not primes. So R is not a UFD.
(d). We claim that the ideal hx2 , x3 i of R is not a principal ideal.
Otherwise suppose that hx2 , x3 i = hf (x)i for some f ∈ R.
It is clear that hx2 , x3 i 6= R. So f is not a unit. There
are u, v ∈ R such that x2 = uf, x3 = vf . Since x2 , x3 are
irreducible, we deduce that u, v ∈ U(R). Thus x2 ∼ x3 which
is impossible. Our claim follows.

(18) Solution. Let D = C[x] which is clearly a UFD. We see that
C[x, y] = D[y].
Note that y 5 + xy 4 − y 4 + x2 y 2 − 2xy 2 + y 2 + x3 − 1 = y 5 +
(x − 1)y 4 + (x − 1)2 y 2 + x3 − 1. Clearly x − 1 is irreducible in D.
Taking p = x − 1 in Schönemann-Eisenstein Criterion we know that
y 5 + xy 4 − y 4 + x2 y 2 − 2xy 2 + y 2 + x3 − 1 is irreducible in C[x, y].
Let f (x, y) = xy 3 + x2 y 2 − x5 y + x2 + 1 which has degree 3 in
y. Consider y 3 f (x, 1/y) = x + x2 y − x5 y 2 + (x2 + 1)y 3 . Taking
p = x in Schönemann-Eisenstein Criterion we know that y 3 f (x, 1/y)
is irreducible in the integral domain C[x, y]. Thus xy 3 + x2 y 2 − x5 y +
x2 + 1 is irreducible in the integral domain C[x, y]. √ 
(21) Solution. Let d = 2, 3. Note that if we define v(a + b d) =
a2 − db2 , v is not a norm (since it can be negative). So we define
March 18, 2022 9:45 amsart-9x6 12819-main page 172

172 RING AND FIELD THEORY



v(a + b d) = |a2 − db2 |. This is clearly a non-negative integer.
Moreover, since d is not a square of an integer, a2 − db2 = 0 if
a = b = 0. So v(α) > 0 if α 6= 0.
√ to see that v(αβ) = v(α)v(β) ≤√v(α)v(β) for nonzero
It is easy
α, β ∈ Z[ d]. We can obviously
√ extend v to Q[ d] such that v(αβ) =
v(α)v(β) for α, β ∈ Q[√ d]. √
For any α, β ∈ Z[ d] with√ β 6= 0, let α/β = a + b d where
a, b ∈ Q. Take q = x + y d where √ x, y ∈ Z with |a − x| ≤ 1/2,
|b − y| ≤ 1/2. Let r = α − qβ ∈ Z[ d]. We know that α = qβ + r,
and
v(r) = v(α − qβ) = v(β)v(α/β − q)
= v(β)|(a − x)2 − (b − y)2 d| ≤ 3v(β)/4 < v(β).

(24) Solution. Suppose that f (x) = g(x)h(x) for some g(x),
h(x) ∈ Z[x] with deg(f (x)) > 0 and deg(g(x)) > 0. We may further
assume that g(x), h(x) are monic. We see that g(ai )h(ai ) = −1, and
furthermore g(ai ) + h(ai ) = 0. Since deg(g(x) + h(x)) ≤ n − 1, we
deduce that g(x) + h(x) = 0, i.e., h(x) = −g(x) which is impossible.

(25) Solution. Suppose that f (x) = g(x)h(x) for some g(x),
h(x) ∈ Z[x] with deg(f (x)) > 0 and deg(g(x)) > 0. We may further
assume that g(x), h(x) are monic. Both g(x) and h(x) do not have
any real zeros. So g(ai ) > 0 and h(ai ) > 0 for all i. We see that
g(ai )h(ai ) = 1, and furthermore g(ai ) = h(ai ) = 1 for all i. If
deg(g(x)) < n, we deduce that g(x) = 1 which is impossible. If
deg(h(x)) < n, we deduce that h(x) = 1 which is impossible. Now we
come to the case that deg(g(x)) = deg(h(x)) = n. Then deg(g(x) −
h(x)) < n, yielding that g(x) − h(x) = 0 since g(ai ) − h(ai ) = 0 for
all i. You see the contradictions. 
(27) Solution. Suppose that g(x) = g1 (x)g2 (x) for some g1 (x),
g2 (x) ∈ Z[x] with deg(g1 (x)) > 0 and deg(g2 (x)) > 0. Using the
above exercise we deduce that g1 (ai ) = g2 (ai ) = 1 for all i, or g1 (ai ) =
g2 (ai ) = −1 for all i. We may assume that g1 (ai ) = g2 (ai ) = 1 for all
i. If deg(g1 (x)) < n, we deduce that g1 (x) = 1 which is impossible.
Similarly the case that deg(g2 (x)) < n does not occur. Now we
come to the case that deg(g1 (x)) = deg(g2 (x)) = n. We see that
g1 (x) = b1 f (x) + 1 and g2 (x) = b2 f (x) + 1 for some b1 , b2 ∈ Z. You
see the contradictions. 
March 18, 2022 9:45 amsart-9x6 12819-main page 173

SAMPLE SOLUTIONS 173

Chapter 3.
 
1 0 0 0
 0 3 0 0 
(7) Answer. B ∼   0 0 21 0 .

0 0 0 0
(6) Answer. 2,  2, 156. 
1 0 0
(8) Answer.  0 x − 1 0 .
0 0 (x − 1)(x − 2)
(9) Answer. 1, 1, 2x2 + 3x.
(15) Proof. Suppose that ϕ̃1 , ϕ̃2 are two such extensions. Con-
sider the module homomorphism ϕ̃1 − ϕ̃2 : M → N . We see that
(ϕ̃1 − ϕ̃2 )(S) = 0. So S ⊆ ker(ϕ̃1 − ϕ̃2 ) ≤ M . Then M = hSi ⊆
ker(ϕ̃1 − ϕ̃2 ) too. So ϕ̃1 − ϕ̃2 = 0, i.e., ϕ̃1 = ϕ̃2 . Therefore ϕ̃ : M → N
is uniquely determined by the map ϕ. 
(18) Proof. (a). The statement that f is one-to-one simply says
C ∩ D = 0, and the statement that g is one-to-one says the same
thing. Suppose f is onto.
(b). To prove g onto we take e ∈ E. Write e = b + c according
to the decomposition B ⊕ C. Since f is onto, there exists d ∈ D
with f (d) = b. Write the decomposition of d as d = b + c1 . Then
c − c1 = −d + e, so that g(c − c1 ) = e as required. 
(22) Hints: Show that every maximal ideal in Z[x] is of the form
(p, f (x)) where p is prime integer and f (x) is primitive integer poly-
nomial that is irreducible modulo p.

Chapter 4.

(12) Solution. Let q = pn where p is a prime. The polynomial


x2 + x + 1 reducible over F, iff there is a ∈ F such that a2 + a + 1 = 0.
This is true for q = 3. Now we assume that q 6= 3.
Then the polynomial x2 + x + 1 reducible over F , iff there is a ∈
F \ {1} such that a3 − 1 = 0, iff 3|φ(q) = pn−1 (p − 1) since (F ∗ , ·)

is cyclic of order φ(q), iff q = 3n , or q = pn with 3|p − 1. 


(15) Proof. Let F be a finite field of pn elements. Let m be a
divisor of n, so that n = mq for some q ∈ N. Then we know that
n
the equation xp −1 = 1 has exactly pn − 1 distinct zeros in F . Let
March 18, 2022 9:45 amsart-9x6 12819-main page 174

174 RING AND FIELD THEORY

pn − 1 = (pm − 1)k, we see that k ≡ 1 (mod p). Then we know that


n −1 m −1 m −1 m −1 m −1)(k−1)
xp −1 = (xp )k −1 = (xp −1)(1+xp +· · ·+x(p )
for all a ∈ F . Since gcd(x pm −1
−1, 1+x pm −1
+· · ·+x (pm −1)(k−1)
) = 1,
m
then there are exactly pm − 1 zeros of xp −1 − 1 in F , i.e., all the pm
m
zeros of xp − x in F form a field of order pm . This is the only such
m
field since any such field satisfies xp − x = 0. 
(21) Solution. Suppose K ≤ F . Then Z13 ≤ K ≤ F , and we
have 11 = [F : Z13 ] = [F : K][K : Z13 ]. We see that [K : Z13 ]|11.
Then [K : Z13 ] = 1 or 11. There are exactly two subfield of F : F
and Z13 . 
(22) Solution. Let r = deg(α, Z13 ), K = Z13 (α) ≤ F . Then
r = [K : Z13 ]. From 11 = [F : Z13 ] = [F : K][K : Z13 ] = r[F : K],
we see that deg(α, Z13 )|11. Then deg(α, Z13 ) = 1 or 11. 
(23) Solution. Let f (x) be an irreducible polynomial of degree 11
over Z13 and E be the splitting field of f (x). Since there is only one
field F1311 of order 1311 we know that [E : Z13 ] = 11 and E ' F1311 .
Thus all 11 distinct zeros of f (x) are in E. There are 1311 − 13
elements in E that has degree 11 over Z13 . So the number of distinct
irreducible polynomials of degree 11 over Z13 is
1311 − 13
.
11

(24) Proof. We know that F = Fpn . Let
rd
YY
g(x) = fd,j (x).
d|n j=1
n
We will show that xp − x = g(x). It is clear that if Zp ≤ K ≤ F
then |K| = pd for some d|n. Each α ∈ F is algebraic over Zp and has
degree d that divides n. Thus irr(α, Zp ) = fd,j (x) for a unique d|n
and for a unique j = 1, 2, · · · , rd , and each zero of fd,j (x) can generate
the same extension field of Zp since they have the same degree over
n
Zp . Note that fd,j (x)|xp − x. Since F is finite, fd,j (x) has exactly
d distinct zeros in F . So each α ∈ F has a unique fd,j (x) as its
irreducible polynomial. Also each fd,j (x) has exactly d zeros in F .
n
We see that F consists of all zeros of g(x). Therefore xp − x = g(x).

2 3 3 3
(30) Answer. irr(e − 2, Q(e )) = x + 6x + 12x + 8 − e . 6
March 18, 2022 9:45 amsart-9x6 12819-main page 175

SAMPLE SOLUTIONS 175

(32) Proof. (a2 + b2 )(c2 + d2 ) = N (a + bi)N (c + di) = N ((a +


bi)(c + di)) = N (ac − bd) + (ad + bc)i) = (ac − bd)2 + (ad + bc)2 . 
(33) Solution. Suppose p is not irreducible in Z[i], and p =
(a + bi)(c + di) where a, b, c, d ∈ Z, and both a + bi and c + di are not
units. Then p2 = (a2 +b2 )(c2 +d2 ). Since (a2 +b2 ) 6= 1 6= (c2 +d2 ), we
know that p = a2 +b2 = c2 +d2 which is not true since p ≡ 3(mod 4).
So p is irreducible in Z[i]. 
3 2
(36) Solution. Note that 1146600 = 13 ∗ 2 ∗ 3 ∗ 5 ∗ 7 . By 2 2

Theorem 4.4.10 we can write the integer as a sum of two integer


squares. Actually
1146600 = 2102 + 10502 .

(38) Proof. If π were algebraic, πi would be algebraic as well. By
Theorem 4.3.13, then eπi = −1 would be transcendental, which is a
contradiction. Therefore π is not algebraic, which means that it is
transcendental. 

Chapter 5.

(2) Solution.
(a). We have the factorization
x4 + 4 = (x2 + 2x + 2)(x2 − 2x + 2),
where the factors are irreducible by Schönemann-Eisenstein
Criterion (p = 2). The roots are ±1 ± i, so the splitting field
is Q(i), which has degree 2 over Q.
An alternate solution is to solve x√4 = −4. To find one
root, use DeMoivre’s theorem to get 4 −4 = ±1 ± i.
(b). So the splitting field is Q(i), which has degree 2 over Q.
(c). The Galois group Gal(Q(i)/Q) must be cyclic of order 2,
which is generated by the conjugation automorphism
φ : Q(i) → Q(i), a + bi 7→ a − bi, ∀a, b ∈ Q.

(3) Solution. Since F = K(u, v) ⊇ K(u) ⊇ K, where [K(u) :
K] = m and [K(u, v) : K(u)] ≤ n, we have [F : K] ≤ mn. But
[K(u) : K] = m and [K(v) : K] = n are divisors of [F : K], and
since gcd(m; n) = 1, we must have mn|[F : K]. So [F : K] = mn. 
March 18, 2022 9:45 amsart-9x6 12819-main page 176

176 RING AND FIELD THEORY

(4) Solution. Since u2 ∈ K(u), we have K(u) ⊇ K(u2 ) ⊇ K.


Suppose that u ∈ / K(u2 ). Then x2 −u2 is irreducible over K(u2 ) since
it has no roots in K(u2 ), so u is a root of the irreducible polynomial
x2 − u2 over K(u2 ). Thus [K(u) : K(u2 )] = 2, and therefore 2 is a
factor of [K(u) : K]. This contracts the assumption that [K(u) : K]
is odd. So u ∈ / K(u2 ) and hence K(u2 ) = K(u). 
(5) Solution. It is clear that the polynomial x3 −√ 11 is irreducible

over
√ the field Q. The roots of the polynomial are 3 11, a 3 11 and
a2 3 11, where a is a primitive
√ cube root of unity. Since a is not
real, it cannot belong to Q( 3 11).
√ Since a is a root √
of the irreducible
2
polynomial x + √ x + 1 over Q( 11) and F = Q( 3 11, a), we have
3

3

3
[F : Q] = [F : Q( 11)][Q( 11) : Q] = 2 · 3 = 6. 
(6) Solution. We know that Gal(F/Q) has odd order. If u is
a nonreal root of f (x), then since f (x) has rational coefficients, its
conjugate u must also be a root of f (x). It follows that F is closed
under taking complex conjugates. Since complex conjugation defines
an automorphism of the complex numbers, it follows that restricting
the automorphism to F defines a homomorphism from F into F.
Because F has finite degree over Q, the homomorphism must be
onto as well as one-to-one. Thus complex conjugation defines an
element of the Galois group of order 2, and this contradicts the fact
that the Galois group has odd order. We conclude that every root of
f (x) must be real. 
(18) Solution.

(a). We have x8 −1 = (x4 −1)(x4 +1) = (x−1)(x+1)(x2 +1)(x4 +


1), giving the factorization over Q. The factor Φ8 (x) = x4 +1
is irreducible over Q. The roots of x4 +1 are thus the primitive
8th roots of unity, ±1±i
√ , and adjoining one of these roots also
2
gives the others, together with i. Thus the splitting field is
obtained in one step, by adjoining one root of x4 + 1, so its
degree over Q is 4.
It is clear that
√ the splitting field can also
√ be obtained by
adjoining first 2 and√then i, so F = Q(i, 2).
(b). These subfields of Q( 2, i) are the splitting fields of x2 −
2, x2 + 1, and x2 + 2, respectively. Any automorphism √ must
take roots to roots,
√ so if
√ θ is an automorphism of Q( 2, i), we
must have θ( 2) = ± 2, and θ(i) = ±i. These possibilities
must in fact define 4 automorphisms of the splitting field.
March 18, 2022 9:45 amsart-9x6 12819-main page 177

SAMPLE SOLUTIONS 177

Since all these automorphisms are of order 2, so the Galois


group is Z2 × Z2 .


Chapter 6.

(7) Proof. It is clear that F ≤ F (x2 ) ∩ F (x2 − x). We need to


show that F (x2 ) ∩ F (x2 − x) ≤ F . To the contrary, suppose not.
Then there is a positive degree f (x) ∈ F (x2 ) ∩ F (x2 − x). We have
the subfields
F ≤ F (f (x)) ≤ F (x2 ) ∩ F (x2 − x) ≤ F (x).
We see that F (x) is a finite algebraic extension of F (f (x)). So any
automorphism in the group Gal(F (x)/F (f (x))) is of finite order.
Consider the following automorphisms τ1 ∈ Gal(F (x)/F (x2 )), τ2 ∈
Gal(F (x)/F (x2 − x) defined by
τ1 (x) = −x, τ2 (x) = 1 − x.
Clearly τ1 τ2 ∈ Gal(F (x)/F (f (x))). But τ1 τ2 (x) = x + 1. Conse-
quently, τ1 τ2 is of infinite order, which is impossible. So F (x2 ) ∩
F (x2 − x) = F . 
(9) Answer. irr(α + α2 , Q) = x3 − 15x − 20.
(16) Solution. Since F has characteristic zero, we know that F
is a normal extension of Q. So we can use Galois theorem. Because
Gal(F/Q) is abelian, every subgroup is normal, and every interme-
diate extension between Q and F must be normal. Therefore if we
adjoin to Q any root u of f (x), the extension Q(u) must contain all
other roots of f (x), since it is normal over Q. Thus Q(u) is a splitting
field for f (x) over Q, so Q(u) = F. 
(18) Solution. Let G be the Galois group of x5 − 2, and let
ω be√a primitive 5th root of unity. Then the roots of x5 − 2 are
b = 5 2 and ω j b, for 1 ≤ j ≤ 4. The splitting field of x5 − 2 over
Q is F = Q(ω, b). Since p(x) = x5 − 2 is irreducible over Q by
Schönemann-Eisenstein Criterion, it is the irreducible polynomial of
b. The element ω is a root of x5 − 1 = (x − 1)(x4 + x3 + x2 + x + 1),
so irr(ω, Q) = x4 + x3 + x2 + x + 1. Thus [F : Q] ≤ 20. Since
[Q(ω, b) : Q(ω)] = 4, [Q(ω, b) : Q(b)] = 5,
the degree [F : Q] must be divisible by 5 and 4. It follows that
[F : Q] ≥ 20. Therefore |G| = 20. 
March 18, 2022 9:45 amsart-9x6 12819-main page 178

178 RING AND FIELD THEORY

y
10

x
−1.5 −1 −0.5 0.5 1 1.5

−5

Figure 2. y = x5 − 4x + 2

(19) Solution. The polynomial f (x) is irreducible over Q since


it satisfies Schönemann-Eisenstein Criterion for p = 2. Consider
y = f (x) as a continuous real»valued function. The derivative f 0 (x)
»=
4 4 4 0
5x − 4 has two real roots ± 5 . We see that f (x) > 0 if |x| > 4 45 ,
»
and f 0 (x) < 0 if |x| < 4 45 . That is, f (x) is strictly increasing
» »
for |x| > 4 45 , and f (x) is strictly decreasing for |x| < 4 45 . Since
 »  » 
f (−∞) = −∞, f − 4 45 > f (−1) = 5 > 0, f 4 45 < f (1) =
−1 < 0, f (∞) = ∞, i.e.,
» »
x −∞ % − 4 45 % 4 4
5 % ∞
f 0 (x) ∞ > 0 0 <0 0 >0 ∞
f (x) −∞ % >0 & <0 % ∞
we see that f (x) must have exactly three real roots. It follows from
a theorem that the Galois group of f (x) over Q is S5 which is not
solvable. Thus f (x) is not solvable by radicals. See Figure 2. 
(36) Hints: Show that Gal(Φ12 (x)/Zp ) is not cyclic, i.e., the unit
group U (Z12 , +, ·) is not cyclic.
March 18, 2022 9:45 amsart-9x6 12819-main page 179

Appendix A. Equivalence Relations and


Kuratowski-Zorn Lemma

In this appendix we mainly recall some concepts and results from Set
Theory, see [L] for details.
Definition A.0.1. Let A, B be nonempty sets. The Cartesian
product of A and B is the set
A × B = {(a, b) : a ∈ A, b ∈ B}.
Definition A.0.2. Let S be a nonempty set. A binary relation
(or a relation) R in S is a subset of S × S. We usually write aRb if
(a, b) ∈ R.
Definition A.0.3. Let R be a relation in a nonempty set S. Then
(1) R is called reflexive if (x, x) ∈ R for all x ∈ S;
(2) R is called symmetric if (x, y) ∈ R ⇒ (y, x) ∈ R;
(3) R is called anti-symmetric if (x, y) ∈ R and (y, x) ∈ R ⇒
x = y;
(4) R is called transitive if (x, y) ∈ R and (y, z) ∈ G ⇒ (x, z) ∈
R.
Definition A.0.4. A relation in a nonempty set S is called an
equivalence relation if it is reflexive, symmetric, and transitive.
Let R be an equivalence relation in a nonempty set S. If (x, y) ∈ R,
we will write simply x ∼ y or x ≡ y (mod R) and say that x is
equivalent to y. Noting that R is an equivalence relation in S, for
any x, y, z ∈ S we have
(1) x ∼ x,
(2) x ∼ y ⇒ y ∼ x,
(3) x ∼ y and y ∼ z ⇒ x ∼ z.
Definition A.0.5. Let S be a nonempty set and let R be an equiv-
alence relation in S. If x ∈ S, then the equivalence class of x
modulo R is defined as follows:
[x]R = {y ∈ S : y ∼ x}.
179
March 18, 2022 9:45 amsart-9x6 12819-main page 180

180 RING AND FIELD THEORY

The collection of all the equivalence classes modulo R:


S/R = {[x]R : x ∈ S}
is call the quotient set of S modulo R.
Definition A.0.6. (1) A nonempty set S is said to be partially
ordered if a given binary relation ≤ in S satisfies:
(a) a ≤ a, for any a ∈ S (reflexive law),
(b) a ≤ b, b ≤ c ⇒ a ≤ c, for any a, b, c ∈ S (transitive
law),
(c) a ≤ b, b ≤ a⇒ a = b, for any a, b ∈ S (antisymmetric
law).
(2) A partially ordered set S is said to be totally ordered if for
every pair a, b ∈ S we have either a ≤ b or b ≤ a.
(3) Let S be a partially ordered set. An elements x ∈ S is called a
maximal element if x ≤ y with y ∈ S ⇒ x = y. Similarly,
we can define a minimal element of S.
(4) Let T be a totally ordered subset of a partially ordered set S.
We say that T has an upper bound in S if there exists c ∈ S
such that x ≤ c for all x ∈ T.
(5) A totally ordered set S is well-ordered if for every nonempty
subset X ⊆ S, there exists x ∈ X satisfying y ≥ x for all
y ∈ X.
Theorem A.0.7 (Kuratowski-Zorn Lemma). Let S be a partially
ordered set. If every totally ordered subset of S has an upper bound
in S, then S contains a maximal element.
Kuratowski-Zorn Lemma is also known as Zorn’s Lemma. It was
proved by Kazimierz Kuratowski (1896–1980) in 1922 and indepen-
dently by Max Zorn (1906–1993) in 1935. Kuratowski-Zorn lemma
is widely used in many situations.
Theorem A.0.8 (The Well Ordering Principle). Any nonempty set
S can be well-ordered, that is, there is a well-ordering on S.
Theorem A.0.9 (The Axiom of Choice). Given a class of nonempty
sets, there exists a “choice function”, i.e., a function which assigns to
each of these sets one of its elements.
In Set Theory, Axiom of Choice is logically equivalent to
Kuratowski-Zorn Lemma which is logically equivalent to the Well
Ordering Principle.
March 18, 2022 9:45 amsart-9x6 12819-main page 181

References

[DF] David S. Dummit, Richard M. Foote, Abstract Algebra, 3rd edition, Dum-
mit and Foote, Wiley, 2003.
[F] John B. Fraleigh, A First Course in Abstract Algebra, 7th edition,
Addison-Wesley, 2003.
[J] Nathan Jacobson, Basic algebra. I. 2nd edition. W. H. Freeman and Com-
pany, New York, 1985. xviii+499 pp.
[LZ] Libin Li, Kaiming Zhao, Introduction to Abstract Algebra, ISBN: 978-7-
03-067958-1, Academic Press, 2021.
[L] Seymour Lipschutz, Set Theory and Related Topics, 2nd edition, McGraw
Hill, 1998.
[M] James S. Milne, Fields and Galois Theory,
https://www.jmilne.org/math/CourseNotes/FT421.pdf, 2013.
[P] Victor V. Prasolov, Polynomials, Translated from the 2001 Russian sec-
ond edition by Dimitry Leites. Algorithms and Computation in Mathe-
matics, 11. Springer-Verlag, Berlin, 2004.
[Z] Kaiming Zhao, Linear Algebra, ISBN: 978-1-7924-6399-0, Kendall Hunt
Publishing Company, 2021.
[ZTL] Kaiming Zhao, Haijun Tan, Genqiang Liu, Group Theory, ISBN: 978-1-
7924-7892-5, Kendall Hunt Publishing Company, 2021.

181
B1948 Governing Asia

This page intentionally left blank

B1948_1-Aoki.indd 6 9/22/2014 4:24:57 PM


March 18, 2022 9:45 amsart-9x6 12819-main page 183

Index

(SX , ·, idX ), 2 algebraic number, 94


Ff (x) , 120 algebraically closed, 100
Fpn , 105 algebraically independent, 102
G/H, 2 anti-symmetric, 179
I  R, 9 associates, 47
M1 ⊕ · · · ⊕ Mn , 70 automorphism of the field, 112
N  G, 3 axiom of choice, 180
R-submodule, 68
R/I, 9 basis, 72
R1 ⊕ · · · ⊕ Rn , 12 binary relation, 179
S(E/F ), 119 Brauer’s irreducibility criterion, 39
[E : F ], 90
[L]γβ , 79 canonical homomorphism, 10, 69
[L]β , 79 Cartesian product, 179
Φn (x), 143 characteristic, 5
Φp (x), 35 classification of finitely-generated
Zn , 4 modules over Euclidean
F , 101 domains, 80
[[G]], 134 Cohen–Macaulay ring, 85
[[K, F ]], 134 Cohn’s irreducibility criterion, 39
[[R, N ]], 10 commutative ring, 3
U(R), 6 composition series, 148
⊕n
P i=1 Mi , 70 conjugate, 111
λ∈Λ Iλ , 12
conjugation isomorphism, 111
{E : F }, 119 content of f (x), 55
L(V ), 78 coset, 9
L(V, W ), 78 cyclic submodule, 71
RP , 120
Rf (x) , 120 Dedekind’s Theorem, 157
EndD (V ), 78 deg(α, F ), 95
GLn (R), 77 depth of a ring, 85
HomD (V, W ), 78 derivative of f (x), 97
direct sum of modules, 70
abelian group, 2 Discriminant, 158
ACC, 50 divide, 6
affine algebraic set, 84 Division Algorithm, 27
algebraic closure, 99, 101 division ring, 7
algebraic element, 93
algebraic extension, 98 elementary matrix, 77

183
March 18, 2022 9:45 amsart-9x6 12819-main page 184

184 RING AND FIELD THEORY

Elementary symmetric height of a prime ideal, 85


polynomials, 156 Hilbert basis theorem, 83
equivalence, 179 homomorphism, 10
equivalence class, 179
equivalence relation, 179 ideal, 9
equivalent matrices, 77 ideal generated by S, 23
Euclidean Algorithm, 54 identity element, 1
Euclidean domain, 52 identity homomorphism, 10
Euclidean norm, 52 integral domain, 6
Euler φ-function, 142 invariant factors, 80
evaluation homomorphism, 93 inverse, 1
evaluation map, 20 invertible matrix, 77
extension field, 90 irr(α, F ), 95
extension of F by radicals, 146 irreducible, 6
external direct sum, 12 irreducible polynomial, 31
irreducible polynomial of α, 95
factor, 6 isomorphism, 10
Factor Theorem, 28
field, 7 Kronecker’s Theorem, 92
field of fractions, 16 Krull dimension of a ring, 85
field of quotients, 16 Kuratowski-Zorn Lemma, 180
field of symmetric rational
Lüroth’s Theorem, 127
functions, 155
left R-module, 67
finite extension of degree n, 90
left coset, 2
finite normal extension, 132
left ideal, 68
finitely generated module, 71
left Noetherian ring, 82
finitely-generated ideal, 82
Lindemann-Weierstrass theorem,
finitely-generated left ideal, 82 102
finitely-generated right ideal, 82 linear independence, 71
First Isomorphism Theorem, 11 linear map, 78
fixed field, 113
free module, 72 maximal element, 180
freely generated, 71 maximal ideal, 22
Frobenius automorphism, 115 minimal element, 180
minimal polynomial, 95
Gal(E/F ), 114 minimal prime ideal, 44
Gal(f (x)/F ), 136 module homomorphism, 69
Galois extension, 132 modulo, 179
Galois group, 114, 136, 150 monic polynomial, 17
Galois Theorem, 136 multiplicative norm, 61
Gauss’ Lemma, 32 multiplicity of root, 122
Gaussian integer, 4, 60
GCD, 49 n-th root of unity, 105
general linear group, 77 nilradical of a ring, 25
greatest common divisor, 49 Noetherian ring, 82
group, 1 norm, 60, 163
March 18, 2022 9:45 amsart-9x6 12819-main page 185

INDEX 185

normal subgroup, 2 solvable by radicals over F , 147


nth cyclotomic extension of F , 142 solvable group, 147
nth cyclotomic polynomial, 143 spectrum, 22
splitting field, 119
ordered basis, 78 standard ordered basis, 78
ordered set, 180 subfield, 8
Ore domain, 17 subfield generated by S, 91
Osada’s irreducibility criterion, 39 subgroup, 2
subring, 4
p-th cyclotomic polynomial, 35 sum, 12
partially ordered, 180 symmetric, 179
partially ordered set, 180 symmetric group, 2
perfect field, 125
Perron’s irreducibility criterion, 37, the matrix of a linear map, 79
38 Third Isomorphism Theorem, 11
polynomial, 17 totally ordered, 180
prime, 48 totally ordered set, 180
prime field, 90 trace, 163
prime ideal, 22 transcendence basis, 102
primitive n-th root of unity, 105 transcendence degree, 102
primitive polynomial, 32, 55 transcendental element, 93
principal ideal, 24 transcendental number, 94
principal ideal domain, 24 transitive, 179
proper ideal, 21 trivial factor, 48

quotient group, 3 UFD, 48


quotient ring, 9 Unique Factorization Theorem, 36
quotient set, 180 unit, 6
unital module, 70
R-isomorphism, 69 unital ring, 5
radical extension, 146 unity, 5
rank, 74 upper bound, 180
reflexive, 179
valuation ring, 45
regular sequence, 85
relation, 179 Wedderburn’s Little Theorem, 8
right R-module, 67 well ordering principle, 180
right ideal, 68 well-ordered, 180
right Noetherian ring, 82
ring, 3 zero divisor, 5
zero homomorphism, 10
Schönemann-Eisenstein Criterion, zero of f (x), 21
34, 59 zero-locus, 84
Second Isomorphism Theorem, 11 Zorn’s lemma, 180
separable extension, 123
simple extension field, 92
simple submodule, 68
Smith normal form, 77

You might also like