Вы находитесь на странице: 1из 247

Lectures on Modern Algebra

Author Jin-Gen Yang


Foreword

For a sufficiently educated person, the word “ algebra” often reminds


him or her a great deal of high school mathematics, such like factor-
ization of quadratic polynomials, solving an equation or a system of
equations, exponential functions, logarithmic functions and so forth.
These subjects are known as pre-calculus algebra or elementary alge-
bra.
This course is quite beyond the precalculus algebra. It emphasizes
the inner structures of groups, rings, fields and vector spaces as well
as the maps between algebraic structures. For this reason this branch
of mathematics is often refered to as “abstract algebra” or “modern
algebra”. The ideas of algebra evolved through many generations of
mathematicians around the turn of 20-th century. Among many promi-
nent mathematicians we mention Emmy Noether and Emil Artin, who
laid the foundation of modern algebra.
In the freshmen linear algebra, the students have encountered the
algebraic structure of vector spaces over real number field or complex
number field, as well as the structures of Euclidean spaces or unitary
spaces after the introduction of inner product. If we look at the set
of integers and the set of real polynomials in one variable closely, we
may find their properties strikingly similar. From the algebraic point of
view, their structures share many common properties. Roughly speak-
ing, a so-called algebraic structure is a set with one or more operations
satisfying certain conditions. This is one of the basic target of study in
mathematics. Another important basic structure is topological struc-
ture, which is not in the scope of this course. The development of
mathematics in the recent century has confirmed that modern algebra

ii
iii

is indispensable.
Due to its importance, modern algebra (or abstract algebra) has
become a basic standard course for math major students, usually of-
fered for sophomores or juniors. An ideal duration of the course is
one whole year. In recent years, only one semester of modern algebra
is available for many major universities of China. It is not easy for
instructors of this course to cover Galois theory in one semester. But
Galois theory is recognized as a milestone of algebra. It will be regret-
ful if a student does not know Galois theory after taking the course of
algebra. Through many years of teaching the author has made careful
selection of the materials for groups, rings and fields so that the stu-
dents can reach the main theorem of Galois theory and the proof of
the insolvability by radicals of equations of degrees greater than four.
Among all algebraic structures, the group is a basic structure. To
shorten the first chapter on groups, I postpone some less fundamental
but more technical subjects such like Sylow’s groups, finitely gener-
ated abelian groups and solvable groups to a later chapter. The first
three chapters aim at basic knowledge on groups, rings and fields. For
ring theory, emphasis is put on the residue ring Z/nZ and polynomial
rings, of one variable and of several variables. Through these concrete
rings students can understand more abstract concepts such like princi-
pal ideal domain and unique factorization ring. The non-commutative
rings are restricted to basic knowledge and a few standard examples
such like matrix rings and quaternions.
The fourth chapter is a brief account of linear algebra over an ar-
bitrary field. Since the students are assumed to have taken the first
course of linear algebra already, the account is often sketchy. The main
purpose of this chapter is to emphasize the difference of vector spaces
over fields of different characteristics and prepare for the theory of ex-
tensions of fields. Students should establish a point of view to regard
the extension of a field as a vector space over the base field.
Chapter 5 covers Sylow groups, finitely generate abelian groups
and solvable groups as I mentioned before. The last three chapters
are devoted to the field theory exclusively. By author experience, it is
reasonable to finish all eight chapters in 15 or 16 weeks.
iv FOREWORD

The appendices are for more enthusiastic students. A proof of


quadratic reciprocity law displays elegant techniques of finite fields.
The proof of the theorem concerning the structure of finite skew fields
displays a clever use of group actions. The proofs of these two theo-
rems are adapted from the ones of J.-P. Serre and A.Weil respectively.
The author wishes that the interested readers can feel the beauty of a
mathematical proof by reading these two appendices.
Due to the limit of time, in the last section only the insolvability
of equations of degree greater than four is proved. Using Galois theory
to deduce the formulas for the solutions of cubic and quartic equations
are put in appendices.
The sections marked with an asterisk can be skipped on the first
reading.
There are adequate amount of exercises throughout the book. The
degree of difficulty varies. Some exercises are chosen from Ph.D qual-
ifying exams. Hints or solutions are provided as an appendix. Most
problems have many different proofs, which are impossible to be in-
cluded. Students are strongly encouraged to find their own solutions
and do not rely on the appendix too heavily. Only by laying one’s
own hand on hard problems, one can feel the charm of mathematical
deduction.
In summary, the author has tried to give a concise and easy to
understand account of basic knowledge and methods in algebra without
loss of rigor. The main perspective readers are sophomores or juniors
of math major.
Professor Li Kezheng offered many valuable comments and pointed
out many mistakes. Many students of Fudan University who have taken
this course in the past several years have provided useful comments
too. I regret that I am not able to list their names. I thank all those
people who have made contributions to this book. The assistance of
Ms. Yao Lili of Science Press in the publication of this book is greatly
appreciated.

J.-G. Yang, Sept., 2008


Contents

Foreword ii

Preliminaries and Notations viii

1 Elements of Groups 1
1.1 Definitions and Examples . . . . . . . . . . . . . . . . 1
1.2 Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 Permutation Groups . . . . . . . . . . . . . . . . . . . 11
1.4 Cosets . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.5 Normal Subgroups and Quotient Groups . . . . . . . . 23
1.6 Alternating Groups . . . . . . . . . . . . . . . . . . . . 28
1.7 Homomorphisms of Groups . . . . . . . . . . . . . . . . 32
1.8 Direct Product of Groups . . . . . . . . . . . . . . . . 39
1.9 * Automorphisms of Finite Cyclic Groups and the Euler
Function . . . . . . . . . . . . . . . . . . . . . . . . . . 43
1.10 Group Action . . . . . . . . . . . . . . . . . . . . . . . 45

2 Elements of Rings and Fields 51


2.1 Basic Definitions . . . . . . . . . . . . . . . . . . . . . 51
2.2 Ideals and Quotient Rings . . . . . . . . . . . . . . . . 55
2.3 Homomorphisms of Rings . . . . . . . . . . . . . . . . 60
2.4 Elementary Properties of Fields . . . . . . . . . . . . . 63

3 Polynomials and Rational Functions 71


3.1 Polynomials in One Variable . . . . . . . . . . . . . . . 71
3.2 Division Algorithm . . . . . . . . . . . . . . . . . . . . 72
3.3 Polynomials in Several Variables . . . . . . . . . . . . . 76

v
vi CONTENTS

3.4 Factorization . . . . . . . . . . . . . . . . . . . . . . . 78
3.5 * Polynomial Functions . . . . . . . . . . . . . . . . . 87

4 Vector Spaces 90
4.1 Vector Spaces and Linear Transformations . . . . . . . 90
4.2 Quotient spaces . . . . . . . . . . . . . . . . . . . . . . 95

5 Topics in Group Theory 98


5.1 The Orbit Formula of an Action by a Finite Group . . 98
5.2 Sylow subgroups . . . . . . . . . . . . . . . . . . . . . 101
5.3 * Structure of finitely generated abelian groups . . . . 106
5.4 Solvable Groups . . . . . . . . . . . . . . . . . . . . . . 116

6 Field Extensions 120


6.1 Definitions and First Properties of Field Extensions . . 120
6.2 Algebraic Extensions . . . . . . . . . . . . . . . . . . . 123
6.3 Constructions of Field Extensions . . . . . . . . . . . . 128
6.4 Algebraically Closed Field . . . . . . . . . . . . . . . . 131
6.5 * Ruler and Compass Construction . . . . . . . . . . . 136

7 Finite fields 142


7.1 Basic Theory . . . . . . . . . . . . . . . . . . . . . . . 142
7.2 The structure of Multiplicative Group of a finite field . 144

8 Finite Galois Theory 148


8.1 Basic theory . . . . . . . . . . . . . . . . . . . . . . . . 150
8.2 * Solvable Extension and Solvability of Algebraic Equa-
tions by Radicals . . . . . . . . . . . . . . . . . . . . . 162

A Quadratic residues 169

B Every finite skew field is a field 174

C Solutions of a cubic equation and Hilbert theorem 90 178

D Solutions of quartic equations 183

E Hints or solutions for exercises 187


CONTENTS vii

Bibliography 232

Index 234
Preliminaries and Notations

Since this course is intended for the undergraduate students in math


major, we assume that the readers have already studied linear alge-
bra and calculus (at least calculus in one variable). The readers are
assumed to have the basic knowledge on sets and maps.
Some conventions and notations are given below.
A map f from one set S to another set T is an injection if f (x) 6=
f (y) wherever x and y are two distinct elements in S. The map f is an
surjection if there exists x ∈ S such that f (x) = z for every z ∈ T. A
map is a bijection if it is injective and surjective. Let U be a subset of
S. Denote by f |U the restriction of f on U.
A map f from S to T carrying x to z is denoted by f : S → T, x 7→
z. For example x 7→ ex is the exponential function. Sometimes the
identity map from a set S into itself is denoted by id or 1.
Let f be a map from S to T and g be a map from T to U. The map
g ◦ f : S → U, x 7→ g(f (x)) is the composite of f and g. It can also be
denoted by gf.
A subset of a set S is often defined by the notation {x ∈ S|P }, where
P is the condition that x should satisfy. For example {x ∈ R|0 ≤ x ≤
1} is the closed interval [0, 1]. The union and the intersection of sets
are denoted by ∪ and ∩ respectively. The difference {x ∈ A|x ∈ / B} of
the sets A and B is denoted by A − B or A\B, and we prefer to use the
latter. A set with a few elements is usually denote by {· · · }, in which
· · · is the list of all elements. For example {a} is a set consisting of a
single element a, {0, 1} is the set consisting of the elements 0 and 1.
The empty set is denoted by ∅.
The knowledge of equivalence relations and equivalence classes is

viii
ix

helpful, but is not indispensable. We will explain them when they are
first used in the text.
Readers are required to know the number fields well. By definition,
a number field is a subset of the complex number field closed under
addition, subtraction, multiplication and division. The most frequently
used number fields are complex number field, real number field and
rational number field.
The equality x = ±a means that x is equal to either a or −a.
Some common notations are listed as follows:

N — set of natural numbers


Z — set of integers
Q — set of rational number
R — set of real numbers
C — set of complex numbers
x PRELIMINARIES AND NOTATIONS
Chapter 1

Elements of Groups

Many objects we encounter in mathematics are sets equipped with one


or more operations. Such objects are usually called algebraic struc-
tures. In this course the algebraic structures we will study include
groups, rings and vector spaces. In terms of the number of operations
the group theory can be considered to be a natural starting point for
this course, since a group involves only one basic operation. In this
chapter the definition, basic properties of groups and homomorphisms
between groups will be explained. More advanced topics on groups will
be covered in later chapters.

1.1 Definitions and Examples

Let S be a set. Denote the Cartesian product of S with itself by S × S.


It consists of all pairs (a, b) with a, b ∈ S. Note that (a, b) and (b, a)
are different elements in S × S if a 6= b. A map f from the set S × S
to S is called a (binary) operation.
For example, let S be the set of all real numbers. Defined the map
f : S × S → S by f (a, b) = a + b. Then f is a binary operation.
We may use all sorts of ways to define operations such as f (a, b) =
a2 + b3 , f (a, b) = ab, . . . .
Since it is awkward to use the notation f (a, b) for an operation,
there are more convenient notations such as a + b, [A, B], u × v, . . . ,
depending on occasions. The simplest notation is ab. We will use this
notation when no confusion will be caused. In practice, it is better

1
2 CHAPTER 1. ELEMENTS OF GROUPS

to use the notations that we are already familiar with. For instance,
the addition of numbers is denoted by “+”, the cross product of two
vectors in R3 is denoted by “×”.
Let’s look at some more examples.

• Let S be a 3-dimensional Euclidean space. The addition u + v


and cross product u × v are binary operations. The dot product
u · v is not a binary operation since the result of the dot product
is not a vector.

• Let S be the set of all n × n matrices. Then A + B, AB, AB −


BA are three different operations on S (the last one is a basic
operation in Lie Algebra).

• Let S be the set of all everywhere well-defined real functions.


Then the composite f ◦ g is a binary operation on S.

Unitary operations (involving one variable) appear quite often too,


such as −a (the negative of a number), a2 (the square of a number),
ā (the conjugate of a complex number), AT (the transpose of a square
matrix), etc..
We say that a binary operation f on a set S satisfies the law of
associativity if
f (a, f (b, c)) = f (f (a, b), c)

holds for any a, b, c ∈ S. By our convention this condition can be written


simply as
a(bc) = (ab)c,

which is exactly the form we are familiar with.


In the previous examples many binary operations satisfy the law of
associativity. The cross product of vectors and the operation AB − BA
for square matrices A, B do not satisfy the law of associativity. It is
easy to verify that the operation f (a, b) = a2 + b3 on real numbers does
not satisfy the law of associativity.
An element e in a set S equipped with a binary operation is called
an identity element if ea = ae for any a ∈ S.
Examples:
1.1. DEFINITIONS AND EXAMPLES 3

• 0 is an identity element in the set of real numbers under addition.

• 1 is an identity element in the set of real numbers under multi-


plication.

• The identity matrix In is an identity element in the set of n × n


matrices under multiplication.

Proposition 1.1.1. A set S with a binary operation has at most


one identity element.
Proof. Let e and e0 be two identity elements of S. Since e is an
identity element, ee0 = e0 holds. On the other hand, ee0 = e holds since
e0 is also an identity element. Therefore e = ee0 = e0 .

Definition 1.1.2. A nonempty set S with a binary operation is


called a semigroup if the law of associativity is satisfied. A semigroup
containing an identity element is called a monoid.
According to Proposition 1.1.1 there is a unique identity element in
a monoid.
Example 1.1.3. The set of natural numbers under addition is a
semigroup but not a monoid.
Definition 1.1.4. Let a be an element in a monoid S and let e
be the identity element of S. If there is an element b ∈ S such that
ba = ab = e then a is called an invertible element and b is called the
inverse of a.
Note that the article in front of the word “inverse” is “the” instead
of “an”. This will be justified by the following proposition.
Proposition 1.1.5. There is a unique inverse for an invertible
element in a monoid.
Proof. Let b, b0 be inverses of a. Then b = be = b(ab0 ) = (ba)b0 =
eb0 = b0 .

It is obvious that the relation “inverse” is symmetric. That is to


say, if a is invertible with b as its inverse, then b is also invertible and
a is the inverse of b.
The inverse of an invertible element a is commonly denoted by a−1 .
4 CHAPTER 1. ELEMENTS OF GROUPS

Definition 1.1.6. A nonempty set G with a binary operation is


defined to be a group if the following three conditions are satisfied:

• The law of associativity is satisfied;

• The identity element exists;

• Every element in G is invertible.

The first two conditions in the definition mean that every group is
a monoid. Although the concepts of semigroup and monoid are more
general than group, but we are primarily interested in groups.
The unitary operation a 7→ a−1 in a group is called the inverse
operation. Since it is totally determined by the binary operation of the
group, we do not consider the inverse as a basic operation of the group.
As we have mentioned before, the binary operation in a group uses
different notations depending on occasions. The default notation is ab,
or a · b occasionally. For this reason, the operation can also be referred
to as “multiplication” and ab is called the product of a and b. In this
case the inverse of a is denoted by a−1 and the identity element e can
also be denoted by 1, or 1G to avoid confusion. For a natural number
n, the n-th power an is defined to be the product of a with itself n
times. The power a−n with negative integral exponent is defined to be
(a−1 )n .

Let G be a group. If ab = ba for any a, b ∈ G, then G is called a


commutative group, or abelian group.
For many abelian groups, the binary operation is usually called
“addition” and denoted by a + b. In this case the identity element is
often called the zero (element) of the group and denoted by 0. The
inverse of an element a is denoted by −a and the n-th power of an
element a becomes na = a + · · · + a.
A groups consisting of a finite number of elements is called a finite
group, otherwise it is called an infinite group. The number of ele-
ments in a group G is denoted by |G|, called the order of G. Hence
|G| is ∞ if G is an infinite group and |G| is a natural number if G is a
finite group.
1.1. DEFINITIONS AND EXAMPLES 5

Let g be an element of a group G. If there is no natural number n


such that g n = 1, then g is called an element of infinite order, otherwise,
define the order of g to be the smallest natural number n such that
g n = 1, denoted by o(g). Hence an element in G has order one if and
only if it is the identity element. By convention, o(g) = ∞ if g is an
element of infinite order.
• Every number field under addition forms an abelian group, called
the additive group of that number field. One thing to keep in mind
is that the multiplication operation of that number field is neglected
when it is regarded as a group, since by definition a group involves only
one basic operation.
• The set Z of all integers under addition is an abelian group, called
the additive group of integers.
• All n × n invertible matrices over a number field K under the
(matrix) multiplication is a group called general linear group and
denoted by GLn (K). It is not an abelian group when n > 1.
• A number field K under multiplication is a monoid, but not a
group since the element 0 in not invertible. Denote K ∗ = K\{0}.
Then K ∗ is an abelian group under multiplication, which is essentially
the same as GL1 (K). Although this group is abelian, but it would be
absurd to use “+” to denote the multiplication.
• Any vector space is an abelian group under addition.
• 3-dimensional Euclidean space R3 is not a semigroup under cross
product, since the law of associativity is not valid.
• The smallest group has only one element. Such a group is called
a trivial group.
• Empty set is not a group.

Proposition 1.1.7 (law of cancelation). If the element a, b, c in a


group satisfy ab = ac or ba = ca, then b = c.

Proof. Assume that ab = ac. By left multiplying a−1 to both sides


of ab = ac the equality a−1 (ab) = a−1 (ac) is obtained, so (a−1 a)b =
(a−1 a)c by the law of associativity. Hence eb = ec by the definition of
inverse. Finally b = c follows from the definition of the identity element
e.
6 CHAPTER 1. ELEMENTS OF GROUPS

The same argument shows that ba = ca implies b = c.

Corollary 1.1.8. Let a, b be two elements of a group G.


1) If ab = a or ba = a then b = 1G ;
2) If ab = 1G or ba = 1G then b = a−1 .

Proof. 1) Apply the law of cancelation to ab = a1G .


2) Apply the law of cancelation to ab = aa−1 .

Proposition 1.1.9. Let a, b be two elements of a group G. Then


(ab)−1 = b−1 a−1 .

Proof. The equality (b−1 a−1 )(ab) = b−1 (a−1 a)b = b−1 b = 1 and
Corollary 1.1.8.2) imply b−1 a−1 = (ab)−1 .

This property is the generalization of the well-known formula (AB)−1 =


B −1 A−1 in linear algebra, where A and B are two invertible n × n ma-
trices.

As we have seen, any number field is a group under addition. For


this reason we may consider addition as the first important operation
of this number field. In fact, this is the first operation taught in ele-
mentary school. The subtraction is not regarded as a basic operation.
Since a − b = a + (−b), the subtraction can be considered to be the
composite of the inverse and the addition.

Exercises
1. In the set G = {a ∈ R|a > 0, a 6= 1} define a binary operation
a ∗ b = aln b . Is G a group under this operation? Here ln b is the natural
logarithm of b.
2. Let A be the set of all strictly increasing continuous functions
on [0, 1] satisfying f (0) = 0, f (1) = 1. For any f, g ∈ A define f g to
be the composite of f and g, i.e., (f g)(x) = f [g(x)] for any x ∈ [0, 1].
Prove that A is a group under this operation.
3. Let G be a semigroup satisfying the following two conditions:
1) there is some e ∈ G such that ea = a for any a ∈ G;
2) for every a ∈ G, there is some b ∈ G such that ba = e.
1.2. SUBGROUPS 7

Prove that G is a group.


4. Prove that every element of a finite group has finite order.
5. Let G be a group and a, b ∈ G. Prove that o(ab) = o(ba).
6. Assume that every element a of a group G satisfies a−1 = a.
Prove that G is an abelian group.

1.2 Subgroups

Definition 1.2.1. Let H be a nonempty subset of a group G sat-


isfying the following two conditions
1) (closure under the multiplication) ab ∈ H for any a, b ∈ H;
2) (closure under the inverse operation) a−1 ∈ H for any a ∈ H.
Then H is called a subgroup of G.
By the closure of multiplication, a subgroup H of G inherits the
binary operation from G. Naturally this inherited operation preserves
the law of associativity. Since H is nonempty, there is some element
a ∈ H. Hence a−1 ∈ H by the closure of inverse operation, which
implies that 1 = a−1 a ∈ H. So H is a group by itself. The readers
may compare the concept of subgroup with that of subspace in linear
algebra.
It is easy to see that any subgroup of an abelian group is abelian.
Example 1.2.2. • Every group G has two subgroups for free. They
are {1} and G. (They become the same if |G| = 1.) They are called
trivial subgroups of G. A subgroup H of G is called a proper subgroup
if H 6= G.
• Let n be a natural number. Let nZ denote the set of all integers
divisible by n. Then nZ is a subgroup of Z.
• Let SLn (K) be the set of all n × n matrices with determinant one
over a number field K. It is a subgroup of GLn (K), called the special
linear group.
• The set of matrices consisting of
" # " #
1 0 −1 0
,
0 1 0 −1
8 CHAPTER 1. ELEMENTS OF GROUPS

is a subgroup of SL2 (K) with |H| = 2.

• The real number field R is a group under addition and R∗ =


R\{0} is a group under multiplication. Although the latter is a subset
of the former, but R∗ = R − {0} is not a subgroup of R because the
binary operations of these two groups are different.

Proposition 1.2.3 (a criterion for subgroup). Let H be a nonempty


subset of a group G. If ab−1 ∈ H for any a, b ∈ H, then H is a subgroup
of G.

Proof. Let c be an arbitrary element in H. Then 1 = cc−1 ∈ H


by the hypothesis of the proposition. For any a ∈ H, we have a−1 =
1a−1 ∈ H. Hence H is closed under the inverse operation.
Since ab = a(b−1 )−1 ∈ H, H is also closed under multiplication.

Proposition 1.2.4. Let {Hi }i∈I be a set (not necessarily finite) of


T
subgroups of a group G. Then H = i∈I Hi is a subgroup of G.

Proof. Since 1 ∈ ∩i∈I Hi the set H is not empty. Let a, b ∈ H.


Then ab−1 ∈ Hi for every i ∈ I, since every Hi is a subgroup of G. Hence
ab−1 ∈ H. It follows from Proposition 1.2.3 that H is a subgroup of
G.

Definition 1.2.5. Let g be an element of a group G. The set


C(g) = {a ∈ G|ag = ga} is called the centralizer of g in G. For any
nonempty subset S of G, the set C(S) = {a ∈ G|ag = ga for all g ∈ S}
is called the centralizer of S in G. In particular, C(G) is called the cen-
ter of G.
T
It is easy to see that C(g) and C(S) = g∈S C(g) are subgroups of
G and C(G) is an abelian group. The group G is abelian if and only if
G = C(G).

Let S be a nonempty subset of a group G. Let hSi denote the in-


tersection of all subgroups of G containing S. Then hSi is the smallest
subgroup of G containing S in the sense that hSi ⊆ H for every sub-
group H of G with S ⊆ H. The subgroup hSi is called the subgroup
generated by S. If S is a finite set consisting of the element a1 , . . . , an ,
1.2. SUBGROUPS 9

then hSi can also be denoted by ha1 , . . . , an i. If G = hSi, then we say


that G is generated by S.
Example 1.2.6. • Z = h1i = h−1i.

• GLn (K) is generated by all n × n elementary matrices.


A group generated by a finite set is called a finitely generated
group. In particular, a group generated by a single element is called a
cyclic group. For instance, Z is a cyclic group while GLn (K) is not.
It is obvious that every cyclic group is abelian.
Let a be an element of a group G. Then hai is a subgroup of G,
called a cyclic subgroup of G. It is easy to verify that o(a) = |hai|. In
fact, if o(a) = ∞ then

hai = {. . . , a−2 , a−1 , 1, a, a2 , . . .}

and if o(a) = n < ∞ then

hai = {1, a, a2 , . . . , an−1 }.

Example 1.2.7. In GLn (K) the matrix


" #
1 1
0 1

generates an infinite cyclic subgroup while the matrix


" √ #
1 3

√2
− 2
3
2
− 21

generates a cyclic subgroup of order 3.


Proposition 1.2.8. Let S be a nonempty subset of a group G. Then

hSi = {ae11 ae22 · · · aenn |a1 , . . . , an ∈ S, e1 , . . . , en = ±1,

n is an arbitrary nonnegative integer}.


Proof. First note that the elements a1 , a2 , . . . , an involved in the
expression are not necessarily distinct.
10 CHAPTER 1. ELEMENTS OF GROUPS

Denote the set of the right hand side by T. Then T is a subgroup


containing S. Hence hSi ⊆ T.
Let H be a subgroup of G such that S ⊆ H. By the definition
of subgroup every element that can be expressed as ae11 ae22 · · · aenn (ai ∈
S, ej = ±1) is in H. Hence T ⊆ H. Therefore T = hSi.

At this point we want to determine all subgroups of Z. First of all,


as we have already seen before, 0, Z, 2Z, 3Z, . . . are all subgroups of Z.
We show that there are no other subgroups.
Let H be a nontrivial subgroup of Z. Then there is some nonzero
integer a in H. Since −a is also in H by the definition of subgroup, there
exists a natural number in H. Let n be the smallest natural number
in H. Then nZ ⊆ H. For any m ∈ H, there are integers q, r such that
m = qn + r in which 0 ≤ r ≤ n − 1. Since r = m − qn ∈ H, r cannot
be a natural number by the minimality of n. Hence r = 0. This implies
that m ∈ nZ. It follows that H ⊆ nZ. Therefore every subgroup of Z
is an infinite cyclic group.

Exercises

1. Let G be the set of all 3 × 3 real upper triangular matrices


with all diagonal elements equal to 1. Show that G is a group under
multiplication and determine the center of G.

2. Let X and Y be two subsets of a group G. Prove that


1) if X ⊆ Y then C(X) ⊇ C(Y );
2) X ⊆ C(C(X));
3) C(X) = C(C(C(X))).
Here C(X) denotes the centralizer of X.

3. Let H be a subgroup of a groups G such that H is contained in


every nontrivial subgroup of G. Prove that H is contained in the center
of G.

4. An element a of a group G is called a perfect square if there


exists b ∈ G such that a = b2 . Assume that G is a cyclic group and
a, b ∈ G are not perfect squares. Show that ab is a perfect square.
1.3. PERMUTATION GROUPS 11

Give an example to show that this statement is not true for non-cyclic
groups.
5. Let H be a nonempty subset of a finite group G. Show that H
is a subgroup of G if ab ∈ H for any a, b ∈ H.
6. Let A be an n × n real invertible matrix and let G be the set
consisting of all n × n real matrices P such that P T AP = A. Show that
G is a subgroup of GLn (R). Here P T is the transpose of P.

1.3 Permutation Groups

Permutations and combinations have already been studied in high


school mathematics. For instance, 2, 1, 4, 5, 3 is a permutation (or re-
arrangement) of the sequence 1, 2, 3, 4, 5. It is known that the number
of permutations of 1, 2, 3, 4, 5 is equal to 5! = 120.
Let’s look at the permutations from another point of view. The re-
arrangement 2, 1, 4, 5, 3 is treated as a bijection σ of the set {1, 2, 3, 4, 5}
into itself, which carries 1 to 2, 2 to 1, 3 to 4, 4 to 5 and 5 to 3. We
can use the following table to specify this bijection
" #
1 2 3 4 5
.
2 1 4 5 3
This interpretation suggests the following definition.

Definition 1.3.1. Let n be a natural number. A bijection from


the set {1, 2, 3, . . . , n} to itself is a permutation of n objects.

As in the example, a permutation can be represented by a table


" #
1 2 3 ··· n
.
σ(1) σ(2) σ(3) · · · σ(n)
Here the numbers 1, 2, 3, . . . , n do not have numerical meaning what-
soever. They are merely convenient labels for distinct objects. Denote
the set of all permutations of n objects by Sn . Then Sn contains n! (the
factorial of n) elements.
12 CHAPTER 1. ELEMENTS OF GROUPS

Introduce a binary operation in Sn in the following way. For any


σ, τ ∈ Sn define στ to be the composite of σ and τ, the map τ followed
by σ. Thus στ (i) = σ[τ (i)] for every i ∈ {1, 2, . . . , n}. Evidently this is
a well-defined binary operation of the set Sn .

Proposition 1.3.2. The set Sn is a group under the binary oper-


ation of composite.

Proof. By the rule of composite of maps σ(τ π) = (στ )π holds for


any σ, τ, π ∈ Sn . Hence the law of associativity is valid.
Let id be the identity map of {1, 2, 3, . . . , n}, i.e., id(i) = i for every
i ∈ {1, 2, 3, . . . , n}. Then id ◦ σ = σ ◦ id = σ for any σ ∈ Sn . Hence id
is the identity element.
Since σ ∈ Sn is a bijection of {1, 2, . . . , n}, its inverse map τ exists.
It means that στ = τ σ = id.

The product of two permutations can be read off from their tables.
For example,
" #" # " #
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
= .
3 1 5 4 2 2 1 4 5 3 1 3 4 2 5
" #−1 " #
1 2 3 4 5 1 2 3 4 5
= .
3 1 5 4 2 2 5 1 4 3

Definition 1.3.3. A subgroup of Sn is called a permutation


group. Sn is called the symmetric group of n objects.

There are two reasons for introducing permutation groups at this


point. First of all, they form a large class of very important finite
groups. Secondly, many non-abelian groups can be found in permuta-
tion groups.
Let’s look at some symmetric groups of low order. When n ≤ 2,
they are too simple to worth studying. The first nontrivial symmetric
group is S3 . Its six elements can be enumerated as
" # " # " #
1 2 3 1 2 3 1 2 3
σ0 = , σ1 = , σ2 = ,
1 2 3 1 3 2 3 1 2
1.3. PERMUTATION GROUPS 13
" # " # " #
1 2 3 1 2 3 1 2 3
σ3 = , σ4 = , σ5 = .
3 2 1 2 1 3 2 3 1
It is easy to verify that σ1 σ2 = σ4 and σ2 σ1 = σ3 . Thus S3 is not
abelian.
It is not hard to verify that S3 has following four nontrivial sub-
groups {σ0 , σ1 }, {σ0 , σ3 }, {σ0 , σ4 }, {σ0 , σ2 , σ5 }.
The relations between the subgroups of S3 can be described by the
following diagram.
V
rrr S4 LLVLVLVVVVV
rrr LLL VVVVVV
rr LLL VVVV
VVVV
rrr L V
{σ0 , σ1 }L {σ0 , σ3 } {σ0 , σ4 } {σ0 , σ2 , σ5 }
LLL h
rr hhhh
LLL
LLL rrrrr hhhhhhhh
h
rhrhrhhhhhh
{σ0 }
If two subgroups are directly connected by a straight line, the sub-
group in the lower position is contained in the one in the upper position.
The diagram makes it easy to see the relations of subgroups at a glance.
Serious readers may try to write all elements of S4 and as many as
possible subgroups and draw a diagram of subgroups.
You may probably notice that this two line notation for permuta-
tions is not economic. The first line is really not necessary. To be
worse, in S9 a permutation of “swapping 1 and 2” is denoted by
" #
1 2 3 4 5 6 7 8 9
,
2 1 3 4 5 6 7 8 9

which is hardly bearable. This suggests the following concepts.

Definition 1.3.4. Let i1 , i2 , . . . , id be d distinct objects in {1, 2, . . . , n}.


Let σ be an element in Sn such that

σ(i1 ) = i2 , σ(i2 ) = i3 , . . . , σ(id ) = i1

and σ(i) = i for all i ∈ / {i1 , i2 , . . . , id }. Then σ is called a d-cycle,


denoted by (i1 i2 · · · id ). The notation of a cycle is not unique. For
instance, (i2 i3 · · · id i1 ) and (i1 i2 · · · id ) are the same cycle. Two cycles
14 CHAPTER 1. ELEMENTS OF GROUPS

are called disjoint if every object in the first cycle does not appear in
the second one. For instance, (142) and (36) are disjoint while (261)
and (3245) are not.
A 2-cycle is called a transposition.

It is not hard to see that every permutation can be expressed as the


product of mutually disjoint cycles. For example,
" #
1 2 3 4 5 6
= (136)(52).
3 5 6 4 2 1
Although the cycle notation is not as straightforward as the table
notation, we will soon see that in some occasions it is indispensable.
One thing we should mention is: (135)(52) can denote an element in S5
as well as an element in any Sn with n > 5. This ambiguity is tolerable,
since the value of n is usually clear from the context.
The multiplication of two permutations under the cycle notation can
be accomplished by changing them into the table notion first. This is
not necessary. One may use his (or her) favorite method to write out the
result directly from the cycle notation. For example, let σ = (152)(34),
τ = (35)(41). Then στ = (152)(34)(35)(41) = (132)(45).

Proposition 1.3.5. Every permutation can be written as the prod-


uct of transpositions, not necessarily disjoint.

Proof. It suffices to show that every cycle can be written as the


product of some transpositions.
It is easy to verify that (i1 i2 · · · id ) = (i1 i2 )(i2 i3 ) · · · (id−2 id−1 )(id−1 id ).

Let σ ∈ Sn . Let r be the number of elements in the set

{(i, j)|1 ≤ i < j ≤ n, σ(i) > σ(j)}.

In other words, r is the number of pairs out of order in the sequence


σ(1), σ(2), · · · , σ(n). The permutation σ is called an even permutation
if r is even, otherwise it is called an odd permutation. The set of all
even permutations in Sn is denoted by An .
1.3. PERMUTATION GROUPS 15

Example 1.3.6. Let


" #
1 2 3 4 5
σ= .
3 1 5 4 2

The sequence 3, 1, 5, 4, 2 has five pairs out of order. Hence σ is an odd


permutation.

Lemma 1.3.7. Let σ ∈ Sn and let τ = (ij) be a transposition with


1 ≤ i < j ≤ n. If σ is an even permutation then στ is an odd permu-
tation. If σ is an odd permutation then στ is an even permutation.

Proof. First assume that j = i + 1. Express σ as


" #
1 2 ··· i i+1 ··· n
.
σ(1) σ(2) · · · σ(i) σ(i + 1) · · · σ(n)

Then
" #
1 2 ··· i i + 1 ··· n
σ ◦ (i, i + 1) = .
σ(1) σ(2) · · · σ(i + 1) σ(i) · · · σ(n)

One may observe that στ has one more (or less) pair out of order than
σ. Hence the multiplication of τ from the right changes the parity of σ.
For the general case 1 ≤ i < j ≤ n, it follows from the equality

(ij) = (i, i + 1)(i + 1, i + 2) · · · (j − 2, j − 1)

(j − 1, j)(j − 2, j − 1) · · · (i + 1, i + 2)(i, i + 1)

and what we have proved that σ ◦ (ij) and σ have different parity.

Corollary 1.3.8. For any natural number n > 1, the number of


even permutations in Sn (n > 1) is equal to that of odd permutations.

Proof. Let A and B denote the sets of all even permutations and
odd permutations respectively. By Lemma 1.3.7 the map σ 7→ σ · (12)
is a one to one correspondence from A to B. Therefore A and B contain
the same number of elements.
16 CHAPTER 1. ELEMENTS OF GROUPS

Corollary 1.3.9. A permutation is even (odd resp.) if and only if


it can be written as a product even (odd resp.) number of transpositions.
Proof. This follows from Proposition 1.3.5 and 1.3.7 immediately.

Permutation groups appear in applications in various forms. One


important application is the symmetry of a geometric figure.
Let Γ be a geometric figure in a Euclidean space. The set of all rigid
transformations of Γ to itself under composite operation form a group,
called the group of symmetries of Γ. Here a rigid transformation is
a bijection from Γ to Γ which is a composite of translations, rotations
and reflections.
As plane figures, the groups of symmetries of the following figures
have order 2 and 1 respectively.

@ @
@ @
@ @

The group of symmetries of a geometric figure reflects how sym-


metric the figure is. The figures in the examples above are not quite
symmetric compared with a circle, whose group of symmetries is an
infinite group.
The group of symmetries of a regular polygon of n sides is denoted
by Dn , called a dihedral group.
We use S = {1, 2, . . . , n} to denote the set of vertices of a regular
polygon of n sides, as illustrated in Figure 1. Since a rigid transforma-
tion carries vertices to vertices and once the images of all vertices are
known the rigid transform is determined, Dn is a subgroup of Sn . In
other words, Dn is a permutation group.
Let us denote by σ the rotation of the polygon by 2π/n in the
counterclockwise direction. Choose an axis of symmetry such as the
x-axis in Figure 1. Denote the reflection across this axis by τ. Then
σ, τ ∈ Dn . It is obvious that o(σ) = n and o(τ ) = 2. Under cycle
notation of permutation groups, we may write σ = (123 · · · n) and
τ = (1n)(2, n − 1) · · · .
1.3. PERMUTATION GROUPS 17

Y
6
3 2
@ 4
@
@ @
4 @ 1   @
  @
  @
- X   @
O 


 @
@
5@ 8 2 
S 

!! 3
@
!!
@ S  !!
@ S  ! !
@ S!!
6 7 1

Figure 1 Figure 2

It is easy to verify that

στ = τ σ n−1 .

Hence Dn is not abelian.


Every rigid transformation carries the adjacent vertices 1, 2 into two
adjacent vertices. The transformation is determined by the images of
1 and 2, which can be enumerated as

(1, 2), (2, 3), . . . , (n − 1, n), (n, 1),

(2, 1), (3, 2), . . . , (n, n − 1), (1, n).

This tells us that |Dn | = 2n.


Let’s investigate the group of symmetries of a regular tetrahedron.
Let 1, 2, 3, 4 denote the vertices of the tetrahedron, as shown in Figure
2. It is easy to see that every permutation of the vertices determines a
rigid transformation of the tetrahedron. Hence the group of symmetries
of tetrahedron is S4 .
18 CHAPTER 1. ELEMENTS OF GROUPS

Exercises

1. Find all natural numbers n such that there is an element of order


n in S7 .

2. Find an element of order 20 in S9 . Show that there is no element


of order 18 in S9 .

3. Show that the group of symmeties of a non-square rectangle


contains four elements.

4. Without explicit enumeration show that the groups of symme-


tries of a cube and a regular octahedron have the same order. Do the
groups of symmetries for regular dodecahedron and icosahedron have
the same number of elements?

1.4 Cosets

Definition 1.4.1. Let H be a subgroup of a group G. Let a be an


element of G. Denote aH = {ah|h ∈ H} and Ha = {ha|h ∈ H}. The
sets aH and Ha are called a left coset and a right coset of H in G
respectively.

Left cosets and right cosets are subsets of the group G, but not
subgroups in general. The subgroup H itself is a left coset as well as a
right coset, since H = 1H = H1.

Example 1.4.2. Let G = GL2 (R) and let H be the subgroup of G


consisting of all 2 × 2 invertible upper-triangular real matrices. Let
" #
1 0
g= .
1 1

Then
1.4. COSETS 19

" #" #
n 1 0 a b o
gH = a 6
= 0, c 6
= 0


1 1 0 c
" #
n a b o
= a 6
= 0, c 6
= 0


a b+c
" #
n a b o
= a 6
= 0, d 6
= b .


a d

" #" #
n a b 1 0 o
Hg = a 6= 0, c 6= 0

0 c 1 1
" #
n a+b b o
= a 6= 0, c 6= 0

c c
" #
n d b o
= c 6= 0, d 6= b .

c c

Note that gH 6= Hg in this example.


Example 1.4.3. H = {e, (123), (132)} is a subgroup of S3 . (12)H =
{(12), (23), (13)} is a left coset of H in S3 . H(12) = {(12), (13), (23)}
is a right coset. Thus (12)H = H(12). In next section we will see that
this phenomenon is not coincidental.
Example 1.4.4. In the dihedral group Dn , let σ be the rotation by
2π/n in counterclockwise direction. Let τ be the reflection across one
axis of symmetry.
Since o(σ) = n, the subset

{1, σ, σ 2 , . . . , σ n−1 }

is a cyclic subgroup of Dn consisting of all rotations. The reflection τ


is not in H. Hence
τ, στ, σ 2 τ, . . . , σ n−1 τ

is a right coset of H which is different from H. Moreover, none of the


elements in this coset is a rotation. Hence H ∩Hτ = ∅. Since |Dn | = 2n,
20 CHAPTER 1. ELEMENTS OF GROUPS

the dihedral group Dn is a disjoint union of H and Hτ. In particular


σ, τ generate Dn . In this way, the group Dn is better understood.

Now let us explore basic properties of cosets.

Proposition 1.4.5. Let H be a subgroup of a group G. For any


element a, b ∈ G, the map f : aH → bH, g 7→ ba−1 g is a bijection from
aH to bH.

Proof. It is easy to see that this map is well-defined. Define a map


φ : bH → aH, g 7→ ab−1 g. Then φ ◦ f and f ◦ φ are identity maps.

Proposition 1.4.6. Two cosets aH and bH are equal if and only


if a−1 b ∈ H.

Proof. Assume that aH = bH. Then b ∈ aH. Hence b = ah for


some h ∈ H, which implies a−1 b = h ∈ H.
Conversely, assume that a−1 b ∈ H. Then h = a−1 b ∈ H. Hence
b = ah ∈ aH, which implies bH ⊆ aH. On the other hand, the equality
b−1 a = (a−1 b)−1 ∈ H implies aH ⊆ bH. Therefore aH = bH.

This proposition provides a convenient way to check whether two


left cosets are equal.

Corollary 1.4.7. If aH ∩ bH 6= ∅, Then aH = bH.

Proof. Assume that g ∈ aH ∩ bH. Then g = ah = bh0 for some


h, h0 ∈ H. Hence a−1 b = hh0−1 ∈ H. It follows from Proposition 1.4.6
that aH = bH.

Since a ∈ aH for every a ∈ G, every element of G belongs to at least


one left coset of H. If follows from Corollary 1.4.7 that G is the disjoint
union of some left costs. In other words, the left cosets of H in G give
a partition of G. The number r of the left cosets in this partition is
called the index of H in G, denoted by (G : H). This number is a
natural number or ∞.
Similarly, the right cosets of H give another partition of G, which
might be different from the partition into left cosets. However the
number of right cosets in this partition is also equal to (G : H). In
1.4. COSETS 21

fact, it is easy to check that the map aH 7→ Ha−1 gives a one to one
correspondence from the set of left cosets of H to the set of right cosets
of H.

Theorem 1.4.8 (Lagrange). Let H be a subgroup of a finite group


G. Then
|G| = |H|(G : H).

Proof. This is because every left coset of H contains |H| elements


due to Proposition 1.4.5.

Lagrange’s theorem tells us that the number of elements in any


subgroup of a finite group G always divides the number of elements in
G.

Corollary 1.4.9. Let a be an element in a group G of order n.


Then o(a)||G|.

Proof. This is because |hai| = o(a).

Corollary 1.4.10. If the order of a group is a prime number then


this group is a cyclic group.

Proof. Denote p = |G|. Choose any a ∈ G such that a 6= 1. Then


o(a) > 1. It follows from 1.4.9 that o(a)|p. Since p is a prime number,
o(a) = p. Hence G = hai.

Corollary 1.4.11. Let a be an element of a finite group G with


n = |G|. Then an = 1.

Proof. Corollary 1.4.9 implies n = o(a)m for some integer m.


Hence an = (ao(a) )m = 1m = 1.

The concept of coset is a special case of the so called “equivalence


class”. In order to deepen our understanding of the partition of a group
into left cosets let us recall equivalence relations and equivalence classes
in the set theory.
22 CHAPTER 1. ELEMENTS OF GROUPS

Let S be a set. A relation ∼ between the elments of S is called an


equivalence relation if the following three conditions hold:

1. (reflexivity) a ∼ a for every a ∈ S;

2. (symmetry) a ∼ b if and only if b ∼ a;

3. (transitivity) a ∼ b and b ∼ c imply a ∼ c.

For example, although the relation “≤” in the set of real numbers
satisfies reflexivity and transitivity, but it does not satisfy symmetry.
Hence it is not an equivalence relation. The relation of similarity in the
set of all triangles in the Euclidean plane is an equivalence relation. The
relation of similarity in the set of n × n matrices is also an equivalence
relation.
Let ∼ be an equivalence relation on a set S. For any a ∈ S, denote
[a] = {x ∈ S|x ∼ a}, called an equivalence class represented by a.
Let b be another element in S. If [a] ∩ [b] 6= ∅, then c ∈ [a] ∩ [b] for some
c ∈ S, i.e.,c ∼ a, c ∼ b. It follows from symmetry and transitivity of ∼
that a ∼ b, which implies [a] = [b]. This means that the equivalence
classes give a partition of S.
Let H be a subgroup G. For any a, b ∈ G, define a ∼ b if a−1 b ∈ H.
It is easy to verify that ∼ is an equivalence relation on G. According to
Proposition 1.4.6 an equivalence class under this relation is a left coset
of H in G.
Exercises
1. Let H = {e, (12)} ⊂ S3 . List all left cosets and right cosets of H.
2. Let H be a subgroup of a group G and let n = (G : H) < ∞.
Determine whether the following statements hold. Prove it or disprove
it by a counterexample.
1) if a ∈ G then an ∈ H;
2) if a ∈ G then there is a natural number k with k ≤ n such that
k
a ∈ H.
3. Let H and K be subgroups of a group G such that (G : H) < ∞.
Show that (K : H ∩ K) ≤ (G : H).
1.5. NORMAL SUBGROUPS AND QUOTIENT GROUPS 23

4. Let K be a nonempty subset of a group G. Denote gK = {gk|k ∈


K} for any g ∈ G. Assume that aK and bK are either identical or
disjoint for any a, b ∈ G. Show that K is a left coset of some subgroup
of G.

5. Let H1 , H2 be two different subgroups of a group G. Let a1 H1 and


a2 H2 be two left cosets of H1 and H2 respectively. Show a1 H1 6= a2 H2 .
Give an example of two different subgroups H1 , H2 in the symmetric
group S3 and a1 , a2 ∈ S3 such that a1 H1 = H2 a2 .

1.5 Normal Subgroups and Quotient Groups

Let H be a subgroup of a group G. Let Γ denote the set of all left


cosets of H in G. For an element aH ∈ Γ, a is a representative of aH.
It is not unique. Every element in aH can be a representative of aH.
Let us try to use the binary operation of G to equip Γ with an
operation under which Γ become a new group.
For aH, bH ∈ Γ, is it possible to define the product of aH and
bH to be (ab)H? This operation makes sense only if the left coset
(ab)H does not depend upon the choices of the representatives a and
b. Let a0 and b0 be two other representatives of aH and bH respec-
tively, i.e., a0 = ah, b0 = bk for some h, k ∈ H. In order that (ab)H =
(a0 b0 )H the element (ab)−1 (a0 b0 ) must be in H by Proposition 1.4.6, i.e.,
b−1 a−1 ahbk = b−1 hbk ∈ H should hold for any b ∈ G and h, k ∈ H.
This is equivalent to saying the b−1 hb ∈ H for any b ∈ G, h ∈ H.
This suggests the following definition.

Definition 1.5.1. Let H be a subgroup of a group G. If b−1 hb ∈ H


for any b ∈ G, h ∈ H, then H is called a normal subgroup of G,
denoted by H  G.

Assume that H  G. By the analysis at the beginning of this section


the binary operation (aH)(bH) = (ab)H on the set Γ is meaningful. It
is straightforward to verify that Γ becomes a group under this oper-
ation, called the quotient group of G by H, denoted by G/H. It is
obvious that |G/H| = (G : H).
24 CHAPTER 1. ELEMENTS OF GROUPS

Similar discussion applies to the right cosets as well. Can we obtain


another “quotient group”? The following property of normal subgroups
answers this question.
Proposition 1.5.2. Let H be a subgroup of a group G. Then H G
if and only if aH = Ha for any a ∈ G.
Proof. Assume that H  G. Then ah = (aha−1 )a ∈ Ha for any
h ∈ H. Hence aH ⊆ Ha. For the same reason Ha ⊆ aH. Therefore
aH = Ha.
Conversely assume that aH = Ha for any a ∈ G. For any a ∈ G
and any h ∈ H, the element ha belongs to aH, i.e,. ha = ah0 for some
h0 ∈ H. Hence a−1 ha = h0 ∈ H.

Thus when H  G there is no distinction between left cosets and


right cosets. We can simply call them cosets.
Keep in mind that an element in the quotient G/H is a coset aH,
which is usually denoted by ā or [a], called the coset represented by
the element a ∈ G.
The operation in many abelian groups is denoted by “+”. In that
occasion a coset is denoted by a + H.
All elements of G/H can be listed as

a1 H, a2 H, . . .

in which a−1
i aj ∈
/ H whenever i 6= j and for every a ∈ G there is
−1
some i such that a ai ∈ H.
Example 1.5.3. • SLn (K)  GLn (K). But the subgroup of
GLn (K) consisting of upper triangular matrices is not a normal
subgroup of GLn (K).

• H = {cIn |c ∈ K, c 6= 0} is a normal subgroup of GLn (K). The


quotient group GLn (K)/H is denoted by P GLn (K), called pro-
jective general linear group.

• Let H = {id, (123), (132)} ⊂ S3 . We know that H S3 in Example


1.4.3. The quotient group S3 /H contains two elements H and
(12)H, of which the first one is the identity element.
1.5. NORMAL SUBGROUPS AND QUOTIENT GROUPS 25

Example 1.5.4. K = {id, (12)} is a subgroup of S3 . The left coset


of K in S3 are
K = eK,

(13)K = {(13), (123)},

(23)K = {(23), (132)}.

The set {ab|a ∈ (13)K, b ∈ (23)K} = {(132), (23), (12), id} is not a left
coset of K.

This example shows that for a non-normal subgroup we can not


define a natural binary operation in the set of left cosets.

Example 1.5.5. • Trivial subgroups are always normal.


• All subgroups of an abelian group are normal.
• Assume that H G and K is a subgroup of G containing H. Then
H  K.

Example 1.5.6 ( an important example). Let n be a natural num-


ber greater than 1. Let nZ be the set of all integers divisible by n. Then
Z/nZ is a cyclic group of order n, whose elements can be enumerated
as
0̄, 1̄, 2̄, . . . , n − 1

in which 1̄ can be chosen to be the generator. Since the group Z/nZ


will be frequently used, it is abbreviated as Zn .

Example 1.5.7. The group Q of all rational numbers under addi-


tion is an abelian group. The quotient group Q/Z is an infinite group
of which every element has finite order.

Let H be a subgroup of a group G. Denote a−1 Ha = {a−1 ha|h ∈


H}. It is easy to verify that a−1 Ha is a subgroup of G, called a conju-
gate subgroup of H. Its order is equal to |H|. but a−1 Ha = H does
not hold in general. Readers may verity that H  G if and only if every
conjugate subgroup of H is equal to H.
In Example 1.4.3 we see that the left cosets {id, (123), (132)} in
S3 coincide with the right cosets. We may explain it without explicit
computation by the following observation.
26 CHAPTER 1. ELEMENTS OF GROUPS

Proposition 1.5.8. Let H be a subgroup of a group G such that


(G : H) = 2. Then H  G.

Proof. It suffices to show that gH = Hg holds for every g ∈ G.


If g ∈ H, then gH = H = Hg. Hence we may assume that g ∈ / H.
Then gH is a left coset different from H. Since (G : H) = 2, there are
only two left cosets of H in G. Hence gH = G\H. For the same reason
Hg = G\H. Therefore gH = Hg.

Example 1.5.9. The center of a group G is always a normal sub-


group of G.

Example 1.5.10. The intersection of arbitrarily many normal sub-


groups of a group is a normal subgroup.

Example 1.5.11. Let S be a subset of a group G. Assume that


g ag ∈ S for any g ∈ G, a ∈ S. Then hSi  G.
−1

Proof. Let b be an arbitrary element in hSi. Then

b = ae11 · · · aenn ,

in which ai ∈ S, ej = ±1. Hence

g −1 bg = (g −1 a1 g)e1 · · · (g −1 an g)en .

By the given condition we have g −1 bg ∈ hSi.

As a special case, let S = {a−1 b−1 ab|a, b ∈ G}. Since

g −1 (a−1 b−1 ab)g = (g −1 ag)−1 (g −1 bg)−1 (g −1 ag)(g −1 bg)−1 ∈ S,

S satisfies the above condition. Denote [G, G] = hSi, called the com-
mutator subgroup of G. It is a normal subgroup of G.

Example 1.5.12. Let H be a subgroup of a group G. Let

NG (H) = {g ∈ G|g −1 Hg = H}.


1.5. NORMAL SUBGROUPS AND QUOTIENT GROUPS 27

Then NG (H) is a subgroup of G containing H such that HNG (H). For


any subgroup K of G containing H and HK, the relation K ⊆ NG (H)
holds. The subgroup NG (H) of G is called the normalizer of H in G.

Proof. Assume that g ∈ NG (H). Then g −1 Hg = H, Hence gHg −1 =


H, which implies that g −1 ∈ NG (H). Hence NG (H) is closed under the
inverse operation.
Let a, b ∈ NG (H). Then (ab)−1 H(ab) = b−1 (a−1 Ha)b = b−1 Hb = H.
Hence ab ∈ NG (H), i.e., NG (H) is closed under multiplication. Hence
NG (H) is a subgroup of G. Obviously it contains H.
It follows from the definition of NG (H) that H  NG (H).
For an arbitrary a ∈ K, a−1 Ha = H since H K. Hence a ∈ NG (H).
This shows that K ⊆ NG (H).

Caution: H1  H2  G does not imply H1  G.

Proposition 1.5.13. For any group G the quotient group G/[G, G]


is an abelian group. If H  G and G/H is an abelian group then
[G, G] ⊆ H.

Proof. Let ā be an element in G/[G, G] represented by a ∈ G.


Then
ā−1 b̄−1 āb̄ = a−1 b−1 ab = 1̄.

Hence āb̄ = b̄ā for any a, b ∈ G. This shows that G/[G, G] is abelian.
Denote by [a] and element in G/H represented by a ∈ G. The
equality [a−1 b−1 ab] = [a]−1 [b]−1 [a][b] = [1] holds for any a, b ∈ G. Hence
a−1 b−1 ab ∈ H, which implies that [G, G] ⊆ H.

Definition 1.5.14. A groups is called a simple group if it contains


at least two elements and does not have a nontrivial normal subgroup.

Exercises
1. Show that the intersection of arbitrarily many normal subgroups
of a a group is normal.
2. Show that {e, (12)(34), (13)(24), (14)(23)} is a normal subgroup
of S4 .
28 CHAPTER 1. ELEMENTS OF GROUPS

3. Let r be a natural number. Assume that a group G has only one


subgroup H of order r. Show that H  G.
4. Let G be a group of order 120 and let H be a subgroup of G with
|H| = 24. Assume that there exist a, b ∈ G such that aH = Hb 6= H.
Show that H  G.
5. Give a counterexample to show that H1  H2  G does not imply
H1  G.
6. Let H, K be two subgroups of a group G. Let HK = {ab|a ∈
H, b ∈ K}.
1. Assume that H  G. Show that HK is a subgroup of G.

2. Let G = S3 , H = {e, (12)}, K = {e, (13)}. Show that HK is not


a subgroup of G.

3. Assume that H  G, K  G. Show that HK  G.

7. Let H be a normal subgroup of a finite group G and a ∈ G. Let


n be the order of aH in G/H and let m be the order of an in G. Show
that mn is the order of a in G.
8. Let N be a cyclic subgroup of a group G. Assume that N  G.
Show that ab = ba for every a ∈ N and every b ∈ [G, G].
9. Let G be a non-abelian finite group. Let Z be the center of G.
Show that 4|Z| ≤ |G|.
10. Let H1 , . . . , Hn be subgroups of a group G. Assume that ai aj =
aj ai for any 1 ≤ i < j ≤ n and any ai ∈ Hi , aj ∈ Hj . If every element
of G can be expressed as b1 b2 · · · bn , in which bi ∈ Hi . Show that Hi  G
for every i.
11. Assume that a finite group G has exactly two subgroup H, K
of order n and G is generated by H ∪ K. Show that H  G, K  G.

1.6 Alternating Groups

In this section we will study more properties of the symmetric group Sn


and its subgroups. Recall that An denotes the subset of Sn consisting
of all even permutations.
1.6. ALTERNATING GROUPS 29

Theorem 1.6.1. An is a normal subgroup of Sn and (Sn : An ) = 2


for all n > 1.
Proof. Assume that σ, τ ∈ An . They can be expressed as products
of even number of transpositions by Corollary 1.3.9. Hence στ can
also be expressed as a product of even number of transpositions. This
implies στ ∈ An . By the same argument An is closed under the inverse
operation. Hence An is a subgroup of Sn .
Since the map σ 7→ (12)σ is a bijection from An into the set (12)An
of odd permutations and every permutation is either even or odd, we
have Sn = An ∪ (12)An . Hence (Sn : An ) = 2. It follows from Proposi-
tion 1.5.8 that An  Sn .

The groups An is called the alternating group of n objects. By


the way, the theorem implies that Sn is not a simple group for n > 2.
The usefulness of the cycle notation for permutations is justified by
the following proposition and its corollaries.
Proposition 1.6.2. Let σ = (i1 · · · ir ) be an r-cyle and τ ∈ Sn .
Then τ στ −1 is also an r-cycle. More precisely,

τ στ −1 = (τ (i1 ) · · · τ (ir )).

Proof. If 1 ≤ i ≤ n does not appear in τ (i1 ), . . . , τ (ir ) then τ −1 (i)


does not appear in i1 , . . . , ir . Hence

τ στ −1 (i) = τ τ −1 (i) = i,

for every i that does not appear in τ (i1 ), . . . , τ (ir ).


The proposition follows from τ στ −1 (τ (ij )) = τ (ij+1 ) for all 1 ≤ j <
r and τ στ −1 (τ (ir )) = τ (i1 ).

Corollary 1.6.3. Let σ = (a1 · · · ar ) · · · (c1 · · · cs ) be the product


of cycles. Then

τ στ −1 = (τ (a1 ) · · · τ (ar )) · · · (τ (c1 ) · · · τ (cs )).

The following proposition shows that Sn and An can be generated


by some simple permutations.
30 CHAPTER 1. ELEMENTS OF GROUPS

Proposition 1.6.4.

Sn = h(12), (13), . . . , (1n)i = h(12), (12 · · · n)i.

An = h(123), (124), . . . , (12n)i.

Proof. The equality (1i)(1j)(1i) = (ij) and Proposition 1.3.5 im-


ply
Sn = h(12), (13), . . . , (1n)i.

It follows from Proposition 1.6.2 that (12 · · · n)(i, i + 1)(12 · · · n)−1 =


(i + 1, i + 2) for all 1 ≤ i ≤ n − 1. Hence

Sn = h(12), (12 · · · n)i.

Every element in An can be written as

(1m1 )(1m2 ) · · · (1m2r ).

It follows from
(1j)(1i) = (1ij) = (12j)(12i)−1

that
An = h(123), (124), . . . , (12n)i.

Theorem 1.6.5. The alternating group An is a simple group if n ≥


5.

Proof. Let K be a normal subgroup of An with |K| > 1. We need


to show that K = An .
First assume that K contains a 3-cycle (rst). For any i > 2, Propo-
sition 1.6.2 implies that σ(rst)σ −1 = (12i), where σ is any permutation
satisfying σ(r) = 1, σ(s) = 2, σ(t) = i. If σ ∈ / An , choose u, v such
that 1 ≤ u < v ≤ n and u ∈ / {r, s, t}, v ∈
/ {r, s, t}. Replace σ by
σ ◦ (uv). Then this new σ is in An and σ(rst)σ −1 = (12i) still holds.
Hence there is always some σ ∈ An such that σ(rst)σ −1 = (12i). So K
1.6. ALTERNATING GROUPS 31

contains (123), (124), . . . , (12n). If follows from Proposition 1.6.4 that


K = An . Hence K = An if K contains a 3-cycle.
For any α ∈ K and any β ∈ An , the element βαβ −1 is in K since
K  An . Hence βαβ −1 α−1 ∈ K. It suffices to find suitable α ∈ K and
β ∈ An such that βαβ −1 α−1 becomes a 3-cycle.
Since |K| > 1, there is an element α in K which is not the identity
element. If α is a 3-cycle, then the proof is done. Assume that α is not
a 3-cycle. Write α as a product of mutually disjoint cycles. At least
one of the following six cases occurs:
1) α contains a cycle of length greater than 4;
2) α contains a 4-cycle;
3) α contains two 3-cycles;
4) α contains four transpositions;
5) α is the product of two transpositions;
6) α is the product of one 3-cycle and two transpositions.
Let’s treat these cases one by one.
1) Without loss of generality assume that α = (12 · · · m)α1 , in
which m ≥ 5 and α1 is the product of mutually disjoint cycles that
do not contain any of the objects 1, 2, . . . , m. Let β = (345). Then
βαβ −1 α−1 = (134) (for m = 5) or (634) (for m > 5).
2) Assume that α contains a cycle (1234). Since α is an even permu-
tation, it contains at least one more cycle, say (56 · · · ). Let β = (345).
Then βαβ −1 α−1 = (15346). This brings us back to Case 1).
3) Assume that α contains two cycles (123), (456). Let β = (124).
Then βαβ −1 α−1 = (12534), reduced to Case 1).
4) Assume that α contains four transpositions (12), (34), (56), (78).
Let β = (135). Then βαβ −1 α−1 = (135)(264), reduced to Case 3).
5) Assume that α = (12)(34). Let β = (135). Then βαβ −1 α−1 =
(13542), reduced to Case 1).
6) Assume that α = (12)(34)(567). Let β = (135). Then βαβ −1 α−1 =
(135)(264), reduced to Case 3.

Exercises
1. Let G be a subgroup of the symmetric group S6 . Assume that G
has an element of order 6. Show that G contains a normal subgroup H
32 CHAPTER 1. ELEMENTS OF GROUPS

such that (G : H) = 2.
2. Show that A4 is not a simple group.
3. Let S be the group of all bijections from N to N under the
composite operation. Let

G = {σ ∈ S| there exists n > 0 such that σ(i) = i ∀ i > n

and the number of pairs out of order is even }.

Show that G is a simple group.


4. Let G be a subgroup of Sn with |G| > 2. Assume that G is a
simple group. Show that G ⊆ An .
5. Let G be a normal subgroup of Sn (n > 1) satisfying (Sn : G) = 2.
Show that G = An .
6. Compute the commutator subgroup [Sn , Sn ] of Sn . Then compute
the commutator subgroup of [Sn , Sn ]. (Hint: Discuss the cases n ≤
2, n = 3, n = 4, n ≥ 5.)
7. Two elements x, y in a group G are called conjugate if there
exists g ∈ G such that g −1 xg = y.
i) Show that the conjugacy in G is an equivalence relation;
ii) Show that every element in Sn is conjugate to its inverse;
iii) Find an element in A4 which is not conjugate to its inverse.
8. Let γ = (a1 , a2 , · · · , an ) be an n-cycle in Sn and β ∈ Sn . Assume
that βγ = γβ. Show that there is some integer k such that β = γ k .

1.7 Homomorphisms of Groups

Definition 1.7.1. Let G1 , G2 two groups with identity elements e1


and e2 respectively. A map f : G1 → G2 is called a homomorphism
if f (ab) = f (a)f (b) for any a, b ∈ G1 .
If f is a homomorphism, then f (e1 ) = e2 and f (a−1 ) = f (a)−1 for
any a ∈ G1 .
For a homomorphism f, denote Ker(f ) = {a ∈ G1 |f (a) = e2 }. This
is a normal subgroup of G1 , called the kernel of f. The homomorphism
1.7. HOMOMORPHISMS OF GROUPS 33

f is injective if and only if Ker(f ) = {e1 }. An injective homomorphism


is called a monomorphism.
Denote Im(f ) = {f (a)|a ∈ G1 }. This is a subgroup of G2 , called the
image of f. The homomorphism f is surjective if and only if Im(f ) =
G2 . A surjective homomorphism is called an epimorphism.
A bijective homomorphism is called an isomorphism.
If there is an isomorphism between two groups G1 , G2 then we say
that G1 and G2 are isomorphic, denoted by G1 ∼ = G2 .
An isomorphism from G to itself is called an automorphism.
That two groups are isomorphic means that they have the same
group structure, which means that they are identical if all other proper-
ties of the underlying sets are ignored. Sometimes two different groups
are considered to be the same if they are isomorphic. For instance, all
groups with two elements are isomorphic, so we can say that there is
only one group of order two. If we say that there are only two groups
of order 6, it means that there exist two non-isomorphic groups G1 and
G2 with |G1 | = |G2 | = 6 and any group of order 6 is isomorphic to
either G1 or G2 .
If G1 ∼
= G2 then there is an isomorphism f : G1 → G2 , but such an
isomorphism f is not necessarily unique.
Let j : H → G be a monomorphism. Then H is isomorphic to
the subgroup j(H) of G. We may regard H as a subgroup of G under
the monomorphism j, or simply say that H is a subgroup of G, if no
confusion will be caused.

Example 1.7.2. The f : G1 → G2 , g 7→ e2 for all g ∈ G1 is a ho-


momorphism with Ker(f ) = G1 , Im(f ) = {e2 }. Such a homomorphism
is called a trivial homomorphism.

Example 1.7.3. Let H be a subgroup of a group G. Then the


injective map i : H → G, h 7→ h is a monomorphism.

Example 1.7.4. Assume that H  G. Then π : G → G/H, a 7→ ā


is an epimorphism, called the canonical homomorphism from G to
G/H.

Example 1.7.5. Let K be a number field. Denote by K ∗ the group


34 CHAPTER 1. ELEMENTS OF GROUPS

of all nonzero elements in K under multiplication. Then

det : GLn (K) → K ∗ , A 7→ |A|

is an epimorphism. Here |A| is the determinant of A.

Example 1.7.6. Let


" #
n 1 x o
G= x ∈ R .

0 1

Then G is a group under multiplication. The map


" #
1 x
f : G → R, 7→ x
0 1

is an isomorphism from G to the additive group of real numbers.

Example 1.7.7. Let g be an element in a group G. Then the map

f : G → G, a 7→ g −1 ag

is an automorphism of G, called an inner automorphism.

For example, if G = GLn (K) then the inner automorphism is the


similarity in linear algebra.
To show that two groups are isomorphic requires the construction
of a homomorphism which is both injective and surjective. The veri-
fication can be straightforward or difficult depending on the problem.
To show that two groups are not isomorphic, it is enough to find a
property that one group possesses but the other group does not. A
simple example is given below.

Example 1.7.8. Although Z/6Z and S3 have the same order 6, but
they are not isomorphic since the former is abelian while the latter is
not.

Theorem 1.7.9 (Fundamental Theorem of Homomorphisms). Let


f : G1 → G2 be a homomorphism of groups. Then
1) Ker(f )  G1 ;
1.7. HOMOMORPHISMS OF GROUPS 35

2) G1 /Ker(f ) ∼
= Im(f ).

Proof. We have mentioned before that Ker(f )  G1 . The verifica-


tion is easy and is left as an exercise.
In order to prove the second statement we construct a map p :
G1 /Ker(f ) → Im(f ) as follows. For an arbitrary aKer(f ) ∈ G1 /Ker(f ),
define p(aKer(f )) = f (a). We need to verify that this definition makes
sense, i.e., p(aKer(f )) does not depend upon the choice of the rep-
resentative a. Assume that a0 ∈ aKer(f ) is another representative of
aKer(f ). Then a0 = ah for some h ∈ Ker(f ). Thus f (a0 ) = f (a)f (h) =
f (a)e2 = f (a). Hence the map p is well-defined.
It is evident that p is surjective. It remains to show that p is injec-
tive. Assume that aKer(f ), bKer(f ) ∈ G/Ker(f ) satisfy f (a) = f (b).
Then f (a−1 b) = e2 . Hence a−1 b ∈ Ker(f ), which implies aKer(f ) =
bKer(f ).

The following example displays a typical application of Theorem


1.7.9.
Example 1.7.10. Let n ∈ N. Every cyclic group of order n is iso-
morphic to Zn . Hence two cyclic groups are isomorphic if and only if
they have the same order.

Proof. Let G be a cyclic group of order n with generator g. Define


a map
f : Z → G, m 7→ g m .

Then f is an epimorphism with nZ as its kernel. It follows from The-


orem 1.7.9 that
Zn ∼
= G.

Hence all cyclic finite groups of the same order are isomorphic. It is
easy to see that every infinite cyclic group is isomorphic to Z.

From this example we see that in order to prove that G/H is isomor-
phic to another group K we need to construct a surjective epimorphism
f : G → K such that H = Ker(f ). The proof of this style is usually
concise and convincing.
36 CHAPTER 1. ELEMENTS OF GROUPS

Lemma 1.7.11. Let f : G1 → G2 be a homomorphism of groups.


Let H2 be a subgroup of G2 . Then f −1 (H2 ) = {g ∈ G1 |f (g) ∈ H2 } is a
subgroup of G1 containing Ker(f ).

Proof. Assume that a, b ∈ f −1 (H2 ). Then f (a), f (b) ∈ H2 . Thus


f (ab−1 ) = f (a)f (b)−1 ∈ H2 . Hence ab−1 ∈ f −1 (H2 ). Therefore f −1 (H2 )
is a subgroup of G1 .
Assume that a ∈ Ker(f ). Then f (a) = e2 ∈ H2 . So a ∈ f −1 (H2 ),
which implies Ker(f ) ⊆ f −1 (H2 ).

The lemma can be stated in plain English: the inverse image of any
subgroup is a subgroup containing the kernel.

Proposition 1.7.12. Assume that H  G and f : G → G/H is


the canonical homomorphism. Let Γ be the set of all subgroups of G
containing H and let Γ0 be the set of all subgroups of G/H. Then the
map f −1 : Γ0 → Γ, K 7→ f −1 (K) is a bijection from Γ0 to Γ.

Proof. By Lemma 1.7.11 the map f −1 : Γ0 → Γ is well-defined.


Construct a map φ : Γ → Γ0 as follows. For any T ∈ Γ define T 0 to be
the set of all cosets aH with a ∈ T. Then T 0 is a subset of G/H. It is
easy to verify that T 0 is a subgroup of G/H. Define φ(T ) = T 0 .
The inclusion T ⊆ f −1 (T 0 ) is obvious. If b ∈ f −1 (T 0 ), then f (b) ∈
T 0 , which means that there is some a ∈ T such that bH = aH. Hence
a−1 b ∈ H. Thus b ∈ T, since H ⊆ T. Hence f −1 (T 0 ) = T, which implies
that f −1 ◦ φ is the identity map. It is easy to verify that φ ◦ f −1 is also
the identity map.

We have seen before that every subgroup of an infinite cyclic group


is an infinite cyclic group. Let us classify the subgroups of finite cyclic
groups. Assume that G = Zn , in which n is a natural number. By
Proposition 1.7.12 there is one-to-one correspondence between the set
of subgroups of G and the set of subgroups of Z containing nZ. A
subgroup mZ of Z contains nZ if and only if m|n. All subgroups of G
are thus determined.
1.7. HOMOMORPHISMS OF GROUPS 37

Theorem 1.7.13 (First Isomorphism Theorem). Assume that H 


G, N  G and H ⊆ N. Then

(G/H)/(N/H) ∼
= G/N.

Proof. It is easy to see that H  N and N/H  G/H. So all the


quotient groups involved in the theorem are well-defined. The inclusion
relations of these groups are described by the following diagram:

G G/H

H N/H
Let φ : G → (G/H)/(N/H) be the composite of the canonical
homomorphisms G → G/H and G/H → (G/H)/(N/H). Then φ is
evidently an epimorphism. It is obvious that N ⊆ Ker(φ). Let g ∈
Ker(φ). Then gH ∈ N/H. Hence gH = sH for some s ∈ N. So g ∈ N.
Hence N = Ker(φ). The theorem follows from Theorem 1.7.9.

Theorem 1.7.14 (Second Isomorphism Theorem). Assume that


H  G and K is a subgroup of G. Let

KH = {ab|a ∈ K, b ∈ H}.

Then KH is a subgroup of G and

KH/H ∼
= K/K ∩ H.

Proof. It follows from Exercise 6 of Section 5 that KH is a sub-


group of G.
The inclusion relations of the five groups involved are described by
the following diagram:
38 CHAPTER 1. ELEMENTS OF GROUPS

KHHH
v
vv HH
vv HH
vv HH
vv H
H HH vK
HH vv
HH vv
HH vv
H vv
K ∩H
Assume that st ∈ KH. Then s = ab, t = cd in which a, c ∈ K and
b, d ∈ H. Hence st−1 = abd−1 c−1 = (ac−1 )(cbd−1 c−1 ) ∈ KH. which
implies that KH is a subgroup of G.
It is clear that H  KH and K ∩ H  K.
Define φ : K → KH/H, a 7→ aH. This is an epimorphism with
Ker(φ) = K ∩ H. The theorem follows from Theorem 1.7.9.

Exercises
1. Let f : G1 → G2 be a homomorphism of groups. Verify
1) f (e1 ) = e2 , where e1 , e2 are the identity elements of G1 , G2 ;
2) f (a−1 ) = f (a)−1 for any a ∈ G1 ;
3) Ker(f )  G1 .
2. Let H be a commutative subgroup of a group G such that n =
(G : H) < ∞. Let
b1 H, b2 H, . . . , bn H

be all left cosets of H in G. Define a map τ : G → H by the following


rule: For every a ∈ G and every 1 ≤ i ≤ n, the element abi belongs to
exact one bj H, i.e. there exists a unique hi ∈ H such that abi = bj hi .
Define n
Y
τ (a) = hi .
i=1

Show that τ is a homomorphism.


3. Show that the additive group C is not isomorphic to GL2 (R).
4. Let G be a finite group. Let N be a normal subgroup of G and
let H be a subgroup of G. Assume that |N | and (G : H) are coprime.
Show that N ⊆ H.
1.8. DIRECT PRODUCT OF GROUPS 39

5. Let G = C−{1}. In G define a binary operation a◦b = a+b−ab.


Verify that G is an abelian group under this operation and show that it
is isomorphic to the multiplicative group of nonzero complex numbers.
( The problems 6-10 involve automorphisms.)
Let G be a group and let Aut(G) denote the set of all automorphism
of G. Define a binary operation on Aut(G) in the following way. For
σ, τ ∈ Aut(G), define στ to be the composite of σ and τ, i.e., (στ )(a) =
σ(τ (a)) for every a ∈ G.
6. Show that Aut(G) is a group under the operation defined above.
This group is called the group of automorphisms of G.
7. Show that the set consisting of all inner automorphism is a
normal subgroup of Aut(G), denoted by Inn(G).
8. Show that Inn(G) ∼
= G/C(G). Here C(G) is the center of G.
9. Let G be an infinite cyclic group. Show that Aut(G) is a group
of order 2.
10. Determine all automorphisms of the additive group Q of rational
numbers.
11. Let N be a normal subgroup of a group G. If a subgroup H of
G satisfies N H = G, N ∩ H = {e}, then H is called a compliment sub-
group of N. Show that all compliment subgroups of N are isomorphic.
12. Let G = Q/Z as in Example 1.5.7. Show that G has a unique
subgroup of order n for every natural number n.
13. Let H be a maximal subgroup of a finite group G, i.e., H 6= G
and there is no subgroup between H and G. Assume that (G : H) is
not a prime number. Show that H is not a normal subgroup of G.

1.8 Direct Product of Groups

So far we have learned two methods to get new groups from old ones:
subgroups and quotient groups. Both methods yield smaller groups. In
this section we introduce the most common construction that combines
more than one groups into a single group.
40 CHAPTER 1. ELEMENTS OF GROUPS

Recall that in linear algebra the space Rn can be regarded as the


direct sum of n one-dimensional spaces.

Rn = R ⊕ · · · ⊕ R
= {(x1 , . . . , xn )|x1 , . . . , xn ∈ R}.

Let G1 , . . . , Gn be groups. Let

G1 × · · · × Gn = {(x1 , . . . , xn )|xi ∈ Gi }.

For arbitrary (a1 , . . . , an ), (b1 , . . . , bn ) ∈ G1 × · · · × Gn , define the prod-


uct
(a1 , . . . , an )(b1 , . . . , bn ) = (a1 b1 , . . . , an bn ).

It is easy to see that G1 × · · · × Gn is a group under this operation,


called the direct product of G1 , . . . , Gn , denoted by G1 × · · · × Gn
or ni=1 Gi . Obviously, the identity element of ni=1 Gi is (1G1 , . . . , 1Gn
Q Q

and the inverse of (a1 , . . . , an ) is (a−1 −1


1 , . . . , an ).
When Gi ’s are abelian groups and the operations are denoted by
“+”, the direct product is also called direct sum, denoted by G1 ⊕
· · · ⊕ Gn or ⊕ni=1 Gi . The addition is given by

(a1 , . . . , an ) + (b1 , . . . , bn ) = (a1 + b1 , . . . , an + bn )

for any (a1 , . . . , an ), (b1 , . . . , bn ) ∈ G1 ⊕· · ·⊕Gn . An element (a1 , . . . , an )


in G1 ⊕ · · · ⊕ Gn can also be denoted by a1 ⊕ · · · ⊕ an .
Let G1 , . . . , Gn be groups. For each index 1 ≤ i ≤ n there are two
homomorphisms:

ji : G i → G 1 × · · · × G n

a 7→ (1G1 , . . . , 1Gi−1 , a, 1Gi+1 , . . . , 1Gn )

and
p i : G1 × · · · × Gn → Gi

(a1 , . . . , an ) 7→ ai .

Obviously ji is a monomorphism such that ji (Gi )  G1 × · · · × Gn and


1.8. DIRECT PRODUCT OF GROUPS 41

pi is an epimorphism. It is clear that (a1 , . . . , an ) = j1 (a1 ) · · · jn (an ) for


any (a1 , . . . , an ) ∈ G1 × · · · × Gn .
Due to the monomorphism ji , we may treat Gi as a subgroup of
G1 × · · · × Gn . For instance, an element in the form (a, 1, . . . , 1) in
G1 × · · · × Gn can be regarded as the element a in G1 . As we have seen,
Gi  G1 × · · · × Gn for every 1 ≤ i ≤ n.
Once we know the meanings of direct product and direct sum, the
next task is the decomposition of a group into direct product or direct
sum, which is harder.
In order to understand the structure of a group, we usually start
with its decomposition into direct product (or direct sum) of some
smaller groups. For this purpose we need to find some groups G1 , . . . , Gn
and an isomorphism

φ : G1 × · · · × Gn → G.

It will be more desirable if every Gi cannot be decomposed further. If


every component Gi is some group that we are familiar with, then we
may consider the structure of G is understood.
How to find these G1 , . . . , Gn ? By the analysis before, if there exists
an isomorphism
φ : G1 × · · · × Gn → G

Then all subgroups H1 = φ ◦ j1 (G1 ), . . . , Hn = φ ◦ jn (Gn ) are normal


subgroups of G and ai aj = aj ai for any ai ∈ Hi , aj ∈ Hj . So all normal
subgroups of G are the candidates of Gi .

Theorem 1.8.1. Let G, G1 , . . . , Gn be groups. Then

G∼
= G1 × · · · × Gn

if and only if there are subgroups H1 , . . . , Hn of G satisfying the follow-


ing three conditions:
1) Hi ∼= Gi for each 1 ≤ i ≤ n;
2) ai aj = aj ai for any ai ∈ Hi , aj ∈ Hj with i 6= j;
3) the map ψ : H1 × · · · × Hn → G, (a1 , . . . , an ) 7→ a1 · · · an is
bijective.
42 CHAPTER 1. ELEMENTS OF GROUPS

Remark 1.8.2. Exercise 1.5.10 shows that Hi  G is a consequence


of 2) and 3).

Proof. We first make a remark on the third condition. H1 × · · · ×


Hn is not a subset of G, although every Hi is a subgroup of G.
Assume that there is an isomorphism φ : G1 × · · · × Gn → G. Let
Hi = φ · ji (Gi ). It is clear that the subgroups H1 , . . . , Hn of G satisfy
all three conditions.
Conversely assume that the three conditions are satisfied for H1 , . . . , Hn .
It suffices to show that ψ is a homomorphism. Let a = (a1 , . . . , an ), b =
(b1 , . . . , bn ) be two elements of H1 × · · · × Hn . Then

ψ(ab) = a1 b1 a2 b2 · · · an bn .

It follows from the second condition that

a1 b1 a2 b2 · · · an bn = a1 a2 · · · an b1 b2 · · · bn = ψ(a)ψ(b).

Hence ψ is a homomorphism.

In application, the surjectivity of ψ is usually easy to verify. The in-


jectivity can be verified by showing that Hi ∩(H1 · · · Hi−1 Hi+1 · · · Hn ) =
{e} for each i.

Example 1.8.3. A cyclic group of order 6 is the direct product of


a cyclic group of order 2 and a cyclic group of order 3.

Proof. Let G = Z/6Z. Then G has a subgroup H1 = {0̄, 3̄} of


order 2 and a subgroup H2 = {0̄, 2̄, 4̄} of order 3. Obviously, H1 and
H2 satisfy all three conditions in the theorem.

Example 1.8.4. The infinite cyclic group is not isomorphic to the


direct product of two proper subgroups.

Proof. Otherwise Z would have two proper subgroups nZ and mZ


satisfying the three conditions of the theorem. According the third
condition nZ ∩ mZ = {0} would hold. This is absurd, since nm ∈
nZ ∩ mZ.
1.9. * AUTOMORPHISMS OF FINITE CYCLIC GROUPS AND THE EULER FUNCTION43

Exercises

1. Let G be a finite group and let G1 = G × G. Assume that G1 has


exactly four normal subgroups. Show that G is not an abelian group.

2. Let G be an abelian group whose operation is denoted by “+”.


Let u, v : G → G be homomorphisms. Define two maps from G to G
by

f (x) = x − v(u(x)), g(x) = x − u(v(x)).

Show that f is surjective if and only if g is surjective.

3. Let H be a subgroup of an abelian group G. Let Ḡ = G/H. Are


the following statements true? If true give a proof, otherwise give a
counterexample.
1) If Ḡ is a finite cyclic group, then G ∼
= Ḡ × H;
2) If Ḡ is an infinite cyclic group, then G ∼
= Ḡ × H.

4. Let N1 , N2 , N3 be normal subgroups of G. Assume that Ni ∩Nj =


{e}, G = Ni Nj for any 1 ≤ i < j ≤ 3. Show that G is an abelian group
and N1 , N2 , N3 are mutually isomorphic.

5. Let S, T be two groups, G = S ×T. Assume that a subgroup H of


G satisfies the condition SH = G = T H. (Here S and T are considered
to be the subgroups of G.)
i) Show that S ∩ H  G, T ∩ H  G;
ii) If S ∩ H = {e} = T ∩ H, show that S ∼
= T.
iii) If S ∩ H = {e} = T ∩ H and H  G, show that G is an abelian
group.

1.9 * Automorphisms of Finite Cyclic Groups and


the Euler Function

Let G = hai be a cyclic group of finite order n. Then

G = {1, a, a2 , . . . , an−1 }.
44 CHAPTER 1. ELEMENTS OF GROUPS

Let m be a natural number smaller than n and coprime with n. The


map
σm : G → G, b 7→ bm

is a homomorphism. Assume ak ∈ Ker(σm ). Then amk = 1. Thus n|mk.


Since m and n are coprime, so n|k. Hence ak = 1, which implies that
σm is a monomorphism. Since every monomorphism from a finite group
to itself is bijective, σm is bijective. Hence σm ∈ Aut(G).
Conversely assume that σ ∈ Aut(G). Then σ(a) = am for some
m with 0 ≤ m ≤ n − 1. Let d be the greatest common divisor of m
and n. Suppose that d > 1. Let n0 = n/d. Then 1 ≤ n0 < n and
0 0 0
σ(an ) = amn = 1. Hence an ∈ Ker(σ). This would contradict the
assumption that σ is an isomorphism. Therefore d = 1, i.e., m and n
are coprime.
Let Z∗n denote the set of all natural numbers less than n and co-
prime with n. The above discussion shows that there is a one-to-one
correspondence between Aut(G) and Z∗n . For any m, m0 ∈ Z∗n , The com-
0
posite of the automorphisms σm and σm0 is the isomorphism b 7→ bmm .
Let r be the remainder of mm0 divided by n. Then σm σm0 = σr .
This means that the set Z∗n becomes a group under the multiplica-
tion mm0 = r and
Aut(G) ∼= Z∗n .

Z∗n is an abelian group, but not necessarily a cyclic group. For


example Z∗8 consists of four elements 1, 3, 5, 7. It is easy to verify that
it does not contain an element of order 4.
The function φ(n) = |Z∗n | is an important arithmetic function, called
the Euler function. A few of its values are

φ(2) = 1, φ(3) = 2, φ(4) = 2, φ(5) = 4, . . .

Lagrange’s theorem implies the celebrated Euler’s Theorem in ele-


mentary number theory:

If two natural numbers m and n are coprime, then the remainder


of mφ(n) divided by n is equal to 1.
1.10. GROUP ACTION 45

The following statement is a special case of Euler’s theorem.

Let p be a prime number and let a be an integer not divisible by p.


Then the remainder of ap−1 divided by p is equal to 1.

This result is known as Fermat’s little theorem.

1.10 Group Action

Let G be a group and let S be a set. If there is a map σ : G × S → S


satisfying the following two conditions:
1) σ(e, x) = x for any x ∈ S;
2) σ(g1 g2 , x) = σ(g1 , σ(g2 , x)) for any x ∈ S and any g1 , g2 ∈ G,
then σ is called a (left) action of G on S.
For a fixed g ∈ G, a map from S into itself is defined by x 7→ σ(g, a).
It is more convenient to denote σ(g, a) by ga. We should be aware
that gx is different from the multiplication in a group, since the two
elements g and x involved in the action are contained in two different
sets G and S. The element g is sometimes called an “operator”. Under
the notation gx, the two conditions in the definition can be rewritten
as
1) ex = x for any x ∈ S;
2) (g1 g2 )x = g1 (g2 x) for any x ∈ S and any g1 , g2 ∈ G.
The expression g1 g2 on the left hand side of 2) is the product of g1
and g2 in G, while the right hand side is the result to two successive
actions. The second condition guarantees that the notation g1 g2 x is
meaningful.
A right action of G on S can be defined similarly. It satisfies the
following conditions:
1) xe = x for any x ∈ S;
2) x(g1 g2 ) = (xg1 )g2 for any x ∈ S and any g1 , g2 ∈ G.
The main difference between the left action and the right action is
the difference of the order in which g1 g2 acts on x ∈ S. For the left
action, x is acted by g2 followed by g1 while the right action takes the
reverse order.
46 CHAPTER 1. ELEMENTS OF GROUPS

Proposition 1.10.1. Assume that G has a left action on a set S.


Then for any g ∈ G the map x 7→ gx is a bijection of S onto itself.

Proof. This follows from g −1 (gx) = (g −1 g)x = ex = x.

The following examples of group actions are easy to verify.

Example 1.10.2. • Every group has a trivial action on any set S.


That is, gx = x, ∀g ∈ G, ∀x ∈ S.
• GLn (K) acts on the space of n-dimensional column vectors by
multiplying from the left. This is a left action. GLn (K) acts on the
space of n-dimensional row vectors by multiplying from the right. This
is a right action.
• An action of G on S induces an action of any subgroup of G on
S.
• G has a left action on itself (the latter is treated as a set instead
of a group):
G×G→G

(g, a) 7→ ga,

called a left translation.


This is an important action, the underlying set of the operator group
is identical to the set on which G acts. The right translation is defined
similarly.
• There is another important left action of a group G on itself:

G×G→G

(g, a) 7→ gag −1 ,

called a conjugation. It is not appropriate to denote this action by


ga. Similarly g −1 ag is a right action of G on itself. The translation and
conjugation are two frequently used actions in group theory.
• Let H be a subgroup of a group G. Let S be the set of all left
cosets of H in G. Then (g, aH) 7→ (ga)H is a left action of G on S.
• Assume that H  G. Then g, aH 7→ (gag −1 )H is a left action of
G on the quotient group G/H.
1.10. GROUP ACTION 47

Due to Proposition 1.10.1 a left action of a group G on a finite set


S induces a natural homomorphism φ : G 7→ Sn , where n is equal to
the number of elements in S. Converse, it is easy to see that every
homomorphism from G to Sn is induced from a left action of G on a
set of n elements. In fancier terminology, a group action is nothing
but a “group representation”. This point of view is useful in dealing
with some problems. A typical example is the so called n! theorem as
follows.
Example 1.10.3. Let H be a subgroup of a finite group G with
n = (G : H) > 1. If |G| > n!, then G is not a simple group.
Proof. Let S be the set of all left cosets of H in G. Then |S| = n.
There is a left action of G on the set S : g(aH) = (ga)H, ∀g, a ∈ G.
This determines a homomorphism φ from G to Sn . Since H is a proper
subgroup of G, Ker(φ) 6= G. Since |Im(φ)| ≤ n!, the fundamental
theorem of homomorphisms implies that (G : Ker(φ)) ≤ n!. Hence
|Ker(φ)| ≥ |G|/n! > 1, which implies that Ker(φ) is a nontrivial normal
subgroup of G.

From now on, by default an action is a left action, unless otherwise


stated.
Definition 1.10.4. Assume that G acts on a set S. For an element
x ∈ S, the set
Gx = {gx|g ∈ G}

is called an orbit. The number of elements in an orbit is called the


length of that orbit. The stabilizer of an element x ∈ S is defined to
be
Stab(x) = {g ∈ G|gx = x}.
It is easy to check that Stab(x) is indeed a subgroup of G. The
stabilizer of x can also be denoted by Gx .
Lemma 1.10.5. For x, y ∈ S, if Gx ∩ Gy 6= ∅ then Gx = Gy.
Proof. Assume that g1 x = g2 y. Then g1−1 g2 y = x. Hence gx =
gg1−1 g2 y ∈ Gy for any g ∈ G. Therefore, Gx ⊆ Gy. For the same
reason, Gy ⊆ Gx.
48 CHAPTER 1. ELEMENTS OF GROUPS

This lemma implies that the orbits give a partition of the set S,
which is useful in counting the number of elements in S.

Definition 1.10.6. An action with only one orbit is called a tran-


sitive action.

In other words, an action is transitive if and only if there exists


g ∈ G such that gx = y for any x, y ∈ S.
For example the left translation of a group is transitive while the
conjugation of a nontrivial group is not transitive.
Assume that a G acts on a set S and x ∈ S. The orbit Gx is closely
related to the stabilizer Gx . Let Γ denote the set of all left cosets of Gx
in G.

Theorem 1.10.7. There is a one-to-one correspondence between the


set Γ and the orbit Gx.

Proof. Define a map

φ : Γ → Gx, gGx 7→ gx.

Let us check that this map is well-defined. Assume that g 0 = gh is


another representative of the left coset gGx in which h ∈ Gx . Then

g 0 x = (gh)x = g(hx) = gx.

Hence φ is well-defined. It is evident that φ is surjective. It remains to


show that φ is injective.
Assume that φ(g1 Gx ) = φ(g2 Gx ). Then g1 x = g2 x. So g1−1 g2 x = x,
i.e., g1−1 g2 ∈ Gx , which implies g1 Gx = g2 Gx . Hence φ is injective.

Corollary 1.10.8. Assume that a group G acts on a set S and


x ∈ S. Then
|Gx| = (G : Gx ).

Example 1.10.9. Let U (2) be the 2 × 2 unitary group. It consists


of all 2 × 2 unitary matrices. Let

S3 = {(u, v) ∈ C2 | |u|2 + |v|2 = 1}.


1.10. GROUP ACTION 49

Then S3 is the unit sphere in the 4-dimensional Euclidean space, i.e.,

S3 = {(x1 , x2 , x3 , x4 ) ∈ R4 |x21 + x22 + x23 + x24 = 1}.

By the definition of unitary matrices,


" #
u
A ∈ S3
v

holds for A ∈ U (2). This gives an action of U (2) on S3 . By the Gram-


Schmidt process in linear algebra, this action is transitive. Hence the
map " #
1
f : U (2) → S3 , A 7→ A
0
is surjective. By Theorem 1.10.7 the inverse image of any (u, v) ∈ S3
in U (2) is a right coset of
" #!
1
Stab .
0

In particular,
" #! " #!
1 1
f −1 = Stab .
0 0

It is not hard to see that


" #!
1
Stab = U (1),
0

which is the set of all complex numbers with modulus 1, which is a


unit circle on the complex plane.
This analysis enables us to visualize U (2) in some degree. Since
3-dimensional sphere is bounded as well as the unit circle. There is
reason to believe that U (2) is also bounded. In the terminology of
topology, U (2) is a compact manifold. In term of Lie theory, U (2) is a
compact Lie group.
50 CHAPTER 1. ELEMENTS OF GROUPS

Exercises
1. Assume that a group G acts on a set S. Define a relation on S
by x ∼ y if and only if there is g ∈ G such that gx = y. Show that
∼ is an equivalence relation and an equivalence class is an orbit of the
given action.
2. A subset S of a group G is called a conjugacy class if there is an
element a ∈ G such that

S = {bab−1 |b ∈ G}.

Assume that G is a finite group.


1) Show that the number of elements in any conjugacy class divides
the order of G;
2) Do all conjugacy classes contain same number of elements ?
3) If G has only two conjugacy classes, show that |G| = 2.
3. Find all conjugacy classes of the special linear group SL2 (C).
4. Assume that a finite group G acts on a finite set S with n =
|S| > 2 such that for any x1 6= x2 , y1 6= y2 ∈ S there exists g ∈ G such
that gx1 = y1 , gx2 = y2 . show that n(n − 1) divides |G|.
5. Assume that a group G acts on a set S with |S| > 1 transitively.
Show that Gx Gy is a proper subset of G for any two distinct elements
x, y of S.
6. Regard S4 as a subgroup of S5 in the natural way ( i.e., the sub-
group of S5 consisting of all permutations fixing the object 5). Define
an action of S4 × S4 on S5 by

(σ1 , σ2 )g = σ1 gσ2−1 .

Find the total number of orbits and the length of each orbit.
Chapter 2

Elements of Rings and Fields

2.1 Basic Definitions

There are more than one operation in many sets that we are familiar
with. These operations obey certain common rules. Typical examples
are set of integers Z and set of all n × n real matrices Mn (R), on which
there are two basic operations — addition and multiplication.
Definition 2.1.1. Let R be a nonempty set. Assume that R has
two binary operations +, · satisfying the following conditions:
1) R is an abelian group under the addition +;
2) a · (b · c) = (a · b) · c for any a, b, c ∈ R;
3) (a + b) · c = a · c + b · c and a · (b + c) = a · b + a · c for any a, b, c ∈ R.
Then R is called a pseudo-ring.
If there is an element e ∈ R such that e · a = a · e = a for any a ∈ R,
then e is called a unity of R (or multiplicative identity). A psudo-ring
with unity is defined to be a ring.
If a ring F satisfies one more condition
4) a · b = b · a for any a, b ∈ R,
then this ring is called a commutative ring, otherwise it is a non-
commutative ring.

Remark 2.1.2. The multiplication a · b in a ring is usually denoted


by ab. The identity element is usually called the zero element and is
denoted by 0, and the unity is usually denoted by 1. They can also be
denoted by 0R and 1R to avoid confusion.

51
52 CHAPTER 2. ELEMENTS OF RINGS AND FIELDS

Let a ∈ R and let n be a natural number. The sum and product of


n copies of a is denoted by na and an respectively.

Example 2.1.3. • By definition a ring contains at least one element.


Is there a ring with only one element? The answer is yes. Let R be a
set containing only one element a. Define a + a = a, aa = a. Then R is
a ring in deed. The element a serves as both zero element and unity.
This ring is simply denoted by 0, called the zero ring.
• Z is a commutative ring under usual addition and multiplication,
called the ring of integers.
• Mn (R) is a ring under the addition and multiplication of matrices,
which is non-commutative when n ≥ 2. Note the difference between
Mn (R) and GLn (R).
• The set of all polynomial K[x] over a number field K is a com-
mutative ring under usual addition and multiplication, called the poly-
nomial ring over K. Later on we will see that the ring of integers and
the polynomial ring share many common properties.

Proposition 2.1.4 (Basic properties of rings). 1) The unity is


unique.
2) 0 · a = a · 0 = 0 for any a ∈ R.
3) (−a) · b = a · (−b) = −(a · b) for any a, b ∈ R.
4) X
(a1 + · · · + am ) · (b1 + · · · + bn ) = ai b j
1≤i≤m,1≤j≤n

for any a1 , . . . , am , b1 , . . . , bn ∈ R.
5) If ab = ba, then
n  
n
X n
(a + b) = ai b j ;
i=0
i

6) In a nonzero ring, i.e, a ring with more than one element, the
unity 1 is not equal to 0.

The proof is left as an exercise.


We have learned in Chapter 1 that na is the sum of n copies of a
for any a ∈ R and any natural number n. There are two interpretation
2.1. BASIC DEFINITIONS 53

of the notation na if n is negative. It can be understood either as the


sum of −n copies of −a or as −[(−n)a]. It is easy to verify that they
are the same. Hence the notation na is unambiguous for any integer n.

It is easy to verify the following properties:


• (n + m)a = na + ma for any integers n, m;
• n(a + b) = na + nb for any integer n;
• n(ma) = (nm)a for any integers n, m.

Definition 2.1.5. Let R be a ring. If a, b ∈ R satisfy ab = 0, then


a is called a left zero-divisor of b and b is called a right zero-divisor
of a. If a, b ∈ R satisfy ab = 1, then a is called a left inverse of b and
b is called a right inverse of a. Moreover if ab = ba = 1, then a is
called the (multiplicative) inverse of b, denoted by b−1 . By symmetry,
b is also the inverse of a, i.e., b = a−1 . A nonzero element a is called a
unit if its multiplicative inverse exists.
Let R be a nonzero commutative ring. If the product of any two
nonzero elements in R is not equal to zero, then R is called an integral
domain.
If every nonzero element in a nonzero ring R has an inverse, then
R is called a division ring or skew field. A commutative skew field
is called a field.

Example 2.1.6. • All number fields are fields.

• Mn (R) is not an integral domain if n ≥ 2.

• Z is an integral domain.

• The polynomial ring K[x1 , . . . , xn ] in n variables over a number


field is an integral domain.

• The set of all rational functions over a number field K is a field.

• The zero ring {0} is not an integral domain.

Let A be a nonzero ring such that the product of any two nonzero
elements is nonzero. The the set of all nonzero elements of R form a
monoid under the multiplication. The set of all units form a group,
54 CHAPTER 2. ELEMENTS OF RINGS AND FIELDS

denoted by A∗ . When A is a division ring, then the group A∗ = A\{0}


is called the multiplicative group of A. The multiplicative group of a
field is an abelian group.

Recall the introduction of a complex number a + bi in which i is


an imaginary number satisfying i2 = −1. A new kind of “number”,
calld quaternion, can be defined in a similar manner. A quaternion
is written as a + bi + cj + dk, in which a, b, c, d are numbers and i, j, k
are three distinct elements satisfying

i2 = j 2 = k 2 = −1, ij = k, ji = −k, jk = i, kj = −i, ki = j, ik = −j.

The addition of two quaternions are given by

(a+bi+cj+dk)+(a0 +b0 i+c0 j+d0 k) = (a+a0 )+(b+b0 )i+(c+c0 )j+(d+d0 )k

and the product is given by

(a + bi + cj + dk)(a0 + b0 i + c0 j + d0 k)
= (aa0 − bb0 − cc0 − dd0 ) + (ab0 + ba0 + cd0 − dc0 )i
+(ac0 + ca0 + db0 − bd0 )j + (ad0 + da0 + bc0 − cb0 )k.

It is easy, though tedious, to verify that the set Q of all quaternions


is a ring under addition and multiplication with 1 as the unity. Since
ij 6= ji, it is not a commutative ring.
Since

(a + bi + cj + dk)(a − bi − cj − dk) = a2 + b2 + c2 + d2 ,

Every nonzero element of Q is a unit. That is to say that Q is a division


ring. This is the simplest non-commutative division ring.
The set {a + bi + cj + dk ∈ Q|a, b, c, d ∈ Z} is a non-commutative
ring.
The definition of the direct product of rings is similar to that of
groups.

Exercises
2.2. IDEALS AND QUOTIENT RINGS 55

1. Verify 1)-3) of Proposition 2.1.4.


2. Show that the inverse of a unit in a ring is unique.
3. Let R be a commutative ring and let u be a unit in R. Assume
that a ∈ R and there exists a natural number n such that an = 0.
Show that u + a is a unit.
4. Let A be the set of maps from the set N of all natural numbers
to the complex number field, i.e., the set of all sequences of complex
numbers. For any f, g ∈ A, define

(f + g)(n) = f (n) + g(n), ∀n ∈ N,

and X n
(f ∗ g)(n) = f (d)g .
d
d|n

Show that A is a commutative ring under the addition and the multi-
plication “*”. What is the unity in this ring?
5. Show that the direct product of two fields is not a field.
6. Let A be a ring such that x2 = x for all x ∈ A. Show that 2x = 0
for all x ∈ A.

2.2 Ideals and Quotient Rings

Many mathematical structures contain substructures. For example,


sets have subsets, vector spaces have subspaces, groups have subgroups.
Naturally rings have subrings which is defined in the following obvious
way.

Definition 2.2.1. Let S be a nonempty additive subgroup of a


ring R. If S is closed under multiplication and contains the unity, then
S is a subring of R.

The subring S itself is a ring.


The subrings inherit most properties of its mother ring, such like
commutativity or being an integral domain, etc, but not all. For ex-
ample, a subring of a non-commutative ring can be commutative, a
56 CHAPTER 2. ELEMENTS OF RINGS AND FIELDS

subring of a field is not necessarily a field. If a subring of a field is a


field then this subring is called a subfield.
Let I be an additive subgroup of a ring R. Since R is an abelian
group under addition, I is a normal subgroup of R. Hence the additive
group R/I is well-defined. We want to know whether the multiplication
in R induces a well-defined multiplication in R/I to make it into a ring.
Write two elements in R/I as a + I, b + I, in which a, b ∈ I. A
reasonable multiplication should be (a+I)(b+I) = ab+I. An important
issue arises as in the discussion of quotient groups: does ab + I depend
upon the choices of the representatives of a + I and b + I?
Change the representatives into a + u and b + v, in which u, v can
be any elements in I. Then

(a + u)(b + v) = ab + av + ub + uv.

In order that the multiplication (a + I)(b + I) = ab + I makes sense,


(a + u)(b + v) and ab must be in the same coset of I. This amounts
to saying that (a + u)v + ub ∈ I for any a, b ∈ R, u, v ∈ I, which is
equivalent to av ∈ I, ub ∈ I for any a, b ∈ R, u, v ∈ I. The following
definition is suggested:

Definition 2.2.2. Let I be an additive subgroup of a ring R. If


au ∈ I, ua ∈ I for any a ∈ R, u ∈ I then I is called a (two-sided) ideal
of R.

Proposition 2.2.3. Let I be an ideal of R. Then R/I forms a ring


under the multiplication (a + I)(b + I) = ab + I, called the quotient
ring of R over I.

Proof. We have already seen that the multiplication is well-defined.


It is evident that R/I is an abelian group under addition. It remains
to check the associative law, the distributive law and the existence of
unity.
Associative law:

(a + I)[(b + I)(c + I)] = (a + I)(bc + I) = abc + I.


2.2. IDEALS AND QUOTIENT RINGS 57

[(a + I)(b + I)](c + I) = (ab + I)(c + I) = abc + I.

Distributive law:

(a + I)[(b + I) + (c + I)] = (a + I)(b + c + I)


= a(b + c) + I
= (a + I)(b + I) + (a + I)(c + I).

The verification of

[(b + I) + (c + I)](a + I) = (b + I)(a + I) + (c + I)(a + I)

is similar. Since (1 + I)(a + I) = 1a + I = a + I for every a ∈ R, the


coset 1 + I is the unity of R/I.

Remark 2.2.4. If the condition au ∈ I, ua ∈ I for all a ∈ R, u ∈ I


is replaced by au ∈ I for all a ∈ R, u ∈ I then I is called a left ideal
of R. The definition of right ideal is similar. For commutative rings,
there is no difference among ideal, left ideal and right ideal.
The role played by ideals in ring theory is similar to normal sub-
groups in group theory. But an ideal is not a subring, except for the
whole ring, since a subring contains unity by definition and an ideal is
the whole ring if and only if it contains the unity.

Two elements a, b in a ring R are called congruent with respect to


the ideal I if a − b ∈ I, denoted by

a≡b (mod I).

In particular a ≡ 0 (mod I) means a ∈ I.

Example 2.2.5. • Let n be a natural number greater than one.


Then nZ is an ideal of Z. The quotient ring Zn = Z/nZ is an
integral domain if and only if n is a prime number. In fact, Zn is
a field if n is a prime number. The notation a ≡ b (mod nZ) can
be abbreviated as a ≡ b (mod n), which is a standard notation
in elementary number theory.
58 CHAPTER 2. ELEMENTS OF RINGS AND FIELDS

• Any nonzero ring R is shipped with two ideals for free. They are
{0} and R, called trivial ideals of R. A nonzero ring is called
a simple ring if it is not a division ring and does not have any
nontrivial ideals.

• Let I be the set of all 2 × 2 matrices whose first row is zero. Then
I is a left ideal of M2 (R), but not an ideal.

Proposition 2.2.6. 1) Let {Iλ }λ∈Λ be a collection of ideals of a


T
ring R. Then λ∈Λ Iλ is also an ideal of R.
2) Let I and J be ideals of a ring R. Let I +J = {a+b|a ∈ I, b ∈ J}.
Then I + J is an ideal of R. More generally, for any collection of ideals
{Iλ }λ∈Λ of R, let
X X
Iλ = { aλ |aλ ∈ Iλ , there are only finitely many nonzero terms }.
λ∈Λ λ

P
Then λ∈Λ Iλ is an ideal of R.
3) Let I, J be ideals of a ring R. Let
n
X
IJ = { ai bi |ai ∈ I, bi ∈ J, n < ∞}.
i=1

Then IJ is an ideal of R.

The proof is left as an exercise.

Example 2.2.7. In Z let I = nZ, J = mZ. Then

I ∩ J = [m, n]Z.

I + J = (m, n)Z.

Here [m, n] and (m, n) denote the least common multiple and greatest
common divisor of m and n respectively.

IJ = mnZ.

This suggests the following definition.


2.2. IDEALS AND QUOTIENT RINGS 59

Definition 2.2.8. Let I, J be ideals of a ring R. If I + J = R, then


I, J are coprime.
Let S be a nonempty subset of a ring R. Let hSi be the intersection
of all ideals of R containing S. Since R contains S, the intersection
makes sense. Then hSi is an ideal of R, called the ideal generated by
S. It is not hard to see that

hSi = {a1 u1 b1 + · · · + an un bn |a1 , . . . , an , b1 , . . . , bn ∈ R, u1 , . . . , un ∈ S}.

When R is commutative then

hSi = {a1 u1 + · · · an un |a1 , . . . , an ∈ R, u1 , . . . , un ∈ S}.

The left and right ideals generated by S are defined in a similar


manner.
Let I be an ideal of a ring R. If there is a finite subset S of R such
that I = hSi , then I is called a finitely generated ideal. Moreover,
if I can be generated by a single element, then I is called a principal
ideal.
If an ideal I is generated by a finite number of elements s1 , . . . , sn ,
the I can be denoted by (s1 , . . . , sn ). For example in the ring of poly-
nomials Q[x], the principal ideal generated by x2 + x is denoted by
(x2 + x).
Exercises
1. Prove Proposition 2.2.6
2. Let A be the ring of all real matrices in the form
" #
a b
.
0 c

Find all ideals of A.


3. Let Mn (K) denote the ring of all n × n matrices over a number
field K. Show that it is a simple ring.
4. Let R = Zk where k is a natural number. Show that the following
two statements are equivalent
60 CHAPTER 2. ELEMENTS OF RINGS AND FIELDS

1) There is an element a ∈ R and a natural number n such that


a 6= 0 and an = 0.
2) k is divisible by the square of a prime number.

5. Let A be a commutative ring and let I1 , . . . , In be mutually


coprime ideals of A. Prove the following statements:
1) For any 1 ≤ i ≤ n there exists bi ∈ A such that

bi ≡ 1 (mod Ii )

and
bi ≡ 0 (mod Ij )

for all j 6= i.
2) (The Chinese Remainder Theorem) For any a1 , . . . , an ∈ A, there
exists b such that
b ≡ ai (mod Ii )

for all i.

6. Let I1 , I2 , . . . be a sequence of ideals of R such that I1 ⊆ I2 ⊆ · · · .


Show that ∞
S
i=0 Ii is an ideal of R.

7. Let I be the ideal generated by three elements 12, 48, 30 in Z.


Find a ∈ Z such that I = (a).

8. Let I be the left ideal generated by all elements in the form


ab − ba(a, b ∈ R) in a ring R. Show that I is an ideal of R.

2.3 Homomorphisms of Rings

Definition 2.3.1. Let R1 and R2 be two rings. A map f : R1 → R2


is called a homomorphism if the following conditions are satisfied:
1) f (a + b) = f (a) + f (b) for any a, b ∈ R1 .
2) f (ab) = f (a)f (b) for any a, b ∈ R1 .
3) f (1R1 ) is equal to 1R2 .

The following example shows that the third condition cannot be


derived from the first two conditions.
2.3. HOMOMORPHISMS OF RINGS 61

Example 2.3.2. Let R = Z × Z. The map f : Z → R, a 7→ (a, 0).


satisfies 1) and 2), but not 3).

For a homomorphism f, denote Ker(f ) = {a ∈ R1 |f (a) = 0}. Then


Ker(f ) is called the kernel of f. f is injective if and only if Ker(f ) =
{0}. Denote Im(f ) = {f (a)|a ∈ R1 }. Then Im(f ) is a subring of R2 .
If Im(f ) = R2 , then f is an epimorphism. A bijective homomorphism
is called an isomorphism. If there is an isomorphism from a ring
R1 to another ring R2 then these two rings are isomorphic, denoted
by R1 ∼ = R2 . A homomorphism from a ring R to itself is called an
endomorphism. A bijective endomorphism is an automorphism.

Theorem 2.3.3 (The Fundamental Theorem of Homomorphisms).


Let f : R1 → R2 be a ring homomorphism. Then Ker(f ) is an ideal of
R1 and
R1 /Ker(f ) ∼
= Im(f ).

Proof. Assume that a ∈ R1 , h ∈ Ker(f ). Then

f (ah) = f (a)f (h) = f (a)0 = 0,

f (ha) = f (h)f (0) = 0f (a) = 0.

Hence ah ∈ Ker(f ), ha ∈ Ker(f ). This implies that Ker(f ) is an ideal


of R1 . Denote I = Ker(f ).
It follows from the fundamental theorem of group homomorphisms
that the group homomorphism

φ : R1 /I → Im(f ), a + I 7→ f (a)

is bijective. It remains to verify φ((a + I)(b + I)) = φ(a + I)φ(b + I),


which is obvious from the explicit formula of φ.

Proposition 2.3.4. Let I be an ideal of a ring R. Let f : R → R/I


be the canonical homomorphism. Let Γ be the set of all ideals of R
containing I. Let Γ0 be the set of all ideals of R/I. The map f −1 is a
one-to-one correspondence from Γ0 to Γ.

Proof. Define a map φ from Γ to Γ0 as follows: For any J ∈ Γ let


62 CHAPTER 2. ELEMENTS OF RINGS AND FIELDS

J 0 be the set of all coset of I in the form a + I(a ∈ J). Then J 0 is an


ideal of R/I. Define φ(J) = J 0 .
Since f −1 (J 0 ) = J the composite f −1 ◦ φ is the identify map. It is
easy to verify that φ ◦ f −1 is also the identity map. Hence f −1 : Γ0 → Γ
is bijective.

Proposition 2.3.5. Let S be a subring of a ring R and let I be an


ideal of R. Then S + I = {a + b|a ∈ S, b ∈ I} is a subring of R, I is
an ideal of S + I, S ∩ I is an ideal of S and

S + I/I ∼
= S/S ∩ I.

Proof. Obviously S +I is an additive subgroup of R. Assume that


a1 , a2 ∈ S, b1 , b2 ∈ I. Then (a1 +b1 )(a2 +b2 ) = a1 a2 +(a1 b2 +b1 a2 +b1 b2 ) ∈
S + I. It is obvious that S + I contains the unity. Hence S + I is a
subring of R. It is easy to check that I is an ideal of S + I and S ∩ I
is an ideal of S.
Construct a map f : S → S + I/I, a 7→ a + I. Then f is an
epimorphism with Ker(f ) = S ∩ I. It follows from the fundamental
theorem of homomorphisms that

S + I/I ∼
= S/S ∩ I.

Exercises

1. Let a be a nonzero element of a commutative ring R and let I


be the principal ideal in R[x] generated by ax − 1. Let f : R → R[x]/I
be the homomorphism defined by f (b) = b + I. Show that Ker(f ) =
{b ∈ R| there exists a natural number n such that an b = 0}.

2. Let I, J be two ideals of a ring R. Define the map

f : R → R/I × R/J, a 7→ (a + I, a + J).

Show that
1) f is a ring homomorphism;
2.4. ELEMENTARY PROPERTIES OF FIELDS 63

2) f is injective if and only if I ∩ J = 0;


3) f is surjective if and only if I + J = R.

3. Denote by F [x, y, z] and F [t] the rings of polynomials in three


variables and in one variable respectively over a field F. Let φ : F [x, y, z] →
F [t] be a homomorphism carrying f (x, y, z) to f (t2 , t3 , t4 ). Show that
Ker(φ) is the ideal generated by y 2 − x3 and z − x2 .

2.4 Elementary Properties of Fields

Fields are special commutative rings in which every nonzero element is


a unit. Recall the definition of a number field. A number field K is a
subset of C satisfying the following three conditions

1. a + b, a − b, ab ∈ K for any a, b ∈ K;

2. 1/a ∈ K for any nonzero a ∈ K;

3. K contains at least one nonzero number.

The generalization is straightforward.

Proposition 2.4.1. A subset E of a field F is a subfield of F if


and only if the following conditions are satisfied:

1. a + b, a − b, ab ∈ E for any a, b ∈ E;

2. 1/a ∈ E for any nonzero a ∈ E;

3. E contains at least one nonzero element.

The homomorphism from a field to a ring is relatively simple, as


stated in the following proposition.

Proposition 2.4.2. Let f : F → A be a homomorphism from a


field to a nonzero ring A. Then f is injective.

Proof. Since f (1F ) = 1A 6= 0, Ker(f ) is a proper ideal of F, which


contains only one element 0. Hence f is injective.
64 CHAPTER 2. ELEMENTS OF RINGS AND FIELDS

The arithmetic of elementary schools starts with integers followed


by fractional numbers. We may imitate this process to construct a field
from a given integral domain R.
Immediately one may say that the set of all a/b with a, b ∈ R(b 6= 0)
is the desired field. On second thought, what does the expression a/b
mean? If a and b are natural numbers, a math teacher of elementary
school may interpret a/b graphically or verbally: cut a cake into b
pieces of equal size and then take a pieces. But for an abstract integral
domain R, such explanation does not make sense, since the elements in
R are not quantities any more. The expression a/b is merely an ordered
pair of elements in R. A better notation is (a, b) with a, b ∈ R, b 6= 0.
The set S of all such pairs are not the field we try to construct, because
there is redundancy in this set. Take fractional numbers as example,
(2, 3), (4, 6), . . . represent the same fraction number, although they are
different pairs. This reminds us to give a rule to equalize the elements in
S. In mathematical jargon, we need to establish an equivalence relation
and consider the set of equivalence classes.
More specifically, for any (a1 , b1 ), (a2 , b2 ) ∈ S, define (a1 , b1 ) ∼
(a2 , b2 ) if and only if a1 b2 = a2 b1 . Let us verify that “∼” is an equiva-
lence relation.
1) reflexivity: ab = ba implies (a, b) ∼ (a, b).
2) symmetry: obvious.
3) (a1 , b1 ) ∼ (a2 , b2 ), (a2 , b2 ) ∼ (a3 , b3 ). So b2 (a1 b3 − a3 b1 ) = (a1 b2 −
a2 b1 )b3 − (a3 b2 − a2 b3 )b1 = 0. Hence (a1 , b1 ) ∼ (a3 , b3 ).
Let a/b denote the equivalence class in S represented by (a, b). This
is the correct interpretation of the notation a/b. In our familiar terms,
a1 /b1 = a2 /b2 if and only if a1 b2 = a2 b1 .
Define addition and multiplication between equivalence classes by
the following rules:

a1 a2 a1 b2 + a2 b1
+ = ,
b1 b2 b1 b2
a1 a2 a1 a2
= .
b1 b2 b1 b2
2.4. ELEMENTARY PROPERTIES OF FIELDS 65

It is indispensable to check that these two operations do not depend


upon the choices of representatives. Assume that a1 /b1 = a01 /b01 and
a2 /b2 = a02 /b02 . Then

(a1 b2 + a2 b1 )b01 b02 = a1 b2 b01 b02 + a2 b1 b01 b02

= a01 b1 b2 b02 + a02 b2 b1 b01 = (a01 b02 + a02 b01 )b1 b2 .

Hence the addition does not depend upon the choices of representatives.
The readers can verify the multiplication easily.
The remaining work is routine: check that the set of equivalence
classes is an abelian group under addition with 0/1 as its zero element,
the law of associativity and commutativity for multiplication and the
law of distributivity are satisfied, 1/1 is the unity and every nonzero
element is a unit.
The field F thus obtained is called the field of fractions of R.
Define a map j : R → F, a 7→ a/1. It is easy to check that j is
injective, which means that the integral domain can be treated as a
subring of its field of fractions.

Remark 2.4.3. This method of construction can be summarized as


follows.
In order to construct a new algebraic structure from a known alge-
braic structure, first construct a set which is larger than needed, then
define an equivalence relation in this large set and finally defined a new
algebraic structure in the set of equivalence classes.
This method is very common in various branches of mathematics.

The method we have used to “enlarge” an integral domain to its field


of fractions is more or less what we are familiar with in elementary and
middle schools. Let’s consider the other direction. Can we “shrink” a
ring R into a field? The first natural candidate might be a subring.
Look at a simple example R = Z. It is easy to see that no subring
of Z is a field. Therefore only limited new fields can be produced as
subrings of known rings.
The correct way of thinking is to consider the quotient ring of a
given ring R. Choose an ideal I of R. Then the quotient ring R/I is
66 CHAPTER 2. ELEMENTS OF RINGS AND FIELDS

“smaller” than R. When I satisfies certain conditions R/I can be a


field.

Definition 2.4.4. Let I be a proper ideal of a ring R. If there is


no other ideal between I and R then I is called a maximal ideal of
R.

Proposition 2.4.5. Let I be an ideal of a commutative ring R.


Then R/I is a field if and only if I is a maximal ideal of R.

Proof. ⇒: Assume that R/I is a field. Let J be an ideal of R


containing I but not equal to I. It suffices to show that J = R. Choose
any a ∈ J\I. Then a + I is a nonzero element in R/I. So there is
b + I ∈ R/I such that (a + I)(b + I) = 1 + I, which means that
ab − 1 ∈ I. So 1 = ab + u for some u ∈ I. Hence 1 ∈ J, which implies
J = R.
⇐: Assume that I is a maximal ideal of R. For any a ∈ R\I let J
be the ideal of R generated by I and a. Then J = R by the definition
of maximal ideals. Hence 1 ∈ J, i.e., 1 = ba + u for some b ∈ R and
u ∈ I. This implies (a + I)(b + I) = 1 + I in R/I. Hence R/I is a
field.

It is easy to find all maximal ideals of Z since its every ideal has
the form nZ for some nonnegative integer n. The relation nZ ⊆ mZ
holds if and only if m|n or n = 0. It follows that nZ is a maximal ideal
if and only if n is a prime number.
Let p be a prime number. Denote Fp = Zp = Z/pZ. Then Fp is a
field containing p elements. This a very important field which must be
kept in mind.
Let F be an arbitrary field. Let

f : Z → F, n → n · 1.

Recall that n·1 = 1+1+· · ·+1 n times. The map f is a homomorphism.


Let Ker(f ) = mZ, in which m is a nonnegative integer. The integer m
is called the characteristic of the field F.
There are two cases:
2.4. ELEMENTARY PROPERTIES OF FIELDS 67

1) the characteristic is 0.
In this case, f is a monomorphism. Thus Z can be treated as a
subring of the field F. More precisely, F contains a subring isomorphic
to Z. Since F is a field, all m/n with m, n ∈ Z, n 6= 0 also belongs to
F. This yields the following
Theorem 2.4.6. Every field of characteristic 0 contains a subfield
isomorphic to Q.
It is immediate that this subfield is the smallest subfield of F, which
is called the prime field of F.
2) the characteristic m > 0.
In this case we want to show that m is a prime number. Suppose
that m = ab in which a and b are natural numbers less than m. Then
f (a)f (b) = f (m) = 0. But both a and b are not in Ker(f ) = mZ.
Hence f (a) 6= 0, f (b) 6= 0, which contradicts the assumption that F is
a field. Hence m is a prime number, which is usually denoted by p.
It is evident that Im(f ) ∼= Fp .
Theorem 2.4.7. Every field of characteristic p > 0 contains a sub-
field isomorphic to Fp .
It is natural to call Fp the prime field of F in this case.
Lemma 2.4.8. Let F be a field of characteristic p > 0 and let a ∈
F. Let n, m be two integers such that p|(m − n). Then ma = na. In
particular, if p|n then na = 0 for any a ∈ F.
Proof. Since p|(m − n), m − n = pr for some r ∈ Z.
First assume that a = 1. Then pr · 1 = 0 for any integer r. Hence
ma − na = m · 1 − n · 1 = pr · 1 = 0.
For an arbitrary a ∈ F, it follows from ma−na = m(1·a)−n(1·a) =
(m · 1 − n · 1)a that ma − na = 0.

The following formula is useful.


Proposition 2.4.9. Let a and b be two elements in a field F of
characteristic p > 0. Let r be a natural number. Then
r r r
(a + b)p = ap + bp .
68 CHAPTER 2. ELEMENTS OF RINGS AND FIELDS

Proof. Apply induction to r.


For any 1 < i < p, the natural number pi is divisible by p. Hence

p i p−i

i
a b = 0 holds for all 0 < i < p. It follows from
p  
p
X p
(a + b) = ai bp−i
i=0
i

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

The remaining step of the proof is simple and left to the readers.

Fields are divided into two big classes: those of characteristic zero
and nonzero. These two classes have quite different properties. Among
the fields of prime characteristics, those of characteristic 2 are very
special, mainly because a quadratic polynomial x2 + bx + c cannot be
changed into the form y 2 + d by a change of variable y = x + e.
In terms of the number of elements, fields are divided into finite
fields and infinite fields. We have already met finite fields Fp , whose
characteristic is a prime number p. Later we will learn that there are
other finite fields. A field of prime characteristic is not necessarily a
finite field. In field theory, the classification of finite fields is totally
solved. But the classification of infinite fields is far from clear.

To conclude this chapter we prove an interesting theorem which


plays an important role in Galois theory.

Theorem 2.4.10 (E. Artin). Let χ1 , . . . , χn be distinct homomor-


phisms from a group G to the multiplicative group F ∗ of a field F. Let
a1 , . . . , an be a set of elements in F not all equal to zero. Then there
exists g ∈ G such that

a1 χ1 (g) + a2 χ2 (g) + · · · + an χn (g) 6= 0.

Proof. Apply induction to n. The statement is obviously true for


n = 1.
Assume that n > 1. If one among a1 , . . . , an is equal to zero, then
the statement holds by induction hypothesis. Hence we may assume
2.4. ELEMENTARY PROPERTIES OF FIELDS 69

that a1 , . . . , an ∈ F ∗ .
Since χ1 , . . . , χn are distinct, there is h ∈ G such that χ1 (h) 6=
χ2 (h). Hence a2 (χ2 (h) − χ1 (h)) 6= 0. Since the n − 1 elements

a2 (χ2 (h) − χ1 (h)), a3 (χ3 (h) − χ1 (h)), . . . , an (χn (h) − χ1 (h))

in F are not all equal to zero, by induction hypothesis there exists


g ∈ G such that

a2 (χ2 (h)−χ1 (h))χ2 (g)+a3 (χ3 (h)−χ1 (h))χ3 (g)+· · ·+an (χn (h)−χ1 (h))χ(g) 6= 0.

Let
u = a1 χ1 (hg) + a2 χ2 (hg) + · · · + an χn (hg),

v = a1 χ1 (g) + a2 χ2 (g) + · · · + an χn (g).

Then

u − χ1 (h)v
= a2 (χ2 (h) − χ1 (h))χ2 (g) + a3 (χ3 (h) − χ1 (h))χ3 (g) + · · · + an (χn (h) − χ1 (h))χ(g)
6= 0.

So u, v cannot be zero at the same time. This shows that the theo-
rem is true for n too.

Exercises

1. Give an example to show that the condition of commutative ring


in Proposition 2.4.5 is indispensable.

2. Let R denote the field of real numbers. Find all maximal ideals
of the ring R × R.

3. Let C denote the field of complex numbers. Show that the ideal
in C[x, y] generated by x + y 2 and y + x2 + 2xy 2 + y 4 is a maximal ideal.

4. Let k be a field with infinitely many elements and let n be a


natural number greater than 1. Assume that (a + b)n = an + bn holds
70 CHAPTER 2. ELEMENTS OF RINGS AND FIELDS

for all a, b ∈ k. Show that the characteristic of k is a prime number


and n is the power of a prime number.
5. Let S be the ring consisting of all real matrices in the form
" #
a b
.
−b a

1) Write out an isomorphism from A to the complex number field


C without proof;
2) Let M = M2 (R) denote the ring consisting of all 2 × 2 real
matrices. Show that there exists a subring N of M containing
" #
0 3
A=
−4 1
such that N ∼ = S;
3) Show that there exists X ∈ M2 (R) such that X 4 + 13X = A.
6. Let A be a commutative ring. An ideal P of A is called a prime
ideal if ab ∈
/ P for any a, b ∈ A\P.
1) Show that an ideal of A is a prime ideal if and only if A/P is an
integral domain;
2)Show that every maximal ideal of A is a prime ideal.
Chapter 3

Polynomials and Rational


Functions

Polynomials over an arbitrary field are defined similar to those over


number fields (real polynomials, complex polynomials, · · · ). In fact
the coefficients of polynomials can be in an arbitrary ring. To avoid
zero-divisors we restrict the discussion to polynomials over an integral
domain.

3.1 Polynomials in One Variable

Definition 3.1.1. Let A be an integral domain. An expression in


the form
an xn + an−1 xn−1 + · · · + a1 x + a0

with
an , an−1 , . . . , a1 , a0 ∈ A.

is called a polynomial over A. For any i with ai 6= 0, ai xi is called a


term of that polynomial and ai is called the coefficient of that term.
The terms of a polynomial can be exchanged at will. When an 6= 0 the
number n is called the degree of that polynomial. The term an xn is
the initial term. A polynomial whose coefficient of the initial term is
equal to one is called a monic polynomial. The symbol x is called
the indeterminate or variable. The term a0 is called the constant

71
72 CHAPTER 3. POLYNOMIALS AND RATIONAL FUNCTIONS

term. The zero polynomial 0 is a special polynomial, whose degree is


equal to −∞ by convention.
The set of all polynomials over A with indeterminate x is denoted
by A[x].

The degree of f (x) ∈ A[x] is denoted by deg(f (x)) or deg(f ). The


degree is equal to zero if and only if f (x) ∈ A\{0}. In this case we may
say that f (x) is a nonzero constant.
Two polynomials are equal if and only if they have the same degree
and their coefficients of the degree i terms are equal for each i.
The addition, subtraction and multiplication between polynomials
are defined in usual way. They satisfy the laws of commutativity, as-
sociativity and distributivity. Hence A[x] is a commutative ring.
The composite f (g(x)) of f (x), g(x) ∈ A[x] is defined as usual. The
following properties are easily verified:
• deg(f g) = deg(f ) + deg(g). Thus the product of two nonzero
polynomials is not equal to zero. Hence A[x] is an integral domain.
• Let c be a nonzero element in A. Then deg(cf (x)) = deg(f ) for
any f (x) ∈ A[x].
• deg(f ± g) ≤ max(deg(f ), deg(g)), and the equality holds when
deg(f ) 6= deg(g).

3.2 Division Algorithm

Theorem 3.2.1. Let A be an integral domain and let f (x), g(x) ∈


A[x]. If g(x) 6= 0 and its leading coefficient is an invertible element in
A then there exist a unique pair of polynomials q(x), r(x) ∈ A[x] such
that
1) f (x) = g(x)q(x) + r(x);
2) deg(r) < deg(g).
The polynomials q(x) and r(x) are called the quotient and re-
mainder of f (x) by g(x) respectively.

Proof. This can be proved by induction on the degree of f (x).


The details are left as an exercise.
3.2. DIVISION ALGORITHM 73

Definition 3.2.2. Let A be an integral domain and let f (x), g(x) ∈


A[x]. If f (x) 6= 0 and there exists h(x) ∈ A[x] such that g(x) =
f (x)h(x), then f (x) divides g(x) or g(x) is divisible by f (x), denoted
by f (x)|g(x) or simply f |g.

Some basic facts are listed as follows:


• Every invertible element of A divides any polynomial;
• Every nonzero polynomial divides the zero polynomial 0;
• 0 does not divide any polynomial;
• If g(x)|f (x), g(x)|h(x), then g(x)|a(x)f (x) + b(x)h(x) for any
a(x), b(x) ∈ A[x];
• If g(x)|f (x) and f (x) 6= 0, then deg(g) ≤ deg(f ).
• g(x)|f (x) implies g(h(x))|f (h(x)) for any h(x) ∈ A[x] such that
g(h(x)) 6= 0.

Lemma 3.2.3. Assume that f (x), g(x) ∈ A[x], g(x) 6= 0, the leading
coefficient of g(x) is an invertible element in A. Then g(x)|f (x) if and
only if the remainder of f (x) by g(x) is zero.

Proof. ⇒: there is some b(x) ∈ A[x] such that f (x) = g(x)b(x). So


f (x) = g(x)b(x) + 0. Since deg(0) < deg(g), the remainder of f (x) by
g(x) is equal to zero.
⇐: the remainder of f (x) by g(x) being zero means that f (x) =
g(x)q(x), i.e., g(x)|f (x). 2

Lemma 3.2.4. Let f (x), g(x) be nonzero polynomials. If f (x)|g(x)


and g(x)|f (x), then f (x) = c · g(x), where c is an invertible element in
A.

Proof. Assume that f (x) = g(x)a(x), g(x) = f (x)b(x). Then f (x) =


f (x)b(x)a(x), so a(x)b(x) = 1. Hence both a(x) and b(x) are invertible
elements in A. 2

The following result is the central theorem in the theory of polyno-


mials in one variable.

Theorem 3.2.5. Every ideal in the polynomial ring F [x] over a


field F is a principal ideal.
74 CHAPTER 3. POLYNOMIALS AND RATIONAL FUNCTIONS

Proof. Let I be an ideal of F [x]. If I = {0}, then I is evidently a


principal ideal. It remains to consider that case I 6= {0}.
Among all nonzero elements in I choose an element g such that
deg(g) reaches the minimum degree. We show that I is generated by
g. For every f ∈ I there are q, r ∈ F [x] such that f = qg + r and
deg(r) < deg(g) by Theorem 3.2.1
Since r = f − qg ∈ I, r must be zero by the choice of g. This means
that f = qg, which implies that f is in the ideal generated by g. 2

We have mentioned that the rings F [x] and Z share many common
properties. The reason is that both of them have division algorithm
and every ideal is a principal ideal. Another interesting ring with this
property is the ring of Gauss integers (cf. Exercise 4). An integral
domain in which every ideal is principal is called a principal ideal
domain, abbreviated as PID.

Definition 3.2.6. Let F be a field and let f (x), g(x), h(x) ∈ F [x].
If h(x)|f (x), h(x)|g(x), then h(x) is called a common divisor of f (x)
g(x). A common divisor d(x) of f (x) and g(x) that divides every com-
mon divisor of f (x), g(x) is called a greatest common divisor of
f (x) and g(x).

Example 3.2.7. There is no greatest common divisor when f (x) =


g(x) = 0.

Example 3.2.8. If f (x) 6= 0, then f (x) is greatest common divisor


of f (x) and 0.

Lemma 3.2.9. If d1 (x), d2 (x) are greatest common divisors of f (x), g(x)
then d1 (x) = c · d2 (x) for some nonzero element c ∈ F.

Proof. The definition of greatest common divisor implies d1 (x)|d2 (x), d2 (x)|d1 (x).
Then apply Lemma 3.2.4. 2

This lemma shows that the greatest common divisor is unique up


to a constant factor. If the leading coefficient is required to be 1, then
it is unique. The monic greatest common divisor of f (x) and g(x) is
denoted by gcd(f (x), g(x)) or gcd(f, g).
3.2. DIVISION ALGORITHM 75

Theorem 3.2.10. If one of f (x), g(x) ∈ F [x] is nonzero then gcd(f (x), g(x))
exists and there are a(x), b(x) ∈ F [x] such that

gcd(f (x), g(x)) = a(x)f (x) + b(x)g(x).

Proof. Let I be the ideal of F [x] generated by f (x) and g(x).


Theorem 3.2.5 implies that I is a principal ideal. Let d(x) be a monic
generator of I. Then there are a(x), b(x) ∈ F [x] such that d(x) =
a(x)f (x) + b(x)g(x).
It remains to show that d(x) = gcd(f (x), g(x)). The definition
of principal ideal implies that d(x)|f (x), d(x)|g(x). Hence d(x) is a
common divisor of f (x) and g(x). Assume that h(x)|f (x), h(x)|g(x),
i.e., there are u(x), v(x) ∈ F [x] such that f (x) = u(x)h(x), g(x) =
v(x)h(x). Then

d(x) = [a(x)u(x) + b(x)v(x)]h(x),

which implies that h(x)|d(x). 2


This proof is not constructive, since if does not tell us how to find
the greatest common divisor. An effective way to compute it is the
Euclidean algorithm, which is based on the division algorithm.
Definition 3.2.11. If gcd(f (x), g(x)) = 1 then f (x) and g(x) are
coprime or relatively prime.
Corollary 3.2.12. Two polynomials f (x) and g(x) over a field
F are coprime if and only if there are u(x), v(x) ∈ F [x] such that
f (x)u(x) + g(x)v(x) = 1.

Exercises
1. Give a detailed proof of Theorem 3.2.1.
2. Let F = F3 , f (x) = 2x4 + 2x + 1, g(x) = x2 − 2x + 2. Find the
quotient and remainder of f (x) by g(x).
3. Prove Lemma 3.2.12
4. The integral domain Z[i] = {m + ni|m, n ∈ Z} is called the ring
of Gauss integers.
76 CHAPTER 3. POLYNOMIALS AND RATIONAL FUNCTIONS

1) For every m + ni ∈ Z[i] let N (m + ni) = m2 + n2 . Show that for


any a, b ∈ Z[i], a 6= 0, there are q, r ∈ Z[i] such that b = qa + r and
N (r) < N (a).

2) Show that Z[i] is a principal ideal domain.

3) Find the greatest common divisor of 11 + 7i and 18 + i.

5. Show that Z[x] is not a principal ideal domain.

6. Show that the quotient ring Q[x]/(x2 −4) is isomorphic to Q×Q.

3.3 Polynomials in Several Variables

Let A be an integral domain and let x1 , x2 , . . . , xn be variables (also


known as indeterminates). An expression in the form cxe11 xe22 · · · xenn
with a nonzero c in A is called a monomial. Here e1 , . . . , en are non-
negative integers.
For fixed variables x1 , x2 , . . . , xn a finite sum of monomials is called
a polynomial in these variables. Every nonzero monomial appearing
in a polynomial is called a term of this polynomial. Note the 0 is also
considered to be a polynomial. The set of polynomials in x1 , . . . , xn is
denoted by A[x1 , . . . , xn ]. The addition,substraction and multiplication
of polynomials are performed in the usual way, keeping in mind that
the arithmetic on the coefficients adopts that in the integral domain
A. It is evident that A[x1 , . . . , xn ] in a commutative ring.
The polynomial ring A[x1 , . . . , xn ] can be regarded as B[xn ] with
B = A[x1 , . . . , xn−1 ]. Thus it is easy to prove that A[x1 , . . . , xn ] is an
integral domain by induction on n.
The degree of a monomial cxe11 xe22 · · · xenn is defined to be e1 +· · ·+en .
The degree of a nonzero polynomial is defined to be the maximum
degree of all its terms. The degree of the zero polynomial 0 is equal to
−∞ by convention. The degree of f (x1 , . . . , xn ) is denoted by deg(f ).
The equality
deg(f g) = deg(f ) + deg(g)

holds for any f (x1 , . . . , xn ), g(x1 , . . . , xn ) ∈ A[x1 , . . . , xn ].


3.3. POLYNOMIALS IN SEVERAL VARIABLES 77

Assume that f (x1 , . . . , xn ), g(x1 , . . . , xn ) ∈ A[x1 , . . . , xn ] with g(x1 , . . . , xn ) 6=


0. If there is some q(x1 , . . . , xn ) ∈ A[x1 , . . . , xn ] such that

f (x1 , . . . , xn ) = g(x1 , . . . , xn )q(x1 , . . . , xn ),

then we say that g divides f, or g is a factor of f, denoted by g|f.


When n > 1 there is no division algorithm in F [x1 , . . . , xn ], even
when F is a field. This is a significant difference between the polynomial
rings in one variable and in several variables.
Let F be a field. The field of fractions of F [x1 , . . . , xn ] is denoted
by F (x1 , . . . , xn ), called the field of rational functions in variables
x1 , . . . , xn . Be careful about the parenthesis and bracket in the nota-
tion. Every element in F (x1 , . . . , xn ) is written as the quotient of two
polynomials.
We have mentioned before that a field of prime characteristic p is not
necessarily a finite field. Here is an example: the field Fp (x1 , . . . , xn ) is
an infinite field of characteristic p.
A polynomial f (x1 , . . . , xn ) ∈ F [x1 , . . . , xn ] is called a symmetric
polynomial if one of the following equivalent conditions is satisfied:
1) f (x1 , . . . , xi , . . . , xj , . . . , xn ) = f (x1 , . . . , xj , . . . , xi , . . . , xn ) for any
1 ≤ i < j ≤ n;
2) f (x1 , . . . , xn ) = f (xσ(1) , . . . , xσ(n) ) for any permutation σ.
Recall the elementary symmetric polynomials:
n
X
σ1 = x1 + · · · + xn = xi ,
i=1
X
σ2 = xi xj ,
1≤i≤j≤n

···

σn = x1 · · · xn .

The following theorem is the most important theorem concerning


symmetric polynomials.

Theorem 3.3.1. For any symmetric polynomials f (x1 , . . . , xn ) ∈


F [x1 , . . . , xn ] there is a unique g(y1 , . . . , yn ) ∈ F [y1 , . . . , yn ] such that
78 CHAPTER 3. POLYNOMIALS AND RATIONAL FUNCTIONS

f (x1 , . . . , xn ) = g(σ1 , . . . , σn ).

The theorem will be used in the last chapter of this book.

3.4 Factorization

The fundamental theorem of arithmetic says that every natural number


greater than one can be factored uniquely (up to the order of the fac-
tors) into a product of prime numbers. For polynomials in one variable,
the irreducible polynomials play the same role as the prime numbers in
the ring of integers and the unique factorization theorem holds in the
ring of polynomials in one variable. One may ask whether the unique
factorization theorem holds for any integral domain. The answer is
negative.
To investigate this property, we need to understand the exact mean-
ing of “unique factorization”.

Definition 3.4.1. Let A be an integral domain. For any a, b ∈ A


with b 6= 0, if a = bc for some c ∈ A, then b is said to divide a (or b is
a factor of a), denoted by b|a
If a|b, b|a for two nonzero elements a and b in A, the a is said to be
associated with b.

The divisibility is the immediate generalization of the divisibility in


the ring of integers and the ring of polynomials.

Example 3.4.2. A natural number n is associated to −n in Z.

Lemma 3.4.3. 1) a is associated to b if and only if a = ub for some


unit u in A.
2) Being associated is an equivalent relation.

Proof. 1) ⇐: Let u0 = u−1 . Then b = u0 a and hence a|b, b|a.


⇒: It follows from a|b, b|a that a = ub, b = u0 a for some u, u0 ∈ A.
Thus a = uu0 a. Since a 6= 0 and A is an integral domain, uu0 = 1,
which implies that u and u0 are units.
2) is a direct consequence of 1).
3.4. FACTORIZATION 79

Definition 3.4.4. A nonzero element p in an integral domain A is


called an irreducible element if
1) it is not a unit and
2) p = ab implies that either a or b is a unit.

In other words, any factor of an irreducible element p is either a


unit or associated with p.

Definition 3.4.5. An integral domain A is a unique factorization


domain (abbreviated as UFD) if the following two conditions are sat-
isfied:
1) Every nonzero element a of A can be expressed as

a = cp1 · · · pr ,

in which c is a unit and p1 , . . . , pr are irreducible elements.


2) If cp1 · · · ps = dq1 · · · qt , with units c, d and irreducible elements
p1 , . . . , ps , q1 , . . . , qt , then s = t and pi is associated with qi for 1 ≤ i ≤ r
after a suitable permutation of q1 , . . . , qt .

It is easy to verify that if p is an irreducible element in a unique


factorization domain A such that p|(a1 · · · an ) then p|aj for some 1 ≤
j ≤ n.
Now we are going to show that the polynomial ring in one variable
over an arbitrary field is a UFD.
Let F be a field. According to definition 3.4.4 a polynomial p(x) ∈
F [x] is irreducible if and only if deg(p) > 0 and p(x) is not a product of
two polynomials of degrees less than deg(p). A polynomial of positive
degree which is not irreducible is called a reducible polynomial.

Lemma 3.4.6. Let f (x), g(x) ∈ F [x]. If f (x) is irreducible and f (x)
does not divide g(x), then gcd(f (x), g(x)) = 1.

Proof. Let d(x) = gcd(f (x), g(x)). Then f (x) = d(x)h(x) for
some h(x) ∈ F [x]. Since f (x) is irreducible, either deg(d) = 0 or
deg(h) = 0. Suppose that deg(h) = 0. Then h(x) is a nonzero ele-
ment in F. It follows from d(x)|g(x) that f (x)|g(x), contradicting the
assumption that f (x) does not divide g(x). Hence deg(d) = 0.
80 CHAPTER 3. POLYNOMIALS AND RATIONAL FUNCTIONS

Theorem 3.4.7. Assume that an irreducible polynomial p(x) di-


vides f (x)g(x). Then p(x)|f (x) or p(x)|g(x).

Proof. Assume that p(x) does not divide f (x). Then Lemma 3.4.6
implies that gcd(p(x), f (x)) = 1. So a(x)p(x) + b(x)f (x) = 1 for some
a(x), b(x) ∈ F [x] by Theorem 3.2.10. Thus

a(x)p(x)g(x) + b(x)f (x)g(x) = g(x).

It follows that p(x)|g(x).

Corollary 3.4.8. Assume that an irreducible polynomial p(x) di-


vides f1 (x) · · · fn (x). Then p(x)|fi (x) for at least one i.

Proof. Apply induction on n.

With these preparations we can prove that F [x] is a UFD.

Theorem 3.4.9. Let f (x) be a polynomial of positive degree over a


field F. Then
f (x) = cp1 (x) · · · pn (x),

in which c is a nonzero element in F, p1 (x), . . . , pn (x) are irreducible


monic polynomials. This equality is called an irreducible decompo-
sition of f (x).
The element c and the sequence p1 (x), . . . , pn (x) of irreducible fac-
tors are uniquely determined by f (x) up to a permutation.

Proof. The existence of the factorization is proved by induction


on deg(f ). When deg(f ) = 1, f (x) is irreducible. Let c be the leading
coefficient of f and let f1 = f /c. Then f (x) = cf1 (x) is the required
irreducible decomposition.
Assume that n > 1 and that every polynomial of degree less than
n has an irreducible decomposition. For a polynomial f of degree n, if
f (x) is irreducible then it is obvious that f has an irreducible decompo-
sition, otherwise f (x) = f1 (x)f2 (x) with deg(f1 ) < deg(f ), deg(f2 ) <
deg(f ). By induction hypothesis we have

f1 (x) = cp1 (x) · · · pr (x), (3.1)


3.4. FACTORIZATION 81

f2 (x) = c0 pr+1 (x) · · · pn (x), (3.2)

in which c, c0 are nonzero elements of F, p1 (x), . . . , pn (x) are irre-


ducible monic polynomials. After plugging (3.2) and (3.1) into f (x) =
f1 (x)f2 (x) an irreducible decomposition is obtained.
The uniqueness is proved by induction on the number n of irre-
ducible factors of f. When n = 1 the factorization of f is obviously
unique. Assume that n > 1. Let

f (x) = c0 q1 (x) . . . qm (x)

be another irreducible decomposition. Since all pi (x), qj (x) are monic


polynomials, both c and c0 are equal to the leading coefficient of f (x),
so c = c0 .
Since p1 (x)|f (x), Corollary 3.4.8 implies that p1 (x) divides some
qj (x), say, p1 (x)|q1 (x). Since both p1 (x) and q1 (x) are irreducible monic
polynomials, so p1 (x) = q1 (x) and

p2 (x) · · · pn (x) = q2 (x) · · · qm (x).

The induction hypothesis implies that m = n and the sequences p2 (x), . . . , pn (x)
and q2 (x), . . . , qm (x) differ by a permutation.

This theorem means that the polynomial ring in one variable over
an arbitrary field is a UFD.

Let
f (x) = cp1 (x) · · · pn (x)

be an irreducible decomposition of f (x). The factors are not necessarily


distinct. By collecting the same factors it can be written as

f (x) = cp1 (x)e1 · · · pr (x)er , (3.3)

in which p1 (x), . . . , pr (x) are distinct irreducible monic polynomials and


e1 , . . . , er are natural numbers. Such an expression is referred to as the
standard irreducible decomposition of f (x). The number ei is called
82 CHAPTER 3. POLYNOMIALS AND RATIONAL FUNCTIONS

the multiplicity of the factor pi (x) in f (x). A factor with multiplicity


greater than one is called a multiple factor.
Let f (x) = a0 xn + a1 xn−1 + · · · + an−1 x + an ∈ F [x]. Define the
formal derivative of f (x) = a0 xn + an−1 xn−1 + · · · + an−1 x + an as

f 0 (x) = na0 xn−1 + (n − 1)a1 xn−2 + · · · + an−1 .

It has no geometric or physical interpretation of the derivative as in


calculus.
The well-known formulae

[f (x) + g(x)]0 = f 0 (x) + g 0 (x)

and
(f (x)g(x))0 = f 0 (x)g(x) + f (x)g 0 (x)

still hold.
Over a field of characteristic zero, the degree of f 0 (x) is is equal
to deg(f ) − 1 as long as deg(f ) > 0. However this is not true for
polynomials over a field of positive characteristic. Let p > 0 be the
characteristic of the field over which the polynomials are defined. Let
m be a natural number divisible by p. Then the derivative of xm is
zero!
The following result is well-known.

Proposition 3.4.10. Let f (x) = cp1 (x)e1 · · · pr (x)er be the standard


irreducible decomposition of a polynomial f (x) with positive degree over
a number field. Then
r
Y
0
gcd(f (x), f (x)) = pi (x)ei −1 .
i=1

This result can be generalized to the fields of characteristic zero but


not to those of positive characteristic. However, the following result
holds for arbitrary fields.

Proposition 3.4.11. If gcd(f (x), f 0 (x)) = 1, then f (x) has no


multiple factors.
3.4. FACTORIZATION 83

Proof. Assume that f (x) = g(x)r h(x), in which deg(g) > 0,


h(x) 6= 0, r > 1. Then

f 0 (x) = rg(x)r−1 g 0 (x)h(x) + g(x)r h0 (x)


= g(x)r−1 [rg 0 (x)h(x) + g(x)h0 (x)].

Hence g(x)|gcd(f (x), f 0 (x)).

So far, we have learned that the ring of integers Z and the ring
of polynomials in one variable over a field are UFD’s. We wish to
construct new UFD’s from known UFD’s.
In the remaining of this section A is always assumed to be a UFD
with F as its field of fractions. Then A[x] is a subring of F [x].
Definition 3.4.12. A nonzero polynomial f (x) = an xn +an−1 xn−1 +
· · · + a1 x + a0 ∈ A[x] is a primitive polynomial if the only nonzero
elements in A dividing all coefficients a0 , a1 , . . . , an are units.
The following theorem is also known as Gauss’s Lemma.
Theorem 3.4.13 ( Gauss ). The product of two primitive polyno-
mials is still primitive.
Proof. Let

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

and
g(x) = bm xm + bm−1 xm−1 + · · · + b1 x + b0

be primitive polynomials. Expand f (x)g(x) as

f (x)g(x) = cn+m xn+m + cn+m−1 xn+m−1 + · · · + c1 x + c0 .

Then X
ck = ai b j
i+j=k

for any k.
Suppose that f (x)g(x) is not primitive. Then there is an irreducible
element p in A dividing every ci . Let r (s resp.) be the largest index
84 CHAPTER 3. POLYNOMIALS AND RATIONAL FUNCTIONS

such that ar (bs resp.) is not divisible by p. Then


X X
cr+s = ar bs + ar−i bs+i + ar+i bs−i
i>0 i>0

is not divisible by p. This leads to a contraction.

Lemma 3.4.14. Let f (x), g(x) be primitive polynomials. Assume


that f (x) = cg(x) for some c ∈ F. Then c is a unit in A. In other
words, two primitive polynomials f (x), g(x) ∈ A[x] are associated in
F [x] if and only if they are associated in A[x].

Proof. Write c as c = a/b where a, b ∈ A satisfy gcd(a, b) = 1.


Then
bf (x) = ag(x). (3.4)

Let ci and di denote the coefficients of the terms of degree i in f (x)


and g(x) respectively. Then bci = adi holds for every i by comparing
the terms of degree i in both sides of (3.4). Suppose that b is not a
unit. Since b is coprime with a, the equality bci = adi implies that b|di ,
so b divides every coefficient of g(x). This contradicts the assumption
that g(x) is primitive. Hence b is a unit in A. The element a is also a
unit in A for the same reason.

Proposition 3.4.15. If a primitive polynomial in A[x] is reducible


in F [x] then it can be factored into a product of two polynomials in
A[x] of lower degrees.

Proof. Let f be a primitive polynomial such that f (x) = g(x)h(x)


for some g(x), h(x) ∈ F [x] with deg(g) < deg(f ), deg(h) < deg(f ).
Choose r, s ∈ F such that rg(x), sh(x) are primitive. This is possible
since A is a UFD.
By Theorem 3.4.13 the right hand side of

rsf (x) = [rg(x)][sh(x)]

is primitive. Then Lemma 3.4.14 implies that rs is a unit in A.

The units and irreducible elements in A[x] are characterized by the


following lemma.
3.4. FACTORIZATION 85

Lemma 3.4.16. 1) Every unit in A[x] is a unit in A.


2) A polynomial f (x) ∈ A[x] of positive degree is irreducible in A[x]
if and only if it is primitive and it is irreducible in F [x].

Proof. 1) Let g(x) be a unit in A[x]. Then g(x)h(x) = 1 for some


h(x) ∈ A[x]. Hence deg(g) = deg(h) = 0, which implies that both g(x)
and h(x) are units in A.
2) First assume that f (x) is an irreducible element in A[x]. Let a
be the greatest common divisor of all coefficients of f (x). Then f (x) =
af1 (x), with f1 (x) being primitive. f1 (x) is not a unit since deg(f1 ) =
deg(f ) > 0. Thus a is a unit in A since f (x) is irreducible in A[x]. This
implies that f (x) is primitive.
By Proposition 3.4.15 if f (x) is reducible in F [x] then it would
be decomposed into the product of two polynomials in A[x] of lower
degrees. This would contradict the assumption that f (x) is irreducible
in A[x]. Hence f (x) is irreducible in F [x].
Conversely assume that a primitive polynomial f (x) ∈ A[x] is irre-
ducible in F [x]. Assume that f (x) = g(x)h(x) with g(x), h(x) ∈ A[x].
Since f (x) is irreducible in F [x], either deg(g) = 0 or deg(h) = 0.
Assume that deg(g) = 0 without loss of generality. This amounts to
saying that g(x) = c is a nonzero element in A. Since f (x) is primitive,
c is a unit. This shows that f (x) is an irreducible element in A[x].

With all these preparation, we can prove the following theorem:

Theorem 3.4.17. A[x] is a UFD if A is a UFD.

Proof. We need to verify for A[x] the two conditions in the defi-
nition of UFD are satisfied.
1) Let f (x) be a nonzero element in A[x]. We show that f (x) can
be decomposed into a product of a unit and some irreducible elements
in A[x] by induction on deg(f ).
If deg(f ) = 0, then f ∈ A and f is decomposed into a product of a
unit and some irreducible elements, since A is a UFD by assumption.
Assume that deg(f ) > 0. Then f = cf1 with c ∈ A and f1 prim-
itive. Let c = dc1 c2 · · · cr be the irreducible decomposition of c in A
where d is a unit and c1 , c2 , . . . , cr are irreducible. If f1 is irreducible
86 CHAPTER 3. POLYNOMIALS AND RATIONAL FUNCTIONS

then f = dc1 c2 · · · cr f1 is the required decomposition already, other-


wise Proposition 3.4.15 implies that f1 = gh, in which g, h ∈ A[x],
deg(g) < deg(f ), deg(h) < deg(f ). By induction hypothesis both g
and h have irreducible decompositions. Hence f has an irreducible
decomposition.
2) Assume that cp1 · · · ps = dq1 · · · qt in which c, d are units and
p1 , . . . , ps , q1 , . . . , qt are irreducible elements in A[x]. Assume that

p1 , . . . , pu , q1 , . . . , qv ∈ A

and
pu+1 , . . . , ps , qv+1 , . . . , qt ∈
/ A.

By Lemma 3.4.16 pu+1 , . . . , ps , qv+1 , . . . , qt are primitive and irreducible


in F [x]. By Theorem 3.4.13 the products pu+1 · · · ps and qv+1 · · · qt are
also primitive polynomials. Hence

p1 · · · pu = λq1 · · · qv (3.5)

and

pu+1 · · · ps = µqv+1 · · · qt , (3.6)

in which λ, µ are units. Since A is a UFD, u = v and after a suitable


permutation of indices pi is associated with qi for 1 ≤ i ≤ u by (3.5).
It follows from (3.6) and Theorem 3.4.9 that s = t and pi is associ-
ated with qi in F [x] after a suitable permutation for 1 < i ≤ s. Lemma
3.4.14 implies that pi and qi are associated in A[x]. The proof of the
uniqueness is concluded.

Corollary 3.4.18. If A is a UFD, then A[x1 , x2 , . . . , xn ] is a UFD.


Proof. Apply the theorem n times.

Corollary 3.4.19. 1) The polynomial ring F [x1 , . . . , xn ] is a UFD


for any field F.
2) Z[x1 , . . . , xn ] is a UFD.

Exercises
3.5. * POLYNOMIAL FUNCTIONS 87

1. Let F3 be the field with three elements. Decompose the polyno-


mial x9 − x over F3 into a product of irreducible polynomials.

2. Let F be a field with four elements. Let α be a nonzero element


of F. Show that x3 + αx + 1 is an irreducible polynomial over F.

3. Let A be the set of all complex numbers in the form m + n 5i
with m, n ∈ Z.
1) Show that A is a subring of the complex number field.
√ √
2) Show that 3, 2 + 5i, 2 − 5i are irreducible elements in A.

3) Show that 3 and 2 + 5i are not associated.
4) Show that A is not a UFD.

3.5 * Polynomial Functions

Definition 3.5.1. Let F be a field and

f (x) = an xn + an−1 xn−1 + · · · + a1 x + a0 ∈ F [x],

b ∈ F. Then

f (b) = an bn + an−1 bn−1 + · · · + a1 b + a0 ∈ F

is called the value of f (x) at b. If f (b) = 0, then b is called a zero of


f (x).

For any field F, a function on F with values in F is a map from F


into F by definition. Any polynomial f (x) over F determines a function
by b 7→ f (b). Such a function is called a polynomial function. Let
f (x) and g(x) be distinct polynomials. We may ask whether they
determine different functions. At the first glance the answer seems to
be affirmative. However this is an illusion.
Let F be a finite field containing q elements. Then xq and x are
two distinct polynomials over F. We show that they determine the same
function on F.
The set F ∗ = F \{0} is a group of order q − 1 under multiplication.
Thus bq−1 = 1 for every b ∈ F ∗ by Lagrange’s theorem. Hence bq = b
88 CHAPTER 3. POLYNOMIALS AND RATIONAL FUNCTIONS

for every b ∈ F. This means that the polynomials xq and x determine


the same polynomial function.
This example tells us that the concepts of polynomial and polyno-
mial functions over a finite field are different. In fact, over a finite field
there are only finitely many functions but there are infinitely many
polynomials. A wonderful result is that every function on a finite field
is a polynomial function (cf. Exercise 3).

Theorem 3.5.2. Assume that f (x) ∈ F [x], b ∈ F. Then f (b) is the


remainder of f (x) divided by x − b.

Proof. By the division algorithm there are q(x), r(x) ∈ F [x] such
that
f (x) = (x − b)q(x) + r(x), (3.7)

with deg(r) < deg(x − b) = 1. Hence r(x) is a constant c ∈ F.


Replace x in (3.7) by b and f (b) = c is obtained.

Corollary 3.5.3. An element b ∈ F is a zero of f (x) if and only


if (x − b)|f (x).

Corollary 3.5.4. A nonzero polynomial of degree n over a field


F has at most n distinct zeros in F.

Proof. Suppose that a polynomial of degree n has n + 1 distinct


elements b1 , . . . , bn+1 . Then (x − bi )|f (x) for i = 1, . . . , n + 1. Hence
(x − b1 ) · · · (x − bn+1 )|f (x), which is impossible.

Corollary 3.5.5. Let f (x), g(x) be nonzero polynomials over F


with degrees at most n. If there are n + 1 distinct b1 , . . . , bn+1 ∈ F such
that f (bi ) = g(bi ) for every i, then f (x) = g(x).

Proof. Let h(x) = f (x) − g(x). Suppose that h(x) 6= 0, then h(x)
would be a nonzero polynomial of degree at most n with n + 1 distinct
zeros b1 , . . . , bn+1 , which is impossible. Hence h(x) = 0.

Definition 3.5.6. If (x − b)d |f (x) but (x − b)d+1 does not divide


f (x), then b is a zero of multiplicity d of f (x). A zero of f (x) of mul-
tiplicity greater than one is called a multiple zero of f (x).
3.5. * POLYNOMIAL FUNCTIONS 89

Theorem 3.5.7. Let f (x) be a nonzero polynomial of degree n and


let b1 , . . . , br be the set of all distinct zeros of f (x) with multiplicities
d1 , . . . , dr . Then
d1 + · · · + dr ≤ n.

Proof. Since

f (x) = (x − b1 )d1 · · · (x − br )dr g(x)

where g(x) is a nonzero polynomial. Hence

d1 + · · · + dr ≤ d1 + · · · + dr + deg(g) = deg(f ) = n.

Example 3.5.8. Let F be a finite field containing q elements. Then


Y
xq − x = (x − a)
a∈F

in F [x].

Proof. Since both sides are monic polynomials of degree q, it suf-


fices to show that the right hand side divides the left hand side. This is
true since aq = a for every element a ∈ F as we have verified before.

Exercises
1. Let F be a field containing infinitely many elements. Let f (x), g(x) ∈
F [x]. If f (x) and g(x) determine the same function on F, then f (x) =
g(x).
2. Let I = {f (x) ∈ R[x]|f (2) = f 0 (2) = f 00 (2) = 0}. Show that I is
an ideal of R[x] and find a monic polynomial that generates I. Is the
set J = {f (x) ∈ R[x]|f (2) = 0, f 0 (3) = 0} an ideal of R[x]? Explain
the reason.
3. Calculate the number of polynomial functions on a finite field
containing q elements. Use this to show that every function on a finite
field is a polynomial function.
Chapter 4

Vector Spaces

The freshman linear algebra is restricted to the linear algebra over


number fields, especially the real and complex number fields. Now that
the concept of abstract fields is at our disposal, most results we have
known in linear algebra can be generalized to arbitrary fields. There
are a few materials that can not be generalized in straight forward way.
The concepts of vector space and its related concepts will be for-
mulated in the framework of algebraic structures. This point of view
will enable us to deepen our understanding of the subject.

4.1 Vector Spaces and Linear Transformations

Definition 4.1.1. Let F be a field and let V be an (additive)


abelian group. If there is an operation

F × V → V, (a, u) 7→ au,

called the scalar multiplication, satisfying the following four conditions


1) 1u = u for any u ∈ V ;
2) (ab)u = a(bu) for any a, b ∈ F, u ∈ V ;
3) (a + b)u = au + bu for any a, b ∈ F, u ∈ V ;
4) a(u + w) = au + aw for any a ∈ F, u, w ∈ V.
then V is called a vector space over the field F. The elements in
V are called vectors.

90
4.1. VECTOR SPACES AND LINEAR TRANSFORMATIONS 91

A vector space is a structure of different nature from groups or rings.


The scalar multiplication is an operation different from those in groups
and rings. Instead of the multiplication of two elements in V, it is the
multiplication of two elements from two different sets F and V.
The concepts of linear independence and basis can be defined in the
same way as in elementary linear algebra.

Definition 4.1.2. 1) A set S of vectors in a vector space V over


a field F is linear independent if for any finite number of vectors
u1 , . . . , un with n > 0 in S and any nonzero elements c1 , . . . , cn in
F, c1 u1 + · · · + cn un is not equal to zero.
2) A subset S spans the vector space V if for every w ∈ V there
exist c1 , . . . , cn ∈ F, u1 , . . . , un ∈ S such that w = c1 u1 + · · · + cn un .
3) A linearly independent subset of a vector space is a basis of the
space if it spans the whole space. The number of elements in a basis is
defined to be the dimension of the space, denoted by dim(V ).

If a vector space is spanned by a finite number of vectors, then the


existence of the basis can be proved in the same way as in elementary
linear algebra. In that case, the number of elements in the basis is
independent of the choice of the basis. Hence the dimension is a well-
defined natural number. If a vector space cannot be spanned by a finite
number of vectors, then we say that it is an infinite dimensional space.
The exists a basis for an infinite dimensional space. The proof uses the
axiom of choice.
Let e1 , . . . , en be a basis of a finite dimensional vector space V.
Every vector u in V can be uniquely expressed as a linear combination
of e1 , . . . , en :
u = c 1 u 1 + · · · + c n un .

This gives a one to one correspondence between V and the space of


column vectors (or row vectors), which depends on the choice of basis.
A subset set W of a vector space V is called a subspace if it is closed
under addition and scalar multiplication.
Let V1 and V2 be two vector spaces over the same field F. A map
f from V1 to V2 is a linear map if it preserves the addition and scalar
multiplication, i.e., f (au + bv) = af (u) + bf (v) fo any a, b ∈ F and any
92 CHAPTER 4. VECTOR SPACES

u, v ∈ V1 . A linear map is also refered to as a homomorphism. When


V1 = V2 = V, a linear map is called a linear transformation. In analysis
a linear map is also called a linear operator. In algebra a linear trans-
form is sometimes called an endomorphism. A bijective endomorphism
is an automorphism, which is an invertible linear transformation.
Let f be a linear transformation of a finite dimensional vector space
V. After fixing a basis e1 , . . . , en of V the linear transformation f de-
termines an n × n matrix with entries in F. The rules of addition and
multiplication of matrices over an arbitrary field are the same as over
a number field. The concepts such like rank of a matrix, determinant
of a square matrix are defined without modification.
The set of all invertible linear transformations forms a group un-
der composite operation, denoted by GL(V ), which is isomorphic to
GLn (F ).
The theory of system of linear equations, characteristic polynomials,
eigenvalues and eigenvectors still holds for arbitrary field.
Besides groups, rings, fields, vector spaces, there is another useful
algebraic structure. Let A be a ring containing a field F as a subring.
Then A is called an F -algebra.
For example, the ring Mn (F ) consisting of all n × n matrices over a
field F is an F -algebra, the skew field of quaternions is an R-algebra.
Every F -algebra A is a vector space over F, in which the scalar mul-
tiplication coincides with the multiplication of A. If a field L contains a
subfield F, then L is an F -algebra. There is an underlying structure of
F -space. This point of view is useful, which we will discuss in greater
details in the future.
Let p be a prime number. An abelian group whose every nonzero
element has order p has a hidden structure of vector space over Fp , as
explained in the next lemma.

Lemma 4.1.3. Let p be a prime number and let G be an (additive)


abelian group. If pa = 0 for every nonzero a ∈ G, then G is a vector
space over the finite field Fp .

Proof. It is known that Fp = {0̄, 1̄, . . . , p − 1}. For every ī ∈ Fp


and u ∈ G, define the scalar multiplication īu = iu. Let us verify that
4.1. VECTOR SPACES AND LINEAR TRANSFORMATIONS 93

G become a vector space under this scalar multiplication


1) 1̄u = 1u = u is obvious.
2) For arbitrary ī, j̄ ∈ Fp , we have īj̄ = k̄, where k = ij − rp for
some nonnegative integer r. Thus (īj̄)u = ku = (ij − rp)u. Since the
order of u divides p, we have rpu = 0. Therefore (īj̄)u = iju = ī(j̄u).
3) Similarly one can easily show that (ī + j̄)u = īu + j̄u
and
4) ī(u + v) = i(u + v) = iu + iv = īu + īv.

The following example shows how useful this lemma is.

Example 4.1.4. Let p be a prime number and let A = Zp2 ⊕ Zp2 ⊕


Zp3 . Find the number of all non-cyclic subgroups of A with order p2 .

Solution:
Every element of A is expressed as (a, b, c) in which a, b ∈ Zp2 and
c ∈ Zp3 . Its order is max{o(a), o(b), o(c)}. Recall that o(a) is the order
of a. A subgroup H of A with order p2 is non-cyclic if and only if the
order of every nonzero element in H has order p. Let V be the subset
of all elements in A with order less than or equal to p. Then V is a
subgroup of order p3 . By Lemma 4.1.3 G is a 3-dimensional space over
Fp . Hence a subgroup H of A with order p2 is non-cyclic if and only if
it is a 2-dimensional subspace of V. Thus the problem is converted into
the counting of 2-dimensional subspaces in a 3-dimensional space over
Fp . This becomes a problem in linear algebra.
There are at least two methods of counting.
A 2-dimensional subspace is spanned by two linearly independent
vectors (a1 , a2 , a3 ) and (b1 , b2 , b3 ). The first vector is a nonzero vector.
Hence there are p3 − 1 choices for (a1 , a2 , a3 ). The second vector is
linear independent from the first if and only if (b1 , b2 , b3 ) 6= λ(a1 , a2 , a3 )
for any λ ∈ Fp \{0}. Hence with fixed (a1 , a2 , a3 ), the second vector
(b1 , b2 , b3 ) has p3 − p choices. Thus the number of choices of ordered
pairs of linearly independent vectors is (p3 − 1)(p3 − p).
Two ordered pairs {(a1 , a2 , a3 ), (b1 , b2 , b3 )} and {(c1 , c2 , c3 ), (d1 , d2 , d3 )}
of linearly independent vectors span the same 2-dimensional subspace
94 CHAPTER 4. VECTOR SPACES

if and only if there is a nonsingular matrix


" #
p11 p12
P =
p21 p22

such that " # " #


a1 a2 a3 c1 c2 c3
=P .
b1 b2 b3 d1 d2 d3
By using the same method as above, the number of 2 × 2 nonsingular
matrices with entries in Fp is (p2 − 1)(p2 − p). Therefore the number of
2-dimensional spaces of F3p is

(p3 − 1)(p3 − p)
2 2
= p2 + p + 1.
(p − 1)(p − p)

The second method is more elegant. Every 2-dimensional subspace


is the space of solutions of a nonzero homogeneous linear equation.
Two homogeneous linear equations

a1 x 1 + a2 x 2 + a3 x 3 = 0

and
b1 x 1 + b2 x 2 + b3 x 3 = 0

have the same solutions if and only if there is a nonzero λ ∈ Fp such


that
(a1 , a2 , a3 ) = λ(b1 , b2 , b3 ).

Hence the total number of 2-dimensional subspaces is equal to

p3 − 1
= p2 + p + 1.2
p−1

Exercises
1. Let V be the real vector space of all continuous real functions
on the closed interval [0, 1]. Show that x3 , sin x and cos x are linearly
independent.
2. Let V be a vector space of finite dimension over a field F and
4.2. QUOTIENT SPACES 95

let f be a linear transformation of V. Show that f is surjective if and


only if f is injective.

3. Let R be an integral domain containing a field F. If R, regarded


as a vector space over F, has finite dimension, then R is a field.

4.2 Quotient spaces

Let W be a subspace of a vector space V over a field F. By disregarding


the scalar multiplication, V becomes and abelian group. Hence W is
a normal subgroup of V. Let V /W the the quotient group. We wish to
construct a scalar multiplication on V /W so that it becomes a vector
space over F.
For any a ∈ F and any x̄ ∈ V /W, in which x ∈ V, define

ax̄ = ax.

Let x0 ∈ V be another representative of x̄. Then x0 − x ∈ W. So


ax0 − ax = a(x0 − x) ∈ W, which means that ax0 = ax. Hence ax̄ does
not depend upon the choice of x. It is easy to verify that V /W becomes
a vector space over F under this scalar multiplication. The space V /W
is called the quotient space of V over the subspace W.
Let w1 , . . . , wr be a basis of the subspace W. Extend it to a basis
w1 , . . . , wr , ur+1 , . . . , un of the whole space V.

Proposition 4.2.1. ūr+1 , . . . , ūn form a basis of V /W. In particu-


lar, dim(V /W ) = dim(V ) − dim(W ).

Proof. Assume that ar+1 , . . . , an ∈ F satisfy

ar+1 ūr+1 + · · · + an ūn = 0̄.

Then
ar+1 ur+1 + · · · + an un ∈ W.

Hence
ar+1 ur+1 + · · · + an un = b1 w1 + · · · + br wr .
96 CHAPTER 4. VECTOR SPACES

It follows that

−b1 w1 − · · · − br wr + ar+1 ur+1 + · · · + an un = 0.

Since w1 , . . . , wr , ur+1 , . . . , un are linearly independent,

ar+1 = · · · = an = 0.

This proves that ūr+1 , . . . , ūn are linearly independent in V /W.


For any x̄ ∈ V /W there exist a1 , . . . , an ∈ F such that

x = a1 w1 + · · · + ar wr + ar+1 ur+1 + · · · + an un .

Hence
x̄ = ar+1 ūr+1 + · · · + an ūn .

This shows that every element of V /W is a linear combination of


ūr+1 , . . . , ūn .

An arbitrary linear transformation f of V generally does not induce


a linear transformation of the quotient space V /W.

Proposition 4.2.2. For an f -invariant subspace W, which means


that f (u) ∈ W for any u ∈ W, define

f¯ : V /W → V /W, x̄ 7→ f (x).

Then f¯ is a linear transformation of V /W.

Proof. An important task is to check that the map f¯ is well-


defined, i.e.,f (x) does not depend upon the choice of the representative
x.
Let x0 be another representative of x̄. Then x0 − x ∈ W. Since W is
invariant under f, f (x0 − x) ∈ W. Hence f (x0 ) = f (x).
The remaining part of the proof is straightforward and is left to the
readers.

Let u1 , . . . , un be a basis of V, in which u1 , . . . , ur form a basis of


4.2. QUOTIENT SPACES 97

the f -invariant subspace W. The matrix of f under this basis is


" #
A B
,
0 C

in which A and C are square matrices of orders r and n−r respectively.


It is easy to see that A is the matrix of f |W under the basis u1 , . . . , ur
and C is the matrix of f¯ under the basis ūr+1 , . . . , ūn .

Exercises
1. Let p be a prime number and let H be a cyclic group of order
p. Let G be the direct sum of n copies of H, i.e., G = H n . Assume
that 1 ≤ n ≤ p. Show that the automorphism group Aut(G) does not
contain elements of order p2 .
Chapter 5

Topics in Group Theory

In this chapter we choose three topics in group theory. The first one
is the Sylow theorem for finite groups. Since its proof involves clever
application of group actions, the basic technique of group actions is
discussed in the first section. The second topic is the structure theorem
of finitely generated abelian groups, which has many applications. It is
indispensable in some branches of mathematics such as number theory,
algebra and topology. The role of this theorem is similar to the Jordan
canonical form in linear algebra. The last topic is the solvable groups,
which will be used in the last chapter on Galois theory.

5.1 The Orbit Formula of an Action by a Finite


Group

The group action is a powerful tool in the theory of finite groups. Its
application is full of tricks. For basic definitions see the last section of
Chapter 1.
Assume that a finite group G acts on a finite set S. The S is the
union of mutually disjoint orbits. Let Gx be an orbit. Then

|Gx| = (G : Stab(x)).

This formula can also be expressed as |Gx| · |Stab(x)| = |G|, which


roughly means that the larger the orbit is the smaller the stabilizer is.
Unlike the coset decomposition of a group, different orbits may contain

98
5.1. THE ORBIT FORMULA OF AN ACTION BY A FINITE GROUP99

different number of elements.


Choose one representative in each orbit and they form a finite set
D = {x1 , . . . , xd }. This means that S is the disjoint union of the orbits
Gx1 , . . . , Gxd . Such a set D is called the complete system of represen-
tatives of the orbits. Thus we obtain the following formula
d
X
|S| = (G : Stab(xi )). (5.1)
i=1

which is referred to as the orbit decomposition formula or simply orbit


formula. This is the basis of a counting technique in finite groups. This
formula will be used to prove the celebrated Sylow theorem in the next
section. In order to get familiar with this technique, let us look at some
simple examples.

Example 5.1.1. The order of the group of symmetry of a cube.

This group acts transitively on the set of eight vertices of the cube.
Fix one vertex. The stabilizer of that vertex is a group of order 6.
Hence the order of the group of symmetry is equal to 48.
The orders of groups of symmetry for other regular polyhedra can
be computed in the same way.

Example 5.1.2. Let G be a finite group with |G| > 2 and the
conjugacy class of one element a ∈ G contains two elements. Then G
is not a simple group.

Proof. The group G acts on G itself by conjugacy. By the given


condition the orbit on which a is located has length 2. Thus (G :
Stab(a)) = 2. So Stab(a) / G by 1.5.8. Since 1 < Stab(a) < |G|,
Stab(a) is a nontrivial normal subgroup of G.

The examples above only use the formular |Gx| = (G : Stab(x)).


The following example is the application of the orbit formula 5.1.

Example 5.1.3. Let p be a prime number and let G be a group with


|G| = pn (n > 1). Assume that H / G and |H| = p. Then H ⊆ C(G),
where C(G) is the center of G.
100 CHAPTER 5. TOPICS IN GROUP THEORY

Proof. Since H is a normal subgroup of G, the group G acts on


H by conjugacy. Let e = x1 , . . . , xm be a complete system of represen-
tatives of the orbits. It follows from the orbit formula that
m
X
(G : Stab(xi )) = |H| = p.
i=1

Since|G| = pn , every number (G : Stab(xi )) is a power of p. Since


Stab(e) = G, i.e., (G : Stab(x1 )) = 1, so m = p and Stab(xi ) = G for
each i. Hence xi ∈ C(G) for every i.

Let p be a prime number. A group of order pn with n > 0 is called


a p-subgroup.

Proposition 5.1.4. Let G be a p-subgroup for some prime number


p. Then the following statements hold:
1) The center Z of G contains more than one element;
2) (Normalizer grows) Let H be a proper subgroup of G. Then H 6=
NG (H).

Proof. 1) Assume that |G| = pn with n > 0. Consider the action


of G on itself by conjugacy. Let Z denote the center of G and let
r = |Z|. Then Z = {x ∈ G|Stab(x) = G}. It follows from 1 ∈ Z that
r > 0. By the orbit formula the sum of the lengths of all orbits is equal
to pn . Since the length of every orbit containing more than one element
is divisible by p, |Z| is divisible by p. Hence r > 1.
2) Suppose that n is the smallest integer such that the statement
fails. By hypothesis there is a proper subgroup H of G such that
H = NG (H). Since Z ⊆ NG (H), the subgroup H = NG (H) contains
Z. Since |H| > 1 by the result of 1), the order of G/Z is strictly less
then pn . Since H/Z is a proper subgroup of G/Z, we have H/Z 6=
NG/Z (H/Z). There is some g ∈ G such that ḡ ∈ NG/Z (H/Z)\H/Z.
Hence g ∈ NG (H)\H, leading to contradiction.

Example 5.1.5. Any group G of order p2 is abelian for any prime


number p.
5.2. SYLOW SUBGROUPS 101

Proof. If G is not abelian, then the center Z of G is a proper


subgroup of G. Then Proposition 5.1.4 implies that |Z| = p. Hence
G/Z is a cyclic group. Let ȳ with y ∈ G be a generator of G/Z. Then
G = {ay i }a∈Z,0≤i<p . It is clear ay i commutes with by j for any a, b ∈ Z
and any i, j.

It follows that the smallest possible non-abelian p-subgroup has or-


der 8. Indeed such a group exists: The eight quaternions ±1, ±i, ±j, ±k
form a non-abelian group of order 8.
An elegant application of the orbit formula is the proof of Wedder-
burn’s theorem of finite skew fields. Refer to Appendix 2.
Exercises

1. Let n = 2m, where m is an odd integer greater than one. Let G


be a subgroup of Sn of order 2r in which r is a natural number. Show
that there are 1 ≤ i < j ≤ n such that σ(i) ∈ {i, j}, σ(j) ∈ {i, j} for
every σ ∈ G.
2. Let G be a finite group of order n. Denote by S(G) the set of all
bijections from G to G. Let στ denote the composite of σ, τ ∈ S(G).
Then S(G) becomes a group.
1) Show that S(G) is isomorphic to Sn ;
2) Prove Cayley’s theorem: every finite group is isomorphic to a
permutation group.
3) Let G be a subgroup of the symmetric group Sn , in which n is
an odd integer. Assume that |G| is a power of 2. Show that there is
some i among 1, 2, . . . , n such that σ(i) = i for all σ ∈ G.

5.2 Sylow subgroups

In this section G is always a fixed finite group with |G| = pr s, in which


p is a prime number, r > 0, and s is not divisible by p.
A subgroup H of G with |H| = pr is called a Sylow p-subgroup of
G, or simply a Sylow group if p is fixed.
The Sylow’s theorem answers the following three questions:
1) Does a Sylow subgroup always exist?
102 CHAPTER 5. TOPICS IN GROUP THEORY

2) How many Sylow subgroups are there?


3) What is the relation between Sylow subgroups?
First we need a preparatory lemma.

Lemma 5.2.1. For a natural number n = pr s, in which r > 0 and s


is not divisible by p, the congruence
 
n
≡ s (mod p)
pr

holds.
r
Proof. First of all pnr is the coefficient of xp in the expansion of


(1 + x)n .
Over the finite field Fp the equality
r r
(1 + x)n = [(1 + x)p ]s = (1 + xp )s
r
holds. Hence the coefficient of xp of (1 + x)n over Fp is equal to s̄.
This means that  
n
≡ s (mod p).
pr

Now we state and prove the main theorem on Sylow subgroups.

Theorem 5.2.2 (Sylow). Let G be a finite group and let p be a


prime factor of |G|. For this fixed p the following statements hold:
1) Sylow subgroups exist in G.
2) Let m be the number of Sylow groups in G, then m ≡ 1 (mod p).
3) All Sylow groups are conjugate.

Proof. Assume that |G| = n = pr s, in which r > 0 and p does


not divide s. Let Γ be the set of all subsets of G containing pr elements
and let N be the number of members in Γ. Then N = pnr . By Lemma


5.2.1 N ≡ s (mod p).


For any A ∈ Γ and any g ∈ G, let gA = {ga|a ∈ A}. Then gA ∈ Γ.
This gives an action of G on Γ.
5.2. SYLOW SUBGROUPS 103

Let A1 , A2 , . . . , Aq ∈ Γ be a complete system of representatives of


the orbits, i.e., they are in different orbits and every member of Γ is
in the same orbit with some Ai . For each 1 ≤ i ≤ q Let Hi denote the
stabilizer of Ai . This means that ha ∈ Ai for any a ∈ Ai , h ∈ Hi . Then
Ai is a union of some right cosets of Hi . Hence |Hi | divides the number
of elements in Ai , which implies that |Hi | = pdi for some integer di ≤ r.
The number of elements in the orbit of Ai is equal to |G|/pdi =
Pq
pr−di s. By Formula (5.1) N is equal to i=1 p
r−di
s. Since N is not
divisible by p, there is some i such that di = r. The corresponding Hi
is a Sylow p-subgroup. Thus 1) is proved.
Moreover, all orbits can be divided into two families: the first family
consists of all orbits of length s and the second family contains the
remaining orbits. This implies that the length of each orbit in the
second family is divisible by p. Assume that there are m orbits in the
first family. Then N ≡ ms (mod p) by the equality N = qi=1 pr−di s.
P

It follows from N ≡ s (mod p) that m ≡ 1 (mod p).


Let A be a member in an orbit of the second family. Then A is not
a subgroup for otherwise A would be contained in its stabilizer, which
would contradict the inequality (G : Stab(A)) > s.
Every orbit of the first family contains an element A such that
1 ∈ A. Let H be the stabilizer of A. Then h = h · 1 ∈ A for each h ∈ H.
Thus H ⊆ A. So H = A since |H| = pr = |A|. Hence A is a Sylow p-
subgroup of G. Moreover the orbit containing A consists all left cosets
of H in G. This shows that every orbit of the first family contains exact
one Sylow p-subgroup. Therefore there are m Sylow p-subgroups in G
and 2) is proved.
Finally we prove a statement stronger than 3):
Let P is a Sylow p-subgroup and let K be a subgroup of G with
|K| = pt for some natural number t. Then there is some g ∈ G such
that K ⊆ gP g −1 .
Let Σ be the set of all subgroups of G conjugate to P. Then gBg −1 ∈
Γ for any g ∈ G, B ∈ Σ. This gives a transitive action of G on the set
Σ. Since P is contained in its stabilizer, the number of elements in Σ
is a factor of s, which is not divisible by p.
This action induces an action of K on Σ. Let Ω ⊆ Σ be an orbit
104 CHAPTER 5. TOPICS IN GROUP THEORY

of the action of K on Σ. For any B ∈ Ω, if the stabilizer of B is not


equal to K then |Ω| is divisible by p. However the sum of lengths of
all orbits is equal to |Σ|, which is not divisible by p. Thus there is
some P 0 = gP g −1 ∈ Σ whose stabilizer is equal to K. This implies that
K ⊆ NG (P 0 ). Since P 0 / NG (P 0 ), the conditions to apply the second
isomorphism theorem 1.7.14 are satisfied (see the figure below). So
KP 0 /P 0 ∼
= K/(K ∩ P 0 ).

NG (P 0 )

KP 0HH
vv HH
vvv HH
vv HH
vv H
0
K HH vP
HH vv
HH vv
HH vv
H vv
K ∩ P0
Hence |KP 0 /P 0 | = |K/K ∩ P 0 ] is a power of p, which implies that
|KP 0 | is also a power of p. Hence |KP 0 | ≤ pr = |P 0 |,which implies
that KP 0 = P 0 . Therefore K ⊆ P 0 . This concludes the proof of the
statement.
It is easy to use this statement to prove 3). Let P and P 0 be
arbitrary Sylow p-subgroups. By the statement we have just proved
there is g ∈ G such that P 0 ⊆ gP g −1 . It follows from |P 0 | = |gP g −1 |
that P 0 = gP g −1 .

Corollary 5.2.3. Let G be a finite group of order pr s, where r > 0,


s is not divisible by the prime number p. Then the number of Sylow p-
subgroups in G divides s.

Proof. Let n be the number of Sylow p-subgroups in G. Since the


conjugate action of G on the set of all Sylow p-subgroups is transitive
by 3) of Sylow’s theorem, the number n divides pr s. On the other hand,
n ≡ 1 (mod p) implies that n is not divisible by p. Hence n|s.

Corollary 5.2.4. Let p be a prime factor of the order of a finite


group G. Then there exists an element of order p.
5.2. SYLOW SUBGROUPS 105

Proof. Take any element a of a Sylow p-subgroup P of G which


is not the identity element. Then the order of a is pt for some natural
t−1
number t and ap is an element of order p.

Remark 5.2.5. In some textbooks this corollary appears as a prepara-


tory lemma to prove Sylow’s theorem.

For a finite group, Corollary 5.2.3 and 2) of Sylow’s theorem can


often be used to prove that this group has only one Sylow p-subgroup
for a prime factor p of the order of the group. This will imply that
the group is not a simple group if p is not the only prime factor of the
given group. Look at the following typical example.

Example 5.2.6. Let G be a group of order pr s, in which p is a


prime number, r > 0. Assume that 1 < s < p. Then G is not a simple
group.

Proof. Let n denote the number of Sylow p-subgroups of G. Then


n ≡ 1 (mod p) by Sylow’s theorem and n divides s by Corollary 5.2.3.
It follows from the hypothesis s < p that n = 1, which means that G
has only one Sylow p-subgroup, which is a nontrivial normal subgroup
of G.

Exercises
1. Let p be a prime number and let Fp be the finite field containing
p elements. Let GLn (Fp ) be the group of all invertible n × n matrices
over Fp and let U be the subgroup of GLn (Fp ) consisting of all upper
triangular matrices whose elements on the diagonal are equal to 1.
Show that U is a Sylow p-subgroup of GLn (Fp ) and find the number
of conjugates of U in GLn (Fp ).
2. Let H be a subgroup of a finite group G such that every element
in G is conjugate to some element in H. Show that H = G.
3. Prove that a group of order 1225 is abelian.
4. Prove that a group of order 640 is not a simple group.
5. Let G be a simple group of order 168 and let S be the set of all
Sylow 7-subgroups of G. Let H ∈ S. Then H acts on S by conjugacy,
106 CHAPTER 5. TOPICS IN GROUP THEORY

i.e., g ∈ H carries K ∈ S into gKg −1 . Find the number of orbits and


the length of each orbit.
6. Find all Sylow 2-subgroups of S4 .
7. Assume that a finite group G has 50 Sylow 7-subgroups and P
is a Sylow 7-subgroup. Let N = NG (P ) = {g ∈ G|gP g −1 = P }.
1) Show that N is a maximal subgroups of G, i.e.,N 6= G and there
is no subgroup between N and G.
2) If N has a normal Sylow 5-subgroup Q, then Q  G. [ Hint Use
2) of Proposition 5.1.4.]
8. Let H be a proper normal subgroup of a finite group G and let p
be a prime factor of |G/H|. Show that the number of Sylow p-subgroups
of G/H divides the number of Sylow p-subgroups of G.

5.3 * Structure of finitely generated abelian groups

Let G be an abelian group, whose operation is always denoted by + in


this section. If there are finitely many elements g1 , . . . , gn in G such that
every element g ∈ G can be expressed as g = k1 g1 + · · · + kn gn for some
integers k1 , . . . , kn , then G is called a finitely generated abelian group
and {g1 , . . . , gn } is a set of generators. If furthermore the coefficients
k1 , . . . kn are uniquely determined by g for every g ∈ G, then G is called
a finitely generated free abelian group and {g1 , . . . , gn } is a basis of G.
Let G be a finitely generated free abelian group with basis {g1 , . . . , gn }.
It is easy to verify that the map

f : Z ⊕ · · · ⊕ Z → G,

(k1 , . . . , kn ) 7→ k1 g1 + · · · + kn gn

is well defined and f is an isomorphism.


This tells us that the structure of a finitely generated free abelian
group is simple. It is the direct sum of a finite number of copies of Z. Its
structure is totally determined by the number of elements in the basis.
This number is called the rank of the group, similar to the dimension
of a finite dimensional vector space.
5.3. * STRUCTURE OF FINITELY GENERATED ABELIAN GROUPS107

Next let us investigate the structure of a finite abelian group, which


is more complicated.

Lemma 5.3.1. Let k1 , . . . , kn be integers, not all zero and gcd(k1 , . . . , kn ) =


1. Then there exists an n × n matrix A with integral entries of deter-
minant ±1 whose first row is k1 , . . . , kn .

Proof. Apply induction on n. The lemma is obviously true for


n = 1. Assume that n > 1.
If k2 = · · · = kn = 0 then k1 ± 1, then replacing the first row of
the n × n identity matrix by k1 , . . . , kn produces a matrix satisfying
the requirement. It remains to consider that case that ki 6= 0 for some
2 ≤ i ≤ n. Let d = gcd(k2 , . . . , kn ). Then there are integers s, t such
sk1 + td = 1.
Let qi = ki /d for 2 ≤ i ≤ n. Then q2 , . . . , qn are integers such
that gcd(q2 , . . . , qn ) = 1. By induction hypothesis there are integers
bij (2 ≤ i ≤ n − 1, 2 ≤ j ≤ n) such that

q2
q3 ··· qn

b
22 b23 ··· b2n
= ±1.

···



bn−1,2 bn−1,3 · · · bn−1,n

Without loss of generality we may assume that this determinant is


equal to 1. Then

k
1 k 2 k 3 · · · k n


−t sq2 sq3 · · ·

sqn

0
b 22 b 23 · · · b 2n
= sk1 + td = 1.


···


0 bn−1,2 bn−1,3 · · · bn−1,n

Lemma 5.3.2. Let A be an abelian group and let x, y ∈ A be two


elements of finite order n and m respectively. If gcd(m, n) = 1, then
the order of x + y is mn.
108 CHAPTER 5. TOPICS IN GROUP THEORY

Proof. It is evident that nm(x + y) = 0. Assume that k(x + y) =


0. Then kx = −ky. Since o(kx)|n and o(−ky)|m, we have o(kx) =
o(−ky) = 1. Hence kx = ky = 0, which implies mn|k.

Lemma 5.3.3. Let A be an abelian group and let x, y ∈ A be two


elements of finite orders n and m respectively. Let lcm(n, m) be the
least common multiple of n and m. There exist integers s and t such
that the order of sx + ty is lcm(n, m).

Proof. The natural numbers n and m can be written as


r
Y r
Y
n= pei i , m = pfi i
i=1 i=1

where p1 , . . . , pr are distinct prime numbers and e1 , . . . , er , f1 , . . . , fr


are non-negative integers. For each 1 ≤ i ≤ r, let
(
ei , if ei ≥ fi
ui =
0, otherwise

and (
fi , if ei < fi
vi = .
0, otherwise

Let u = ri ui and v = ri vi . Then u|n, v|m, gcd(u, v) = 1 and uv =


Q Q

lcm(n, m). Since o((n/u)x) = u, o((m/v)y) = v, the order of (n/u)x +


(m/v)y is uv by Lemma 5.3.2.

Lemma 5.3.4. Let x and y be two elements of finite orders of an


abelian group A. Let n and m be the orders of x and y respectively.
Assume that m|n and dy ∈ hxi for some natural number d. Then dy ∈
hdxi.

Proof. It is easy to verify the m/ gcd(m, d) divides n/ gcd(n, d).


Hence o(dy)|o(dx), which implies dy ∈ hdxi.

Theorem 5.3.5. Let A be a finite abelian group with |A| > 1 gen-
erated by a1 , . . . , an . Then A can be decomposed into a direct sum of
cyclic groups B1 , . . . , Bm of orders d1 , . . . , dm respectively, satisfying
the following conditions:
5.3. * STRUCTURE OF FINITELY GENERATED ABELIAN GROUPS109

1) m ≤ n;
2) dm > 1 and di divides di−1 for 2 ≤ i ≤ m.

Proof. Apply induction on |A|. If |A| = 2 then A is a cyclic group


and the theorem holds. Assume that |A| > 2.
Take an element b01 in A of highest order d1 . Write

b01 = s1 a1 + · · · + sn an ,

where si ∈ Z. Let q = gcd(s1 , . . . , sn ), ki = si /q, and let

b1 = k1 a1 + · · · + kn an ∈ A.

Then b01 = qb1 and the order of b1 is greater than or equal to d1 . Since
d1 is an upper bound of orders of elements in A, the order b1 is equal
to d1 . Note that gcd(k1 , . . . , kn ) = 1.
Lemma 5.3.3 implies that the order of every element in A divide d1 .
Since gcd(k1 , . . . , kn ) = 1, Lemma 5.3.1 implies that there are inte-
gers cij (2 ≤ i ≤ n − 1, 1 ≤ j ≤ n) such that

k1
· · · kn
c
21 · · · c2n
= ±1.
···



cn1 · · · cnn

Let
e2 = c21 a1 + · · · + c2n an ,

···

en = cn1 a1 + · · · + cnn an .

Then b1 , e2 , . . . , en generate A.
Let B1 be the cyclic subgroup of A generated by b1 and let Ā =
A/B1 . If |Ā| = 1, then A = B1 and there is nothing more to prove.
Assume that |Ā| > 1. Denote by ē2 , . . . , ēn the images of e2 , . . . , en
in Ā. Then ē2 , . . . , ēn generate Ā. By induction hypothesis there are
f2 , . . . , fm ∈ A such that
1) m ≤ n;
110 CHAPTER 5. TOPICS IN GROUP THEORY

2) Ā = hf¯2 i ⊕ · · · ⊕ hf¯m i;
3) dm > 1 and di divides di−1 for 3 ≤ i ≤ m, where di is the order
¯
of fi in Ā for 2 ≤ i ≤ m.
The element di fi is in B1 for 2 ≤ i ≤ m. By Lemma 5.3.4 di fi ∈
hdi b1 i, since the order of fi divides that of b1 . Hence di fi = ui di b1 for
some natural number ui .
Let bi = fi − ui b1 . Then b̄i = f¯i and the order of bi is equal to di .
For each 2 ≤ i ≤ m let Bi be the cyclic subgroup of A generated by
bi . We claim that A = B1 ⊕ B2 ⊕ · · · ⊕ Bm . It is obvious that b1 , . . . , bm
generate A. Assume that c1 , . . . , cm ∈ Z satisfy

c1 b1 + · · · + cm bm = 0.

Then
c2 b̄2 + · · · + cm b̄m = 0̄,

i.e.,
c2 f¯2 + · · · + cm f¯m = 0̄.

By the conditions that f2 , . . . , fm satisfy, di |ci for 2 ≤ i ≤ m. Hence


ci bi = 0 for 2 ≤ i ≤ m, which implies c1 b1 = 0. This proves the
claim.

Corollary 5.3.6. Let A be a finite abelian group. Then A can


be decomposed into the direct sum of some cyclic groups B1 , . . . , Bm of
orders d1 , . . . , dm respectively such that

1 < dm |dm−1 | · · · |d2 |d1 .

The number m and the decreasing sequence of natural numbers d1 , d2 , . . . , dm


are totally determined by A.
Proof. The first half is a consequence of Theorem 5.3.5. Assume
that A is decomposed into the direct sum of another set of cyclic groups
C1 , . . . , Cn of orders e1 , . . . , en respectively such that

1 < en |en−1 | · · · |e2 |e1 .

Both d1 and e1 are equal to maxa∈A o(a). Hence d1 = e1 . Suppose


5.3. * STRUCTURE OF FINITELY GENERATED ABELIAN GROUPS111

that the decreasing sequences d1 , d2 , . . . , dm and e1 , e2 , . . . , en are dif-


ferent. Since m
Q Qn
i=1 di = i=1 ei , there is r > 1 such that di = ei for
1 ≤ i < r and dr 6= er . Without loss of generality we may assume
that dr < er . Let K = dr A = {dr a|a ∈ A}. It is obvious that K is a
subgroup of A. The decomposition A ∼ = B1 ⊕ · · · ⊕ Bm yields
r−1
Y di
|K| = ,
i=1
dr

while the second decomposition A ∼


= C1 ⊕ · · · ⊕ Cn implies that
r r−1
Y ei Y ei
|K| = > .
d
i=1 r
d
i=1 r

This leads to contradiction.

The proof of Theorem 5.3.5 is constructive. It gives an algorithm


to find the explicit decomposition.

Example 5.3.7. Let

B = Z8 ⊕ Z12 ⊕ Z16

and let A be the subgroup of B generated by

u = 2̄ ⊕ 3̄ ⊕ 2̄,

v = 4̄ ⊕ 4̄ ⊕ 8̄,

w = 4̄ ⊕ 6̄ ⊕ 6̄.

Find the decomposition of A and generators of cyclic components sat-


isfying the conditions in Theorem 5.3.5.

Solution:
The orders of the elements u, v, w are 8, 6, 8 respectively. Hence
u + v = 6̄ ⊕ 7̄ ⊕ 10 is an element of highest order in A. Its order is equal
112 CHAPTER 5. TOPICS IN GROUP THEORY

to 24. Due to
1 1 0

0 1 0 = 1,


0 0 1

A is generated by u + v, v, w.

Let z1 = u + v. Then

z1 = 6̄ ⊕ 7̄ ⊕ 10,

2z1 = 4̄ ⊕ 2̄ ⊕ 4̄,

3z1 = 2̄ ⊕ 9̄ ⊕ 14,

4z1 = 0̄ ⊕ 4̄ ⊕ 8̄,
¯ ⊕ 2̄,
5z1 = 6̄ ⊕ 11

6z1 = 4̄ ⊕ 6̄ ⊕ 12,

7z1 = 2̄ ⊕ 1̄ ⊕ 6̄,

8z1 = 0̄ ⊕ 8̄ ⊕ 0̄,

9z1 = 6̄ ⊕ 3̄ ⊕ 10,
¯ ⊕ 4̄,
10z1 = 4̄ ⊕ 10

11z1 = 2̄ ⊕ 5̄ ⊕ 14,

12z1 = 0̄ ⊕ 0̄ ⊕ 8̄,

13z1 = 6̄ ⊕ 7̄ ⊕ 2̄,

14z1 = 4̄ ⊕ 2̄ ⊕ 12,

15z1 = 2̄ ⊕ 9̄ ⊕ 6̄,

16z1 = 0̄ ⊕ 4̄ ⊕ 0̄,
¯ ⊕ 10,
17z1 = 6̄ ⊕ 11

18z1 = 4̄ ⊕ 6̄ ⊕ 4̄,

19z1 = 2̄ ⊕ 1̄ ⊕ 14,
5.3. * STRUCTURE OF FINITELY GENERATED ABELIAN GROUPS113

20z1 = 0̄ ⊕ 8̄ ⊕ 8̄,

21z1 = 6̄ ⊕ 3̄ ⊕ 2̄,
¯ ⊕ 12,
22z1 = 4̄ ⊕ 10

23z1 = 2̄ ⊕ 5̄ ⊕ 6̄.

Denote B̄ = B/hz1 i.
Since

v = 4̄ ⊕ 4̄ ⊕ 8̄ ∈
/ B/Zz1 , 2v = 0̄ ⊕ 8̄ ⊕ 0̄ = 8z1 ,

the element [v] ∈ B̄ has order 2. Similarly [w] ∈ B̄ has order 4 and
4w = 12z1 .
Since 2w−v = 4̄⊕ 8̄⊕ 4̄ ∈/ hz1 i, so 2[w] 6= [v]. Hence B̄ = h[w]i⊕h[v]i.
Let v = v − 4z1 , w = w − 3z1 . Then the orders of v 0 and w0 in A are
0 0

equal to 2 and 4 respectively. Let z2 = w0 = 2̄⊕ 9̄⊕ 8̄, z3 = v 0 = 4̄⊕ 0̄⊕ 0̄.
Then A = hz1 i ⊕ hz2 i ⊕ hz3 i with o(z1 ) = 24, o(z2 ) = 4, o(z3 ) = 2. In
particular |A| = 24 × 4 × 2 = 192.

Lemma 5.3.8. Let A be a finitely generated abelian group. If every


element of A has finite order then A is a finite group.

Proof. Assume that A is generated by a1 , a2 , . . . , an of orders


d1 , d2 , . . . , dn respectively. Then every element of A can be written
as
k1 a1 + k2 a2 + · · · + kn an

with 0 ≤ ki < di , although not necessarily uniquely. Hence |A| ≤


d1 d2 · · · dn .

Next we discuss the structure of a finitely generated abelian group.

Theorem 5.3.9. A finitely generated abelian group A is isomorphic


to Zn ⊕ Zd1 ⊕ · · · ⊕ Zdm , in which

1 < dm |dm−1 | · · · |d2 |d1 .

The natural numbers n, m and the sequence d1 , d2 , . . . , dm are uniquely


determined by A.
114 CHAPTER 5. TOPICS IN GROUP THEORY

Proof. Since Corollary 5.3.6 covers the case where A is finite, we


may assume that |A| = ∞.
Let T be the set of all elements of finite order in A. Then T is a
normal subgroup of A and K = A/T is also a finitely generated abelian
group. Due to Lemma 5.3.8, K 6= {0}.
We claim that the only element of finite order is K is 0̄. Assume
that a ∈ A satisfies mā = 0̄ for some natural number m. Then ma ∈ T.
So there is some natural number k such that kma = 0. This means
that a ∈ T and ā = 0̄. The claim is proved.
Choose a set of generators x1 , . . . , xn of K so that n reaches the
minimum value.
Assume that
m1 x1 + · · · + mn xn = 0

for some m1 , . . . , mn ∈ Z. Let d = gcd(m1 , . . . , mn ), mi = dki . Then

d(k1 x1 + · · · + kn xn ) = 0.

Since the only element of finite order in K is 0,

k1 x1 + · · · + kn xn = 0.

Since gcd(k1 , . . . , kn ) = 1 there is an integral n × n matrix A = (aij )


with determinant ±1 whose first row is (k1 , . . . , kn ).
Let yi = ai1 x1 + · · · + ain xn (2 ≤ i ≤ n). Then
 
  0
x1
   y2 
 
A  ...  =  ..  .

.
xn
 
yn

Hence  
  0
x1
 y2 
 
 ..  −1
 . =A   ..  ,

.
xn
yn
5.3. * STRUCTURE OF FINITELY GENERATED ABELIAN GROUPS115

which implies that K is generated by y2 , . . . , yn , contradicting the min-


imality of n. Hence K is a free abelian group with basis x1 , . . . , xn .
Choose zi ∈ A such that the image of in K = A/T is xi for 1 ≤ i ≤
n.
For an arbitrary element a ∈ A, there is a unique set of integers
k1 , . . . , kn such that

ā = k1 x1 + · · · + kn xn .

This means that a − (k1 z1 + · · · + kn zn ) ∈ T. Hence

A∼
= Zn ⊕ T.

In particular, T ∼
= A/Zn . Hence T is finitely generated. It follows from
Lemma 5.3.8 and Corollary 5.3.6 that T has the required decomposi-
tion.
The uniqueness is obvious.

Remark 5.3.10. An element of finite order in a abelian group is


called a torsion element. The part T in the above theorem is called
the torsion part of A and n is called the rank of A.

Exercises
1. Show that the additive group Q of rational numbers is not a
finitely generated abelian group.
2. Determine the structure of the abelian group A = Z2 /Z(6 ⊕ 30).
3. Determine all abelian groups of order 80.
4. Assume that an abelian group A is isomorphic to the direct sum
of three cyclic groups of orders 24, 6, 2 respectively. Determine the
structures of Sylow 2-subgroup and Sylow 3-subgroup of A.
5. Let

A = {(a1 , a2 ) ∈ Q2 |4a1 b1 + 2a1 b2 + 2a2 b1 + 4a2 b2 ∈ Z ∀(b1 , b2 ) ∈ Z2 }.

Determine the structure of the quotient group A/Z2 .


116 CHAPTER 5. TOPICS IN GROUP THEORY

6. Assume that a finite group G satisfies the following property:


For any pair of subgroups H, K either H ⊆ K or K ⊆ H. Show that
G is a cyclic group of order pn where p is a prime number.

7. Determine all finitely generated abelian group F satisfying |Aut(G)| <


∞.

5.4 Solvable Groups

One cannot expect a structure theorem of non-abelian groups as simple


as that of finitely generated abelian groups.

Definition 5.4.1. Let G be a group. A chain

{e} = G0 / G1 / · · · / Gr = G

is called a subnormal series of G. The quotient groups Gi /Gi−1 (1 ≤


i ≤ r) are called the factor groups of that series. The subnormal series
is called a composition series if every quotient factor is simple.

In a subnormal series, Gi−1 is a normal subgroup of Gi for 1 ≤ i ≤ n,


but Gi is not necessarily a normal subgroup of Gj if j − i > 1.

Example 5.4.2. The set of eight quaternions G = {±1, ±i, ±j, ±k}
is a non-abelian group under multiplication. Set G0 = {1}, G1 =
{±1}, G2 = {±1, ±i}, G3 = G. Then G0 / G1 / G2 / G3 is a compo-
sition series.
Set G00 = {1}, G01 = {±1}, G02 = {±1, ±j}, G03 = G. Then G00 / G01 /
G02 / G03 is another composition series.

Lemma 5.4.3. Let

{e} = G0 / G1 / · · · / Gr = G

be a composition series of a group G. Then Gi−1 is a maximal normal


subgroup of Gi , which means that Gi−1 6= Gi and there is no proper
normal subgroup H of Gi such that Gi−1 ⊂ H ⊂ Gi .
5.4. SOLVABLE GROUPS 117

Proof. If H / Gi and Gi−1 ⊆ H, then H/Gi−1 / Gi /Gi−1 . Since


Gi /Gi−1 is a simple group, H/Gi−1 is a trivial subgroup of Gi /Gi−1 ,
which means that H = Gi−1 or H = Gi .

Remark 5.4.4. A group may have more than one maximal normal
subgroup.

Proposition 5.4.5. Every finite group has a composition series.

Proof. The induction on the order of the group works.

Remark 5.4.6. This proposition is not valid for infinite groups.


For example, Z does not have composition series, since every nontrivial
subgroup of Z is isomorphic to Z.

Theorem 5.4.7 (Jordan-Hölder). Let

{e} = H0 / H1 / · · · / Hs = G

and
{e} = G0 / G1 / · · · / Gr = G

be two composition series of a finite group G. Then r = s and there


exists a permutation σ of {1, . . . , r} such that

Gi /Gi−1 ∼
= Hσ(i) /Hσ(i)−1

for 1 ≤ i ≤ r.

Proof. Apply induction on |G|. When |G| = 1 the theorem holds


obviously. Assume that |G| > 1. Then r ≥ 1, s ≥ 1. If Gr−1 ∼ = Hs−1
then the induction hypothesis implies the required result. Assume that
Gr−1 6= Hs−1 .
Since Gr−1 Hs−1 / G and Gr−1 is a maximal normal subgroup of G,
Gr−1 Hs−1 = G.
Let K = Gr−1 ∩ Hs−1 . Theorem 1.7.14 implies

Gr−1 /K ∼
= G/Hs−1 , Hs−1 /K ∼
= G/Gr−1 .
118 CHAPTER 5. TOPICS IN GROUP THEORY

Choose an arbitrary composition series

{e} = K0 / K1 / · · · / Km = K

of K. The theorem is true by induction hypothesis and by compareing


the following four composition series of G :

{e} = H0 / H1 / · · · / Hs−2 / Hs−1 / G,

{e} = G0 / G1 / · · · / Gr−2 / Gr−1 / G,

{e} = K0 / K1 / · · · / Km / Hs−1 / G,

{e} = K0 / K1 / · · · / Km / Gr−1 / G.

Definition 5.4.8. If a group G has a subnormal series whose all


factor groups are abelian groups, then G is called a solvable group.

By convention, the trivial group {1} is solvable.

Proposition 5.4.9. Every factor group of a composition series of


a finite solvable group is a cyclic group of prime order.

Proof. This is because a finite abelian group is simple if and only


if it is a cyclic group of prime order.

Proposition 5.4.10. Any subgroup or quotient group of a solvable


group is solvable.

Proof. Let H be subgroup of a solvable group G. Let

{e} = G0 / G1 / · · · / Gr = G

be a subnormal series such that all Gi /Gi−1 are abelian. Then

{e} = G0 ∩ H / G1 ∩ H / · · · / Gr ∩ H = H

is a subnormal series. Since Gi ∩H/Gi−1 ∩H is isomorphic to a subgroup


of Gi /Gi−1 , it is abelian. Hence H is solvable.
5.4. SOLVABLE GROUPS 119

Let G/N be a quotient group of G. Then

{e} = G0 /G0 ∩ N / G1 /G1 ∩ N / · · · / Gr /Gr ∩ N = G/N

is a subnormal series of G/N. Since (Gi /Gi ∩ N )/(Gi−1 /Gi−1 ∩ N ) is a


quotient group of Gi /Gi−1 for every i, G/N is solvable.

Theorem 5.4.11. The symmetric group Sn is solvable if and only


if n ≤ 4.

Proof. S1 and S2 are obviously solvable.


Since S3 has a normal subgroup of order 3, S3 is solvable.
As for S4 , let G0 = {e}, G1 = {1, (12)(34), (13)(24), (14)(23)}, G2 =
A4 , G3 = S4 . Then
G 0 / G1 / G2 / G3

is a subnormal series all whose factor groups are abelian.


When n ≥ 5 the alternating group An is simple by Theorem 1.6.5.
Hence
{e} / An / Sn

is a composition series and the factor group An is not abelian.

Exercises
1. Give a complete proof of Proposition 5.4.5.
2. Let F be a field an let G be the group of all 2 × 2 invertible
upper triangular matrices over F. Show that G is a solvable group.
Chapter 6

Field Extensions

6.1 Definitions and First Properties of Field Ex-


tensions

Definition 6.1.1. Let F be a subfield of a field E. Then E is an


extension of F. This relation is denoted by E/F. We may also say that
E/F is a field extension (or simply extension if the context makes its
meaning clear). A subfield E 0 of E containing F is called an interme-
diate field of E/F.

Let E/F be a field extension. Under the addition and multiplication


of E, E is a vector space over F, whose dimension is denoted by [E : F ],
called the degree of the extension E/F. If [E : F ] = ∞, then E/F
is called an infinite extension, otherwise a finite extension. A finite
extension E/F has degree one if and only if E = F.

Proposition 6.1.2. Let E/F, L/E be field extensions (or equiva-


lently, E is an intermediate field of L/F ). Then

[L : F ] = [L : E][E : F ].

Proof. First assume that E/F, L/E are finite extensions. Let
a1 , . . . , an be a basis of the F -space E and let b1 , . . . , bm be a basis
of the E-space L. It suffices to show that {ai bj }1≤i≤n,1≤j≤m is a basis
of the F -space L.

120
6.1. DEFINITIONS AND FIRST PROPERTIES OF FIELD EXTENSIONS121

Assume that there are cij ∈ F, (1 ≤ i ≤ n, 1 ≤ j ≤ m) such that


X
cij ai bj = 0.
1≤i≤n,1≤j≤m

Then !
m
X n
X
cij ai bj = 0.
j=1 i=1

Since b1 , . . . , bm are linearly independent over E,

n
X
cij ai = 0
i=1

holds for 1 ≤ j ≤ m. Hence cij = 0 for 1 ≤ i ≤ n, 1 ≤ j ≤ m


since a1 , . . . , an are linearly independent over F. Hence the mn elements
{ai bj }1≤i≤n,1≤j≤m are linearly independent over F.

Let u be an arbitrary element in L. Then u can be expressed as


m
X
u= dj bj ,
j=1

in which dj ∈ E. Express dj as
n
X
dj = eij ai ,
i=1

in which eij ∈ F. Then


m X
X n
u= eij ai bj ,
j=1 i=1

in other words, u is a linear combination of the elements {ai bj }1≤i≤n,1≤j≤m .


Hence {ai bj }1≤i≤n,1≤j≤m is a basis of L over F, which implies that
[L : F ] = mn = [L : E][E : F ].

If [E : F ] = ∞ or [L : E] = ∞ Then there are arbitrarily many


elements in L that are linearly independent over F, which means that
[L : F ] = ∞. Hence [L : F ] = [L : E][E : F ] holds in this case too.
122 CHAPTER 6. FIELD EXTENSIONS

The following lemma is obvious.


Lemma 6.1.3. Let {Kλ }λ∈Λ be a collection of intermediate fields of
T
E/F in which Λ 6= ∅. Then λ∈Λ Kλ is an intermediate field of E/F.
Proposition 6.1.4. Let E/F be a field extension and let S be a
subset of E. There exists an intermediate field F (S) of E/F such that
1) S ⊆ F (S);
2) F (S) ⊆ K for any intermediate field K of E/F containing S.
Such an intermediate field F (S) is uniquely determined by S.
Proof. Let Γ be the set of all intermediate field of E/F containing
S. Since E ∈ Γ, the set Γ is not empty. Let F (S) = E 0 ∈Γ E 0 . Then
T

F (S) clearly satisfies the two conditions.


Assume that L is a field such that S ⊆ L and L ⊆ K for any
intermediate field K of E/F containing S. Then F (S) ⊆ L since F (S)
is an intermediate field of E/F containing S. For the same reason, L ⊆
F (S). Therefore L = F (S). This proves the uniqueness of F (S).

Remark 6.1.5. The field F (S) is understood to be the smallest


intermediate of E/F containing S, called the subfield of E generated
by the set S over F.
Another presentation of F (S) is

 f (a1 , . . . , an )
F (S) = n ≥ 0, f, g ∈ F [x1 , . . . , xn ],
g(a1 , . . . , an )

a1 , . . . , an ∈ S, g(a1 , . . . , an ) 6= 0 .

Definition 6.1.6. Let E/F be a field extension and let S be a


subset of E. If E = F (S), then we say that the extension E/F is
generated by S. If there is a finite subset S = {t1 , . . . , tn } of E such
that E = F (S), then E/F is called a finitely generated extension,
denoted by E = F (t1 , . . . , tn ). In particular if E/F is generated by a
single element α ∈ E, i.e., E = F (α), then E/F is called a simple
extension.

Remark 6.1.7.
 f (α)
F (α) = f, g ∈ F [x], g(α) 6= 0 .
g(α)
6.2. ALGEBRAIC EXTENSIONS 123

Remark 6.1.8. Let E/F be a field extension, t1 , . . . , tn ∈ E. Denote

F [t1 , . . . , tn ] = {f (t1 , . . . , tn )|f ∈ F [x1 , . . . , xn ]}.

It is a subring of E containing F and the elements t1 , . . . , tn , but not


necessarily a subfield.
For example in the extension R/Q, the ring

Q[π] = {a0 + a1 π + · · · + an π n |a0 , a1 , . . . , an ∈ Q}

is different from Q(π), since the element 1/π is in Q(π) but not in Q[π].

6.2 Algebraic Extensions

Let E/F be a field extension and let α ∈ E. There is a ring homomor-


phism
φ : F [x] → E, f (x) 7→ f (α).

If Ker(φ) = 0, then α is called a transcendental element over F. In


other words, a transcendental element is not a zero of any nonzero
polynomial with coefficients in F.
Any non-transcendental element in a field extension is called an
algebraic element (over F ). In other words, an algebraic element is the
zero of some nonzero polynomial over F. In particular, every element
in F is algebraic.
Let α ∈ E be an algebraic element over F. Then Ker(φ) 6= 0, i.e.,
Ker(φ) is a nonzero ideal of the polynomial ring F [x]. Since F [x] is a
principal ideal domain, the ideal Ker(φ) is generated by some monic
polynomial f (x). If follows from Ker(φ) 6= F [x] that deg(f ) ≥ 1.
We claim that f (x) is irreducible. Otherwise f (x) = g(x)h(x) with
deg(g) < deg(f ), deg(h) < deg(f ). Then g(x) ∈ / Ker(φ), h(x) ∈
/ Ker(φ).
Let ḡ and h̄ denote the images of g(x) and h(x) in E respectively. Then
ḡ 6= 0, h̄ 6= 0. However, ḡ h̄ = φ(gh) = φ(f ) = 0. This would contradict
the fact that E is a field. Hence f (x) is an irreducible polynomial
over F and so Ker(φ) is a maximal ideal of F [x]. By Proposition 2.4.5
Im(φ) ∼ = F [x]/(f (x)) is a field, which is nothing but F (α).
124 CHAPTER 6. FIELD EXTENSIONS

The polynomial f (x) above is called the minimal polynomial of


the algebraic element α, whose degree is defined to be the degree of the
element α over the base field F.
Proposition 6.2.1. Let f (x) be the minimal polynomial of an al-
gebraic element in a field extension E/F and let n = deg(f ). Then
[F (α) : F ] = n and 1, α, α2 , . . . , αn−1 form a basis the the F -space
F (α).
Proof. It is clear that the map

φ : F [x] → F (α), g(x) 7→ g(α)

is an epimorphism. Hence every element β ∈ F (α) can be written as


β = g(α) for some g(x) ∈ F [x]. Let

g(x) = q(x)f (x) + r(x),

where q(x), r(x) ∈ F [x] and deg(r) < n. Then β = g(α) = r(α). This
shows that β is a linear combination of 1, α, α2 , . . . , αn−1 over F.
It remains to show that 1, α, α2 , . . . , αn−1 are linear independent
over F. Suppose that there are not all zero elements c0 , c1 , . . . , cn−1 ∈ F
such that
c0 + c1 α + c2 α2 + · · · + cn−1 αn−1 = 0.

Let g(x) = c0 +c1 x+c2 x2 +· · ·+cn−1 xn−1 . Then g(x) is a nonzero poly-
nomial over F such that g(α) = 0 and deg(g) < n, which contradicts
the assumption that f (x) is the minimal polynomial of α.

Let α ∈ E and let F [α] = {f (α)|f (x) ∈ F [x]}. Then F [α] is a


subring of E.
The proposition above shows that F [α] = F (α) when α is an alge-
braic element.
When α is transcendental, F [α] is isomorphic to the polynomial
ring F [x], which is not a field. Hence F [α] = F (α) if and only if α is
algebraic.
Definition 6.2.2. A field extension E/F is called an algebraic
extension if every element in E is algebraic over F.
6.2. ALGEBRAIC EXTENSIONS 125

Lemma 6.2.3. Let K be an intermediate field of an extension E/F


and let α ∈ E be an algebraic element over F. Then α is also an
algebraic element over K.

Proof. By definition there is a nonzero polynomial f (x) ∈ F [x]


such that f (α) = 0. Since F ⊆ K, f (x) ∈ K[x]. Hence α is also an
algebraic element over K.

Theorem 6.2.4. A field extension E/F is a finite extension if and


only if it is a finitely generated algebraic extension.

Proof. Assume that E/F is a finitely generated algebraic ex-


tension, i.e., there are algebraic elements α1 , . . . , αn such that E =
F (α1 , . . . , αn ). Denote Ei = F (α1 , . . . , αi ) for 1 ≤ i ≤ n. Lemma
6.2.3 implies that αi is algebriac over Ei−1 for each i > 1. Hence
[Ei : Ei−1 ] < ∞. It follows that

[E : F ] = [En : En−1 ] · · · [E2 : E1 ][E1 : F ] < ∞.

Next assume that [E : F ] = d < ∞. For any α ∈ E the d + 1


elements
1, α, α2 , . . . , αd

are linearly dependent over F, i.e., there are elements

c0 , c1 , . . . , cd ∈ F,

not all zero, such that

c0 + c1 α + c2 α2 + · · · + cd αd = 0.

Hence α is algebraic. This proves that E/F is an algebraic exten-


sion. Choose a basis b1 , . . . , bd of the F -space E. It is clear that
E = F (b1 , . . . , bd ). Hence E/F is finitely generated.

Proposition 6.2.5. Assume that E/F, L/E are algebraic exten-


sions. Then L/F is an algebraic extension.

Proof. We need to show that every element γ ∈ L is an algebraic


126 CHAPTER 6. FIELD EXTENSIONS

element over F. For this purpose, it suffices to show that [F (γ) : F ] <
∞ by virtue of Theorem 6.2.4.
Since γ is algebraic over E, there is a nonzero f (x) ∈ E[x] such that
f (γ) = 0. Let

f (x) = bn xn + bn−1 xn−1 + · · · + b1 x + b0 ,

in which b0 , b1 , . . . bn ∈ E. By hypothesis, all coefficients b0 , b1 , . . . , bn


are algebraic elements over F. Hence F (b0 , b1 , . . . , bn ) is a finite exten-
sion of F by Theorem 6.2.4.
Since f (γ) = 0, γ is algebraic over F (b0 , b1 , . . . , bn ). Hence

[F (b0 , b1 , . . . , bn , γ) : F (b0 , b1 , . . . , bn )] < ∞,

which implies that


[F (b0 , b1 , . . . , bn , γ) : F ]

= [F (b0 , b1 , . . . , bn , γ) : F (b0 , b1 , . . . , bn )][F (b0 , b1 , . . . , bn ) : F ] < ∞.

Therefore γ is algebraic over F.

Let E/F, K/F be two extensions of a field F and let f be a ho-


momorphism from E to K such that f (a) = a for all a ∈ F. Then
f is called a homomorphism from E/F to K/F, also referred to as a
homomorphism from E to K over F.

Proposition 6.2.6. Let E/F be an algebraic extension and let f


be a homomorphism from E/F to E/F itself. Then f : E → E is an
isomorphism.

Proof. The homomorphism is injective by Proposition 2.4.2. Hence


it suffices to show that f is surjective.
Let α ∈ E, By assumption α is algebraic over F. Let p(x) be the
minimal polynomial of α and let α = α1 , α2 , . . . , αn be all zeros of
p(x) in E. Let K = F (α1 , α2 , . . . , αn ). Then K is a finitely generated
algebraic extension of F. Theorem 6.2.4 implies that K/F is a finite
extension. Thus K is an F -space of finite dimension.
6.2. ALGEBRAIC EXTENSIONS 127

Since f is a homomorphism,

p(f (αi )) = f (p(αi )) = f (0) = 0

for all 1 ≤ i ≤ n. Hence f (αi ) ∈ K. This shows that f (K) ⊆ K. Since


any injective linear transformation of a vector space of finite dimension
is surjective, f : K → K is surjective. Hence there is some β ∈ K such
that f (β) = α). Therefore f : E → E is surjective.

Example 6.2.7. The condition of algebraic extension in the pro-


postion is indispensable. For example,

a(x) a(x2 )
f : F (x) → F (x), 7→
b(x) b(x2 )

is a monomorphism over F, but not an epimorphism.

Exercises
√ √
1. Show that u = 2 + 3 is an algebraic number, i.e., there is
f (x) ∈ Q[x] such that f (u) = 0. Find the ideal I of Q[x] such that
Q[u] ∼= Q[x]/I.

2. Let α is a complex root of the equation x3 − 3x + 4 = 0. Express


the multiplicative inverse of α2 + α + 1 in Q[α] in the form a + bα + cα2
in which a, b, c are rational numbers.

3. Let F (α) be a simple extension of a field F and let β ∈ F (α)\F.


Show that α is an algebraic element over F (β).

4. Let E/F be a field extension and α ∈ E such that [F (α) : F ] = 5,


Show that F (α2 ) = F (α).

5. Let E/F be a finite extension and let K be an intermediate


field. Let α ∈ E and let p(x) be the minimal polynomial of α over F.
Assume that deg(p) is relative prime to [K : F ]. Show that p(x) is an
irreducible polynomial in K[x].
128 CHAPTER 6. FIELD EXTENSIONS

6.3 Constructions of Field Extensions

Thus far we have discussed the properties of the intermediate fields of


a given extension E/F. To construct an intermediate field it suffices to
give a set of generators.
Now we may ask how to construct an extension E/F from a given
F without any knowledge of E. For example, we may ask what kind of
fields can be extensions of the finite field Fp . To answer these questions
we need to learn a new method.
Let F be a field. We consider to add a new element and the addition
and multiplication of this new element with the elements in F. Of course
a lot of other elements must also be thrown in to form a field.
For example, starting from the field Q of rational number, if we

want to add 2 to obtain a larger field we must add all elements in

the form a + b 2, (a, b ∈ Q).
This example is easy, because the fields of real numbers and com-

plex numbers are available and the arithmetic involving 2 is well-
known. This construction is essentially the construction of an interme-
diate field.
For an abstract field F, there is no larger field such like complex
number field available. This makes the construction of extensions of F
more difficult.
As the first step, we wish to construct a simple extension E/F, say
E = F (α) such that the element α satisfies some preassigned condi-
tions.
If we wish α to be transcendental over F, then the construction is
easy. Let E = F (x) be the field of all fractions over F in one variable
x. Then E is a simple extension of F generated by a transcendental
element x over F.
Suppose that we wish α to be algebraic over F. If E = F (α) has
been constructed then there will be a surjective homomorphism φ :
F [x] → E, f (x) → f (α), with Ker(φ) = (g(x)) where g(x) is the
minimal polynomial of α over F. Then

E∼
= F [x]/(g(x)).
6.3. CONSTRUCTIONS OF FIELD EXTENSIONS 129

The key observation that g(x) is an irreducible polynomial in F [x]


suggests us to start the construction of E from an arbitrary monic
irreducible polynomial g(x) ∈ F [x], and then define E = F [x]/(g(x)).
Since (g(x)) is a maximal ideal, E is a field by Proposition 2.4.5. Every
element in E can be expressed as f (x), in which f (x) ∈ F [x]. Recall
that f (x) is a coset of the principal ideal I = (g(x)) represented by
f (x). The addition and multiplication in E follow the rules

f1 (x) + f2 (x) = f1 (x) + f2 (x),

f1 (x)f2 (x) = f1 (x)f2 (x).

We are facing another issue: F and E are two different fields because
they contain elements of different nature. In what sense is F regarded
as a subfield of E?
The answer is that E contains a subfield isomorphic to F. This view
of point can hardly be over-emphasized.
Every element in F is also an element in F [x]. Let F 0 = {ā|a ∈ F }.
It is easy to verify that F 0 is a subfield of E.
The map
σ : F → F 0 , a 7→ ā

is obviously an isomorphism.

Example 6.3.1. x2 + x + 1 is an irreducible polynomial over the


finite field F2 .

Proof. Every reducible polynomial of degree two can be decom-


posed into the product of two polynomials of degree one. There are
only two linear polynomials x and x + 1 over F2 . Hence there are only
three reducible polynomials x2 , x2 + x, x2 + 1 of degree two. Hence
x2 + x + 1 is irreducible over F2 .

It follows that E = F2 [x]/(x2 +x+1) is a simple algebraic extension


of F2 , with x2 + x + 1 being the minimal polynomial of its generator x̄.
Hence E/F is an extension of degree 2, which implies that E is a field
containing four elements. Denote α = x̄. Then the four elements of E
130 CHAPTER 6. FIELD EXTENSIONS

are listed as 0, 1, α, α + 1. The addition and multiplication tables in E


are

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

and

× 0 1 α α+1
0 0 0 0 0
1 0 1 α α+1
α 0 α α+1 1
α+1 0 α+1 1 α

respectively.
By the same method it can be verified that x3 +x+1 is an irreducible
polynomial over F2 . So E = F2 [x]/(x3 + x + 1) is a field containing 8
elements. Let α = x̄. The eight elements of E are

0, 1, α, α + 1, α2 , α2 + 1, α2 + 1, α2 + α + 1.

The product (α2 + 1)(α2 + α + 1) can be computed in the following


way. Let

f (x) = (x2 + 1)(x2 + x + 1) = x4 + x3 + x + 1.

Use division algorithm to compute the remainder of f (x) divided by


x3 + x + 1 :
f (x) = (x + 1)(x3 + x + 1) + x2 + x.

Thus
(α2 + 1)(α2 + α + 1) = f (α) = α2 + α.

Exercises
6.4. ALGEBRAICALLY CLOSED FIELD 131

1. Let E/F be an algebraic extension and let f (x) ∈ F [x] be an


irreducible polynomial over F. Assume that α, β ∈ E satisfy f (α) =
f (β) = 0. Show that there exists a homomorphism φ : F [α] → E such
that
1) φ(a) = a for all a ∈ F ;
2) φ(α) = β.
Show that φ induces an isomorphism from F [α] to F [β].
2. Show that x2 + x + 2 is an irreducible polynomial over F3 . Let
E = F3 [x]/(x2 + x + 2). Write out the multiplication table of E.

6.4 Algebraically Closed Field

Let F be a field. According to the previous section, in order to con-


struct a simple algebraic extension F (α)/F of degree d it is enough to
find an irreducible polynomial of degree d over the field F. This exten-
sion F (α)/F is called trivial if F (α) = F, which is equivalent to d = 1.
Thus the field F has a nontrivial algebraic extension if and only if there
is an irreducible polynomial of degree greater than one over F. In other
words, the following three statements are equivalent.
• The only algebraic extension of F is F/F.
• A polynomial f (x) ∈ F [x] is irreducible if and only if deg(f ) = 1.
• Every polynomial of positive degree over F has a zero in F.
A field satisfying the above equivalent conditions is called an alge-
braically closed field.
Our most familiar algebraically closed field is the complex number
field due to the Fundamental Theorem of Algebra.

Definition 6.4.1. Let L/F be an algebraic extension. If L is al-


gebraically closed then L is called the algebraic closure of F.

Proposition 6.4.2. Let E/F be an extension. Let E 0 be the set of


all algebraic elements (over F ) in E. Then
1) E 0 is an intermediate field of E/F, refered to as the algebraic
closure of F in E.
2) If E is algebraically closed then E 0 is the algebraic closure of F.
132 CHAPTER 6. FIELD EXTENSIONS

Proof. 1) Let α, β ∈ E 0 . Then [F (α, β) : F ] < ∞, which implies


that F (α, β)/F is an algebraic extension. Hence F (α, β) ⊆ E 0 . So
α + β ∈ E 0 , α − β ∈ E 0 , αβ ∈ E 0 , and α/β ∈ E 0 if β 6= 0. This
shows that E 0 is a subfield of E. Obviously F ⊆ E 0 . Hence E 0 is an
intermediate field of E/F.
2) Let f (x) be a polynomial of positive degree over E 0 . Since E
is algebraically closed by assumption, there is some α ∈ E such that
f (α) = 0. This implies that E 0 (α)/E 0 is an algebraic extension. Hence
E 0 (α)/F is an algebraic extension, which implies that α is algebraic
over F, i.e., α ∈ E 0 . This shows that E 0 is algebraically closed. It is
evident that E 0 /F is an algebraic extension. Hence E 0 is the algebraic
closure of F by definition.

The condition of E being algebraically closed in 2) of Proposition


6.4.2 is indispensable. For example the algebraic closure of Q in R is
not an algebraically closed field.
For any subfield F of the complex number field C, the algebraic
closure of F in C is algebraically closed by Proposition 6.4.2. A typical
example is the algebraic closure of Q in C, commonly denoted by Q̄.
The numbers in Q̄ are called algebraic numbers while the other number
are transcendental numbers.
Now a fundamental problem we are facing is: For an arbitrary field
F, does its algebraic closure exist? If exists, is it unique?
Unlike the cases of number fields, there is no “universal” field for F
to play the role of the complex number field. The fields such like finite
fields or fields of rational functions are not contained in the complex
number field. A new method to resolve this problem is required.
We may try to solve the existence problem naively. If F is already
algebraically closed, then F itself is the algebraic closure of F. Other-
wise there exists a nontrivial simple algebraic extension F (α)/F using
the approach in the previous section. Then treat the new field F (α) in
the same way. If the same process is repeated again and again, prob-
ably after infinitely many times, an algebraically closure of F will be
obtained. This argument looks like induction. But this is not a solid
proof since the term “infinitely many times” is ambiguous.
6.4. ALGEBRAICALLY CLOSED FIELD 133

Similar issues of this nature arise in many fundamental definitions


and theorems in mathematics. It forces us to take a few axioms as
granted, upon which the normal deduction in mathematics is based.
Here we need to use the celebrated Zorn’s lemma.

Definition 6.4.3. Let ≺ be a relation in a set S, i.e, a ≺ b holds


for some pairs a, b of elements in S. If the following conditions
1) a ≺ a for any a ∈ S;
2) a ≺ b, b ≺ a implies a = b;
3) a ≺ b, b ≺ c, implies a ≺ c,
then S is called a partially-ordered set under the partial order ≺ .
It is denoted by (S, ≺), or simply S if no confusion arises.

Remark 6.4.4. A partially-ordered set (S, ≺) is a totally-ordered


set if it satisfies the following additional condition
4) For any pair a, b of elements in S at least one of a ≺ b and b ≺ a
holds.

Example 6.4.5. • Any set of real numbers is a totally-ordered set


under the relation “≤”.
• For any set S let Ω be the set of all subsets of S. For any U, V ∈ Ω,
define U ≺ V if and only if U ⊆ V. Then Ω is a partially-ordered set
under “≺”, but not a totally-ordered set if S contains more than one
element.

Definition 6.4.6. Let (S, ≺) be a partially-ordered set and let T


be a subset of S. An element b ∈ S is called an upper bound of T if
a ≺ b for any a ∈ T.
An element b in S is called a maximal element of S if b ≺ a ∈ S
implies b = a.

It is possible that a partially-ordered set has no maximal element


or has more than one maximal elements.
Zorn’s Lemma
Let (S, ≺) be a non-empty partially-ordered set. If every totally-
ordered subset of S has an upper bound in S then there exists a max-
imal element in S.
134 CHAPTER 6. FIELD EXTENSIONS

The following example is a typical application of Zorn’s lemma in


algebra.

Example 6.4.7. Let I be an ideal of a commutative ring A such


that I 6= A. Then there exists a maximal ideal m of A such that I ⊆ m.

Proof. Let Γ be the set of all ideals of I containing I but not


equal to A. Since I ∈ Γ, Γ is not empty.
For any J1 , J2 ∈ Γ, define J1 ≺ J2 if and only if J1 ⊆ J2 . Then
(Γ, ≺) is a partially-ordered set.
Let T be a totally-ordered subset of Γ. Since 1 ∈
/ J for every J ∈ T,
S S S
J∈T J 6= A. Hence J∈T J ∈ Γ and J ≺ J∈T J for every J ∈ T,
which means that the condition for Zorn’ lemma is satisfied. Hence
there is a maximal element m in Γ.
Let J be an ideal of A such that m ⊆ J and J 6= m. The maximality
of m implies that J ∈ / Γ. This forces J = A. Hence m is a maximal
ideal of A.

Theorem 6.4.8. Let σ : E → L be a homomorphism of fields with


L algebraically closed. Let K/E be an algebraic extension. Then there
exist homomorphism τ : K → L such that τ |E = σ.

Proof. Let Γ = {(H, γ)| H is an intermediate field of K/E, γ :


H → L is a homomorphism such that γ|E = σ.}. Define a partial order
in Γ by (H1 , γ1 ) ≺ (H2 , γ2 ) if and only if H1 ⊆ H2 and γ2 |H1 = γ1 .
S
Let T be a totally-ordered subset of Γ and let B = (H,γ)∈T H.
Then B is an intermediate field of K/E. Define a map β : B → L by
the following way: for an arbitrary x ∈ B, choose (H, γ) ∈ T such that
x ∈ H, let β(x) = γ(x). It is easy to verify that β(x) does not depend
upon the choice of (H, γ) and β is a homomorphism from B to L such
that β|H = γ for any (H, γ) ∈ T. This means that (B, β) ∈ Γ is an
upper bound of T.
By Zorn’s lemma, there is a maximal element (A, η) in Γ. It suffices
to show that A = K. Suppose that it were not the case, there would be
some element α ∈ K\A. Let f (x) ∈ A[x] be the minimal polynomial
of α over A. There is an isomorphism q : A[α] → A[x]/(f (x)). Identify
A and η(A) via η, so that f (x) is regarded as a polynomial over η(A).
6.4. ALGEBRAICALLY CLOSED FIELD 135

Since L is assumed to be algebraically closed, f (x) has a zero a in L.


Thus there is an homomorphism p : A[x]/(f (x)) → L such that p(x̄) =
a. Let τ = p ◦ q : A[α] → L. Then (A[α], τ ) ∈ Γ and (A, η) ≺ (A[α], τ ).
This contradicts the hypothesis that (A, η) is a maximal element of Γ.
Hence A = K.

Theorem 6.4.9. Let F be a field. Then the following statements


hold.
1) The algebraic closure L of F exists.
2) Let L1 , L2 be two algebraic closures of F. There exists isomor-
phism f : L1 → L2 such that f (a) = a for every a ∈ F.
Proof. 1) Let Ω be the set of all algebraic extensions of F. For
any E/F and K/F in Ω, define E/F ≺ K/F if E ⊆ K. This set Ω
becomes a partially-ordered set under the relation “≺”.
Let ∆ be a totally-ordered subset of Ω. In order to give an upper
bound of ∆, it is natural to consider the union of all members of ∆
as done in proofs of Example 6.4.7 and Theorem 6.4.8. But a new
difficulty pops up. The members of ∆ are not necessarily contained in
a common large set, so it does not make sense to take the union. We
need to be careful to construct an upper bound of ∆.
For any E/F ≺ K/F let jE,K be the injection of E into K. There
exists a set A satisfying the following two conditions:
1) For any E/F ∈ ∆ there is an injective map σE : E → A such
that σE = σK ◦ jE,K for any E/F ≺ K/F in ∆.
2) A = E/F ∈∆ Im(σE ). 1
S

For any x, y ∈ A there exist E/F, K/F ∈ ∆ such that x = σE (a), y =


σK (b) for some a ∈ E, b ∈ K. Since ∆ is totally-ordered, E/F ≺ K/F
or K/F ≺ E/F. Without loss of generality we may assume that E/F ≺
K/F. Define x + y = σK (a + b), xy = σK (ab). By the condition 1) that
A satisfies, x + y and xy do not depend upon the choices of E/F and
K/F. It is easy to verify that the set A under the addition and mul-
tiplication just defined is a field and σE : E → A is a monomorphism
1
The set A can be constructed in the following way: Let B = tE/F ∈∆ E be the disjoint
union of all members of ∆, i.e., B = {(a, E/F )| E/F ∈ ∆, a ∈ E}. Define a relation in B
by (a, E/F ) ∼ (b, K/F ) if and only if E/F ≺ K/F and b = jE,K (a) or K/F ≺ E/F and
b = jK,E (a). This is an equivalence relation. Let A be the set of all equivalence classes.
For any E/F ∈ ∆, define σE : E → A, a 7→ [(a, E/F )]. Then all conditions are satisfied.
136 CHAPTER 6. FIELD EXTENSIONS

for each E/F ∈ ∆. Hence we may identify the elements of E and those
of σE (E). The condition 2) guarantees that every element of A is alge-
braic over F, so A/F ∈ Ω. Since E/F ≺ A/F for each E/F ∈ ∆, A/F
is an upper bound of ∆.
By Zorn’s lemma, there exists a maximal element L/F in Ω. It
remains to show that L is algebraically closed. Let L0 /L be an algebraic
extension. Then L0 /F ∈ Ω and L/F ≺ L0 /F. Since L/F is maximal in
Ω, L = L0 . This proves that L is algebraically closed.
2) Let L1 and L2 be two algebraic closures of F. By Theorem 6.4.8
there is a monomorphism φ : L1 → L2 such that φ(a) = a for all a ∈ F.
For the same reason there is a monomorphism ψ : L2 → L1 such that
ψ(a) = a for all a ∈ F. Since ψ ◦ φ is a homomorphism from L1 /F to
L1 /F, Proposition 6.2.6 implies that ψ ◦ φ is an automorphism. For
the same reason φ ◦ ψ is also an automorphism. Hence φ and ψ are
isomorphisms.

Exercises

1. Let A be a commutative ring and let S be a subset of A such


that 0 ∈
/ S and ab ∈ S for any a, b ∈ S. Use Zorn’s lemma to show the
existence of a prime ideal P of A such that P ∩ S = ∅.
2. Show that Q is not isomorphic to Q(x).
3. Let F = {a ∈ R|a is algebraic over Q}. Show that F is not
isomorphic to Q.
4. Let E/F be an algebraic extension. Show that Ē is also the
algebraic closure of F.

6.5 * Ruler and Compass Construction

A fascinating type of problems in plane geometry is ruler and compass


construction. One can only use a pair of compass and an unmarked
ruler to construct geometric figures with preassigned properties.
Many problems are impossible. A notorious example is the im-
possibility of trisection of an arbitrary angle. In this section the field
6.5. * RULER AND COMPASS CONSTRUCTION 137

extension theory will be used to settle this issue. This is a remarkable


example of the application of algebra.
First of all the construction problem should be converted into a
problem of algebra. Most problems can be reduced to the construction
of one or more points. For example, to construct a line or a circle it is
enough to construct two points on the line or the center of the circle
and a point on the circle. We concentrate on the construction of one
point. The rules of the construction are formulated as follows.
Let S0 be a finite set of points on the plane that contains at least two
points. These points are the initial data of the relevant construction
problem.
Then the set S0 is extended inductively to a finite sequence S1 , S2 , . . . , Sn
of sets of points such that

S0 ⊂ S1 ⊂ · · · ⊂ S n ,

where Si = Si−1 ∪ {Pi } and Pi is a point obtained from either one of


the following three methods:
1) Choose any four distinct points Q1 , Q2 , Q3 , Q4 in Si−1 and draw
a line ` passing Q1 and Q2 by ruler and another line `0 passing Q3 and
Q4 . Take the intersection point of ` and `0 to be Pi .
2) Choose any two distinct points Q1 , Q2 in Si−1 and draw a line
` passing Q1 and Q2 . Choose any three points A, B1 , B2 in Si−1 , not
necessarily distinct. Draw a circle C by compass with A as the center
and the distance from B1 to B2 as the radius. Take an intersection
point (if any) of ` and C to be Pi .
3) Draw two distinct circles by the rule in 2). Take an intersection
point of these two circles to be Pi .
Such a sequence is the process of construction.
A point P is said to be constructible from S0 if there exists a process
of construction
S0 ⊂ S 1 ⊂ · · · ⊂ Sn

such that P ∈ Sn .
Example: Construct a square of a given side length.
This is a typical construction problem. We may assume that the
138 CHAPTER 6. FIELD EXTENSIONS

given length is 1 and one side lies on the x-axis from 0 to 1. Then S0 =
{(0, 0), (1, 0)}. The object is to construct two points (0, 1) and (1, 1).
We may set P = (0, 1), since the other point (1, 1) can be constructed
in the similar way.
Draw a circle C with (0, 0) as its center and the distance from (0, 0)
to (1, 0) as its radius. It intersects the x-axis ( the line passing (0, 0)
and (1, 0)) at a new point P1 = (−1, 0). Let S1 = {(0, 0), (1, 0), P1 }.
Let d be the distance from (−1, 0) to (1, 0). Draw two circles of radius
d with centers at (1, 0) and (−1, 0) respectively. Take an intersection
point P2 of these two circles. Let S2 = {(0, 0), (1, 0), P1 , P2 }. Draw a
line L passing (0, 0) and P2 . Take the intersection point P3 of L and
the circle C in the upper half plane. Let S3 = {(0, 0), (1, 0), P1 , P2 , P3 }.
Then P = P3 . This means that P is a constructible point.
Here are some more examples.
• the construction of a regular polygon of n sides : S0 = {(0, 0), (1, 0)}, P =
(cos(2π/n), sin(2π/n)).
• the trisection of an arbitrary angle: S0 = {(0, 0), (1, 0), (cos α, sin α),
(cos α, − sin α)} and P = (cos(α/3), sin(α/3)).
• the construction of a square having the same area as a given circle:

S0 = {(0, 0), (1, 0)} , P = ( π, 0).
• the construction of the edge of a cube having twice the volume of

a given cube: S0 = {(0, 0), (1, 0)} and P = ( 3 2, 0).

Assume that S0 = {(x1 , y1 ), (x2 , y2 ), . . . , (xm , ym )}. Let F0 be the


field Q(x1 , y1 , x2 , y2 , . . . , xm , ym ). It is a subfield of R.

Theorem 6.5.1. A point P = (x, y) on the plane is constructible


from the given set S0 = {(x1 , y1 ), . . . , (xm , ym )} only if [F0 (x, y) : F0 ]
is a power of 2.

Proof. Let S0 ⊂ S1 ⊂ · · · ⊂ Sm be a process of construction


such that Si = Si−1 ∪ {Pi } for 1 ≤ i ≤ m, where Pi = (ui , vi ). Let
Fi = F0 (u1 , v1 , . . . , ui , vi ). Then we obtain a chain of field extensions

F0 ⊆ F1 ⊆ · · · ⊆ Fm .

If P is constructible, then both x and y are in the field Fm , which


6.5. * RULER AND COMPASS CONSTRUCTION 139

implies that F0 (x, y) is an intermediate field of the extension Fm /F0 .


Hence it suffices to show that [Fm : F0 ] is a power of 2. For this purpose
we need to show that [Fi : Fi−1 ] is a power of 2 for 1 ≤ i ≤ m. This
can be achieved by treating three separate cases according to how Pi
is constructed.
Case 1) Pi is the intersection of two lines.
In this case (ui , vi ) is the solution of a system

ax + by = c,

dx + ey = f

where a, b, c, d, e, f ∈ Fi−1 such that ae − bd 6= 0. Hence ui , vi ∈ Fi−1 ,


which implies Fi = Fi−1 .
Case 2) Pi is the intersection of a circle and a line.
Assume that the equations of the line and circle are ax + by = c
and
(x − e1 )2 + (y − e2 )2 = r

where a, b, c, e1 , e2 , r ∈ Fi−1 . Then a 6= 0 or b 6= 0. We may assume that


b 6= 0 without loss of generality. Then vi = (c − aui )/b and
c a 
2
(ui − e1 ) + − e2 − ui = r.
b b
Hence [Fi−1 (ui , vi ) : Fi−1 ] is less than or equal to 2. This implies that
[Fi : Fi−1 ] is a power of 2.
Case 3) Pi is the intersection of two circles. Assume that the equa-
tions of the two circles are

(x − a1 )2 + (y − b1 )2 = c1 (6.1)

and
(x − a2 )2 + (y − b2 )2 = c2 (6.2)

respectively. Subtracting (6.1) from (6.2) yields a linear equation.


Hence this case is reduced to Case 2).
This concludes the proof of the theorem.
140 CHAPTER 6. FIELD EXTENSIONS

Example 6.5.2. • If S0 = {(0, 0), (1, 0)} then the point P = ( 3 2, 0)

is not constructible, since [Q( 3 2) : Q] = 3 is not a power of 2. Hence
it is impossible to double the volume of a cube by ruler and compass.

• If S0 = {(0, 0), (1, 0)} then the point P = ( π, 0) is not con-

structible, since [Q( π) : Q] = ∞ is not a power of 2. Hence it is
impossible to construct a square whose area is equal to that of a given
circle.
• Let S0 = {(0, 0), (1, 0), (cos α, sin α), (cos α, − sin α)} and P =
(cos(α/3), sin(α/3)). Let F0 = Q(cos α, sin α). It follows from
 α α 3
cos + i sin = cos α + i sin α
3 3
that
α α α
cos3 − 3 cos (1 − cos2 ) = cos α
3 3 3
and
α α α
− sin3 + 3 sin (1 − sin2 ) = sin α.
3 3 3
Except for a few special values of α, the polynomial x3 − 3x(1 − x2 ) −
cos α is irreducible. Hence [F0 (cos α3 , sin α3 ) : F0 ] is not a power of
3. Hence the trisection of an arbitrary angle by ruler and compass is
impossible.

Remark 6.5.3. The power of 2 in Theorem 6.5.1 is only a nec-


essary condition of the constructibility. For the sufficient conditions,
interested readers may consult references such like [5].

Exercises

1. Solve the following problems:


1) Let C be a unit circle with its center at (0, 0). Let D be the circle
with (n, 0) as its center and a positive integer n as its radius. Find the
x-coordinate of the intersection points of C and D.
2) Design a method to subdivide a given line segment into three
equal parts.
3) Given two points (0, 0), (1, 0). For any rational number r show
that the point (r, 0) is constructible.
6.5. * RULER AND COMPASS CONSTRUCTION 141

2. Let ζ = cos(2π/5) + i sin(2π/5).



1) Show that ζ + ζ̄ = ( 5 − 1)/2.
2) Construct a regular pentagon by ruler and compass.
Chapter 7

Finite fields

7.1 Basic Theory

Lemma 7.1.1. The number of elements of a finite field is a power


of a prime number.

Proof. Let E be a finite field. Its characteristic must be a prime


number p. Hence E contains a subfield isomorphic to Fp . Let n = [E :
F ]. Then |E| = pn .

Let p be a prime number. Denote by F̄p the algebraic closure of Fp .

Theorem 7.1.2. Let n be an arbitrary natural number. Then there


is a unique subfield E of F̄p containing pn elements.
n
Proof. Let E be the set of all zeros of the polynomial xp − x in
F̄p .
Let a1 , a2 ∈ E. Then
n n n
(a1 − a2 )p − (a1 − a2 ) = ap1 − ap2 − a1 − a2 = 0.
n
(a1 a2 )p − (a1 a2 ) = 0.

If a is a nonzero element in E, then


n n
(1/a)p − 1/a = (1/ap ) − 1/a = 0.

Hence E is a subfield of F̄p .

142
7.1. BASIC THEORY 143

n n
Since the formal derivative of xp − x is −1, the polynomial xp − x
has no multiple zeros by Proposition 3.4.11. Hence E contains exactly
pn elements. The existence of E is proved.
Assume that K is a subfield of F̄p containing pn elements. By
n
example 3.5.8 every element in K is a zero of xp − x. Therefore K =
E.

Corollary 7.1.3. F̄p is an infinite field.

Proposition 7.1.4. Let E, F be subfields of F̄p containing pn and


pm elements respectively. Then F ⊆ E if and only if m|n.
n m
Proof. Since E, F are zeros of xp − x and xp − x in F̄p respec-
tively,
F ⊆E⇔
m n
xp − x|xp − x ⇔
m −1 n −1
xp − 1|xp −1⇔

pm − 1|pn − 1 ⇔

m|n.

Remark 7.1.5. Since every finite extension of Fp can be embedded


into F̄p , the finite field containing q = pn elements is unique up to
isomorphism. This field is usually denoted by Fq .

Exercises

1. How many monic irreducible polynomials of degree 2 are there


over F5 ? of degree 3?
2. Discuss the irreducibility of x4 +1 over R, Q and F16 respectively.
3. Decompose the polynomials x4 − 4 and x3 − 2 into the products
of irreducible polynomials over the fields R, C and F3 respectively.
4. Show that every finite integral domain is a finite field. Give an
example of a finite commutative ring which is not a field.
144 CHAPTER 7. FINITE FIELDS

7.2 The structure of Multiplicative Group of a fi-


nite field

Lemma 7.2.1. Let G be a finite abelian group. If G has at most


one subgroup of order m for every natural number m then G is a cyclic
group.

Proof. Let n = |G| and let n = pe11 . . . perr be the standard prime
decomposition of n in which p1 , . . . , pr are distinct prime numbers.
For each 1 ≤ i ≤ r, let Gi be the Sylow pi subgroup of G. By the
structure theorem of finite abelian groups, every Gi is a direct product
of some cyclic subgroups, say, Gi = H1 × · · · × Ht , where each Hj is
cyclic. According to Lagrange’s theorem, the order of every Hj is a
power of pi . So every Hj contains a subgroup of order pi and G has
at least t subgroups of order p. Hence t = 1, which means that every
Sylow subgroup of G is a cyclic group. Lemma 5.3.2 implies that G is
cyclic.

Theorem 7.2.2. Let p be a prime number and let n be a natural


number. If p|n then there is no subgroup of order n in F̄∗p , otherwise
there is a unique subgroup H of order n in F̄∗p . Moreover, this subgroup
H is cyclic.

Proof. First assume that p|n. Suppose that H is a subgroup of


F̄∗pwith |H| = n. Then there is an element α of order p in H. Hence
(α − 1)p = αp − 1 = 0, which leads to contradiction. Hence there is no
subgroup of order n in F̄∗p .
Next assume that n is not divisible by p. Since the polynomial xn −1
is coprime to its formal derivative, it has no multiple zeros in F̄p . Since
F̄p is algebraically closed, xn − 1 has exactly n zeros in F̄p . Let H be
the set of all zeros of xn − 1 in F̄p . It is obvious that H is a subgroup
of F̄∗p with |H| = n.
Let K be a subgroup of F̄∗p with |K| = n. Lagrange’s theorem implies
that an = 1 for every a ∈ K, which implies that every element in K is
a zero of xn − 1 in F̄p . So K = H. This proves the uniqueness of the
subgroup of order n in F̄∗p .
7.2. THE STRUCTURE OF MULTIPLICATIVE GROUP OF A FINITE FIELD145

In particular, H contains at most one subgroup of order m for any


natural number m. It follows from 7.2.1 that H is a cyclic group.

Corollary 7.2.3. Assume that q = pn , where p is a prime number


and n is a natural number. Then the multiplicative group F∗q = Fq \{0}
is a cyclic group.

Corollary 7.2.4. Let F be a subfield of a finite field E. Then E/F


is a simple extension.

Proof. Take any generator α of the cyclic group E ∗ . Then E =


F [α].

Corollary 7.2.5. There exist irreducible polynomials of any given


degree over a finite field Fq .

Proof. For any natural number n, Corollary 7.2.4 implies that


Fqn = Fq [α] for some α ∈ Fqn . The minimal polynomial of α over Fq is
an irreducible polynomial of degree n.

Definition 7.2.6. Let F be a finite field. Any generator of the


cyclic group F ∗ is called a primitive element of F.

Example 7.2.7. Let p be a prime number and let m be a natural


number not divisible by p. Then there is a natural number n such that
m divides pn − 1.

Proof. According to Theorem 7.2.2 there is a cyclic subgroup of


order m in the multiplicative group F̄∗p . Let α be the generator of this
cyclic subgroup. and let E = Fp [α]. Then |E| = pn for some natural
number n. Hence m|pn − 1 by Lagrange’s theorem.

This example shows that some problems in elementary number the-


ory can be solved by using finite fields. The following example gives
one more simple application of finite fields in number theory. For more
advanced application see the proof of quadratic reciprocity in Appendix
1.
146 CHAPTER 7. FINITE FIELDS

Theorem 7.2.8 (Wilson). For any prime number p the congruence

(p − 1)! ≡ −1 (mod p)

holds.

Proof. The theorem holds evidently when p = 2. Assume that p


is an odd prime. Then 1 6= −1 in Fp and ±1 are all solutions of the
equation x2 = 1 in Fp . Hence t 6= t−1 for any t ∈ Fp \{1, −1}. This
implies that Y
t = 1.
t∈F∗p \{1,−1}

So Y
t = −1,
t∈F∗p

which can be rewritten as


p−1
Y
j̄ = −1.
j=1

This is obviously equivalent to

(p − 1)! ≡ −1 (mod p).

Exercises

1. Let Fp be a field of p elements in which p is a prime number.


Give an irreducible polynomial of degree p over Fp .
2. Let a be a nonzero element in a finite field F. Show that there
are x, y ∈ F such that x2 + y 2 = a.
3. Let p be a prime number and f (x) ∈ Fp [x], g(x) = xp − x. Show
that gcd(f, g) is the product of all distinct factors of degree one of f (x).
4. Decompose x4 + x3 + x + 3 ∈ F5 [x] into the product of irreducible
polynomials.
7.2. THE STRUCTURE OF MULTIPLICATIVE GROUP OF A FINITE FIELD147

5. Let q be a power of a prime number. Find the number of solutions


of the equation y q z + yz q = xq+1 in the finite field Fq2 .
Chapter 8

Finite Galois Theory

One classical problem in the history of mathematics is to find formulas


to solve a polynomial equation in one variable.
More precisely, let

f (x) = xn + a1 xn−1 + · · · + an−1 x + an = 0 (8.1)

be an equation of degree n. Is there any formula which can give a


solution in terms of the coefficients a1 , . . . , an involving addition, sub-
traction, multiplication, division and radicals.
It is known that equations of degrees upto 4 has such formulas. The
solutions of x2 + a1 x + a2 = 0 are
p
−a1 ± a21 − 4a2
x=
2
for arbitrary elements a1 , a2 . This solution is actually a function with
a1 , a2 as its variables. More generally the solution of the equation 8.1
is a multi-valued function in the variables a1 , . . . , an .
We may consider to approach the problem by using field extensions
as in the ruler and compass construction.
First we need to formulate the problem properly. What we are
given is the complex number field and the variables a1 , . . . , an . So we
start with the field C(a1 , . . . , an ) in which a1 , . . . , an are indeterminates.
Then we can add new elements by ordinary arithmetic “+, −, ×, /”.
They are contained in C(a1 , . . . , an ).

148
149

However, adding an n-th root will extend the field. Let F be some
field. Let n be a natural number greater than 1 and let a ∈ F. The
zeros of xn − a = 0 are not necessarily in F. It is natural to consider a
finite extension E of F containing all zeros of xn − a = 0, preferably
minimal. Unlike the ruler and compass construction, the degree [E : F ]
is not necessarily a power of 2. It can be an arbitrary natural number.
To handle the problem in the way similar to the ruler and compass
construction we formulate a recursive process searching for a solution
of (8.1) as follows. Let L be the algebraic closure of F0 = C(a1 , . . . , an ).
For i ≥ 0 let Fi+1 = Fi (α1 , . . . , αr ), in which α1 , . . . , αr are all zeros of
xni − a = 0 in L for some a ∈ Fi and some natural number ni . Such an
extension Fi+1 /Fi is called a radical extension.
If there is a sequence of radical extensions

F0 ⊆ F1 ⊆ F2 ⊆ · · · ⊆ Fm

such that Fm contains all zeros of (8.1) then we say the the equation
(8.1) is solvable by radicals.
Since L is algebraically closed,

f (x) = (x − α1 ) . . . (x − αn )

where α1 , . . . , αn ∈ L. Let E = F0 (α1 , . . . , αn ). Hence the equation


f (x) = 0 is solvable by radicals if and only if there is a finite sequence
of radical extensions

F0 ⊆ F1 ⊆ F2 ⊆ · · · ⊆ Fm

such that E ⊆ Fm .
It is desirable to test the solvability by radicals directly from the
extension E/F0 . This is not as simple as the ruler and compass con-
struction problem in which the degree of the extension can give a neg-
ative answer. The Galois theory relates the problem of solvability by
radicals to the group theory. The extension E/F0 above corresponds to
a finite group called the Galois group of E/F0 . The equation f (x) = 0
is solvable by radicals if and only if the corresponding Galois group is
150 CHAPTER 8. FINITE GALOIS THEORY

solvable. Since the Galois group of every polynomial equation of degree


greater than or equal to five with general coefficients is not solvable,
there is no formula for the zeros of f (x) = 0 involving addition, sub-
traction, multiplication, division and radicals only.
This is only a typical application of Galois theory, which has promi-
nent significance in mathematics. It becomes an important tool in
modern algebra, number theory and other branches of mathematics.
In this chapter we will discuss the main theorem of Galois theory and
its classical application to the solution of polynomial equations by rad-
icals.

8.1 Basic theory

Definition 8.1.1. Let E/F be an extension. Denote by Aut(E/F )


the group of all automorphisms σ of E such that σ(a) = a ∀a ∈ F,
called the automorphism group E over F.

It is straightforward to check that Aut(E/F ) is a group under the


operation of composite.
Let K be an intermediate field of E/F. Then Aut(E/K) is a sub-
group of Aut(E/F ). This gives a map from the set of all intermediate
fields of E/F to the set of all subgroups of Aut(E/F ) by

K 7→ Aut(E/K).

On the other hand, for any subgroup H of Aut(E/F ) let

E H = {b ∈ E|σ(b) = b for all σ ∈ H}.

Then E H is an intermediate field of E/F, called the fixed subfield of


H. It gives a map from the set of subgroups of Aut(E/F ) to the set of
all intermediate fields of E/F by

H 7→ E H .

Lemma 8.1.2. 1) Let K1 , K2 be intermediate fields of E/F with


K1 ⊆ K2 . Then Aut(E/K1 ) ⊇ Aut(E/K2 );
8.1. BASIC THEORY 151

2) Let H1 , H2 be subgroups of Aut(E/F ) with H1 ⊆ H2 . Then E H1 ⊇


E H2 ;
3) K ⊆ E Aut(E/K) holds for any intermediate field K of E/F ;
4) H ⊆ Aut(E/E H ) holds for any subgroup H of Aut(E/F ).

The proof is left to the readers.


1) and 2) of Lemma 8.1.2 mean that the corresponding subgroup
shrinks when the intermediate field grows and vice versa. But be aware
that it is possible that Aut(E/K1 ) and Aut(E/K2 ) can be the same
when K1 ⊂ K2 .

Definition 8.1.3. Let E/F be a finite extension and let L/F be


an arbitrary extension of F. A homomorphism σ : E → L such that
σ(a) = a for all a ∈ F is called an embedding of E/F into L/F.

Proposition 8.1.4. The number of embeddings of a finite extension


E/F to L/F does not exceed [E : F ] for any extension L/F.

Proof. Let n = [E : F ]. Suppose that there are n + 1 distinct


embeddings σ1 , . . . , σn , σn+1 from E/F to L/F. Let u1 , u2 , . . . , un be a
basis of E/F. Since the number of rows of the matrix
 
σ1 (u1 ) σ2 (u1 ) · · · σn+1 (u1 )
···
 
 
σ1 (un ) σ2 (un ) · · · σn+1 (un )

is less than the number of columns, there are not all zero elements
a1 , . . . , an+1 ∈ L such that
  
σ1 (u1 ) σ2 (u1 ) · · · σn+1 (u1 ) a1
  .. 
···   .  = 0.


σ1 (un ) σ2 (un ) · · · σn+1 (un ) an+1

Hence
  
i σ1 (u1 ) σ2 (u1 ) · · · σn+1 (u1 ) a1
  .. 
h
c1 · · · cn  ···  .  = 0

σ1 (un ) σ2 (un ) · · · σn+1 (un ) an+1
152 CHAPTER 8. FINITE GALOIS THEORY

for arbitrary c1 , . . . , cn ∈ F. This implies that

a1 σ1 (c1 u1 + · · · + cn un ) + · · · + an+1 σn+1 (c1 u1 + · · · + cn un ) = 0

for arbitrary c1 , . . . , cn ∈ F. Since u1 , . . . , un form a basis of E/F,

a1 σ1 (x) + · · · + an+1 σn+1 (x) = 0

for any x ∈ E. This contradicts Theorem 2.4.10.

Corollary 8.1.5. For any finite extension E/F, the inequality


|Aut(E/F )| ≤ [E : F ] holds.

Proof. If follows from Proposition 6.2.6 that Aut(E/F ) consists


all embeddings of E/F into E/F. So the corollary is a consequence of
Proposition 8.1.4

This corollary tells us that [E : F ] is an upper bound of |Aut(E/F)|.


The extension E/F should be of special interest when this upper bound
is reached. Thus it is natural to introduce the following definition.

Definition 8.1.6. Let E/F be a finite extension. If [E : F ] =


|Aut(E/F)| then E/F is called a Galois extension and the group
Aut(E/F) is called the Galois group of E/F, denoted by Gal(E/F)
or G(E/F).

Example 8.1.7. Assume the the characteristic of a field F is not


equal to 2. Then any quadratic extension E/F (i.e., [E : F ] = 2) is a
Galois extension.

Proof. Choose any α ∈ E\F. Then E = F [α] and the minimal


polynomial p(x) of α is a quadratic polynomial

p(x) = x2 + ax + b ∈ F [x].

Since the characteristic of F is not equal to 2 by assumption there


exists c ∈ F such that a = 2c. Hence p(x) = (x + c)2 + b − c2 . Since
f (x) is irreducible over F, b − c2 6= 0. Since p0 (x) = 2(x + c) 6= 0, p(x)
does not have multiple zeros. Hence p(x) = (x − α)(x − β) in which
8.1. BASIC THEORY 153

β ∈ E and β 6= α. It is evident that 1, α and 1, β are two bases of E as


a vector space over F. Define a map

σ : E → E, c + dα 7→ c + dβ, (c, d ∈ F ).

Then σ is an invertible linear transformation of the F -space E. To verify


that σ ∈ Aut(E/F) it suffices to check that σ((c1 + d1 α)(c2 + d2 α)) =
σ(c1 + d1 α)σ(c2 + d2 α) for any c1 , d1 , c2 , d2 ∈ F. It follows from

(c1 + d1 α)(c2 + d2 α) = (c1 c2 − bd1 d2 ) + (c1 d2 + c2 d1 )α,

that

σ((c1 +d1 α)(c2 +d2 α)) = (c1 c2 −bd1 d2 )+(c1 d2 +c2 d1 )β = σ(c1 +d1 α)σ(c2 +d2 α),

So σ ∈ Aut(E/F) and Aut(E/F) contains at least two elements id and


σ. It follows from |Aut(E/F)| ≤ [E : F ] = 2 that |Aut(E/F)| = 2.

Example 8.1.8. The cubic extension Q[ 3 2]/Q is not a Galois ex-

tension, in which 3 2 is the real cubic root of 2.
√ √ √
Proof. Let σ ∈ Aut(Q[ 3 2]/Q) and let α = σ( 3 2). Since ( 3 2)3 =
√ √
2, α3 = 2. But 3 2 is the only number in Q[ 3 2] ⊂ R whose cubic

power is equal to 2, so α = 3 2. Hence σ is the identity map. Hence

|Aut(Q[ 3 2]/Q)| = 1.

Proposition 8.1.9. Every finite extension of a finite field is a Ga-


lois extension with a cyclic group as its Galois group.

Proof. Let F = Fq be a finite field containing q elements and let


E/F be a finite extension with [E : F ] = n. Then E ∼
= Fqn . Let

φ : E → E, a 7→ aq .

Then φ ∈ Aut(E/F) and φ generates a cyclic subgroup of order n of


Aut(E/F). Hence |Aut(E/F)| = n = [E : F ] by Corollary 8.1.5, which
implies that E/F is a Galois extension and G(E/F) is a cyclic group
of order n.
154 CHAPTER 8. FINITE GALOIS THEORY

Proposition 8.1.10. Let K be an intermediate field of a finite


Galois extension E/F. Then E/K is a Galois extension.

Proof. Let n = [E : F ], m = [E : K], d = [K : F ]. Let A be the


set of all embeddings from K/F to E/F. For any g ∈ G(E/F), let g|K
denote the restriction of g on K. Then g 7→ g|K gives a map

f : G(E/F) → A.

For an arbitrary h ∈ A such that f −1 (h) is not empty, there is some


u ∈ G(E/F) such that u|K = h. It is easy to verify the f −1 (h) =
uAut(E/K), i.e., f −1 (h) is a left coset of the subgroup Aut(E/K) in
G(E/F). This means that f −1 (h) contains at most |Aut(E/K)| ele-
ments for any h ∈ A. Since G(E/F ) is the disjoint union of f −1 (h)
where h runs over all elements of A, the inequality

|G(E/F)| ≤ |Aut(E/K)| · |A| (8.2)

holds. It follows from Corollary 8.1.5 that |Aut(E/K)| ≤ [E : K].


Lemma 8.1.4 tells us that |A| ≤ [K : F ]. Hence the inequality (8.2)
implies |G(E/F)| ≤ [E : K][K : F ] = [E : F ]. Since E/F is a Galois
extension by assumption, |G(E/F )| = n. Hence |Aut(E/K)| = [E : K],
which means that E/K is a Galois extension.

Lemma 8.1.11. Let G be a finite subgroup of Aut(E) for a field E.


Let F = E G = {a ∈ E|g(a) = a for all g ∈ G}. Then E/F is a
finite Galois extension with G(E/F) = G.

Proof. Let g1 , g2 , . . . , gn be all elements of G. It suffices to show


that [E : F ] ≤ n. Suppose that there are n + 1 elements u1 , . . . , un+1 in
E linearly independent over F. The system of homogeneous equations

g1 (u1 )x1 + · · · + g1 (un+1 )xn+1 = 0,

··· (8.3)

gn (u1 )x1 + · · · + gn (un+1 )xn+1 = 0


8.1. BASIC THEORY 155

has a nonzero solution

x1 = a1 , . . . , xn+1 = an+1

in the field E. Choose one solution with minimum number of nonzero


elements. Without loss of generality, assume that a1 = 1. If ai ∈ F for
1 ≤ i ≤ n + 1, then

g1 (a1 u1 + · · · + an+1 un+1 ) = 0,

so
a1 u1 + · · · + an+1 un+1 = 0,

contradicting the hypothesis that u1 , . . . , un+1 are linearly independent


over F. Hence there is at least one ai that is not in F. Assume that
a2 ∈
/ F without loss of generality.
According to the definition of F there is some g ∈ G such that
g(a2 ) 6= a2 . By applying g to each equation in (8.3) we obtain

gg1 (u1 )g(a1 ) + · · · + gg1 (un+1 )g(an+1 ) = 0,

···

ggn (u1 )g(a1 ) + · · · + ggn (un+1 )g(an+1 ) = 0.

Since gg1 , gg2 , . . . , ggn are still all elements of G, these equations are
merely a rearrangement of the equations

g1 (u1 )g(a1 ) + · · · + g1 (un+1 )g(an+1 ) = 0,

···

gn (u1 )g(a1 ) + · · · + gn (un+1 )g(an+1 ) = 0.

By subtracting the equations

g1 (u1 )a1 + · · · + g1 (un+1 )an+1 = 0,


156 CHAPTER 8. FINITE GALOIS THEORY

···

gn (u1 )a1 + · · · + gn (un+1 )an+1 = 0

one by one we obtain equations

g1 (u1 )(g(a1 ) − a1 ) + g1 (u2 )(g(a2 ) − a2 ) + · · · = 0,

···

gn (u1 )(g(a1 ) − a1 ) + gn (u2 )(g(a2 ) − a2 ) + · · · = 0,

which means that g(a1 ) − a1 , g(a2 ) − a2 , . . . , g(an+1 ) − an+1 form a new


nonzero solution of the system (8.3). Since g(a1 )−a1 = 0, g(a2 )−a2 6= 0
and g(ai ) − ai = 0 whenever ai = 0 for 3 ≤ i ≤ n + 1, the number
of nonzero elements in {g(a1 ) − a1 , g(a2 ) − a2 , . . . , g(an+1 ) − an+1 )} is
less than that in {a1 , a2 , . . . , an+1 }. A contradiction is reached. Hence
[E : F ] ≤ n.

Corollary 8.1.12. A finite extension E/F is Galois if and only


if for every α ∈ E\F there is some σ ∈ Aut(E/F) such that σ(α) 6= α.

Proof. ⇐: Let G = Aut(E/F). Then E G = F and E/F is a Galois


extension by Lemma 8.1.11.
⇒: Assume that E/F is a Galois extension. Let K = E G(E/F) .
Then E/K is a Galois extension with G(E/F) as its Galois group.
Hence [E : K] = |G(E/F)| = [E : F ]. It follows from F ⊆ K that
K = F.

Remark 8.1.13. Due to Lemma 8.1.11 the item 4) of Lemma 8.1.2


can be improved as: H = Aut(E/E H ) for every subgroup H of Aut(E/F).

Theorem 8.1.14 (The fundamental theorem of Galois theory). Let


E/F be a finite Galois extension with n = [E : F ]. Then the following
three statements hold:
1) The maps
φ : K 7→ Aut(E/K)
8.1. BASIC THEORY 157

and
ψ : H 7→ E H

give a one-to-one correspondence between the set of all intermediate


fields of E/F and the set of all subgroups of G(E/F).
2) [K : F ] = (G(E/F) : G(E/K)) for any intermediate subfield K
of E/F.
3) For any intermediate field K, K/F is a Galois extension if and
only if G(E/K) / G(E/F). In this case, G(K/F) ∼
= G(E/F)/G(E/K).

Proof. To prove 1) we need to verify that both ψ ◦ φ and φ ◦ ψ


are identity maps.
First of all, Lemma 8.1.11 means exactly that φ ◦ ψ is the identity
map.
According to Proposition 8.1.10 E/K is a Galois extension for any
intermediate field K. Hence |Aut(E/K)| = [E : K]. Let B = E Aut(E/K) .
Then B ⊇ K. By Lemma 8.1.11 E/B is a Galois extension, so [E :
B] = |Aut(E/K)| = [E : K] and B = K. This proves that ψ ◦ φ is the
identity map.
2) Since both E/F and E/K are Galois extensions, [E : F ] =
|G(E/F)|, [E : K] = |G(E/K)|. Hence [K : F ] = [E : F ]/[E : K] =
(G(E/F) : G(E/K)).
3) Assume that H / G(E/F). Let K = E H . and let g be an arbitrary
element in G(E/F). We claim that g(b) ∈ K for every b ∈ K. Otherwise
there would be some b ∈ K such that g(b) ∈ / K. So there would be
some h ∈ H such that h(g(b)) 6= g(b). It follows from g −1 hg ∈ H that
g −1 hg(b) = b, i.e., hg(b) = g(b), contradiction.
Thus a homomorphism

φ : G(E/F) → Aut(K/F ), g 7→ g|K

is well defined. It is evident that Ker(φ) = G(E/K). By the funda-


mental theorem of homomorphism we have

|Imφ| = (G(E/F) : G(E/K)) = [E : F ]/[E : K] = [K : F ].


158 CHAPTER 8. FINITE GALOIS THEORY

Since
[K : F ] = |Imφ| ≤ |Aut(K/F )| ≤ [K : F ],

the homomorphism φ is surjective and |Aut(K/F )| = [K : F ]. This


means that K/F is a Galois extension and G(K/F) ∼= G(E/F)/G(E/K).
Conversely, assume that K is an intermediate field such that K/F
is Galois. Then |G(K/F)| = [K : F ]. Since there are at most [K : F ]
embeddings from K/F into E/F, G(K/F) contains all embeddings from
K/F to E/F. For every σ ∈ G(E/F), σ|K is an embedding of K/F into
E/F. Hence σ ∈ G(K/F). This give a homomorphism

G(E/F) → G(K/F), σ 7→ σ|K

whose kernel is G(E/K). Hence G(E/K) / G(E/F).

Theorem 8.1.15. A finite extension E/F is Galois if and only if


there is some α ∈ E satisfying the following two conditions:
1) E = F [α];
2) the minimal polynomial f (x) ∈ F [x] of α is decomposed in E[x]
into
f (x) = (x − α1 )(x − α2 ) · · · (x − αn )

for some distinct α1 , α2 , . . . , αn in E.

Proof. When F is a finite field the statement is true due to Propo-


sition 8.1.9 and Corollary 7.2.4. Hence we may assume that F is an
infinite field.
⇐: Assume that E/F is a Galois extension of degree n. For any
u ∈ E let
Hu = {g ∈ G(E/F)|g(u) = u}.

Then Hu is a subgroup of G(E/F).


For any u, v ∈ E we claim that there is some w ∈ E such that
Hw = Hu ∩ Hv . Since G(E/F) has only finitely many subgroups but F
contains infinitely many elements, there are c1 , c2 ∈ F, c1 6= c2 such that
Hu+c1 v = Hu+c2 v by pigeonhole principle. Let w = u + c1 v. It is obvious
that Hu ∩ Hv ⊆ Hw . Since g(u + c1 v) = u + c1 v, g(u + c2 v) = u + c2 v
for every g ∈ Hw , (c1 − c2 )g(v) = (c1 − c2 )v, which implies g(v) = v
8.1. BASIC THEORY 159

and thus g(u) = u. So g ∈ Hu ∩ Hv . This shows that Hw = Hu ∩ Hv .


This concludes the proof of the claim.
By induction, for any u1 , . . . , um ∈ E there is w ∈ E such that
Hw = Hu1 ∩ · · · ∩ Hum . In particular, let u1 , . . . , un ∈ E be a basis of
E over F. Then

Hu1 ∩ · · · ∩ Hun = {g ∈ G(E/F)|g(x) = x ∀x ∈ E}.

Hence
Hu1 ∩ · · · ∩ Hun = {1}

by the fundamental theorem of Galois theory. Therefore there exists


α ∈ E such that Hα = {1}.
Let f (x) be the minimal polynomial of α over F. Then deg(f ) ≤ n.
Suppose that deg(f ) < n. Since f (g(α)) = 0 for any g ∈ G(E/F), by
pigeonhole principle there exist two distinct elements g1 , g2 ∈ G(E/F)
such that g1 (α) = g2 (α), so g1−1 g2 ∈ Hα and thus g1−1 g2 = 1. This
contradicts with g1 6= g2 . Therefore E = F [α].
Let 1 = g1 , g2 , . . . , gn be all elements of G(E/F) and let αi =
gi (α), 1 ≤ i ≤ n. Then α1 , . . . , αn are distinct elements in E and each
of them is a zero of f (x). Hence

f (x) = (x − α1 )(x − α2 ) · · · (x − αn ).

⇒: For every αi there is a surjective homomorphism

fi : F [x] → E, h(x) 7→ h(αi )

such that Ker(fi ) = (f (x)). Hence fi induces an isomorphism φi :


F [x]/(f (x)) → E. Denote gi = φi ◦ φ−1
1 , (1 ≤ i ≤ n). Then g1 , . . . , gn
are distinct elements of Aut(E/F). Hence |Aut(E/F)| ≥ n. It follows
from Corollary 8.1.5 that |Aut(E/F)| = [E : F ]. Therefore E/F is a
Galois extension.

Let E and K be subfields of a field L. The compositum of E and


K is the smallest subfield of L containing E and K, denoted by EK. It
is the extension of E generated by the set K or that of K generated by
160 CHAPTER 8. FINITE GALOIS THEORY

E. The compositum of more than two subfields is defined in a similar


way.

Proposition 8.1.16. Let E, K and F be subfields of a field L such


that F ⊆ E and F ⊆ K. Assume that E/F is a finite Galois extension.
Then EK/K is a finite Galois extension and G(EK/K) ∼ = G(E/E ∩
K).

Proof. If E = F then the proposition is trivial. We may assume


that [E : F ] > 1.
According to Theorem 8.1.15 E = F [α] such that the minimal poly-
nomial f (x) of α over F is equal to

(x − α1 ) · · · (x − αn )

for distinct elements α1 , . . . , αn ∈ E. We may assume that α1 = α.


Then KE = K[α] is a finite extension of K.
Let p(x) = xm + cm−1 xm−1 + · · · + c1 x + c0 ∈ K[x] be the minimal
polynomial of α over K. Then p(x) divides f (x) in K[x]. Hence every
zero of p(x) in L is a zero of f (x). Without loss of generality, we may
assume that p(x) = (x − α1 ) · · · (x − αm ). It follows from Theorem
8.1.15 again that EK/K is a finite Galois extension.
Since every coefficient ci of p(x) belongs to F [α1 , . . . , αm ], so ci ∈
E ∩ K for 0 ≤ i < m. This means that p(x) ∈ E ∩ K[x]. It is clear
that p(x) is irreducible over E ∩ K. Hence p(x) is also the minimal
polynomial of α over E ∩ K. Therefore [EK : K] = [E : E ∩ K].
The restriction map g 7→ g|E gives a homomorphism φ : G(EK/K) →
G(E/E ∩ K). Assume that g ∈ Ker(φ). Then g|E is the identity map
from E to E, which implies that g(α) = α. It follows from EK = K[α]
that g : EK → EK is the identity map. Hence φ is injective. Since
[EK : K] = [E : E ∩ K], φ is a bijection.

Exercises

1. Show that the condition that the characteristic of the field


is not equal to 2 is indispensable in Example 8.1.7 by the example
F2 (x)/F2 (x2 ).
8.1. BASIC THEORY 161

2. Let ζ = (−1 + −3)/2 and let F = Q[ζ]. Show that the cubic

extension F [ 3 2]/F is Galois.

3. Let L/F be a finite Galois extension and let H1 , H2 be two


subgroups of G(L/F) with E1 , E2 as their fixed fields respectively.
1) Show that the fixed field of H1 ∩ H2 is E1 E2 .
2) Show that E1 ∩ E2 is the fixed field of H1 H2 where H1 H2 is the
subgroup of G(L/F) generated by H1 ∪ H2 .

4. Let K/E, E/F be finite Galois extensions. Show by example


that K/F is not necessarily a Galois extension.

5. Let E/F be a finite extension and let L be the algebraic closure of


E. If there exist [E : F ] distinct embeddings from E/F into L/F, then
E/F is called a separable extension; if every embedding from E/F
into L/F carries E into E then E/F is called a normal extension.
Prove that E/F is a Galois extension if and only if E/F is separable
and normal.

6. Let F be a field and let f (x) ∈ F [x]. Let α1 , . . . , αn be all zeros


of f (x) in the algebraic closure F̄ of F. Then the field F [α1 , . . . , αn ] is
referred to as the splitting field of f (x).
Show that a finite extension E/F is a Galois extension if and only if
it is the splitting field of some f (x) ∈ F [x] that does not have multiple
zeros in the algebraic closure of F.

7. Let p be an odd prime number and let E/F be a Galois extension


whose Galois group G(E/F) is a non-abelian group of order 2p. Let K
be an intermediate field of E/F such that [E : K] = 2. Show that K/F
is not a Galois extension.

8. Let E/F be a finite Galois extension and let α ∈ E. Assume


that g(α) 6= α for every g ∈ G(E/F)\{1}. Show that E = F [α].

9. Let E/F be a finite extension. Show that the integer |Aut(E/F)|


divides [E : F ].
162 CHAPTER 8. FINITE GALOIS THEORY

8.2 * Solvable Extension and Solvability of Alge-


braic Equations by Radicals

Definition 8.2.1. A finite extension E/F if a solvable extension if


there is a finite chain of intermediate fields

F = E0 ⊂ E1 ⊂ · · · ⊂ En = E

such that Ei /Ei−1 is Galois and G(Ei /Ei−1 ) is abelian for 1 ≤ i ≤ n.

Let g be an automorphism of a field K. For any f (x) = an xn +


an−1 xn−1 + · · · + a0 ∈ K[x] denote by f g (x) the polynomial g(an )xn +
g(an−1 )xn−1 + · · · + g(a0 ) ∈ K[x].

Lemma 8.2.2. For a fixed automorphism g of a field K, the map


K[x] → K[x], f (x) 7→ f g (x) is a ring isomomorphism. A polynomial
f (x) ∈ K[x] is irreducible if and only if f g (x) is irreducible over K.

Proof. Let f (x) = an xn + an−1 xn−1 + · · · + a0 and h(x) = bn xn +


bn−1 xn−1 + · · · + b0 be two polynomials in K[x] (in which an or bn are
not necessarily nonzero). Then

f g (x)+hg (x) = g(an +bn )xn +g(an−1 +bn−1 )xn−1 +· · ·+g(a0 +b0 ) = (f h)g (x)

and the coefficient of xr in (f h)g (x) is i+j=r g(ai bj ) = i+j=r g(ai )g(bj ),
P P

which is exact the coefficient of xr in f g (x)hg (x). Hence (f h)g (x) =


f g (x)hg (x). This shows that the map Φg : f (x) 7→ f g (x) is a homomor-
phism.
Since Φg−1 is the inverse of Φg , the homomorphism Φg is an isomor-
phism.
The second statement is clear.

Lemma 8.2.3. Let E/K, K/F be finite Galois extensions and let Ē
be an algebraically closed field containing E. There exists a subfield L
of Ē such that
1) E ⊆ L;
2) L/F is a finite Galois extension;
8.2. * SOLVABLE EXTENSION AND SOLVABILITY OF ALGEBRAIC EQUATIONS BY RADICALS1

3) there is a chain of intermediate fields

K = L0 ⊆ L1 ⊆ · · · ⊆ L n = L

such that Li /Li−1 is a Galois extension and G(Li /Li−1 ) is isomorphic


to a subgroup of G(E/K) for 1 ≤ i ≤ n.

Proof. Let {σ1 , . . . , σr } be the set of all embeddings of E/F into


Ē/F in which σ1 is the identity map. Denote Ei = σi (E) for 1 ≤ i ≤ r.
Let L = E1 · · · Er .
Since K/F is Galois, σi (K) = K. The group Aut(Ei /K) is isomor-
phic to G(E/K) via the correspondence τ 7→ σi−1 τ σi . Hence

|Aut(Ei /K)| = |G(E/K)| = [E : K] = [Ei : K],

which implies that Ei /K is a Galois extension for each i.


Let Li = E1 · · · Ei for 0 ≤ i ≤ n. Since Li = Li−1 Ei , Proposition
8.1.16 implies that the extension Li /Li−1 is Galois and G(Li /Li−1 ) ∼=
G(Ei /Ei ∩ Li−1 ), which is isomorphic to a subgroup of G(E/K). So 3)
is proved.
Let σ be an embedding of L/F into Ē/F. Then σ(Ei ) = σσi (E) ⊆
L, since σσi is an embedding from E/F into Ē/F. This shows that
σ ∈ Aut(L/F). Hence every embedding of L/F is an automorphism of
L over F.
It remains to show that L/F is a Galois extension. It suffices to
show that for any a ∈ L\F there is some σ ∈ Aut(L/F) such that
σ(a) 6= a due to Corollary 8.1.12. This is proved in two separate cases.
Case 1) a ∈ K\F.
Since K/F is Galois, there is some g ∈ G(K/F ) such that g(a) 6= a.
There exists an embedding σ : L → Ē from L/F into Ē/F such that
σ|K = g by Theorem 6.4.8. Then σ ∈ Aut(L/F) and σ(a) 6= a.
Case 2) a ∈ L\K.
Then a ∈ Li \Li−1 for some 1 ≤ i ≤ n. Since Li /Li−1 is Galois, there
is some g 0 ∈ G(Li /Li−1 ) such that g 0 (a) 6= a. Then the same argument
as in Case 1) shows that there is σ ∈ Aut(L/F) such that σ|Li = g 0 .
164 CHAPTER 8. FINITE GALOIS THEORY

Lemma 8.2.4. For any finite solvable extension E/F there is a finite
extension L/E such that L/F is a finite solvable Galois extension.

Proof. Let

F = E0 ⊂ E1 ⊂ · · · ⊂ En = E

be a finite chain of intermediate fields such that Ei /Ei−1 is Galois and


G(Ei /Ei−1 ) is abelian for 1 ≤ i ≤ n.
The lemma is proved by induction on n. When n = 1 the lemma is
obviously true. Assume that n > 1. Since En−1 /F is solvable, by induc-
tion hypothesis there is a finite extension K 0 /En−1 such that K 0 /F is a
finite solvable Galois extension. Let Ē be the algebraic closure of E and
let σ be an embedding of K 0 /En−1 into Ē/En−1 . Denote K = σ(K 0 ).
Then K is a finite extension of En−1 such that K/F is a finite solvable
Galois extension. It is obvious that En−1 ⊆ E ∩ K.
Since E/En−1 is a Galois extension, so is KE/K and G(KE/K) ∼ =
G(E/E ∩K) by Proposition 8.1.16. Since E ∩K is an intermediate field
of E/En−1 , G(E/E ∩K) is a subgroup of the abelian group G(E/En−1 ).
Hence G(KE/K) is an abelian group.
By Lemma 8.2.3 there is a finite Galois extension L of F in Ē
containing KE and a chain of intermediate fields

KE = L0 ⊆ L1 ⊆ · · · ⊆ Ln = L

such that Li /Li−1 is Galois with an abelian Galois group for each 1 ≤
i ≤ n.
Put all these extensions together and we obtain a chain of extensions

F ⊆ K ⊆ KE = L0 ⊆ L1 ⊆ · · · ⊆ Ln = L

in which K/F is a solvable extension, G(KE/K) is abelian and each


G(Li /Li−1 ) is abelian for 1 ≤ i ≤ n. Hence L/F is solvable.

Proposition 8.2.5. A finite Galois extension E/F is solvable if


and only if G(E/F) is a solvable group.
8.2. * SOLVABLE EXTENSION AND SOLVABILITY OF ALGEBRAIC EQUATIONS BY RADICALS1

Proof. ⇒: If there are intermediate subfields

F = K0 ⊂ K 1 ⊂ K2 · · · ⊂ Kn = E

such that Ki /Ki−1 is Galois with an abelian Galois group for 1 ≤ i ≤ n.


By the fundamental theorem of Galois theory, there is a subnormal
series

{1} = G(E/Kn ) / G(E/Kn−1 ) / · · · / G(E/K1 ) / G(E/K0 ) = G(E/F).

Since G(E/Ki−1 )/G(E/Ki ) ∼ = G(Ki /Ki−1 ) is an abelian group for each


i, G(E/F) is a solvable group.
⇐: Assume that G(E/F) is a solvable group. Then there is a sub-
normal series

{1} = H0 / H1 / · · · / Hn−1 / Hn = G(E/F)

such that Hi /Hi−1 is abelian for each 1 ≤ i ≤ n. It corresponds to a


sequence of intermediate fields

F = E Hn ⊆ E Hn−1 ⊆ · · · ⊆ E H1 ⊆ E H0 = E.

Since Hi / Hi+1 , E Hi /E Hi+1 is a Galois extension with G(E Hi /E Hi+1 ) ∼


=
Hi+1 /Hi by the fundamental theorem of Galois theory. Therefore E/F
is a solvable extension.

Let us investigate a special type of Galois extension, which is closely


related to the solvability of polynomial equations by radicals.

Proposition 8.2.6. Let p be a prime number and let F be a field of


characteristic zero. Assume that F contains all p-th roots of unity. Let
E/F be an extension and let α be an element of E such that αp ∈ F.
Then F [α]/F is a Galois extension with a cyclic Galois group.

Proof. If α ∈ F then the proposition becomes trivial. Hence we


assume that α ∈
/ F.
Since the characteristic of F is zero, gcd(xp − 1, pxp−1 ) = 1, the
polynomial xp − 1 has p distinct zeros 1, ζ, ζ 2 , . . . , ζ p−1 in the algebraic
166 CHAPTER 8. FINITE GALOIS THEORY

closure of E by Lemma 3.4.11. Since 1, ζ, ζ 2 , . . . , ζ p−1 ∈ F and αp =


a ∈ F by the given conditions, there is the following decomposition

xp − a = (x − α)(x − ζα)(x − ζ 2 α) · · · (x − ζ p−1 α).

Let f (x) be the minimal polynomial of α over F. Then deg(f ) > 1 by


the assumption α ∈ / F. Since f (x) is a factor of xp − a, at least one ζ d α
is a zero of f (x) in which 1 ≤ d ≤ p − 1. Hence there is an embedding
g from F [α] to Ē such that g(α) = ζ d α. Since g 2 , g 3 , . . . , g p−1 are also
embeddings of F [α]/F into Ē/F and g i (α) = ζ di α for 1 ≤ i ≤ p − 1,
there are p distinct embeddings from F [α]/F into Ē/F. If follows from
Lemma 8.1.4 that p ≤ [F [α] : F ]. Since [F [α] : F ] = deg(f ) ≤ deg(xp −
a) = p, we have p = [F [α] : F ] and f (x) = xp − a. By Theorem
8.1.15 F [α]/F is a Galois extension and its Galois group consists of the
elements 1, g, g 2 , g 3 , . . . , g p−1 , which is a cyclic group of order p.

Let C be the field of complex numbers and let x1 , x2 , . . . , xn be


indeterminates. Let L = C(x1 , . . . , xn ) be the field of rational functions
over C with variables x1 , x2 , . . . , xn .
Every element σ in the symmetric group Sn determines a permuta-
tion of variables x1 , . . . , xn , which induces an automorphism of the field
L. All such automorphisms form a subgroup G of Aut(L) isomorphic
to Sn . Let
F = {a ∈ L|f (a) = a ∀f ∈ G}.

It follows from Lemma 8.1.11 that L/F is a Galois extension and


G(L/F) = G ∼ = Sn .

Proposition 8.2.7. If n > 4 then the Galois extension L/F is not


solvable.

Proof. By Theorem 1.6.5 Sn is not solvable for n > 4, L/F is not


solvable by Proposition 8.2.6.

Let
σ1 = x1 + · · · + xn ,

σ2 = x1 x2 + x1 x3 + · · · + xn−1 xn ,
8.2. * SOLVABLE EXTENSION AND SOLVABILITY OF ALGEBRAIC EQUATIONS BY RADICALS1

···

σn = x1 x2 · · · xn

be the elementary symmetric polynomials.

Proposition 8.2.8. Under the above notation,

F = K(σ1 , σ2 , . . . , σn ).

Proof. Since F consists of all fractions u(x1 , . . . , xn )/v(x1 , . . . , xn )


such that u and v are symmetric polynomials. It is well-known that
every symmetric polynomial can be written as h(σ1 , . . . , σn ) for some
polynomial h(y1 , . . . , yn ). Hence the field F is generated by elementary
polynomials.

Corollary 8.2.9. A polynomial equation in one variable with de-


gree greater or equal to 5 and with general coefficients is not solvable
by radicals.

Proof. A monic polynomial of general coefficients can be written


as
f (x) = xn − σ1 xn−1 + · · · + (−1)i σi xn−i + (−1)n σn .

Starting from E0 = C(σ1 , . . . , σn ) construct a chain of extensions

E0 ⊂ E1 ⊂ E2 · · · ⊂ EN ,

where Ei = Ei−1 [αi ], αipi ∈ Ei−1 . We may assume that each pi is a prime
number, since any radical is the composite of prime order radicals. For
√ p√
example 6 a = 3 a. By Proposition 8.2.6 each extension Ei /Ei−1 is a
Galois extension with cyclic Galois group. Hence EN /E0 is a solvable
extension. By Lemma 8.2.4 there is a finite solvable Galois extension
L/E0 containing EN as an intermediate field.
Suppose that EN contains all zeros x1 , . . . , xn of f (x). Then C(x1 , . . . , xn )
is an intermediate field of the solvable Galois extension L/C(σ1 , . . . , σn ).
Hence G(C(x1 , . . . , xn )/C(σ1 , . . . , σn )) is a quotient group of the solv-
able group G(L/E0 ). This contradicts Proposition 8.2.7.

To conclude this course we introduce the inverse Galois problem.


168 CHAPTER 8. FINITE GALOIS THEORY

Proposition 8.2.10. Every finite group G is isomorphic to the


Galois group of some finite Galois extension.

Proof. By Cayley’s theorem, G is isomorphic to a subgroup of


some symmetric group Sn .
Since the Galois group of C(x1 , . . . , xn )/C(σ1 , . . . , σn ) is Sn , there is
an intermediate subfield E such that G(C(x1 , . . . , xn )/E) is isomorphic
to G by the fundamental theorem of Galois theory.

This proof is easy. But if the base field is required to be the field of
rational numbers, the problem becomes a long standing open problem
in algebra. The precise formulation is as follows:
Is any finite group isomorphic to the Galois group of some finite
Galois extension of Q?

Exercises

1. Let E/F be a finite Galois extension and let H be a subgroup of


G(E/F). Show that there is β ∈ E such that H = {g ∈ G(E/F)|g(β) =
β}.
2. Construct two finite Galois extensions of Q whose Galois groups
are cyclic groups of orders 3 and 4 respectively.
3. Let p be a prime number and let F = Fp (x) be the field of
rational functions in one variable over the finite field Fp . Let E =
F [y]/(y p − y + x). Show that E/F is a Galois extension of degree p.
Appendix A

Quadratic residues

Let p be a prime number. The map sq : F∗p → F∗p , t → t2 is a group


homomorphism.
Definition A.0.11. An integer n not divisible by p is called a
quadratic residue modulo p if n̄ ∈ Im(sq), otherwise it is a quadratic
nonresidue modulo p.
For any prime number p, the Legendre symbol of an integer n
module p is defined as


  
n  0, if p|n
= 1, if p does not divide n and n is a quadratic residue modulo p
p 
 −1, if p does not divide n and n is a quadratic nonresidue modulo p

Example A.0.12. 1, 2, 4 are quadratic residues modulo 7 while


3, 5, 6 are quadratic nonresidues modulo 7.
Lemma A.0.13. Let p be an odd prime and let 0, a1 , . . . , ap−1 be a
complete residue system modulo p. Then there are exactly (p − 1)/2
quadratic residues among a1 , . . . , ap−1 .
 
Proof. This follows from api = 1 if and only if ai ∈ Im(sq) and

|F∗p | p−1
|Im(sq)| = = .
|Ker(sq)| 2

169
170 APPENDIX A. QUADRATIC RESIDUES

Lemma A.0.14. The equality


    
mn m n
=
p p p

holds for any integer m, n and any prime number p.


Proof. If p|mn then both sides are equal to zero.
Assume that none of m, n is divisible by p. By the definition of
Legendre symbol,  
n
n 7→
p
is a group homomorphism from F∗p to {1, −1}. Hence the equlity holds.

Lemma A.0.15. Let n be an integer not divisible by an odd prime


p. Then ( np ) = 1 if and only if
p−1
n 2 ≡1 (mod p).

Proof. Since F∗p is a cyclic group of order p − 1,


 
n
= 1 ⇔ n ∈ Im(sq) ⇔ n|Im(sq)| = 1.
p

Corollary A.0.16. Let p be an odd prime. Then


 
−1 p−1
= (−1) 2 .
p

 Lemma A.0.17. Let p be an odd prime. If p ≡ ±1 (mod 8), then


2
p
= 1, otherwise p2 = −1.
Proof. Since 8 is coprime with p, Theorem 7.2.2 implies that there
exists an element α of order 8 in F̄∗p , i.e., α8 = 1 and αd 6= 1 for any
1 ≤ d < 8. Hence α4 = −1 and thus α2 + α−2 = 0. So 2 = (α + α−1 )2 .
If p ≡ ±1 (mod 8), then

(α + α−1 )p = αp + α−p = α + α−1 .


171

Hence (α + α−1 )p−1 = 1. It follows that


p−1
2 2 = (α + α−1 )p−1 = 1.

Lemma A.0.15 implies that


 
2
= 1.
p

If p ≡ ±5 (mod 8) then

(α + α−1 )p = α5 + α−5 = −(α + α−1 ).

So (α + α−1 )p−1 = −1. Hence


p−1
2 2 = (α + α−1 )p−1 = −1.

It follows from Lemma A.0.15 that


 
2
= −1.
p

The following theorem is one of the fundamental theorem in number


theory.

Theorem A.0.18 ( Gauss reciprocity law ). Let p, p0 be two distinct


prime numbers. Then
   0
p (p−1)(p0 −1) p
0
= (−1) 4 .
p p

Choose an element w of order p in F̄∗p0 . For any integers n, m such


that n ≡ m (mod p), the equality wn = wm holds. Hence f (x) = wx
is a function on Fp .
The element
X x
y= wx
x∈F
p
p

is in F̄p0 .
172 APPENDIX A. QUADRATIC RESIDUES

Lemma A.0.19.
y 2 = (−1)(p−1)/2 p.

Proof.
X  xz 
2
y = wx+z
x,z∈Fp
p
 
X X  x(u − x) 
= wu  .
u∈F x∈F
p
p p

When u = 0,
X  x(u − x)  
−1

= (p − 1) .
x∈Fp
p p

When u 6= 0,
X  x(u − x)  X  −x2   1 − x−1 u 
=
x∈Fp
p x∈F∗p
p p

1 − x−1 u
 X   
−1 −1
= =− .
p x∈F∗p
p p

It follows from
p−1
X X
wu = wi = −1
u∈F∗p i=1

that  
2 −1
y =p = (−1)(p−1)/2 p.
p

Lemma A.0.20.  0
p0 −1 p
y = .
p
173

Proof. Since y ∈ F̄p0 ,


X x
p0 0
y = wxp
x∈Fp
p
 0 X  0
p xp 0
= wxp
p x∈F p
p
 0
p
= y.
p

Theorem A.0.18 follows from

 
p0 −1 2 (p0 −1)/2
(p−1)(p0 −1)
(p0 −1)/2
(p−1)(p0 −1) p
y = (y ) = (−1) 4 p = (−1) 4
p0

and Lemma A.0.20.

Example A.0.21.
        
29 43 14 2 7
= = =
43 29 29 29 29
   
7 29
= − =−
29 7
 
1
= − = −1.
7
Appendix B

Every finite skew field is a


field

Let n be a natural number greater than 1 and let ζ = e2mπi/n =


cos(2mπ/n) + i sin(2mπ/n), in which m is a natural number such that
m < n and gcd(m, n) = 1. Then ζ is is called an n-th primitive root of
unity. Let C denote the set of all n-th primitive roots of unity.
The equation xn − 1 = 0 has n zeros 1, e2πi/n , e4πi/n , . . . , e2(n−1)πi/n
in the complex number field. They form a cyclic group of order n under
the multiplication. Then C is exactly the set of all elements of order n
in this cyclic group. The number of elements in C is the Euler function
φ(n).
If ζ, η ∈ C, then there is a natural number d such that gcd(d, n) = 1
and ζ d = η.
Q
Theorem B.0.22. Let f (x) = ζ∈C (x − ζ). Then f (x) is a monic
irreducible polynomial in Z[x] such that f |xn − 1 and deg(f ) = φ(n).

Proof. Choose any ζ ∈ C. Since ζ n − 1 = 0, the complex number


ζ is algebraic over Q. Let g(x) be the minimal polynomial of ζ over Q.
Then xn − 1 = g(x)h(x), for some monic polynomial h(x) over Q. It
follows from Gauss’s lemma that g(x), h(x) ∈ Z[x].
Let p be a prime that does not divide n. Suppose that g(ζ p ) 6= 0.
Then h(ζ p ) = 0, since g(ζ p )h(ζ p ) = ζ pn − 1 = 0. For any

u(x) = am xm + am−1 xm−1 + · · · + a1 x + a0 ∈ Z[x]

174
175

let
ū(x) = am xm + am−1 xm−1 + · · · + a1 x + a0 ∈ Fp [x],

in which aj is the residue class aj modulo p. Then

xn − 1 = ḡ(x)h̄(x)

holds in Fp [x]. Hence h̄(x)p = h̄(xp ), since F (xp ) = F (x)p for any
F (x) ∈ Fp [x]. Since p does not divide n, the polynomial xn − 1 has
no multiple zeros in Fp [x]. Hence ḡ(x) is coprime with h̄(x), so ḡ(x)
is coprime with h̄(xp ). By Lemma 3.2.12 there exist ū(x), v̄(x) ∈ Fp [x]
such that
ū(x)ḡ(x) + v̄(x)h̄(xp ) = 1,

i.e., there exist u(x), v(x), w(x) ∈ Z[x] such that

u(x)g(x) + v(x)h(xp ) = 1 + pw(x).

Substitute ζ for x and we obtain pw(ζ) + 1 = 0. Hence g(x)|pw(x) + 1


and so ḡ(x)|1, which leads to contradiction. Hence g(ζ p ) = 0 for any
prime number p that does not divide n.
Every η ∈ C can be written as η = ζ m for some integer m that is
coprime with n. Let m = p1 p2 . . . ps be the prime decomposition of m.
Then every pj does not divide n. Hence g(η) = 0, which implies that
Q
f (x) = ζ∈C (x − ζ) divides g(x). In particular, deg(f ) ≤ deg(g).
On the other hand, since g(x) is irreducible, Q[α] ∼= Q[ζ] holds for
every zero α of g(x). Hence every zero of g(x) is in C. So deg(g) ≤
deg(f ), which implies deg(f ) = deg(g). Since both f (x) and g(x) are
monic and f (x)|g(x), it follows that f (x) = g(x).

The polynomial f (x) in Theorem B.0.22 is called the n-th cyclo-


tomic polynomial, denoted by Φn (x).

Lemma B.0.23. Let a be a real number greater than or equal to 2.


Then |Φn (a)| > a − 1.

Proof. Since Y
Φn (a) = (a − ζ)
ζ∈C
176 APPENDIX B. EVERY FINITE SKEW FIELD IS A FIELD

by the definition of Φn (x), so |a − ζ| > a − 1 ≥ 1 for each ζ ∈ C.

Theorem B.0.24 (MacLagen-Wedderburn). Every finite skew field


is a field.

Proof. Suppose that E is a finite non-commutative skew field. Let

F = {a ∈ E|ab = ba for any b ∈ E}.

We claim that F is a field. For any a1 , a2 ∈ F and any b ∈ E, the


equalities

(a1 + a2 )b = a1 b + a2 b = ba1 + ba2 = b(a1 + a2 )

and

(a1 a2 )b = a1 (a2 b) = a1 (ba2 ) = (a1 b)a2 = (ba1 )a2 = b(a1 a2 )

hold. Hence a1 + a2 , a1 a2 ∈ F. For any a ∈ F \{0} the equality ab = ba


implies ba−1 = a−1 b and so a−1 ∈ F. Hence F is a field. It is evident
that the multiplication in F is commutative. Hence F is a field. Let
q = |F |. Then q is a power of a prime number.
The skew field E has a natural structure of vector space over F. So
|E| = q n , where n is the dimension of E over F.
The set E ∗ of all nonzero elements of E is a finite group of order
q n − 1 under multiplication. Consider the conjugation action g(a) =
gag −1 of E ∗ on itself. Let x1 , . . . , xr be a complete set of representatives
of the orbits. Since E is not commutative by hypothesis, there exists
at least one orbit of length greater than one. We may assume that the
length of the orbit containing xi for 1 ≤ i ≤ m is equal to one and the
remaining orbits has length greater than one. Let Hi be the stabilizer
of xi for i > m. Then F ∗ ⊆ Hi . It is easy to verify that Hi ∪ {0} is a
subspace of vector space E over F. Hence |Hi | = q ni − 1 with ni < n.
Since x1 , . . . , xr are exactly all the elements of F ∗ , we have
s
X qn − 1
q−1+ n
= q n − 1. (B.1)
j=1
q −1
j
177

It follows from nj < n that Φn (x)(xnj − 1)|xn − 1. Hence


s
n
X xn − 1
x −1− = Φn (x)u(x).
j=1
xnj − 1

Since Φn (x) is a monic polynomial with integral coefficients by Theorem


B.0.22, we have u(x) ∈ Z[x]. The equality

q − 1 = Φn (q)u(q)

holds by (B.1). Therefore q −1 = |Φn (q)||u(q)| ≥ |Φn (q)|, contradicting


Lemma B.0.23.
Appendix C

Solutions of a cubic equation


and Hilbert theorem 90

By a standard change of variables x = y + b1 /3 a cubic equation y 3 +


b1 y 2 + b2 x + b3 = 0 is transformed into the form

x3 + a2 x − a3 = 0. (C.1)

Let x1 , x2 , x3 be three roots of (C.1). Then

x1 + x2 + x3 = 0,

x 1 x 2 + x 1 x 3 + x 2 x 3 = a2 ,

x 1 x 2 x 3 = a3 .

Let F = C(a2 , a3 ), E = C(x1 , x2 , x3 ). Then E/F is a Galois exten-


sion with the symmetric group S3 as its Galois group. Denote by σ the
element in G(E/F ) carrying x1 , x2 , x3 to x2 , x3 , x1 respectively. Then σ
generates a cyclic normal subgroup H of G(E/F ). Let K be the fixed
field of H. By the fundamental theorem of Galois theory, K/F is a
quadratic extension and E/K is a cubic Galois extension. Let τ be the
element of G(E/F ) carrying x1 , x2 into x2 , x1 respectively and keeping
x3 unchanged. Then the image of τ in G(K/F ) ∼ = G(E/F )/G(E/K)
is a generator of G(K/F ).
As a first step, we try to find a generator of the extension K/F. A

178
179

standard choice is

δ = (x1 − x2 )(x2 − x3 )(x3 − x1 ) ∈ K\F,

since σ(δ) = δ and τ (δ) = −δ. Hence K = F [δ]. In fact, δ 2 = −4a32 −


27a23 is the discriminant of the equation (C.1).
The next step is to find the minimal equation of x1 in the Galois
cubic extension K[x1 ]/K. We introduce a useful technique.

Let ζ = e2πi/3 = −1/2 + i 3/2 be a cubic root of unity. Then

1 + ζ + ζ 2 = 0. (C.2)

Let
u = x1 + ζx2 + ζ 2 x3 , v = x1 + ζ 2 x2 + ζx3 .

Then

σ(u) = x2 + ζx3 + ζ 2 x1 = ζ 2 u, σ(v) = x2 + ζ 2 x3 + ζx1 = ζv.

Hence
σ(u3 ) = u3 , σ(v 3 ) = v 3 .

This means that u3 , v 3 ∈ F [δ]. Hence u, v are cubic roots of some


elements in F [δ].
By linear algebra, x1 , x2 , x3 are linear combinations of u, v. By the
definition of u, v, there is a system of linear equations
    
1 1 1 x1 0
2   =   .
1 ζ ζ  x2  u

1 ζ2 ζ x3 v

Multiplying  
1 1 1
1
1 ζ 2 ζ 

3
1 ζ ζ2
to the both sides yields
u v u v u v
x1 = + , x2 = ζ 2 + ζ , x3 = ζ + ζ 2 . (C.3)
3 3 3 3 3 3
180APPENDIX C. SOLUTIONS OF A CUBIC EQUATION AND HILBERT THEOREM 90

It remains to compute u, v. A simple manipulation shows that uv =


−3a2 , so
u3 v 3 = −27a32 . (C.4)

In terms x1 , x2 , x3 express u3 and v 3 as

u3 = 9a3 + 3ζx21 x2 + 3ζ 2 x1 x22 + 3ζx22 x3 + 3ζ 2 x2 x23 + 3ζx23 x1 + 3ζ 2 x3 x21 ,

v 3 = 9a3 + 3ζ 2 x21 x2 + 3ζx1 x22 + 3ζ 2 x22 x3 + 3ζx2 x23 + 3ζ 2 x23 x1 + 3ζx3 x21 .

Then

u3 + v 3 = 18a3 − 3(x21 x2 + x1 x32 + x22 x3 + x2 x23 + x23 x1 + x3 x21 ) = 27a3 ∈ F.

Since u3 , v 3 ∈ F [δ], they can be written as

u3 = α + βδ, v 3 = α0 + β 0 δ,

with α, β, α0 , β 0 ∈ F and β, β 0 6= 0, for u3 , v 3 ∈


/ F. So β 0 = −β since
u3 + v 3 ∈ F. Then u3 v 3 ∈ F implies α0 = α. Hence

2α = u3 + v 3 = 27a3 ,

which implies α = 27a3 /2. It follows from (C.4) that

α2 − β 2 (−4a32 − 27a23 ) = −27a32 .

So
27 3
(4a32 + 27a23 )β 2 = − (4a2 + 27a23 ).
4

Thus β = ± −27/2. By the symmetry of u3 and v 3 we may assume

that β = −27/2. Hence we have
r
a3 a2  3  a3  2
(u/3)3 = + + ,
2 3 2
r
3 a3 a2  3  a3  2
(v/3) = − + .
2 3 2
Since every nonzero complex number has three distinct cubic roots, we
181

are facing the problem of the choices. We can choose any cubic root
for u/3, which is always a generator of E/K. So u/3 can be written as
s r
3 a3 a2  3  a3  2
u/3 = + + .
2 3 2

Once u/3 is chosen, v/3 is determined. It must satisfy the condition


uv a2
=− .
33 3
By using (C.3) we obtain

Theorem C.0.25 ( Cardano’s formula). The three roots of the cubic


equation x3 + a2 x − a3 = 0 are
s r  s r 
3 a3 a2 3 a3 2 3 a3 a2 3  a3  2
 
x1 = + + + − + ,
2 3 2 2 3 2
s r  s r
2 3 a3 a 3 2
 a 2
3 3 a3 a2  3  a3  2
x2 = ζ + + +ζ − + ,
2 3 2 2 3 2
s r  s r
3 a3 a 3
2
 a 2
3 2 3 a3 a2  3  a3  2
x3 = ζ + + +ζ − + ,
2 3 2 2 3 2
in which the two cubic root satisfy
s r  s r 
3 a3 a2 3 a3 3 a3
2 a2 3  a3  2 a2
 
+ + − + =− .
2 3 2 2 3 2 3

A key point in the proof is the existence of u such that σ(u) = ζ 2 u,


i.e. ζ 2 = σ(u)/u. This can be regarded as a special case of the following
celebrated theorem.

Theorem C.0.26 (Hilbert theorem 90). Let E/F be a Galois ex-


tension of degree n such that G(E/F ) is a cyclic group generated by σ.
For any η ∈ E ∗ the following two statements are equivalent
1) ησ(η)σ 2 (η) · · · σ n−1 (η) = 1;
2) there exists u ∈ E ∗ such that η = σ(u)/u.
182APPENDIX C. SOLUTIONS OF A CUBIC EQUATION AND HILBERT THEOREM 90

Proof. 2) ⇒ 1): Assume that η = σ(u)/u. Then

σ(u) σ 2 (u) σ n−1 (u) u


ησ(η)σ 2 (η) · · · σ n−1 (η) = · · · n−2 = 1.
u σ(u) σ (u) σ n−1 (u)

1) ⇒ 2): Assume that ησ(η)σ 2 (η) · · · σ n−1 (η) = 1.


Let f : E → E be map carrying ξ ∈ E into

ξ + ησ(ξ) + ησ(η)σ 2 (ξ) + · · · + ησ(η) · · · σ n−2 (η)σ n−1 (ξ).

It is evident that σ(f (ξ)) = ηf (ξ) for arbitrary ξ ∈ E. It suffices to


show the existence of ξ ∈ E such that f (ξ) 6= 0. But this is an easy
consequence of Theorem 2.4.10.

When η ∈ F ∗ and η n = 1 The condition 1) of Hilbert theorem 90


always holds. So u = ξ + ησ(ξ) + η 2 σ(ξ) + · · · + η n−1 σ(ξ) 6= 0, and
ησ(u) = u. For this reason, the choice of u = x1 + ζx2 + ζ 2 x3 and
v = x1 + ζ 2 x2 + ζx3 in solving the cubic equation is closely related to
Hilbert theorem 90.
Appendix D

Solutions of quartic
equations

The formula for the solutions of a quartic equation can be obtained in


a similar way as for the cubic equation by using Galois theory.
The coefficient of x3 can be eliminated by a suitable change of vari-
able. So a quartic equation can be written as

x4 + a2 x2 − a3 x + a4 = 0. (D.1)

Let x1 , x2 , x3 , x4 be the four roots of (D.1). Then

x1 + x2 + x3 + x4 = 0,

x 1 x 2 + x 1 x 3 + x 1 x 4 + x 2 x 3 + x 2 x 4 + x 3 x 4 = a2 ,

x 1 x 2 x 3 + x 1 x 2 x 4 + x 1 x 3 x 4 + x 2 x 3 x 4 = a3 ,

x 1 x 2 x 3 x 4 = a4 .

Let F = C(a2 , a3 , a4 ), E = C(x1 , x2 , x3 , x4 ). Then E/F is a Ga-


lois extension with G(E/F ) ∼ = S4 . Every element in S4 determines a
permutation of x1 , x2 , x3 , x4 .
Start as a first step with a composition series of the symmetric
group S4 :
{e} ⊂ K ⊂ H ⊂ A4 ⊂ S4 ,

183
184 APPENDIX D. SOLUTIONS OF QUARTIC EQUATIONS

in which K is the cyclic subgroup generated by (12)(34) and H =


{e, (12)(34), (13)(24), (14)(23)}. The factor group A4 /H is of order 3
and all other factors have order 2.
According to the fundamental theorem of Galois theory, this com-
position series determines a chain of intermediate fields

E ⊃ E1 ⊃ E2 ⊃ E3 ⊃ F.

Since [E : E1 ] = 2 and G(E/E1 ) = {e, (12)(34)}, the minimal


polynomial of x1 over E1 is

(x − x1 )(x − x2 ) = x2 − (x1 + x2 ) + x1 x2 . (D.2)

whose coefficients x1 + x2 and x1 x2 belong to E1 but not to E2 , for


they are changed into x3 + x4 and x3 x4 respectively under the action
of (13)(24).
Since (13)(24) is a generator of G(E1 /E2 ), the minimal polynomials
of x1 + x2 and x1 x2 over E2 are

(x − x1 − x2 )(x − x3 − x4 ) = x2 + (x1 + x2 )(x3 + x4 ) (D.3)

and
(x − x1 x2 )(x − x3 x4 ) = x2 − (x1 x2 + x3 x4 )x + a4 (D.4)

respectively.
We need to find (x1 + x2 )(x3 + x4 ), x1 x2 + x3 x4 and their conjugates
(x1 + x3 )(x2 + x4 ), (x1 + x4 )(x2 + x3 ), x1 x3 + x2 x4 , x1 x4 + x2 x3 over
F. In fact, it is enough to know x1 x2 + x3 x4 , x1 x3 + x2 x4 , x1 x4 + x2 x3 ,
because
(x1 + x2 )(x3 + x4 ) + x1 x2 + x3 x4 = a2 ,

(x1 + x3 )(x2 + x4 ) + x1 x3 + x2 x4 = a2 ,

(x1 + x4 )(x2 + x3 ) + x1 x4 + x2 x3 = a2 .

Since H  S4 and S4 /H ∼ = S3 , E2 /F is a Galois extension of degree


6 with the Galois group S3 . This coincides with the Galois extension of
185

a cubic equation. The cubic equation satisfied by the elements x1 x2 +


x3 x4 , x1 x3 + x2 x4 , x1 x4 + x2 x3 is

(x − (x1 x2 + x3 x4 ))(x − (x1 x3 + x2 x4 ))(x − (x1 x4 + x2 x3 ))


= x3 − a2 x2 − 4a4 x + 4a2 a4 − a23
= 0.

This cubic equation is called the resolvent cubic of the quartic


equation (D.1). Let λ1 , λ2 , λ3 be its three roots. By the symmetry of
the three elements x1 x2 +x3 x4 , x1 x3 +x2 x4 , x1 x4 +x2 x3 , we may assume
that
x 1 x 2 + x 3 x 4 = λ1 ,

x 1 x 3 + x 2 x 4 = λ2 ,

x 1 x 4 + x 2 x 3 = λ3 .

(x1 + x2 )(x3 + x4 ) = a2 − λ1 ,

(x1 + x3 )(x2 + x4 ) = a2 − λ2 ,

(x1 + x4 )(x2 + x3 ) = a2 − λ3 .

By (D.3) the equation satisfied by x1 + x2 and x3 + x4 is x2 = λ1 − a2 .


Hence p p
x 1 + x 2 = λ 1 − a2 , x 3 + x 4 = − λ 1 − a2 .

For the same reason


p p
x1 + x3 = λ 2 − a2 , x 2 + x 4 = − λ 2 − a2 .

It is not hard to see that the 4 difference combinations of the choices


of two square roots only affect the order of x1 , x2 , x3 , x4 . So these

two square roots can be chosen arbitrarily. But once λ1 − a2 and
186 APPENDIX D. SOLUTIONS OF QUARTIC EQUATIONS

λ2 − a2 are chosen, x1 + x4 and x2 + x3 are determined, because

(x1 + x2 )(x1 + x3 )(x1 + x4 )


1
= (x1 + x2 − x3 − x4 )(x1 + x3 − x2 − x4 )(x1 + x4 − x2 − x3 )
8
= a3 .

So the element x1 + x4 = λ3 − a2 satisfies
p p p
λ 1 − a2 λ 2 − a2 λ 3 − a2 = a3 .

It follows from
p p p
2x1 = (x1 +x2 )+(x1 +x3 )+(x1 +x4 ) = λ1 − a2 + λ 2 − a2 + λ 3 − a2

that
1 p p p
x1 = ( λ1 − a2 + λ2 − a2 + λ3 − a2 ),
2
1 p p p
x2 = ( λ1 − a2 − λ2 − a2 − λ3 − a2 ),
2
1 p p p
x3 = (− λ1 − a2 + λ2 − a2 − λ3 − a2 ),
2
1 p p p
x4 = (− λ1 − a2 − λ2 − a2 + λ3 − a2 ).
2
Appendix E

Hints or solutions for


exercises

Exercises 1.1
1. Check that the binary operation in G satisfies the three condition
for a group.
(1) Law of associativity: (a ∗ b) ∗ c = [aln(b) ]ln(c) = aln(b)ln(c) and
ln(c)
a ∗ (b ∗ c) = aln[b ] = aln(b)ln(c) imply (a ∗ b) ∗ c = a ∗ (b ∗ c).
(2) Existence of the identity element: Let e be the base of the
natural logarithm. Then e ∗ a = eln(a) = a, a ∗ e = aln(e) = a for every
a ∈ G. Hence e is the identity element.
(3) Existence of the inverse: For any a ∈ G, let b = e1/ln(a) . Then

a ∗ b = aln(b) = (eln(a) )ln(b) = eln(a)ln(b) = e,

b ∗ a = bln(a) = (eln(b) )ln(a) = eln(b)ln(a) = e.

Hence b is the inverse of a. 2


2. The composition of two strictly increasing continuous functions
f (x), g(x) with f (0) = g(0) = 0, f (1) = g(1) = 1 is still a function sat-
isfying all conditions required by the set A. Hence the binary operation
is well-defined.
The associativity holds in A, since the composition of functions
satisfies the associative law. Define the function e(x) = x, (0 ≤ x ≤ 1).
Then e ∈ A, and f e = ef for any f ∈ A. Hence e is the identity element

187
188 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

of A. Finally, the inverse function f −1 (x) of any strictly increasing


function f (x) exists and f −1 (x) is also strictly increasing. Hence f −1 ∈
A. It is obvious that f f −1 (x) = x = f −1 f (x) for any x ∈ [0, 1]. Hence
f f −1 = f −1 f = e. 2

3. It suffices to show that


(a) ae = a for any a ∈ G;
(b) If a, b ∈ G satisfy ba = e, then ab = e.
Assume that a, b ∈ G satisfy ba = e. The first condition implies

bab = eb = b. (E.1)

By the second condition there exists c ∈ G such that cb = e. Multiply-


ing both sides of (E.1) by c from the left yields cbab = cb. It follows
that ab = e, which proves (b).
For any a ∈ G, the second condition and (b) implies the existence
of b ∈ G such that ab = ba = e. Hence ae = a(ba) = (ab)a = ea = a,
which proves (a). 2

4. Assume that n = |G| < ∞. For any a ∈ G, there are two elements
among a, a2 , a3 , . . . , an+1 that are equal by the pigeon hole principle.
Assume that ai = aj (i < j). Then aj−i = e. 2

5. Let n be the order of ab. Then (ab)n = 1. Multiplying both


sides of (ab)n = 1 by a−1 from the left and by a from the right yields
(ba)n = 1. Hence the order of ba does not exceed that of ab. For the
same reason the order of ab does not exceed that of ba. Hence they are
equal. 2

6. The given condition implies ab = (ab)−1 = b−1 a−1 for any a, b ∈


G. It follows from a−1 = a, b−1 = b that ab = ba. 2

Exercises 1.2
1. Let A, B ∈ G. Then
   
1 a b 1 a0 b 0
A = 0 1 c  , B = 0 1 c0  .
   
0 0 1 0 0 1
189

So  
1 a + a0 b + ac0 + b0
AB = 0 1 c + c0  ∈ G,
 
0 0 1
 
1 −a ac − b
A−1 = 0 1 −c  ∈ G.
 
0 0 1
Hence G is a subgroup of GL2 (R).
The center of G consists of all
 
1 x y
C = 0 1 z 
 
0 0 1

such that
     
1 a b 1 x y 1 x y 1 a b
0 1 c  0 1 z  = 0 1 z  0 1 c 
     
0 0 1 0 0 1 0 0 1 0 0 1

for any a, b, c ∈ R. This means that


   
1 a + x b + az + y 1 x + a y + xc + b
0 1 c + z  = 0 1 z + c .
   
0 0 1 0 0 1

Hence az = xc for any a, c ∈ R, which holds if and only if x = z = 0.


Hence the center of G consists of all real matrices in the form
 
1 0 y
0 1 0
 
0 0 1

2. 1) Let a ∈ C(Y ). Then ab = ba for any b ∈ X, since X ⊆ Y. This


means a ∈ C(X). Hence C(X) ⊇ C(Y ).
2) Let a ∈ X. Since ab = ba for any b ∈ C(X) by the definition of
C(X), we have a ∈ C(C(X)). Hence X ⊆ C(C(X)).
190 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

3) It follows by 2) that C(X) ⊆ C(C(C(X))). Let Y = C(C(X)).


Then X ⊆ Y by 2). Thus C(X) ⊇ C(Y ) = C(C(C(X))) by 1). Hence
C(X) = C(C(C(X))). 2
3. Assume that a ∈ H, b ∈ G, b 6= 1. Than a ∈ hbi. Hence ab = ba,
which implies that every element H is in the center of G. 2
4. Let t be a generator of the cyclic group G. Then a = tn , b = tm ,
in which n, m are non-negative integers. Suppose that neither a nor b
is a perfect square. Then n, m are odd numbers. So n + m is even, i.e.,
n + m = 2r for some integer r. Hence ab is the square of tr .
Let Q∗ be the group of all nonzero rational numbers under the
multiplication. Then 2, 3, 2 · 3 = 6 are not perfect squares. 2
Remark: This example shows that Q∗ is not a cyclic group.
5. First proof: Let a be an arbitrary element in H. Since H is closed
under multiplication, every power of a is in H. Since G is a finite group,
some power of a is equal to the identity element 1. Hence 1 ∈ H.
Assume that an = 1. Let b = an−1 . Then ab = ba = 1. Hence H is
a subgroup of G.
Second proof: For any a ∈ H, construct a map f : H → H, b 7→ ab.
The cancelation law implies that f is injective. Since H is a finite
set, f is surjective too. Hence there exists c ∈ H such that f (c) = a,
i.e., ac = a, which implies that c = 1. There exists d ∈ H such that
f (d) = 1 for the same reason. So ad = 1, and d = a−1 . Therefore H is
a subgroup of G. 2
6. For any P, Q ∈ G the equality (P Q)T A(P Q) = QT (P T AP )Q =
QT AQ = A holds. Hence P Q ∈ G. Since P T AP = A for any P ∈ G,
we have A = (P T )−1 AP −1 = (P −1 )T AP −1 . Hence P −1 ∈ G. 2

Exercises 1.3
1. Hint: Every element of S7 which is not the identity element
can be expressed as the product of disjoint cycles. It is of one of the
following patterns:
(*******),(******),(*****),(****),(***),(**),
(*****)(**),(****)(**),(***)(**),(**)(**),
191

(****)(***),(***)(***),
(***)(**)(**), (**)(**)(**).
Answer1,2,3,4,5,6,7,10,12. 2
2. Hint: 20 = 4 × 5, 18 = 3 × 6 = 2 × 9 = 2 × 2 × 3 × 3.
Solution: The element (1234)(56789) is an element of order 20.
Writhe an element σ of order 18 as the product of disjoint cycles
σ = σ1 · · · σd such that the length of σi is not smaller than that of σi+1
for i = 1, . . . , d − 1. The sum of the lengths of all factors is less than or
equal to 9. The length of each σi is a factor of 18. So the length m1 of
σ1 is one of 9, 6, 3, 2. If m1 = 9, then σ is a 9-cycle, whose order is not
equal to 18. If m1 = 6, then σ has at most one more factor, which is
either a 3-cycle or a 2-cycle. In this case the order of σ is 6, not equal
to 18. If m1 ≤ 3, then the order of σ does not exceed 6. Hence there is
no element of order 18 in S9 . 2
3. Let A, B, C, D be the upper-left, upper-right, lower-left and
lower-right corners of the rectangle respectively. The permutations

id, (AB)(CD), (AC)(BD), (AD)(BC)

represent the identity map, left-right symmetry, top-bottom symmetry


and rotation by 180 degrees respectively. These elements are contained
in the symmetric group of the rectangle.
Since every isometry carries the long side to long side, it must carry
the ordered pair (A, B) to (A, B), (B, A), (C, D) or (D, C). It must be
one of the transformation as listed above. 2
4. Hint: Connect the centers of the six faces of the cube properly to
form an octagon located inside the cube. So the cubic and the regular
octagon has the same group of symmetry. 2

Exercises 1.4
1. Left cosets: H, {(13), (123)}, {(23), (132)}
Right cosets: H, {(13), (132)}, {(23), (123)} 2
2. The statement 1) does not hold. A counter-example is G = S3
and H is the subgroup of the previous exercise with a = (23). Then
(G : H) = 3 but a3 ∈
/ H.
192 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

The key of the proof of 2) is to use pigeon hold principle. If H has


n left cosets and none of a, a2 , . . . , an is in H, then at least two of them
are in the same left coset. Assume that as ∈ ar H, where 1 ≤ r < s ≤ n.
Then as−r = (ar )−1 as ∈ H, contradiction. 2
3. Let A be the set of all left cosets of H ∩ K in K and let B be
the set of all left cosets of H in G. Construct a map f : A → B which
carries a(H ∩ K) to aH. Need to show that this map is well-defined.
Assume that a(H ∩K) = a0 (H ∩K). Then a−1 a0 ∈ H ∩K. So a−1 a0 ∈ H,
and thus aH = a0 H. Hence f is well-defined.
Check that f is injective. Assume that f (a(H ∩K)) = f (a0 (H ∩K)),
i.e., aH = a0 H. Then a−1 a0 ∈ H. Since a, a0 ∈ K, so a−1 a0 ∈ K. Hence
a−1 a0 ∈ H ∩ K, which implies that a(H ∩ K) = a0 (H ∩ K). Hence f is
injective. Since B is a finite set, so is A. 2
4. Choose an arbitrary a ∈ K. Let H = a−1 K. Then H is a subset
of G containing the identity element. It suffices to show that H is a
subgroup of G.
For any h ∈ H define a map f : H → G, g 7→ hg. The image of f
is ha−1 K. Since h = ha−1 a ∈ ha−1 K, so h ∈ a−1 K ∩ ha−1 K. By the
hypothesis of the problem ha−1 K = a−1 K = H. Hence f is a bijection
from H to H. This implies that H is closed under multiplication. Since
1 = a−1 a ∈ H, the element 1 is in the image of f. Thus there exists
b ∈ H such that hb = 1. 2
5. Assume that a1 H1 = a2 H2 , then H1 = a−1 1 a2 H2 . Hence H1 is
a left coset of H2 containing the identity element, which implies that
H1 = H2 .
In S3 let H1 = {id, (12)}, H2 = {id, (13)}, a1 = (132), a2 = (132).
Then
a1 H1 = {(132), (23)} = H2 a2 .

2
Exercises 1.5
1. Let {Hi }i∈Λ be a collection of normal subgroups of a group G.
Let H = ∩i∈Λ Hi . For any g ∈ G and any h ∈ H, since h ∈ Hi for every
i ∈ Λ, g −1 hg ∈ Hi for every i ∈ Λ. Hence g −1 hg ∈ H. This shows that
H  G. 2
193

2. Hint: the conjugate of the product of two transpositions is still


the product of two transpositions. 2
3. Let g −1 Hg be a conjugate subgroup of H. Then |g −1 Hg| = r.
Since G has only one subgroup of order r, g −1 Hg = H. Hence H  G.
2
4. The key point is to use the normalizer NG (H) = {g ∈ G|g −1 Hg =
H}. It contains H. Since (G : H) = 5 is a prime number, Lagrange’s
theorem implies NG (H) is either H or G. It suffices to exclude the
possibility NG (H) = H.
It follows from aH = Hb that b−1 aH = b−1 Hb. Hence b−1 aH con-
tains the identity element. So b−1 Hb = H, which implies b ∈ NG (H).
The condition Hb 6= H implies b ∈ / H. 2
5. It is known in Exercise 2) that H2 = {e, (12)(34), (13)(24), (14)(23)}
is a normal subgroup of S4 . Let H1 = {e, (12)(34)}. Then H1  H2 but
H1 is not a normal subgroup of S4 . 2
6. Let a, a0 ∈ H, b, b0 ∈ K. Then

aba0 b0 = aba0 b−1 bb0 .

Since H  G, so ba0 b−1 ∈ H. Hence aba0 b−1 ∈ H, bb0 ∈ K, which implies


aba0 b0 ∈ HK. Hence HK is closed under multiplication.
Let a ∈ H, b ∈ K. Then (ab)−1 = b−1 a−1 = (b−1 a−1 b)b−1 ∈ HK.
Hence HK is a subgroup of G and 1) is proved.
For 2), since HK = {e, (12), (13), (132)} contains 4 elements, HK
is not a subgroup of S3 by Lagrange’s theorem.
Assume that H  G, K  G. Let a ∈ H, b ∈ K, g ∈ G. Then
g (ab)g = (g −1 ag)(g −1 bg) ∈ HK. Hence HK  G. 2
−1

7. Divide the proof into two steps: 1) amn = e; 2) If k > 0 and


ak = e, then mn|k.
1) Since m is the order of an in G, so (an )m = e. Hence anm = e.
2) Since āk = ak = ē, so n|k. Assume that k = nq. Then (an )q = e.
Since m is the order of an in G, so m|q. Therefore mn|k. 2
8. Hint: Let g be a generator of N. It suffices to show that g(a−1 b−1 ab) =
(a−1 b−1 ab)g for any a, b ∈ G.
194 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

Since N G, there exist integers m, n such that aga−1 = g n , bgb−1 =


gm. 2
9. Hint: Otherwise |G/Z| < 4, which would imply that G/Z is a
cyclic group. This would imply that G is an abelian group. 2
10. Hint: Any element g ∈ G can be written as bi b1 · · · bi−1 bi+1 · · · bn .
Hence gai g −1 = bi ai b−1
i for any ai ∈ Hi . 2
11. For any a ∈ K, since |aHa−1 | = n, either aHa−1 = H or
aHa−1 = K. If aHa−1 = K, then H = a−1 Ka = K, leading to contra-
diction. Hence aHa−1 = H. Since G is generated by H ∪K, g −1 Hg = H
for every g ∈ G. 2
Exercises 1.6
1. Every element of order 6 in S6 is either a 6-cycle or the product
of a 3-cycle and a transposition. So it is an odd permutation. Thus
G 6⊆ A6 . Hence (G : G ∩ A6 ) ≥ 2. Since (G : G ∩ A6 ) ≤ (S6 : A6 ) by
Ex.1.4.3, it follows that (G : G ∩ A6 ) = 2.
2. By Exercise 1.5.2, G = {e, (12)(34), (13)(24), (14)(23)} is a nor-
mal subgroup of S4 . Since every element of G is an even permutation,
G is also a normal subgroup of A4 . 2
3. For any natural number n let

Hn = {σ ∈ G|σ(i) = i for all i > n}.

Then H1 ⊆ H2 ⊆ H3 · · · , and G = ∞
S
n=1 Hn . It is evident that Hn = An
is simple when n ≥ 5.
Assume that N  G and N 6= G. There exists g ∈ Hm \N for some
m. Hence g ∈ Hn for every n > m. Since N ∩Hn Hn , so N ∩Hn = {e}
for every n ≥ max{5, m}. Therefore

[
N= (N ∩ Hn ) = {e}.
n=1

This shows that G is a simple group. 2


4. Hint: |G/(G∩An )| ≤ 2 implies |G∩An | > 1. Then use G∩An G.
2
195

5. Hint: The order of each 3-cycle has order 3 or 1 in the quotient


group Sn /G. 2
6. When n < 3, Sn is abelian, so [Sn , Sn ] and its subgroup of
commutators are trivial.
For n ≥ 3, it is obvious that [Sn , Sn ] is normal subgroup of G
contained in An . Since An is a non-abelian simple group when n ≥ 5,
[Sn , Sn ] and its subgroup of commutators are equal to An .

[S3 , S3 ] = A3 , [[S3 , S3 ], [S3 , S3 ]] = {1}.

It remains to treat the case n = 4. Since (12)(13)(12)(13) = (123),


so [S4 , S4 ] = A4 . Since A4 has only one non-trivial normal subgroup
N = {id, (12)(34), (13)(24), (14)(23)}, [A4 , A4 ] = A4 or N.
Let σ, τ ∈ A4 . Claim that σ −1 τ −1 στ ∈ N. Since A4 is generated by
3-cycles, we may assume that σ, τ are 3-cycles such that στ 6= τ σ. We
may assume that σ = (ijk) and τ = (ijm) or (jim), in which i, j, k, m
are distinct. It follows from

(ijk)−1 (ijm)−1 (ijk)(ijm) = (ij)(km) ∈ N

and
(ijk)−1 (jim)−1 (ijk)(jim) = (ik)(jm) ∈ N

that [A4 , A4 ] = N. 2
7. It is routine to check that conjugacy is an equivalence relation.
Express an element σ ∈ Sn as a product of disjoint cycles τ1 , . . . , τr
whose lengths are m1 , . . . , mr respectively. Then σ −1 is still a product
of r disjoint cycles of lengths m1 , . . . , mr . Hence σ is conjugate to σ −1 .
If (123) ∈ A4 is conjugate to (132) = (123)−1 , then there exists
τ ∈ A4 such that τ (123)τ −1 = (132). However such a τ can only be
(12), (13) or (23). None of them belongs to A4 . 2
8. Without loss of generality may assume γ = (123 · · · n). Then
βγβ −1 = (β(1)β(2) · · · β(n)). Since βγ = γβ, the equality βγβ −1 = γ
holds. So (β(1)β(2) · · · β(n)) = (123 · · · n), which implies that β(1) =
k, β(2) = k + 1, . . . , β(n − k + 1) = n, β(n − k) = 1, . . . , β(n) = k − 1.
Hence β = γ k . 2
196 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

Exercises 1.7

1. [f (e1 )]2 = f (e21 ) = f (e1 ) implies f (e1 ) = e2 .


e2 = f (e1 ) = f (a−1 a) = f (a−1 )f (a) implies f (a)−1 = f (a−1 ).
Assume that h ∈ Ker(f ), g ∈ G1 . Then f (g −1 hg) = f (g)−1 f (h)f (g) =
f (g)−1 f (g) = e2 . Hence Ker(f )  G1 . 2

2. For any a ∈ G define the map

sa : {1, 2, . . . , n} → {1, 2, . . . , n}

by abi ∈ bsa (i) H. The key point is to show that sa is bijective. Assume
that sa (i) = sa (j). Then abi = bsa (i) hi , abj = bsa (j) hj for some hi , hj ∈
H. Hence
bi h−1 −1 −1
i = a bsa (i) = a bsa (j) = bj hj .
−1

This means that bi , bj are in the same left coset. So i = j. Therefore sa


is bijective.
It remains to verify that τ (a0 a) = τ (a0 )τ (a) for any a0 , a ∈ G. As-
sume that a0 bi = bsa0 (i) h0i , with h0i ∈ H. Then
n
Y
0
τ (a )τ (a) = h0i hi .
i=1

It follows from

(a0 a)bi = a0 bsa (i) hi = bsa0 (sa (i)) h0sa (i) hi

that n
Y
τ (a0 a) = h0sa (i) hi .
i=1

Since sa is a bijection from the set {1, 2, . . . , n} to itself and H is an


abelian group,
Yn n
Y
h0sa (i) hi = h0i hi
i=1 i=1

holds, which implies τ (a0 a) = τ (a0 )τ (a). 2

3. The additive group C is abelian while GL2 (R) is not. 2


197

4. The second isomorphism theorem implies H/(N ∩ H) ∼


= HN/N.
So (H : N ∩ H) = (HN : N ). Since H and N are subgroups between
N ∩ H and HN, as illustrated in the following diagram,

v
HN HH
vv HH
vv HH
vv HH
vv H
H HH vN
HH vv
HH vv
HH vv
H vv
H ∩N

we have

(HN : N ∩ H) = (HN : H)(H : N ∩ H) = (HN : N )(N : N ∩ H).

Hence (HN : H) = (N : N ∩ H). Lagrange’s theorem implies (HN :


H)|(G : H), (N : N ∩ H)||N |. Since (G : H) is coprime with |N |, the
equality (HN : H) = (N : N ∩ H) = 1 holds. Hence N ⊆ H. 2
5. Hint: It is routine to check that G is an abelian group.
Then check that the map f : C∗ → G, z 7→ 1 − z is an isomorphism.
2
6. Since the composite of two maps satisfies the associative law,
so does the binary operation in Aut(G). Let σ ∈ Aut(G). Since σ is a
bijection, its inverse map τ exists, i.e., τ (σ(g)) = g and σ(τ (g)) = g
for any g ∈ G. In order to show that τ is the inverse of σ in Aut(G),
we need to verify that τ is a homomorphism. Let g, g 0 ∈ G. Then

σ(τ (g)τ (g 0 )) = σ(τ (g))σ(τ (g 0 )) = gg 0 .

So τ (gg 0 ) = τ (g)τ (g 0 ), which implies that τ is a homomorphism. 2


7. Let σ ∈ Aut(G) be an inner automorphism, i.e., there exists
a ∈ G such that σ(g) = a−1 ga for any g ∈ G. Let τ ∈ Aut(G). Then

τ −1 στ (g) = τ −1 (a−1 τ (g)a) = τ −1 (a)−1 gτ −1 (a)

for any g ∈ G. So τ −1 στ ∈ Inn(G). This means that Inn(G)  Aut(G).


198 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

8. Construct a map f : G → Inn(G) in the following way: Every


a ∈ G determines an inner automorphism g 7→ aga−1 . Define this inner
automorphism to be f (a). It is obvious that f is surjective.

f (aa0 )(g) = (aa0 )g(aa0 )−1 = a(a0 ga0−1 )a−1 = f (a)[f (a0 )(g)]

holds for any a, a0 ∈ G. Hence f (aa0 ) = f (a)f (a0 ). So f is a surjective


homomorphism. Since

a ∈ Ker(f ) ⇔ f (a) is the identity map ⇔ aga−1 = g∀g ∈ G ⇔ a ∈ C(G),

Inn(G) ∼
= G/C(G) holds by the fundamental theorem of homomor-
phism. 2

9. May assume that G = Z, which has exact two generators 1 and


−1. Since an automorphism of Z is totally determined by the image of
the generator 1, there are at most two automorphisms. Since n 7→ −n
is a nontrivial automorphism, we have |Aut(G)| = 2 2

10. For every nonzero rational number r, define

f : Q → Q, a 7→ ra.

It is easy to see that f ∈ Aut(Q).


Assume that g ∈ Aut(Q). Then r = g(1) is a nonzero rational
number. Since g is a group homomorphism, the equality

g(n) = g(1 + 1 + · · · + 1) = g(1) + g(1) + · · · + g(1) = ng(1) = nr

holds for any natural number n. Moreover, g(n) = nr holds for any
n ∈ Z. Let n/m ∈ Q, in which m, n ∈ Z, m > 0. Then
 n n
rn = g(n) = g m = mg .
m m
n n
Hence g( m ) = rm , which implies that g(a) = ra for any a ∈ Q. It
follows that there is one to one correspondence between Aut(Q) and
199

the set Q∗ of nonzero rational numbers. Furthermore it is easy to show


that Aut(Q) is isomorphic to the multiplicative group Q∗ . 2

11. N ∩ H = {e} implies H ∼ = H/(H ∩ N ). The second isomor-


phism theorem implies H/(H ∩ N ) ∼= N H/N = G/N. Hence every
complement subgroup of N is isomorphic to G/N. 2

12. Hint: First show that S = {r ∈ Q|nr ∈ Z} is a cyclic subgroup


of G of order n.
Next show that S is the unique subgroup of order n by using Propo-
sition 1.7.12.

13. Suppose that H  G. Then G/H contains a nontrivial subgroup


which corresponds with a nontrivial subgroup of G containing H but
not equal to H, contradicting to the hypothesis that H is a maximal
subgroup of G. 2

Exercises 1.8

1. Hint: Write out four normal subgroups of G1 and show that G


is a simple group.
It G is abelian, then every subgroup of G1 is normal. In particular
the subgroup generated by any element (a, a) with a ∈ G is a normal
subgroup.

2. Assume that f is surjective. Then there exists a map (not


required to be a homomorphism) h : G → G such that f ◦ h = 1. Here
1 stands for the identity map. Since

u ◦ f = u − u ◦ v ◦ u = g ◦ u,

the equality u = g ◦ u ◦ h holds. Then 1 = g + u ◦ v implies

1 = g + g ◦ u ◦ h ◦ v = g ◦ (1 + u ◦ h ◦ v),

Hence g is surjective.
By symmetry, the surjectivity of g implies that of f. 2

3. Let G = Z and H = 2Z. The G is not isomorphic to the direct


product of H and Z/2Z.
200 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

Assume that Ḡ is an infinite cyclic group with generator ā with


some a ∈ G. Define a map f : H × Ḡ → G by f (h, ān ) = han . Let
0 0 0
(h0 , ān ) be another element of H × Ḡ. Then f (hh0 , ān ān ) = hh0 an+n =
0
f (h, ān )f (h0 , ān ), for Ḡ is an infinite cyclic group. Hence f is a ho-
momorphism. Let (h, ān ) ∈ Ker(f ). Then han = 1. Hence an ∈ H, so
ān = 1, which implies n = 0, h = 1. Hence f is a monomorphism. Since
every element of G belongs to a coset of H f is surjective. 2
4. Hint: Every element of Ni commutes with every element of Nj for
i 6= j. In order to show that G is abelian, it suffices to show that every
Ni is abelian. Assume that a, b ∈ N1 . Then b = uv, u ∈ N2 , v ∈ N3 .
Hence ab = auv = uva = ba.
G = N1 × N2 implies N2 ∼ = G/N1 . For the same reason, N3 ∼ = G/N1 .
5. i) Let g = (a, e) ∈ S ∩ H. Express any k ∈ G as k = th, (t ∈
T, h ∈ H. Then t = (e, b). Hence k −1 gk = h−1 gh ∈ S ∩ H. Hence
S ∩ H  G. By symmetry, T ∩ H  G too.
ii) Construct a homomorphism f : H 7→ T, (a, b) 7→ b. Since G =
SH, f is surjective. The condition S ∩ H = {e} implies that f is
injective. Hence H ∼ = T. For the same reason H ∼ = S.
iii) Let s1 , s2 ∈ S. Then there exists t1 ∈ T such that (s1 , t1 ) ∈ H.
It follows from H  G that (s−1 −1
2 s1 s2 , t1 ) = (s2 , e) (s1 , t1 )(s2 , e)
−1
∈ H.
−1 −1 −1 −1 −1 −1
Hence (s1 s2 s1 s2 , e) = (s1 , t1 ) (s2 s1 s2 , t1 ) ∈ S ∩H. So s1 s2 s1 s2 =
e and S is abelian. For the same reason , T is abelian too. 2

Exercises 1.10
1. ex = x ⇒ x ∼ x.
x ∼ y ⇒ gx = y ⇒ g −1 y = x ⇒ y ∼ x.
x ∼ y, y ∼ z ⇒ gx = y, g 0 y = z ⇒ (g 0 g)x = z ⇒ x ∼ z
Let [x] be the equivalence class represented by x ∈ S. For any y ∈ [x]
there exists g ∈ G such that gx = y, so y ∈ Gx. Hence [x] ⊆ Gx.
Conversely z ∈ Gx implies z = gx for some g ∈ G. Hence Gx ⊆ [x].
Therefore [x] = Gx. 2
2. Since every conjugacy class C(a) is an orbit under the conjugacy
action of G on itself, the order of G is the product of the length of C(a)
and the order of the stabilizer of a. This proves 1).
201

2) Since C(1G ) contains only one element 1, so all conjugacy classes


contain same number of elements if and only if G is abelian.
3) Since {1G } is a conjugacy class, G − {1G } must be the other
conjugacy class. Since (|G| − 1)||G| by the result of 1), |G| is equal to
2. 2
3. Two square matrices are conjugate if and only if they are similar.
If P −1 AP = B for some P ∈ GL2 (C) then there exists Q ∈ SL2 (C)
such that Q−1 AQ = B. In fact, Q can be taken to be
"p #
|P | 0
Q= p P
0 |P |

where |P | stands for the determinant of P. Hence it suffices to enumer-


ate all Jordan canonical forms of determinant one. They are
" #
a 0
,
0 a−1
where a is a nonzero complex number, and
" #
1 1
0 1
or " #
−1 1
.
0 −1

4. Let T = {(a1 , a2 )|a1 , a2 ∈ S, a1 6= a2 }. Then |T | = n(n − 1). The


group acts on T by g(a1 , a2 ) = (ga1 , ga2 ).
Let (x1 , y1 ), (x2 , y2 ) be two distinct elements of T. Then there exists
g ∈ G such that g(x1 , y1 ) = (x2 , y2 ). Hence the action of G on T is
transitive. Therefore |T | divides |G|. 2
5. Suppose that G = Stab(x)Stab(y). Choose g ∈ G such that
g(y) = x. Assume that g = hk, h ∈ Stab(x), k ∈ Stab(y). Then h(y) =
hk(y) = g(y) = x, contradicting h(x) = x. 2
6. It is easy to see that S4 is an orbit.
Let (5i1 · · · in ) be an n+1-cycle. If n > 1 then (i1 · · · in )−1 (5i1 · · · in ) =
202 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

(5in ). In this case (5i1 · · · in ) and the transposition (5in ) are in the same
orbit.
Assume that 1 ≤ i < j ≤ 4. Then (ij)(5i)(ij) = (5j). Hence all
transpositions in the form (5i) are in the same orbit.
Let g, h ∈ S5 \S4 . Then g = (5i1 · · · in )g1 , h = (5j1 · · · jm )h1 , in
which g1 , h1 ∈ S4 . There exist p, q ∈ S4 such that p(5i1 · · · in )q −1 =
(5j1 . . . jm ). Let r = g1−1 q −1 h1 . Then pgr = h.
In summary, there are only two orbits S4 and S5 \S4 with lengths
24 and 96 respectively. 2

Exercises 2.1

1. If e, e0 are unities, then e = ee0 = e0 .


2) 0 · a + 0 · a = (0 + 0) · a = 0 · a implies 0 · a = 0. a · 0 = 0 is proved
in a similar way.
3) (−a) · b + ab = (−a + a)b = 0 · b = 0 implies (−a) · b = −(ab).
a · (−b) = −(ab) is proved in a similar way. 2

2. Let b, b0 be the inverses of an invertible element a. Then b =


b · 1 = b(ab0 ) = (ba)b0 = 1 · b0 = b0 . 2

3. (u + a) n−1 i n−1−i
= un − (−a)n = un implies that
P
i=0 u (−a)
u−n n−1 i n−1−i
is the inverse of u + a. 2
P
i=1 u (−a)

4. It is obvious that A is an abelian group under addition. The


equality
X n
[(f ∗ g) ∗ h](n) = (f ∗ g)(d)h
d
d|n
   
XX d n
= f (e)g h
e d
d|n e|d
X
= f (r)g(s)h(t)
rst=n
= [f ∗ (g ∗ h)](n).

holds for any f, g, h ∈ A. Hence the multiplication of A is associative.


203

The equalities
X n
[(f + g) ∗ h](n) = (f (d) + g(d))h
d
d|n
X n X n
= f (d)h + g(d)h
d d
d|n d|n

= (f ∗ h)(n) + (g ∗ h)(n).

and f ∗ g = g ∗ f for all f, g ∈ A imply that A satisfies the law of


commutativity for multiplication and the law of distributivity.
Let (
1, n = 1
u(n) = .
0, n > 1
Then (u ∗ f )(n) = f (n) for any f ∈ A, which implies that u is the unity
of A. Hence A is a commutative ring. 2

5. Let F, E be two fields. Then (1F , 0E )(0F , 1E ) = (0F , 0E ). Hence


F × E is not an integral domain, far from being a field. 2

6. 2x = (2x)2 = 4x2 = 4x for any x ∈ A. Hence 2x = 0. 2

Exercises 2.2
T
1. It is known from Chapter 1 that λ∈Λ Iλ is a subgroup of R
T
as an additive group. Let a ∈ λ∈Λ Iλ , b ∈ R. Since Iλ is an ideal
for every λ ∈ Λ and a ∈ Iλ , so ab, ba ∈ Iλ for every λ ∈ Λ. Hence
T T
ab, ba ∈ λ∈Λ Iλ . This shows that λ∈Λ Iλ is an ideal.
Let I and J be ideals of R. Then I + J is an additive subgroup of
R. Since c(a + b) = ca + cb ∈ I + J, (a + b)c = ac + bc ∈ I + J for any
c ∈ R, a ∈ I, b ∈ J, I + J is an ideal of R.
Assume that ni ai bi ∈ IJ(ai ∈ I, bi ∈ J) and c ∈ R. Then cai ∈
P

I, bi c ∈ J for every i. Hence


n
! n
X X
c ai bi = (cai )bi ∈ IJ,
i i

n
! n
X X
ai b i c = ai (bi c) ∈ IJ.
i i
204 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

Hence IJ is an ideal of R.
2. Hint: The three elements
" # " # " #
0 1 1 0 0 0
, ,
0 0 0 0 0 1

generate three nontrivial principal ideals


(" # )
0 b
I1 = |b ∈ R ,
0 0
(" # )
a b
I2 = |a, b ∈ R ,
0 0
(" # )
0 b
I3 = |b, c ∈ R
0 c
respectively.
Let I be a nonzero ideal. Then I contains a nonzero element
" #
a b
.
0 c

By discussing the cases a 6= 0, c 6= 0 and a = c = 0 one can see that I


is one of I1 , I2 , I3 and A. Hence there are five ideals.
2
3. Hint: Let I be a nonzero ideal. Need to show I = Mn×n (F ). Let
A = (aij ) be a nonzero element in I. Then aij 6= 0 for some i, j. By
multiplication of suitable elementary matrices from left and(or) from
the right one can shows that I contains all Ers , where Ers is the matrix
whose (r, s) entry is 1 and all other entries are zero. 2
4. Assume that k = pn s, in which n > 1 and (p, s) = 1. Let a = ps.
Then a 6= 0 and an = 0.
Assume that k = p1 · · · pm where p1 , . . . , pm are distinct prime num-
bers. Then ān = 0 (with a ∈ Z) implies pi |an for all i. Then pi |a for all
i and so ā = 0̄. 2
5. Since Ii + Ij = A for any i 6= j, there exist aj ∈ Ii , cj ∈ Ij such
205
Q
that aj + cj = 1. Let bi = j6=i cj . Then

bi ≡ 0 (mod Ij )

for all j 6= i and


Y
bi = (1 − aj ) ≡ 1 (mod Ii ).
j6=i

Let b1 , . . . , bn be the elements satisfying the condition in 1). Then


b = a1 b1 + · · · + an bn satisfies the required conditions. 2

6. Let a, b ∈ ∪∞i=0 Ii . There exists n > 0 such that a, b ∈ In . Hence


a + b ∈ In ⊆ ∪i=0 Ii . Moreover ac, ca ∈ In ⊆ ∪∞ i=0 Ii for any c ∈ In . 2

7. The greatest common divisor 6 of 12, 48, 30. 2

8. Since I is a left ideal, a(bc − cb) ∈ I for any a, b, c ∈ R. It follows


from
(ab − ba)c = a(bc − cb) + (ac)b − b(ac) ∈ I

that I is an ideal. 2

Exercises 2.3

1. Let b ∈ Ker(f ). Then b ∈ I, i.e., b = (ax − 1)g(x) for some

g(x) = cm xm + cm−1 xm−1 + · · · + c1 x + c0

with ci ∈ R. Hence

acm xm+1 + (acm−1 − cm )xm + · · · + (ac0 − c1 ) − c0 = b

holds. By compare the coefficients of x of both sides we obtain

acm = 0,

cm = acm−1 ,

···

c1 = ac0 ,
206 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

b = −c0 .

Therefore

am+1 b = −am+1 c0 = −am c1 = · · · = −a2 cm−1 = −acm = 0.

Conversely assume that an b = 0. Then

−(ax − 1)(an−1 xn−1 + an−2 xn−2 + · · · + ax + 1)b = b.

Hence b ∈ Ker(f ). 2

2. Let a, b ∈ R. Then

f (a+b) = (a+b+I, a+b+J) = (a+I, a+J)+(b+I, b+J) = f (a)+f (b),

f (ab) = (ab + I, ab + J) = (a + I, a + J)(b + I, b + J) = f (a)f (b).

Hence f is a homomorphism.
Assume that a ∈ I ∩ J. Then a + I = I, a + J = J, so a ∈ Ker(f ).
Conversely a ∈ Ker(f ) implies a ∈ I, a ∈ J, which implies a ∈ I ∩ J.
Hence Ker(f ) = I ∩ J.
Assume that I +J = R. There exist u ∈ I, v ∈ J such that u+v = 1.
The equality

f (bu + av) = (bu + av + I, bu + av + J) = (av + I, bu + J) = (a + I, b + J)

holds for any a, b ∈ R. Hence f is an epimorphism. Conversely assume


that f is surjective. Then there exists a ∈ R such that f (a) = (1+I, J),
i.e., a ∈ J, 1 − a ∈ I. Hence 1 = (1 − a) + a ∈ I + J, which implies
I + J = R. 2

3. Assume that f (x, y, z) ∈ Ker(φ). There exists q(x, y, z) ∈


F [x, y, z] and r(x, y) ∈ F [x, y] such that f (x, y, z) = q(x, y, z)(z −
x2 ) + r(x, y). There also exist q1 (x, y) ∈ F [x, y] and u(x), v(x) ∈ F [x]
such that r(x, y) = q1 (x, y)(y 2 − x3 ) + u(x)y + v(x). It follows from
f (t2 , t3 , t4 ) = 0 that u(t2 )t3 + v(t2 ) = 0. Since u(t2 )t3 has only terms
of odd degrees and v(t2 ) has terms of even degrees, u(x) = v(x) = 0.
Hence f is contained in the ideal generated by y 2 − x3 and z − x2 . 2
207

Exercises 2.4
1. 0 is a maximal ideal of the simple ring M2 (R) but M2 (R) is not
a field. 2
2. Let I = R × 0, J = 0 × R. Then I, J are ideals of R × R. Since
R × R/I ∼= R × R/J ∼ = R, both ideals I and J are maximal ideals.
Let K be a maximal ideal of R × R. Since 0 is not a maximal ideal,
K contains a nonzero element (a, b). If a 6= 0, b 6= 0, then K = R × R,
contradicting the assumption that K is a maximal ideal. Hence either
a = 0 or b = 0. It follows that K is either I or J. 2
3. Let p(x, y) = x + y 2 , q(x, y) = y + x2 + 2xy 2 + y 4 and let I
be the ideal generated by p(x, y) and q(x, y). Every element f (x, y) in
I can be written as f (x, y) = u(x, y)p(x, y) + v(x, y)q(x, y) for some
u(x, y), v(x, y). Hence f (0, 0) = 0 for every f (x, y) ∈ I. Hence 1 ∈
/ I
and I is a proper ideal.
Since

y = q(x, y) − p(x, y)2 ∈ I, x = p(x, y) − y 2 ∈ I

and x, y generate a maximal ideal, so I is a maximal ideal. 2


4. The characteristic of k cannot be zero, otherwise (1 + 1)n = 2n 6=
2 = 1n + 1n . So the characteristic of k is a prime number p. Write n as
n = pe m, in which m is not divisible by p. Suppose that m > 1. Let
e e (m−1)
f (x) = (x + 1)n − xn − 1 = (xp + 1)m − xn − 1 = mxp + g(x),

in which deg(g) < pe (m − 1). Since the degree of f (x) is pe (m − 1)


and k is an infinite field, there exists a ∈ k such that f (a) 6= 0, which
implies (a + 1)n 6= an + 1n , contradicting the hypothesis of the problem.
2
5. 1) " #
a −b
7 a + bi.

b a

2) Any P ∈ GL2 (R) induces an automorphism M 7→ P −1 M P of the


ring M2 (R). It suffices to find some P ∈ GL2 (R) such that P −1 AP ∈ S.
208 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

We may try " #


1 y
P = .
0 x
Then
" #" #" # " #
4y 2
1 − xy 0 3 1 y 4y
3x − y + x
P −1 AP = = x4 .
0 1
x
−4 1 0 x −x 1 − 4y
x

The condition P −1 AP ∈ S yields the system of equations

4y 4y
=1− ,
x x
4y 2 4
3x − y +
= .
x x
√ √
One solution of this system is {x = 8/ 47, y = 1/ 47}.
3) is a consequence of the fundamental theorem of algebra. 2

6. 1) Let P be a prime ideal. Assume that (a + P )(b + P ) = P.


Then ab ∈ P, so a ∈ P or b ∈ P, which implies that either a + P or
b + P is zero in A/P. Hence A/P is an integral domain.
Conversely assume that A/P is an integral domain. Assume that
a∈ / P, b ∈
/ P. Then a + P, b + P are nonzero elements in the quotient
ring A/P. Hence (a + P )(b + P ) 6= P. This implies that ab ∈/ P. Hence
P is a prime ideal.
2) Let I be a maximal ideal of A. Then A/I is a field, which is an
integral domain. It follows from 1) that I is a prime ideal. 2

Exercises 3.1

1. 1) existence: If deg(f ) < deg(g), then q = 0, r = f will work.


Assume that deg(f ) ≥ deg(g). Apply induction on deg(f ). Let axn
and bxm be the leading terms of f and g respectively. Let h = f −
ab−1 xn−m g. Then deg(h) < deg(f ). By induction hypothesis there exist
q1 , r ∈ A[x] such that h = q1 g + r, deg(r) < deg(g). Let q = q1 +
ab−1 xn−m . Then f = qg + r.

2) uniqueness: Assume that q1 , q2 , r1 , r2 ∈ A[x] satisfy f = q1 g +


r1 = q2 g + r2 , where deg(r1 ) < deg(g), deg(r2 ) < deg(g). Then (q1 −
209

q2 )g = r2 − r1 . Hence deg(r2 − r1 ) = deg(q1 − q2 ) + deg(g). It follows


from deg(r2 − r1 ) < deg(g) that r2 − r1 = q1 − q2 = 0. 2
2.

f (x) = 2x2 g(x) + x3 + 2x2 + 2x + 1


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

Hence f (x) = (2x2 + x + 1)g(x) + 2x + 2. 2


3. The necessity is the consequence of Theorem 3.2.10. Assume
that f (x)u(x) + v(x)g(x) = 1. If h(x) is a common divisor of f (x) and
g(x), then h(x) divides f (x)u(x)+v(x)g(x), i.e., h(x)|1. So deg(h) = 0.
Hence f (x) and g(x) are coprime. 2
4. 1) Assume that b/a = s + ti, in which s, t ∈ Q. There exist
m, n ∈ Z such that |s − m| ≤ 1/2, |t − n| ≤ 1/2. From
p
|b/a − (m + in)| = |(s − m) + i(t − n)| = (s − m)2 + (t − n)2 < 1

it follows that N (b − a(m + in)) < N (a).


2) Let I be an ideal of Z[i]. Since the ideal {0} is a principal ideal, we
need to consider nonzero ideals only, i.e., we may assume that I 6= {0}.
Choose a nonzero element g in I\{0} such that N (g) reaches the
minimum value. Then N (g) > 0. For any f ∈ I, there exist q, r ∈ Z[i]
such that f = qg + r and N (r) < N (g) by the result of 1). Since
r = f − qg ∈ I, r can only be 0 by the minimality of N (g). Hence
f ∈ qg, which implies that f is in the ideal generated by g. Therefore
I is the principal ideal generated by g.
3)
18 + i = (11 + 7i) + (7 − 6i),

11 + 7i = i(7 − 6i) + 5,

7 − 6i = (1 − i)5 + (2 − i),

5 = (2 − i)(2 + i).

Hence gcd(11 + 7i, 18 + i) = 2 − i.


210 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

Another method: N (18 + i) = 325 and N (11 + 7i) = 170 imply


that gcd(N (18 + i), N (11 + 7i)) = 5. If h = gcd(11 + 7i, 18 + i), then
N (h) = 1 or 5. If 18 + i and 11 + 7i are not coprime, then h is one of
2 + i, 2 − i, −2 + i, −2 − i. By checking these four possible solution one
obtains h = 2 − i or −2 + i. 2

5. Let I be the ideal generated by 2 and x. It consists of all elements


in Z[x] whose constant term is even. So I is a nontrivial ideal. Suppose
that I is a principal ideal generated by f (x) ∈ Z[x]. Then there are
g(x), h(x) ∈ Z[x] such that 2 = f (x)g(x), x = f (x)h(x). The equality
2 = f (x)g(x) implies that f (x) is equal to 2 or −2, contradicting the
equality x = f (x)h(x). 2

6. Construct an epimorphism φ : Q[x] → Q×Q, f 7→ (f (2), f (−2)).


Then Ker(φ) = (x2 − 4). Hence Q[x]/(x2 − 4) ∼
= Q × Q. 2

Exercises 3.4

1.

x9 − x = x(x − 1)(x + 1)(x6 + x4 + x2 + 1)


= x(x − 1)(x + 1)(x2 + 1)(x4 + 1).

Let g(x) = x2 + 1. Since g(1) = g(2) = 2 6= 0, g(x) is irreducible.


Let f (x) = x4 + 1. Since f (1) = f (2) 6= 0, f (x) has not linear factors.
Suppose that f (x) = (x2 + ax + b)(x2 + cx + d). Then

bd = 1,

a + c = 0,

b + d + ac = 0,

ad + bc = 0.

The equalities a + c = 0 and ad + bc = 0 imply c(b − d) = 0. If c = 0,


then b = −d, and b2 = −1 due to bd = 1. But there is no element b ∈ F3
such that b2 = −1. Hence c 6= 0, b = d, a = −c. Since b + d + ac = 0, so
211

2b − 1 = 0, which implies b = 2. It follows that

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

Therefore

x9 − x = x(x − 1)(x + 1)(x2 + 1)(x2 + x + 2)(x2 + 2x + 2)

is the irreducible decomposition of x9 − x. 2

2. First proof: let x3 + αx + 1 = (x + a)(x2 + bx + c). Then

a + b = 0, c + ab = α, ac = 1.

Hence
1 − a3 = aα.

It is obvious that a 6= 0. Since F ∗ = F − {0} is a group of order 3


under the multiplication, a3 = 1 for any a ∈ F ∗ by Lagrange’s theorem.
Hence 1 − a3 = aα does not hold.

Second proof: since the characteristic of F is equal to 2, the equality


a + a = 0 holds for any a ∈ F. Denote f (x) = x3 + αx + 1. Then f (0) =
1 6= 0. Assume that b ∈ F \{0}. Then b3 = 1. Hence f (b) = αb 6= 0. So
f (a) 6= 0 for any a ∈ F. 2
√ √
3. 1) Let m + n 5i, m0 + n0 5i ∈ A. Then
√ √ √
(m + n 5i) + (m0 + n0 5i) = (m + m0 ) + (n + n0 ) 5i ∈ A,
√ √ √
(m + n 5i)(m0 + n0 5i) = (mm0 − 5nn0 ) + (mn0 + nm0 ) 5i ∈ A.

Hence A is a subring of C.
√ √
2) Assume that 3 = (m + n 5i)(m0 + n0 5i). Then

9 = (m2 + 5n2 )(m02 + 5n02 ).

It follows that m2 +5n2 = 3, m02 +5n02 = 3 or m2 +5n2 = 1, m02 +5n02 =


9 or m2 + 5n2 = 9, m02 + 5n02 = 1. It is easy to see that m2 + 5n2 =
3 has no integral solution. The integral solutions of m2 + 5n2 = 9
212 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

are m = ±3, n = 0 and the integral solutions of m02 + 5n02 = 1 are


√ √
m0 = ±1, n0 = 0. Hence one of m + n 5i and m0 + n0 5i is equal to
±1. Hence 3 is irreducible. It can be shown by the same method that
√ √
2 + 5i and 2 − 5i are irreducible too. 2

3) This is because (2 + 5i)/3 ∈ / A.
√ √
4) Because 9 = 3 · 3 = (2 + 5i)(2 − 5i). 2

Exercises 3.5

1. Let h(x) = f (x) − g(x). If f (a) = g(a) for all a ∈ F, the h(a) = 0
for all a ∈ F. If h(x) 6= 0, then it has only finitely many zeros in F.
Hence h(x) = 0. 2

2. First proof: direct verification.


Let f (x), g(x) ∈ I. Then f (2) = f 0 (2) = f 00 (2) = g(2) = g 0 (2) =
g 00 (2) = 0, Hence f (2) + g(2) = (f + g)0 (2) = (f + g)00 (2) = 0. So
f (x) + g(x) ∈ I. For any h(x) ∈ R[x], Since

(h(x)f (x))0 = h0 (x)f (x) + h(x)f 0 (x),

(h(x)f (x))00 = h00 (x)f (x) + 2h0 (x)f 0 (x) + h(x)f 00 (x),

the values of the polynomials h(x)f (x), (h(x)f (x))0 , (h(x)f (x))00 at 2
are all equal to zero. Hence h(x)f (x) ∈ I. Therefore I is an ideal.

Second proof: Since the characteristic of R is zero, f (x) ∈ I if and


only if (x−2)3 |f (x). Hence I is the principal ideal generated by (x−2)3 .

/ J. 2
J is not an ideal, since f (x) = (x − 3)2 − 1 ∈ I, while xf (x) ∈

3. Let f (x) ∈ Fq [x]. Then f (x) = q(x)(xq − x) + r(x), in which


deg(r) < r. Since aq − a = 0 for all a ∈ Fq , f (x) and r(x) represent
the same function on Fq . Hence every polynomial function on Fq is
represented by a polynomial of degree less than q. By Corollary 3.5.4
any two distinct polynomials of degrees less than q represent distinct
functions. It is obvious that q q is the total number of the maps from
Fq to Fq , as well as the number of polynomials of degree less than q. 2

Exercises 4.1

1. Assume that a, b, c ∈ R such that ax3 + b sin(x) + c cos(x) = 0.


213

By substitution x = 0 we obtain c = 0. Hence ax3 + b sin(x) = 0. Let


x = π and a = 0 is obtained. Hence b = 0. 2
2. Let n = dim(V ). Then f is represented by an n × n matrix A
over F under a basis of V. Then f is injective ⇔ the rank of A is equal
to n and f is surjective ⇔ the rank of A is equal to n. Hence f is
injective ⇔ f is surjective. 2
3. For any nonzero element a of R, define a map f : R → R, x 7→ ax.
Since f is a linear transform and since R is an integral domain, f is
injective. Hence f is surjctive by the previous exercise. Hence there
exists b ∈ R such that f (b) = 1, which means that ab = 1. 2
Exercises 4.2
1. First proof: Every element f ∈ Aut(G), can be represented by
2 2
an n × n invertible matrix A over Fp . If Ap = I, then (A − I)p = 0,
so all eigenvalues of A − I are equal to zero, which implies that A − I
is similar to an upper triangular matrix with zeros on the diagonal.
Hence (A − I)p = 0, i.e., Ap = I. Therefore there is no element of order
p2 in Aut(G).
Second proof: Since

|Aut(G)| = |GL2 (Fp )| = (p2 − 1)(p2 − p)

is not divisible by p2 , there is no element in Aut(G) of degree p2 by


Lagrange’s theorem. 2
Exercises 5.1
1. Let Γ be the collection of all subsets of {1, 2, . . . , n} containing
two elements. Then G acts on Γ by g{i, j} = {g(i), g(j)}. Since m is
odd, the number of elements 2m

2
= m(2m − 1) of Γ is also odd.
r
Since |G| = 2 , the length of every orbit is a power of 2. Since |Γ|
is odd, there is at least one orbit whose length is equal to 1. 2
2. 1) is obvious.
2) Let G be an arbitrary finite group. For any a ∈ G, let f (a) ∈
S(G) be the map carrying g ∈ G to ag. Then f determines a map
from G to S(G). It is easy to verify the f is a monomorphism. Hence
214 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

G is a subgroup of S(G). It follows from 1) that G is isomorphic to a


subgroup of Sn . 2
3. The group G acts on the set S = {1, 2, . . . , n} in a natural way.
Let i1 , . . . , ir be the complete system of representatives of the orbits.
Then r
X |G|
n= .
j=1
|Stab(ij )|
|G|
Since |G| is a power of 2, as long as Stab(ij ) 6= G, the number |Stab(ij )|
is even. Since n is odd, Stab(ij ) = G for some j. 2
Exercises 5.2
1. Since

|GLn (Fp )| = (pn − 1)(pn − p)(pn − p2 ) · · · (pn − pn−1 ),

|U | = pn(n−1)/2 ,

the index (GLn (Fp ) : U ) = (pn − 1)(pn−1 − 1) · · · (p − 1) is not divisible


by p. Hence U is a Sylow p-subgroup of GLn (Fp ).
Let Γ be the collection of all Sylow p-subgroups of GLn (Fp ). By
Sylow’s theorem the action of GLn (Fp ) on Γ is transitive. Hence |Γ| =
(GLn (Fp ) : Stab(U )). Since Stab(U ) consists of all nonsingular upper
triangular matrices (see the proof below), so

|Stab(U )| = (p − 1)n |U |.

Therefore

|Γ| = (p + 1)(p2 + p + 1) · · · (pn−1 + pn−2 + · · · + p + 1).

Finally show that Stab(U ) is the set of all nonsingular upper trian-
gular matrices. It is obvious that every nonsingular upper triangular
matrix is in Stab(U ). Assume that P = (pij )1≤i,j≤n ∈ Stab(U ) is not
upper triangular, then there exist 1 ≤ s < r ≤ n such that prs 6= 0. Let
s be the minimal index such that prs 6= 0 for some s < r. Then pij = 0
for all 1 ≤ j < s, j < i ≤ n. Let M be a matrix whose diagonal entries
are equal to 1 and the (s, r)-entry is equal to 1 and the remaining en-
215

tries are zero. Then M ∈ U. Let N = P −1 M P. Suppose that N ∈ U,


then the (s, s) entries of P N and M P are pss and pss + prs respectively.
This would contradict P N = M P. Hence N = P −1 M P ∈ / U, which
means that P ∈ / Stab(U ). 2
2. Suppose that H 6= G. Let n = |G|, m = |H|. Let S be the
collection of all subgroups of G conjugate to H. Then G acts transitively
on S by conjugation. Since H is contained in the stabilizer of H, the
inequality |Stab(H)| ≥ m holds. hence |S| ≤ n/m. Since H ∩ gHg −1 6=
∅ for every g ∈ G, the inequality |∪K∈S K| < m(n/m) = n holds. Hence
there exists g ∈ G − ∪K∈S K, which implies that g is not conjugate to
any element in H. 2
3. 1225 = 52 · 72 . Let N be the number of Sylow 7-subgroups. Then
N ≡ 1 (mod 7) and N |52 . Since the remainder of 5 and 25 divided
by 7 are not equal to 1, the equality N = 1 holds. Hence the Sylow
7-subgroup is normal.
The number r of Sylow 5-subgroups divides 72 and r ≡ 1 (mod 5).
Hence r = 1. Hence the Sylow 5-subgroup is normal.
Since the intersection of the Sylow 7-subgroup and the Sylow 5-
subgroup is {1}, a group of order 1225 is isomorphic to the direct
product of its Sylow 7-subgroup and its Sylow 5-subgroup. Hence this
group is abelian, because every group whose order is the square of a
prime number is abelian by Example 5.1.5. 2
4. Let G be a group of order 640 = 27 5. Let n be the number of
its Sylow 2-subgroups. Then n|5. Hence n = 1 or 5. If n = 1 then the
Sylow 2-subgroup is nontrivial normal subgroup of G.
Assume that n = 5. Let P be a Sylow 2-subgroup. Let M be
the collection of all left cosets of P in G. Then |M | = (G : P ) = 5.
The group G acts on M by left multiplication. This action gives a
homomorphism φ : G → S5 . It is obvious that |Im(φ)| > 1. Since
(G : Ker(φ)) = |Im(φ)| divides |S5 | = 5! = 23 · 3 · 5, so |Ker(φ)| > 1.
Hence G contains a nontrivial normal subgroup Ker(φ). 2
5. 168 = 23 · 3 · 7. Let N be the number of Sylow 7-subgroups of G.
Since G is simple, N > 1. The condition N ≡ 1 (mod 7), N |48 implies
that N = 8. Hence |S| = 8.
216 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

Let H 0 be a Sylow 7-subgroup different from H. Then H 6⊆ NG (H 0 ),


because H 0 is the unique Sylow 7-subgroup of NG (H 0 ). Hence there
exists a ∈ H, such that aH 0 a−1 6= H 0 . It follows that {H} is the only
orbit of length one of S under the action of H. Since the length of every
other orbit is divisible by 7 and |S| = 8, there are only two orbits with
lengths 1 and 7 respectively. 2
6. |S4 | = 24 = 23 3. S4 has three Sylow 2-subgroups. They are

P1 = {id, (12)(34), (13)(24), (14)(23), (12), (34), (1423), (1324)},

P2 = {id, (12)(34), (13)(24), (14)(23), (13), (24), (1432), (1234)},

P3 = {id, (12)(34), (13)(24), (14)(23), (14), (23), (1342), (1243)}.

7. 1) Sylow’s theorem implies (G : N ) = 50. Let H be a subgroup


between N and G. Then (H : N )|50. Assume that H 6= G. Then
(H : N ) < 50. So (H : N ) is one of 1, 2, 5, 10, 25. Since P is also
a Sylow 7-subgroup of H with NH (P ) = N, we have (H : N ) ≡ 1
(mod 7), which implies (H : N ) = 1. Hence N is a maximal subgroup
of G.
2) Let K = NG (Q). Then N ⊆ K. It follows from i) that K = N
or G. It suffices to show that K 6= N. Let R be a Sylow 5-subgroup of
G containing Q. Then (R : Q) > 1. By Proposition 5.1.4 Q is a proper
subgroup of NR (Q). Assume that |Q| = 5r , |NR (Q)| = 5r+s , s > 0.
Since Q is a sylow 5-subgroup of N, |N | = 5r t for some t not divisible
by 5. Hence |NR (Q)| does not divide |N |. So NR (Q) 6⊆ N, which implies
K 6= N. 2
8. Denote by A and B the sets of Sylow p-subgroups of G and G/H
respectively. Then G acts on A by conjugation. On the other hand,
for any g ∈ G, Q ∈ B, let gQ = ḡQḡ −1 , in which ḡ is the element in
G/H represented by g. This defines an action of G on B. By Sylow’s
theorem both actions are transitive.
Let π : G → G/H denote the natural homomorphism. Pick Q ∈ B
and a Sylow p-subgroup P of π −1 (Q). Since (G : π −1 (Q)) = (G/H :
π −1 (Q)/H) = (G/H : Q) is not divisible by p, P is a Sylow p-subgroup
217

of G, i.e., P ∈ A. Since H  G, P H is a subgroup of π −1 (Q). Since P


is a Sylow p-subgroup of both π −1 (Q) and P H, (π −1 (Q) : P H) is not
divisible by p. On the other hand, (π −1 (Q) : P H)|(π −1 (Q) : H) = |Q|
implies that (π −1 (Q) : P H) is a power of p. Hence P H = π −1 (Q), which
implies π(P ) = Q. So StabG (P ) ⊆ StabG (Q) and (G : StabG (Q))| (G :
StabG (P )). Hence the number of Sylow p-subgroups of G/H divides
that of G. 2
Exercises 5.3
1. Suppose that Q is generated by m1 /n1 , . . . , mr /nr in which
m1 , . . . , mr , n1 , . . . , nr ∈ Z. Then every natural number is equal to

a1 m1 /n1 + a2 m2 /n2 + · · · + ar mr /nr

for some a1 , . . . , ar ∈ Z. This would imply that tn1 · · · nr is an integer


for any rational number t, which is absurd. Hence Q is not finitely
generated. 2
2. An element [(n, m)] in A has finite order if and only if there exist
r ∈ N, t ∈ Z such tht rn = 6t, rm = 30t. This is equivalent to the
condition m = 5n. Hence the torsion subgroup of A is a cyclic group
of order 6 generated by [(1, 5)]. Since (1, 5) and (0, 1) generate Z2 and
[(0, 1)] is not an element of finite order, we have A ∼= Z ⊕ (Z/6Z). 2
3. 80 = 24 5. Hence a abelian group of order 80 is the direct sum of
a cyclic group of order 5 and an abelian group of order 24 . The answer
is

Z/24 Z ⊕ Z/5Z,

Z/23 Z ⊕ Z/2Z ⊕ Z/5Z,

Z/22 Z ⊕ Z/22 Z ⊕ Z/5Z,

Z/22 Z ⊕ Z/2Z ⊕ Z/2Z ⊕ Z/5Z,

Z/2Z ⊕ Z/2Z ⊕ Z/2Z ⊕ Z/2Z ⊕ Z/5Z.

4. The Sylow 2-subgroup of A is the direct sum of a cyclic group of


order 8 and two cyclic groups of order 2. The Sylow 3-subgroup of A
218 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

is the direct sum of two cyclic groups of order 3. 2

5. Let A be the set of all (a1 , a2 ) ∈ Q2 such that


" #" #
4 2 b1
[a1 , a2 ] ∈ Z ∀b1 , b2 ∈ Z.
2 4 b2
Since " #−1 " #
4 2 1/3 −1/6
= ,
2 4 −1/6 1/3
A is generated by u = (1/3, −1/6) and v = (−1/6, 1/3). The order of
the cosets A/Z2 represented by u and v are equal to 6. Since |A/Z2 | =
42 − 22 = 12, the quotient group A/Z2 is the direct sum of a cyclic
group of order six and a cyclic group of order 2. 2

6. Suppose that |G| has two distinct prime factors p and q. Let P be
a Sylow p-subgroup of G and let Q be a Sylow q-subgroup of G. Then
P 6⊆ Q and Q 6⊆ P hold. This would contradict the hypothesis of the
problem. Hence |G| = pn for some prime number p. Suppose that G is
not abelian. Then there exist x, y ∈ G, xy 6= yx. So x ∈ / hyi, y ∈
/ hxi,
contradiction. Hence G is abelian. If G were not cyclic, then G =
G1 ⊕ G2 for some nontrivial abelian groups G1 and G2 . Since inside G,
G1 does not contain G2 and G2 does not contain G1 , a contradiction is
reached. Hence G is cyclic. 2

7. All finitely generated abelian groups of rank less than or equal


to 1. 2

Exercises 5.4

1. Apply induction on |G|. When |G| = 1 the proposition is obvi-


ously true. Assume that |G| = n > 1. Choose any maximal normal
subgroup G1 of G. That means that G1  G, G1 6= G, and there is no
other normal subgroups between G1 and G. This implies that G/G1 is
a simple group. Since |G1 | < |G|, there exists a composition series

1 = Gr  Gr−1  · · ·  G2  G1
219

by induction hypothesis. Hence

1 = Gr  Gr−1  · · ·  G2  G1  G

is a composition series of G. 2

2. Let (" # )
1 a
H= a ∈ F .
0 1
Then H  G. Since G/H ∼ = F ∗ × F ∗ , both G/H and H are abelian
groups. Hence G is solvable. 2

Exercises 6.2
√ √ √ √
1. Since [Q( 2) : Q] = 2, [Q( 2, 3) : Q( 2)] = 2, the exten-
√ √ √ √
sion Q( 2, 3)/Q is finite. Hence every element in Q( 2, 3) is an
√ √ √ √
algebraic number. So 2 + 3 ∈ Q( 2, 3) is an algebraic number.
√ √
To find I it suffices to find the minimal polynomials of 2 + 3.
First contruct
√ √ √ √
g(x) = (x − 2− 3)(x − 2+ 3)
√ 2
= (x − 2) − 3
2

= x − 2 2x − 1.

Next construct
√ √
f (x) = (x2 − 1 − 2 2)(x2 − 1 + 2 2)
= (x2 − 1)2 − 8
= x4 − 2x2 − 7.

So I is the principal ideal generated by x4 − 2x2 − 7 2

2. Since x3 − 3x + 4 is irreducible, it is coprime with x2 + x + 1. Use


Euclidean algorithm to find a(x), b(x) ∈ Q[x] such that

(x3 − 3x + 4)a(x) + (x2 + x + 1)b(x) = 1.


220 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

We obtain
3x 8 3x2 5x 17
a(x) = + , b(x) = − − + .
49 49 49 49 49
Hence
3α2 5α 17
(α2 + α + 1)−1 = − − + .
49 49 49
2
3. Since β ∈ F (α), there exist a(x), b(x) ∈ F [x] such that

a(α)
β= .
b(α)

Let f (x) = a(x) − βb(x). Since β ∈ / F, f (x) is a nonzero element in


F (β)[x]. It follows from f (α) = 0 that α is algebraic over F (β). 2
4. Since α2 ∈ F (α), the field F (α2 ) is an intermediate field of the
extension F (α)/F. For the polynomial f (x) = x2 − α2 ∈ F (α2 )[x] the
equality f (α) = 0 holds. Hence [F (α) : F (α2 )] ≤ 2. Since

[F (α) : F (α2 )][F (α2 ) : F ] = [F (α) : F ] = 5,

so [F (α) : F (α2 )] = 1. 2
5. Since both K and F [α] are intermediate fields of K[α]/F,

[K[α] : K][K : F ] = [K[α] : F [α]][F [α] : F ].

Since [F [α] : F ] = deg(p) and [K : F ] are coprime, so deg(p) divides


[K[α] : K]. Since p(α) = 0 implies [K[α] : K] ≤ deg(p), so deg(p) =
[K[α] : K]. Hence p(x) is irreducible over K. 2
Exercises 6.3
1. The kernel of the homomorphism λ : F [x] → E, g(x) 7→ g(α)
is a principal ideal generated by an irreducible polynomial. Since
f (x) ∈ Ker(λ) is irreducible, so f (x) generates Ker(λ). By the fun-
damental theorem of homomorphism λ induces an isomorphism λ̄ :
F [x]/(f (x)) ∼
= F (α).
For the same reason the homomorphism µ : F [x] → E, g(x) 7→ g(β)
221

induces an isomorphism µ̄ : F [x]/(f (x)) ∼


= F (β). The composite µ◦λ−1
is an isomorphism from F [α] to F [β] keeping every element in F fixed.
2

2. Since

02 + 0 + 2 = 2 6= 0, 12 + 1 + 2 = 1 6= 0, 22 + 2 + 2 = 2 6= 0,

the polynomial x2 + x + 2 has no zeros in F3 . Hence x2 + x + 2 is an


irreducible polynomial over F3 . Note that this method works only for
polynomials of degree not exceeding 3.

There are nine elements in E :

0, 1, 2, α, α + 1, α + 2, 2α, 2α + 1, 2α + 2.

Here α is a root of x2 + x + 2 = 0.

The table of addition is

0 1 2 α α+1 α+2 2α 2α+1 2α+2

0 0 1 2 α α+1 α+2 2α 2α+1 2α+2


1 1 2 0 α+1 α+2 α 2α+1 2α+2 2α
2 2 0 1 α +2 α α+1 2α+2 2α 2α+1
α α α+1 α+2 2α 2α+1 2α+2 0 1 2
α+1 α+1 α+2 α 2α+1 2α+2 2α 1 2 0
α+2 α+2 α α+1 2α+2 2α 2α+1 2 0 1
2α 2α 2α+1 2α+2 0 1 2 α α+1 α+2
2α+1 2α+1 2α+2 2α 1 2 0 α+1 α+2 α
2α+2 2α+2 2α 2α+1 2 0 1 α+2 α α+1

The multiplication table (by using α2 = 2α + 1) is


222 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

0 1 2 α α+1 α+2 2α 2α+1 2α+2

0 0 0 0 0 0 0 0 0 0
1 0 1 2 α α+1 α+2 2α 2α+1 2α+2
2 0 2 1 2α 2α+2 2α+1 α α+2 α+1
α 0 α 2α 2α+1 1 α+1 α+2 2α+2 2
α+1 0 α+1 2α+2 1 α+2 2α 2 α 2α+1
α+2 0 α+2 2α+1 α+1 2α 2 2α+2 1 α
2α 0 2α α α+2 2 2α+2 2α+1 α+1 1
2α+1 0 2α+1 α+2 2α+2 α 1 α+1 2 2α
2α+2 0 2α+2 α+1 2 2α+1 α 1 2α α+2
Exercises 6.4
1. Let Γ be the set of all ideals of A disjoint from S. Then Γ is a
partially ordered set under the inclusion relation. Since 0 ∈
/ S, the zero
ideal is a member of Γ, so Γ 6= ∅. Let ∆ be a totally ordered subset of
Γ. Let I = ∪J∈∆ J. Then I ∩ S = ∅. Thus I is an upper bound of ∆.
By Zorn’s lemma, there exists a maximal element P in Γ.
Assume that a ∈ A\P, b ∈ A and ab ∈ P. Let J be the ideal
generated by P and a. Since P ⊆ J and P 6= J, the maximality of
P implies that J = A. So 1 ∈ J, which implies that 1 = ca + p with
c ∈ A and p ∈ P. Multiplying both sides by b yields b = cab + pb ∈ P.
Therefore P is a prime ideal. 2
2. Suppose there is an isomorphism f : Q(x) → Q. Then f (a) = a
for all a ∈ Q. So f (x) is transcendental over Q, leading to contradiction.
2
3. Since x2 + 1 has no zero in F, the field F is not algebraically
closed. Hence F is not isomorphic to Q. 2
4. Since Ē/E and E/F are algebraic extensions, the extension Ē/F
is also algebraic. Since Ē is algebraically closed, Ē is the algebraic
closure of F by definition. 2
Exercises 6.5
1. 1) Let (x, y) ∈ C ∩ D. Then x, y satisfy the system

x2 + y 2 = 1,
223

(x − n)2 + y 2 = n2 .

Subtracting both sides yields x = 1/2n.


2) Double the length of the given line segment and then subdivide
the new line segment into six equal segments.
3) same method.

2. 1) Since (ζ + ζ̄)2 = ζ 2 + 2 + ζ̄ 2 , the equality

(ζ + ζ̄)2 + (ζ + ζ̄) − 1 = 0

holds. Hence √
−1 ± 5
ζ + ζ̄ = .
2
It follows from ζ + ζ̄ > 0 that

−1 + 5
ζ + ζ̄ = .
2

2) Let two points O, A be given. Constructing the regular pentagon


following the following steps:
• Draw a line passing O and A by a ruler.
• Construct a circle C with O as its center and OA as its radius.
• Use standard method to construct a line passing O and perpen-
dicular to OA. It meets the circle C at two points P, P 0 .
• Construct a circle with A as it center and OA as its radius. This
circle meets the extension of the line segment OA at a new point Q.
• Connect the points P, Q by the ruler. The length of the line

segment P Q is equal to 5.
• Construct a circle with Q as its center and OA as its radius. It
intersects the line segment P Q at a point R.
• Use standard method to construct the middle point S of the line

segment P R. Then the length of P S is equal to ( 5 − 1)/2.
• Construct a circle with O as its center and P S as its radius. It
intersects the line segment OA at a point T.
• Use standard method to construct a line passing T and perpen-
dicular to OA. It intersects the circle C at two points U, U 0 .
224 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

• The points A, U, U 0 are three adjacent vertices of a regular pen-


tagon. It is easy to construct the remaining two vertices. 2
Exercises 7.1
1. The number of monic polynomials of degrees 1, 2, 3 are 5, 25, 125
respectively. The number of reducible monic polynomials of degree two
is 5 + 10 = 15. Hence the number of irreducible monic polynomials of
degree two is 25 − 15 = 10.
A reducible monic polynomial of degree three is either the product
of three linear monic polynomials or the product of a linear monic
polynomial and an irreducible quadratic monic polynomial. They can
be easily counted. So the number of irreducible monic polynomials of
degree three is 40. 2
2. Since the degree of an irreducible real polynomial is at most two,
the polynomial x4 + 1 is reducible over R. We have

x4 + 1 = (x − a)(x − a3 )(x − a5 )(x − a7 )

in which √
2
a= (1 + i).
2

Hence x4 + 1 has two real factors (x − a)(x − a7 ) = x2 − 2x + 1 and

(x − a3 )(x − a5 ) = x2 + 2x + 1. Both factors are not in Q[x]. Hence
x4 + 1 is irreducible over Q.
Since the characteristic of F16 is 2, the polynomial

x4 + 1 = (x + 1)4

is reducible over F16 . 2


3.
√ √ √ √
x4 − 4 = (x − 2)(x + 2)(x − i 2)(x + i 2)

and
√ √ √
x3 − 2 = (x − 2)(x − ζ 2)(x − ζ 2 2)
3 3 3

over C, in which √
1 3i
ζ=− + .
2 2
225

√ √
x4 − 4 = (x − 2)(x + 2)(x2 + 2)

and
√ √ √
x3 − 2 = (x − 2)(x2 +
3 3 3
2x + 4)

over R.

x4 − 4 = x4 − 1 = (x − 1)(x + 1)(x2 + 1)

and
x3 − 2 = x3 + 1 = (x + 1)3

over F3 . 2
4. Let A be a finite integral domain. Every nonzero element a ∈ A
determines an injective map

f : A → A, x 7→ ax.

Since A is a finite set, the map f is surjective. Hence there exists


b ∈ A such that ab = 1. This means that every nonzero element of A
is invertible.
The commutative ring F2 × F2 is not a field, but it contains 4 ele-
ments.
2
Exercises 7.2
1. Let f (x) = xp − x − 1 and let α be a zero of f (x) in F̄p . Then

αp = α + 1.

Take p-th power of the both sides repeatedly and we obtain


2
αp = α + 2,
3
αp = α + 3,

···
p−1
αp = α + p − 1,
226 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

p
αp = α.
p i
So α is a zero of the polynomial xp − x. But α is not a zero of xp − x
for any 1 ≤ i ≤ p − 1. This implies that [Fp [α] : Fp ] = p. Hence f (x) is
an irreducible polynomial over Fp . 2

2. Let S = {x2 |x ∈ F }. Since the equation x2 = t has at most two


solutions in F for any t ∈ F and it has a unique solution when t = 0,
so |S| > |F |/2. Let T = a − S = {a − b|b ∈ S}. Then |T | = |S| > |F |/2.
By the pigeon hole principle, S ∩ T 6= ∅. This implies that there are
x, y ∈ F such that x2 = a − y 2 . 2

3. This is because
Y
xp − x = (x − a).
a∈Fp

4. To decide whether there is a linear factor of f (x) = x4 +x3 +x+3


we need to check whether f (x) has a zero in F5 . Since

f (0) = 3, f (1) = 1, f (2) = 4, f (3) = 4, f (4) = 2,

f (x) has no linear factor.


Suppose that x4 + x3 + x + 3 = (x2 + ax + b)(x2 + cx + d). By
comparing coefficients we obtain the relations

a + c = 1,

b + d + ac = 0,

ad + bc = 1,

bd = 3.

The equality c = 1 − a and d = 3/b can be used to eliminate c, d to


obtain
b + 3/b = a2 − a,

3a/b + b(1 − a) = 1.
227

Let u(a) = a2 − a, v(b) = b + 3/b. Then

u(0) = 0, u(1) = 0, u(2) = 2, u(3) = 1, u(4) = 2,

v(1) = 4, v(2) = 1, u(3) = 4, u(4) = 1.

Hence only a = 3, b = 2 or 4 satisfy the condition b + 3/b = a2 − a.


Let g(a, b) = 3a/b + b(1 − a). Then

g(3, 2) = 2 + 1 = 3 6= 1, g(3, 4) = 0 6= 1.

Hence x4 + x3 + x + 3 is not a product of two quadratic polynomials.


Hence it is irreducible. 2
5. Assume that z 6= 0. Let x0 = x/z, y 0 = y/z. Then y 0q + y 0 = x0q+1 .
For any β ∈ Fq2 , let α1 , . . . , αn be all zeros of the equation y 0q +y 0 = β q+1
in F̄q2 . It is easy to check that this equation has no multiple zeros.
Hence n = q.
The equality
2 2 +q
αiq + αiq = (αiq + αi )q = β q = β q+1 = αiq + αi
2
implies αiq − αi = 0. So αi ∈ Fq2 . Hence the number of solutions of the
equation y 0q + y 0 = x0q+1 is equal to q 3 .
When z = 0 the solution is x = 0, z = 0 and y arbitrary. Therefore
the original equation has q 3 + q 2 solutions. 2
Exercises 8.1
1. Let σ ∈ Aut(F2 (x)/F2 (x2 )). Then σ(x)2 = σ(x2 ) = x2 . Hence

(σ(x) − x)2 = σ(x)2 − x2 = 0.

It follows that σ(x) = x, which implies that σ is an identity map. Hence


Aut(F2 (x)/F2 (x2 )) contains only one element. However F2 (x)/F2 (x2 )
is a quadratic extension. Therefore, F2 (x)/F2 (x2 ) is not a Galois ex-
tension. 2

2. Since [F : Q] = 2 and [Q( 3 2) : Q] = 3, the polynomial x3 − 2 is

irreducible over F. So the minimal polynomial of 3 2 over F is x3 − 2.
228 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

It can be decomposed into


√ √ √
x3 − 2 = (x − 2)(x − ζ 2)(x − ζ 2 2)
3 3 3


in F [ 3 2]. Hence it is a Galois extension by Theorem 8.1.15. 2

3. 1) Let K be the fixed subfield of H1 ∩ H2 . Since (H1 ∩ H2 ) ⊆ Hi


for i = 1, 2, so Ei ⊆ K for i = 1, 2. Hence E1 E2 ⊆ K.
Since Ei ⊆ E1 E2 for i = 1, 2, Aut(L/E1 E2 ) ⊆ Aut(L/Ei ) = Hi for
i = 1, 2. Hence Aut(L/E1 E2 ) ⊆ H1 ∩ H2 .
It follows from Proposition 8.1.10 that L/E1 E2 is a Galois exten-
sion. The relation Aut(L/E1 E2 ) ⊆ H1 ∩ H2 implies K ⊆ E1 E2 by the
fundamental theorem of Galois theory.
2) Let K be the fixed field of H1 H2 . Since Hi ⊆ H1 H2 for i = 1, 2,
so K ⊆ Ei for i = 1, 2. Hence K ⊆ E1 ∩ E2 . Since E1 ∩ E2 ⊆ Ei for
i = 1, 2, so Hi = Aut(L/Ei ) ⊆ Aut(L/E1 ∩ E2 ) for i = 1, 2. Hence
H1 H2 ⊆ Aut(L/E1 ∩ E2 ). Therefor E1 ∩ E2 ⊆ K, for E1 ∩ E2 is the
fixed field of Aut(L/E1 ∩ E2 ). 2

4. Let F = Q, E = Q[ 2], K = Q[21/4 ]. Since [E : F ] = [K : E] =
2, the extensions K/E, E/F are Galois extensions. However K/F is
not a Galois extension. 2

5. ⇒: Assume that E/F is a Galois extension, then |Aut(E/F )| =


[E : F ] by definition. By Proposition 8.1.4 the number of embed-
dings from E/F into L does not exceed [E : F ]. Since every element
in Aut(E/F ) is an embedding from E/F into L, the number of em-
beddings of E/F into L is exactly equal to [E : F ]. Hence E/F is a
separable extension. Since the number of embeddings from E/F into
L is equal to Aut(E/F ) and every such embedding carries E into E,
the extension E/F is normal.
⇐: Let σ1 , . . . , σn be the set of all embeddings from E/F into L,
in which n = [E : F ]. Since E/F is normal, σ1 , . . . , σn ∈ Aut(E/F ).
Hence [E : F ] ≤ |Aut(E/F )|. Therefore E/F is a Galois extension. 2

6. The necessity follows from Theorem 8.1.15. The sufficiency is


proved by induction on [E : F ].
If [E : F ] = 1 then there is nothing to be proved. Assume that
229

[E : F ] > 1. Then at least one αi is not in F. Let p(x) ∈ F [x] be its


minimal polynomial. Since p(x)|f (x), p(x) is decomposed into linear
factors in E[x]. Assume that p(x) = (x − α1 ) · · · (x − αr ) without loss of
generality. Let K = F [α1 , . . . , αr ]. Theorem 8.1.15 implies that K/F is
a Galois extension and [E : K] < [E : F ]. By the induction hypothesis
E/K is a Galois extension.
Let f ∈ G(K/F ). Theorem 6.4.8 implies that there exists an em-
bedding g : E → Ē from E to the algebraic closure Ē of E such
that g|K = f. For any αj , the image σ(αj ) is still a zero of f (x), so
σ(αj ) ∈ F [α1 , . . . , αn ]. This implies g ∈ Aut(E/F ). Hence the homo-
morphism
φ : Aut(E/F ) → G(K/F ), g 7→ g|K

is surjective. Since Ker(φ) = G(E/K), we have

|Aut(E/F )| = |G(E/K)| · |G(K/F )| = [E : K][K : F ] = [E : F ].

Hence E/F is a Galois extension. 2


7. Since G(E/K) is a subgroup of G(E/F ) of order 2, it suffices
to show that G(E/K) is not a normal subgroup of G(E/F ). This is a
consequence of the following proposition.
If a finite group of order 2p, with p being prime, has a normal
subgroup of order 2, then G is abelian.
In fact, the both the Sylow p-subgroup and the Sylow 2-subgroup
are normal subgroups. Hence G is their direct product. Hence G is
abelian.
8. Since E/F [α] is a Galois extension, the equality [E : F [α]] =
|G(E/F [α])| holds. The given condition means that G(E/F [α]) is triv-
ial. Hence [E : F [α]] = 1. 2
9. Let K be the fixed field of Aut(E/F ). Lemma 8.1.11 implies
that E/K is a Galois Galois extension. Hence |Aut(E/F )| = [E : K]
divides [E : F ]. 2

Exercises 8.2
1. If F is a finite field, then every finite extension of F is a simple
230 APPENDIX E. HINTS OR SOLUTIONS FOR EXERCISES

extension, which implies that E H = F [β] for some β ∈ E H . It follows


from the fundamental theorem of Galois theory that H = G(E/F [β]) =
{g ∈ G(E/F )|g(β) = β}.
Assume that F is an infinite field. For any α ∈ E H , let Gα = {g ∈
G(E/F )|g(α) = α}. Then Gα is a subgroup of G(E/F ) containing H.
Choose β ∈ E H such that |Gβ | reaches the minimum. It suffices to
show that Gβ = H.
Assume that Gβ 6= H. Choose g ∈ Gβ \H. Then there exists α ∈ E H
such that g(α) 6= α.
Let h1 , . . . , hr be all elements of G(E/F )\Gβ . Since F is an infinite
field, there exists a ∈ F such that

hi (aβ − α) 6= aβ − α

for all 1 ≤ i ≤ r. This means that Gaβ−α ⊆ Gβ . It follows from

g(aβ − α) 6= aβ − α

that Gaβ−α is a proper subgroup of Gβ , leading to contradiction. 2

2. Let ζ = e2πi/5 . Then

x4 + x3 + x2 + x + 1 = (x − ζ)(x − ζ 2 )(x − ζ 3 )(x − ζ 4 ).

Hence Q[x]/(x4 + x3 + x2 + x + 1) ∼ = Q[ζ] is a quartic Galois extension


of Q whose Galois group is cyclic.
The cubic Galois extension is a little more complicated. The poly-
nomial f (x) = x3 −12x+8 is a primitive polynomial (all coefficients are
integers with 1 as their greatest common divisor). If f (x) is reducible
in Q[x] then it can be decomposed into the product of two primitive
polynomials of lower degrees and so it has a zero in Z. Since every zero
of f (x) in Z divides 8, it can only be one of ±1, ±2, ±4, ±8. It is easy
to check that none of these integers is a zero of f (x). Hence f (x) is
irreducible over Q. So E = Q[x]/(x3 − 12x + 8) is a cubic extension
of Q. In order to show that it is a Galois extension, we need to check
that f (x) has three distinct zeros in E. In fact, it is easy to check that
2 2
α, α 2−8 and α 2+8 are zeros of f (x).
231

3. Define the homomorphism

φj : F [y] → F [y], f (y) 7→ f (y + j).

for j = 0, 1, . . . , p − 1. They keep y p − y + x unchanged. So φj induces


an automorphism of E. It is obvious that φ0 , . . . , φp−1 are p distinct
elements in Aut(E/F ).
It follows from [E : F ] ≤ p that |Aut(E/F )| = p = [E : F ].
Therefore E/F is a Galois extension of degree p. 2
Bibliography

[1] Apostol,T., Introduction to analytic number theory, Springer


Verlag, New York, Berlin, Heidelberg, 1976, ISBN 0-387-90163-9.

[2] Artin,M., Algebra, Prentice Hall, Englewood Cliffs, New Jersey,


1991, ISBN 0-13-004763-5.

[3] Bourbaki,N., Algebra, Chapters 1-3 (English translation),


Addison-Wesley, Reading, Massachusetts, 1974, ISBN 0-201-
00639-1.

[4] Bourbaki,N., Algebra, Chapters 4-7 (English translation),


Springer Verlag, New York, Berlin, Heidelberg, 1988, ISBN 0-387-
19375-8.

[5] Feng,K., Li,S., Za, J. and Zhang, P., Introduction to modern alge-
bra (in Chinese), Press of the University of Science and Technology
of China, Hofei, 2002ISBN 7-312-00041-X/O.47

[6] Lang,S., Algebra, Addison-Wesley, Reading, Massachusetts, 1971,


ISBN 0-201-04177-4.

[7] Li, K., Basic alstract algebra (in Chinese), Qinghua University
Press, Beijing, 2007ISBN 978-7-302-14407-6

[8] Moh. Z., Lan, Y. and Zhao, C., Algebra I and II (in Chinese),
Beijing University Press, Beijing, 1986ISBN 7-301-01371-x/O.222

[9] Serre, J-P., A course in arithmetic, Springer Verlag, Berlin, New


York, Heidelberg, 1973, ISBN 0-387-90041-1.

[10] Shafarevich,I., Algebra I, Springer Verlag, Berlin New York Hei-


derberg.

232
BIBLIOGRAPHY 233

[11] de Sousa,P., Silver,J., Berkeley problems in mathematics, Springer


Verlag, 2004, ISBN 10: 0387008926.

[12] Vinberg,E., A course in algebra (translated from Russian) , Amer-


ican Mathematical Society, 2003, ISBN 0-8218-3318-9.

[13] Wan,Z., Algebra and Coding ,3rd ed.(in Chinese), Higher Educa-
tion Press, Beijing, 2007ISBN 978-7-04-021717-9

[14] Weil, A. Basic number theory, 3rd edition, Springer Verlag, New
York, Berlin, Heidelberg, 1988, ISBN 0-387-19375-8.

[15] Yao, M., Abstract algebra (in Chinese), Fudan University Press ,
Shanghai, 1998, ISBN 7-309-02096-0/O.183
Index

abelian group, 4, 7 degree of a polynomial, 71


algebraic closure, 131 dihedral group, 16
algebraic element, 123 direct product, 40
algebraic extension, 123, 125 direct sum, 40
algebraic number, 132 division algorithm, 72
algebraic structure, 1 division ring, 53
algebraically closed field, 131
embedding, 151
alternating group, 29
endomorphism, 61
associated elements, 78
epimorphism, 33
automorphism, 33
equivalence class, 22
binary operation, 1 Euler function, 44, 174
Euler’s theorem, 44
Cardano’ formula , 181
even permutation, 14
Cayley’s theorem, 101
center, 8, 10, 26 factor group, 116
centralizer, 8 Fermat’s little theorem, 45
characteristic, 66 field, 53
Chinese Remainder Theorem, 60 field extension, 120
commutative ring, 51 field of fractions, 65
commutator subgroup, 26 field of rational functions, 77
composition series, 116 finite fields, 142
congruence, 57 finite group, 4
conjugate subgroup, 25 finitely generated extension, 122
coprime ideals, 59 finitely generated group, 9
coset, 18, 24 finitely generated ideal, 59
cubic equation, 181 First Isomorphism Theorem, 37
cycle, 14 fixed subfield, 150
cyclic group, 9, 11, 21, 35 fundamental theorem of homomor-
cyclotomic polynomial, 175 phisms, 35

234
INDEX 235

Galois extension, 152 left inverse, 53


Galois group, 152 left translation, 46
Galois theory, 148 left zero-divisor, 53
Gaussian integer, 75
general linear group, 5 maximal ideal, 66
greatest common divisor, 74 minimal polynomial, 124
group, 4 monic polynomial, 71
group action, 45 monoid, 3
group of automorphisms, 39 monomial, 76
group of symmetries, 16 monomorphism, 33
multiple factor, 82
Hilbert theorem 90, 181 multiple zero, 88
homomorphism, 32, 60 multiplicity, 82

ideal, 56 non-commutative ring, 51


identity element, 2 normal extension, 161
image, 33 normal subgroup, 23, 26
indeterminate, 72 normalizer, 27
index, 20
infinite group, 4 odd permutation, 14
inner automorphism, 34 orbit, 47
integral domain, 53 orbit formula, 99
intermediate field, 120, 150 order of a group, 4
invariant subspace, 96
partition, 20
inverse, 3, 53
permutation group, 12
invertible element, 3
polynomial, 71
irreducible decomposition, 80
polynomial function, 87
irreducible element, 79
prime field, 67
isomorphism, 33, 61
prime ideal, 70
kernel, 33, 61 primitive element, 145
primitive polynomial, 83
Lagrange’s theorem, 21 principal ideal, 59
law of associativity, 2 principal ideal domain, 74
law of cancelation, 5 projective general linear group, 25
left coset, 18
left ideal, 57 quadratic nonresidue, 169
236 INDEX

quadratic residue, 169 Sylow subgroup, 101


quartic equation, 183 Sylow’s theorem, 102
quaternion, 54 symmetric group, 12
quotient, 72 symmetric polynomial, 77
quotient group, 23
tetrahedron, 17
quotient ring, 56
torsion, 115
quotient space, 95
transcendental element, 123
rank, 115 transcendental number, 132
reciprocity law, 171 transitive action, 48
regular polyhedron, 18 transposition, 14
remainder, 72 trivial ideal, 58
resolvent cubic, 185 trivial subgroups, 7
right coset, 18
unit, 53
right ideal, 57
unitary group, 49
right inverse, 53
unitary operation, 2
right zero-divisor, 53
unity, 51
ring, 51
unque factorization domain, 79
Second Isomorphism Theorem, 37
vector space, 90
semigroup, 3, 7
separable extension, 161 Wilson’s theorem, 146
simple extension, 122
zero-divisor, 53
simple group, 27, 30
Zorn’s lemma, 133
simple ring, 58
skew field, 53
solvable extension, 162
solvable group, 118
special linear group, 7
splitting field, 161
stabilizer, 47
subfield, 56
subgroup, 7
subnormal series, 116
subring, 55
subspace, 91
INDEX 237

Вам также может понравиться