Download as pdf or txt
Download as pdf or txt
You are on page 1of 150

Abstract Algebra

Dr. Jason Juett


April 7, 2014
Table of Contents
1.1: Sets and Basic Notation
1.2: Functions
1.3: Equivalence Relations and Partial Orders
1.4: Well-Ordered Sets and Induction/Recursion
1.5: Ordinal Numbers
1.6: Cardinal Numbers
2.1: Semigroups, Monoids, and Groups
2.2: Subgroups and Cosets
2.3: Homomorphisms and Isomorphisms
2.4: Cyclic Groups
2.5: Cauchys Theorem and Other Assorted Facts
2.6: Permutations
3.1: Rings, Integral Domains, and Fields
3.2: More Ideals
3.3: The Construction of Z and
3.4: The Construction of 1
3.5: Polynomial Rings
1
2
3.6: The Ascending Chain Condition
3.7: Divisibility
Appendix to 3.7: The Euclidean Algorithm
3.8: Unique Factorization Domains
3.8: Cohens Theorem
4.1: Vector Spaces
4.2: Field Extensions
4.3: Splitting Fields
Introduction
These notes are more or less a self-contained introductory abstract algebra
course. However, while only an introduction to this subject, this course is
still fairly ambitious and expects a signicant amount of eort from the stu-
dent who wishes to master these topics. Very little pre-requisite knowledge is
assumed, beyond a general level of familiarity with pre-calculus mathematics
that is necessary in order to understand references made in examples. We will
start with a development of set theory, then move on to discuss groups, rings,
and elds. Along the way, we will carefully dene and construct the familiar
number systems of the integers, rational numbers, real numbers, and complex
numbers. Some things that will set this course apart from the typical one are:
(i) we are developing topics with a fair degree of rigor, (ii) we spend longer on
background topics than usual at the start but then make up for it by moving
more eciently through the main material, and (iii) my various biases about
which topics are interesting or important.
When references are made to Durbin, I am referring to the textbook Mod-
ern Algebra: An Introduction, 6th Edition, by Durbin [1]. At the beginning
of each section of the notes, I have indicated the most closely corresponding
section(s) of Durbin. It is not necessary to possess or read Durbins book, but
you may nd it a useful additional learning aid, because it provides a greater
quantity of concrete examples than I have, and is written in a dierent style.
Be advised, though, that Durbin and I have developed our topics in a dierent
order, and there are some small dierences in notation and denitions, so you
will have to accommodate for this if reading Durbin.
In order to be able to do the exercises, you will need to have some grasp of
how to write proofs. I suggest that you use the style in the proofs and theorems
in these notes (or in Durbin) as a model of roughly how your mathematical
writing should look. I have written many hints to help guide you through
the more challenging exercises, and we will be working with you during the
discussion sections and oce hours on helping you develop your mathematical
reasoning skills and your uency in writing mathematics. Here are a few more
words of advice. When one reads your mathematical writing, it should ow
naturally, as if reading a normal sentence, and it is usually easier to process
written words than a huge mess of symbols. For this reason, you should also
use complete sentences, proper grammar, well-organized thoughts, and so on.
3
Chapter 1
Set Theory
During the course of this introductory chapter, we will learn the basics about
sets, as well as some slightly advanced things about the foundations of set theory
and mathematics, building up to a development of the ordinal and cardinal
numbers, and tools such as mathematical induction and Zorns Lemma. The
study of abstract algebra proper will begin in the next chapter. While this
chapter will certainly provide a foundation for and have direct applications to
the algebraic structures we study later, we admittedly could go into somewhat
less depth and still have enough background to achieve a basic understanding
of algebraic structures. Therefore, this chapter is not only about giving you
information you will specically need for the algebra to come, but also about
exposing you to some (hopefully) interesting and challenging ideas and getting
you used to thinking mathematically.
1.1 Sets and Basic Notation
[Durbin: Appendices A and B]
Notation.
We will occasionally use the following logic symbols.
: implies or only if
: is implied by or if
: if and only if or is equivalent to
Denition. Roughly speaking, a set is a collection of objects. Two sets are
equal if they have the same members.
Notation.
1. We reserve special symbols for the following sets.
4
5
: empty set (the set with no elements)
N: set of natural numbers (including 0)
Z: set of integers
: set of rational numbes
1: set of real numbers
C: set of complex numbers
We will sometimes add a superscript + (resp., ) to denote the modied
version of that set containing only the positive (resp., nonzero) elements.
(The use of resp., which stands for respectively, above illustrates an-
other convention. That sentence consists of two very similar sentences
combined together to save space. To interpret such a sentence, rst read
it without any of the resp. things, then read it with the substitutions
indicated by the places that say resp..)
2. We use to indicate that an object is a member of a set, and / to
indicate that it is not. For example, 2 Z and

2 / .
3. There are three main ways to describe a set.
(a) We can represent a set by listing its elements inside , separated
by commas, e.g., S = 1, 2, 5 or T = 2, 4, 6, 8, . . ..
(b) Alternatively, we can represent sets by stating a rule describing which
elements are in the set, e.g., T = x Z
+
[ x is even, where the
symbol [ means such that in this context. (Alternatively, you
may use a colon, as Durbin does.) This notation is called set builder
notation.
(c) Finally, we may simply describe a set in words, such as the set of all
odd integers.
Denition.
1. If every member of a set X is a member of a set Y , we say that X is a
subset of Y , and write X Y . Alternatively, we say Y is a superset of X
and write Y X. (Note that X = Y if and only if X Y and Y X.)
If X Y and X ,= Y , then X is a proper subset of Y , and we may write
X _ Y or Y _ X. (Some people use the symbols and , but we will
avoid this, because it is not universally agreed upon if these mean and
or _ and _.) The power set of X is the set T(X) of subsets of X.
2. An indexed family is a set x

that contains an object x

for each
in the index set . Note that there is no requirement here that distinct
indices give distinct elements. Also, every set can be written as an indexed
family, because X = x
xX
for any set X.
6
3. The union of sets X and Y is X Y = x [ x X or x Y . More
generally, if X

is an indexed family of sets, then

= x [
x X

for some . The empty union is the union

with an empty
index set; it equals .
4. The intersection of sets X and Y is X Y = x [ x X and x Y .
More generally, if X

is a nonempty indexed family of sets, then

= x [ x X

for every . The empty intersection is the


intersection

with an empty index set, and what this equals is a matter


of convention. (Like how the expression 0
0
is undened in general, but in
some contexts may be dened to be 0 or 1.) We say X and Y are disjoint
if X Y = . A collection of sets is called pairwise disjoint if each pair of
distinct members of the collection are disjoint.
5. The complement of a set Y in a set X is X Y = x X [ x / Y .
(Note that this is a backslash.) Thus X and Y are disjoint if and only if
X Y = X.
Notation. We will often use notation akin to sigma-notation as a shorthand
when dealing with certain index sets. For example,

n
i=1
X
i
=

i{1,...,n}
X
i
and

j=0
Y
j
=

jN
Y
j
.
The approach to set theory described so far, where a set is simply any
collection of objects, is called nave set theory, and in practical applications it is
usually sucient. However, if one is creative, then inconsistencies within nave
set theory can be found, such as Russells Paradox: if ! is the collection of all
sets that are not a member of themselves (this collection is called the Russell
class) and ! is a set, then ! is a member of itself if and only if it is not. In order
to avoid such problems, we need to place some restrictions on which collections
of objects are allowed to be sets.
This leads us to our rened set theory, which works as follows. Every object
in our mathematical universe is a class, which is a collection of objects called
sets. Sets are then also classes, hence also collections of sets. But not every
class is a set; the ones that are not are called proper classes. Note then that
a proper class cannot be a member of a class. Russells Paradox is now no
longer a paradox, but merely a proof that ! is a proper class. (And there
is no analogous paradox with a class of all classes that are not a member of
themselves, because that is not a valid denition of a class.) We want all the
constructions above to give sets, provided that the objects we started with were
sets, so we make axioms that this is so. More explicitly:
1. If X is a set, then any subclass of X is a set. (Hence complements in sets
and intersections of arbitrary nonempty families of sets are sets.)
2. If is a set and X

is a set for each , then X

and

are sets.
3. If X is a set, then so is T(X).
7
These laws give us ways to create new sets or verify that certain classes are
sets, but in order to do anything at all, we have to make some axiom assuming
that a set exists. We will do this with the Axiom of Innity, which, roughly
speaking, asserts that N is a set. (We will discuss this in more detail later.)
In view of (1), this implies that is a set. Now you may be wondering, if N
is a set, and members of sets are also sets in our foundational system, then
how is a number such as 5 a set? For a look at whats to come, the way
that these numbers are built up from literally nothing is to dene 0 = , 1 =
, 2 = , , 3 = , , , , and so on, with each natural number
being the set containing the previous ones. (We are able to dene any given
natural number in this fashion starting only with , but the Axiom of Innity
is necessary in order for N itself to be a set.) In fact, we will see that every
object in standard mathematics is denable as a set (or a proper class at worst),
so nothing is really lost by our apparently restricted point of view that every
object is a class.
Remark.
1. Most normal math that one would do does not necessitate dealing with
proper classes, so throughout these notes I will generally frame our def-
initions in terms of sets rather than classes, and in those cases when we
do wish to apply those terms to classes, I will trust that what is meant is
clear.
2. Proper classes can intuitively be thought of as classes that are too big
to be a set. We will eventually see that a class is proper if and only if it is
the same size as the class V of all sets. (We will make this more precise
later.)
3. This course (like almost all math courses) assumes the Global Axiom of
Choice: Given any class of nonempty sets, there is a way to simultaneously
choose one element from each set. (This is the global version because it
allows the collection of sets to be a class. We will state the Global Axiom
of Choice more precisely in the next section.) This statement is provable
when there are only nitely many sets involved; it is the case where there
are innitely many when this cannot be proven and we need to take it as
an axiom. In some specialized parts of mathematical logic, the (Global)
Axiom of Choice is not assumed, but it is such an intuitive assumption
that it is quite easy to use it without even realizing it, so we will use it
freely without necessarily mentioning it each time.
4. We will also use the Axiom of Regularity: Every nonempty set has a
member that is disjoint from it, i.e., there is no set X whose members are
all nonempty subsets of X. This is actually a rather natural assumption,
since we would not even be able to properly describe such a set anyway,
since in attempting to describe what one of its members was, you would
be referred to one of the other members that was in that member, then in
describing that second member, you would be referred to a third, and so
8
on forever, without ever achieving a real description of what the members
of your set actually are. In an exercise, you will see that a consequence
of the Axiom of Regularity is that no set can be a member of itself, and
thus V = !. (This is one explicit way to show that V is a proper class,
but even without the Axiom of Regularity it would still be true, since any
class containing a proper class must be proper.) Another exercise will
show that, if you are given a set, then you pick a member of that set, then
you pick a member of that member, and continue doing this, then in a
nite number of steps you will always reach .
Exercises.
1. Let A and X be sets. Show that X(XA) A, and that equality holds
if and only if A X. (Hint: It suces to prove that X(XA) = XA.)
2. (Distributive Law) Show that, if X

and Y

are indexed fami-


lies of sets, then
_

_
=

,
(X

).
3. (De Morgans Laws) Show that, if X is a set and A

is a nonempty
family of sets, then X

(X A

) and X

(X A

).
4. (a) Show that there is no set that is a member of itself. (Hint: Suppose
X X. Note why X is a set, and then apply the Axiom of
Regularity to X to get a contradiction.)
(b) Show that there is no innite sequence X
n

n=0
of sets with each
X
n+1
X
n
. (Hint: Show that this sequence is a set that violates the
Axiom of Regularity.)
(c) Give an example of a set X
0
such that, for each N Z
+
, there is a
sequence X
n

N
n=0
with each X
n+1
X
n
. (This shows that, while
no such sequence can go on forever, it may be the case that there is
no upper bound on the length of such nite sequences.)
5. A class is transitive if each of its members is a subset of it. We will denote
the class of transitive sets by TR.
(a) Prove that a class T is transitive

tT
t T a T whenever
a b and b T. (This explains the name transitive.)
(b) Let X be a transitive set. Prove that T(X) is transitive. (Hint:
The denition of transitive can be rephrased as T is transitive
T T(T).)
(c) Prove that every nonempty transitive class has as a member. (Hint:
Use the Axiom of Regularity.)
(d) Show that TR is not transitive, i.e., that members of transitive sets
are not necessarily transitive. (Hint: The smallest possible coun-
terexample has three members.)
(e) Let T

be a family of transitive sets. Prove that

and

are transitive. (For the latter statement, assume ,= .)


9
1.2 Functions
[Durbin: Sections 1-2]
Denition.
1. A function (or mapping or map) from a set X (called the domain) into a
set Y (called a codomain) is a correspondence that assigns to each x X a
unique element of Y , denoted f(x) and called the image of x. (Note that,
although maps are often dened by a formula, this need not be the case.
Also, some sources use the word range for codomain, but some use it
to mean something else, so we avoid the word entirely.) We indicate that f
is a function from X into Y by writing f : X Y or X
f
Y . We denote
the set of maps X Y by M(X, Y ), and abbreviate M(X, X) = M(X).
2. Let f : X Y . For A X, the image of A is f[A] = f(a) [ a A;
we say f maps A onto f[A]. It is also common to write f(A) instead of
f[A] (as Durbin does), but we will occasionally encounter situations where
both A X and A X, so we have adopted the latter notation in order
to avoid any possible confusion.
3. For B Y , the pre-image of B is f
1
[B] = x X [ f(x) B. (This
notation is dened even if f does not have an inverse function.)
4. If f : X Y and g : Z W, then we say f = g if X = Z and
f(x) = g(x) for each x X. Note that with our denition there are
multiple dierent possible codomains for a function f : X Y ; any set
containing f[X] will do. (This is dierent from Durbins convention, but
has certain advantages, since we are really only concerned with the domain
and the correspondence, not the rather arbitrary choice of codomain.)
Example. f : 1 1 : x x
2
is a function. (We use this notation as a
shorthand to indicate that f is given by the formula f(x) = x
2
.)
1. f[1] = f[[0, )] = f[(, 0]] = [0, ).
2. f[(0, 2)] = (0, 4).
3. f
1
[(0, 4)] = (2, 0) (0, 2).
Example. If Y is a set, then there is a unique function Y , called the empty
function, which is the function that makes no assignments. However, there are
no functions from a nonempty set into .
Remark.
1. Functions from classes into classes can be dened in an analogous way.
(Though M(X, Y ) does not exist if X is a proper class, because in the
technical denition of a function as a class, a function is a set if and only
if its domain is. See exercises.)
10
2. The indexed family of sets X

can be identied with the function


X

, so the notion of an indexed family is thus just a notational


convenience. Along these lines, the axiom that a family of sets indexed by
a set is a set can be rephrased as: the image of a set under a function is a
set. However, pre-images of sets are not necessarily sets (if the functions
domain is a proper class). For example, if f : V N is any function, then
f
1
[N] = V.
Denition.
1. If X Y , the inclusion map : X Y is given by (x) = x. If we wish to
think of this map as a map X X, we refer to it as the identity map, and
denote it id. If there is any danger of confusion about to which inclusion
or identity map we are referring, we may add subscripts, e.g.,
X
or id
X
.
2. If f : X Y and g : Y Z, then the composition of g with f is the
function g f : X Z : x g(f(x)). Note that function composition
is associative, i.e., if f : X Y , g : Y Z, and h : Z W, then
(h g) f = h (g f). (Exercise.)
3. If A X and f : X Y , then the restriction of f to A is the function
f
A
= f
A
, i.e., f
A
: A Y : x f(x). In this case, we say the former
function is an extension of the latter to X.
Denition. The Cartesian product (or direct product) of an indexed family
X

of sets is the set

of all functions f :

with each
f() X

.
1. For n Z
+
, the Cartesian product X
1
X
n
=

n
i=1
X
i
can be
considered as the set of all ordered n-tuples (x
1
, . . . , x
n
) with each x
i
X
i
,
by associating each f

n
i=1
X
i
with the n-tuple (f(1), . . . , f(n)).
2. Similarly, the Cartesian product

i=1
X
i
can be thought of as the set of
sequences (x
1
, x
2
, . . .) with each x
i
X
i
.
3. The empty Cartesian product by denition consists only of the empty
function . We can represent this as a 0-tuple:

= ().
4. When writing functions with an input of an ordered n-tuple or sequence,
we use the abbreviations f((x
1
, . . . , x
n
)) = f(x
1
, . . . , x
n
) and f((x
1
, x
2
, . . .)) =
f(x
1
, x
2
, . . .).
Remark. We now have the terminology to more precisely state the two forms
of the Axiom of Choice.
1. Axiom of Choice:

= if and only if some X

= .
2. Global Axiom of Choice: There is a g : V V with each g(x) x.
(It is a good idea to ponder for a moment how these two statements repre-
sent choice.) Recall that we have agreed to freely use these axioms without
necessarily mentioning that we are doing so.
11
Denition.
1. A function f is injective (or one-to-one or an injection) if f(x) = f(y)
x = y.
2. A surjection X Y is a function f : X Y with f[X] = Y .
3. A bijection (or one-to-one correspondence) X Y is an injection with
f[X] = Y .
4. In the phrases surjection X Y and bijection X Y , we may omit
the X Y if it is clear from context what is meant, e.g., let f : X Y
be a surjection.
Example.
1. f : 1 1 : x x
2
is neither an injection nor a surjection.
2. f : 1 1 : x e
x
is injective but not a surjection.
3. f : 1 1 : x x
3
3x is a non-injective surjection.
4. f : 1
+
1 : x ln x is a bijection.
Denition. Let f : X Y be a function. A function g : Y X is a left
(resp., right) inverse function of f if g f = id
X
(resp., f g = id
Y
). We say
g is an inverse function of f if it is both a left and a right inverse function of
f. If a function f has an inverse, then the inverse is unique (we will prove this
shortly), and we denote the inverse function by f
1
. (Note that in this case we
have (f
1
)
1
= f.)
Example.
1. The identity function on any set is its own inverse.
2. The functions exp : 1 1
+
and ln : 1
+
1 are inverses.
3. The empty function is its own inverse.
4. Consider the functions sin : 1 [1, 1] and arcsin : [1, 1] 1. We
have sin arcsin = id
[1,1]
, so sin is a left inverse of arcsin, and arcsin
is a right inverse of sin, but the two functions are not inverses, because
arcsin(sin ) = 0. However, one can modify the domains/codomains so
that the functions are inverses: sin : [/2, /2] [1, 1] and arcsin :
[1, 1] [/2, /2].
Theorem 1. Let f : X Y .
1. f is injective it has a left inverse Y X or X = .
2. f is a surjection it has a right inverse Y X. Hence a right inverse
function of f must have domain f[X].
12
3. If f has a left inverse g : Y X and a right inverse h : Y X, then
g = h and Y = f[X]. In particular, inverse functions are unique when
they exist.
4. f is a bijection it has a left and a right inverse Y X it has an
inverse Y X.
5. f has an inverse it has an inverse f[X] X it is injective.
Proof.
1. (): Assume f is injective and X ,= . Dene g : Y X so that, for
each y f[X], g(y) is the unique element of X with f(g(y)) = y. Then
for each x X we have g(f(x)) = x, so g f = id
X
. (): If f has a left
inverse g, then f(x) = f(y) x = g(f(x)) = g(f(y)) = y.
2. (): Assume f[X] = Y . For each y Y , dene g(y) to be an element
such that f(g(y)) = y. Then f g = id
Y
. (): If f has a right inverse
g : Y X, then for each y Y we have f(g(y)) = y, and hence f[X] = Y .
3. In this case, we have g = g id
Y
= g (f h) = (g f) h = id
X
h = h.
4. The second equivalence follows from (3). The case of the rst equiv-
alence is immediate from (1) and (2), and, if the domain is nonempty, so
is the case. For the remaining case, assume f : Y is a bijection.
Then Y = f[] = , and f is its own inverse.
5. The rst equivalence follows from (2), and the second follows from (4).
Remark. It follows that there is a bijection X Y if and only if there is
a bijection Y X. In this case, we say that X and Y are in one-to-one
correspondence.
Theorem 2. Let f : X Y and g : Y Z.
1. g f is injective f and g f[X] are.
2. g f is a surjection g f[X] is.
3. g f is a bijection f is injective and g f[X] is a bijection.
Proof.
1. (): Assume g f is injective. Then f(x
1
) = f(x
2
) g(f(x
1
)) =
g(f(x
2
)) x
1
= x
2
, so f is injective. Also, if g(y
1
) = g(y
2
) for some
y
1
, y
2
f[X], then there are x
1
, x
2
X with g(f(x
1
)) = g(y
1
) = g(y
2
) =
g(f(x
2
)), so x
1
= x
2
and f(x
1
) = f(x
2
). Therefore g f[X] is injective.
(): If X ,= and f and g f[X] are injective, then they have left
inverses f

: Y X and g

: Z Y , respectively, and (f

) (g f) =
(f

) ((g f[X]) f) = id
X
, so g f is injective. On the other hand, if
X = , then f = g f[X] = g f is the empty function, which is injective.
13
2. This follows from the observation that (g f)[X] = g[f[X]] = (g
f[X])[f[X]].
3. Follows from (1) and (2).
Exercises.
1. Prove that function composition is associative.
2. Let f : X Y , A
1
, A
2
X, B
1
, B
2
Y , A

be an indexed
family of subsets of X, and B

be an indexed family of subsets of Y .


Correctly replace the question marks with either , , or =. If you
use one of the former two symbols, give an example where the inclusion
is proper.
(a) f[f
1
[B
1
]] ? B
1
.
(b) f
1
[f[B
1
]] ? B
1
.
(c) f[

] ?

f[A

].
(d) f[

] ?

f[A

] (here ,= ).
(e) f[A
1
A
2
] ? f[A
1
] f[A
2
].
(f) f
1
[

] ?

f
1
[B

].
(g) f
1
[

] ?

f
1
[B

] (here ,= ).
(h) f
1
[B
1
B
2
] ? f
1
[B
1
] f
1
[B
2
].
3. Let X and Y be classes. Dene (x, y)
K
= x, x, y for x X and
y Y , and dene X
K
Y = (x, y)
K
[ x X, y Y . (The subscript
K is to distinguish these from our denitions of the analogous concepts,
where the K is in honor of these versions inventor Kuratowski.) Our
ocial denition of a function f : X Y is an object of the form
(x, f(x))
K
[ x X, where each f(x) Y . (Intuitively, we are dening
a function to be its graph.)
(a) Show that (a, b)
K
= (c, d)
K
a = c and b = d. (Hint: You will
probably need to break this down into at least a couple cases.)
(b) Show that the above denition is equivalent to the more informal
one given at the beginning of the section. (That is, show that two
functions are equal under one denition if and only if they are equal
under the other.)
(c) Prove that f : X Y is a set if and only if X is. (Hint: Show that
they can each be written as a family indexed by the other.)
(d) Let X be a set. Prove that M(X, Y ) is a set if and only if Y is a set
or X = . (Hint: If X ,= , then Y can be indexed by the constant
functions. If Y is a set, then show that M(X, Y ) X
K
Y
T(T(X Y )).)
14
(e) Let X

be a family of sets indexed by a set. Show that

is a set.
(Note: Even though this is our ocial denition of a function, we will
not explicitly use it again, due to its incredible unwieldiness.)
4. Show that the following are equivalent for a map f : X Y .
(a) f is a surjection.
(b) f[f
1
[B]] = B for each B Y .
(c) f
1
[B] _ f
1
[C] for each B _ C Y .
5. Show that the following are equivalent for a map f : X Y .
(a) f is injective.
(b) f
1
[f[A]] = A for each A X.
(c) f[

] =

f[A

] for each indexed family A

of sub-
sets of X.
(d) f[A] _ f[B] for each A _ B X.
6. Let X and Y be classes, with X ,= . Show that there is an injection
X Y if and only if there is a surjection Y X.
1.3 Equivalence Relations and Partial Orders
[Durbin: Sections 16 and 63]
Denition.
1. A relation on a set X is a subset of XX. We write a b if (a, b) ,
and otherwise we write a ~ b.
2. If is a relation on a set X and A X, then the restriction of to
A is
A
= (A A) . In other words, the relation
A
is dened by
a
A
b a b. (In the future, when we dene a relation on a set,
and then refer to it as a relation on some subset, what we are technically
referring to is the restriction of that relation to that subset.) In this case,
we say the former relation is an extension of the latter to X.
3. A relation on a set X is:
(a) reexive if x x for all x X,
(b) irreexive if x ~ x for all x X,
(c) symmetric if x y y x,
(d) antisymmetric if x y and y x x = y, and
(e) transitive if x y and y z x z,
15
4. An equivalence relation is a reexive, symmetric, and transitive relation.
5. If is an equivalence relation on a set X, then the equivalence class of an
element x X is [x] = a X [ a x = a X [ x a. (If necessary,
we will add a subscript to avoid ambiguity, e.g., [x]

.) We denote the set


of equivalence classes of by X/ .
Example.
1. If X is any set, then = is an equivalence relation on X. The equivalence
classes are the singleton subsets x.
2. The relation is in one-to-one correspondence with is an equivalence re-
lation on V. The equivalence classes consist of sets that are the same
size.
3. For each n Z, congruence modulo n is an equivalence relation on Z.
(Recall that a, b Z are congruent modulo n, written a
n
b or a b
(mod n), if n [ (b a). Recall also that for a, b Z, we say a divides b,
and write a [ b, if b is a multiple of a.) This is not hard to verify directly,
but we will prove it as a special case of a more general theorem later.
4. The only relation on is the empty relation , which is an equivalence
relation. There are no equivalence classes.
Proposition 3. Let be an equivalence relation on a class X. The following
are equivalent for x, y X.
1. [x] = [y].
2. x y.
3. x [y].
4. y [x].
5. [x] [y] ,= .
Proof. Exercise.
Denition. A partition of a set X is a collection P of pairwise disjoint nonempty
subsets of X such that X =

AP
A.
Remark.
1. If X is a nonempty set, then a collection P of subsets of X is a partition
if and only if each element of X is a member of exactly one set in P.
2. The only partition of is the empty partition .
Theorem 4. Let X be a set. Then the set of equivalence relations on X is in
one-to-one correspondence with the set of partitions of X, via [x]
xX
.
The inverse map is P
P
, where x
P
y there is an A P with x, y A.
16
Proof. We need to show three things: (1) the set described in the second sen-
tence is a partition, (2) the relation described in the last sentence is an equiva-
lence relation, and (3) the two maps described are inverses.
(1) Let be an equivalence relation on X. Since each x [x], we have X =

xX
[x], and [x]
xX
is a pairwise disjoint collection of sets by Proposition
3, as desired.
(2) Let P be a partition of X. The relation
P
is reexive since each x X
is a member of some A P, and the fact that
P
is symmetric is clear. It
only remains to show transitivity. Assume x
P
y and y
P
z. Then there
are A, B P with x, y A and y, z B. Because the elements of P are
pairwise disjoint and y A B, we have A = B. So x, z A and x
P
z.
(3) We need to show that =
{[x]}
xX
and P = [x]

xX
for each equiv-
alence relation on X and each partition P of X. The former equation
states that two elements are -related if and only if there is an equivalence
class of in which they are both members; this follows from Proposition
3. The latter equation can be phrased as the equivalence classes of
P
are
the members of P, which is clear.
Remark. We can state a version of the above theorem for classes as follows. If
is an equivalence relation on a class X, then every element of X is a member
of exactly one equivalence class.
Denition.
1. A partial order is a reexive, antisymmetric, and transitive relation.
2. We say x and y are comparable with respect to a partial order if x y
or y x; otherwise, they are incomparable. A total order or linear order
is a partial order for which every pair of elements are comparable.
3. A partially (resp., totally) ordered set is a set together with a partial (resp.,
total) order on it. (Formally, we dene a partially ordered set as an ordered
pair (X, ), where X is a set and is a partial order on it, but we will often
abbreviate this as simply X if there is no danger of confusion.) Sometimes
a partially ordered set is called a poset for short. A totally ordered subset
of a partially ordered set is called a chain.
Example.
1. and are total orders on 1.
2. Any set X is partially ordered by =. This is a total order if and only if X
has at most one element.
3. The relations and are partial orders on V. They are not total. The
relation is often referred to as the inclusion relation.
17
4. The relation [ is a partial order on N but not a total order. The chains are
those sequences (either nite or innite) where each term is a multiple of
the previous one.
5. If is a partial order, then the reverse partial order is the partial order
given by a b b a. Note that is the reverse partial order of .
Denition.
1. A strict partial order is an irreexive and transitive relation.
2. If < is a partial order, then its corresponding strict partial order is the
partial order > given by a > b b < a.
3. Given a partial order , its corresponding strict partial order is the relation
< given by a < b a b and a ,= b. Conversely, given a strict partial
order <, its associated partial order is the relation given by a b
a < b or a = b. (In the theorem below, we verify that these relations
are indeed the kinds of relations their names suggest, and that the maps
< and < are inverses, so the terminology corresponding makes
sense.)
4. We dene terms for strict partial orders by analogy with the corresponding
ones for partial orders.
Remark. In a totally ordered set, we have a _ b a > b, but this does not
hold if the partial order is not total.
Theorem 5. The partial orders and strict partial orders on a set X are in
one-to-one correspondence via <. The inverse map is <.
Proof. We need to verify that (1) the rst map takes partial orders to strict
partial orders, (2) the second map takes strict partial orders to partial orders,
and (3) the maps are inverses.
1. Let be a partial order on X. The irreexiveness of < is immediate from
its denition. For transitivity, assume a < b and b < c. Then a b c,
a ,= b, and b ,= c. If a = c, then we have a b and b a, so a = b by the
antisymmetric property, a contradiction. Therefore a < c.
2. Assume < is a strict partial order. The reexiveness of is immediate
from its denition. If a b and b a, then since we cannot have both
a < b and b < a (exercise), we have a = b. Therefore is symmetric.
For transitivity, assume a b and b c. If a = b or b = c, then a c.
Otherwise, we have a < b < c, and certainly a c.
3. We need to verify that for every partial order and strict partial order <
we have: (i) x y (x y and x ,= y) or x = y, and (ii) x < y (x <
y or x = y) and x ,= y. Both of these statements are clear.
18
Denition. Let (X, ) be a partially ordered set.
1. We say m X is maximal (resp., minimal) if x m (resp., x m)
x = m.
2. A maximum (resp., minimum) element of a partially ordered set X is an
element m such that m x (resp., m x) for all x X.
3. If x < y and there is no z X with x < z < y, then we say x is
a predecessor of y, and y is a successor of x. In a totally ordered set,
successors and predecessors and unique when they exist. (Exercise.) When
an element x has a unique successor (resp., predecessor), we denote it by
s(x) (resp., p(x)).
Example.
1. It is important to understand the distinction between maximal and
maximum.
(a) Maximum means that it is the largest element, but maximal simply
means that there are no larger elements. Thus, in a totally ordered
set, the notions of maximal and maximum are the same.
(b) A partially ordered set can have at most one maximum element, but
may have arbitrarily many maximal elements. For example, if X is
a set, then every element of X is maximal with respect to =.
The above comments hold with maximal and maximum replaced with
minimal and minimum, respectively.
2. Z, , and 1 with the usual orders have no maximal or minimal elements.
3. Let X be a set. Then (T(X), ) has maximum element X and minimum
element .
4. (N, [) has minimum element 1 and maximum element 0. If you remove
those two elements, then the minimal elements are the primes and there
are no maximal elements.
Denition.
1. A map f : X Y between partially ordered sets is an order embedding if
f(a) f(b) a b. Order embeddings are injective. (Exercise.)
2. An order isomorphism between partially ordered sets is an order embed-
ding of one onto the other. (Important: Note the use of onto here.) We
say a partially ordered set X is order isomorphic to a partially ordered
set Y if there is an order isomorphism f : X Y . In an exercise you will
show that order isomorphic to is an equivalence relation on the class of
partially ordered sets.
19
Remark. If two partially ordered sets are order-isomorphic, then, as far as
partial order properties go, we can think of these partially ordered classes as
being the same, except merely for their elements being renamed. For example,
the order-isomorphic sets 0, 1, 2 and 10, 6, 105 (with the usual orders) have
the same partial order properties, and we can think of the latter partially ordered
set being the same as the former, but with 0 renamed to 10, 1 renamed to 6,
and 2 renamed to 105. Thus, if two partially ordered sets are order-isomorphic,
then any sort of partial order property of one set is also true for the other
set, with the appropriate substitutions made if the property specically names
elements. On a similar note, if there is an order embedding f : X Y , we may
think of f[X] as being a copy of X contained in Y that possesses all the same
partial order properties as X.
Exercises.
1. Prove Proposition 3.
2. Let < be a strict partial order. Prove that a < b b a.
3. Prove that, in a totally ordered set, an element can have at most one
successor and at most one predecessor.
4. (a) Prove that order embeddings are injective.
(b) Prove that a map f : X Y between totally ordered classes is an
order embedding f(x) < f(y) whenever x < y.
5. Show that is order isomorphic to is an equivalence relation on the class
of partially ordered sets.
1.4 Well-Ordered Sets and Induction/Recursion
[Durbin: Appendix C]
Denition. A well-ordered set is a totally ordered set in which every nonempty
subset has a minimum element. A well-ordering on a set is a partial order with
respect to which it is well-ordered. Note that every non-maximum element of a
well-ordered set has a unique successor, and that every subset of a well-ordered
set is well-ordered.
Example. All sets in this example are given their usual orders.
1. N is well-ordered.
2.
+
is not well-ordered, since it is non-empty and has no least element.
3.
+
0 is also not well-ordered, because even though it has a minimum
element, it has a nonempty subset
+
that does not.
4. is well-ordered (by the empty relation). It has no non-empty subsets, so
it is vacuously true that every non-empty subset has a minimum element.
20
Theorem 6 (Principle of Induction). Let X be a nonempty well-ordered class,
and for each x X let P(x) be a statement about x. Then P(x) is true for all
x X if and only if the following two statements hold.
1. (Base Case) P(a) is true for the minimum element a of X.
2. (Inductive Step) If b > a and P(x) is true for all x < b, then P(b) is true.
Proof. (): Clear. (): By contrapositive. Assume that P(b) is false for some
b X. We can choose a minimum such b, and by minimality P(x) is true for
all x < b. Therefore (1) and (2) cannot both hold.
Remark.
1. The most common usage of induction is applying it to the well-ordered
set X = Z
+
to show that a statement is true for all positive integers.
2. Statements (1) and (2) could equivalently be combined into If b X and
P(x) is true for all x < b, then P(b) is true, because if this statement
holds, then P(a) must be true. However, in practice it is often simplest
to verify the cases b = a and b > a separately, which is why induction is
usually formulated as above.
Example. It is essential that you are comfortable reading and writing proofs
by induction, so I will explain in some detail how this is done. For an example,
we prove the summation formula

n
k=1
k =
n(n+1)
2
. I will rst write the proof
in a manner that explicitly makes reference to the Principle of Induction, so
you can see how the theorem is being used, and then I will rewrite it in an
abbreviated form that is more like how mathematicians write in practice. (You
can decide which method of writing suits you best.)
1. (Explicit use of Principle of Induction:) For each n N, let P(n) be the
statement:

n
k=1
k =
n(n+1)
2
. We will show by induction that P(n) is true
for all n N. For the base case, we have

0
k=1
k = 0 =
0(0+1)
2
, so P(0)
is true. For the inductive step, if n > 0 and P(m) is true for all m < n,
then

n
k=1
k =

n1
k=1
k + n =
(n1)n
2
+ n =
n
2
n+2n
2
=
n
2
+n
2
=
n(n+1)
2
,
so P(n) is true. By the Principle of Induction, the statement P(n) is true
for all n N.
2. (Abbreviated form:) We will show that

n
k=1
k =
n(n+1)
2
for all n N
by induction on n. For the base case, we have

0
k=1
k = 0 =
0(0+1)
2
.
So assume n > 0. By induction, we have

n
k=1
k =

n1
k=1
k + n =
(n1)n
2
+n =
n
2
n+2n
2
=
n
2
+n
2
=
n(n+1)
2
.
The abbreviated form may seem almost nonsensical if read literally and without
being accustomed to such things, so I will explain the conventions/understandings
behind writing proofs in this way. You verify the base case, and then assume
21
that n is larger than the base case. Then, for the rest of the proof, you are
allowed to assume the statement is true for everything smaller than n. (The
Principal of Induction justies this.) Each time you use this assumption, you
use a phrase like by induction so the reader understands what you are doing.
If it is not obvious from context what the variable in your statement is, you
should say it prior to commencing the inductive proof, e.g., by induction on
n. In the above example, it would also have been stylistically acceptable to
omit the by induction on n, and I could have also left out the words for the
base case. If the setup for your proof is exceptionally complicated (like if you
are doing an induction within an induction within an induction or something
similarly crazy), then it may be necessary to write out everything in a more
explicit form so that what exactly you are doing is 100% clear.
Theorem 7 (Recursive Denition). Let X be a well-ordered class, let X
b
=
x X [ x < b for each b X, and let G : V V. Further assume that each
X
b
is a set. Then there is a unique f : X V with f(x) = G(f X
x
) for each
x X.
Remark.
1. In other words, in this case, it is valid and unambiguous to dene functions
X V recursively, i.e., we can specify a value for the rst point and a
way to determine the value at a point given the values at the previous
points. (We do not necessarily need to specify the value for the rst point
if our rule is phrased in such a way that it makes sense if there are no
previous points.) The function G represents the rule for determining f(x)
based on fs values at previous points.
2. If X = N, then we may reword this theorem as: it is valid and unambigu-
ous to dene a sequence a
n

n=0
recursively.
Proof. Uniqueness immediately follows from induction. To prove existence, it
suces to show that for each b X there is such a function f
b
: X
b
b Y ,
because then by uniqueness these functions agree where their domains overlap
and they can thus be extended to the desired function. By induction, there is
such a function f
a
for each a < b, and by uniqueness these functions can be
extended to such a function f
b
: X
b
Y , and dening f
b
(b) = G(f
b
X
b
)
extends f
b
to the desired function.
Example. The factorial function on N can be dened recursively by 0! = 1 and
n! = n(n 1)! for n > 0. (Note that it is perfectly ne that our recursive rule
has n in it, because n = (n1) +1 can be derived from the numbers preceding
it.)
As an example of recursive denition, we prove the following useful result.
Theorem 8 (Regularity of Classes). Every nonempty class has a member that
is disjoint from it.
22
Proof. By contradiction. Suppose that there is a nonempty class A that has
a nonempty intersection with each of its members. We recursively dene a
sequence X
n

n=0
of members of A with each X
n+1
X
n
. Let X
0
A, and
for n > 0, the set X
n1
A has some member X
n
A. The sequence we have
constructed violates the Axiom of Regularity (past exercise).
Another way of phrasing the above theorem is the following: if there is
a set satisfying a certain property P, then there is a set that satises P but
its members do not. Replacing P in the above statement with its negation
and then taking the contrapositive yields the following theorem describing a
powerful proof technique.
Theorem 9 (Hereditary Induction). Let P be a statement about sets. If P is
true for a set whenever it is true for all of its members, then P is true for all
sets.
One of the most famous, and perhaps surprising, results of set theory is the
Well-Ordering Theorem: every class has a well-ordering. (If you do not consider
this somewhat strange, try to imagine a way to place a well-ordering on 1. But
do not try too hard, because it turns out that, even though one exists, there
is not one that is explicitly denable and provably correct.) We will put o
proving the Well-Ordering Theorem until the next section. For now, we will
examine a couple of its consequences.
Theorem 10 (Hausdor Maximal Principle). Every chain in a partially ordered
set is contained in a maximal chain.
Proof. Let ( be a chain in a partially ordered set (X, ). By the Well-Ordering
Theorem, there is a well-ordering _ on X. Recursively dene f : X V
by f(x) = ( x if ( x

ax
f(a) is -totally ordered, and f(x) =
otherwise.
We claim that T =

xX
f(x) is a -chain. To see this, pick x y in T.
By the denition of T, each of its elements is -comparable to every element of
(, so we may assume x, y / (. Then x and y are both in the -totally ordered
set ( y

ay
f(a), hence comparable, as desired.
Now we show that ( T. If T ,= , then some f(x) ,= , which means
( ( x = f(x) T. On the other hand, if T = , then for each x ( the
set ( x

ax
f(a) = ( is not totally ordered, and since this is impossible
we conclude that ( = = T.
The fact that T is a maximal -chain follows once we observe that, if x / T,
then ( x

ax
f(a) T x is not -totally ordered.
The proof of the following very useful result will be an exercise.
Theorem 11 (Zorns Lemma). Let X be a nonempty partially ordered set. If
every nonempty chain in X has an upper (resp., lower) bound, then X has a
maximal (resp., minimal) element.
Exercises.
23
1. Show that, given sets X and Y , there is an injection or a surjection X
Y . (Hint: Assume there is no surjection X Y . Well-order X, and
recursively dene an injection f : X Y .)
2. An initial segment of a partially ordered set X is a subset A such that,
for every x X and a A, we have x a A x A.
(a) Show that unions and intersections of families of initial segments are
initial segments.
(b) Show that every proper initial segment of a well-ordered class X is
of the form a X [ a < b for some b X.
(c) Let f : X Y be an order isomorphism and A be an initial segment
of X. Show that f[A] is an initial segment of Y .
(d) Prove that a well-ordered class cannot be order isomorphic to one of
its proper initial segments. (Hint: Let A be an initial segment of a
well-ordered class X and f : A X be an order isomorphism. Use
induction to show that f = .)
3. (a) Prove the upper version of Zorns Lemma. (Hint: Use the Haus-
dor Maximal Principle to get a maximal chain.)
(b) Let X be a nonempty partially ordered set. Prove that, if each
nonempty well-ordered subset of X has an upper bound, then every
element of X is bounded above by a maximal element. (Hint: Pick
a X and consider the subset X
a
= x X [ x a. Show that the
set ( of well-ordered subsets of X
a
is partially ordered by the relation
A _ B A is an initial segment of B, and apply the upper version
of Zorns Lemma to ((, _). Verifying the requirement about chains
in ( having upper bounds can be reduced to showing that the union
of such a chain is in (.)
(c) Let X be a nonempty partially ordered set. Show that, if each
nonempty chain in X has an upper (resp., lower) bound, then each
element of X is bounded above (resp., below) by a maximal (resp.,
minimal) element. (As a special case of this part, the lower ver-
sion of Zorns Lemma is now proved. Hint: The upper version
follows immediately from part (b). For the lower version, apply
the upper version to the reverse partial order.)
1.5 Ordinal Numbers
[Not in Durbin.]
In the nal two sections of this chapter we will begin the project of con-
structing the number systems. Roughly speaking, one extends N to Z by adding
additive inverses, then extends that to by forming fractions, then extends that
to 1 by lling in the holes between rational numbers, then extends that to C
by adding an element i =

1. (By far the biggest jump is from to 1.) Often


24
the set N is taken as the starting point, and then everything else is built up out
of that, but I would like to share with you a brilliant way that mathematicians
have shown that this set (and hence all of standard mathematics) can be re-
cursively constructed from a starting point of only the empty set. We will rst
dene the ordinal numbers, which generalize the notion of natural numbers cor-
responding to an order, i.e., rst, second, third, and so on. In the next section,
we will dene the cardinal numbers, which are numbers to measure the size of
a set, and we will dene arithmetic on the cardinal numbers, of which natural
number arithmetic will be a special case. This way, we will be able to carefully
and thoroughly prove the basic properties of the natural numbers (including the
commutative, associative, distributive properties, and so on).
Before we dene what an ordinal number is, we will discuss the general idea of
the ordinals. We want to dene ordinal numbers so that the class ON of ordinals
is strictly well-ordered by , and so that each ordinal number consists precisely
of the smaller ordinals, or in other words is an initial segment of ON. So the
smallest ordinal should be , the next smallest should be , then , , then
, , , , and so on. In general, the successor of should be ,
which you will verify in the exercises. We will label the smallest ordinal as 0, the
next smallest as 1, the one after that as 2, and so on. So we understand how to
form the set representing each of the natural numbers, but the above description
of how to construct the ordinals would not be a very precise denition of what
an ordinal is. The following denition, due to von Neumann, very elegantly
and unambiguously denes the ordinals without having to resort to any sort of
(possibly vague or circular) recursive process.
Denition. An ordinal number (or simply ordinal) is a transitive set of transi-
tive sets. We denote the class of ordinal numbers by ON.
The next theorem will show that this very abstract denition is indeed the
one we were looking for.
Theorem 12 (Properties of Ordinals).
1. The class ON is transitive, i.e., members of ordinals are ordinals.
2. The class ON (and hence any ordinal) is strictly well-ordered by .
3. The relations and _ are the same on ON.
4. A set is an ordinal number if and only if it is an initial segment of ON.
5. Every well-ordered set is order isomorphic to a unique ordinal number.
Proof.
1. It is immediate from the denition that any member of an ordinal is a
transitive set consisting of transitive sets, hence an ordinal.
2. The fact that is a strict partial order on ON follows from the fact that
ordinals are transitive. Furthermore, if is an ordinal in a subclass A
25
of ON, then by regularity the set A has an -minimal element ,
and by transitivity is in fact an -minimal element of A. So the proof
will be complete once we show that is a strict total order on ON. By
two successive hereditary inductions, we reduce to proving that distinct
ordinals and are -comparable if each member of (resp., ) is -
comparable to (resp., ). [Technical note: In this specic proof, we are
not using the full version of the hereditary induction theorem as stated
in the previous section. Using that theorem in this specic proof would
be circular, because in the course of proving it we made implicit use of
some properties of ordinals by recursively dening a sequence indexed by
N. Instead, our use of hereditary induction on ON is justied by the fact
that each of its nonempty subclasses has an -minimal element.] Without
loss of generality, let us say there is a . Then = or ,
and in either case we have .
3. If _ are ordinals, then by regularity we cannot have , and from
(2) we conclude .
4. By (1), an ordinal is the set of ordinals less than it, hence an initial segment
of ON. Conversely, if X is any set that is an initial segment of ON, then
it consists of transitive sets, and each A consists of smaller ordinals,
hence is a subset of A.
5. First we prove that every well-ordered set X is order isomorphic to an
ordinal number. For each b X let X
b
= x X [ x < b. Recursively
dene f : X V : x f[X
x
]. We claim that f is an order-isomorphism
onto an ordinal. Since any pair of elements of X is contained in a set
of the form X
b
b, and since a union of a set of ordinals is an ordinal
(exercise), it suces to show that the restriction of f to each of these sets
is an order-isomorphism onto an ordinal. By induction, the set f(b) =
f[X
b
] =

x<b
f[X
x
x] is a union of a set of ordinals, hence an ordinal,
and f[X
b
b] = f(b) f(b) is its successor. Also, for each a < b, we
have f(a) f[X
b
] = f(b), as desired. For uniqueness, we note that two
distinct ordinals _ cannot be order-isomorphic, because in this case
is a proper initial segment of .
Henceforward, we will typically write < in place of when referring to the
ordering of the ordinal numbers.
Denition. The smallest ordinal number is denoted 0, its successor is denoted
1, and so on. The smallest ordinal number greater than all of 0, 1, 2, . . . is
denoted or N, depending on whether we wish to think of it as an ordinal
number or as the set of natural numbers.
Remark.
26
1. The above denition of the natural numbers is consistent with our intu-
itive notion of them as counting numbers, and is the standard precise
denition used in rigorous mathematics.
2. One way to state the Axiom of Innity is: there is an ordinal such that
(i) 0 < and (ii) if < , then so is its successor. (You will show in an
exercise that every ordinal number has a successor.) A more precise way
to state the denition of informally given above is that is the least
such .
3. The ordinal numbers are not merely indicative of the size of a well-
ordered set, but also of how it is ordered. Finite sets (a concept we will
dene precisely in the next section) can only be well-ordered in one way
(up to order isomorphism), hence only correspond to one ordinal, but an
innite set can correspond to innitely many dierent ordinals, depending
on how it is well-ordered. (Future exercise.) The numbers that are used
to measure the size of a set are the cardinal numbers, and we will study
them in the next section.
Theorem 13 (Well-Ordering Theorem for Sets). Every set has a well-ordering.
Proof. It suces to show that every set is in one-to-one correspondence with an
ordinal, since then the ordinals well-ordering corresponds to a well-ordering on
that set, and, since = 0, we only need to consider nonempty sets. Let X be
a nonempty set, and let g : T(X) X be a function with g(A) A for each
A ,= . Recursively dene f : ON X by f() = g(X f[]). Note that the
denition implies that f() ,= f() for < with f[] ,= X. If there is no
ON with f[] = X, then f is an injection and X is a set containing the
proper class f[ON], a contradiction. (An exercise shows that ON is a proper
class.) Therefore there is some minimum ON with f[] = X, and X is in
one-to-one correspondence with .
Another application of the ordinal numbers is showing that every set can be
recursively built up from .
Theorem 14 (Rank of Sets). Recursively dene V

<
T(V

) for
ON. Then:
1. The V

s are transitive sets, and V

for .
2. V =

ON
V

. In particular, for each set X, there is an ordinal with


X V

, and we may dene the rank of X to be the least such ordinal.


3. The rank of a set is the least ordinal greater than the rank of all its ele-
ments.
4. The rank of an ordinal is itself.
Proof.
27
1. By induction, any V

is a union of a set of powersets of transitive sets,


which is transitive by past exercises. The second statement is clear.
2. By hereditary induction it suces to show that if every member of a set
X is in

ON
V

, then so is X. In this case, we have each x X in


some V
x
. Let =

xX

x
, which is a union of ordinals and hence
an ordinal (exercise). Then each
x
, so X

xX
V
x
V

and
X T(V

) = V

, where is the successor of .


3. Let X be a set of rank , and let be the least ordinal greater than the
rank of all the elements of X. For each x X, we have x V
x
for some

x
< , so X

xX
T(V
x
) V

. Therefore . On the other


hand, the inclusion X V

<
T(V

) shows that every element of


X has rank less than , so , and thus = .
4. Let ON. By hereditary induction, the rank of each of s elements
is itself, so by the previous part the rank of is the least ordinal greater
than all the ordinals less than , which of course is .
Theorem 15 (Hereditary Denition). Let T be a transitive class and G : V
V. Then there is a unique f : T V with f(x) = G(f x) for each x T.
Remark.
1. Note that we need each x T to be a subset of T in order for the equation
f(x) = G(f x) to make sense, so this is why we take T to be transitive.
2. This theorem allows us to recursively dene a function on a transitive
class (such as V or ON) by specifying how it determines the value of a
set, given the values of that sets members.
Proof. Uniqueness is immediate from hereditary induction. For existence, it
suces to prove the case T = V, and we can in turn reduce this to proving
that for each ON there is such a f

: V

V (note that V

is transitive),
because by uniqueness the f

s agree where their domains overlap (recall from


an exercise that an intersection of transitive sets is transitive), and they can
thus be extended to the desired function. By induction, there is such a function
f

: V

V for each < . Dene f

: V

V : x G(f

x), where
< and x V

. (Note that f

is well-dened since the f

s agree where
their domains overlap.) We observe that each f

= f

, and consequently
f

(x) = G(f

x) for each x V

, as desired.
Example.
1. We could have used the equivalent statement of rank in Theorem 14 part
(3) above as its denition.
28
2. We can also make recursive denitions like dening a hereditarily nite set
to be a nite set of hereditarily nite sets (we will give a precise denition
of nite in the next section). This is because a denition for a kind of
set could be viewed as a function f : V V, where f(X) = 1 if X is
that kind of set and f(X) = 0 if X is not. The hereditarily nite sets are
interesting from a logicians point of view, because they can be used to
show that the Axiom of Innity cannot be proved from the other axioms
(hence must be taken as an axiom), but we will not talk about them any
further in this class.
Theorem 16 (Well-Ordering Theorem for Proper Classes). Every proper class
has a well-ordering that makes it order isomorphic to ON.
Proof. We show that V has a well-ordering in which every proper initial segment
is a set. This then induces such a well-ordering for all other classes, and by an
exercise this proves the desired conclusion. For each ordinal , let

be a
well-ordering on the set of sets of rank . (We know that this is a set because
by denition it is a subset of T(V

).) Dene a relation on V by X Y


rank(X) < rank(Y) or rank(X) = rank(Y) and X
rank(X)
Y . It is simple to
verify that is a total order on V. If A is any nonempty subclass of V, then it
has an element of minimum rank , and the set of elements of A of rank has
a

-minimum element, which is easy to verify is a -minimum element of A.


Therefore is a well-ordering on V. Finally, if C is any proper initial segment
of V, then C has some upper bound Y , and C T(V
rank(Y)
), showing that C
is a set.
An immediate corollary is the following, which gives the previously alluded
to result that a class is proper if and only if it is the same size as V.
Corollary 17 (Limitation of Size). A class is proper if and only if it is in
one-to-one correspondence with V.
Exercises.
1. (a) Prove that the successor of an ordinal number is .
(b) Explicitly write out what set the ordinal number 5 is, without using
any numerals.
(c) Explicitly write out what set V
3
is, this time using numerals to rep-
resent the ordinals in it.
2. Prove that every union or intersection of a family of ordinal numbers is
an ordinal number. (Hint: Refer to a past exercise.)
3. Show that ON is a proper class. (Hint: Show that if ON is a set, then
ON ON.)
4. (a) Show that every nonzero ordinal falls under exactly one of these two
classications.
29
i. Successor ordinal: a successor of an ordinal.
ii. Limit ordinal: a nonzero ordinal that is the union of the ordinals
less than it.
(b) Prove that is the smallest limit ordinal. (Hint: Show that any limit
ordinal satises conditions (i) and (ii) in the precise denition of .)
5. This exercise will show that ON can be characterized as the unique (up
to order isomorphism) well-ordered proper class whose proper initial seg-
ments are sets.
(a) Prove that every well-ordered proper class has an initial segment
order isomorphic to ON. (Hint: Let W be a well-ordered proper
class and recursively dene a function f : ON W so that f() is
the least element of W not in f[]. [Note that f[] ,= W since f[] is
a set.] Show that f[ON] is an initial segment. To show that f is an
order embedding, by a previous exercise [cite it] it suces to show
that if < , then f() < f().)
(b) Prove that a well-ordered proper class is order isomorphic to ON if
and only if each of its proper initial segments is a set. (Hint: For
, use the fact that the initial segments of such a class are images
of the initial segments of ON. Conversely, use induction to show
that the order embedding from (a) is a surjection if the proper initial
segments are sets. More specically, if x W, then by induction
a W [ a < x is the image of some subclass of ON. Show that
this subclass is an initial segment that is a set, hence an ordinal ,
and that x = f().)
(c) Prove that every proper class contained in ON is order-isomorphic to
ON.
(d) Give an example of a well-ordered proper class not order isomorphic
to ON.
1.6 Cardinal Numbers
[Not in Durbin.]
Denition. The cardinality of a set X is the smallest ordinal number [X[
that is in one-to-one correspondence with X. (This exists because X can be
well-ordered, hence is in one-to-one correspondence with an ordinal, and ON is
well-ordered, so there is a least such ordinal.) The class CN of cardinal numbers
(or simply cardinals) consists of the ordinal numbers that are cardinalities of
sets. Note that for each ON we have [[ , and equality holds if and only
if CN.
Remark.
30
1. It is immediate that [[ = 0, and we will soon see that [n[ = n for n .
Thus all ordinal numbers are cardinal numbers.
2. The way that mathematicians have dened cardinal and ordinal arith-
metic, the expression + 1 equals in cardinal arithmetic and the suc-
cessor of in ordinal arithmetic. For this reason, it is customary to write

0
instead of when thinking of it as a cardinal number.
3. With this denition, proper classes do not have a cardinality, because they
cannot be in one-to-one correspondence with any set. We could generalize
the denition of cardinality to classes by dening [X[ to mean the smallest
initial segment of ON that is in one-to-one correspondence with X (so all
proper classes would have cardinality ON), but this is nonstandard.
Theorem 18 (Cantor-Bernstein Theorem). The following are equivalent for
classes X and Y .
1. X and Y are in one-to-one correspondence.
2. There is an injection X Y and a surjection X Y .
3. There are injections (resp., surjections) X Y and Y X.
Proof. Note that the theorem is trivially true if X or Y is empty, so we may
assume X, Y ,= .
(1) (2): Clear. (2) (3) (4): Immediate from a past exercise.
(2) (1): Assume there are injections f : X Y and g : Y X.
Let A
0
= X g(Y ), and recursively dene A
n
= g[f[A
n1
]] for n 1. Let
A =

n=0
A
n
, and let h : X Y be the function with h(x) = f(x) for x A
and h(x) = g
1
(x) for x / A. (Note that X A X A
0
= g[Y ], so g
1
is
dened on X A.) We wish to show that h is a bijection.
For injectivity, assume h(a) = h(b) for some a, b X. If exactly one of
these elements is in A, say a A and b / A, then a A
n
for some n 0, and
b = g(g
1
(b)) = g(f(a)) g[f[A
n
]] = A
n+1
A, a contradiction. So a, b A
or a, b / A, and it follows from the injectivity of f and g
1
that a = b.
For surjectivity, pick any y Y . If y f[A], then there is an x A with
f(x) = y, and hence h(x) = y, so let us assume y / f[A]. If g(y) A
n
=
g[f[A
n1
]] for some n Z
+
, then y f[A
n1
] f[A], a contradiction. Because
g(y) / A
0
, we conclude that g(y) / A, and hence h(g(y)) = g
1
(g(y)) = y.
Corollary 19. Let X and Y be sets.
1. [X[ = [Y [ X and Y are in one-to-one correspondence.
2. [X[ [Y [ there is an injection X Y . In particular, if X Y , then
[X[ [Y [.
Proof.
1. Follows directly from the denition of cardinality.
31
2. Let f : X [X[ and g : Y [Y [ be bijections. (): If [X[ [Y [,
then [X[ [Y [ and g
1

|X|
f : X Y is an injection. (): By
contrapositive. Assume [X[ > [Y [. Then X and Y are not in one-to-one
correspondence, and by () there is an injection Y X, so by the
Cantor Bernstein Theorem there is no injection X Y .
We want to regard two sets as being the same size if they are in one-to-one
correspondence, so the above corollary shows that the cardinality of a set is the
number one should think of as a measurement of the size of a set.
Theorem 20 (Characterization of Innite Classes). The following are equiva-
lent for a class X.
1. There is an injection X X that is not a surjection.
2. There is a surjection X X that is not an injection.
3. There is a bijection between X and one of its proper subsets.
4. There is a surjection X N.
5. There is an injection N X.
Remark. We dene a class to be innite if it satises one (equivalently, all)
of the above conditions, and nite otherwise. Observe that two classes in one-
to-one correspondence are either both nite or both innite. Since V is clearly
innite (by (5)), it follows that every proper class is innite and every nite
class is a set.
Proof. (2) (1): Assume there is a surjection f : X X that is not injective.
Let g : X X be a right inverse of f. Then f g = id, so g is injective.
However, the function g cannot be a surjection, because then it would be a
bijection and so would f = g
1
. (1) (3): If there is a map f : X X
that is injective but not a surjection, then f is a bijection between X and
its proper subset f[X]. (3) (4): The case where X is a proper class (and
hence in one-to-one correspondence with V) is clear, so let us assume X is a
set and that there is a bijection f from X onto a proper subset. Recursively
dene a sequence X
n

n=0
by X
0
= X and X
n
= f[X
p(n)
] for n > 0. We will
prove by induction on n that each X
s(n)
_ X
n
. For the base case, we have
X
1
= f[X
0
] = f[X] _ X = X
0
. So assume n 1. By induction, we have
X
n
_ X
p(n)
, and thus X
s(n)
= f[X
n
] _ f[X
p(n)
] = X
n
, as desired. (Note that
we are using the fact that f is injective to conclude that the last containment
is proper.) It follows that X
n
X
s(n)

n=0
is a collection of pairwise disjoint
nonempty subsets of X. Dene g : X N to be a function that takes elements
in X
n
X
s(n)
to n (it makes no dierence where g sends elements not in any
X
n
X
s(n)
). Then g is a surjection. (4) (5): Immediate from an exercise.
(5) (1): Assume there is an injection f : N X. Dene g : X X by
32
g(x) = f(s(f
1
(x))) for x f[N] and g(x) = x for x / f[N]. Note that g
takes f[N] into f[N] and takes X f[N] onto itself. If there is an x X with
g(x) = f(0), then x f[N] and f(0) = f(s(f
1
(x))), so 0 = s(f
1
(x)), a
contradiction to the fact that 0 is the smallest ordinal. Therefore g is not a
surjection. To show that g is injective, it suces to show that its restrictions to
f[N] and X f[N] are injective, which follows from observing that the former
is a composition of injections (note that the successor function is injective) and
the latter is the identity map. (1) (2): Similar to (2) (1).
Example.
1. Examples of innite sets include Z
+
and 1.
2. Examples of nite sets include and 1, . . . , n, where n Z
+
.
Proposition 21.
1. An innite cardinal number is a limit ordinal.
2. [n[ = n for n .
3. A set X is innite if and only if [X[
0
. In other words, a set is nite
if and only if it is in one-to-one correspondence with some n < .
Proof.
1. If is any innite successor ordinal, then [[ = [p() p()[ = [p()[
p() < , since by an exercise removing a single point from an innite
set does not change its cardinality. Since 0 is nite (there is certainly no
injection N ), this means that any innite cardinal number must be a
limit ordinal.
2. By a past exercise, the smallest limit ordinal is , so any ordinal less than
is nite. Since an innite set cannot be in one-to-one correspondence
with a nite set, it follows that [[ = . On the other hand, if n < , then
n is nite and not in one-to-one correspondence with a proper subset, so
[n[ = n.
3. Follows from (2) and Theorem 20.
Denition. In practice, when one encounters an innite set, it is often not
really important which innite cardinality it has, but merely that it is innite,
so mathematicians have adopted the following lazy notation. For a cardinal
number , we write = to indicate that it is innite, and < to indicate
that it is nite. (Note that < has the same meaning as <
0
, and
= has the same meaning as
0
.) It is important to emphasize
that is not a cardinal number, but is just notation, and does not make sense
as an object on its own, only in expressions like the preceding two.
33
Denition. We dene addition, multiplication, and exponentiation operations
on the cardinal numbers as follows. (We will not discuss ordinal arithmetic in
this course.)
1. If X and Y are disjoint, then [X[ +[Y [ = [XY [. (One can always choose
disjoint representative sets X and Y , by replacing X with 0 X and Y
with 1 Y , if necessary.)
2. [X[[Y [ = [X Y [.
3. [X[
|Y |
= [M(Y, X)[.
(An exercise shows that these denitions are well-dened, i.e., they do not de-
pend on which sets of a particular cardinality we choose.)
Remark.
1. The denitions of cardinal addition, multiplication, and exponentiation
are consistent with the usual denitions of these operations on N. (Pre-
calculus counting techniques verify this.) Thus one may think of cardinal
arithmetic as an extension of standard arithmetic.
2. While the expression 0
0
is considered to be an indeterminate form in
calculus, in the context of cardinal arithmetic it is dened to be 1.
3. One can extend the denitions of cardinal addition and multiplication to
accommodate sums and products of arbitrarily many terms:

[X

[ =
[

[ and

[X

[ = [

[, where in the former case the X

s
must be chosen to be disjoint. Induction shows that these denitions are
consistent with the above ones.
Theorem 22 (Addition and Multiplication of Innite Cardinals). Let and
be cardinal numbers, at least one of which is innite. Then + = =
max(, ).
Proof. Without loss of generality, we may assume , and thus is innite.
It is simple to check that + and are bounded between and
2
, so it
will suce to show that =
2
.
Suppose to the contrary that there is an innite cardinal with <
2
;
because CN is well-ordered, we can pick to be the least innite cardinal
with this property. Dene a relation _ on by (x, y) _ (z, w) if (i)
max(x, y) < max(z, w), or (ii) max(x, y) = max(z, w) and x < z, or (iii)
max(x, y) = max(z, w), x = z, and y w. It is slightly tedious but not
hard to show that _ is a well-ordering. (You may ll in the details if you
wish.) Thus there is an order isomorphism f from onto some ordinal ,
and <
2
= [[ . Since , there are , with f(, ) = .
Let be the successor of max(, ), and note that [[ < since
is a limit ordinal, so by the minimality of we have either [[
2
= [[ or
[[
2
<
0
, and in either case [[
2
< . (The intuitively obvious fact that a
34
nite cardinal number squared is nite is noted in the theorem below.) But
= f[(, ) [ (, ) (, )] f[ ], so [[
2
< , a
contradiction.
Theorem 23 (Properties of Cardinal Arithmetic). The cardinal numbers satisfy
the following arithmetical properties.
1. Addition and multiplication are commutative and associative.
2. The distributive property holds.
3. + 0 = 1 =
1
= .
4. 0 = 0 and
0
= 1.
5. If and , then + +, , and

.
6. n = + +
. .
n copies
for 0 < n <
0
.
7.
n
=
. .
n copies
for 0 < n <
0
.
8. ()

.
9.
+
=

.
10. (

.
11. If m, n <
0
, then m+n is the nth successor of m.
12. If m, n <
0
, then m+n, mn, m
n
<
0
.
13. For < max(
0
, , ), + = + = .
14. For 0 < < max(
0
, , ), = = .
15. If < or <
0
, then there is a unique with + = .
Remark. You would not be expected to memorize all of these properties. Just
look them over once and be content in the knowledge that most of the familiar
facts about natural number arithmetic are now solidly proven.
Proof. [I have only written a sketch of this proof, since it is very tedious but
not too dicult once the key observations are made.] Properties (1)-(5) follow
from the denitions in a straightforward manner. Properties (6) and (7) can be
proved by observing that n = [

n
k=1
(k )[ and
n
= [

n
k=1
[.
8. It is straightforward to verify that the map : M(, ) M(, )
M(, ) given by (f, g)(x) = (f(x), g(x)) is a bijection.
9. Follows from (7) if and are nite, and the case where 1 is trivial.
Otherwise, we have
+
=
max(,)
= max(

) =

.
35
10. It is straightforward to verify that the map : M(, ) M(, M(, ))
given by (f)(x)(y) = f(y, x) is a bijection.
11. Because contains the successor of each of its elements, we have m+1 =
[m m[ = m m, which is the successor of m (past exercise). The
full result can be proven with induction.
12. The m + n case follows from (11) and induction. One can now use this
result and (6) to prove the mn case by induction, and in turn use that
result and (7) to prove the m
n
case by induction.
13. Assume . If <
0
, the result follows from (11) and induction, so
assume
0
< . Then max(, ) = + = + = , and hence
= .
14. Assume . The cases = 0 and = 1 are trivial, and the case

0
< is dealt with as in the proof of (13), so let us assume 1
and 2 <
0
. If
0
, then (11) implies that
0
and p() =
p()+ = (p() +1) = = = p(), and then = by induction
(on ). So let us assume <
0
. Then p() + = (p() + 1) =
= = p() + , and p() = p() by (13). Thus p() = p() by
induction (on ), and hence = .
15. Let = [ [ and note that + = and . Uniqueness in the
case <
0
follows from (13). On the other hand, if < and
0
,
then = max(, ), and hence = .
In the last section, we rigorously constructed N, and the above theorem
demonstrates most of its important arithmetical properties. We will delay the
ocial construction of the other number systems Z, , 1, and C for the moment,
because it will be useful to learn some algebra before commencing this project.
For now, we will be content with our informal/intuitive understanding of these
number systems so that we can use them to make examples.
Theorem 24 (Cantors Theorem). [X[ < [T(X)[ = 2
|X|
for any set X.
Proof. The second equality follows from observing that there is a natural one-to-
one correspondence between T(X) and M(X, 0, 1), where A X corresponds
to the function that takes elements in A to 1 and elements not in A to 0.
Now suppose that [T(X)[ [X[. Since T(X) ,= , this means there is a
surjection f : X T(X), and there is a y X with f(y) = x X [ x / f(x).
Hence y f(y) y / f(y), a contradiction.
Since CN is well-ordered, Cantors Theorem shows that each cardinal number
has a cardinal successor, i.e., a least cardinal number greater than it. It follows
that there is no largest cardinal number, and that CN is innite. In fact, an
exercise will show that the class of (innite) cardinal numbers is a proper class,
36
and, since it is contained in ON, it is order isomorphic to ON (past exercise).
Explicitly, the order isomorphism is as follows. For > 0, we dene

to be the
smallest cardinal number greater than

for all ordinals < . Then

is an order isomorphism from ON onto the class of innite cardinal numbers.


In an exercise, you will see that [N[ = [Z[ = [[ =
0
.
Theorem 25. [1[ = 2
0
.
Proof. We rst show that [[0, 1)[ = 2
0
. Consider the function f : T(Z
+
)
[0, 1] that takes a set to the number 0.a
1
a
2
a
3
. . .
2
whose binary expansion has
zeros in the digits in the subset and 1s elsewhere. For example, f(Z
+
) = 0,
f(2) = 0.101111 . . .
2
, f(3, 4, 5, . . .) = 0.11
2
, and 1, 3, 5, 7, . . . 0.010101 . . .
2
.
Each number in [0, 1) has a binary expansion, and the expansion is unique, if
we agree to rewrite expressions of the form 0.a
1
. . . a
n
01111 . . .
2
as 0.a
1
. . . a
n
1
2
.
So, if we restrict f to the set T

(Z
+
) of innite subsets of Z
+
, we get a bijection
T(Z
+
) [0, 1). By an exercise, we thus have [[0, 1)[ = [T

(Z
+
)[ = 2
0
.
Finally, the map (n, x) n + x is a one-to-one correspondence between
Z [0, 1) and 1, so [1[ =
0
2
0
= 2
0
.
Remark. The famous Continuum Hypothesis is that
1
= 2
0
, or, in other
words, that there are no cardinal numbers between [N[ and [1[. The Generalized
Continuum Hypothesis is that
n
= 2
n1
for all n Z
+
. Logicians have
proven that both hypotheses are impossible to either prove or disprove from the
standard axioms of mathematics.
Denition. A set is countable if its cardinality is at most
0
; otherwise, it is
uncountable. (The name comes from the fact that a set is countable if and only
if there is a way to well-order it so that you could count to any given element
in a nite number of steps.) A set that is of cardinality equal to
0
is called
countably innite.
Exercises.
1. Let X be an innite set. Show that adding or subtracting a single point
does not change its cardinality. (Do not use any results occurring after
Theorem 20 in your proof, because most of their proofs rely on this ex-
ercise. Hint: First show that it does not matter which point is removed,
then use the fact that X is in one-to-one correspondence with a proper
subset.)
2. (This exercise will show that the class of (innite) cardinals is a proper
class.) Prove that every class with members of arbitrary large cardinality
is a proper class. (Hint: Suppose that such a class is a set, then arrive
at a contradiction by constructing a set of greater cardinality than all the
sets in it.)
3. Show that the cardinal arithmetic denitions are well-dened.
4. (a) Prove that [[ = [Z[ = [Z
+
[ =
0
. (Hint: Use cardinal addition and
multiplication to reduce this to nding an injection
+
Z
+
Z
+
.)
37
(b) Prove that [C[ = 2
0
. (Hint: Find a bijection C 1 1.)
5. (a) Let X

be a family of disjoint sets of the same cardinality .


Show that

= [[.
(b) Let X be a nonempty set. Show that the cardinality of the set of
nite sequences with elements from X has cardinality
0
if X is nite
and cardinality [X[ if X is innite. (Hint: This set can be written as

n=1

n
k=1
X.)
(c) Let X be an innite set and T
f
(X) (resp., T

(X)) be the set of


its nite (resp., innite) subsets. Prove that [T
f
(X)[ = [X[ and
[T

(X)[ = 2
|X|
. (Hint: Use (b) to prove the rst equation, and then
use cardinal addition to derive the second from the rst.)
6. (This exercise will prove the earlier comments about how every innite
set can be made order-isomorphic to innitely many dierent ordinals,
depending on how we choose to well-order it.) Let X be a set and ( be
the class of ordinal numbers in one-to-one correspondence with X. Show
that (a) ( is a set, (b) [([ = 1 if X is nite, and (c) [([ is the successor
cardinal of [X[ if X is innite. (Hint: Let be the successor cardinal of
[X[, and show that ( = ON [ [X[ < and [X[ +[([ = .)
Chapter 2
Group Theory
In this chapter, we will begin our study of algebraic structures. We will go
through the most elementary parts of group theory, covering topics such as
quotient groups, direct products, and isomorphism, and then consider some im-
portant special kinds of groups, mainly the cyclic groups and the permutation
groups. Group theory is kind of a peculiar topic in that the basic, most impor-
tant parts are fairly simple, but studying it any further beyond that becomes
incredibly intricate. So we will be able to do a thorough study of the key things
fairly quickly (in about three weeks), and then move on to studying ring theory.
As I mentioned before, my research area is commutative rings, so you can expect
some bias in that direction later in the course.
2.1 Semigroups, Monoids, and Groups
[Durbin: Sections 3-5, 14]
Denition.
1. An operation on a set S is a function : S S S. We abbreviate
(a, b) = a b.
2. An operation on a set S is:
(a) associative if a (b c) = (a b) c for all a, b, c S; and
(b) commutative if a b = b a for all a, b S.
3. A semigroup is a pair (S, ), where S is a set and is an associative oper-
ation on S. (For simplicity, we will often just refer to S as the semigroup
when it is understood what the operation is.) A semigroup is commutative
if its operation is. Sometimes we refer to the cardinality of a semigroup
as its order.
Remark.
38
39
1. In cases where there is no danger of confusion with some sort of standard
multiplication operation, we will usually name our operation instead of
. (This is standard in actual mathematical practice; Durbin prefers to
use in all cases, as a pedagogical tool to emphasize that these opera-
tions do not necessarily correspond to any sort of familiar multiplication.)
When we are using this multiplicative notation, we will make use of the
standard abbreviations a b = ab and a
n
= a a
. .
n copies
. (We will see below
that it is unambiguous to write products a
1
a
n
in a semigroup with-
out parentheses, and that in a commutative semigroup the order of the
factors does not matter.) It turns out that the familiar exponentiation
properties hold in semigroups, i.e., if S is a semigroup, then for x S and
m, n Z
+
, we have (x
m
)
n
= x
mn
and x
m
x
n
= x
m+n
. These equations
are really just a special case of the fact that it does not matter how one
groups parentheses with an associative operation. For similar reasons, if
S is a commutative semigroup, x, y S, and n Z
+
, then (xy)
n
= x
n
y
n
.
2. The next most common name for an operation is + (especially when
the operation is commutative). When we are using this additive notation,
we will make use of the abbreviation na = a + +a
. .
n copies
.
3. Any operation on a nite set x
1
, . . . , x
n
can be represented in table
form as follows.
x
1
x
2
x
n
x
1
x
2
1
x
1
x
2
x
1
x
n
x
2
x
2
x
1
x
2
2
x
2
x
n
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
x
n
x
n
x
1
x
n
x
2
x
2
n
This is called a Cayley table for the operation. Note the convention re-
garding order of multiplication: the (i, j) entry is x
i
x
j
.
4. The class of all semigroups is a proper class. In fact, the corresponding
classes for all the major algebraic structures we will discuss in this course
(semigroups, monoids, groups, rings, integral domains, and elds) are all
proper, because it will turn out that each of these classes has members of
arbitrarily large cardinality. (We will delay the proof of this for quite a
while.)
Example.
1. Addition and multiplication are commutative and associative operations
on C, but the subtraction operation is neither.
2. Division and exponentiation are operations on 1
+
that are neither com-
mutative nor associative.
40
3. Addition modulo 4 is an operation on 0, 1, 2, 3, with the following Cayley
table.
+
4
0 1 2 3
0 0 1 2 3
1 1 2 3 0
2 2 3 0 1
3 3 0 1 2
It is commutative and associative.
Theorem 26 (Generalized Commutative and Associative Properties).
1. In a semigroup, the values of expressions are unaected by how one groups
parentheses.
2. In a commutative semigroup, the values of expressions are unaected by
order of factors.
Proof. Let S be a semigroup and a
1
, . . . , a
n
S.
1. We need to show that any product of a
1
, . . . , a
n
, with the terms writ-
ten in that order and parentheses inserted in any legal way, is equal
to ( ((a
1
a
2
)a
3
)a
4
)a
n
. The case n 3 is already covered by the
associative property, so assume n 4. This product is an expression
of the form bc, where b (resp., c) is some sort of product of a
1
, . . . , a
k
(resp., a
k+1
, . . . , a
n
), with the terms written in that order, for some k
1, . . . , n 1. By induction, we have b = (( ((a
1
a
2
)a
3
)a
4
)a
k
) and
c = (( ((a
k+1
a
k+2
)a
k+3
)a
k+4
)a
n
) = (a
k+1
(a
k+2
(a
k+3
( a
n2
(a
n1
a
n
))))).
If n = k+1, then we are done, so let us assume n k+2. Then by the asso-
ciative property bc = (ba
k+1
)(a
k+2
(a
k+3
( a
n2
(a
n1
a
n
)))), which by in-
duction (viewing ba
k+1
as one factor) equals ( ((ba
k+1
)a
k+2
)a
k+3
)a
n
,
as desired.
2. Assume that S is commutative. We need to show that any product of
a
1
, . . . , a
n
, with the terms written in any order, is equal to a
1
a
n
. The
case n = 1 is trivial, so let us assume n > 1. This product is an expression
of the form ba
k
, where b is a product whose factors consist of the a
i
s for
i ,= k. By induction, we have b = a
1
a
k1
a
k+1
a
n
. If k = n, then
we are done. Otherwise, by the commutative and associative properties
we have ba
k
= (a
1
a
k1
a
k+1
a
n1
a
k
)a
n
, which by induction equals
(a
1
a
n1
)a
n
, as desired.
Denition.
1. An element 1 (resp., 0) of a semigroup S is an identity (resp., absorbing)
element if 1 a = a 1 = a (resp., 0 a = a 0 = 0) for each a S. (It
is also extremely common to use the symbol e for an identity element, as
Durbin does. If we are using additive notation, then we denote identity
41
and absorbing elements with 0 and , respectively. Then their dening
properties would be written as 0+a = a+0 = a and +a = a+= .)
A semigroup with an identity element is called a monoid. Note that a
semigroup has at most one identity (resp., absorbing) element, because
if x and y are identity (resp., absorbing) elements, then x = x y = y.
When necessary, we will add subscripts to indicate for which semigroup
an element is an identity or absorbing element, e.g., 1
S
or 0
S
.
2. Let S be a monoid. If a, b S and ab = 1, then we say a is a left inverse
of b and b is a right inverse of a. If ab = ba = 1, then a and b are
inverses. An element with an inverse (resp., right inverse, left inverse)
is called invertible (resp., right invertible, left invertible). The invertible
elements are also called units, and the set of units of S is denoted S

.
(The notation U(S) is also common.) We say S is a group if S = S

; a
commutative group is called abelian. In an exercise you will show that if
an element of S has a left inverse a and a right inverse b, then a = b. This
shows that inverses are unique when they exist, so we may denote the
inverse of a S

by a
1
. (In additive notation, we use a for the inverse
of a, and we abbreviate x a = x +a.) Also, in order to determine the
inverse of an element of a group, it suces to nd a left or right inverse.
We note that 1 S

, that (a
1
)
1
= a for each a S

, and that the


inverse of a product of units is given by (u
1
u
n
)
1
= u
1
n
u
1
1
. (This
can be easily proven with induction. The order is important if S is not
commutative.) Thus S

is a group, and is for that reason often called the


group of units of S.
3. If x is a member of a monoid S, we dene x
0
= 1. (In additive notation,
this would be written 0 x = 0.) Observe that with this denition the
previously noted exponentiation rules now apply for all natural number
exponents. If additionally x is a unit, then for each n Z
+
we have
(x
n
)
1
= (x
1
)
n
, and we dene x
n
to be this element. (In additive
notation, this denition would be written (n)x = (nx) = n(x).)
Remark.
1. Note that the symbols 1 and 0 now have two possibly dierent meanings:
the natural numbers 1 and 0, or identity/absorbing elements for the semi-
group we are discussing. (Occasionally, like in (N, ), the two meanings
coincide.) We have to determine from context which interpretation of
these symbols is appropriate.
2. An absorbing element of a monoid is a unit if and only if it is the only
element. (Exercise.) Thus groups with more than one element (called
nontrivial groups) do not have an absorbing element.
Proposition 27 (Exponentiation Rules). Let G be a group, g G, and m, n
Z.
1. (g
m
)
n
= g
mn
. In particular, (g
n
)
1
= g
n
.
42
2. g
m
g
n
= g
m+n
.
We note that the exponentiation rule (xy)
n
= x
n
y
n
is valid only in abelian
groups. (Exercise.)
Proof.
1. We have already noted that this is true if m, n > 0 or if m = n = 1,
and if m = 0 or n = 0, then both sides equal 1. If m > 0 and n < 0,
then (g
m
)
n
= ((g
m
)
n
)
1
= (g
mn
)
1
= g
mn
. If m < 0 and n > 0,
then (g
m
)
n
= ((g
1
)
m
)
n
= (g
1
)
mn
= g
mn
. Finally, if m, n < 0, then
(g
m
)
n
= (((g
m
)
1
)
1
)
n
= (g
m
)
n
= g
mn
.
2. We have already noted that this is true if m, n > 0, and if m = 0 (resp.,
n = 0), then both sides equal g
n
(resp., g
m
). If m > 0, n < 0, and
m + n 0, then n > 0 and g
m
g
n
= (g
m+n
g
n
)g
n
= g
m+n
. If
m > 0, n < 0, and m + n < 0, then (m + n) > 0 and g
n
g
m
=
(g
(m+n)
g
m
)g
m
= g
(m+n)
, and taking inverses yields g
m+n
= g
m
g
n
.
We have now established all cases where m 0. If m < 0, then m > 0
and g
m
g
n
= (g
1
)
m
(g
1
)
n
= (g
1
)
(m+n)
= g
m+n
.
Example. In the following example, I will list several examples of groups and
discuss the various concepts we have dened in this section in relation to these
groups. I will list several facts without proof, and when you are reading through
them you should be thinking about why they are true, in order to develop a
feeling for these concepts.
1. is a commutative semigroup whose operation is the empty function.
However, a monoid cannot be empty.
2. The set 2Z of even integers is a commutative semigroup under , but it is
not a monoid.
3. Let R Z, , 1, C. Then both (R, ) and (R, +) are commutative
monoids. The former has identity 1, absorbing element 0, and its group
of units is 1, 1 if R = Z and R

otherwise. The latter is an abelian


group with identity 0; the inverses are the familiar additive inverses.
4. Let n Z
+
and R Z, , 1, C. The set /
n
(R) of n n matrices
over R forms a monoid under either the usual matrix multiplication or
the usual matrix addition. The latter is an abelian group whose identity
is the zero matrix. The former is commutative if and only if n = 1, its
identity is the identity matrix (hence the name), its absorbing element is
the zero matrix, and its group of units is the general linear group of degree
n over R, which is denoted GL
n
(R) and consists of the matrices whose
determinant is a unit in R.
43
5. If X is a set, then (T(X), ) and (T(X), ) are commutative monoids.
The former has identity X and absorbing element , and in the latter
those elements roles are reversed. These monoids are groups if and only
if X = .
6. If X is a set, then (M(X), ) is a monoid. The identity is the identity
map (hence the name), and there is an absorbing element [X[ 1
M(X) is commutative. The units are the bijections, which are called
permutations of X; thus M(X) is a group if and only if [X[ 1. The
group of units is denoted Sym(X) and called the symmetric group on X;
it is abelian [X[ 2. We will be studying the symmetric groups later.
7. If X is a nonempty set and (S, ) is a semigroup (resp., monoid, group),
then so is (M(X, S), ), where is dened on S in the obvious way:
(fg)(x) = f(x)g(x). If S is a monoid, then 1
M(X,S)
: X S : x 1
S
,
and M(X, S)

= M(X, S

), where the inverse of f M(X, S

) is the
map x f(x)
1
. Note that M(X, S) is commutative if and only if S is.
8. If S

is a family of semigroups (resp., monoids, groups), then so is


the direct product

with the operation (fg)() = f()g(), i.e.,


multiplication is done coordinate-wise. The direct product is commutative
if and only if each S

is. If these are monoids, then the identity of the


direct product is the map 1
S

, and (

. In
other words, the units are the elements with units in each coordinate, and
the inverse of a unit u

is the map u
1

.
9. Let n 1 and Z
n
= 0, 1, . . . , n1. Let +
n
and
n
represent addition and
multiplication modulo n. Then (Z
n
, +) is an abelian group with identity
zero, and (Z
n
, ) is a commutative monoid with identity 1 and absorbing
element 0. The invertible elements of (Z
n
, ) are the ones relatively prime
to n, i.e., those that have no common divisors with n other than 1. We
will return to Z
n
later when we know enough algebra to eciently prove
all of these properties and others as special cases of more general theorems.
10. For another example of a non-abelian group, consider the quaternion group
Q
8
= 1, i, j, k, whose operation is as follows. (I only included the
positive elements in the table; one follows the usual rules for signs to
gure out the other possible multiplications.)
1 i j k
1 1 i j k
i i 1 k j
j j k 1 i
k k j i 1
11. Consider a semigroup with exactly one element x. Then we must have
x
2
= x, so x is the identity element, and x is its own inverse. Therefore
any semigroup with exactly one element is an abelian group. This group
is called the trivial group, and usually written 1 (or 0 if using additive
44
notation). (In some sense, there is only one trivial group, because the
only dierence between these groups is what we name the element, which
is inconsequential.)
So you have now seen that a wide variety of interesting structures are groups.
Part of the beauty and usefulness of abstract algebra is that we will be proving
theorems that can be applied to so many dierent things.
Remark. A semigroup S is left- (resp., right-) cancellative if, whenever a, b, c
S, a is not an absorbing element, and ab = ac (resp., ba = ca), then b = c. A
semigroup is cancellative if it is both left- and right-cancellative. All groups are
cancellative (one can derive the right-hand equations by multiplying by a
1
on
the appropriate side), but a cancellative monoid (even one without an absorbing
element) need not be a group, e.g., (Z

, ).
Exercises.
1. (It follows from this exercise that an element of a monoid is invertible
if and only if it is both left- and right-invertible, and that inverses are
unique when they exist.) Prove that, if an element x of a monoid has a
left inverse a and a right inverse b, then a = b.
2. Show that an absorbing element of a monoid is a unit if and only if it is
the only element.
3. Let G be a group, x, y G, and n Z.
(a) Prove that (xy)
n
= x
n
y
n
if xy = yx. (Hint: The case n > 0 is
already covered in this section, and the case n = 0 is rather trivial.
Deduce the n < 0 case by writing (xy)
n
= ((xy)
1
)
n
.)
(b) Show that, if (xy)
2
= x
2
y
2
or (xy)
1
= x
1
y
1
, then xy = yx. (Thus
the law in (3) holds for all x, y, n G is abelian. Hint: Multiply on
the left and right by appropriate things.)
(c) Show that G is abelian if g
2
= 1 for all g G.
4. Let X be a set. Show that (M(X), ) has an absorbing element [X[ 1
M(X) is commutative. (Hint: (a) (b): Show that an absorbing
element must equal every constant function. (b) (c): [M(X)[ = 1 in
this case. (c) (a): Show that any constant function is an absorbing
element.)
2.2 Subgroups and Cosets
[Durbin: Sections 7 and 16-17, 22]
Denition. A subset H of a monoid S is called a submonoid (resp., subgroup)
of S if it contains 1 and is a monoid (resp., group) with respect to the operation
on S. (Thus 1
H
= 1
S
, and it also follows that a unit of H has the same inverse
45
in H as its inverse in S.) We call H a proper submonoid (resp., subgroup) if
H ,= S. We write H S to indicate that H is a subgroup of S.
Remark.
1. A submonoid/subgroup of a monoid is not the same thing as a subset that
is a monoid/group with respect to the monoids operation. For example,
consider the monoid (1, ). Then 0 is a group under , but it does not
contain 1. However, we will soon see that the contains 1 requirement is
redundant for a subgroup of a group.
2. If S is a monoid, then is a partial order on the set of subgroups of S.
The maximum element is S

and the minimum element is 1. Sometimes


this partially ordered set is called the lattice of subgroups of S, because it
is a special type of partially ordered set called a lattice. We will not be
discussing lattice theory in this course, but occasionally I may ask you to
draw the lattice of subgroups of a group, which means to draw a diagram
illustrating what all the subgroups are and which ones are contained in
which. I will not draw one in these notes because it takes quite a lot of
eort to typeset, but I will draw a few examples in class.
3. A very simple theorem is: Let T be a monoid and _ S T. Then S is
a submonoid of T 1 S and for each a, b S we have ab S. (This
is immediate since the operation is automatically associative on S.) This
result could be called the Submonoid Test.
Theorem 28 (Subgroup Test). Let G be a group and _ H G. The following
are equivalent.
1. H G.
2. H is a group with respect to the operation on G.
3. For each a, b H, (i) ab H, and (ii) b
1
H.
4. For each a, b H, ab
1
H.
Proof. (2) (1): If (2) holds and 1
1
H
denotes the inverse of 1
H
in G, then
1
G
= 1
1
H
1
H
= 1
1
H
(1
2
H
) = (1
1
H
1
H
)1
H
= 1
G
1
H
= 1
H
H. (1) (4): Follows
from the denition of group and the fact that inverses of elements in H coincide
with inverses in G. (4) (3): Assume (4). Because H ,= , there is an h H,
and hence 1 = hh
1
H. Thus for every a, b H we have b
1
= 1 b
1
H
and ab = a(b
1
)
1
H. (3) (2): Assume (3). Then (i) shows that the
associative operation for G is an associative operation on H, and (ii) shows that
every element of H is invertible in H, so all that remains is to show that 1 H,
which follows as in the proof of (4) (3).
Example. The set Z[i] of Gaussian integers consists of the complex numbers
with integer real and imaginary parts, i.e., Z[i] = a + bi [ a, b Z. We could
verify that (Z[i], +) is a group directly, but it is even quicker (assuming we
46
already know that (C, +) is a group) to show that Z[i] C using the Subgroup
Test. For this, we note that Z[i] ,= (for example, 0 Z[i]), and that, for any
a +bi, c +di Z[i], we have (a +bi) (c +di) = (a c) + (b d)i Z[i].
Theorem 29. Let G be a group and H G. Let
H
be the relation on G given
by a
H
b ab
1
H. Then
H
is an equivalence relation. The equivalence
classes are given by [g] = Hg = hg [ h H, and each has cardinality [H[.
Proof. We rst show that
H
is an equivalence relation.
1. (Reexive): For each a G, we have aa
1
= 1 H, so a
H
a.
2. (Symmetric): If a
H
b, then ab
1
H, so ba
1
= (ab
1
)
1
H, and
thus b
H
a.
3. (Transitive): If a
H
b and b
H
c, then ab
1
, bc
1
H, so ac
1
=
(ab
1
)(bc
1
) H, and thus a
H
c.
Now consider an equivalence class [g]. We have a [g] a
H
g ag
1

H a Hg. Therefore [g] = Hg. Finally, we note that the map h hg is a


bijection from H onto Hg, so [Hg[ = [H[.
Denition. Let G be a group and H G. We abbreviate G/
H
= G/H, and
call the members of this set right cosets of H.
Remark. If G is a group and H G, an analogous thing can be done with
the left cosets of H, which are the subsets of the form gH = gh [ h H. We
have a bH b
1
a H a
1
b H, so the left cosets are the equivalence
classes of an equivalence relation analogous to
H
. Analogously to the proof
above, we have [gH[ = [H[ = [Hg[. Finally, the map gH Hg
1
is a bijection
from the set of left cosets of H onto the set of right cosets of H. (Exercise.)
Denition. Let G be a group and H G. The index of H in G is [G : H] =
[G/
H
[, i.e., the number of right (equivalently, left) cosets of H.
Theorem 30 (Lagranges Theorem). Let G be a group and H G. Then
[G[ = [H[[G : H]. Hence [H[ [ [G[ if G is nite.
Proof. Because the right cosets of H are equivalence classes of the equivalence
relation
H
, they form a partition of G. Each right coset has cardinality [H[,
and the cardinality of the set of right cosets is [G : H], so [G[ = [H[[G : H].
Proposition 31. Let H
1
H
2
H
n
be groups. Then [H
n
: H
1
] =

n1
k=1
[H
k+1
: H
k
].
Proof. The result is trivial if n 2, so let us assume n > 2. Let A (resp., B) be a
set consisting of one element from each set in H
n
/H
n1
(resp., H
n1
/H
1
). De-
ne f : A B H
n
/H
1
: (a, b) H
1
(ba). If f(a
1
, b
1
) = f(a
2
, b
2
), then
(b
1
a
1
)(b
2
a
2
)
1
H
1
H
n1
, so H
n1
a
1
= H
n1
(b
1
a
1
) = H
n1
(b
2
a
2
) =
H
n1
a
2
and a
1
= a
2
, and thus H
1
b
1
= H
1
b
2
and b
1
= b
2
. Therefore f is injec-
tive. In fact, since H
n
=

aA
H
n1
a =

aA
_
bB
H
1
b
_
a =

(a,b)AB
H
1
(ba),
47
the latter union must be over all cosets of H
1
in H
n
, so f is a bijection. Thus
[H
n
: H
1
] = [AB[ = [H
n
: H
n1
][H
n1
: H
1
] = [H
n
: H
n1
]

n2
k=1
[H
k+1
: H
k
],
where the last equality is by induction.
Denition. A congruence relation on a group is an equivalence relation such
that ab cd whenever a c and b d.
Example.
1. For each n Z, the relation
n
of congruence modulo n is a congruence
relation on Z (hence the name). (Because the group operation is +, the
property that needs to be veried is that a +b

=n
c +d whenever a

=n
c
and b

=n
d. It is not hard to verify this directly, as you may have done in a
pre-calculus class, but we will wait for a moment to discuss the congruence
modulo n relation as a special case of a more general phenomenon.) We
will later see that these are in fact the only congruence relations on Z.
2. For any semigroup S, the relations = and S S are congruence relations.
3. The only congruence relation on is the empty relation. The empty
relation is not even reexive on non-empty sets, though.
Proposition 32. Let be a congruence relation on a group. If a
1
b
1
, . . . ,
a
n
b
n
, then a
1
a
n
b
1
b
n
.
Proof. By induction on n. The base case n 2 is covered by the denition
of congruence relation, so assume n 3. By induction, we have a
1
a
n1

b
1
b
n1
, and thus a
1
a
n
= (a
1
a
n1
)a
n
(b
1
b
n1
)b
n
= b
1
b
n
.
Theorem 33. Let be a congruence relation on a group G. Then G/ is
a group under the operation given by [a][b] = [ab]. (This group is called the
quotient group or factor group of G by .) The identity is [1] and the inverses
are given by [g]
1
= [g
1
]. If G is abelian, then so is G/ .
Proof. We rst need to check that this operation is well-dened. That is, we
need to show that, if [a] = [c] and [b] = [d], then [ab] = [cd]. (Otherwise, the
denition above would be ambiguous.) But this is simply another way to phrase
the denition of a congruence relation. The rest of the facts listed above are
very simple to verify.
Notation. We usually use the same symbol for the operation on G/ as we
do for G. For example, if the operation on G is +, then we write [a] +[b] instead
of [a][b].
Proposition 34. Let G be a group and N G. The following are equivalent.
1. gN = Ng for each g G.
2. gN Ng for each g G.
3. Ng gN for each g G.
48
4. gNg
1
N for each g G.
Proof. (1) (2): Clear. (2) (4): If gN Ng, then gNg
1
(Ng)g
1
= N.
(4) (3): If g
1
N(g
1
)
1
N, then Ng = g(g
1
N(g
1
)
1
) gN. (3) (1):
If Ng gN and Ng
1
g
1
N, then gN = g(Ng
1
)g g(g
1
N)g = Ng, and
hence gN = Ng.
Denition. A subgroup N of a group G is normal, written N G, if it satises
one (equivalently, all) of the conditions in the above theorem.
Example.
1. For any group G, the subgroups 1 and G are normal. A group in which
these are the only normal subgroups is called simple.
2. Every subgroup of an abelian group is normal.
3. The relation is not transitive. The simplest example of this involves
groups of permutations, which we will study later.
Theorem 35. Let G be a group. The map N
N
is a bijection between the
normal subgroups of G and the congruence relations on G. The inverse map is
[1]

.
Proof. We need to show that, for each N G and each congruence relation
on G: (1)
N
is a congruence relation, (2) [1]

G, (3) [1]

N
= N, and (4)

[1]
=.
1. We have already seen that
N
is an equivalence relation. Now assume
a
N
c and b
N
d. Then ac
1
N and bd
1
N, so (ab)(cd)
1
=
abd
1
c
1
= (a(bd
1
)a
1
)(ac
1
) N, and hence ab
N
cd.
2. 1 [1]

, and for each a, b [1]

, we have ab
1
= (ab
1
) 1 ab
1
b =
a 1. Therefore [1]

G by the Subgroup Test. For each g G and


x [1]

, we have gxg
1
g 1 g
1
= 1, so [1]

G.
3. [1]

N
= N 1 = N.
4. If a b, then ab
1
bb
1
1, and hence a
[1]
b. On the other hand,
if a
[1]
b, then ab
1
1, and thus a = (ab
1
)b eb = b. Therefore

[1]
=.
Denition. Let G be a group and N G. The quotient group or factor group
of G by N is G/
N
. (Note that the identity is N and the inverses are given
by (Ng)
1
= Ng
1
.)
49
Example. In Z, we have n [ (a b) a b nZ, so
n
=
nZ
. (It is easy
to check that nZ Z.) Therefore
n
is a congruence relation, and performing
addition modulo n in Z is the same thing as doing addition in the group Z/nZ.
When working with elements of Z/nZ, it is cumbersome to write expressions of
the form nZ +a, so it is customary to write [a] instead, where from context we
know that this denotes [a]
nZ
.
Remark. If N G and N H G, then H/N G/N, and for g G we
have Ng H/N g H. (To see this, note that, if g G, h H, and
Ng = Nh, then gh
1
N H, so g = (gh
1
)h H.)
Theorem 36. Let G be a group, H, K G, and HK = hk [ h H, k K.
1. HK G HK = KH.
2. If H G or K G, then HK G.
3. If H, K G, then HK G.
4. [HK[[H K[ = [H[[K[.
The equation in (4) is sometimes useful for proving that a subgroup of a
certain size must exist. For example, if we know that G has normal subgroups
H and K of orders 5 and 7, then [H K[ = 1 by Lagranges Theorem and HK
is a normal subgroup of order
57
1
= 35.
Proof.
1. (): Assume HK G. Each element of HK is the inverse of an element
of HK, which by the inverse formula is in KH. Conversely, each element
of KH is the inverse of an element of HK, hence in HK. Therefore
HK = KH. (): Assume HK = KH. Then 1 = 1 1 HK, and for
each h
1
, h
2
H and k
1
, k
2
K we have (h
1
k
1
)(h
2
k
2
)
1
= h
1
k
1
k
1
2
h
1
2

(HK)(KH) = HK
2
H = HKH = H
2
K = HK. By the Subgroup Test,
we conclude HK G.
2. If H G, then HK =

kK
Hk =

kK
kH = KH, so HK G. The
case where K G is similar.
3. Assume H, K G. Then HK G by (2). Moreover, for each g G we
have gHK = HgK = HKg, so HK G.
4. We have H K =

xHK
(h, k) H K [ hk = x, where the sets
in the union are disjoint, so it suces to show that for each h H and
k K there is a bijection HK (h

, k

) HK [ h

= hk. Dene
f : H K H K : t (ht, t
1
k). By the cancellative properties,
we see that f is an injection. Lastly, we observe that if hk = h

, then
h
1
h

= k(k

)
1
H K, h

= h(h
1
h

), and k

= (k(k

)
1
)
1
k. This
shows that f[H K] = (h

, k

) H K [ h

= hk, as desired.
50
Exercises.
1. (a) Prove that the union of a nonempty chain of subgroups of a monoid
is a subgroup.
(b) Let H and K be subgroups of a group G. Show that H K G
H K or K H.
(c) Show that every subgroup of a monoid that is not a group is contained
in a maximal subgroup. (A maximal subgroup of a monoid is a
maximal element among the proper subgroups.)
2. (a) Prove that every nonempty intersection of subgroups of a monoid is
a subgroup.
(b) Prove that, for any subset A of a group, there is a smallest subgroup
containing A.
3. Let G be a group and A G. The centralizer of A is the set C(A) of
elements of G that commute with each element of A, and the center of G
is Z(G) = C(G).
(a) Show that C(A) G. (This is not part of the problem, but note
that Z(G) G.)
(b) Show that A C(C(A)).
4. Let H be a subgroup of a group G. Prove that the map gH Hg
1
is a
bijection between the set of left cosets of H onto the set of right cosets of
H.
5. Let G be a group. The conjugacy relation on G is the relation given by
a b there is a g G with gag
1
= b.
(a) Show that the conjugacy relation is an equivalence relation.
(b) The conjugacy class of an element a G is its equivalence class Cl(a)
under the conjugacy relation. Elements in the same conjugacy class
are called conjugates. Show that [Cl(a)[ = [G : C(a)]. (Here C(a) is
an abbreviation for C(a). Hint: Show that the map f : G/C(a)
Cl(a) : C(a)g g
1
ag is a well-dened bijection.)
(c) Show that a subgroup of G is normal it contains all the conju-
gates of its elements the conjugacy classes of its elements form a
partition of it.
(d) Prove the Class Equation: [G[ = [Z(G)[ +

[G : C(a)], where the


sum is taken over a set consisting of one representative a for each
conjugacy class of an element not in the center. (Hint: What is the
cardinality of the conjugacy class of an element in the center?)
6. Show that every subgroup with index 2 is normal.
51
2.3 Homomorphisms and Isomorphisms
[Durbin: Sections 18, 19, 21, and 23]
Denition.
1. A function : S T between monoids is a homomorphism if (1
S
) = 1
T
and (ab) = (a)(b) for each a, b S. A (pre-)image under a homomor-
phism is called a homomorphic (pre-)image.
2. A homomorphism is a monomorphism (resp., epimorphism, isomorphism)
if it is an injection (resp., surjection, bijection).
3. An endomorphism (resp., automorphism) of a monoid S is a homomor-
phism (resp., isomorphism) S S. We denote the set of endomorphisms
(resp., automorphisms) of S by End(S) (resp., Aut(S)).
Remark.
1. In the second half of the course, we will move from group theory to dis-
cussing another algebraic structure known as a ring, and we will also dene
the various morphisms for rings. If necessary, we will use phrases such
as monoid homomorphism, group homomorphism, or ring homomor-
phism to distinguish between the concepts, if what is meant is not clear
from context.
2. Analogous to the situation we encountered with the denitions of sub-
monoid and subgroup, the requirement that (1) = 1 is an essential
part of the denition of a monoid homomorphism, but may equivalently
be left out of the denition of a group homomorphism. We will prove this
momentarily.
3. The importance of isomorphisms is that they give us a precise way to say
when two groups that are technically unequal are actually the same group,
for all algebraic intents and purposes. When there is an isomorphism from
one group onto another, the second groups elements are essentially just
the rst groups elements, given dierent names, but interacting with each
other in the same way.
Example.
1. The groups 1 and 1 0 are, for all algebraic intents and purposes,
the same group. The element x 1 corresponds to (x, 0) 1, and
addition is carried out analogously: (x, 0) + (y, 0) = (x + y, 0). These
observations are expressed in a more precise manner by saying that the
map 1 1 0 : x (x, 0) is an isomorphism.
2. Every group has the identity map as an automorphism.
3. For any groups G and H, the map g 1
H
is a homomorphism, called the
trivial homomorphism.
52
4. The map Z Z : n n is an automorphism. More generally, if G is
a group, then the map g g
1
is an automorphism if and only if G is
abelian. To see this, note that the map is always a bijection (its inverse
is itself), and then apply Exercise 2.1.3.
5. If G is a group and g G, then the map x gxg
1
is an automorphism
of G (exercise), called an inner automorphism.
6. If G is a group and H G, then the inclusion map : H G is a
monomorphism.
7. If G is a group and N G, then the natural epimorphism of G onto G/N
is the epimorphism : G G/N : g Ng.
8. If R Z, , 1, C and n Z
+
, then det : GL
n
(R) R is a homomor-
phism.
Proposition 37. Let S be a monoid and G be a group. A map : S G is a
homomorphism if and only if (ab) = (a)(b) for each a, b S.
Proof. (): Clear. (): Assume (ab) = (a)(b) for every a, b S. Then
(1
S
) = (1
2
S
) = (1
S
)
2
, and multiplying by (1
S
)
1
yields 1
G
= (1
S
). There-
fore is a homomorphism.
Proposition 38. Let : S T be a homomorphism.
1. (a
1
a
n
) = (a
1
) (a
n
) for each a
1
, . . . , a
n
S.
2. [S

] T

. The inverses are given by (u)


1
= (u
1
).
3. (u
n
) = (u)
n
for each u S

and n Z.
Proof.
1. The case n = 1 is trivial, and if n > 1, then by induction we have
(a
1
a
n
) = (a
1
a
n1
)(a
n
) = (a
1
) (a
n1
)(a
n
).
2. For u S

, we have (u)(u
1
) = (uu
1
) = (1) = 1, so (u)
1
=
(u
1
).
3. The case n > 0 is a special case of (1), and (u
0
) = (1) = 1 = (1)
0
. If
n < 0, then by (1) and (2) we have (u
n
) = ((u
n
)
1
) = (u
n
)
1
=
((u)
n
)
1
= (u)
n
.
Proposition 39.
1. A composition of homomorphisms (monomorphisms, epimorphisms, iso-
morphisms) is a homomorphism (resp., monomorphism, epimorphism,
isomorphism).
53
2. The inverse of an isomorphism is an isomorphism.
3. If S is a monoid, then End(S) is a submonoid of (M(S), ) and Aut(S) =
End(S)

. (In particular, Aut(S) is a group.)


4. Homomorphic images and pre-images of monoids (resp., groups) are monoids
(resp., groups).
5. Homomorphic pre-images of normal subgroups are normal subgroups.
Proof.
1. We have already shown that a composition of injections (resp., surjections,
bijections) is an injection (resp., surjection, bijection), so it suces to show
that a composition of homomorphisms is a homomorphism. If : S
1
S
2
and : S
2
S
3
are homomorphisms, then ( )(1) = (1) = 1, and for
each a, b S
1
, we have ( )(ab) = ((a)(b)) = ( )(a)( )(b),
as desired.
2. Let : S T be an isomorphism. Then for each a, b T we have
(
1
(a)
1
(b)) = (
1
(a))(
1
(b)) = ab = (
1
(ab)), and by in-
jectivity
1
(a)
1
(b) =
1
(ab). Because
1
(1) = 1 by injectivity, we
conclude that
1
is an isomorphism.
3. Of course we have id End(S), and a composition of homomorphisms
is a homomorphism by (1), so End(S) is a submonoid of M(S). Finally,
we have Aut(S) = End(S) M(S)

= End(S)

, where the rst equality


follows from the denition of an automorphism and the second follows
from (2).
4. We prove the group case. A small modication produces the proof of
the monoid case. Let : G
1
G
2
be a group homomorphism. If
H
1
G
1
, then 1 = (1) (H
1
), and, for each a, b H
1
, we have
(a)(b)
1
= (ab
1
) (H
1
), so [H
1
] G
2
by the Subgroup Test.
If H
2
G
2
, then 1
1
[H
2
], and, for each a, b
1
[H
2
], we have
(ab
1
) = (a)(b)
1
H
2
, so
1
[H
2
] G
1
by the Subgroup Test.
5. Let : G H be a group homomorphism and N H. We have

1
[N] G by (4). For each g G an x
1
[N], we have (gxg
1
) =
(g)(x)(g)
1
N, so
1
[N] G.
Denition. Let S and T be monoids. We say S is isomorphic to T, written
S

= T, if there is an isomorphism S T. It follows from the above proposition
that

= is an equivalence relation on the class of all monoids. (Also on the class
of all groups.)
Remark. Assume there is a monomorphism : S T. Then S

= [S] T.
So, in some sense, there is a submonoid of S that is a copy of T. For this reason,
we say that S is embedded in T via , and that is an embedding.
54
Denition. The kernel of a homomorphism : S T is ker =
1
[1
T
].
Proposition 40. A group homomorphism is a monomorphism ker =
1.
Proof. (): We have (1) = 1, so ker = 1 if is a monomorphism. (): As-
sume ker = 1. If (a) = (b), then (ab
1
) = (a)(b)
1
= (a)(a)
1
=
1, so ab
1
= 1, and hence a = b. Therefore is a monomorphism.
Proposition 41. Let G be a group. A subset of G is the kernel of a homomor-
phism if and only if it is a normal subgroup of G.
Proof. (): The kernel of a homomorphism is the homomorphic pre-image of
the normal subgroup 1. (): Let N G and : G G/N : g Ng. Then
is a homomorphism and ker = N.
Theorem 42 (Fundamental Homomorphism Theorem). Let : G H be a
group homomorphism and N = ker . Then : G/N [G] : Ng (g) is
an isomorphism. Consequently, we have [[G][ = [G : N].
Proof. Note that Na = Nb ab
1
N (a)(b)
1
= (ab
1
) = 1
H

(a) = (b). This shows that is well-dened and injective. Because is
clearly a surjection, all that remains is to show that it is a homomorphism. If
a, b G, then ((Na)(Nb)) = (N(ab)) = (ab) = (a)(b) = (Na)(Nb),
as desired.
Theorem 43 (Diamond Isomorphism Theorem). Let G be a group, H G,
and N G. Then HN/N

= H/(H N).
Proof. (Note that N HN G, and that HN H, so the statement of the
theorem makes sense.) Dene : H HN/N : h Nh. By the Fundamental
Homomorphism Theorem, it suces to show that is an epimorphism with
kernel H N. For each h H and n N, we have (h) = Nh = (Nh)(Nn) =
N(hN), so is a surjection. For each a, b H, we have (ab) = N(ab) =
(Na)(Nb) = (a)(b), so is an epimorphism. Finally, we have (h) = N
hN = N h H N, so ker = H N.
Theorem 44 (Cancelation Isomorphism Theorem). Let G be a group, K
N G, and K G. Then
G/K
N/K

= G/N.
Proof. Dene : G/K G/N : Kg Ng. By the Fundamental Homomor-
phism Theorem, it suces to show that is a well-dened epimorphism with
kernel N/K. (This will also show that N/K G/K, so the statement of the the-
orem does indeed make sense.) If Ka = Kb, then ab
1
K N, so Na = Nb.
Therefore is well-dened. The map is clearly a surjection, and for each a, b
G we have ((Ka)(Kb)) = (K(ab)) = N(ab) = (Na)(Nb) = (Ka)(Kb), so
is an epimorphism. Finally, we have (Kg) = N Ng = N g N, so
ker = N/K.
55
Theorem 45 (Correspondence Theorem). Let G be a group and N G. The
map H H/N is an order-preserving one-to-one correspondence between the
subgroups of G containing N and the subgroups of G/N. The inverse map is
H
1
[H], where : G G/N : g Ng. Both maps take normal subgroups
to normal subgroups.
Proof. Note that (i) is an epimorphism, (ii) H/N = [H] G/N whenever
N H G, and (iii) N =
1
[N]
1
[H] G for H G/N. We already
know that homomorphic pre-images of normal subgroups are normal subgroups,
and the identity (Ng)(H/N)(Ng
1
) = (gHg
1
)/N shows that the former map
also takes normal subgroups to normal subgroups. It is clear that the maps are
order-preserving, so all that remains is to show that the two maps are inverses.
Because is a surjection, for each H G/N we have H = [
1
[H]] =

1
[H]/N. The proof will be complete once we show that
1
[H/N] = H
whenever N H G. Certainly H
1
[H/N]. Now pick any x
1
[H/N].
Then Nx = Nh for some h H, so xh
1
N. Therefore x = (xh
1
)h H, as
desired.
Exercises.
1. Verify that inner automorphisms are actually automorphisms.
2.4 Cyclic Groups
[Durbin: 15, 17]
Denition. Let G be a group.
1. If A G, then the smallest subgroup subgroup of G containing A (which
exists by an exercise) is called the subgroup generated by A, and denoted
A. We abbreviate a
1
, . . . , a
n
= a
1
, . . . , a
n
. If G = A for some
nite subset A, we call G nitely generated, and if G = g for some g G,
we say G is cyclic, and that g is a generator. Thus, for any g G, we call
g the cyclic subgroup generated by g.
2. Let g G. If there is an n Z
+
with g
n
= 1, then the order or period of
g, denoted [g[ or o(g), is the least such n. Otherwise, we write o(g) =
0
.
(This is somewhat nonstandard, but it will make some formulas work out
nicely. The notation o(g) = is much more common.)
Proposition 46. Let G be a group and g G.
1. o(g) <
0
if and only if there are distinct integers m, n Z with g
m
= g
n
.
In other words, o(g) =
0
if and only if the map Z G : n g
n
is
injective.
2. g
n
= e o(g) [ n.
56
3. There are exactly o(g) distinct powers of g. If o(g) is nite, these are
1, g, g
2
, . . . , g
o(g)1
.
4. g = g
n

nZ
.
5. [g[ = o(g), so [G[ = [G : g]o(g).
6. If G is nite, then g
|G|
= 1.
Proof.
1. (): If o(g) <
0
, then o(g) Z
+
and g
0
= g
o(g)
. (): If m < n and
g
m
= g
n
, then g
nm
= g
m
g
m
= 1, so o(g) n m <
0
.
2. (): Assume g
n
= 1. We can write n = qo(g) + r, where q, r Z
and 0 r < o(g). (This result from basic algebra is called the Division
Algorithm. We will prove it later.) Then 1 = g
n
= (g
o(g)
)
q
g
r
= g
r
, so
r = 0 by the denition of o(g), and hence o(g) [ n. (): If o(g) [ n, then
there is a q Z with n = qo(g), and g
n
= (g
o(g)
)
q
= 1.
3. The case o(g) =
0
is covered by (1), so let us assume o(g) <
0
. By the
proof of (2), we see that each power of g is one of 1, g, g
2
, . . . , g
o(g)1
. If
0 m < n o(g) 1 and g
m
= g
n
, then g
nm
= 1 and n m < o(g), a
contradiction. Therefore these elements are distinct.
4. Because g
n

nZ
g, by minimality it suces to show g
n

nZ
G,
which follows from the Subgroup Test after observing that g
m
(g
n
)
1
=
g
mn
.
5. Follows from (3) and (4).
6. Follows from (2) and (5).
Example.
1. The group Z is cyclic, with generator 1.
2. For n Z
+
, the group Z/n = Z/nZ is cyclic. It is generated by [1],
which has order n, so Z/nZ = [0], . . . , [n 1].
3. More generally, if G = g, then any homomorphic image of G is generated
by the image of g. (This is a special case of an exercise.)
Theorem 47 (Isomorphism Classes of Cyclic Groups). Let G be a cyclic group.
1. If [G[ = n Z
+
, then G

= Z/nZ.
2. If G is innite, then G

= Z.
Thus there is (up to isomorphism) exactly one cyclic group of each countable
positive cardinality, and there are no uncountable cyclic groups.
57
Proof. Let g be a generator for G, and note that [G[ = o(g). Consider the map
: Z G : n g
n
. By the above proposition, this map is a surjection, and it
is in fact an epimorphism since (m+n) = g
m+n
= g
m
g
n
for each m, n Z. If
G is innite, then is an isomorphism by the above proposition. On the other
hand, if [G[ = n Z
+
, then the above proposition gives ker = nZ, and thus
G

= Z/nZ.
Proposition 48. Every subgroup of a cyclic group is cyclic.
Proof. Let H G = g. If H = 1, then it is cyclic, so assume H ,= 1.
Then there is an n Z

with g
n
H, and we can pick such an n with [n[
minimum. If g
m
H, then the Division Algorithm gives m = nq + r, where
0 r < [n[, so g
m
= g
nq+r
= g
r
, and by minimality we have r = 0 and
g
m
= (g
n
)
q
. Therefore H = g
m
.
Theorem 49. Let G be a group, g G, and k Z
+
.
1. If o(g) =
0
, then o(g
k
) =
0
.
2. If o(g) = n Z
+
, then o(g
k
) =
n
gcd(k,n)
.
Proof.
1. By contrapositive. If o(g
k
) <
0
, then g
ko(g
k
)
= 1 and o(g) [ ko(g
k
).
2. Assume o(g) = n Z
+
. We have g
ko(g
k
)
= 1, so n [ ko(g
k
) and
n
o(g
k
)
[ k.
(Some notation: for a, b Z, we say a divides b, and write a [ b, if b is a
multiple of a.) Hence
n
o(g
k
)
[ gcd(k, n) and
n
gcd(k,n)
[ o(g
k
). On the other
hand, we have (g
k
)
n/ gcd(k,n)
= (g
n
)
k/ gcd(k,n)
= 1, so o(g
k
) [
n
gcd(k,n)
, and
thus o(g
k
) =
n
gcd(k,n)
.
Theorem 50 (Fundamental Theorem of Finite Cyclic Groups). Let n Z
+
.
1. The generators of Z/nZ are the elements of the form [a] with gcd(a, n) = 1.
2. There is exactly one subgroup of Z/nZ of each order dividing n.
Proof. Part (1) is immediate from the above proposition. Now assume d [ n.
By the proposition, we have o([n/d]) =
n
gcd(n/d,n)
=
n
n/d
= d, so [[n/d][ = d.
For uniqueness, let H be any subgroup of order d. Because H is cyclic, we
may write H = [k], and d =
n
gcd(k,n)
, so gcd(k, n) =
n
d
and
n
d
[ k. Thus
[k] [n/d], and in fact equality holds because the subgroups are nite sets
of the same cardinality.
Eulers totient function is the function : Z
+
Z
+
that takes n to the
number of positive integers n that are relatively prime to it. The above
theorem then says that the number of generators of Z/nZ is (n). (We can
rewrite [0] as [n], so that Z/nZ = [1], . . . , [n].) Also, if d [ n, then Z/nZ has
58
exactly one subgroup of order d, so each element of order d generates that same
subgroup, and there are (d) such elements. Because all the elements of Z/nZ
have order dividing n, this leads us to the following number theoretic result.
Corollary 51 (Eulers Formula). n =

d|n
(d) for n Z
+
.
The following powerful result characterizes all nitely generated abelian
groups up to isomorphism. Henceforward, to save a little space on writing
we will abbreviate Z
n
= Z/nZ, and we will abbreviate the coset k + nZ as [k]
(or [k]
n
if necessary to avoid ambiguity).
Theorem 52 (Fundamental Theorem of Finite Abelian Groups). Let G be a
nontrivial nite abelian group.
1. G is isomorphic to a unique (up to order) nite direct product of the form
Z
m1
Z
mn
, where n Z
+
and m
i
[ m
i+1
for 1 i n 1. The
m
i
s are called the invariant factors of G.
2. G is isomorphic to a unique (up to order) nite direct product of the form
Z
q1
Z
qn
, where n Z
+
and the q
i
s are powers of primes. The q
i
s
are called the elementary divisors of G.
This is such an important result that you would be cheated if I did not
mention it, but on the other hand its proof is fairly long and intricate (by
our standards) and would require about two full sections to itself, and none of
our future results will rely on this theorem, so we will be omitting its proof.
(Probably the best way to prove it is to develop it as a corollary to a more
general structure theorems about modules, which we will only briey discuss in
this class.) However, you will be expected to be able to apply it to determine
how many abelian groups there are (up to isomorphism) of a particular order,
as I do in the following example.
Example. List all the abelian groups of order 12 (up to isomorphism).
The rst step is to completely factor 18 = 2 3
2
. Then write out all possible
ways to express 12 as a product of prime powers: 18 = 233 and 18 = 29. Using
the elementary divisor representations, we see that the only abelian groups of
order 12 are (up to isomorphism) Z
2
Z
3
Z
3
and Z
2
Z
9
.
This answers the question, but suppose that for some reason you would prefer
to know the invariant factor representations. You will show in an exercise that
Z
m
Z
n

= Z
mn
if and only if gcd(m, n) = 1. So this allows us to rewrite
Z
2
Z
3
Z
3

= Z
3
Z
6
and Z
2
Z
9

= Z
18
. The invariant factor representations
have the advantage that they are shorter and always present a cyclic group as
a single factor.
Exercises.
1. Let G be a group and A G.
59
(a) Show that A is the set of all elements that can be written as prod-
ucts involving only 1, elements of A, and inverses of elements of A.
(b) Prove that, if : G H is a group homomorphism and A G, then
[A] = [A].
(c) Let A be a subset of a group. Prove that [A[
0
if A is nite, and
that [A[ = [A[ if A is innite. (Hint: Part (a) gives an injection
from A into the set of nite sequences with entries taken from the
set A1a
1
[ a A. Use previous exercises about cardinalities
of sets of nite sequences.)
(d) Prove that the group 1 is not countably generated, i.e., not generated
by a countable subset.
(e) Prove that there is no group epimorphism 1. (Hint: This can
be proven by noting that is countably generated, but 1 is not by
(c), and then citing (b). However, I realized that this is a dumb
problem and there is an even easier solution that does not need any
group theory: There is not even a surjection 1, due to the
cardinalities of and 1.)
2. (a) Prove that every nitely generated subgroup of (, +) is cyclic. (Hint:
Given a nite subset of , put its elements over a common denomina-
tor, then consider the subgroup of Z generated by the numerators.)
(b) Prove that is not nitely generated. (Hint: Show that is not
cyclic and use part (a).)
(c) Prove that there is no epimorphism Z .
3. (a) Prove that a group G is abelian if and only if G/Z(G) is cyclic.
(Hint: For the nontrivial direction , let x, y G. Use the fact
that G/Z(G) is cyclic to show that there is a g G and m, n Z
with xg
m
, yg
n
Z(G). Then write xy = (xg
m
)(g
m
)(yg
n
)(g
n
)
and show that this equals yx.)
(b) Show that every group of prime order is cyclic. (Hint: Use Lagranges
Theorem.)
(c) Show that every group of prime-squared order is abelian. (Hint: Use
the Class Equation to show that [Z(G)[ = p or p
2
.)
(d) Give an example of a non-abelian group of prime-cubed order. (Hint:
Look at the examples of groups in Section 2.1.)
(e) Prove that every group of order at most 5 is abelian.
4. Let G be a group of even order. Show that the number of elements in g
of order 2 is odd. (Hint: Show that g, g
1

gG
is a partition of G.)
5. (a) Let G
1
, . . . , G
n
be groups. Show that the order of an element (a
1
, . . . , a
n
)
G
1
G
n
is lcm(o(a
1
), . . . , o(a
n
)). (Recall that the least common
multiple of positive integers x
1
, . . . , x
k
is the unique positive multiple
60
of x
1
, . . . , x
k
that divides all other such multiples. Our convention
for the case where o(a
i
) =
0
is to dene lcm(o(a
1
), . . . , o(a
n
)) =
0
.
We will discuss GCDs and LCMs in more detail later.)
(b) Show that Z
m
Z
n

= Z
mn
if and only if gcd(m, n) = 1. (Hint: You
may use the fact that gcd(m, n)lcm(m, n) = mn. We will prove this
at some later point.)
6. List all abelian groups of order 120 (up to isomorphism), in both elemen-
tary divisor and invariant factor form.
7. Prove that a group is innite if and only if it has innitely many subgroups.
(Hint: For , let G be an innite group. If G has an element of
innite order, then G has a subgroup isomorphic to Z, which in turn has
innitely many subgroups. If Ghas no element of innite order, recursively
construct an innite sequence of distinct cyclic subgroups.)
2.5 Cauchys Theorem and Other Assorted Facts
[Not in Durbin.]
This short section consists of Cauchys Theorem and a couple other random
important facts that we will soon need but which I neglected to include in earlier
sections.
Theorem 53 (Internal Direct Products). Let G be a group and H
1
, . . . , H
n

G. The following are equivalent.
1. H
i
C(H
j
) for i ,= j, and every element of G can be uniquely written in
the form h
1
h
n
, where each h
i
H
i
.
2. H
i
C(H
j
) for i ,= j, G = H
1
H
n
, and each H
i
H
1
H
i1
H
i+1
H
n
=
1.
3. H
1
, . . . , H
n
G, G = H
1
H
n
, and each H
i
H
1
H
i1
H
i+1
H
n
=
1.
In this case, we say that G is an internal direct product of the H
i
s, which
we call direct factors of G, and we indicate this by writing G = H
1
H
n
.
(This is a slight abuse of notation since G is not literally equal to that direct
product, but the two are isomorphic via h
1
h
n
(h
1
, . . . , h
n
).)
Proof. (1) (2): Assume (1) holds. Then clearly G = H
1
H
n
. If h
i
=
h
1
h
i1
h
i+1
h
n
, where each h
j
H
j
, then 1 1 h
i
1 1 = h
1
h
i1

1 h
i+1
h
n
, and by the uniqueness of the representations we have h
i
= 1.
This shows that each H
i
H
1
H
i1
H
i+1
H
n
= 1. (2) (3): Assume
(2) holds. Because elements from distinct H
j
s commute, for x H
i
, h
1
H
1
,
. . . , and h
n
H
n
, we have:
(h
1
h
n
)x(h
1
h
n
)
1
= (h
1
h
i1
h
i+1
h
n
)(h
i
xh
1
i
)(h
1
h
i1
h
i+1
h
n
)
1
= h
i
xh
1
i
H
i
.
61
Therefore each H
i
G. (3) (1): Assume (3) holds. For i ,= j, a H
i
, and
b H
j
, we have aba
1
b
1
= (aba
1
)b
1
H
j
and aba
1
b
1
= a(ba
1
b
1
)
H
i
, so aba
1
b
1
= 1 and ab = ba. Therefore H
i
C(H
j
) for i ,= j. Certainly
every element of G can be expressed in the form h
1
h
n
, where each h
i
H
i
.
For uniqueness, assume a
1
a
n
= b
1
b
n
, where each a
i
, b
i
H
i
. Then
a
n
b
1
n
= (a
1
a
n1
)
1
(b
1
b
n1
) H
n
H
1
H
n1
= 1, so a
n
= b
n
,
and hence a
1
a
n1
= b
1
b
n1
. Note that H
1
H
n1
G by a the-
orem from a previous section, that H
1
, . . . , H
n1
H
1
H
n1
, and that
H
i
H
1
H
i1
H
i+1
H
n1
= 1 for 1 i n 1. So by induction we
have a
i
= b
i
for 1 i n 1, as desired.
Here is another useful but very simple result concerning centralizers.
Proposition 54. Let G be a group and A, B G.
1. For A, B G, we have A C(B) B C(A).
2. For A, B G, we have A C(B) A C(B).
3. For A G, we have C(A) = C(A).
Proof.
1. Both statements simply say that every element of A commutes with every
element of B.
2. Immediate from the fact that C(B) is a group (past exercise).
3. From (1) and (2), we have x C(A) A C(x) A C(x) x
C(A).
Example. Recall that we have proven that a nite cyclic group has an element
of each order dividing its order. This result is not true for all groups (or even all
abelian groups), since [Z
2
Z
2
[ = 4 but by a homework exercise the elements
of Z
2
Z
2
are all of order 1 or 2. However, we do have the following weaker,
but still incredibly useful, result.
Theorem 55 (Cauchys Theorem). Let G be a nite group and p be a prime
dividing [G[. Then G has an element of order p.
Proof. If [G[ = p, then G is cyclic, so let us assume [G[ > p. The Class Equation
gives [G[ = [Z(G)[ +

[G : C(a)], where the sum is taken over a set consisting


of one representative of each conjugacy class of an element not in the center.
If p does not divide some [G : C(a)], then the equation [G[ = [C(a)[[G : C(a)]
shows that p [ [C(a)[, and we have [C(a)[ < [G[ since a / Z(G), so by induction
C(a) (and hence G) has an element of order p. (Recall this fact about primes:
if p [ mn, then p [ m or p [ n. We will discuss this later when we get to
factorization theory.) So let us assume p divides each [G : C(a)]. Then p [
62
[G[

[G : C(a)] = [Z(G)[. So [Z(G)[ p > 1, and there is some non-identity


element a Z(G). Then [Z(G)[ = [Z(G) : a]o(a), so p [ o(a) or p [ [G : a].
If p [ o(a), then a (and hence G) has an element of order p, so let us assume
p [ [Z(G) : a]. Because o(a) > 1, we have [Z(G)/a[ =
|Z(G)|
o(a)
< [Z(G)[ [G[,
so by induction the group Z(G)/a has an element ax of order p. (Note that
Z(G) is abelian, so a Z(G).) But (ax)
o(x)
= a, so p [ o(x), and thus x
(and hence G) has an element of order p.
Exercises. Note: These exercises do not use any results from this section.
They are both fairly simple consequences of denitions/results from the section
on subgroups.
1. Let G be a group, H G be a subgroup of nite index, and a H be
an element with a nite number m of conjugates in G. Prove that the
number of conjugates of a in H is a number of the form
km
[G:H]
, where
1 k [G : H]. (Hint: Recall that m = [Cl
G
(a)[ = [G : C
G
(a)] and
the number n of conjugates of a in H is [H : C
H
(a)]. [As usual, we add
subscripts to symbols to indicate which group it is with respect to.] Write
[G : C
H
(a)] in two dierent ways to get an equation relating m, n, and
[G : H].)
2. Let G be a group, H G, and N G. Show that H N H. (Hint:
We get H N H more or less immediately from a past exercise about
intersections of subgroups. Normality is simple to verify directly from any
of the equivalent denitions of a normal subgroup.)
2.6 Permutations
[Durbin: Sections 6 and 20]
Denition. The symmetric group on a set X is the group Sym(X) = Perm(X) =
M(X)

; its elements are called permutations on X. For x X and Sym(X),


we say xes x if (x) = x, and otherwise say moves x. We will often use
multiplicative notation for the group Sym(X), with the understanding that the
operation is actually function composition but writing things multiplicatively
saves some eort.
Proposition 56. Let X and Y be sets.
1. If [X[ [Y [, then Sym(X) can be embedded in Sym(Y).
2. If [X[ = [Y [, then Sym(X)

= Sym(Y).
Proof.
1. Assume [X[ [Y [. Then there is an injection h : X Y . Dene
: Sym(X) Sym(Y) so that (f) = h f h
1
on h(X) and equals
id
Y
elsewhere. (Note that each (f) is indeed a permutation on Y .) If
63
(f
1
) = (f
2
), then h f
1
h
1
= h f
2
h
1
, and composing by h on
the right and h
1
on the left gives f
1
= f
2
. For each f
1
, f
2
Sym(X), we
have (f
1
f
2
) = h(f
1
f
2
) h
1
= (hf
1
h
1
) (hf
2
h
1
) on h(X)
and equals id
Y
= id
Y
id
Y
elsewhere, and thus (f
1
f
2
) = (f
1
) (f
2
).
Therefore is a monomorphism.
2. Assume [X[ = [Y [. Then there is a bijection h : X Y . Dene a
monomorphism as in the proof of (1). For each g Sym(Y), we have
h
1
g h Sym(X), and (h
1
g h) = h (h
1
g h) h
1
= g, so
is in fact an isomorphism.
Remark. Let X Y be sets. We can extend f Sym(X) to a permutation on
Y by dening f(y) = y for y Y X. (This corresponds to choosing h = in the
proof of (1) above.) This gives us a natural way to identify each permutation
on X with a permutation on Y , and thus we will typically regard Sym(X) as a
subgroup of Sym(Y).
Theorem 57 (Cayleys Theorem). Any group G can be embedded in Sym(G).
Proof. Dene : G Sym(G) so that (g)(x) = gx. (Note that each (g) is
indeed a permutation; its inverse is the map (g
1
).) For each g, h, x G, we
have (gh)(x) = (gh)x = g(hx) = ((g) (h))(x), so is a homomorphism.
Finally, if (g)(x) = x, then gx = x and g = 1, so is a monomorphism.
Denition. For n Z
+
, we abbreviate Sym(1, . . . , n) = S
n
. For m n,
we may consider S
m
to be a subgroup of S
n
by the identication given above.
More explicitly, we may consider S
m
to be the subgroup of S
n
consisting of
those permutations that x the elements m+ 1, . . . , n.
Notation. We can represent an element of S
n
in two-row form by writing a
2n matrix whose (1, j) entry is j, and whose (2, j) entry is (j). For example,
_
1 2 3
2 3 1
_
is the element of S
3
that takes 1 to 2, 2 to 3, and 3 to 1.
Theorem 58. [S
n
[ = n!.
Proof. From the two-row form for permutations, we see that S
n
is in one-to-one
correspondence with the re-orderings of the list 1, 2, . . . , n (i.e., permutations
in the pre-calculus sense). In pre-calculus math courses, you learned that the
number of such re-orderings is n!.
Corollary 59. For each n Z
+
, there are (up to isomorphism) only nitely
many groups of order at most n.
Proof. Let n Z
+
. If G is a group and [G[ = m n, then G embeds into
Sym(G)

= S
m
S
n
. Thus every group of order at most n is isomorphic to a
subgroup of S
n
, and there are only nitely many subgroups of S
n
.
64
Remark. From the above proof, an upper bound for the number of isomorphism
classes of groups of order at most n is 2
n!
, since that is the number of subsets of
S
n
. But this is an extremely poor upper bound, in the sense that it is far larger
than necessary. For example, there are (up to isomorphism) only six groups of
order at most 5 (namely 1, Z
2
, Z
3
, Z
4
, Z
2
Z
2
, and Z
5
), but 2
5!
1.310
36
.
The reason why this upper bound is so much bigger than necessary is that not
nearly every subset of S
n
is a subgroup, and because many of the subgroups of
S
n
have order larger than n.
In a past homework exercise, you showed that any subgroup of index 2 is
normal. The following is a generalization.
Theorem 60. Let G be a nite group, H G, and [G : H] = p, where p is the
smallest prime dividing [G[. Then H G.
Proof. Dene : G Sym(G/H) by (g)(Hx) = Hxg
1
. (Note that each (g)
is indeed a permutation on G/H; its inverse is (g
1
).) Then for g, h, x G
we have (gh)(Hx) = Hx(gh)
1
= (Hxh
1
)g
1
= ((g) (h))(x), so is a
homomorphism. Also note that ker H, because if (g) = id, then Hg
1
= H
and hence g H. We have p[H : ker ] = [G : H][H : ker ] = [G : ker ] =
[[G][ [ [Sym(G/H)[ = p!, so [H : ker ] [ (p 1)!. Also, we have [H : ker ] [
[H[ [ [G[. As we will discuss when we get to factorization theory, every positive
integer can uniquely (up to order) be written as a (possibly empty) product
of primes, and the unique such representation of (p 1)! involves no primes
greater than or equal to p. Because every prime dividing [G[ is greater than or
equal to p, we thus have gcd((p 1)!, [G[) = 1, and hence [H : ker ] = 1 and
H = ker G.
Denition.
1. Let k 2 and a
1
, . . . , a
k
be distinct elements of 1, . . . , n. Then
(a
1
, a
2
, . . . , a
k
) denotes the permutation in S
n
that moves a
1
to a
2
, a
2
to a
3
, . . . , a
k1
to a
k
, a
k
to a
1
, and xes all other elements. We call this
element a cycle of length k, or a k-cycle.
2. Two cycles (a
1
, . . . , a
m
) and (b
1
, . . . , b
k
) are disjoint if a
i
,= b
j
for all i, j.
3. A 2-cycle is called a transposition.
4. The identity permutation is written as (1), but is not considered a cycle.
Remark. Note that a cycle (a
1
, . . . , a
k
) can be written in exactly k distinct
ways, namely (a
1
, . . . , a
k
) = (a
2
, a
3
, . . . , a
k
, a
1
) = = (a
k1
, a
k
, a
1
, . . . , a
k2
) =
(a
k
, a
1
, . . . , a
k1
). That is, the element that one selects to write rst does not
matter, only which element follows which.
Example.
1. S
1
= (1).
65
2. S
2
= (1), (1, 2) = (1, 2)

= Z
2
.
3. S
3
= (1), (1, 2), (1, 3), (2, 3), (1, 2, 3), (1, 3, 2). We have (1, 2)(1, 3) =
(1, 3, 2) and (1, 3)(1, 2) = (1, 2, 3), so S
3
is non-abelian. All groups of
order 5 are abelian (past exercise), so S
3
is the smallest non-abelian
group (in terms of cardinality). It turns out that Z
6
and S
3
are (up to iso-
morphism) the only groups of order 6. More generally, if p < q are primes
with p [ q 1, then there are (up to isomorphism) exactly two groups
of order pq, namely Z
pq
and a unique non-abelian group of order pq. If
p [ q 1, then Z
pq
is the only such group. We will not be proving these
facts since they are better left to a course that discusses more advanced
group theory.
4. It follows from these three examples that a symmetric group Sym(X) is
abelian if and only if [X[ 2. (If [X[ 2, then Sym(X)

= 1 or Z
2
. On
the other hand, if [X[ 3, then the non-abelian group S
3
embeds into
Sym(X), and thus Sym(X) is itself non-abelian.)
Theorem 61. Every element of S
n
can be written uniquely (up to order of
factors) as a (possibly empty) product of disjoint cycles.
Such a representation of a cycle is called its cyclic decomposition. Note that
the cyclic decomposition of (1) is (1) = (1), and the cyclic decomposition of a
non-identity permutation is a nonempty product of disjoint cycles.
Proof. We rst prove by induction on n that such a representation exists. The
base case n 2 is clear, so assume n 3. Let S
n
. If (n) = n,
then S
n1
, and the representation exists by induction. So let us as-
sume (n) ,= n. Consider the sequence
m
(n)

m=0
. Let k be the least
positive integer such that
k
(n) =
m
(n) for some m < k. (Such a k ex-
ists because the sequence is contained in the nite set 1, . . . , n. Also note
that k 2 since (n) ,= n, and that by the minimality of k the elements
n, (n),
2
(n), . . . ,
k1
(n) are distinct.) We have
km
(n) = n, and by the
minimality of k we have k m = k, and thus m = 0 and
k
(n) =
0
(n) = n.
So and the k-cycle = (n, (n),
2
(n), . . . ,
k1
(n)) agree on the elements
n, (n),
2
(n), . . . ,
k1
(n), and, because = (
1
), the permutation
1
must x those elements and hence be in S
n1
, and by induction it has a cyclic
decomposition, and any cycles in this decomposition must be disjoint from ,
in order for
1
to x the elements moved by . Therefore = (
1
) has a
cyclic decomposition.
For uniqueness, let =
1

m
be any cyclic decomposition of some S
n
.
It suces to show that, for any
j
and any b moved by
j
, the cycle
i
is the
(necessarily unique) cycle in S
n
satisfying
k
j
(b) =
k
j
(b) for each k 0. By the
disjointness of the
i
s, we see that and
j
agree on the elements they move.
We have
0
j
(b) = b =
0
(b), and for k > 0 the element
k1
j
(b) is xed by
j
(since
k
j
(b) =
k1
j
(b) leads to the contradiction
j
(b) = b), and by induction

k
j
(b) =
j
(
k1
(b)) =
k
(b), as desired.
66
Remark. Carefully examining the proof of the above theorem gives the fol-
lowing algorithm for determining the cyclic decomposition of any non-identity
permutation.
1. Let S
n
and C = .
2. If xes all elements that are xed by every cycle in C, then =

C
.
3. Otherwise, pick some element a 1, . . . , n that moves and every cycle
in C xes, and form a cycle (a, (a),
2
(a), . . . ,
k1
(a)), where k is the
smallest positive integer with
k
(a) = a. Add this cycle to C and return
to step (2).
We can determine a number of facts about a permutation from its cyclic
decomposition, as shown in the following proposition.
Proposition 62.
1. (a
1
, . . . , a
k
)
1
= (a
k
, . . . , a
1
).
2. Disjoint cycles commute.
3. For = (a
0
, . . . , a
k1
), m Z, and i 0, . . . , k 1 we have
m
(a
i
) =
a
i+
k
m
. Consequently, the order of a cycle is its length.
4. If
1
, . . . ,
m
are disjoint cycles, then o(
1

m
) = lcm(o(
1
), . . . , o(
m
)).
5. For k n, the number of k-cycles in S
n
is
n!
k(nk)!
.
6. For S
n
, we have (a
1
, . . . , a
k
)
1
= ((a
1
), . . . , (a
k
)).
7. The conjugates of S
n
consist of those permutations with the same
(disjoint) cycle structure, i.e., those whose cyclic decomposition has the
same number of cycles of each length as s cyclic decomposition does.
Proof.
1. This is fairly simple to check. (It would be good for your understanding
of permutations and cycle notation to do a quick mental check.)
2. Let and be disjoint cycles. If m is moved by , then m and (m) are
xed by and ( )(m) = (m) = ( )(m). If m is moved by , then
m and (m) are xed by and ( )(m) = (m) = ( )(m). Finally,
if m is xed by both and , then ( )(m) = m = ( )(m).
3. We have
0
(a
i
) = id(a
i
) = a
i
= a
i+
k
0
and
1
(a
i
) = (a
i
) = a
i+
k
1
by denition. If m > 1, then by induction
m
(a
i
) = (
m1
(a
i
)) =
(a
i+
k
(m1)
) = a
i+
k
(m1)+
k
1
= a
i+
k
m
. If m < 0, then
m
(a
i+
k
m
) =
a
i+
k
m+
k
(m)
= a
i
=
m
(
m
(a
i
)), and hence
m
(a
i
) = a
i+
k
m
by the
injectivity of
m
.
67
4. For each i, j we have
j
C(
i
), hence
j
C(
i
). It follows from
disjointness that all the permutations in
1

i1

i+1

m
x
all elements moved by any permutation in
i
, and hence the intersection
of these two subgroups is (1). Thus
1

m
=
1

m
,
and the result now follows by a past exercise.
5. By pre-calculus counting techniques, we know there are
n!
(nk)!
ways to
write an ordered list of k elements from 1, . . . , n. But each k-cycle can
be written in exactly k ways as such an ordered list, so there are exactly
n!
k(nk)!
k-cycles in S
n
.
6. This is another thing that is routine to check, and checking this on your
own would be of some benet.
7. The result is trivial for = (1), so let us assume ,= (1). Let =
1

k
be the cyclic decomposition of . The for each S
n
we have
1
=
(
1

1
)(
2

1
) (
k

1
), where each
i

1
is a cycle of the same
length as
i
, and it follows from the previous part and the fact that is
injective that the
i

1
s are disjoint. Therefore each conjugate of has
the same cycle structure as . Conversely, let

S
n
be any permutation
with the same cycle structure as . We can write the cyclic decomposition
of

as

k
, where each

i
is a cycle of the same length as
i
.
Let A
i
(resp., C
i
) the the elements moved by
i
(resp.,

i
), and let B
(resp., D) be the elements xed by (resp.,

). Then A
i

k
i=1
B
and C
i

k
i=1
D form partitions of 1, . . . , n, and there are bijections
f
i
: A
i
C
i
and g : C D, and it is easily checked that the extension
of these bijections to 1, . . . , n forms a permutation of 1, . . . , n. Thus

= (
1

1
) (
k

1
) =
1
is a conjugate of .
Example. We will determine the conjugacy classes of S
4
and their sizes. The
cycle structures for permutations in S
4
are: the identity, transpositions, 3-
cycles, 4-cycles, and products of two disjoint transpositions. The number of
permutations with these cycle structures are: 1,
43
2
= 6,
432
3
= 8,
4321
4
= 6,
and
1
2
(
43
2
) = 3, respectively. (To double-check our computations, we can add
1 + 6 + 8 + 6 + 3 = 24 = 4! = [S
4
[. The only non-trivial case where we could
not simply refer to our formula for the number of m-cycles was calculating the
number of products of two disjoint transpositions. For this, note that there
are
43
2
ways to choose the rst transposition, and then the second is already
chosen for you because there are only two possible entries left. Then we divide
by 2 because it does not matter which transposition we write rst, so we were
counting every element twice.) The conjugates of a cycle are just the ones with
the same cycle structure, so we now also know the conjugacy classes and their
sizes.
Theorem 63. For n 2, every permutation in S
n
can be written as a product
of transpositions of the form (k, k + 1).
68
Proof. Because (1) = (1, 2)
2
, and because every non-identity permutation in S
n
has a cyclic decomposition, it suces to prove the proposition for cycles. It is
straightforward to verify that (a
1
, . . . , a
m
) = (a
1
, a
2
)(a
2
, a
3
)(a
3
, a
4
) (a
m1
, a
m
),
so the result is now proven for cycles with adjacent entries diering by 1, and we
have reduced to proving the proposition for transpositions. Finally, for m < k
we have (m, k) = (m, m + 1, . . . , k 2, k 1)(k, k 1, . . . , m + 1, m), which as
previously noted can be written in the desired form.
Theorem 64. If
1

m
=
1

k
are products of transpositions, then m
and k have the same parity, i.e., they are either both even or both odd.
Proof. Let n 2. Let T be the set of two-element subsets of 1, . . . , n, and
for each S
n
and A T, let s
A
() = 1 if is order-preserving on A and 1
if it is order-reversing on A. (In other words, if A = a, b with a < b, then
s
A
() = 1 if (a) < (b), and s
A
() = 1 if (a) > (b).) We note that, for
each A T and , S
n
, we have s
A
() = s
(A)
()s
A
(). Indeed, this follows
from observing that is order-preserving on A if and only if (i) is order-
preserving on A and is order-preserving on (A), or (ii) is order-reversing
on A and is order-reversing on (A).
Dene sgn : S
n
1, 1 :

AT
s
A
(). Then for each , T, we
have sgn() =

AT
s
A
() =
_
AT
s
(A)
()
_ _
AT
s
A
()
_
= sgn()sgn(),
so sgn is a homomorphism. (The last equality follows from observing that the
map A (A) is a permutation of T, so s
(A)
()
AT
= s
A
()
AT
.)
For each k < n, the only two-element subset of 1, . . . , n that (k, k + 1)
is order-reversing on is k, k + 1, so sgn((k, k + 1)) = 1. Examining the
proof of the previous theorem, we see that every transposition can be written
as a product of an odd number of transpositions of that form, so sgn takes all
transpositions to 1, hence takes products of an odd number of transpositions
to 1 and products of an even number of transpositions to 1. Therefore there
is no permutation that can be written as both a product of an odd number of
transpositions and a product of an even number of transpositions.
Denition. A permutation in S
n
is called even (resp., odd) if it can be writ-
ten as a product of an even (resp., odd) number of transpositions. This is
referred to as the permutations parity. Note that (1) is even, including in
the case where n = 1 and there are no transpositions, because it is still a
product of an even number (namely, zero) of transpositions. The alternating
group on 1, . . . , n is the normal subgroup A
n
of S
n
that consists of the even
permutations. (It is a normal subgroup because it is the kernel of the homo-
morphism sgn : S
n
1, 1 used in the proof above. Again, for emphasis,
sgn takes odd permutations to 1 and even permutations to 1. Also, we have
sgn(
1
) = sgn()
1
= sgn(), so a permutation and its inverse have the same
parity.)
Proposition 65. A cycle of odd length is an even permutation, and a cycle
of even length is an odd permutation. (This allows one to easily nd the par-
ity of any permutation by nding its cyclic decomposition =
1

k
and
evaluating sgn() = sgn(
1
) sgn(
k
).)
69
Proof. Writing (a
1
, . . . , a
m
) = (a
1
, a
2
)(a
2
, a
3
) (a
m1
, a
m
), we see that an m-
cycle is a product of m1 transpositions.
Proposition 66. Let H be a subgroup of S
n
. Then either H A
n
or half
of the permutations in H are even and half are odd. Thus for n 2 we have
[A
n
[ =
n!
2
.
Proof. Assume H contains an odd permutation . Dene f : HA
n
HA
n
:
and g : H A
n
H A
n
:
1
. (Note that right-multiplication
by an odd permutation changes the parity.) It is easy to see that f and g are
inverse maps, so [H A
n
[ = [H A
n
[, or in other words H consists of half even
permutations and half odd.
Example. We will determine the conjugacy classes of A
4
and their sizes. The
cycle structures of permutations in A
4
are: the identity, 3-cycles, and products
of two disjoint transpositions. As before, there are 1, 8, and 3 permutations of
these types, respectively. (We can check: 1+8+3 = 12 =
4!
2
= [A
4
[.) So we now
know the conjugacy class of each even permutation in S
4
, but not necessarily
in A
4
. Because the conjugacy class of an element in A
4
is either the same size
as in S
4
or half of that (since [S
4
: A
4
] = 2 see past exercises), the conjugacy
classes in A
4
are either these same three sets, or the set of 3-cycles splits into
two conjugacy classes of size 4. The following lemma shows that the former is
the case.
Lemma 67.
1. For n 3, the group A
n
is generated by the set of 3-cycles in S
n
.
2. For n 5, all 3-cycles are conjugates in A
n
. Hence the only normal
subgroup of A
n
containing a 3-cycle is A
n
itself.
Proof.
1. Let n 3. It suces to show that any product (a, b)(c, d) of two (not
necessarily disjoint) transpositions is a product of 3-cycles. If (a, b) =
(c, d), then (a, b)(c, d) = (1) = (1, 2, 3)
3
, so let us assume (a, b) ,= (c, d).
Then without loss of generality we have b / c, d and c / a, b, and we
have (a, b)(c, d) = (a, b)(b, c)(b, c)(c, d) = (a, b, c)(b, c, d).
2. Let n 5, ,

be 3-cycles, and be a transposition disjoint from .


(Here we use the fact that n 5.) Because

is a conjugate of in
S
n
(they have the same cycle structure), we have

=
1
for some
S
n
. If A
n
, we are done. Otherwise, we have A
n
, and
()()
1
= (
1
)
1
=
1
, as desired. (The last equality
follows since and commute.)
Theorem 68. A
n
is simple if and only if n ,= 4.
70
Proof. The cases n = 1, 2, 3 are trivial, because A
1
= A
2
= (1) and [A
3
[ =
3!
2
= 3, which is prime. You will do the cases 4 and 5 in the exercises. So now let
us assume n 6. Suppose that A
n
has a nontrivial proper subgroup N. Then
N A
n1
A
n1
(past exercise). By induction, the group A
n1
is simple, so
N A
n1
= (1) or A
n1
. Because N is a proper normal subgroup of A
n
, it
cannot contain a 3-cycle, so A
n1
_ N, and thus N A
n1
= (1). Now pick
some non-identity N. Observe that s cyclic decomposition must either
contain a cycle of length at least 3 or contain at least two transpositions. In the
former case, the permutation has at least (n1)(n2) conjugates in S
n
. (For
a xed a 1, . . . , n, there are at least (n 1)(n 2) m-cycles that move a,
and must have at least one dierent conjugate for each of these m-cycles.) In
the latter case, it has at least (n 2)(n 3) conjugates in S
n
. (For xed a, b
1, . . . , n, there are (n 2)(n 3) ways to pick two transpositions of the form
(a, x) and (b, y), and must have at least one dierent conjugate for each such
choice.) So either way has at least (n2)(n3) 3(n2) = 2n+n6 2n
conjugates in S
n
, hence at least
2n
2
= n conjugates in A
n
. Since N contains
all the conjugates of in A
n
, and also contains (1), which is not a conjugate
of , we have [N[ > n. Thus [NA
n1
[ =
|N||An1|
|NAn1|
>
n(n1)!/2
1
=
n!
2
= [A
n
[, a
contradiction.
Corollary 69. For n ,= 4, the only normal subgroups of S
n
are (1), A
n
, and
S
n
.
Proof. Let n ,= 4. To avoid dealing with special cases later, we note that the
result is trivially true for n = 1 or n = 2, so let us assume n 3. We have
already noted that (1), A
n
, S
n
S
n
. Now let N be any normal subgroup of
S
n
. Then N A
n
A
n
, so by the simplicity of A
n
we have N A
n
= (1)
or A
n
. Recall that N A
n
is either equal to N or exactly half the size of
N. So N is either (1), A
n
, S
n
, or has order 2. But it is easily checked that
every conjugacy class of a non-identity element of S
n
consists of more than one
element, so the last case cannot happen.
As a bit of trivia, it turns out A
5
is the smallest (in terms of cardinality)
simple group of non-prime order. One of the largest and most impressive clas-
sications in abstract algebra is the work done on classifying all nite simple
groups. It has been shown that every nite simple group is (isomorphic to) one
of the following:
1. 1,
2. Z
p
for p prime,
3. A
n
for n 5,
4. a group from one of sixteen other countably innite families that have
been precisely described but which we will not discuss,
5. or one of 26 nite simple groups called the sporadic groups that t no
particular pattern.
71
The biggest of the sporadic groups is called the monster group or the friendly
giant, which has order approximately 8 10
53
. The monster is unbelievably
complex and contains at least 60 dierent (i.e., non-isomorphic) non-abelian
groups as subgroups, including A
12
and all but six of the sporadic groups. The
main point in bringing this up is to give you an idea of how intricate the study
of group theory can get. However, we are more or less done with our study of
group theory at this point, and you are now essentially equipped with all the
basic knowledge of groups that you will need in order to go on and study the
other algebraic structures.
Exercises.
1. (a) Convert this permutation from 2-row form to its cyclic decomposi-
tion.
_
1 2 3 4 5
3 5 4 1 2
_
(b) Write the cyclic decomposition of (1, 2)(1, 4, 3)(2, 3, 5).
(c) Write the permutation (1, 2, 3)(4, 5) S
5
in 2-column form.
2. Show that every group of order pq, where p q are primes, has a normal
subgroup of order q. (Hint: Use Cauchys Theorem and a theorem from
this section.)
3. Let K
4
= (1), (1, 2)(3, 4), (1, 3)(2, 4), (1, 4)(2, 3) A
4
. (This is called
the Klein-4 group, which also happens to be the name of a cool music
group.)
(a) (This exercise shows that A
4
is not simple.) Show that K
4
A
4
.
(Hint: Since K
4
is so small, in order to show K
4
A
4
, it may be
simplest just to explicitly use the Subgroup Test. Alternatively, it is
somewhat less tedious but requires a bit more knowledge to do the
following. Verify that (1, 2)(3, 4)(1, 3)(2, 4) is a subgroup of A
4
by
showing that the generators commute, and calculate that it contains
K
4
and that its order is 4 by the order formula, so this subgroup is
A
4
. To show normality, note that K
4
is a union of conjugacy classes
in S
4
, hence normal in S
4
, hence normal in A
4
.)
(b) (This exercise shows that normality is not transitive.) Show that
(1, 2)(3, 4) K
4
, but that (1, 2)(3, 4) _ A
4
. (Hint: For the rst
statement, calculate the index. For the second statement, gure out
the conjugacy class of (1, 2)(3, 4) in S
4
, then use an exercise from the
previous section to deduce its conjugacy class in A
4
.)
(c) Show that K
4

= Z
2
Z
2
. (Hint: If you used the alternate hint
method in (a), it follows nearly immediately from that proof that
K
4
= (1, 2)(3, 4) (1, 3)(2, 4)

= Z
2
Z
2
. If you did not do this
method, then another way to prove this would be to cite the past
result that shows that every group of order 4 is abelian, show that
72
K
4
Z
4
by noting the orders of its elements, and appeal to the
Fundamental Theorem of Finite Abelian Groups.)
4. (a) List all the cycle structures for permutations in S
5
. Indicate which
of them correspond to even permutations.
(b) Determine how many permutations there are with each of these cycle
structures. (You are allowed to use any pre-calculus counting tech-
niques without proof. Hint: Look at the example in the notes where
this is done for A
4
.)
(c) Show that A
5
is simple. (Hint: Recall from a past exercise that a
subgroup is normal if and only if the conjugacy classes of its elements
form a partition of it. You determined the sizes of the conjugacy
classes of S
5
in part (b). Use this and a result of this section to narrow
down the possible sizes of conjugacy classes of A
5
. Use Lagranges
Theorem to narrow down the possible sizes of subgroups of A
5
. Using
this knowledge of sizes, show that no nontrivial proper subgroup of
A
5
can be a disjoint union of conjugacy classes of A
5
.)
Chapter 3
Ring Theory
In this section, we will study ring theory. One might think of a group as the
basic algebraic structure with one operation, and a ring as the basic algebraic
structure with two operations. We will learn the basic denitions of concepts
such as rings, ideals, and ring homomorphisms. We will discuss these things
in full generality, but our bias will be (surprise) in favor of commutative rings,
and particularly integral domains. This is not only my preference, but is typical
of basic algebra courses, because the familiar number systems are all integral
domains, and the abstract notion of an integral domain captures most of their
algebraic properties. Around the end of this chapter we will nish our explicit
precise construction of the rest of the number systems Z, , 1, and C. We will
save factorization properties (such as every integer being a product of primes)
and related topics for the next chapter.
3.1 Rings, Integral Domains, and Fields
Note: This section is at least two lectures worth, so your homework assignment
will probably be this plus one more section.
[Durbin: Sections 24-27, 38, 39]
Denition.
1. A ring is a set R together with two operations + and on R such that
(R, +) is an abelian group, (R, ) is a monoid, and the distributive laws
a(b +c) = ab +ac and (b +c)a = ba +ca are satised. (Formally, the ring
is a triple (R, +, ), but as we did with groups, we will typically just refer
to R as the ring if what we mean is clear from context.)
2. Let R be a ring. The zero (or additive identity) of R is the identity 0
of (R, +), and its multiplicative identity (or simply identity for short) is
the identity 1 of (R, ). In an exercise you will show that the denitions
imply that 0 is the absorbing element of (R, ). (This is fortunate, because
otherwise we would have a conict of notation.)
73
74
3. Recall from a past exercise that the absorbing element of a monoid is a
unit if and only if the monoid is trivial. Thus a ring is the zero ring (or
trivial ring) 0 1 = 0 0 has a multiplicative inverse. There is little
to nothing interesting about the zero ring, and a lot of theorems are true
for every ring but the zero ring, so from now on by ring we will always
mean a nonzero ring unless explicitly stated otherwise.
4. Let R be a ring. We say R is commutative if (R, ) is. (Remember: addition
is always commutative in a ring.) The center of R is the center C(R) of
(R, ).
5. We abbreviate R

= R 0, and we use R

to denote the group of


units of (R, ). Thus units of R are elements with a multiplicative inverse.
(Remember: Every element has an additive inverse.) We always have
1 R

. We say R is a eld if it is commutative and R

= R

.
6. A ring D is an integral domain (or simply a domain) if it is commutative
and (D, ) is cancellative. (Recall that this means that for every a, b, c D
with a ,= 0, if ab = ac, then b = c.) Thus every eld is an integral domain.
(But not every domain is a eld; see the examples below.) An element a
of a commutative ring R is called a zero divisor if there is a b R

with
ab = 0. A commutative ring is an integral domain if and only if it has no
nonzero zero divisors (exercise).
7. A subring of a ring is a subset that is a ring with respect to the same
operation and which contains 1. Thus a subset A of a ring R is a subring
if and only if (A, ) is a submonoid of (R, ) and (A, +) is a subgroup of
(R, +). We do not have a subring symbol analogous to the symbol
we used for subgroups.
Remark. I will briey note the dierences between our denitions/notation
and those of Durbin. He denotes the identity element of rings by e. This is a
pretty common notation for groups, but is pretty unusual for rings, at least in
actual mathematical practice. Also, in his denition of ring, he only requires
(R, ) to be a semigroup, and he refers to a ring (in our sense) as a unital ring
with e ,= 0. His denition of not requiring a ring to have an identity element
is the more traditional, old-fashioned denition. It is what would be called
a rng (pronounced rung) by most modern ring theorists. (The name rng
originally started as a joke it is a ring without the identity requirement, hence
the i is missing.) At some point mathematicians realized that the version of
the denition with an identity included was the more useful one to take as the
basic denition.
Example.
1. With the usual operations, , 1, and C are elds.
2. Any subring of a eld is a domain. For example, with the usual operations,
Z and Z[i] are integral domains. They are not elds, since Z

= 1
and Z[i]

= 1, i.
75
3. N (resp.,

) is not a ring, due to the lack of additive inverses (resp., of a


zero element).
4. If R is any ring and n Z
+
, then /
n
(R) is a ring, with matrix addition
and multiplication dened in the usual way. It is commutative if and only
if R is commutative and n = 1. The center of /
n
(R) is the set of diagonal
matrices.
5. Let R be any commutative ring. The polynomial ring R[X] consists of
all polynomials with coecients in R. Two polynomials are considered
equal if and only if they are equal as formal expressions, i.e., correspond-
ing terms have the same coecients. One adds and multiplies polynomi-
als according to the usual rules. (Important note: Polynomials here are
treated as formal expressions, not as functions. For example, the poly-
nomials X, X
2
Z
2
[X] are distinct even though as functions they agree
on Z
2
.) The ring R[X] is a domain if and only if R is, and in this case
R[X]

= R

. (Modern commutative ring theorists tend to write capital


letters for indeterminates as I have, to make it more clear that they are
indeterminates, but it is ne if you like lowercase letters better.) We will
talk a lot more about polynomials later.
6. Similarly, if R is any commutative ring, we can dene the power series
ring R[[X]] to consist of all formal power series with coecients in R.
The operations are formal addition and multiplication of power series in
the usual way. (Again, these are formal expressions, not functions. We
do not worry about convergence, even if R is a ring like 1 where such a
notion would make sense.) The ring R[[X]] is a domain if and only if R is,
and in this case R[[X]]

is (perhaps surprisingly) the set of power series


with unit constant term. We will also discuss these later, though not in
nearly so much detail as polynomial rings.
7. Let R be a ring and X be a nonempty set. Then M(X, 1) is a ring,
with addition and multiplication given by (f + g)(x) = f(x) + g(x) and
(fg)(x) = f(x)g(x), respectively.
Proposition 70. Let R be a ring and x, a
1
, . . . , a
n
R.
1. (a
1
+ +a
n
)x = a
1
x + +a
n
x.
2. x(a
1
+ +a
n
) = xa
1
+ +xa
n
.
3. x 0 = 0 x = 0.
4. x 1 = 1 x = x.
(Note that by (4) the analogous versions of (1) and (2) hold with one or more
+s replaced by s.)
Proof. 1. The case n = 1 is trivial, so let us assume n 2. By induction and
the distributive law we have (a
1
+ +a
n
)x = (a
1
+ +a
n1
)x+a
n
x =
a
1
x + +a
n1
x +a
n
x.
76
2. Similar to (1).
3. Exercise.
4. x + x 1 = x 1 + x 1 = x(1 + 1) = x 0 = 0, so x 1 = x. The
proof that 1 x = x is similar.
Denition. Let R be a ring.
1. An ideal of R is a subgroup I of (R, +) such that ra, ar I for every a I
and r R; we say I is a proper ideal if I ,= R. An ideal is proper it
does not contain 1 is does not contain a unit (exercise). Note that R
must have at least two ideals, namely 0 and R. In a eld, those are the
only ideals. Note also that every ideal is a rng, and the only ideal that is
a ring is R itself.
2. Any intersection of a nonempty collection of ideals is an ideal (exercise),
and it follows that for each A R there is a smallest (with respect to
inclusion) ideal containing A, called the ideal generated by A, and denoted
(A). We abbreviate (a
1
, . . . , a
n
) = (a
1
, . . . , a
n
) and (a
1
, a
2
, . . .) =
(a
1
, a
2
, . . .). An ideal generated by a single element is called principal. In
an exercise you will show that, if R is commutative, then (a
1
, . . . , a
n
) =
Ra
1
+ +Ra
n
. (This is shorthand for the set

n
i=1
r
i
a
i
[ each r
i
R.)
3. The set of ideals of R is partially ordered by inclusion, with minimum
element 0 and maximum element R. A maximal ideal of R is a maximal
element among the proper ideals. In an exercise you will show that every
proper ideal of R is contained in a maximal ideal.
Example.
1. The ideals of Z are the subsets of the form nZ for n Z, i.e., each ideal
of Z is principal. (These are easily veried to be ideals, and we know
this is all of them because these are all the subgroups of (Z, +), as every
subgroup of a cyclic group is cyclic. Technically, our proof of all this relies
on the Division Algorithm holding in Z, a fact that we will prove at some
later point.) Such a domain in which each ideal is principal is called a
principal ideal domain (PID).
2. Another (rather trivial) example of a PID is a eld, since the only ideals
of a eld are (0) and (1).
3. Another example of a PID is K[X], where K is any eld. (We will prove
this later.)
4. An example of a domain that is not a PID is Z[X]. The ideal (2, X) is
not principal. To see this, note that any generator of (2, X) would have
to be a common divisor of 2 and X, hence would have to be 1, but every
polynomial in (2, X) = XZ[X] + 2Z has even constant term.
77
Theorem 71 (Quotient Rings). Let I be a proper ideal of a ring R. Then
R/I forms a ring under the operations (I + a) + (I + b) = I + (a + b) and
(I +a)(I +b) = I +ab. The zero is I, the identity is I +1, and additive inverses
are given by (I +a) = I a. The ring R/I is commutative if R is. This ring
is called a quotient ring.
Proof. All the statements will easily follow once we show that the stated oper-
ations are well-dened. We already know that this is the case for the addition
operation from group theory. So assume I +a = I +a

and I +b = I +b

. Then
aa

I and bb

I, so aba

= abab

+ab

= a(bb

)+(aa

)b

I,
and hence I +ab = I +a

, as desired.
Example. Let n 2. The above construction makes Z
n
= Z/nZ into a ring.
We have [a]
n
= [b]
n
if and only if a
n
b, so addition/multiplication in Z
n
correspond to adding/multiplying integers modulo n. Note that [a]
n
is a zero
divisor there is a b 1, . . . , n 1 with n [ ab lcm(a, n) < an
gcd(a, n) > 1. (For the last step we used the fact that an = gcd(a, n)lcm(a, n),
which we will prove later.) An exercise shows that every element of a nite ring
is either a unit or a zero divisor, so the units of Z
n
are the elements [a]
n
with
gcd(a, n) = 1. Hence Z
n
is a eld if and only if it is an integral domain if and
only if n is prime.
Denition. A map : R S between rings is called a ring homomorphism (or
simply homomorphism if it is clear from context which kind of homomorphism
we mean) if it is both a group homomorphism from (R, +) to (S, +) and a
monoid homomorphism from (R, ) to (S, ). More explicitly, what this means is
that (1) = 1 and for every x, y R we have (xy) = (x)(y) and (x +y) =
(x)+(y). (Recall that since (R, +) is a group, we automatically have (0) = 0
if these other properties are satised.) The various types of ring homomorphisms
(ring isomorphisms, ring monomorphisms, etc.) are dened analogously to how
they were for groups. The kernel of a ring homomorphism is
1
[(0)], which
is an ideal by the proposition below.
Proposition 72.
1. A composition of ring homomorphisms (monomorphisms, epimorphisms,
isomorphisms) is a ring homomorphism (resp., monomorphism, epimor-
phism, isomorphism).
2. The inverse of a ring isomorphism is a ring isomorphism.
3. If R is a monoid, then End(R) is a submonoid of (M(R), ) and Aut(R) =
End(R)

. (In particular, Aut(R) is a group.)


4. Homomorphic images and pre-images of rings are rings.
5. Homomorphic pre-images of ideals are ideals.
78
Proof. Statements (1)-(4) are immediate consequences of the corresponding
properties of group/monoid homomorphisms. Now let : R T be any ring
homomorphism and J be any ideal of T. We know that
1
[J] is a normal
subgroup of R from group theory. For any r, a R with (a) J, we have
(ra) = (r)(a) J and (ar) = (a)(r) J, so
1
[J] is an ideal of R.
Theorem 73 (Fundamental Homomorphism Theorem for Rings). Let : R
T be a ring homomorphism. Then R/ ker

= [R].
Proof. Dene f : R/ ker [R] : ker + x (x). We know from group
theory that this is a well-dened group isomorphism, so all that remains is to
show that it is a ring homomorphism. Firstly, note that f(ker +1) = (1) = 1.
Secondly, for each a, b R we have f((ker + a)(ker + b)) = f(ker + ab) =
(ab) = (a)(b) = f(ker +a)f(ker +b). Thus f is a ring isomorphism.
Theorem 74 (Correspondence Theorem for Rings). Let R be a ring and I
be a proper ideal of R. The map T T/I is an order-preserving one-to-
one correspondence between the subrings (resp., ideals) of R containing I and
the subrings (resp., ideals) of R/I. The inverse map is T
1
[T], where
: R R/I : r I +r.
Proof. From group theory we already know that these are order-preserving bi-
jections between the subgroups of (R, +) containing I and the subgroups of
(R/I, +), and that the maps are inverses. Thus it will suce to show that both
maps take subrings to subrings and ideals to ideals. The subring case is shown
exactly analogously to the subgroup case of the Correspondence Theorem from
group theory, as is the fact that the latter map takes ideals to ideals. Now let
J be any ideal of R containing I. We already know that J/I is a subring of
T/I, and for a J and t T we have (I + a)(I + t) = I + at J/I and
(I +t)(I +a) = I +ta J/I, so J/I is an ideal of T/I, as desired.
Proposition 75. Let R be a ring. Then : Z R : n n 1 is a ring
homomorphism. (Recall the denition of n 1 as an element of (R, +) this is
exponentiation in additive notation.) Therefore Z or some Z
n
(n 2) embeds
into R.
Proof. We have (1) = 1 1 = 1, and for m, n Z, we have (m + n) =
(m+n) 1 = m 1 +n 1 = (m) +(n) and (mn) = (mn) 1 = m (n 1) =
(m 1)(n 1) = (m)(n). (The equality (m+n) 1 = m 1 +n 1 comes from
the exponentiation law in additive notation, as the distributive properties of R
are not applicable here. The equality (mn) 1 = m (n 1) also comes from
an exponentiation law in additive notation, as the associative property of (R, )
is again not applicable here. Finally, the equality m (n 1) = (m 1)(n 1)
follows from the distributive laws.) Therefore is a homomorphism. The last
statement follows from the Fundamental Homomorphism Theorem and the fact
that the kernel of , being an ideal of Z, must be 0 or some nZ with n 2.
79
Denition. Let R be a ring. The characteristic of R is the least n Z
+
with
n 1 = 0, if one exists, and is otherwise 0. (Compare and contrast this with
the order of 1 in (R, +).) If R has characteristic n, then it follows from the
distributive property that n x = 0 for all x R. The proof of the above
proposition shows that Z embeds into R if R has characteristic 0, and that Z
n
embeds into R if R has characteristic n Z
+
. A domain has either zero or
prime characteristic (exercise).
Proposition 76 (Freshmans Dream). Let R be a commutative ring of prime
characteristic p and x, y R. Then (x +y)
p
= x
p
+y
p
.
Proof. The familiar Binomial Theorem for expanding (x + y)
p
is proven using
only properties that R possesses, so we have (x + y)
p
=

p
k=0
_
p
k
_
x
nk
y
k
. We
have
_
p
k
_
=
p!
k!(pk)!
, so k!(p k)! [ p!. So, if 0 < k < p, all the prime factors of
k!(pk)! are less than p, which leads us to the conclusion that k!(pk)! [ (p1)!
and p [
_
p
k
_
, and thus
_
p
k
_
x
nk
y
k
= 0. Hence (x +y)
p
= x
p
+y
p
.
Example. It is essential for this theorem that the characteristic be prime. For
example, in Z
4
we have ([1] + [1])
4
= [16] = [0] ,= [2] = [1]
4
+ [1]
4
.
Exercises.
1. Let R

be an indexed family of rings. Verify that

is a ring
under the operations (fg)() = f()g() and (f + g)() = f() + g().
(This ring is called the direct product of the R

s. Hint: You may use


our previous knowledge of direct products of groups/monoids to conclude
that

is an abelian group under + and a monoid under . [Review


this if you do not remember this.] So all that is left is to show that the
distributive laws hold.)
2. Let R be a ring. Show that 0 x = x 0 = 0 for all x R. (This is
actually true in rngs too. Hint: Use the distributive law to show that
0 x+0 x = 0 x, and then cancel to get 0 x = 0. The other one is proven
similarly.)
3. Prove that a commutative ring is an integral domain if and only if it has
no nonzero zero divisors. (Hint: Note that ab = ac a(b c) = 0.)
4. Let R be a ring.
(a) (Ideal Test) Let _ I R. Prove that I is an ideal of R if and only
if a b, ra, ar I for all a, b I and r R. (The way that I have
this stated applies to rngs as well, where ideals of rngs are dened
analogously to ideals of rings. When dealing with rings, you may
equivalently replace the a b with a + b in the Ideal Test, due
to the identities a b = a + (1)b and a + b = a (1)b. In the
future, you may use either of these two versions. Hint: This follows
more or less immediately from the Subgroup Test and the denition
of an ideal.)
80
(b) Show that an ideal of R is proper it does not contain 1 it does
not contain any units.
(c) [Extra credit] Prove that every intersection of a nonempty collection
of ideals of R is an ideal of R. (Hint: Use the Ideal Test.)
(d) [Extra credit] Prove that every union of a nonempty chain of ideals
of R is an ideal of R. (Hint: Use the Ideal Test. Alternatively, recall
that such a union is a subgroup by a past exercise, and then verify
the ra, ar I requirement.)
(e) Prove that every proper ideal of R is contained in a maximal ideal.
(Hint: Use (d) and Zorns Lemma. Note that a union of proper ideals
is necessarily proper by (b).)
(f) [Extra credit] Show that (A) =

n
i=1
r
i
a
i
s
i
[ n N, each r
i
, s
i

R, and a
i
A. (Note that if R is commutative this simplies to
(A) =

n
i=1
r
i
a
i
[ n N, each r
i
R, and a
i
A. Hint: Let I
be the latter set. From the denitions we have A I (A), so it
will suce to show that I is an ideal. Use the Ideal Test. When you
show that I ,= , note that 0 I, even if A = , because I is dened
to contain the empty sum, which is 0.)
(g) Assume R is commutative and let a
1
, . . . , a
n
R. Show that (a
1
, . . . , a
n
) =
Ra
1
+ +Ra
n
. (Hint: Use the previous part.)
(h) [Extra credit] Let I and J be ideals of R. Show that (I J) =
I +J. (Hint: This follows more or less directly from part (f), but the
notation is slightly cumbersome. Alternatively, from the denitions
we have I J I +J (I J), so it will suce to show that I +J
is an ideal. Use the Ideal Test.)
(i) [Extra credit] Show that every eld homomorphism is a monomor-
phism. (Hint: Deduce from the part (b) that its kernel is (0), and
then injectivity follows from group theory.)
5. [Extra credit] Let R be a ring.
(a) (Subring Test) Let T R and 1 T. Prove that T is a subring of R
if and only if ab, ab T for every a, b T. (Note that if we change
the requirement 1 T to merely T ,= , we get a Subrng Test.
Hint: If you cite the Subgroup and Submonoid Tests appropriately,
this is more or less immediate.)
(b) Show that C(R) is a subring of R. (Recall C(R) is the center of R.
Hint: Use the Subring Test.)
(c) Show that any intersection of a nonempty collection of subrings of a
ring is a subring. (Hint: Use the Subring Test.)
(d) Let A R. Show that there is a smallest (with respect to inclusion)
subring of R containing A. (This is called the subring generated by
A. Hint: Use part (c).)
81
6. [Extra Credit] Prove that every element of a nite ring is either a unit
or a zero divisor. (It follows that every nite integral domain is a eld.
Hint: Let R be a nite ring and let a R be any nonunit. Then a cannot
have both a left and a right inverse [past exercise], so one of the maps
f : R R : x ax or g : R R : x xa is not surjective, hence by
niteness not injective. Show why this implies that a is a zero divisor.)
7. [Extra Credit] Prove that every domain has either zero or prime character-
istic. (Hint: If D is a domain of positive characteristic p, then Z
p
embeds
in D, hence Z
p
is a eld.)
3.2 More Ideals
[Mostly not in Durbin.]
From now on, we will be focusing our eorts entirely on commutative ring
theory. Certain denitions and theorems that we will state for commutative
rings also apply to non-commutative rings, but keeping track of which ones do
and which do not is too cumbersome to be worthwhile to us. For brevity, we will
thus adopt the following convention: from now on, whenever I say ring,
I mean commutative ring.
Denition. Let R be a ring.
1. For a, b R, we say that a is a divisor of b, or that a divides b, written
a [ b, if there is an x R with ax = b. We have a [ b b (a) (a) (b)
(exercise). (Carefully note the direction of inclusion here.) If (a) _ (b),
then we call a a proper divisor of b. We say a, b R are associates if
(a) = (b), or, equivalently, a [ b and b [ a. Thus every divisor of an
element is either an associate or a proper divisor. It is also immediate
that the associate relation is an equivalence relation. If R is an integral
domain, then (a) = (b) if and only if there is a u R

with b = au
(exercise).
2. Let I and J be ideals of R. We dene IJ =

n
i=1
a
i
b
i
[ n Z
+
, a
i

I, b
i
J. (We dene it this way instead of in the obvious way because we
want to make sure that IJ is an ideal. We will verify this in a moment.)
We will see that this multiplication makes the set J(R) of ideals of R into
a commutative monoid with identity R, and that I
1
I
n
= (a
1
a
n
[
a
i
I
i
) I
1
I
n
, and that (a
1
) (a
n
) = (a
1
a
n
).
Example. In this example we explore the dierence between products and
intersections of ideals.
1. In any PID (such as Z), we have (a
1
) (a
n
) = (lcm(a
1
, . . . , a
n
)).
(Future exercise.) A least common multiple (LCM) of a
1
, . . . , a
n
is dened
to be a common multiple that divides all other common multiples. There
is typically more than one LCM: the LCMs consist of the associates of
82
a given LCM. We will discuss this more later when we get to divisibility
properties.
2. Let D be a domain and R be the subring of D[X] consisting of the poly-
nomials with no degree 1 term. (Verifying that this is indeed a subring
is a good example of the Subring Test. The polynomials in R are pre-
cisely those that can be written in the form X
2
f + a for some f D[X]
and a D. Then 1 R, and for f, g D[X] and a, b D we have
(X
2
f +a)+(X
2
g+b) = X
2
(f +g)+(a+b) R and (X
2
f +a)(X
2
g+b) =
X
2
(X
2
fg + bf + ag) + ab R, so R is indeed a subring of D[X] by the
subring test.) One can show that X
5
= (X
2
)(X
3
) is an LCM of X
2
and
X
3
, but (X
5
) _ (X
2
) (X
3
), since X
6
(X
2
) (X
3
) but X
6
/ (X
5
).
The purpose of the next couple results is to make computations with the
addition or multiplication or ideals a bit easier to handle.
Lemma 77. Let R be a ring.
1. If I is an ideal of R, then (I) = I.
2. For A R and x R, we have x(A) = (xA). In particular, if I is an
ideal of R, then so is xI.
3. For a family A

of subsets of R, we have (

) = (

(A

)).
4. Ideal multiplication is associative, and for A
1
, . . . , A
n
R, we have (A
1
) (A
n
) =
(a
1
a
n
[ a
i
A
i
).
Proof.
1. Clearly I is the smallest ideal containing I.
2. By a past exercise,
x(A) = x
n

i=1
r
i
a
i
[ r
i
R, a
i
A =
n

i=1
r
i
(xa
i
) [ r
i
R, a
i
A = (xA).
3. The inclusion is clear, and follows once we observe that

(A

)
(

).
4. The case n = 1 is trivial, so we consider the case n = 2. Using parts (2) and
(3), we obtain (a
1
a
2
[ a
i
A
i
) = (

a1A1
a
1
A
2
) = (

a1A1
(a
1
A
2
)) =
(

a1A1
a
1
(A
2
)) = (

a2(A2)
A
1
a
2
) = = (

a2(A2)
(A
1
)a
2
) = (a
1
a
2
[
a
i
(A
i
)). But by a past exercise we have (a
1
a
2
[ a
i
(A
i
)) =

n
i=1
r
i
(a
i
a

i
) [ r
i
R, a
i
(A
1
), a

i
(A
2
) = (A
1
)(A
2
). (For the last
equality, note that we have since each r
i
a
i
(A
1
), and we have by
setting each r
i
= 1.) With this case established, the associativity of ideal
multiplication now follows from the associative of multiplication in R, and
the case n 3 easily follows by induction.
83
Theorem 78. The set J(R) of ideals of a ring R forms a commutative semiring
under ideal multiplication and addition as we have dened them. (A semiring
is a ring with the requirement that it is an abelian group under + weakened to
merely being a commutative monoid under +. In other words, a semiring is like
a ring except you are not allowed to subtract.) The zero is (0) and the identity
is R.
Proof. We know that the sum of two ideals is an ideal by a past exercise, and
it is then obvious that (J(R), +) is a commutative monoid with identity (0).
From the lemma we see that (J(R), ) is a monoid with identity R, and it is
obvious that multiplication is commutative. So all that remains is to check that
the distributive law holds. For I, J, K J(R), the lemma yields I(J + K) =
(a(b +c) [ a I, b J, c K) = (ab +ac [ a I, b J, c K) IJ +IK,
but also IJ, IK I(J +K), so IJ +IK = (IJ IK) I(J +K), as desired.
Denition.
1. A prime ideal of R is a proper ideal P such that whenever IJ P for
ideals I and J of R, we have I P or J P. The set of prime ideals
of R is called the prime spectrum (or simply spectrum) of R, and denoted
Spec(R).
2. We say p R is prime if (p) is. Note that this means that any associate of
a prime element is prime, and that prime elements are nonunits. We will
momentarily see that p R R

is prime whenever p [ ab, then p [ a


or p [ b. (Note that a principal ideal is proper if and only if its generator
is a nonunit.)
3. We say a R R

is an atom (or irreducible) if whenever a = bc we have


(a) = (b) or (a) = (c). In an integral domain, this is equivalent to saying
that a cannot be written as a product of two nonzero nonunits. (Exercise.)
Every prime is irreducible (exercise), and, when we discuss factorization,
we will see that in Z (or more generally any PID) the notions of prime
and irreducible are equivalent. (In fact, there is a good chance that the
denition of prime you learned in basic math classes is the denition
algebraists use for irreducible.)
Remark. Any associate of an atom is an atom, but the proof requires concepts
that I do not plan to get into in this course. However, the proof of this statement
for the integral domain special case (which is really the case we are interested
in when it comes to these concepts related to factorization) is straightforward,
and will be a future exercise.
Theorem 79. The following are equivalent for a proper ideal P of a ring R.
1. P is prime.
2. Whenever I
1
, . . . , I
n
are ideals with I
1
I
n
P, then some I
i
P.
84
3. Whenever a
1
a
n
P, then some a
i
P.
4. Whenever ab P, then a P or b P.
5. R P is multiplicatively closed.
6. There is a multiplicatively closed S R such that P is maximal among
the ideals of R disjoint from S.
7. R/P is an integral domain.
Consequently, a nonunit p is prime (p [ ab p [ a or p [ b) (p [ a
1
a
n
p divides some a
i
).
Proof. (1) (2): Assume P is prime, I
1
, . . . , I
n
are ideals, and I
1
I
n
P.
The case n = 1 is trivial, so let us assume n 2 and I
n
_ P. Then I
1
I
n1

P, and by induction I
i
P for some i n 1. (2) (3): If (2) holds and
a
1
a
n
P, then (a
1
) (a
n
) = (a
1
a
n
) P, so some a
i
(a
i
) P (3)
(4): Clear. (4) (5): Statements (4) and (5) are contrapositives. (5) (6):
If (5) holds, then take S = R P. (6) (1): Assume (6) holds, and let I and
J be ideals of R with IJ P. If I _ P and J _ P, then P _ I + P, J + P,
so by maximality there are s S (I + P) and t S (J + P), and st is an
element of S in (I +P)(J +P) = IJ +IP +PJ +P
2
P +P +P +P = P,
a contradiction. Therefore I P or J P. (5) (7): Note that the nonzero
elements of R/P are precisely those of the form P +a with a R P. So (5) is
saying that the product of any two nonzero elements of R/P is nonzero, or in
other words that R/P is an integral domain.
The last statement is simply a rewording of (1), (3), and (4) for the ideal
(p).
Another consequence of the above theorem is that a ring is a domain if and
only if (0) is prime. (Since R/(0)

= R by the Fundamental Homomorphism
Theorem, or more explicitly via (0) +x x.)
Example. In this example we explore the dierence between irreducible and
prime elements of a domain.
1. The prime elements of Z are 2, 3, 5, . . .. (Though in prior sections
when I have said prime I was using the everyday usage and only includ-
ing the positive primes.) In Z (or any PID), every nonunit is a product
of primes, and these products are unique up to order and associates. We
will discuss this more later.
2. Let D be a domain. Then X is prime in D[X]. This can be seen directly
by noting that the product of two polynomials with nonzero constant
term has nonzero constant term. A slightly fancier way to see this is the
following: the map : D[X] D : f f(0) is an epimorphism with
kernel (X), so D[X]/(X)

= D is an integral domain. One can similarly
show that X is prime in D[[X]].
85
3. Consider the domain D = XC[X] +1 of polynomials over C with real con-
stant term. From degree considerations, we can see that X is irreducible
in D. (If X = fg, then at least one of f and g must be a nonzero constant,
hence a unit.) However, because X [ X(X) = (iX)
2
and X [ iX, the
element X is not prime in D. (If Xf = iX in D, then by cancelation in
the domain C[X] we obtain f = i, a contradiction to the fact that the
constants in D are real.)
4. Let D be a domain and R be the subring of D[X] consisting of polynomials
with no degree 1 term. It is not too dicult to verify that X
2
and X
3
are irreducible in R (no degree 1 polynomials in R), and that neither one
divides the other in R (because X / R). But the equation (X
2
)
3
= (X
3
)
2
then shows that neither X
2
nor X
3
is prime in R.
Example. We have (p
1
) (p
n
) = (p
1
p
n
) for non-associate primes
p
1
, . . . , p
n
of a domain. (Future exercise.)
Theorem 80. Let M be a proper ideal of a ring R. Then M is maximal if and
only if R/M is a eld. Therefore every maximal ideal is prime.
Proof. Exercise.
Example. The ideal (X) is a prime but not maximal ideal of Z[X], since
Z[X]/(X)

= Z is not a eld. For any nonzero prime p of Z, the ideal (p, X)
is maximal, because Z[X]/(p, X)

= Z
p
is a eld. (To see that those rings are
isomorphic, note that : Z[X] Z
p
: f [f(0)] is a ring epimorphism whose
kernel is the set of polynomials with constant term a multiple of p, or in other
words ker = (p, X).)
Theorem 81. Let S be a multiplicatively closed subset of a ring R, and let I be
an ideal disjoint from S. Then I is contained in some (necessarily prime) ideal
P maximal among the ideals of R disjoint from S.
Proof. The union of a nonempty chain of ideals disjoint from S is an ideal (past
exercise), and it is necessarily disjoint from S. So every nonempty chain in the
partially ordered set of ideals disjoint from S has an upper bound, namely its
union. The conclusion now follows by Zorns Lemma.
Lemma 82. Every union or intersection of a nonempty chain of prime ideals
is prime.
Proof. Exercise.
Example. The requirement that it is a chain is important. For example, in Z
we have (2) (3) = (6), which is not prime, and (2) (3) is not even an ideal,
since it does not contain 5 = 2 + 3.
Theorem 83. Let I be a proper ideal of a ring R. Every prime ideal containing
I contains a prime ideal minimal among the prime ideals containing I.
86
Proof. By the above lemma, the intersection of any nonempty chain of prime
ideals containing I is a prime ideal (and it necessarily contains I), which is a
lower bound for that chain. Thus the conclusion follows by Zorns Lemma.
Denition.
1. The minimal ideals in the above theorem are called the minimal primes of
I. Note that I necessarily has at least one minimal prime, because there
is some maximal (hence prime) ideal containing it. The minimal primes
of R are the minimal primes of (0). Note that (0) is the unique minimal
prime of an integral domain.
2. Let I be an ideal of a ring R. The radical of I is

I = x R [ x
n

I for some n Z
+
. Note that

R = R. From the following theorem it


follows that

I is an ideal. We call
_
(0) the nilradical of R; its elements
are called nilpotent elements. In other words, an element is nilpotent if
some positive power of it is zero. We say I is radical if I =

I. The
radical of an ideal is (fortunately) radical (exercise), so the map I

I
is a surjection from the set of ideals of R onto the set of radical ideals of
R.
Example.
1. All nilpotent elements are zero divisors. (Certainly 0 is a zero divisor,
and, if x is a nonzero nilpotent and n is the minimum positive integer
with x
n
= 0, then x
n1
,= 0 and x(x
n1
) = x
n
= 0.) Therefore the only
nilpotent element in a domain is 0.
2. The element [2] in Z
4
is nilpotent because [2]
2
= [4] = [0].
Theorem 84. Let I be a proper ideal of a ring R. Then

I is the intersection of
the prime ideals containing I, or, equivalently, is the intersection of the minimal
primes of I. Therefore every prime ideal is radical.
Proof. Let J be the intersection of the prime ideals containing I. (Since every
prime ideal containing I contains a minimal prime of I, any element in every
minimal prime is also in J, so J is equal to the intersection of the minimal
primes of I.) If x

I and P is a prime ideal containing I, then there is an


n Z
+
with x
n
I P, and hence x P. Therefore

I J. On the other
hand, if x /

I, then x
n

n=1
is a multiplicatively closed subset of R disjoint
from I, so I is contained in a prime ideal disjoint from x
n

n=1
, and thus x / J.
So J

I, and therefore

I = J.
Example. We calculate the radical ideals of Z. Obviously
_
(0) = (0) and

Z = Z. By our knowledge of prime factorization in Z (which we will ocially


prove later), every other ideal can be written in the form (p
m1
1
p
mn
n
), where
p
1
, . . . , p
n
are distinct positive primes and m
1
, . . . , m
n
Z
+
. (Changing the
generator by a factor of 1 does not change the principal ideal, because unit
multiples are associates by an exercise.) The prime ideals containing this ideal
87
are those generated by positive primes dividing p
m1
1
p
mn
n
, or in other words
p
1
, . . . , p
n
. Thus
_
(p
m1
1
p
mn
n
) = (p
1
) (p
n
) = (p
1
p
n
).
Theorem 85 (Cancelation Isomorphism Theorem for Rings). Let R be a ring
and I J be ideals of R. Then
R/I
J/I

= R/J.
Proof. (Note that J/I is an ideal of R/I by the Correspondence Theorem, so
the statement of this theorem makes sense.) Dene : R/I R/J : I +
x J + x. We saw in the proof of the Cancelation Isomorphism Theorem for
groups that this is a well-dened group epimorphism with kernel J/I. By the
Fundamental Homomorphism Theorem, it will now suce to show that is a
ring homomorphism. We have (I + 1) = J + 1, and for a, b R, we have
((I +a)(I +b)) = (I +ab) = J +ab = (J +a)(J +b) = (I +a)(I +b), as
desired.
Theorem 86 (Correspondence Theorem for Prime Ideals). Let R be a ring and
I be a proper ideal of R. The map P P/I is an order-preserving one-to-one
correspondence between the prime ideals of R containing I and the prime ideals
of R/I. The inverse map is Q
1
[Q], where : R R/I : r I +r.
Proof. We already know the version of this theorem with prime ideals replaced
by ideals. So all that we need to do is show that an ideal P containing I is
prime in R if and only if P/I is prime in R/I. We have
R/I
P/I

= R/P by the
Cancelation Isomorphism Theorem, hence one of those rings is a domain if and
only if the other is, and hence P is prime if and only if P/I is.
Denition. A ring is reduced if its nilradical is zero, or in other words it has
no nonzero nilpotents. Note that an ideal I of a ring R is radical if and only if
R/I is reduced, since there is an x R I with x
n
I if and only if there is a
nonzero element of R/I whose nth power is zero.
Corollary 87 (Correspondence Theorem for Radical Ideals). Let R be a ring
and I be a proper ideal of R. The map J J/I is an order-preserving one-to-
one correspondence between the radical ideals of R containing I and the radical
ideals of R/I. The inverse map is Q
1
[Q].
Proof. All that remains to be veried is that an ideal J containing I is radical
in R if and only if J/I is radical in R/I. We have R/J

=
R/I
J/I
, hence one ring
is reduced if and only if the other is, hence J is radical if and only if I is.
Example.
1. Integral domains are reduced.
2. Let n 2 and p
1
, . . . , p
n
be distinct positive primes of Z. Then Z
p1pn
=
Z/(p
1
p
n
) is reduced (since (p
1
p
n
) is radical in Z) but not an integral
domain.
Exercises.
88
1. (a) Show that a [ b b (a) (a) (b). (Hint: Recall that (a) = Ra.)
(b) Prove that, in a domain, (a) = (b) a = bu for some unit u.
(Hint: You will only need the domain hypothesis for the direction
. For this direction, by part (a) you have a = bu and b = av for
some u, v R. Remember that in a domain you can cancel nonzero
elements.)
(c) Show that a prime element is irreducible. (Hint: If p = ab, then
(a)(b) = (p).)
(d) Prove that a nonunit in a domain is irreducible if and only if it
cannot be written as a product of two nonzero nonunits. (Hint: The
direction where you use the domain hypothesis is . For the case
where you need to show that a nonzero irreducible cannot be written
as a product of two nonzero nonunits, use part (b).)
2. Show that a proper ideal M of a ring R is maximal if and only if R/M
is a eld. (Hint: Note that a ring is a eld if and only if the zero ideal
is maximal, then use the Correspondence Theorem. Alternatively, note
that the nonzero elements of R/M are of the form M +a for a / M. By
maximality, we have 1 R = M + (a); use that to give an inverse for
M +a.)
3. [Extra Credit] Prove that a union or intersection of a nonempty chain of
prime ideals is prime. (Hint: You already know this is an ideal by a past
exercise, so all you need to show is primeness. The more dicult one is
the intersection. For this, assume ab is in the intersection and b is not.
Pick one of the prime ideals not containing b, show that it contains a, and
show that all the prime ideals contain a. You will need to use the fact
that prime ideals form a chain in the last part.)
4. [Extra Credit]
(a) Show that an ideal is radical it is an intersection of prime ideals
it is an intersection of radical ideals. (Hint: Cite a theorem to get
(1) (2) (3), then show (3) (1) using the denition of radical.)
(b) Show that the radical of an ideal is radical. (Hint: Use the previous
part.)
(c) Show that two ideals are contained in the same prime ideals they
have the same minimal primes they have the same radical.
(d) Prove that an ideal I is radical if and only if x I whenever x
2
I.
(Note that it follows from this that a ring is reduced if and only if
every nonzero element has a nonzero square. Hint: For the nontrivial
direction , prove by induction on n that for each n Z
+
we have
x
n
I x I. For n 3, show that x
n/2
or x
(n+1)/2
in I.)
89
3.3 The Construction of Z and
[Durbin: Section 30]
We have been making free use of various properties of the number systems Z
and (along with those of 1 and C) so that we could give interesting concrete
examples of the abstract concepts we have been discussing, but now is time to
go back and nish our construction of them. Recall our construction of the
ordinals in Chapter 1. Starting with literally nothing (i.e., the empty set),
we rigorously constructed all the ordinals, including N = . We proved that
N forms a commutative semiring with both operations cancellative (although
we did not use that terminology). (The operations are dened to be cardinal
arithmetic.) Starting from here, this section will give a precise denition of the
domain Z and the eld . Along the way, we will learn about groupoids of
fractions, elds of fractions, and localizations.
Denition. Let H be a monoid. If H has no absorbing element, or if H = 0,
we set H

= H, and otherwise we set H

= H 0. (Note that this denition


is consistent with the notation for rings, and that H

.)
Theorem 88 (Groupoid of Fractions). Let H be a cancellative commutative
monoid. Dene a relation on H H

by (a, s) (b, t) at = sb. This


relation is a congruence relation (this is dened the same way for monoids
as for groups), and we use the symbol
a
s
to denote [(a, s)]. The groupoid of
fractions of H is QG(H) = (HH

)/ , which is a commutative monoid under


the operation
a
s
b
t
=
ab
st
. Furthermore:
1. The identity of QG(H) is
1
1
=
s
s
for any s H

. In fact, for any a H


and s, t H

we have
at
st
=
a
s
.
2. H embeds into QG(H) via a
a
1
. (We thus typically think of H as a
subset of QG(H) by identifying a with
a
1
.)
3. H has an absorbing element if and only if QG(H) does. In this case, the
absorbing element of QG(H) is
0
1
=
0
s
for any s H

.
4. QG(H)

=
a
s
[ a, s H

= QG(H)

. (The inverses are given by (


a
s
)
1
=
s
a
.) Therefore QG(H) is cancellative.
5. If M is any cancellative commutative monoid containing H and in which
every element of H

is invertible, then QG(H) embeds in M via


a
s
as
1
.
In other words, we may think of QG(H) as the unique smallest cancellative
commutative monoid containing H in which each element of H

is invert-
ible. In particular, if H has no absorbing element, then QG(H) is the
unique smallest group containing H, and is called the group of fractions
of H.
Proof. We rst show that is a congruence relation on the commutative monoid
H H

.
90
(Reexive:) For any (a, s) H H

, we have as = sa, so (a, s) (a, s).


(Symmetric:) If (a, s) (b, t), then at = sb, so bs = ta and (b, t) (a, s).
(Transitive:) Assume (a, s) (b, t) and (b, t) (c, u). Then at = sb and
bu = tc. Thus a(tu) = (at)u = (sb)u = (bu)s = (tc)s, and hence au = sc
and (a, s) (c, u).
(Congruence:) Assume (a, s) (a

, s

) and (b, t) (b

, t

). Then as

=
sa

and bt

= tb

, so (ab)(s

) = (as

)(bt

) = (sa

)(tb

) = (a

)(st) and
(a, s)(b, t) = (ab, st) (a

, s

) = (a

, s

)(b

, t

).
The fact that QG(H) = (HH

)/ is a commutative monoid under the given


operation now follows as did the analogous result from group theory (look up
when we constructed groups from congruence relations if you do not remember
this).
1. The fact that the identity of QG(H) is [(1, 1)] =
1
1
is clear from the deni-
tions. We also note that for any a H and s, t H

we have (at)s = (st)a


and hence
at
st
=
a
s
.
2. The map a
a
1
is clearly a monoid homomorphism from H into H
S
. In
fact, it is a monomorphism since if
a
1
=
b
1
for a, b H, then a = a 1 =
1 b = b.
3. (): Assume H has an absorbing element 0. Then for each s S we
have 0 s = 0 = 1 0 and hence
0
1
=
0
s
. Thus for every a H and s S
we have
0
1
a
s
=
0
s
=
0
1
, showing that
0
1
is the absorbing element of H
S
.
(): Assume H
S
has an absorbing element
a
s
. Then for each x H we
have
xa
s
=
x
1
a
s
=
a
s
, so xas = sa and hence xa = a, showing that a is the
absorbing element of H.
4. If
a
s
b
t
= 1, then ab = st H

, hence a H

. Conversely, if a, s H

,
then
a
s

s
a
=
as
as
= 1, so
a
s
QG(H)

. Thus QG(H)

=
a
s
[ a, s H

=
QG(H)

, where the last equality follows from the previous part.


5. It is easy to check that that
a
s
as
1
is a homomorphism. In fact, it is
a monomorphism since if as
1
= bt
1
, then at = sb and
a
s
=
b
t
.
Remark. With the denitions above, the set H is not literally a subset of
QG(H), but it is natural to think of a H as the same as
a
1
H
S
, so algebraists
typically consider H to be a subset of H
S
. If this bothers you to call H a subset
of QG(H), you can imagine that after QG(H) is dened, we throw out the
elements of the form
a
1
and replace them with a, so that H truly is literally a
subset of QG(H).
The abelian group (Z, +) of integers is the group of fractions of (N, +). Note
that the notation of the above theorem has to be adjusted to additive notation,
and we write a b instead of
a
b
.
91
1. When we proved the properties of cardinal arithmetic, we showed that for
each a < b in N there is a c N with a + c = b, hence b a = c and
c a = c. So every integer is exactly one of the following: 0, an element
of Z
+
, or additive inverse of an element of Z
+
. (If a 0 and b > 0, then
a + b > 0 + 0 = 0, so no element of Z
+
has an additive inverse in N, and
the three categories are distinct.)
2. For any a Z and member b of an additive group, the symbol ab is dened
(this is exponentiation in additive notation). Thus the multiplication on Z
is now dened. The multiplication restricted to N is the same as cardinal
arithmetic, and (N, +, ) forms a commutative semiring under cardinal
addition and multiplication. (We have already veried these facts back
in Chapter 1, but of course we did not phrase it this way back then.)
The distributive laws a(b + c) = ab + ac and (b + c)a = ba + ca are
a special case of the exponential rules for groups in additive notation.
(Our proof of these rules did not involve any more properties of Z than
we have already developed in this section.) We also have the identities
ab = (a)(b) = ((a)b) = a(b) from exponentiation properties
(whose proof again does not involve any properties of Z beyond the basic
ones developed thus far); an appropriate use of these identities allows us
to deduce the commutativity of on Z from its commutativity on N. A
similar (though somewhat more tedious) type of trick allows us to prove
that is associative on Z, and we have now proven that (Z, +, ) is a
commutative ring. Finally, the same kind of trick allows one to deduce
that Z is a domain, i.e., that is cancellative on Z.
To sum everything up: we have now nished our precise construction of Z and
proven that it is an integral domain.
Theorem 89 (Field of Fractions). Let D be an integral domain. The quotient
groupoid of (D, ) can be made into a eld, called the eld of fractions of D and
denoted QF(D), by dening
a
s
+
b
t
=
at+bs
st
. Its identity (resp., zero) is
1
1
=
s
s
(resp.,
0
1
=
0
s
) for any s S, and inverses are given by (
a
s
)
1
=
s
a
for a, s D

.
We also have the identity
a
s
+
b
s
=
a+b
s
for a, b D and s D

. Furthermore:
1. D is a subring of QF(D). (Recall that we are identifying a with
a
1
.)
2. If R is any domain containing D in which every element of D

is invert-
ible, then QF(D) embeds in R via
a
s
as
1
. In other words, we may
think of QF(D) as the unique smallest domain containing H in which ev-
ery element of H

is invertible, and hence as the unique smallest eld


containing D.
Proof. First note that the addition operation is well-dened, because if
a
s
=
a

and
b
t
=
b

, then as

= sa

and bt

= tb

, so (at +bs)(s

) = (as

)tt

+(bt

)ss

=
(sa

)tt

+ (tb

)ss

= (st)(a

+ b

), and hence
at+bs
st
=
a

+b

. We already
know from the quotient groupoid construction that (QF(D), ) is a commutative
monoid with every nonzero element invertible, and it is a routine (though mildly
92
tedious) process to check that (QF(D), +) is an abelian group with zero
0
1
=
0
s
for all s D

, and that the distributive law holds, so QF(D) is a eld. The


statements about the identity and inverses are already known by the quotient
groupoid theorem, and for the last identity we compute
a
s
+
b
s
=
as+bs
s
2
=
(a+b)s
s
2
=
a+b
s
. We already know that the maps in (1) and (2) are monoid
monomorphisms, so all we need to verify is that they respect addition.
1.
a+b
1
=
a1+b1
11
=
a
1
+
b
1
.
2. (at +bs)(st)
1
= as
1
+bt
1
.
Denition. The eld of rational numbers is = QF(Z).
Denition. Let R be a ring.
1. A saturated subset of R is a subset that contains all divisors of its elements.
(In other words, a subset S is saturated if ab S a, b S.)
2. By a multiplicative set of R, we mean a nonempty saturated multiplica-
tively closed subset of R contained in R

.
3. It follows from the denitions that any intersection of a nonempty family
of saturated subsets (resp., multiplicative sets) of R is saturated (resp.,
multiplicative). Hence every subset A of H is contained in a unique small-
est saturated subset (resp., multiplicative set), called the saturation of A
(resp., called the multiplicative set generated by A).
Example.
1. R, R

, and are saturated subsets of R. A saturated subset is equal to


R if and only if it contains 0.
2. R

is the smallest (with respect to inclusion) nonempty saturated subset


of R, and also the smallest multiplicative set of R. If R is a domain, then
R

is a multiplicative set of H, and it contains all other multiplicative


sets.
3. If I is an ideal of a ring R, then R I is a saturated subset of R. If P is a
prime ideal of R, then RP is a multiplicative set of R. We will generalize
these statements momentarily.
4. The multiplicative set generated by the primes of Z is 1, 2, 3, 5, 7, . . ..
Remark. Let R be a ring. The following facts are simple consequence of the
denitions. (It would be a good idea to think about why each is true while
reading.)
1. A R is saturated if and only if every multiple of an element of R A is
in R A.
93
2. The saturation of A R consists of all divisors of elements of A.
3. The multiplicative set generated by A R consists of all divisors of prod-
ucts a
1
a
n
with each a
i
A.
Proposition 90. Let R be a ring and _ A R. Then A is a saturated subset
(resp., a multiplicative set) of R if and only if R A is a union of ideals (resp.,
prime ideals).
Proof. (): Assume A is a saturated subset of R. Then for each x R A we
have (x) = Rx R A. Thus R A =

xR\A
(x). Now let us further assume
that A is a multiplicative set. Then for each x R A we have (x) disjoint
from the multiplicatively closed set A, so (x) is contained in some prime ideal
P
x
disjoint from A. Thus R A =

xR\A
P
x
. (): Assume R A is a union of
ideals (resp., prime ideals). If this is the empty union, then A = R is certainly
a multiplicative set, so let us assume that this is a union of a nonempty family.
Then A is an intersection of complements of ideals (resp., prime ideals) by De
Morgans Laws, hence an intersection of saturated subsets (resp., multiplicative
sets) of R, hence a saturated subset (resp., multiplicative set) of R.
Theorem 91 (Localization). Let S be a multiplicative set of a domain D, and
let D
S
=
a
s
QF(D) [ a D, s S. Then D
S
is a domain between D and
QF(D). Furthermore:
1. An element
a
s
of D
S
is a unit if and only if (a) S ,= .
2. If R is any domain containing D and in which every element of S is
invertible, then D
S
embeds in R via
a
s
as
1
. In other words, we may
think of D
S
as the unique smallest domain containing D in which each
element of S is invertible.
3. The map I I
S
=
a
s
[ a I, s S is a surjection from the set of ideals
of D disjoint from S onto the set of proper ideals of D
S
. A right inverse
is the map J J D.
4. The map P P
S
is an order-preserving one-to-one correspondence be-
tween the prime ideals of D disjoint from S and the prime ideals of D
S
.
Its inverse is the map Q Q D.
Proof. We have 1 =
1
1
D
S
, and for a, b D and s, t S we have
a
s
b
t
=
ab
st

D
S
and
a
s
+
b
t
=
at+bs
st
D
S
. (Note that 1 S since S is saturated and st S
since S is multiplicative.) Therefore D
S
is a subring of QF(D) by the Subring
Test. Since 1 S, we have D D
S
QF(D), and D
S
is a domain since every
subring of a domain is a domain.
1. If
a
s
b
t
= 1, then st = ab Ha S. Conversely, if there is a b H with
ab S, then
a
s
bs
ab
=
abs
abs
= 1.
2. It is easy to check that that
a
s
as
1
is a homomorphism. In fact, it is
a monomorphism since if as
1
= bt
1
, then at = sb and
a
s
=
b
t
.
94
3. It is routine to check that the maps take ideals to ideals, and the rst
takes ideals disjoint from S to proper ideals of D
S
by part (1). It is also
simple to verify that (J D)
S
= J for each ideal J of D
S
, so the rst map
is surjective and the second map is a right inverse.
4. If P is a prime ideal of D disjoint from S and
a
s
b
t
P
S
, then there are
c P and u S with abu = stc P, hence a P or b P. So the rst
map takes prime ideals to prime ideals, and it is even simpler to verify that
the second map does. All that remains is to show that P
S
D = P for each
prime ideal P of D disjoint from S. We certainly have P P
S
D. On
the other hand, for each a P and s S with
a
s
D, we have bs = a P
for some b D, hence b P, hence
a
s
= b P. Thus P = P
S
D, as
desired.
Denition. Let P be a prime ideal of a domain D. The localization of D at
P is D
P
= D
R\P
. (Note that R P is a multiplicative set and P is not. For
emphasis: The denominators of fractions in D
P
are the elements not in P.)
Example. Let p be a prime of Z. The domain Z
(p)
consists of all rational
numbers of the form
a
s
with p [ s. It is not a eld, because
a
s
is a unit if and
only if p [ a.
Remark.
1. The prime ideals of D
P
are those of the form Q
P
= Q
R\P
, where Q is a
prime ideal disjoint from R P, or in other words a prime ideal contained
in P. Thus P
P
is the unique maximal ideal of D
P
.
2. A ring with only one maximal ideal is called quasilocal. (Some people
simply say local, but that sometimes means something else.) Fields are
of course quasilocal, and localizations at prime ideals provide some more
interesting examples of quasilocal rings.
3.4 The Construction of 1
[Durbin: Sections 28-31]
Unfortunately, before we can construct 1, we need to develop a precise
denition of the usual total order on . (Because extending to 1 is more
of an analytic concept than an algebraic one, and thus involves notions such as
limits and so on, and we need a total order for those to make sense.)
Denition.
1. A partial order on a ring is called compatible if (i) a b a+c b +c
and (ii) a b and c 0 ac bc. Number (i) could equivalently be
stated as: a < b a + c < b + c. If the ring is a domain, then (ii) could
be stated as: a < b and c > 0 ac < bc. Note that (i) and (ii) imply
that if a, b 0, then a +b 0 and ab 0.
95
2. A partially (resp., totally) ordered ring is a pair (R, ) such that R is a
ring and is a compatible partial (resp., total) order on R. (As usual,
we will often simply refer to R as the ordered ring if there is no potential
ambiguity.)
3. An element x of a partially ordered ring is called positive if x > 0 and
negative if x < 0.
4. An order embedding (resp., order isomorphism) between partially ordered
rings is a map that is both an order embedding between partially ordered
sets and a monomorphism (resp., isomorphism) of rings. The relation or-
der isomorphic to forms an equivalence relation on the class of partially
ordered rings.
5. A positive subset of a ring R is a subset such that (i) 0 P, (ii) P is closed
under addition and multiplication, and (iii) P (P) = 0. (Here P
denotes the set p [ p P.)
Remark.
1. In this section we will precisely dene the usual orders on Z, , and 1,
and show that they are totally ordered domains with these orders.
2. In a partially ordered ring, we have a 0 a 0, and a 0 a 0.
(Add a or a to the inequalities as appropriate.) Thus, for each element
a of a totally ordered ring we have either a 0 and a 0, or a 0 and
a 0. In the preceding statements, we could have equivalently replaced
each with < and each with >.
3. If a, b 0 in a totally ordered ring, then a, b 0 and ab = (a)(b)
0. Hence, if a, b < 0 in a totally ordered domain, then ab > 0.
4. If a 0 and b 0 in a totally ordered ring, then ab = (a)b 0 and
hence ab 0. Hence, if a < 0 and b > 0 in a totally ordered domain, then
ab < 0.
5. If u is a unit in a totally ordered ring, then we have either s, s
1
> 0 or
s, s
1
< 0. (If one is positive and one is negative, then 1 = ss
1
< 0, a
contradiction.)
6. It follows from (3) that any even power of an element in a totally ordered
ring is non-negative. Thus, in any totally ordered integral domain (or
more generally in any totally ordered reduced ring), any even power of a
nonzero element is positive. In particular, in any totally ordered ring we
have 1 = 1
2
> 0 and 1 < 0. Thus n 1 > 0 for all n Z
+
, and it follows
that any totally ordered ring must have characteristic 0.
7. On any totally ordered ring R, one can dene the absolute value function
[ [ : R R
+
0
by [x[ = x if x 0 and [x[ = x if x < 0. Checking a
couple cases gives the rules (i) [x[ = 0 x = 0, (ii) [ x[ = [x[ x,
96
and (iii) [xy[ = [x[[y[ for all x, y R. Hence [x[ + [y[ x + y and also
[x[ +[y[ xy = (x+y), and the Triangle Inequality [x[ +[y[ [x+y[
follows.
8. Since Z embeds into every ring of characteristic 0, and thus embeds into
every eld of characteristic 0, every totally ordered ring contains a copy
of Z and every totally ordered eld contains a copy of . (In other words,
Z is the smallest totally ordered ring and is the smallest totally ordered
eld. In particular, every totally ordered ring is innite.) In fact, these
copies are not just isomorphic as rings, but isomorphic as totally ordered
rings. To see this, recall that Z embeds into any ring of characteristic 0
via n n 1. This is in fact an order embedding, because if m < n, then
n m Z
+
and (n m) 1 > 0 and hence m 1 < n 1. The fact that
is order-embedded into any totally ordered eld follows from the fact
that there is only one way to extend the total order on Z to a compatible
total order on . (We will prove this in a moment.)
9. In any totally ordered eld, if x < y, then x =
x+x
2
<
x+y
2
<
y+y
2
= y.
(Here we use 2 to denote the element, namely 1 + 1, that corresponds to
2 in the copy of inside the totally ordered eld.) Thus, in a totally
ordered eld, there is an element between every two distinct elements.
Theorem 92. Let R be a ring. For each positive subset P of R dene a relation

P
on R by a
P
b b a P, and for each compatible partial order
on R dene P

= x R [ x 0. Then the map P


P
is a one-to-
one correspondence between the positive subsets of R and the compatible partial
orders on R, with inverse map P

. The total orders are the images of the


positive subsets P with P (P) = R.
Proof. First we show that
P
is a compatible partial order on R for each positive
subset P of R.
1. (Reexive:) For a R we have a a = 0 P and hence a
P
a.
2. (Anti-symmetric:) If a
P
b and b
P
a, then b a P and b a =
(a b) P, so b a = 0 and a = b.
3. (Transitive:) If a
P
b and b
P
c, then b a, c b P, so c a =
(c b) + (b a) P.
4. (Compatible with addition:) If a
P
b, then (b +c) (a +c) = b a P
and a +c
P
b +c.
5. (Compatible with multiplication:) If a
P
b and c
P
0, then b a P
and c = c 0 P, so bc ac = (b a)c P and ac
P
bc.
Next we show that P

is a positive subset for each compatible partial order


on R. Of course 0 P

, and we have already noted that P

is closed under
addition and multiplication. If a P

(P

), then a 0 and a 0 (since


a P

), so 0 = a +a a + 0 = a and thus a = 0.
97
If is a compatible partial order on R, then a b b a 0 b a
P

a
P

b, so =
P

. If P is a positive subset of R, then x P x0


P 0
P
x x P

P
, so P = P

P
. The previous two sentences show that
the maps are bijections and are inverses of each other.
Finally, a partial order is total a b or b a for all a, b R
b a 0 or a b 0 for all a, b R a 0 or a 0 for each a R
P

(P

) = R.
Denition.
1. In the notation of the above theorem, we call
P
the partial order induced
by P, and P

the positive cone of . The positive cone of a partially


ordered ring (R, ) is the positive cone of , which we denote by R
+
0
. We
abbreviate R
+
= R
+
0
0, i.e., R
+
is the set of positive elements of R.
(While it is convenient to consider 0 to be part of the positive cone, the
element 0 is not typically considered positive.)
2. We make Z into a totally ordered domain by dening to be the par-
tial order induced by N. (So this makes N the positive cone of Z and
the notation Z
+
we have been using is consistent with the use in (1).)
Fortunately, it follows from the properties of cardinal arithmetic that the
partial order that N inherits from Z is the same as that it inherits from
ON. (Also, after a little consideration one can see that any compatible
total order on Z extending the previously dened order on N must have N
as its positive cone, and thus the total order we have dened on Z is the
unique compatible total order that extends the order on N.)
3. A well-ordered integral domain is a totally ordered domain in which the
positive cone is a well-ordered set. (For emphasis, the domain itself is
not a well-ordered set, just its positive cone.) So Z is one example of a
well-ordered domain. We will see in a moment that it is (up to order iso-
morphism) the only well-ordered domain. This is a nice way to know that
the denition we have chosen for Z is the correct one, because, however
one imagines Z, it is a well-ordered domain, and hence order isomorphic
to the ordered integral domain we have dened.
Theorem 93. Every well-ordered integral domain is order isomorphic to Z.
Proof. Let D be any well-ordered integral domain. As we have previously dis-
cussed, the map : Z D : n n 1 is an order-embedding, and all that
remains is to show that it is surjective. The smallest x D
+
(here we use the
fact that D is a well-ordered domain) must satisfy x x
2
(since x
2
> 0)
and also x
2
x 1 = x, so x
2
= x and canceling gives x = 1. Thus 1
is the smallest element of D
+
. We have 0 = 1 0 = (0) [N], and for
x D
+
we have x 1 D
+
0
(since x 1) and by induction x 1 = (n)
for some n N, and hence x = (n) + 1 = (n + 1) [N]. Finally, for each
x < 0 in D, we have x D
+
and x = (n) for some n N, and hence
x = (n) = (n) [Z].
98
Theorem 94. Let (D, ) be a totally ordered integral domain. Then can be
extended to a unique compatible total order on QF(D). With this total order,
we have QF(D)
+
=
a
s
[ a, s > 0, and for s, t > 0 we have
a
s

b
t
at sb.
Proof. Let P =
a
s
[ a 0, s > 0 and note that 0 P. If a, b 0 and
s, t > 0, then ab, at + bs 0 and st > 0. Thus P is closed under addition and
multiplication. Note that any element of QF(D) can be written in the form
a
s
where s > 0. (Multiply the numerator and denominator by 1 if necessary.) It
follows that QF(D) = P (P). Moreover, if
a
s
=
b
t
and s, t > 0, then at = sb,
so a and b must either be both 0 or both 0. So the only element both in
P and P is 0. Therefore P is a positive cone for some compatible total order
on QF(D). Note that b a P b a D
+
0
, so this total order extends .
The last sentence follows from the denition of the partial order induced by a
positive subset.
For uniqueness, it suces to show that P is the positive cone of any com-
patible total order

on QF(D) extending . To see this, note that for a D


and s D
+
, we have
a
s

0 a

0 a 0. (Multiply by s to get from the


rst to the second statement, and by s
1
, which is also positive, to get from the
second to the rst.)
Denition. We make into a totally ordered eld by dening
a
s

b
t
at sb
(where the fractions are written with s, t > 0). By the above theorem, this is
the unique compatible total order on that extends the order on Z, and thus
the unique compatible order on that extends the order on N.
Now we dene the reals as equivalence classes of special kinds of sequences
of rationals.
Denition.
1. Now that we have a total order on , we can dene limits of sequences
in in the usual way. Let a
n

n=1
and L . We say that the
limit of the sequence a
n

n=1
is L, or that the sequence converges to L,
written lim
n
a
n
= L, if for every
+
there is an N Z
+
such
that [a
n
L[ < whenever n N. It is a consequence of this denition
that a sequence can have at most one limit. (One can use the Triangle
Inequality shows that, if L
1
and L
2
limits of the same sequence, then for
each
+
we have 0 [L
1
L
2
[ < , and thus L
1
= L
2
.) One can
prove that the usual Limit Laws from calculus also hold for sequences
in . (These proofs are essentially the exact same as the ones from an
advanced calculus class, so I will not prove them.)
2. A Cauchy sequence in is a sequence a
n

n=1
such that for every

+
there is an N Z
+
such that [a
m
a
n
[ < whenever m, n N.
For convenience, I will use CS() to denote the set of all Cauchy sequences
in .
3. We will say a sequence a
n

n=1
is eventually positive (resp., negative)
if there is an x
+
and N Z
+
with a
n
> x (resp., a
n
< x) for all
99
n N. (Take a moment to note the precise meaning of this. A sequence
could have every single term positive without being eventually positive, if
it converges to 0. Also, this terminology is by no means standard, it is
just a convenient temporary denition to have for when we construct 1.)
We will say it is eventually zero if there is an N Z
+
with a
n
= 0 for all
n N.
Remark.
1. The notions of limits and Cauchy sequences may be dened analogously
on any totally ordered ring, and to a certain extent one can do calculus
on any totally ordered ring, at least as far as the really basic Limit Laws
go.
2. You may remember from calculus that a sequence in 1 converges if and
only if it is Cauchy. (This is one way to state the completeness property
of 1.) One can show that a convergent sequence in must be Cauchy,
but not necessarily the other way around. (Once we have 1 constructed,
it is easy to come up with examples of non-convergent Cauchy sequences,
because we simply take an irrational number, and use its decimal repre-
sentation to construct a Cauchy sequence that does not converge in .)
3. A Cauchy sequence a
n

n=1
is always bounded, i.e., there is some M

+
with [a
n
[ M for all n Z
+
. To see this, note that there is an
N Z
+
with [a
m
a
n
[ < 1 for m, n N. Hence for n N we have
[a
n
[ [a
N
[ + [a
n
a
N
[ [a
N
[ + 1. Thus for all n Z
+
we have [a
n
[
max([a
1
[, . . . , [a
N1
[, [a
N
[ + 1).
4. The following is a convenient observation: lim
n
a
n
= L there is a
k
+
such that for every
+
there is an N Z
+
with [a
n
L[ < k
whenever n N. The analogous observation holds for Cauchy sequences
as well, and in these four equivalent denitions of limits/Cauchy sequences
one may equivalently replace the < with .
5. Another useful trick: lim
n
a
n
= 0 lim
n
[a
n
[ = 0.
Theorem 95 (Construction of 1).
1. CS() is a subring of

n=1
.
2. Dene a relation on CS() by a
n

n=1
b
n

n=1
lim
n
(a
n

b
n
) = 0. Then is a congruence relation on CS(). (A congruence
relation on a ring is a relation that is a congruence relation with respect
to both operations.) Thus CS()/ forms a ring, which we dene to be
the ring 1 of real numbers.
3. Every Cauchy sequence is either eventually positive, eventually negative,
or converges to zero. Two equivalent Cauchy sequences must both fall into
the same one of these three categories.
100
4. Dene 1
+
0
= [a
n

n=1
] [ each a
n
0. Then 1
+
0
forms a positive
subset of 1 and 1 = 1
+
0
(1
+
0
), so 1
+
0
forms the positive cone of a
compatible total order on 1, which we dene to be the standard total order
on 1. One explicit way of dening this order is: [a
n

n=1
] [b
n

n=1
]
lim
n
(a
n
b
n
) = 0 or there is an N Z
+
with a
n
b
n
for all n N.
5. order-embeds into 1 via x [x

n=1
]. We thus regard as a subset
of 1 by identifying each element of with the corresponding equivalence
class of a constant sequence.
Proof.
1. We use the Subring Test. Clearly any constant sequence is a Cauchy
sequence, so the zero and identity elements are in CS(). Let a
n

n=1
and b
n

n=1
be Cauchy sequences, with bounds M
1
and M
2
, respectively.
Let
+
. There is an N
1
Z
+
with [a
m
a
n
[ < for m, n N
1
, and
there is an N
2
Z
+
with [b
m
b
n
[ < for m, n N
2
. So, for m, n
max(N
1
, N
2
), we have [(a
m
b
m
) (a
n
b
n
)[ [a
m
a
n
[ +[b
n
b
m
[ < 2
and [a
m
b
m
a
n
b
n
[ = [a
m
b
m
a
m
b
n
+ a
m
b
n
a
n
b
n
[ [a
m
[[b
m
b
n
[ +
[a
m
a
n
[[b
n
[ < (M
1
+M
2
). Therefore a
n
b
n

n=1
and a
n
b
n

n=1
are
Cauchy sequences, as desired.
2. The fact that the relation is an equivalence relation is a fairly easy con-
sequence of the limit laws, as is the fact that it respects addition. To
show that it respects multiplication, note that if lim
n
(a
n
b
n
) =
lim
n
(c
n
d
n
) = 0, then lim
n
(a
n
c
n
b
n
d
n
) = lim
n
(a
n
b
n
)c
n
+
lim
n
(c
n
d
n
)b
n
= 0 + 0 = 0. (To get that lim
n
(a
n
b
n
)c
n
= 0,
pick M
+
with [c
n
[ M for all n, bound M[a
n
b
n
[ (a
n
b
n
)c
n

M[a
n
b
n
[, and use the Squeeze Theorem from calculus. The other limit
is done similarly.)
3. Let a
n

n=1
be any Cauchy sequence in that does not converge to 0.
Then there is some
+
such that for each N Z
+
there is an n N
with [a
n
[ . There is some N Z
+
with [a
m
a
n
[ <

2
for m, n N.
If we pick a larger N, this is still true, so we may choose this N so that
[a
N
[ . For n N, we have [a
N
a
n
[ <

2
, and thus a
n
and a
N
are
either both greater than

2
or less than

2
, as desired.
The last statement is a simple consequence of the Limit Laws.
4. It is clear that 1
+
0
is closed under addition and multiplication. Note that
the members of 1
+
0
(resp., 1
+
0
) are precisely those equivalence classes
represented by either the zero sequence or an eventually positive (resp.,
negative) sequence. Thus 1
+
0
(1
+
0
) = [0

n=1
] and 1
+
0
(1
+
0
) = 1,
as desired.
5. It is very straightforward to check that this map is a ring monomorphism.
To see that it is an order embedding, note that if x < y, then the sequence
y x

n=1
is eventually positive.
101
Of course, now that 1 is precisely dened and constructed, we will not
typically represent a real number as an equivalence class of a Cauchy sequence
of rationals, except if we are doing some extremely technical proof where such
a thing is required. We will see below that real numbers correspond to how
one usually thinks of them: decimal representations (or representations in any
base b 2, for that matter). We could have equivalently dened 1 as the set
of all decimal representations (this is known as Stevins Construction after the
16th century mathematician who was one of the rst promoters of the use of
decimals), but it is a bit trickier to dene the operations this way. (For example,
how precisely does one multiply two innite, non-repeating decimals?)
Theorem 96 (Some Properties of 1).
1. 1 is a eld. Division by a nonzero element (i.e., multiplying by that
elements inverse) can be done by [a
n

n=1
]/[b
n

n=1
] = [a
n
/b
n

n=1
].
(Note that only nitely many of the b
n
s can be zero, so the latter number
is dened even if some nitely many a
n
/b
n
terms are not dened. This is
because two Cauchy sequences that only dier in nitely many terms are
equivalent, so it does not matter what the terms up through the last index
with b
n
= 0 are.)
2. Every real number is bounded above by an integer. (This is the Archimedean
Property.) Also, every positive real number is bounded below by a rational
of the form
1
n
.
3. Both and 1 are dense subsets of 1, i.e., there is a rational (resp.,
irrational) number between every two distinct real numbers.
4. If a
n

n=1
is a Cauchy sequence in , then lim
n
a
n
= [a
n

n=1
] in 1.
5. A sequence in converges in if and only if it converges in 1 to a
rational limit. In this case, its limits in and 1 coincide.
6. Let b 2 be an integer. Then each non-negative real number has a unique
base b representation, i.e., it has a unique representative Cauchy sequence
of the form

m
j=0
c
j
b
m
+

n
k=1
a
k
b
k

n=1
, where m N, each c
j
, a
k
N,
c
m
> 0 if m ,= 0, and the sequence a
n

n=1
is not equivalent to b
1

n=1
. Conversely, every such sequence is Cauchy, so there is a one-to-one
correspondence between the real numbers and the base b representations.
We denote the real number with this representation as c
m
c
0
.a
1
a
2
. . .
b
,
and we omit the subscript in the case b = 10. To determine which of two
distinct non-negative real numbers is greater, we can simply compare the
rst digit in which they dier. (Of course, if one can compare two non-
negative real numbers, one can compare any two real numbers.) A real
number is rational if and only if its base b representation ends with some
repeating nite string of digits.
102
7. Every nonempty subset with an upper (resp., lower) bound has a least upper
bound (resp., greatest lower bound). (This property is called Dedekind
completeness or simply completeness.)
8. A sequence in 1 converges if and only if it is Cauchy.
9. 1 is, up to order isomorphism, the only complete totally ordered eld.
Proof.
1. Let [a
n

n=1
] 1

. We must show a
1
n

n=1
is Cauchy, because, if so, its
equivalence class will obviously be the desired inverse. (Again, note that
some nite number of a
1
n
s may be undened, but this does not matter.)
Because a
n

n=1
is eventually positive or eventually negative, there is an
N
1
Z
+
with x
+
with [a
n
[ > x for n N
1
. Let
+
be given.
Then there is an N
2
Z
+
with [a
m
a
n
[ < for m, n N
2
. So, for
n max(N
1
, N
2
), we have [a
1
m
a
1
n
[ =
|anam|
|am||an|
<

x
2
, as desired.
2. Let [x
n

n=1
] 1. Because x
n

n=1
is Cauchy, it has an upper bound
a
s

+
, and we may take a, s Z
+
. Thus each x
n

a
s

a
s
s = a, and
it easily follows that [x
n

n=1
] a. For the second statement, if y 1
+
,
then there is an N Z
+
with y
1
< N, and thus
1
N
< y.
3. It suces to show that there is a rational (resp., irrational) between every
two positive real numbers x < y. By (2), there is an n Z
+
with
1
n
< yx.
Also, there is a least m N with m ny. Then m 1 < ny, so x <
y
1
n

m
n

1
n
=
m1
n
< y. So we have established the rational version,
and there is a rational c with x

2 < c < y

2, hence x <
c

2
< y, and
c

2
is irrational. (If
c

2
=
a
b
for some a, b Z
+
, then

2 =
bc
a
, a contradiction
to the irrationality of

2. This irrationality can be proven directly, but


this is not really necessary, because

2 could be replaced by any positive


irrational in this argument, and from cardinality considerations we know
that such a thing must exist.)
4. Let a
n

n=1
CS() and 1
+
. By (2), there is a positive rational less
than , so there is an N Z
+
with [a
m
a
n
[ < for m, n N. Thus for
n N we have [a
n
[a
m

m=1
][ = [[a
n
a
m
]

m=1
[ = [[a
n
a
m
[

m=1
]
, as desired. (The rule [b
n

n=1
[ = [b
n
[

n=1
follows fairly easily from
the fact that a Cauchy sequence is either eventually positive, negative, or
zero.)
5. Because there is a positive rational less than every positive real, the de-
nitions of the relevant statements are equivalent.
6. Let y 1
+
0
. By the Archimedean property, there is a least positive integer
greater than y, hence a greatest x N with x y. Thus x y <
x + 1. We recursively construct the sequence a
n

n=1
as follows. Given
a
1
, . . . , a
n1
0, . . . , b1 with x+

n1
k=1
a
k
b
k
y < x+

n1
k=1
a
k
b
k
+
103
b
(n1)
, there is a greatest a
n
0, . . . , b 1 with x +

n
k=1
a
k
b
k
y,
and necessarily y < x+

n
k=1
a
k
b
k
+b
n
. (If y x+

n
k=1
a
k
b
k
+b
n
,
then the maximality of a
n
would force a
n
= b1, so y x+

n1
k=1
a
k
b
k
+
(b 1)b
n
+b
n
= x+

n1
k=1
a
k
b
k
+b
(n1)
, a contradiction.) Let y
n
=
x +

n
k=1
a
k
b
k
. For N m n, we have [y
m
y
n
[

n
k=m
[a
k
[[b
k
[ <

n
k=m
b
(k1)
=
b
(m1)
(1b
(nm+1)
)
1b
1
< b
2m
b
2N
. (Here we use
the geometric series formula. Its proof does not involve anything beyond
the basic properties we have developed.) For N > 2, we have N 2 <
2
N2
b
N2
by Cantors Theorem, so taking N large enough gets b
2N
less than any given positive real number, and thus the sequence y
n

n=1
is Cauchy. To see that y = [y
n

n=1
], note that the dierence between y
and [y
n

n=1
] is by construction less than any b
n
<
1
n
, hence less than
any positive real number, hence zero. One can use the geometric series
formula to show that lim
n

n
k=m
(b 1)b
k
=
b
m
(b1)
1b
1
= b
(m1)
.
Hence the maximality of the construction of x and the a
n
s prevents us
from ending the sequence a
n

n=1
with an innite string of (b 1)s.
We will discuss base b representations of integers later when we study
the Division Algorithm in Z, and this will give the unique representation
of x. For uniqueness of the other digits, let x

n
k=1
a

k
b
k

n=1
be
any other Cauchy sequence of the given form that is equivalent to x +

n
k=1
a
k
b
k

n=1
. By the maximality of x, we have x

x. Again using
the geometric series formula, one shows that lim
n

n
k=1
a

k
b
k
< 1,
and it follows that x = x

. One inductively proceeds through all the


terms in this manner to show that each a
n
= a

n
. The statement about
comparing non-negative real numbers follows from using the geometric
series formula to show that, if z, w 1 rst dier in digit k, with the
kth digit of z being greater, then all the terms of the Cauchy sequence
of the above form representing w are less than the corresponding ones
for the sequence of z. If a real numbers base b representation has an
innite string of the same digit at the end, then the corresponding Cauchy
sequence representing it converges to a rational number by the geometric
series formula, and that real number is equal to that rational limit. On
the other hand, one may nd the decimal representation of a rational
p
q
by
doing the usual algorithm of repeated long division by q. Because there
are only nitely many possible remainders (namely 0 through q 1), there
must be a repetition at some point, which leads to the same string of digits
being repeated again and again for the remainder of the representation.
(I will not write out the full details for this, but if you havent seen this
before, then try doing an example or two and you will see what I mean.)
7. Let A be a nonempty subset of 1 that is bounded above. (The other case
is similar.) One may reduce to the case where A has some non-negative
elements, so there is some non-upper bound that is a non-negative rational.
Let a
0
be the largest natural number that is not an upper bound for A.
(This exists because there is some smallest natural number that is an upper
104
bound, and a
0
is this minus 1.) For n 1, one recursively constructs a
n
to be the largest rational number representable with n 1 decimal digits
that is not an upper bound for A. (As one would expect, it turns out that
a
n
is simply a
n1
plus an appropriate (n1)st digit.) By the maximality
of the construction, the number [a
n

n=1
] must be an upper bound for A,
and by examining the decimal representation of any lesser real number,
one sees that it is the least upper bound.
8. (): Let x
n

n=1
be any convergent sequence in 1, say lim
n
x
n
= L.
Let 1
+
be given. Then there is an N Z
+
with [x
n
L[ < whenever
n N. So, for m, n N we have [x
m
x
n
[ [x
m
L[ +[L x
n
[ < 2,
showing that x
n

n=1
is Cauchy. (): Assume x
n

n=1
is Cauchy. For
each n Z
+
, let L
n
be the least upper bound of x
k

k=n
, and let L be
the greatest lower bound of L
n

n=1
. To see that lim
n
x
n
= L, let
1
+
be given. There is an N
1
Z
+
with L
N1
L < (otherwise L+
would be a greater lower bound of the L
n
s), and thus 0 L
n
L <
for n N
1
(since the L
n
s are decreasing). There is an N
2
Z
+
with
[x
m
x
n
[ < whenever m, n N
2
. For n max(N
1
, N
2
), there is some
k n with 0 L
n
x
k
< (otherwise L
n
would be a lesser lower
bound of x
k

k=n
), and [x
n
L[ [x
n
x
k
[ +[x
k
L
n
[ +[L
n
L[ < 3,
showing that lim
n
x
n
= L.
9. Let K be any complete totally ordered eld. Then K contains . The
same proof as above shows that a sequence in K converges if and only
if it is Cauchy. Moreover, the Archimedean property holds for K as a
simple consequence of completeness. (For x K, consider the set A =
n N [ n x. If A = , then 0 > x. [Actually A cannot be empty by
the Archimedean property we are about to prove, but that is besides the
point.] So let us assume A ,= , so A has a least upper bound y. Then
y 1 is not an upper bound for A, so y 1 n for some n A, and
hence n + 1 y, so n + 1 / A and n + 1 > x.) With these two facts, one
can prove similarly to how we did above that every element of K has a
decimal expansion. Dene : 1 K : [a
n

n=1
] lim
n
a
n
. (This
map is well-dened, because lim
n
(a
n
b
n
) = 0 implies lim
n
a
n
=
lim
n
b
n
. Note here that we are taking limits in K, not .) It is easy
to see that this map is a eld homomorphism, hence by a past exercise
a eld monomorphism. Also, if [a
n

n=1
] < [b
n

n=1
], then b
n
a
n
is
eventually positive, and it follows that lim
n
(b
n
a
n
) > 0 and hence
lim
n
b
n
> lim
n
a
n
. Lastly, for surjectivity, given y K, one nds
the decimal representation of y, and the corresponding element of 1 maps
onto y.
105
3.5 Polynomial Rings
[Durbin: Sections 34 and 35]
Denition. Let S

be a nonempty family of additive monoids (resp.,


additive groups). The direct sum of this family is

= f

[
f() = 0 for all but nitely many s. Using the Submonoid/Subgroup test,
one easily sees that

is the appropriate type of sub-object of

.
Denition. Let (S, +) be a nontrivial commutative monoid and R be a ring.
1. Dene R[X; S] =

sS
R, with the following multiplication: (fg)(s) =

s1+s2=s
f(s
1
)g(s
2
). (This is dened because it is really only a nite sum,
since there are only nitely many pairs (s
1
, s
2
) with f(s
1
) and g(s
1
) not
both zero.) Then R[X; S] is a ring (exercise), called the monoid ring of S
over R. We call the elements of R[X; S] polynomials (for lack of a better
word), but the polynomial ring over R is R[X; N].
2. We denote the element of R[X; S] with entry a in coordinate s and all
other entries zero by aX
s
. (We omit the a if a = 1). In the case a ,=
0, an element of this form is called a monomial. (The zero polynomial
is not typically considered a monomial.) This elements coecient is a
and its exponent is s. The ring R, which is sometimes called the base
ring, embeds into R[X; S] via a aX
0
(exercise). We thus make the
indicated identication and consider R to be a subring of R[X; S]. When
discussing R[X; S], we call the elements of R constants. Note that from
the denitions we have (aX
s
)(bX
t
) = abX
s+t
, and by a simple inductive
proof it follows that (aX
s
)
n
= a
n
X
ns
for n N.
3. From the denitions, we see that every nonzero polynomial has a unique
representation as a sum of monomials with distinct exponents. These
monomials are called its terms, and the coecients of a polynomial are
the coecients of its terms. The zero polynomial is considered to have
one term, namely 0, with coecient 0. The constant term of a polynomial
is its term that is a constant (if it has one), and otherwise we consider the
constant term to be 0. In practice, how one usually multiplies polynomials
is by writing them as sums of monomials, using the distributive law, and
then multiplying the monomials as above.
We will mainly be concerned with the special case R[X] = R[X; N], which
is the polynomial ring over R, but general monoid rings are very useful in some
advanced commutative ring theory research, because they allow one to construct
rings with a wide variety of properties. For example, the ring C[X;
+
0
] is one
of the simpler examples of an antimatter domain (a domain with no atoms),
and the ring R[X; Z] of Laurent polynomials has a connection to the study of
complex variables in the case R = C. (Though it is more so the ring of Laurent
power series that has the connection.) When dealing with polynomial rings, we
typically abbreviate X = X
1
.
106
Denition. Let R be a ring. If we wanted to form a polynomial ring over R[X],
we would write R[X][Y ] (instead of R[X][X]). The important thing is that
whatever symbol we are adjoining has (roughly speaking) no algebraic relation
to anything in the ring; such a thing is called an indeterminate. One could
recursively dene polynomials in multiple variables, but we will do it by a more
ecient means. For a nonempty set , we dene R[X

] = R[X;

N].
For s

N with nonzero entries m


1
, . . . , m
n
in coordinates
1
, . . . ,
n
,
respectively, we write X
m1
1
X
mn
n
instead of X
s
. If , then R[X

]
is a subring of R[X

] with the obvious identications. We abbreviate


R[X
1
, . . . , X
n
] = R[X
1
, . . . , X
n
].
Remark. Let
1
, . . . ,
n
be disjoint nonempty sets and =

n
i=1

i
. Then
R[X

] = R[X
1

11
][X
2

22
] . . . [X
n

nn
] with the obvious
identications. In particular, we have R[X
1
, . . . , X
n
] = R[X
1
][X
2
] . . . [X
n
], or we
could adjoin the indeterminates in any order and arrive at the same thing. Also,
we have R[X

] =

1,...,n
R[X
1
, . . . , X
n
]. (This union is often taken
as the denition of a polynomial ring over an arbitrary set of indeterminates.)
Denition. Let A

be a family of partially ordered sets indexed by a


nonempty well-ordered set . The lexicographic partial order on

is
the partial order given by f g f = g or f() < g() for the smallest with
f() ,= g().
Proposition 97 (Lexicographic Order). Let A

be a family of nonempty
partially ordered sets indexed by a nonempty well-ordered set . Dene the
lexicographic order on

by f g f = g or f() < g() for the


smallest with f() ,= g().
1. The lexicographic order is a partial order. It is a total order if and only if
each A

is totally ordered.
2. If the A

s are partially (resp., totally) ordered cancellative commutative


additive monoids, then so is

with respect to the lexicographic


order. (A partially ordered commutative additive monoid is dened analo-
gously to a partially ordered ring, just with the multiplicative requirement
on compatibility dropped.)
In particular, this result tells us that the lexicographic order on the exponents
in R[X

] is compatible with their addition.


Proof.
1. We rst verify that is a partial order, or, equivalently, that < is a
strict partial order. Irreexivity is clear. For the transitive property,
assume f < g and g < h. Let , , be the minimum elements with
f() ,= g(), g() ,= h(), and f() ,= h(). Then for < min(, ) we
have f() = g() = h(), so min(, ) . On the other hand, we have
f(min(, )) g(min(, )) h(min(, )), with at least one inequality
strict, so = min(, ). If = , then f() < g() h(), and if
107
= , then f() g() < h(). Either way, we conclude f < h.
Now we prove the equivalence. (): Follows immediately once we observe
that each A

is order-embedded into the Cartesian product in the obvious


way. (Fix all the other coordinates). (): Assume each A

is totally
ordered. Then for f ,= g that rst dier in coordinate , we have either
f() < g() or g() < f(), hence either f < g or g < f.
2. Assume the A

s are partially ordered cancellative commutative additive


monoids. Then

is easily seen to be cancellative. For compati-


bility, let f < g and h be elements of

, and let be the smallest


coordinate where f and g (or, equivalently, f +h and g +h) dier. Then
(f + h)() = f() + h() < g() + h() = (g + h)() by compatibility,
hence f +h < g +h.
Denition. Let R be a ring and be a nonempty set. The total degree (or
simply degree) of a monomial aX
mn
1
X
mn
n
in R[X

] is m
1
+ + m
n
,
and the total degree (or simply degree) of a nonzero polynomial f is the largest
degree of any of its terms, written deg f. The leading (resp., trailing) term of
f R[X]

is its term of largest degree (resp., smallest degree). The leading


(resp., trailing) term of 0 is 0, and deg 0 = . A polynomial in R[X

can possibly have multiple terms of maximum (resp., minimum) degree; the
term of maximum (resp., minimum) degree whose exponent is largest (resp.,
smallest) in the lexicographic order is designated as the leading (resp., trailing)
term. (Recall that exponents of monomials in R[X

] are elements of

N.) Here the lexicographic order is with respect to some well-


ordering on , or, equivalently, on the set of the variables. We make N
into a commutative monoid with absorbing element by dening n =
+ n = for all n N, and we dene < n for all n N. (Note,
however, that this does not make N into a totally ordered monoid,
because the operation is no longer cancellative. It may be a good idea to pause
for a second to think about why a totally ordered monoid must be cancellative.)
The leading (resp., trailing) coecient of a polynomial is the coecient of its
leading (resp., trailing) term.
Example. If our well-ordering of the variables is done alphabetically, then
XY
5
+X
3
Y
3
+X
2
Y
4
+X
5
+Y
5
+X
2
Y
2
+X
4
R[X, Y ] has leading term X
3
Y
3
and trailing term X
2
Y
2
. The exponents of the terms of maximum degree are
(1, 5), (3, 3), and (2, 4), with (3, 3) being the largest in the lexicographic order,
and the exponents of the terms of smallest degree are (2, 2) and (4, 0), with (2, 2)
being the smallest in the lexicographic order. Note that it is quite possible for
the term with the largest (resp., smallest) exponent in the lexicographic order
not to be the leading (resp., trailing) term, since here (5, 0) (resp., (0, 5)) is not
the leading (resp., trailing) term.
For the remainder of the section, we will be using to denote a nonempty
set.
108
Theorem 98. Let R be a ring. The following are equivalent.
1. R is a domain.
2. For f, g R[X

], the leading (resp., trailing) term of fg is the


product of the leading (resp., trailing) terms of f and g. (Here we are
dening leading/trailing terms with respect to some xed well-ordering on
.)
3. deg is a monoid epimorphism from (R[X

], ) onto N .
4. R[X

] is a domain.
Proof. For notational simplicity, we write R[X

] = R[X; S] in this proof,


where S =

N. (1) (2): Assume R is a domain. The statement is


trivially true when f = 0 or g = 0, so let f, g R[X; S]

have leading terms


aX
s
and bX
t
, respectively. Each of the terms of fg of maximum degree has an
exponent of the form s

+ t

, where s

s and t

t, and hence has exponent


at most s + t, with equality only occurring when s

= s and t

= t. (Here we
are using the compatibility of the lexicographic order on (S, +).) Because R is
a domain and a, b ,= 0, we have ab ,= 0, and this means abX
s+t
= (aX
s
)(bX
t
) is
the leading term of fg. The proof for lowest degree terms is similar. (2) (3):
First note that the two versions of (2) are equivalent, because each version
implies that R is a domain, hence implies the other version by (1) (2). (To
see this, note that if the leading (resp., trailing) version of (2) holds, then for
a, b R

and s S
+
the leading (resp., trailing) term of (aX
s
+1)b = abX
s
+b
(resp., (X
s
+ a)b = bX
s
+ ab) is abX
s
(resp., ab), hence ab ,= 0.) Now assume
(2) holds. Clearly deg is surjective, and by denition deg 1 = 0. Certainly the
identity deg fg = deg f + deg g holds for monomials whose coecients do not
multiply to be zero, and it also holds rather trivially if f = 0 or g = 0. It
then follows from (2) that the identity holds for all f, g R[X; S], so deg is an
epimorphism. (3) (4): Assume (3). Then fg = 0 deg f +deg g = deg fg =
deg f = or deg g = f = 0 or g = 0. Therefore R[X; S] is a
domain. (4) (1): Any subring of a domain is a domain.
Corollary 99. If D is a domain, then D[X

= D

.
Proof. Exercise.
Example. As we have previously seen, this is false for non-domains, because
([2]X + 1)
2
= [4]X
2
+ [4]X + [1] = [1] in Z
4
[X].
Theorem 100. Let I be an ideal of a ring R. Then the ideal of R[X

] gen-
erated by I is I[X

] = polynomials in R[X

] with coecients in I.
If I is proper, then R[X

]/I[X

= (R/I)[X

]. In particular,
the ideal I[X

] is prime if and only if I is, and an element of R is prime


in R if and only if it is prime in R[X

].
Proof. Exercise.
109
The following proposition simplies the process of verifying that a map from
a monoid ring is a homomorphism.
Proposition 101. Let R and T be rings and S be a nontrivial commutative
monoid. A map : R[X; S] T is a homomorphism if and only if (i) (1) = 1,
(ii) (f + g) = (f) + (g) for all f, g R[X; S], and (iii) (fg) = (f)(g)
for all monomials f and g.
Proof. (): Clear. (): Assume that satises the given proper. Then is a
group homomorphism between (R[X; S], +) and (T, +), so (0) = 0. Hence for
f R[X; S] we have (f 0) = (0) = 0 = (f)(0). So all that remains is to
show that (fg) = (f)(g) for f, g R[X; S]

. We proceed by induction on
m+n, where m (resp., n) is the number of terms in f (resp., g), and without loss
of generality we take m n. The base case m+n = 2 is simply property (iii), so
let us assume m+n > 2, and consequently f = f
1
+f
2
, where f
1
, f
2
R[X; S]

have less terms than f. So by induction we have (fg) = (f


1
g + f
2
g) =
(f
1
g) + (f
2
g) = (f
1
)(g) + (f
2
)(g) = ((f
1
) + (f
2
))(g) = (f)(g), as
desired.
Denition. Let R be a ring, f R[X
1
, . . . , X
n
], and a
1
, . . . , a
n
R. From now
on, we will use f(a
1
, . . . , a
n
) to denote the result of evaluating f at the point
(a
1
, . . . , a
n
), i.e., replacing all instances of X
i
with a
i
. (We will no longer use
f(a
1
, . . . , a
n
) to denote the coecient of X
a1
1
X
an
n
, since we are thinking
of polynomials as their usual representations now instead of as elements of

n
k=1
N
R.) The map f f(a
1
, . . . , a
n
) is an epimorphism of R[X
1
, . . . , X
n
]
onto R. (This is very easily veried by using the above proposition.) These
comments hold for R[X

] as well, although in that case we do not have a


convenient notation like f(a
1
, . . . , a
n
), but would have to write f(a

) to
denote the result of replacing all instances of X

with a

. For convenience, we
will write f(

0) to denote f(0

). Note that f(

0) is simply the constant term


of f.
Denition.
1. A ring R is a Euclidean ring if there is a : R ON such that for every
a R and b R

there are q, r R with a = bq + r and (r) < (b).


In this case, we call an algorithm for R. (There are multiple slightly
dierent ways to dene Euclidean ring. The old-fashioned way to dene
them is with N in place of ON, which is actually not equivalent to how we
are dening them. But eventually people realized that this more inclusive
way of dening Euclidean rings is arguably better.) Note that necessarily
(0) < (x) for all x R

. (Exercise.) A Euclidean domain is simply a


domain that is a Euclidean ring. An example of a Euclidean domain is Z.
(Exercise.)
2. A principal ideal ring is a ring in which every ideal is principal. Obviously
PIDs are PIRs. Every Euclidean ring is a PIR. (Exercise.)
110
Theorem 102 (Division Algorithm for Polynomials). Let R be a ring, f
R[X], and g R[X]

be an element with unit leading coecient. Then there


are q, r R[X] with f = gq + r and deg r < deg g. If R is a domain, then q
and r are unique.
Proof. Let aX
m
and bX
n
be the leading terms of f and g, respectively. The
equation f = g 0 + f covers the case deg f < deg g, so let us assume deg f
deg g. Then deg(f ab
1
X
mn
g) < deg f, hence by induction we have f
ab
1
X
mn
g = gq + r for some q, r R[X] with deg r < deg g. Thus f =
g(q +ab
1
X
mn
) +r.
For uniqueness, assume R is a domain and gq+r = gq

+r

with deg r, deg r

<
deg g. Then g(q q

) = r

r, so deg g + deg(q q

) = deg(r

r). Since
deg(r

r) < deg g, this implies deg(q q

) = = deg(r

r), hence q = q

and r = r

.
Example. The theorem would be false if we did not require the leading co-
ecient of g to be a unit. For example, in Z[X], there is no way to write
X = 2Xq +r with r Z.
Corollary 103 (Remainder Theorem/Factor Theorem). Let R be a ring, f
R[X], and a R. Then f = (X a)q + f(a) for some q R[X]. Therefore
f(a) = 0 if and only if X a [ f.
Proof. From the Division Algorithm, we have f = (X a)q +r for some q, r
R[X] with deg r < 1. Thus r 1, and evaluating both sides of the equation at
a, we obtain f(a) = (a a)q(a) +r = r.
Corollary 104 (Polynomial Ring over a Field is a Euclidean Domain). Let R
be a ring. Then R[X] is a Euclidean domain R[X] is a PID R is a eld.
Proof. Label the statements (1)-(3), respectively. (1) (2): Every Euclidean
domain is a PID. (2) (3): Assume R[X] is a PID and let a R

. Then
(a, X) = (b) for some b R[X]. From degree considerations, we see that b
is constant, and it divides 1 (since that is the coecient of X), hence is a
unit. Thus (a, X) = R[X] and af + Xg = 1 for some f, g R[X]. Hence
af(0) = af(0) + 0 g(0) = 1 and a is a unit. (3) (1): Dene : K[X] ON
by (0) = 0 and (f) = degf + 1 for f ,= 0. By the Division Algorithm, for
f R[X] and g R[X]

, there are unique q, r R[X] with f = gq + r and


deg r < deg g. If r = 0, then (r) = 0 < 0 +1 deg g +1 = (g), and otherwise
(r) = deg r + 1 < deg g + 1 = (g).
Exercises.
1. Let (S, +) be a nontrivial commutative monoid and R be a ring. In this
problem, use only the denition of R[X; S] without any of the later prop-
erties we developed.
(a) Prove that R[X; S] is a ring. (Hint: You already know it is an abelian
group, so all that needs to be veried is that it is a commutative
monoid under and that the distributive law holds.)
111
(b) Prove that R embeds in R[X; S] via a aX
0
.
2. Let D be a domain. Show that D[X

= D

. (Hint: Use the fact


that deg(fg) = deg f + deg g.)
3. Let R be a ring. Prove that R[X]/(X)

= R. (Note: For future reference,


essentially the same proof shows that R[X

]/(X

= R.)
4. Let I be an ideal of a ring R. (Hint: In this problem it may be notation-
ally easier at times to use the notation of R[X;

N] rather than the


customary multiple-variable notation.)
(a) Show that the ideal of R[X

] generated by I is I[X

] =
polynomials in R[X

] with coecients in I. (Hint: You have


one containment by the closure properties of ideals, and the other
containment once you show that I[X

] is an ideal. If you show


that it is a kernel in part (b), you can cite that.)
(b) Prove that, if I ,= R, then R[X

]/I[X

= (R/I)[X

].
(Note that this shows that an element of R is prime in R if and only
if it is prime in R[X

]. Hint: Find a homomorphism with the


appropriate image and kernel.)
5. (a) Show that a Euclidean ring is a PIR. (Hint: Let I be a nonzero ideal
of a Euclidean ring R with algorithm . Show that I = (b), where
b I

with (b) minimum.)


(b) Show that Z is a Euclidean domain. (Hint: Use the absolute value
function as your algorithm. See if you can mimic the Euclidean
Domain proof in this section.)
(c) [Extra credit] Show that, if R is a Euclidean ring with algorithm ,
then (0) < (x) for all x R

.
(d) [Extra credit] Let b 2. Show that every natural number has a
unique base b representation. (Hint: Use induction for existence.
For uniqueness, show that the coecient of b
0
is the same in any
two representations [you can use the following part], and then use
induction.)
(e) [Extra credit] (Division Algorithm for Z) Let a Z and b Z

. Show
that there are unique q, r N with a = bq +r and 0 r < b. (Hint:
For existence, you already have a = bq + r with 0 [r[ < b by part
(b). In the case r < 0, consider the equation a = b(q1)+(b+r). For
uniqueness, let a = bq

+r

be another such representation and try to


deduce that r = r

and q = q

from the equation [r r

[ = b[q q

[.)
6. [Extra credit] Let R be a ring. Dene R[[X]] =

nN
R, denote the
element (a
0
, a
1
, a
2
, . . .) by

n=0
a
n
X
n
, and redene the multiplication by
(

m=0
a
m
X
m
)(

n=0
b
n
X
n
) =

k=0
(

m+n=k
a
m
b
n
)X
k
. We call R[[X]]
the power series ring over R, and call the elements of R[[X]] power series.
112
For f R[[X]], dene f(0) to be its constant term (terms are dened
analogously to the polynomial case). Note that, once we prove R[[X]] to
be a ring, it is immediate that it has R[X] as a subring.
(a) Prove that R[[X]] is a ring. (Hint: This is extremely similar to one
of the other exercises. Just an outline of what needs to be shown and
a reference to the corresponding similar proof is enough.)
(b) Prove that the map f f(0) is an epimorphism of R[[X]] onto R.
(c) Prove that R[[X]]/(X)

= R. (Hint: Use part (b).)


(d) Prove that R[[X]] is a domain if and only if R is. (Hint: If R is a
domain and f, g R[[X]]

, show that the lowest degree term of fg


is nonzero.)
(e) Prove that, if R is a domain, then R[[X]]

= f R[[X]] [ f(0)
R

. (Hint: That f(0) R

for f R[[X]]

follows from part (b).


For the other inclusion, let

n=0
a
n
X
n
R[[X]] with a
0
R

, let
g =

n=0
b
n
X
n
, write out what needs to happen for fg = 1, and
see if you can recursively construct the sequence b
n

n=0
that makes
this happen. This is a fun exercise so I will not spoil it further.)
(f) Show that, if R is a eld, then R[[X]] is a Euclidean domain. (Hint:
Use the previous part to show that every nonzero power series is
associate to some power of X. Then pick the algorithm accordingly.)
(g) Show that, if R is a eld, then R[[X]] is a valuation domain, i.e., that
for f, g R[[X]], we have f [ g or g [ f. (Hint: This follows from the
observation in the previous hint.)
3.6 The Ascending Chain Condition
[Mostly not in Durbin.]
In some ways, the next best thing to a PIR is a ring in which every ideal
is nitely generated. Emily Noether was the rst to point out the relationship
between these rings and the ascending chain condition, so the rings are named
in her honor. They are without question one of the biggest and most-studied
topics in commutative ring theory.
Denition. A partially ordered set satises the ascending chain condition
(ACC) if it has no strictly increasing innite sequence. A partially ordered
set satisfying the ACC is called Noetherian.
Theorem 105 (Characterization of Noetherian Posets). Let following are equiv-
alent for a partially ordered set A.
1. A is Noetherian.
2. Every non-decreasing innite sequence in A is eventually constant. (In
other words, whenever a
1
a
2
, there is an N Z
+
with a
n
= a
N
for all n N.)
113
3. Every nonempty subset of A has a maximal element.
4. Every nonempty chain in A has a maximum element.
Proof. (1) (3): By contrapositive. Assume there is a nonempty B A with
no maximal element. We show A is non-Noetherian by recursively constructing
a strictly increasing sequence in B as follows: let b
1
B, and for n > 1, we
know b
n1
is not a maximal element of B, so there is some b
n
B greater than
b
n1
. (3) (4): Immediate once we recall that maximal and maximum are
equivalent in totally ordered sets. (4) (2): Assume every nonempty chain in
A has a maximum element. If a
1
a
2
, then there is some maximum a
N
,
and necessarily a
n
= a
N
for n N. (2) (1): Clear.
Denition. We say a ring is Noetherian if its set of ideals is Noetherian. (This
set is partially ordered by inclusion.)
Theorem 106 (Characterization of Noetherian Rings). The following are equiv-
alent for a ring R.
1. R is Noetherian.
2. R satises the ACC on nitely generated ideals.
3. Every ideal of R is nitely generated.
Proof. (1) (2): Clear. (2) (3): Assume R satises the ACC on nitely
generated ideals and let I be an ideal of R. Then the set of nitely generated
ideals contained in I is nonempty (since (0) is in it), so it has a maximal element
J. If there is an a I J, then J + (a) is a larger nitely generated ideal
contained in I than J, a contradiction. Therefore I = J. (3) (1): Assume
every ideal of R is nitely generated, let I
1
I
2
be an ascending sequence
of ideals, and let I =

n=1
I
n
. (Recall that I is an ideal by a past exercise.)
Write I = (a
1
, . . . , a
m
), and for each i pick k
i
with a
i
I
ki
. Without loss of
generality, let us say k
m
is the largest k
i
. Then I
km
contains each I
ki
, hence
each a
i
, hence I = I
km
. Consequently I
n
= I
ki
for all n k
i
.
Example. We discuss a few examples of Noetherian rings and ways to construct
new Noetherian rings from old ones.
1. Every PIR is Noetherian. In particular, this includes Z, Z[i], every eld
(such as , 1, and C), and every polynomial ring K[X] with K a eld.
This last example is a special case of the famous Hilbert Basis Theorem
below. We will prove that Z[i] is a Euclidean Domain (hence a PID) later.
2. If R is a Noetherian ring, then so is R/I. (Exercise.) In particular, this
includes each Z
n
.
3. Every localization of a Noetherian ring is Noetherian. (Exercise.) Thus
for each prime p the ring Z
(p)
is a local ring, which we dene to be a
Noetherian quasilocal ring. (Refer to the end of the localization section if
you do not recognize what this ring is.)
114
4. A nite direct product

n
i=1
R
i
of rings is Noetherian if and only if each
R
i
is Noetherian. (Exercise.)
Theorem 107 (Hilbert Basis Theorem). If R is a Noetherian ring and n Z
+
,
then R[X
1
, . . . , X
n
] is Noetherian.
Proof. Assume R is Noetherian. By induction, we reduce the result to proving
that R[X] is Noetherian, which amounts to proving that every nonzero ideal J
of R[X] is nitely generated. For each k N, let I
k
be the ideal of R consisting
of the leading coecients of the polynomials in J of degree at most k. (This is
easily veried to be an ideal with the Ideal Test.) Then I
0
I
1
I
2
and
the sequence eventually stabilizes, say at index N. Each I
i
is nitely generated,
say I
i
= (a
i,1
, . . . , a
i,ni
), where n
i
0 and each a
i,j
,= 0. For each i, j, let f
i,j
be
a polynomial in J with degree i and leading coecient a
i,j
. (We can always nd
one of degree i by multiplying by an appropriate power of X if necessary.) We
claim that J = (A), where A = f
0,1
, . . . , f
0,n0
, . . . , f
N,1
, . . . , f
N,n
N
. Certainly
(A) J. Now let g J, and let a be its leading coecient. If g = 0, then g
(A), so let us assume deg g 0. Let d = min(deg g, N). Then a I
d
, so we may
write a =

n
d
k=1
c
k
a
d,k
for some c
1
, . . . , c
n
d
R, and g X
deg gd

n
d
k=1
c
k
f
d,k
is a polynomial in J of lesser degree than g, which by induction is in (A), so
g = (g X
deg gd

n
d
k=1
c
k
f
d,k
) +X
deg gd

n
d
k=1
c
k
f
d,k
(A), as desired. (Note
that we can use induction since the set N of possible degrees is well-
ordered.)
Remark.
1. The analogous result for power series rings is also true, but we will not
be proving it, since its proof is a bit more intricate and we are mainly
concerned with polynomial rings anyway.
2. The Hilbert Basis Theorem is not true in innitely many variables. Indeed,
if R is any ring, then R[X
1
, X
2
, . . .] is non-Noetherian. (Exercise.)
3. The converse to the Hilbert Basis Theorem is true. This is merely a special
case of the more general result: If I is a proper ideal of a Noetherian ring,
then R/I is Noetherian. (You should pause for a second to recall why this
implies the converse to the Hilbert Basis Theorem. This is referencing an
exercise from the previous section.)
When one studies factorization theory, it is the following weaker version of
the Noetherian property that is of interest.
Denition. A ring satises the ascending chain condition on principal ideals
(ACCP)) if its set of principal ideals is Noetherian. In other words, for any
innite sequence a
n

n=1
in the ring with each a
n+1
[ a
n
, there is an N Z
+
with (a
n
) = (a
N
) for all n N.
Example. Of course, every Noetherian ring satises the ACCP. However, the
converse is false. For example, we will shortly see that, if D is any domain
satisfying ACCP, then D[X
1
, X
2
, . . .] satises ACCP but is not Noetherian.
115
Proposition 108. Let R be a domain and D be a subring of R satisfying
D R

= D

. (In particular this holds if D

= R

.) If R satises ACCP,
then so does D.
Proof. Exercise.
Theorem 109. Let D be a domain and be a nonempty set. Then D satises
the ACCP if and only if D[X

] does.
Proof. (): Since D[X

= D

, this is immediate from the above propo-


sition. In fact, this direction is true even if D is not a domain, with the following
alternate proof. It suces to observe that two elements of D are associates in D
if and only if they are associates in D[X

]. To see this, note that if af = b


and bg = a for a, b D and f, g D[X

], then af(

0) = b and bg(

0) = a.
(): Assume D satises the ACCP and let (f
1
) (f
2
) be an ascending
sequence of principal ideals in D[X

], say f
n
= f
n+1
g
n
. It will suce to
consider the case where all the f
n
s have the same degree, because if f
m
has
minimum degree among the f
n
s, then by the additive degree formula (which
holds since D is a domain) we have deg f
n
= deg f
m
for n m. Also, we may
assume that all the f
n
s are nonzero, because if any f
n
is nonzero, then all the
remaining ones must be too. From the equation deg f
n
= deg f
n+1
+deg g
n
, we
conclude each deg g
n
= 0, or in other words each g
n
D

. Let a
n
denote the
leading coecient of f
n
. Then each a
n
= a
n+1
g
n
(again using the fact that D
is a domain), so we have an ascending sequence (a
1
) (a
2
) of principal
ideals in D, which eventually stabilizes, say at index N. Because associates in a
domain are unit multiples, for n N we have g
n
=
an+1
an
D

= D[X

.
Therefore (f
N
) = (f
N+1
) = , as desired.
Remark.
1. In contrast to the Hilbert Basis Theorem, the direction is not true
without the domain hypothesis. (Examples showing this are slightly com-
plicated.)
2. The alternate version of this theorem for power series is also true. (Exer-
cise.)
Denition. A ring is atomic if every nonzero nonunit has an atomic factor-
ization, meaning that every nonzero nonunit can be written as a product of
atoms. (This includes trivial products with one factor, so an atom is considered
a product of atoms.) Note that 0 is also a product of atoms in an atomic ring.
(If the ring is a domain, then 0 is prime, hence irreducible, while if the ring is
not a domain, then 0 is a product of two nonzero nonunits.)
Theorem 110. Every ring satisfying the ACCP is atomic.
Proof. Suppose to the contrary that there is a non-atomic ring satisfying the
ACCP. Then it has some nonzero proper principal ideal with a generator that
is not a product of atoms, hence by ACCP a maximal such principal ideal I.
116
Let I = (a), where a is a nonzero nonunit that is not a product of atoms. Then
a is not irreducible, so we have a = bc for some b and c with (a) _ (b), (c), and
by maximality b and c are products of atoms (note that (b) and (c) are proper
since otherwise a would be an associate to b or c), hence a is a product of atoms,
a contradiction.
Example. There are atomic domains that do not satisfy ACCP, but they are
quite strange and complicated.
Denition. Let D be a domain. We use D
#
to denote its set of nonzero
nonunits, i.e., D
#
= D

. Note that D
#
is multiplicatively closed (in a
domain a product of nonzero elements is nonzero, and in any monoid a product
of two nonunits is a nonunit).
1. A factorization of x D
#
is a product x = a
1
a
n
with each a
i
D
#
.
We call n the length of the factorization, and the a
i
s are the factors in
the factorization. An atomic factorization is a factorization whose factors
are atoms. Recall that in a domain a nonunit is irreducible if and only
if it cannot be written as a product of two nonzero nonunits. Calling a
factorization of length 1 trivial, we thus see that a nonzero nonunit is an
atom if and only if it has no nontrivial factorizations. Thus an atomic
factorization is a complete factorization in the sense that it cannot be
factored any further.
2. A unique factorization domain (UFD) is a domain in which every nonzero
nonunit has a unique atomic factorization up to order and associates.
The up to order and associates part means that, whenever a
1
a
m
=
b
1
b
n
are products of nonzero atoms, then m = n and each (a
i
) = (b
i
)
after a suitable reordering. (This last part could be stated more precisely
as there is some S
m
with each (a
i
) = (b

(i)) but in practice the


more informal, less cumbersome, description is used.) We will later give
several equivalent characterizations of UFDs. For now, we remark that a
domain is a UFD if and only if it is atomic and every atom is prime. (You
are allowed to use this result without proof if you need it. We will prove
it later.)
UFDs are the nicest kind of domain with respect to factorization properties,
because they are the domains where you can completely factor any given nonzero
nonunit and there is (up to order and associates) a unique way to do so. High-
school type problems where you are asked to completely factor an integer or
polynomial rely on the fact that the domain in which you are working is a UFD.
Example.
1. We will later see that any PID (such as Z, Z[i], 1[X], C, etc.) is a UFD.
2. If D is a UFD, then D[X
1
, X
2
, . . .] is a non-Noetherian UFD. (We will
later prove that any polynomial ring in arbitrarily many variables over a
UFD is a UFD. However, interestingly, the analogous statement for power
117
series is not true.) However, every UFD does satisfy the ACCP, as we will
see below.
3. Also, a Noetherian domain need not be a UFD. Recall the domain 1 +
XC[X] discussed in the examples in Section 3.2. Its element X is irre-
ducible but not prime, so 1 + XC[X] is not a UFD. However, it can be
shown to be Noetherian. (But doing so is best done using module theory,
and we will not be going into module theory beyond the most basic of
denitions in this course.)
4. The polynomial f = 3X
2
+6X+3 has unique (up to order and associates)
atomic factorization 3(X + 1)
2
in Z[X]. (It has other possible atomic
factorizations in Z[X] such as (X + 1)(3)(X 1), but these other
factorizations are the same up to order and associates as the given one.
Here 3 and 3 are associates, and X 1 and X + 1 are associates.) In
[X], the unique (up to order and associates) atomic factorization of f is
(3(X + 1))(X + 1). Note that 3(X + 1) is irreducible in [X] but not in
Z[X]. (The dierence is that nonzero constants are units in [X], so by
degree considerations any degree 1 polynomial must be irreducible.) We
will talk about UFDs and related topics in much more detail later.
Example. We have already remarked that a Noetherian domain need not be
a UFD, but did not provide a proof. The main thing we are interested in is
the weaker statement that atomic factorizations in an atomic domain need not
have the desired uniqueness. For this, it suces to give an example of a domain
satisfying the ACCP that is not a UFD. We give two such examples here.
1. Let R = 1 + XC[X], i.e., R is the subring of C[X] consisting of the
polynomials with constant term in 1. In the examples in Section 3.2 we
showed that X is an irreducible element of R that is not prime, so R is
not a UFD.
2. Let K be a eld and R = K[X
2
, X
3
], which denotes the subring of K[X]
consisting of all polynomials with no degree 1 term. In the examples in
Section 3.2 we showed that X
2
and X
3
are non-prime atoms, so R is not
a UFD.
The fact that the domains in (1) and (2) satisfy the ACCP is a special case of
one of the exercises: if D is a domain satisfying the ACCP and ,= , then any
ring between D and D[X

satises the ACCP.


Remark. Note that, if a nonzero nonunit in a domain has a factorization of
maximum length, then that factorization is necessarily atomic, since otherwise
we could replace one of the factors with a nontrivial factorization and obtain a
longer factorization. A similar argument shows that in an atomic domain, given
a factorization of a nonzero nonunit, one can nd an atomic factorization that
is at least as long. Thus, in a UFD every nonzero nonunit has a nite upper
bound on the lengths of its factorizations, namely the length of its unique (up
to order and associates) atomic factorization.
118
Theorem 111. Every UFD satises the ACCP.
Proof. By contrapositive. Let D be a domain with a properly ascending se-
quence (a
1
) _ (a
2
) _ of principal ideals, where we may assume that each
a
i
is a nonzero nonunit. (Only a
1
could possibly be zero, and none of the a
i
s
can be units or else the sequence would have stabilized.) By an exercise, this
means that each a
n
= a
n+1
b
n
for some nonzero nonunit b
n
. Repeated substi-
tutions yield a
1
= a
2
b
1
= a
3
b
2
b
1
= a
4
b
3
b
2
b
1
= , so a
1
has arbitrarily long
factorizations. Therefore D is not a UFD.
We have the following diagram of implications between some of the kinds of
domains we have discussed in this chapter.
Euclidean Domain
//
PID
//
&&
UFD
//
ACCP
//
atomic
Noetherian
88
(Again, we have not yet proven PID UFD, but we will later.) The exam-
ples in these notes and the exercises show that there are no further nontrivial
implications, i.e., none of the implications reverse, neither UFD nor Noetherian
implies the other.
Exercises.
1. Let R be a ring.
(a) Prove that, if R is Noetherian, then so is each homomorphic image
of R. (Hint: Note that by the fundamental homomorphism theorem
this is the same as showing that if R is Noetherian, then so is R/I for
each proper ideal I. Show this using the Correspondence Theorem.)
(b) Show that, if R[X

] is Noetherian, then so is R. (Hint: Use


part (a).)
(c) Show that R[X
1
, X
2
, . . .] is not Noetherian. (Hint: Show that (X
1
) _
(X
1
, X
2
) _ (X
1
, X
2
, X
3
) _ . To show that these inclusions are
proper, derive a contradiction from X
n+1
= X
1
f
1
+ + X
n
f
n
by
evaluating both sides at an appropriate point.)
(d) Give an example of a non-Noetherian domain with a Noetherian ho-
momorphic image.
(e) Show that every localization of a Noetherian domain is Noetherian.
(Hint: Review the results about the ideals of localizations.)
(f) Show that every domain has a Noetherian localization. (This means
that the localization being Noetherian does not imply that the origi-
nal domain is Noetherian. Hint: Think of the most basic example of
localization.)
119
2. (a) Let D be a domain and a, b D

. Show that a is a proper divisor of


b if and only if b = ax for some nonzero nonunit x. (Hint: Reference
the past exercise on associates in domains.)
(b) Let R be a domain and D be a subring of R satisfying D

= R

D. Prove that D satises the ACCP if R does. (Hint: Proceed by


contrapositive. If Da
1
_ Da
2
_ Da
3
_ is a properly ascending
sequence of principal ideals in D, then it follows from (a) that Ra
1
_
Ra
2
_ Ra
3
_ is a properly ascending sequence of principal ideals
in R.)
3. Let D be a domain satisfying ACCP and be a nonempty set. Prove
that every ring between D and D[X

] satises ACCP. (Hint: Let


R be such a ring, show that R

= D[X

= D

, and nd an
appropriate result from this section to cite.)
4. Give a specic example of a UFD that is not a PID. Prove that your
example is such a thing. (Do not use any facts about UFDs that we have
not proven except for the following: if D is a UFD, then so is D[X]. Hint:
Briey note why a eld is a UFD. Then pick a eld and use the previous
fact [possibly more than once] to construct a UFD that is not a PID.
Refer to the previous section for a result about when polynomial rings are
PIDs.)
5. [Extra Credit] Let R

be a nonempty family of rings and R =

.
(a) Show that, if each I

is an ideal of R

, then

is an ideal of
R. (Hint: Use the Ideal Test.)
(b) Prove that, if is nite, then every ideal is of the form in (a). (Hint:
For notational convenience, write = 1, . . . , n, so R =

n
i=1
R
i
and the elements of R can be written as n-tuples. Let I be an ideal
of R. Note why each
i
(I) is an ideal of R
i
, where
i
: R R
i
:
(x
1
, . . . , x
n
) x
i
. Show that I =

n
i=1

i
(I). For the inclusion,
if a
i

i
(I), then show that the element f
i
R with a
i
in the
ith coordinate and zeros elsewhere is in I, and thus (a
1
, . . . , a
n
) =
f
1
+ +f
n
I.)
(c) Prove that R is Noetherian if and only if is nite and each R

is
Noetherian. (Hint: For , proceed by contrapositive, using part
(a) to give innite properly ascending chains of ideals in both cases
where is innite or some R

is non-Noetherian. For , use part


(b) show that every ideal of that form is nitely generated.)
6. [Extra Credit] Show that a domain D satises the ACCP if and only if
D[[X]] does. (Note that by induction this can be generalized from D[[X]]
to D[[X
1
, . . . , X
n
]] = D[[X
1
]][[X
2
]] [[X
n
]]. The version with innitely
many variables is true as well, but there are at least four non-equivalent
ways to dene power series in innitely many variables, so we will not be
120
getting into this. Hint: Mimic the proof of the polynomial version, but
use trailing terms in place of leading terms.)
3.7 Divisibility
[Durbin: Section 12]
Denition. Let D be an integral domain and _ A D.
1. A common divisor of A is an element of D that divides each element
of A; the set of common divisors of A is denoted CD(A). Note that
necessarily D

CD(A) D

. A greatest common divisor (GCD) of A


is a common divisor that every other common divisor divides; the set of
greatest common divisors of A is denoted GCD(A).
2. A common multiple of A is an element of D that every element of A
divides; the set of common multiples of A is denoted CM(A). A least
common multiple (LCM) of A is a common multiple that divides every
other common multiple; the set of least common multiples of A is denoted
LCM(A).
3. We use our usual abbreviations for the case when Ais nite, e.g., GCD(a
1
, . . . , a
n
) =
GCD(a
1
, . . . , a
n
), etc. We say a, b D

are relatively prime if CD(a, b) =


D

, or, equivalently, GCD(a, b) = D

.
4. If there is more than one domain that we are talking about, then it may be
ambiguous which domain we are taking these denitions with respect to,
so to avoid this potential ambiguity we will use subscripts when necessary.
For example, GCD
Z
(2, 4) = 2, 2 and GCD
Q
(2, 4) =

.
5. We say D is a GCD domain if GCD(a, b) ,= for each a, b D

.
Remark.
1. For a, b, c D

, we have a [ b ac [ bc. To see this, note that the


cancellative property shows that b = ax bc = (ac)x.
2. Without being given any more information, there is no guarantee that a
GCD or LCM of A exists. However, it is immediate from the denitions
that GCDs (resp., LCMs) of A are unique up to associates when they
exist, i.e., if a and b are GCDs (resp., LCMs) of A, then (a) = (b).
3. If x is in CD(A), GCD(A), CM(A), or LCM(A), then so is any associate
of x.
4. Every singleton subset of D

has a GCD and LCM. More specically, for


a D

we have GCD(a) = LCM(a) = aD

. More generally, if A D

and a A CD(A) (resp., a A CM(A)), then GCD(A) = LCM(A) =


aD

. (These observations follow directly from the denitions and the fact
that associates in integral domains are unit multiples of each other.)
121
Example.
1. We will soon see that every PID is a GCD domain. In the next section,
we will generalize this by showing that a domain is a UFD if and only if
it is an atomic GCD domain.
2. 1 + XC[X] is not a GCD domain, because there is no GCD of X
2
and
iX
2
. (Exercise.)
Theorem 112 (Properties of GCDs and LCMs). Let D be a domain and
_ A D

.
1. CM(A) =

aA
(a).
2. LCM(A) = m D [ (m) =

aA
(a).
3. CD(A) = d D [ (A) (d).
4. d GCD(A) if and only if (d) is minimum among the principal ideals
containing (A).
5. If x D

and GCD(xA) ,= , then GCD(A) ,= and GCD(xA) =


xGCD(A).
6. If d GCD(A), then GCD(d
1
A) = D

.
Proof. Exercise.
Using (2) and (4) of the theorem above, we immediately have the following.
Corollary 113. If D is a PID, then every nonempty subset of D

has a GCD
and LCM.
Proposition 114. Let D be a domain, A

be a nonempty family of
nonempty subsets of D

, and A =

.
1. If each m

LCM(A

), then LCM(A) = LCM(m

).
2. If each d

GCD(A

), then GCD(A) = GCD(d

).
Proof.
1. Assume each m

LCM(A

). From Theorem 112(2) and the equal-


ity

aA
(a) =

aA

(a) =

(m

), we see that LCM(A) =


LCM(m

).
2. Assume each d

GCD(A

). Then each (d

) is minimum among the


principal ideals containing (A

), so a principal ideal is minimum among


the principal ideals containing (A) if and only if it is minimum among the
principal ideals containing (d

). From Theorem 112(4) we conclude


that GCD(A) = GCD(d

).
122
Recall the formula gcd(a, b)lcm(a, b) = ab we have been using for a, b Z
+
.
The following is a generalization.
Theorem 115. Let D be a domain.
1.
ab
m
GCD(a, b) for each a, b D

and m LCM(a, b).


2. If D is a GCD domain, then
ab
d
LCM(a, b) for each a, b D

and
d GCD(a, b).
Proof.
1. Let a, b D

and m LCM(a, b). Then


ab
m

m
b
= a and
ab
m

m
a
= b, so
ab
m
CD(a, b). For x CD(a, b), we have
ab
x
= a
b
x
= b
a
x
CM(a, b), so
ab
x
[ m, hence
ab
m
[ x. Therefore
ab
m
GCD(a, b).
2. Assume D is a GCD domain, a, b D

, and d GCD(a, b). Then


ab
d
=
a
b
d
= b
a
d
CM(a, b). Now let x CM(a, b), say x = ay = bz. Then
xa = (ab)z and xb = (ab)y, so ab CD(xa, xb). Since xd GCD(xa, xb),
we have ab [ xd, hence
ab
d
[ x, as desired.
In particular, the above theorem shows that if two nonzero elements have an
LCM, then they have a GCD. However, the other way around is not necessarily
true (if our domain is not a GCD domain).
Example. Let D = 1+XC[X]. Then GCD(X, iX) = R

but LCM(X, iX) = .


(Exercise.)
Theorem 116 (Characterization of GCD Domains). The following are equiv-
alent for a domain D.
1. D is a GCD domain.
2. Every nonempty nite subset of D

has a GCD.
3. Every 2-element subset of D

has an LCM.
4. Every nonempty nite subset of D

has an LCM.
5. For each a, b D

, there is a minimum principal ideal among those con-


taining (a, b).
6. For each nonzero nitely generated ideal I, there is a minimum principal
ideal among those containing I.
7. Every intersection of two principal ideals is principal.
8. Every intersection of nitely many principal ideals is principal.
123
Proof. Theorem 112 shows that (1) (resp., (2),(3),(4)) is equivalent to (5) (resp.,
(6),(7),(8)), so it will suce to show the equivalence of (1)-(4). (1) (2):
Assume D is a GCD domain, and let a
1
, . . . , a
n
D

. If n 2, then cer-
tainly GCD(a
1
, . . . , a
n
) ,= , so let us assume n 3. By induction there is a
d GCD(a
2
, . . . , a
n
), hence GCD(a
1
, . . . , a
n
) = GCD(a
1
, d) ,= . (2) (3):
Theorem 115. (3) (4): Similar to the proof of (1) (2). (4) (1):
Theorem 115.
Theorem 117 (More Properties of GCD Domains). Let D be a GCD domain.
1. If GCD(a, b) = GCD(a, c) = D

, then GCD(a, bc) = D

.
2. If a [ bc and GCD(a, c) = D

, then a [ b.
3. If a [ b
1
b
n
, then we may write a = b

1
b

n
, where each b

i
[ b
i
.
4. An element of D is irreducible if and only if it is prime.
Proof.
1. Assume GCD(a, b) = GCD(a, c) = D

, and let d CD(a, bc). Then d


CD(ac, bc) and GCD(ac, bc) = cGCD(a, b) = cD

, so d CD(a, c) = D

,
as desired.
2. Assume a [ bc and GCD(a, c) = D

. Then a CD(ba, bc) and b bD

=
bGCD(a, c) = GCD(ba, bc), so a [ b.
3. Assume a [ b
1
b
n
. The case n = 1 is trivial, so let us assume n 2.
Let b

1
GCD(a, b
1
), and note that b

1
[ b
1
. Then
a
b

1
[
b1
b

1
(b
2
b
n
) and
GCD(
a
b

1
,
b1
b

1
) = D

, so
a
b

1
[ b
2
b
n
by (2). By induction, we have
a
b

1
=
b

2
b

n
, where each b

i
[ b
i
. Writing a = b

1
b

n
completes the proof.
4. We already know that prime implies irreducible in any ring. Now let a be
any irreducible element of D. Because 0 is prime in an integral domain,
we may assume a D
#
. If a [ b
1
b
2
, then a = b

1
b

2
where each b

i
is a
nonzero element dividing b
i
, and by irreducibility one of the b

i
s, say b

1
,
is a unit, and (a) = (b

2
) (b
2
). Therefore a is prime.
Denition. Let D be a GCD domain. The content of f D[X]

, denoted
c(f), is the set of GCDs of fs coecients. A polynomial in D[X] is primitive
if its content is D

, i.e., if its coecients have no nonunit common divisor.


Remark. Let D be a GCD domain and K = QF(D).
1. Every polynomial in K[X]

is an associate of a primitive polynomial in


D[X]. To see this, let f K[X]

, write the coecients of f as fractions,


let a be the product of their denominators (so af D[X]

), and s c(af).
Then
a
s
f is an associate of f in K[X], and it is a primitive polynomial in
D[X]. In fact, we can obtain an associate of f K[X]

with arbitrary
content by multiplying
a
s
f by the appropriate element of D

.
124
2. From Theorem 112(5), we immediately have c(fg) = c(f)c(g) for f, g
D[X]

with f or g a monomial. This is a special case of Gausss Lemma


below.
Theorem 118 (Gausss Lemma). Let D be a GCD domain and K = QF(D).
1. Then c(f
1
f
n
) = c(f
1
) c(f
n
) for f
1
, . . . , f
n
D[X]

. Consequently,
the primitive polynomials form a multiplicative set of D[X].
2. Let f, g D[X]

. Then f [ g in D[X] if and only if f [ g in K[X] and the


elements of c(f) divide those of c(g) in D.
3. A non-constant polynomial in D[X] is irreducible in D[X] if and only if
it is primitive in D[X] and irreducible in K[X].
Proof.
1. Note that the second conclusion follows from the rst, since D

is a mul-
tiplicative set of D. It will suce to prove the rst statement for the
case n = 2, because one can then easily use induction to prove the rest
of the cases. Let f, g D[X]

. We proceed by induction on the total


number of terms in f and g. We have already noted the case where f
or g is a monomial, so let us assume that the total number of terms is
at least 4. If d
1
c(f) and d
2
c(g), then d
1
1
f and d
1
2
g are primitive
and c(fg) = d
1
d
2
c((d
1
1
f)(d
1
2
g)). This observation shows that it suces
to prove the case where f and g are primitive. Suppose that fg is not
primitive. Then the coecients of fg have some nonunit common divisor
x. Because x and the leading coecient of fg are not relatively prime, x
must have a nonunit common divisor y with the leading coecient of f or
g, say f. Let h be the leading term of f. Then y [ fg hg = (f h)g,
so y divides each coecient of (f h)g, hence divides every member of
c((f h)g), which by induction is c(f h)c(g) = c(f h). So y in fact
divides every coecient of f, a contradiction to the fact that f is primitive.
2. (): If g = fh in D[X], then c(g) = c(f)c(h) by (1). Thus, for d
1
c(f)
and d
2
c(g), some element of c(g) is divisible by d
1
, hence d
1
[ d
2
since
the elements of c(g) are the associates of d
2
. (): Assume g = fh in K[X]
and the elements of c(f) divide those of c(g) in D. Write the coecients
of h as fractions and let a be the product of their denominators. Then
ag = f (ah) is a product in D[X], and taking contents and applying (1)
gives ac(g) = c(f)c(ah). Since the elements of c(f) divide those of c(g)
(in D), we conclude that a divides c(ah) (in D), hence divides (in D) each
of the coecients of ah, so h =
ah
a
D[X], as desired.
3. Let f D[X] D. (): Assume f is irreducible in D[X]. Then f must
be primitive since f = d(d
1
f) for d c(f). Now assume f = gh, where
g, h K[X]

. Let a be the product of the denominators of the coecients


of g and s c(ag), and note that g
0
=
a
s
g is a primitive polynomial in
D[X]. Let h
0
=
s
a
h. Since g
0
divides f in K[X] and c(g
0
) = D

= c(f),
125
we conclude by (2) that g
0
[ f in D[X], hence f = g
0
h
0
is a product in
D[X], hence either g
0
or h
0
is a unit in D[X], hence either g or h is a unit
in K[X], as desired. (): By contrapositive. Assume f is reducible in
D[X] but irreducible in K[X]. (An element of a domain is called reducible
if it is a product of two nonzero nonunits.) Then we may write f = gh,
where g, h D[X]
#
, but necessarily one of those factors, say g, is a unit
in K[X]. This implies that g is constant, so g D
#
, hence f is not
primitive.
Theorem 119. Let D be a domain and ,= . Then D is a GCD domain if
and only if D[X

] is.
Proof. (): First note that D[X

is a saturated subset (in fact a multi-


plicative set) of D[X

] for each . (The additive degree formula


shows that the nonzero constants form a multiplicative set of any polyno-
mial ring over a domain, and the above conclusion then follows from writ-
ing D[X

] = D[X

][X

\
].) Thus f, g D[X
1
, . . . , X
n
]
have the same sets of (greatest) common divisors in D[X
1
, . . . , X
n
] as in
D[X

]. Because any two polynomials in D[X

] involve only nitely


many variables between them, this observation shows that it suces to show
that any polynomial extension of D by nitely many variables is a GCD do-
main, and using induction one shows that it suces to prove that D[X] is a
GCD domain. Let f, g D[X]

, d be a GCD of the coecients of f and g, and


K = QF(D). Then K[X] is a PID and hence a GCD domain, so there is an
h GCD
K[X]
(f, g), and since associates of GCDs are GCDs we may choose h
with d c(h). We claim that h GCD
D[X]
(f, g). Since h [ f, g in K[X] and
the elements of c(h) = dD

divide those of c(f) and c(g) in D, we conclude


by Gausss Lemma that h CD
D[X]
(f, g). Now let q CD
D[X]
(f, g), and note
that q [ h in K[X]. By Gausss Lemma, every element in c(q) is a common
divisor of c(f) c(g), hence divides every element of dD

= c(h). So in fact
q [ h in D[X], as desired. (): Assume D[X

] is a GCD domain. The


additive degree formula shows that D

is a saturated subset of D[X

], so
for a, b D

we have GCD
D
(a, b) = GCD
D[{X

]
(a, b) ,= , showing that D
is a GCD domain.
Exercises.
1. Show that, in a domain, every associate of an atom is an atom. (The
domain hypothesis is not necessary, but it makes the proof a lot easier.
Hint: Recall these two facts about domains from past exercises: (i) a
nonunit is an atom if and only if it cannot be written as a product of two
nonzero nonunits, and (ii) two elements are associates if and only if one
is a unit multiple of the other.)
2. Prove Theorem 112. (This is worth the same number of points as a six-part
problem. Hint: (1)-(4) follow from the denitions in a fairly straightfor-
ward way if you remember that (x) (y) y [ x. For (5), refer to the
126
denitions to show that if d GCD(xA), then x [ d and x
1
d GCD(A),
hence GCD(A) ,= . Also, you can derive GCD(xA) = xGCD(A) from
this. Keep in mind while doing (5) the rule y [ z yw [ zw, and also that
the GCDs of a set consist of the associates of any given GCD. (6) follows
from (5).)
3. Let D = 1 +XC[X].
(a) Show that GCD
D
(X
2
, iX
2
) = . (Hint: Suppose f GCD
D
(X
2
, iX
2
).
Note that X CD
D
(X
2
, iX
2
), hence X [ f, hence if D. From
fg = X
2
and fh = iX
2
, conclude that if CD
D
(X
2
, iX
2
). State
why this is a contradiction.)
(b) Show that GCD
D
(X, iX) = 1

and LCM
D
(X, iX) = . (Hint: For
the GCD, using degrees, show that every non-constant divisor of X
in D is of the form aX for some a 1

, and that aX does not divide


iX in D. For the LCM, suppose that m LCM
D
(X, iX). Use the
formula relating LCMs and GCDs to conclude that m is of the form
aiX
2
for some a 1

, then arrive at a contradiction by nding a


common multiple of X and iX that m does not divide.)
4. [Extra credit] A Bezout domain is a domain in which every nitely gen-
erated ideal is principal. (Thus a PID is a Bezout domain.) Prove that a
domain D is a Bezout domain if and only if it is a GCD domain and for
every a, b D

and d GCD(a, b) there are x, y D with ax + by = d.


(Hint: Recall the relationship between GCD(a, b) and (a, b).)
Appendix to 3.7: The Euclidean Algorithm
[Durbin: Section 12]
This extremely brief section contains a useful practical way to compute
GCDs in a Euclidean Domain that I neglected to
Theorem 120 (Euclidean Algorithm for Finding GCDs). Let D be a Euclidean
domain with algorithm , and let a, b D

. Dene r
1
= a, r
0
= b, and for
k 1 recursively dene q
k
D and r
k
D

so that r
k2
= r
k1
q
k
+ r
k
and (r
k
) < (r
k1
), stopping at the rst index n 0 with r
n
[ r
n1
. Then
(a, b) = (r
n
), and hence r
n
GCD(a, b).
Proof. First note that the construction is well-dened because the stopping
condition ensures that all the r
k
s are nonzero. Next note that the process must
indeed stop at some point, because otherwise (b) > (r
1
) > (r
2
) > is an
innite descending sequence of ordinals, violating the Axiom of Regularity. If
n = 0, then b = r
0
[ r
1
= a and (a, b) = (b) = (r
n
), so let us assume n 1.
Then (a, b) = (b, r
1
) = (r
n
), where the rst equality results from the equation
a = bq
1
+r
1
and the second equality is by induction.
127
Example. We use the Euclidean Algorithm to compute gcd(171, 112).
171 = 112 1 + 59
112 = 59 1 + 53
59 = 53 1 + 6
53 = 6 8 + 5
6 = 5 1 + 1.
Since 1 [ 5, we stop and conclude that gcd(117, 112) = 1. Since (117, 112) = (1),
we can write 1 as a linear combination of 117 and 112, i.e., in the form 1 =
117x + 112y. We can nd such an x and y by successively substituting:
1 = 6 5
= 6 (53 6 8) = 9 6 53
= 9(59 53) 53 = 9 59 10 53
= 9 59 10(112 59) = 19 59 10 112
= 19(171 112) 10 112 = 19 171 29 112.
(This is not to say that x = 19 and y = 29 is the only solution. In fact, there
are innitely many possible choices that work, due to the equality 117x+112y =
117(x 112) + 112(y + 117).)
The Euclidean Algorithm in the context of Z dates back to Euclids El-
ements. From a computing point of view, it is a rather ecient algorithm,
always completing in time polynomial to the number of digits of the number of
smaller absolute value. I mention this for computer scientists and others who
may be interested in algorithmic eciency, but we will not be studying eciency
in this course.
Exercises.
1. Use the Euclidean Algorithm to nd gcd(192, 82), then write this GCD as
a linear combination of 192 and 82. (Show your work for both steps.)
3.8 Unique Factorization Domains
[Durbin: Section 37]
Remark. We recall some of the previously discussed facts related to factoriza-
tion. Let R be a ring.
1. Elements a, b R are called associates if (a) = (b), or, equivalently, if a [ b
and b [ a. We have a [ b (b) (a), so elements that are associates have
the same divisors and the same multiples. From the equality (a)(b) = (ab)
it easily follows that the associate relation is a congruence relation on the
monoid (R, ). If R is a domain, then for a, b R

, a is an associate (resp.,
a proper divisor) of b if and only if there is a unit (resp., nonunit) c with
b = ac.
128
2. A nonunit a R is called irreducible or an atom if a = bc (a) = (b)
or (a) = (c), or, equivalently (by induction), if a = b
1
b
n
some
(b
i
) = (a). Associates of atoms are atoms, though we have only proven
this for the domain special case. In a domain, the atoms are precisely those
nonunits that cannot be written as a product of two nonzero nonunits.
Equivalently, an element a of a domain is an atom if and only if (a) is
maximal among the proper principal ideals. A ring is called atomic if
every nonzero nonunit is a product of atoms. Any ring satisfying the
ACCP is atomic.
3. A nonunit p of a ring is called prime if p [ ab p [ a or p [ b, or, equiv-
alently (by induction), if p [ a
1
a
n
p divides some a
i
. In verifying
the the rst denition, it suces to consider the case where a and b are
nonunits, since, if, say b R

, then a and ab have the same divisors. The


element p is prime if and only if the corresponding principal ideal (p) is a
prime ideal. Hence associates of primes are prime, and it also follows that
R is a domain if and only if 0 is prime. Every prime is irreducible, but we
have seen examples showing that not every irreducible is prime.
4. A unique factorization domain (UFD) is a domain in which every nonzero
nonunit can be uniquely (up to order and associates) written as a product
of atoms. (Such an expression is called an atomic factorization.) We have
UFD ACCP atomic, and we have seen examples showing that none
of these implications reverse.
Lemma 121. Let p
1
p
m
= q
1
q
n
be products of nonzero primes in an
integral domain. Then m = n and, after re-ordering if necessary, each (p
i
) =
(q
i
). In other words, prime factorizations of a given nonzero nonunit in an
integral domain are unique up to order and associates when they exist.
Proof. Exercise.
Lemma 122. Let p
1
, . . . , p
n
be non-associate primes of a domain D, and let
m
1
, . . . , m
n
, k
1
, . . . , k
n
N.
1. (p
m1
1
) (p
mn
n
) = (p
m1
1
p
mn
n
).
2. (p
m1
1
p
mn
n
) (p
k1
1
p
kn
n
) = (p
max(m1,k1)
1
p
max(mn,kn)
n
).
3. p
min(m1,k1)
1
p
min(mn,kn)
n
GCD(p
m1
1
p
mn
n
, p
k1
1
p
kn
n
) and p
max(m1,k1)
1
p
max(mn,kn)
n

LCM(p
m1
1
p
mn
n
, p
k1
1
p
kn
n
).
Proof. Exercise.
Example. 48 = 2
4
3 and 90 = 2 3
2
5, so gcd(48, 90) = gcd(48, 90) =
2
1
3
1
5
0
= 6 and lcm(48, 90) = lcm(48, 90) = 2
4
3
2
5 = 720. (Check:
6 720 = 48 90.)
Theorem 123 (Characterization of UFDs). The following are equivalent for
a domain D.
129
1. D is a UFD.
2. D is atomic and every atom is prime.
3. Every (nonzero) nonunit in D is a product of primes.
4. D is an atomic GCD domain.
Because UFD ACCP atomic, in (2) and (4), the atomic property could
equivalently be replaced by the ACCP property.
Proof. First we note that it does not matter whether or not (3) contains the
nonzero, because 0 is always prime in a domain. (1) (2): We must show
that if D is a UFD, p D is irreducible, a, b D
#
, and p [ ab, then p [ a or p [ b.
Write pc = ab, and note that necessarily c / D

(otherwise p = (ac
1
)b is a
product of two nonzero nonunits, contradicting irreducibility). Replace a, b, and
c in the above equation with atomic factorizations, and conclude by uniqueness
that p is an associate of one of the atoms in the factorizations of a or b, hence
divides a or b. (2) (4): Assume (2) holds. Then every nonzero nonunit
is associate to a product of the form p
m1
1
p
mn
n
, where p
1
, . . . , p
n
are non-
associate primes and each m
i
Z
+
. (Write that nonzero nonunit as a product
of primes and group together the factors that are associates.) Thus, given two
nonzero nonunits, one can nd associates of them that are written as in Lemma
122 and nd a GCD and LCM, so D is a GCD domain. (4) (2): Follows
immediately once we recall that atoms are prime in GCD domains. (2) (1):
Lemma 121. (2) (3): Clear. (3) (2): If (3) holds, then D is atomic since
primes are irreducible, and, moreover, every atom is a product of primes, hence
is prime by irreducibility.
Exercises.
1. Prove Lemma 121. (Hint: Proceed by induction on m.)
2. Prove Lemma 122. (Hint: For part (1), you can by induction reduce to
proving the case n = 2. For x (p
m1
1
) (p
m2
2
), write x = p
m1
1
a = p
m2
2
b.
Show that p
m2
2
[ a and x p
m1
1
p
m2
2
. Deduce part (2) from part (1), and
part (3) from part (2).)
3. A ring has (Krull) dimension if there are strictly ascending sequences
P
0
_ P
1
_ _ P
n
of prime ideals for arbitrarily large n N; otherwise,
its (Krull) dimension is the maximum possible such n. We denote the
dimension of a ring R by dimR.
(a) Prove that a ring is a eld if and only if it is a zero-dimensional
domain. (Hint: Recall that a ring is a eld if and only if (0) is
maximal.)
(b) Prove that a ring is a PID if and only if it is a UFD of dimension
at most 1. (Do not use Kaplanskys Theorem in your solution unless
you completed the extra credit exercise where it was proven. Hint:
130
[A solution without Kaplanskys Theorem.] For show from the
denitions that a PID has dimension at most 1. Showing that a PID
is a UFD can be done in multiple ways. For example: (i) cite the
appropriate results that show that a PID is an atomic GCD domain,
or (ii) cite results that show a PID is atomic and directly show that
atoms in a PID are prime. For , it suces by the results of the
appendix below to prove that every prime ideal is principal, which
will follow once you show that every nonzero prime ideal of a UFD
contains a prime element.)
4. Let D be a domain and ,= . Prove that D is a UFD if and only if
D[X

] is. (Hint: Use the fact that a UFD is the same thing as a
GCD domain satisfying the ACCP.)
5. [Extra Credit] Let D be a domain.
(a) Let S be the subset of D consisting of the units and the products of
nonzero primes. Show that S is a multiplicative set of D. (Recall
that a multiplicative set of D is a nonempty multiplicatively closed
saturated subset of D

, or in other words: (i) _ S D

and (ii)
ab S a, b S.)
(b) (Kaplanskys Theorem) Prove that D is a UFD every nonzero
prime ideal contains a nonzero prime element. (Hint: For , you
may cite your proof of Exercise 3(b) if you already proved this there.
For , proceed by contrapositive. Assuming that D is not a UFD,
there is some a D
#
that cannot be written as a product of primes.
Show that (a) is contained in some prime ideal disjoint from S. You
may need to revisit section 3.2 if it is not immediately clear how to
do this.)
(c) Prove that every localization of a UFD is a UFD. (Hint: Use Ka-
planskys Theorem.)
Appendix to 3.8: Cohens Theorem
[Not in Durbin]
The purpose of this appendix is to prove some lemmas that will simplify one
of the homework problems for 3.8. We return to discuss some topics dealing
with nitely generated and principal ideals. In particular, it turns out that, in
order to show that a ring is Noetherian (resp., a PIR), it suces to show that
every prime ideal is nitely generated (resp., principal).
Lemma 124. An ideal that is maximal among the non-nitely generated ideals
of a ring is prime.
Proof. Suppose to the contrary that there is a (necessarily proper) ideal P
maximal among the non-nitely generated ideals of a ring R, and that there are
131
a, b R P with ab P. By the maximality of P, the ideal P + (a) must be
nitely generated, say P + (a) = (x
1
+ r
1
a, . . . , x
n
+ r
n
a), where each x
i
P
and r
i
R. Let [P : a] = x R [ ax P, and note that P [P : a]
and b [P : a] (so [P : a] ,= ). For x, y [P : a] and r R, we have
a(x y) = ax ay P and a(rx) = r(ax) P, hence x y, rx [P : a].
Therefore [P : a] is an ideal by the Ideal Test. Since it properly contains P, it is
nitely generated, and so is [P : a](a). We arrive at the desired contradiction by
showing that P is the nitely generated ideal (x
1
, . . . , x
n
)+[P : a](a). Certainly
(x
1
, . . . , x
n
) + [P : a](a) P. On the other hand, for c P, we have c
P +(a) and c = y
1
(x
1
+r
1
a) + +y
n
(x
n
+r
n
a) for some y
1
, . . . , y
n
R, and
(y
1
r
1
+ +y
n
r
n
)a = cy
1
x
1
y
n
x
n
P, hence y
1
r
1
+ +y
n
r
n
[P : a]
and c = y
1
x
1
+ + y
n
x
n
+ (y
1
r
1
+ + y
n
r
n
)a (x
1
, . . . , x
n
) + [P : a]a, as
desired.
Theorem 125 (Cohens Theorem). Every non-nitely generated ideal of a ring
is contained in a maximal such ideal. Consequently, a ring is Noetherian if and
only if every prime ideal is nitely generated.
Proof. Note that the second statement will follow from the rst in view of the
lemma. By Zorns Lemma, it will suce to show that a union of a chain of
non-nitely generated ideals is non-nitely generated. A minor modication to
an argument used in the Noetherian rings section shows that, if the union of a
chain of ideals is nitely generated, then at least one of the ideals in the chain
is nitely generated (and in fact is equal to the union). The desired conclusion
immediately follows.
Remark. Cohens Theorem does not mean that the Noetherian property is
equivalent to the ascending chain condition on prime ideals. The latter property
is in general a much weaker property.
Lemma 126. An ideal maximal among the non-principal ideals of a ring is
prime.
Proof. Suppose to the contrary that there is a (necessarily proper) ideal P
maximal among the non-principal ideals of a ring R, and that there are a, b
RP with ab P. By maximality, we have P +(a) = (c) for some c R. Then
[P : c] is an ideal of R (see the proof of the other lemma), and (P, b) [P : c]
(we have b [P : c] since (bc) = (b)(P + (a)) P), so by maximality [P : c]
is principal. We have [P : c]c P by denition, and the other inclusion holds
since for x P (c) we may write x = yc, where by denition y [P : c].
Hence P = [P : c]c is principal, a contradiction.
Theorem 127. A ring is a PIR if and only if every prime ideal is principal.
Proof. (): Clear. (): Let R be a ring with every nonzero prime ideal
principal. Then R is Noetherian by Cohens Theorem. If the set of non-principal
ideals is nonempty, then the Noetherian property implies that this set has a
maximal element, which by the lemma is prime, a contradiction. Therefore R
is a PIR.
Chapter 4
Field Theory
In this chapter we will touch upon the basics of eld theory. While elds are
the simplest and most well-behaved type of ring, there is a surprising amount of
depth to their theory. Our ultimate goal will be to show how abstract algebra
nally settled the classic ancient Greek straightedge and compass construction
problems. Along the way, we will study such topics as vector spaces, eld
extensions, algebraic/transcendental elements, and more.
4.1 Vector Spaces
[Not in Durbin.]
In this section, we will touch upon the basics of linear algebra, mostly con-
ning ourselves to those facts that we will later make use of.
Denition. Let R be a ring.
1. An R-module or module over R is an additive abelian group M together
with a map : R M M (called the scalar multiplication) such that
for each r, s R and a, b M:
(a) (r +s)a = ra +sa,
(b) (rs)a = r(sa),
(c) r(a +b) = ra +rb, and
(d) 1 a = a.
(Note that the usual multiplicative notation conventions apply to scalar
multiplication.)
2. When one is talking about an R-module, the elements of R are called
scalars.
3. A submodule of an R-module M is a subset of M that is a module under
the same addition and scalar multiplication, or, equivalently, a subset of
M closed under addition and scalar multiplication.
132
133
4. Let M be an R-module. An element of the form c
1
x
1
+ + c
n
x
n
with
n N and each c
i
R and x
i
M is called a linear combination of
x
1
, . . . , x
n
. For X M, the span of X is the set span(X) of all linear
combinations of elements of X. (Note that span() = 0 and span(M) =
M by denition.) We abbreviate span(a
1
, . . . , a
n
) = span(a
1
, . . . , a
n
).
The smallest submodule of M containing X is span(X), and consequently
span(span(X)) = span(X). (Exercise.) We say that X spans M if M =
span(X).
5. If K is a eld, then a K-vector space or vector space over K is a K-module
V . We call the elements of V vectors, and we call the submodules of V
subspaces.
Proposition 128. Let R be a ring and M be an R-module. The following
properties hold for a M and r R:
1. 0 a = r 0 = 0.
2. (r)a = ra. In particular, we have (1)a = a.
3. If R is a eld (or in other words M is a vector space), then ra = 0 r = 0
or a = 0.
Proof.
1. We have 0 a +0 a = (0 +0)a = 0 a, and canceling yields 0 a = 0. The
proof that r 0 = 0 is similar.
2. We have ra + (r)a = (r r)a = 0 a = 0 by (1), hence (r)a = ra.
3. (): If ra = 0 and r ,= 0, then a = r
1
(ra) = r
1
0 = 0 by (1). ():
Follows from (1).
Example.
1. Every ring is a module over itself. It submodules are simply its ideals.
2. Any eld is a vector space over itself. Its only subspaces are the zero
subspace and the whole space.
3. To generalize (1) and (2), any ring (resp., eld) is a module (resp., vector
space) over any subring (resp., subeld).
4. If M

is a nonempty family of R-modules, then so are

and

, under the scalar multiplication given by (rf)() = rf().


As a special case, we have the familiar vector spaces

n
i=1
1 studied in an
elementary linear algebra course.
134
5. If R is any ring (resp., eld), then /
n
(R) is an R-module (resp., R-
vector space) under the usual scalar multiplication. (If r R and A
/
n
(R), then rA /
n
(R) is the matrix whose elements are r times the
corresponding element of A.)
6. Any abelian group G is a Z-module, where the scalar multiplication ng
has the familiar denition. The submodules are simply the subgroups.
Thus an abelian group can be viewed as a special case of a module over
a PID, and there are theorems about the structure of such things (which
we will not discuss). This is what I meant earlier when I said that the
Fundamental Theorem of Finite Abelian Groups is most naturally proven
in the context of module theory.
Denition. Let R be a ring. An R-module homomorphism is a map : M N
between R-modules that satises (rm) = r(m) and (m+n) = (m) +(n)
for r R and m, n M. (Note that this means an R-module homomorphism is
also a group homomorphism.) We use this to dene phrases such as R-module
isomorphism, etc., in the usual way. In the case where R is a eld, we call
R-module homomorphisms linear transformations.
Example.
1. The 1-vector spaces 1 1 and C are isomorphic via : 1 1 C :
(a, b) a +bi.
2. It is useful to point out that there is a big dierence between the denitions
of module homomorphism and ring homomorphism. The former satises
(rm) = r(m), while the latter satises (rm) = (r)(m).
3. In (1) we see an example of a module homomorphisms between rings that
is not a ring homomorphism, since ((0, 1)(0, 1)) = ((0, 1)) = i but
((0, 1))
2
= i
2
= 1.
4. Similarly, a ring homomorphism between modules need not be a module
homomorphism. For example, the ring homomorphism : Z[X] [X] :
f f(0) is not a Z[X]-module homomorphism because (X 1) = 0 ,=
X = X(1).
Denition. Let V be a vector space over a eld K.
1. A subset L of V is called linearly independent if for c
1
, . . . , c
n
K and
distinct x
1
, . . . , x
n
L we have c
1
x
1
+ + c
n
x
n
= 0 each c
i
= 0. A
subset L is linearly independent if and only if x / span(L x) for each
x L. (Exercise.) A subset of V is linearly dependent if it is not linearly
independent.
2. A basis of V is a linearly independent spanning subset. Equivalently, a
subset B is a basis of V if and only if for each nonzero v V there is a
unique (up to order) way to write v = c
1
b
1
+ +c
n
b
n
with c
1
, . . . , c
n
K

and b
1
, . . . , b
n
distinct elements of B. (Exercise.)
135
Example.
1. In any vector space, is linearly independent and 0 is linearly depen-
dent.
2. Any subset (resp., superset) of a linearly independent (resp., dependent)
set is linearly independent (resp., dependent). Thus no linearly indepen-
dent set contains 0.
3. (1, 0), (0, 1) is a basis for 1 1. Another basis is (1, 1), (2, 3).
The next theorem shows that every vector space has a basis. In fact, it
shows that every linearly independent subset can be expanded to a basis, and
every spanning subset can be shrunk to a basis.
Theorem 129 (Vector Spaces Have Bases). Let V be a vector space over a eld
K, L be a linearly independent subset of V , and S be a spanning subset of V
containing L. Then there is a basis B of V with L B S.
Proof. Let ( be the set of linearly independent subsets of S that contain L.
We claim that the nonempty partially ordered set ((, ) has a maximal
element B. (Note that L (.) By Zorns Lemma, it suces to show that
the union of any nonempty chain L

in ( is in (. Since L L

S
for each , we have L

S. For linear independence, assume


c
1
x
1
+ +c
n
x
n
= 0 for c
1
, . . . , c
n
K and distinct x
1
, . . . , x
n

, say
x
i
L
i
. Because L
1
, . . . , L
n
is a nite (hence Noetherian) totally ordered
set, it has a maximum element, say L
j
. We have x
1
, . . . , x
n
L
j
, and by the
linear independence of L
j
we have each c
i
= 0, as desired.
We have L B S, and B is linearly independent, so all that remains is to
show that span(B) = V. It suces to show that S span(B), because then we
have V = span(S) span(span(B)) = span(B). Let s S. If s B, then we
are done, so let us assume s / B. Since L B B s S, the set B s
must be linearly dependent by maximality, so there are c
1
, . . . , c
n
K not all
zero and b
1
, . . . , b
n1
B with c
1
b
1
+ +c
n1
b
n1
+c
n
s = 0. If c
n
= 0, then
necessarily c
i
,= 0 for some i n1, and the equation c
1
b
1
+ +c
n1
b
n1
= 0
contradicts the linear independence of B. Thus c
n
,= 0, and s = (c
1
n
c
1
)b
1
+
+ (c
1
n
c
n1
)b
n1
span(B), as desired.
Corollary 130. A subset of a vector space is a basis it is a minimal spanning
subset it is a maximal linearly independent subset.
Proof. Label the statements (1)-(3), respectively. (1) (2), (3): By an exercise,
any linearly independent set has a strictly larger span than each of its proper
subsets. Hence no proper subset of a basis can span the vector space, and each
proper superset of a basis, being a subset with the same span as one of its
proper subsets, cannot be linearly independent. Therefore a basis is a minimal
spanning subset and a maximal linearly independent subset. (2), (3) (1):
Every spanning subset can be shrunk to a basis, and every linearly independent
set can be expanded to a basis. Hence a minimal spanning subset or a maximal
linearly independent set must be a basis.
136
Theorem 131. Let K be a eld, V be a K-vector space, and B be a basis for
V . Then [V [ =
_
[K[
|B|
, B is nite
max([B[, [K[), B is innite.
Proof. If B = , then [V [ = [0[ = 1 = [K[
0
= [K[
|B|
, so we will now assume
B ,= . If B is nite, then each element of V has a unique representation
of the form

bB
c
b
b with each c
b
K, so [V [ = [K[
|B|
. So let us assume
B is innite. If K is nite, then each nite X B has nite span by the
nite basis case, and the identity V = span(B) =

XP
f
(B)
span(X) shows that
[B[ [V [ [T
f
(B)[
0
= [B[, hence [V [ = [B[ = max([B[, [K[). Finally, assume
K is innite. Then [span(X)[ = [K[
|X|
= [K[ for each nite X B, hence [B[
[V [ [T
f
(B)[[K[ = max([B[, [K[). On the other hand, picking some b B, we
obtain [K[ = [K[
1
= [span(b)[ [V[, hence [V [ = max([B[, [K[).
Corollary 132.
1. For any ring R and nonempty set , we have [R[X

][ = max([R[, [[,
0
).
2. Every localization of a domain is the same cardinality as the original do-
main. In particular, a domain has the same cardinality as its eld of
quotients.
3. There is a eld of every innite cardinality.
4. Any nite eld has prime-power cardinality.
Proof.
1. For notational convenience, let S =

N, so R[X

] = R[X; S].
Then R[X; S] is an R-module and each of its nonzero elements has a unique
expression as an linear combination of the elements from the set X
s

xS
of cardinality [S[. The proof of the above lemma only really used the
uniqueness of the representations (not the fact that the scalars are from
a eld), so we have [R[X; S][ = max([R[, [S[) (since [S[ is innite), and it
suces to show that [S[ = max([[,
0
). Also, S is a semimodule (or
whatever one would call a module over a semiring) over N in which every
element has a unique expression as a linear combination of the elements
from the set of elements with exactly one nonzero coordinate (a set which
has size [[[N[), we conclude by the argument of the lemma that [S[ =
max([N[, [[[N[) = max([[,
0
), as desired.
2. Let D be a domain and K = QF(D). For any localization D
S
of D, we
have D D
S
K, so it suces to show that [K[ [D[. If D is nite,
then D is a eld and D = K, so let us assume D is innite. The map
D D

K : (a, s)
a
s
is a surjection, hence [K[ [D D

[ =
[D[[D

[ = [D[
2
= [D[, as desired.
3. Let be any innite cardinal. Then by (1) and (2) we have [QF(Z[X

<
])[ =
[Z[X

<
][ = max([Z[, [[,
0
) = .
137
4. Let K be any nite eld. Then Z
p
embeds into K, so we may regard K as
a Z
p
-vector space. By the above theorem, it necessarily has a nite basis
B, and [K[ = [Z
p
[
|B|
= p
|B|
.
Remark.
1. It follows from (3) above that the class of all elds is a proper class.
(Past exercise.) Since every type of algebraic structure we have discussed
contains the elds as a special case (for example, every eld is a group
under addition, and every eld is a vector space over itself), it follows
that, for any of the algebraic structures we have discussed, the class of all
such structures is a proper class.
2. We will study nite elds in a bit more detail later. In particular, we
will show that for each prime power cardinality there is a unique (up to
isomorphism) nite eld of that cardinality.
3. The uniqueness does not extend to innite elds. For example, for any
prime p Z
+
, the elds and QF(Z
p
[X]) both have cardinality
0
, but
they are not isomorphic because the former has characteristic zero and
the latter has characteristic p. A more familiar example is given by the
elds 1 and C. We have [1[ = [C[ by a past exercise, but the two elds
are not isomorphic because 1 has a square root in C but not in 1.
Theorem 133 (Replacement Theorem). Let V be a vector space, B be a basis
for V of nite cardinality n, and L be a linearly independent subset. Then
[L[ n and there are distinct b
|L|+1
, . . . , b
n
B L with L b
|L|+1
, . . . , b
n
a
basis.
Proof. The case [L[ = 0 is trivial, so let us assume [L[ > 0. Each nite
proper subset of L has cardinality less than [L[, hence by induction has car-
dinality at most n, and it follows that L is nite. Write L = a
1
, . . . , a
m
,
where m Z
+
. By induction, we have m 1 n, and there are distinct
b
m
, . . . , b
n
B a
1
, . . . , a
m1
with B
1
= a
1
, . . . , a
m1
, b
m
, . . . , b
n
a basis.
Write a
m
= c
1
a
1
+ + c
m1
a
m1
+ c
m
b
m
+ + c
n
b
n
, where c
1
, . . . , c
n
are
scalars. By the linear independence of the a
i
s, we see that m n and c
i
,= 0
for some i m (otherwise c
1
a
1
+ + c
m1
a
m1
+ (1)a
m
= 0, a contradic-
tion). If necessary, re-index so that i = m. If b
j
= a
m
for some j > m, then
a
m
= c
1
a
1
+ +c
m1
a
m1
+c
m
b
m
+ +c
n
b
n
and a
m
= b
j
are two distinct rep-
resentations of a
m
as a linear combination of the elements of the basis B
1
, a con-
tradiction. Thus b
m+1
, . . . , b
n
B L. Let B
2
= a
1
, . . . , a
m
, b
m+1
, . . . , b
n
.
The above equation for a
m
shows that a
m
span(B
1
) and b
m
span(B
2
), hence
span(B
2
) = span(B
1
) = V. It only remains to show that B
2
is linearly indepen-
dent. For this, take any scalars d
1
, . . . , d
n
with d
1
a
1
+ +d
m
a
m
+d
m+1
b
m+1
+
+ d
n
b
n
= 0. Substituting the above equality for a
m
, collecting like terms,
and using the linear independence of B
1
, we see from that the coecient c
m
d
m
138
of b
m
is 0, hence d
m
= 0 since c
m
,= 0. By the linear independence of B
2
, this
in turn implies that we also have d
i
= 0 for each i ,= m, as desired.
Theorem 134 (Dimension of a Vector Space). Every basis of a given vector
space has the same cardinality.
Proof. Let V be a vector space. It follows from the Replacement Theorem that
if V has a nite basis, then every basis is nite and of the same cardinality. So
we only need to consider the case where B
1
and B
2
are innite bases. Every
b B
1
can be written as a linear combination of elements from some nite
B
b
B
2
, hence span(

bB1
B
b
) = span(span(

bB1
B
b
)) span(B) = V, and
by the fact that B
2
is a minimal spanning set we obtain B
2
=

bB1
B
b
and
[B
2
[ [B
1
[
0
= [B
1
[. The reverse inequality holds by symmetry.
Denition. By the above theorem, we can unambiguously dene the dimension
of a vector space to be the cardinality of one of its bases. We write the dimension
of a K-vector space V as dim
K
V . (We may drop the subscript if what we
mean is obvious from context, but be careful not to mix this up with the Krull
dimension of a ring.)
Corollary 135. If dimV = n < , then an n-element subset of V is a basis
it is linearly independent it spans V .
Proof. We only need to show that an n-element linearly independent or spanning
set B is a basis. Then some subset or superset of B is a basis. All bases of V
have n-elements, so no proper subset or superset of B is a basis, hence B itself
must be a basis.
Example.
1. A zero-dimensional vector space is the same thing as the zero vector space.
(Since span() = 0.)
2. dim
R
1 1 = 2, since (1, 0), (0, 1) is a basis.
3. More generally, if K is any eld and ,= , then dim
K

K = [[,
since the set of elements with one coordinate equal to 1 and the rest equal
to zero forms a basis.
The following theorem shows that vector spaces have an almost disappoint-
ingly simple structure.
Theorem 136 (Structure of Vector Spaces). Let K be a eld and be a positive
cardinal. Then every K-vector space of dimension is isomorphic to

<
K.
Proof. Let V be a K-vector space with a basis B of cardinality . Index B =
b

<
, and dene :

<
K V : f

<
f()b

. (Note that this


function is well-dened since at most nitely many f()s are nonzero, so the
sum is really a nite sum.) It is readily veried that is a linear transformation,
and it is a bijection due to the uniqueness of the representations, so is a vector
space isomorphism.
139
Remark. Starting with this theorem, one can develop the theory of linear
transformations between nite-dimensional vector spaces as corresponding to
matrices, but we will not be doing that in this course. Depending on your linear
algebra course, you may have already seen this theory developed with some
degree of rigor for nite-dimensional vector spaces over 1 and/or C. A good
portion of the theory carries over to arbitrary nite-dimensional vector spaces
with essentially the same proofs.
Proposition 137. Let K
1
K
n
be elds. Then
dim
K1
K
n
= (dim
Kn1
K
n
)(dim
Kn2
K
n1
) (dim
K1
K
2
).
Proof. By induction, we may reduce to proving the case n = 3. Let B
1
be a
basis for K
2
over K
1
, B
2
be a basis for K
3
over K
2
, and B = b
1
b
2
[ b
i

B
i
. We will show that B is a basis for K
3
over K
1
of cardinality [B
1
[[B
2
[ =
(dim
K1
K
2
)(dim
K2
K
3
). For the cardinality claim, it suces to show that each
element of B has a unique representation of the given form. If b
1
b
2
= b

1
b

2
for
some b
1
, b

1
B
1
and b
2
, b

2
B
2
, then b
1
b
2
b

1
b

2
= 0, so b
2
= b

2
(otherwise
the linear independence of B
2
is contradicted), and canceling yields b
1
= b

1
.
To see that B spans K
3
, note that each element of K

3
can be written in the
form c
1
b
1
+ + c
n
b
n
with each b
i
B
2
and c
i
K
2
. Writing each c
i
as a
K
1
-linear combination of elements of B
1
, we may rewrite the above expression
as the desired K
1
-linear combination of elements of B. It only remains to show
that B is linearly independent. For this, take any distinct b
1
, . . . , b
m
B
1
and
b

1
, . . . , b

n
B
2
and any c
1,1
, . . . , c
m,n
K with

i,j
c
i,j
b
i
b

j
= 0. Rearranging
this to obtain

n
j=1
(

m
i=1
c
i,j
b
i
)b

j
= 0, we see by the linear independence of
B
2
that each

m
i=1
c
i,j
b
i
= 0, and then by the linear independence of B
1
that
each c
i,j
= 0, as desired.
Exercises. In the rst two exercises, do not use any results later than the
relevant denitions. (Some of proofs in this section use the results of these
exercises.)
1. Let R be a ring, M be an R-module, and X M.
(a) Show that span(X) is the smallest submodule of M containing X.
(b) Show that span(span(X)) = span(X). (Hint: Use part (a).)
(c) Assume that R is a eld. (So M is an R-vector space.) Show that
X is linearly independent if and only if x / span(X x) for each
x X.
2. Let K be a eld, V be a K-vector space, and B V . Prove that B is a
basis for V if and only if for each nonzero v V a unique (up to order)
way to write v = c
1
b
1
+ + c
n
b
n
with c
1
, . . . , c
n
K

and b
1
, . . . , b
n
distinct elements of B.
3. Let W be a subspace of a vector space V . Show that dimW dimV .
(Hint: Find the result to cite that shows that a basis for W can be extended
to a basis for V .)
140
4. [Extra Credit] Let R be a ring and dene
d
dx
: R[X] R[X] analogously
to the polynomial rule from calculus, i.e.,
d
dx
(a
n
X
n
+ +a
0
) = na
n
X
n1
+ (n 1)a
n1
X
n2
+ +a
1
.
For f R[X], abbreviate
d
dx
(f) =
df
dx
= f

. (For this problem, do not


try to quote facts from calculus. Strictly speaking, calculus facts do not
apply here since these are not necessarily polynomials over 1.)
(a) Prove that
d
dx
is an R-module homomorphism.
(b) Show that
d
dx
is not injective.
(c) Prove the product rule:
d
dx
(f
1
f
n
) =

n
k=1
f
1
f
k1
f

k
f
k+1
f
n
.
(Note that an immediate consequence is the power rule
d
dx
(f
n
) =
nf
n1
f

. Hint: Proceed by induction on n. The hardest case is


n = 2. This can be done straight from the denitions but it is
tedious. A more elegant method is the following. First prove the
rule (*):
d
dx
(Xf) = f + Xf

directly from the denitions [you will


use this later]. Then proceed by induction on deg f
1
. The base case
deg f
1
0 follows fairly easily from (a). For deg f
1
1, let h be
the leading term of f, and use (*) and induction to nd
d
dx
(hf
2
) and
d
dx
((f
1
h)f
2
).)
(d) Let f R[X] and a be a root of f (meaning that f(a) = 0). We say
a is a repeated root of f if (X a)
2
[ f. Show that a is a repeated
root of f if and only if f

(a) = 0. (Hint: Use the Factor Theorem


and the product rule.)
Appendix to 4.1: Quotient Modules
Denition.
1. Let R be a ring, M be an R-module and N be a submodule of M. Then
the abelian group (M/N, +) can be made into an R-module, called the
quotient module of M by N, by dening the scalar multiplication by r(m+
N) = rm+N. (The only nontrivial detail to check is that this operation is
well-dened. To see this, note that if m+N = m

+N, then mm

N,
so rmrm

= r(mm

) N, hence rm+N = rm

+N, as desired.)
2. The kernel of a module homomorphism is the same as its kernel when
considered as a group homomorphism. Thus a module homomorphism is
injective if and only if its kernel is the trivial submodule.
3. The kernel of a module homomorphism is a submodule, and module ho-
momorphisms take submodules to submodules. (The proofs of these facts
are similar to the analogous facts from ring theory.)
141
Theorem 138 (Fundamental Homomorphism Theorem for Modules). Let R
be a ring, M and N be R-modules, and : M N be an R-module homomor-
phism. Then M/ ker

= [M].
Proof. From group theory we already know that (M/ ker , +)

= ([M], +) via
: M/ ker [M] : m + ker (m). But for each r R we have
(r(m + ker )) = (rm + ker ) = (rm) = r(m) = r(m + ker ), so is
also an R-module isomorphism.
Theorem 139 (Rank-Nullity Theorem). Let : V W be a linear transfor-
mation. Then dimV = dimker + dim[V ].
Proof. Let B be a basis for ker , and expand B to a basis B

for V . We claim
that the set b + ker [ b B

B is a basis for V/ ker . Each element


of can be written as a linear combination of elements of B

, and the cosets


of elements of B are zero in V/ ker , so it follows that the given set spans
V/ ker . For linear independence, assume (c
1
b
1
+ +c
n
b
n
) +ker = 0+ker
for some scalars c
1
, . . . , c
n
and distinct b
1
, . . . , b
n
B

B. Then the element


c
1
b
1
+ +c
n
b
n
, being in ker = span(B) can be written as a linear combination
of distinct elements of B. Writing out this equation and moving all elements to
one side, we see by the linear independence of B

that each c
i
= 0, establishing
the claim. Thus dimV = [B

[ = [B[ + [B

B[ = dimker + dimV/ ker =


dimker + dim[V ].
Exercises.
1. Let : V W be a linear transformation. Show that the image of a
basis for V is a basis for W.
4.2 Field Extensions
[Durbin: Section 42]
Denition.
1. Let R T be rings and c T. For A T, we use R[A] to denote
the subring of T generated by R A, i.e., the smallest subring of T con-
taining R A. (This exists and equals the intersection of all subrings
of T containing R A.) Observe that R R[A] T by denition. If
we index A = a

, it is easily checked that R[A] = f(a

) [
f R[X

]. (The ring R[A] contains this set by the appropriate


closure properties, and the Subring Test shows that this is a subring of
T containing R A, hence equality holds.) Therefore the ring R[A] is in
fact dened independently of T, so we do not need to worry about any
ambiguity resulting from the notation R[A] not indicating what T is, and
we could refer to R[A] as simply the smallest ring containing R and A.
Also, the identity R[A
1
A
n
] = R[A
1
] [A
n
] holds. (Exercise.)
We abbreviate R[c
1
, . . . , c
n
] = R[c
1
, . . . , c
n
]. As a special case of our
observations above, we note that R[c] = a
n
c
n
+ +a
1
c +a
0
[ a
i
R.
142
2. Let K L be elds and A L. We use K(A) to denote the subeld of L
generated by KA, i.e., the smallest subeld of L containing RA. (Again,
this exists and equals the intersection of all subelds of L containing RA.)
Observe that K K(A) L by denition. Also, we have K(A) =
QF(K[A]) (exercise), so K(A) is in fact dened independently of L, and
could simply be referred to as the smallest eld containing K and A. We
again have the identity K(A
1
A
n
) = K(A
1
) (A
n
). (Exercise.)
We abbreviate K(c
1
, . . . , c
n
) = K(c
1
, . . . , c
n
).
Remark. This notation is consistent with our notation for polynomial rings,
because R[X

] is the smallest ring containing R and each indeterminate.


Denition. Let K L be elds and c L. We say c is algebraic over K if c
is a root of a nonzero polynomial with coecients in K. Otherwise, we say c is
transcendental over K.
Example.
1. Any element of K is algebraic over K. (Since a K is a root of X a
K[X].)
2. The element

2 is algebraic over since it is a root of X


2
2 (which is
its minimal polynomial). The elements and e are transcendental over
, but this is exceedingly dicult to prove and well beyond the scope of
this course.
In the following theorem, by a monic polynomial, we mean a polynomial
whose leading coecient is 1.
Theorem 140. Let c be algebraic over K. Then there is a unique monic irre-
ducible m
c,K
K[X] with c as a root, and f K[X] [ f(c) = 0 = (m
c,K
).
Proof. Exercise.
The polynomial m
c,K
in Theorem 140 is called the minimal polynomial of c
over K.
Theorem 141. Let K be a eld and f K[x] be a polynomial of degree n 1.
Then there is a eld containing K in which f has a root. If f is irreducible and
c is a root of f (in some extension of K), then K[c] = K(c) = a
n1
c
n1
+
+a
1
c +a
0
[ a
0
, . . . , a
n1
K

= K[X]/(f) (both as rings and vector spaces)


and dim
K
K(c) = n.
Proof. Writing f as a product of irreducibles (recall that K[X] is a UFD),
we see that f has some non-constant irreducible divisor p, and any root of p
is a root of f. So we only need to consider the case where f is irreducible.
Because K[X] is a PID, hence one-dimensional (past exercise), the nonzero
prime ideal (f) is maximal, and thus K[X]/(f) is a eld. Moreover, note that
K embeds into K[X]/(f) via a a + (f). (It is easily checked that this
map is a ring homomorphism, and it is injective since if f [ a b in K[X] for
143
some a, b K, then a b = 0 by degree considerations, hence a = b.) Write
f = b
n
X
n
+ +b
0
, where b
0
, . . . , b
n
K. Then X+(f) K[X]/(f) is a root of
f, since f(X+(f)) = b
n
(X+(f))
n
+ +b
1
(X+(f))+b
0
= f +(f) = 0+(f). We
show that dim
K
K[X]/(f) = deg f by verifying that 1 + (f), . . . , X
n1
+ (f)
is a basis for L over K. This is a spanning set since by the Division Algorithm
every element of L is of the form g + (f) with deg g n 1. It is linearly
independent over K since if c
n1
(X
n1
+(f)) + +c
0
(1+(f)) = 0+(f), then
f [ c
n1
X
n1
+ +c
0
in K[X], and by degree considerations we conclude that
each c
i
= 0.
Finally, let c be any root of f. Then (f) (m
c,K
), and by maximality we
have (f) = (m
c,K
). The map : K[X] K[c] : g g(c) has kernel (m
c,K
) is
both a ring and a K-vector space epimorphism, hence K[c]

= K[X]/(m
c,K
) =
K[X]/(f) (via g+(f) g(c)) both as rings and as vector spaces. So K[c] is actu-
ally a eld, and thus K[c] = K(c). Finally, because 1+(f), . . . , X
n1
+(f) is a
basis for K[X]/(f), its image 1, c, . . . , c
n1
under the K-vector space isomor-
phism g + (f) g(c) is a basis for K[c], hence K[c] = span
K
1, c, . . . , c
n1
=
a
n1
c
n1
+ + a
1
c + a
0
[ a
0
, . . . , a
n1
K.
Denition. Let i denote a root of X
2
+ 1 1[X]. We dene C = 1[i].
Theorem 142. Let K L be elds and c L. Then c is algebraic over K if
and only if dim
K
K(c) is nite. In this case, we have deg m
c,K
= dim
K
K(c).
Proof. (): If c is algebraic over K, then it is a root of the irreducible poly-
nomial m
c,K
, and by Theorem 141 we have dim
K
K(c) = deg m
c,K
. ():
Assume dim
K
K(c) = n < . If c
m
= c
k
for some natural numbers m < k,
then c is a root of X
k
X
m
, so let us assume all the nonnegative powers of
c are distinct. Then 1, c, c
2
, . . . , c
n
, being a subset of K(c) of cardinality
greater than n, is linearly dependent, so there are a
0
, . . . , a
n
K not all zero
with a
n
c
n
+ + a
1
c + a
0
= 0, and c is a root of the nonzero polynomial
a
n
X
n
+ +a
0
K[X].
Example. One way to see that dim
R
C = 2 is to note that m
i,R
= X
2
+1. (By
an exercise, a non-constant polynomial over a eld is irreducible if and only if
it has no roots.) Thus dim
R
C = deg(X
2
+ 1) = 2.
Theorem 143. Let K L be elds. Then the set of elements of L that are
algebraic over K form a subeld of L.
Proof. The only nontrivial detail to check is that if a, b L are algebraic
over K and b ,= 0, then a b, ab, and
a
b
are algebraic over K. (The ele-
ments a b and ab being algebraic over K is trivial in the case where b = 0.)
Since a and b are algebraic over K, the element b is certainly algebraic over
K(a), so dim
K
K(a) and dim
K(a)
K(a)(b) are nite, and thus dim
K
K(a, b) =
(dim
K(a)
K(a)(b))(dim
K
K(a)) <
0
. Because the K-vector spaces K(a
b), K(ab), and K(
a
b
) are all subspaces of K(a, b), we conclude that they are
nite-dimensional, hence a b, ab, and
a
b
are algebraic over K, as desired.
144
Theorem 144. Let K L be elds and c L be transcendental over K. Then:
1. K[c] _ K(c).
2. K[c]

= K[X] and K(c)

= K(X) (both as rings and as K-vector spaces).


Proof. Consider the map : K(X) K(c) :
f
g

f(c)
g(c)
. To see that it is
well-dened, note that g(c) ,= 0 for any nonzero g K[X], and it is also
easily checked that if
f1
g1
=
f2
g2
, then
f1(c)
g1(c)
=
f2(c)
g2(c)
. It is also readily veried
that is a ring and a K-vector space homomorphism, and the fact that it is
an epimorphism follows from the previously noted fact that K[c] = [K[X]].
Since K is transcendental, we have ker = 0, so is in fact an isomorphism.
Thus K(X)

= K(c) and K[X]

= [K[X]] = K[c] (both as rings and vector
spaces). Because K(c)

= K(X) is a eld and K[c]

= K[X] is not, we have
K[c] _ K(c).
Exercises.
1. Prove the identities K(A) = QF(K[A]), R[A
1
A
n
] = R[A
1
] [A
n
],
and K(A
1
A
n
) = K(A
1
) (A
n
) noted in the denitions at the start
of this section. (Hint: For the latter two, use induction. Their proofs are
pretty similar.)
2. (a) Prove Theorem 140. (Hint: Recall that K[X] is a PID.)
(b) Let K L be elds and c be algebraic over L. Prove that m
c,L
[ m
c,K
in L[X]. (Hint: Use Theorem 140.)
3. Let K be a eld and f K[X] satisfy 2 deg f 3. Prove that f is
irreducible if and only if it has no root in K. (Hint: Use the additive
degree formula.)
4. (a) Let D be a domain and f D[X] be a non-constant polynomial of
degree n. Prove that f has at most n roots. (Hint: Show that if
a
1
, . . . , a
m
are distinct roots of f, then (X a
1
) (X a
m
) [ f.)
(b) [Extra Credit] Let K L be innite elds. Prove that the set of
elements of L that are algebraic over K has the same cardinality as
K. (Note that this provides an indirect proof that there are real
numbers transcendental over . However, proving that any specic
real number is transcendental over is quite hard.)
4.3 Splitting Fields
Lemma 145. Let K L be elds. Then K = L if and only if dim
K
L = 1.
Proof. Exercise.
Denition. Let K L be elds. We say L is an algebraic eld extension of K
if every element of L is algebraic over K.
145
Theorem 146 (Finite Degree Algebraic Extensions). The following are equiv-
alent for elds K L.
1. dim
K
L <
0
.
2. L = K(a
1
, . . . , a
n
) for some elements a
1
, . . . , a
n
L (n 0) algebraic
over K.
In this case, the eld L is algebraic over K.
Proof. (): Assume dim
K
L <
0
. If there is a c L transcendental over K,
then dim
K
L dim
K
K[c] = dim
K
K[X] =
0
, a contradiction. Therefore L is
algebraic over K. Suppose that (2) does not hold. We will recursively construct-
ing a sequence a
n

n=1
of elements of L with each dim
K
K(a
1
, . . . , a
n
) 2
n
,
which will lead to a contradiction by implying that dim
K
L
0
. Because
L ,= K, there is an a
1
L K, and dim
K
K(a
1
) 2 = 2
1
. Having constructed
a
1
, . . . , a
n1
L with dim
K
K(a
1
, . . . , a
n1
) 2
n1
, we have some a
n
L
K(a
1
, . . . , a
n1
), and dim
K
K(a
1
, . . . , a
n
) = (dim
K(a1,...,an1)
K(a
1
, . . . , a
n
))(dim
K
K(a
1
, . . . , a
n1
))
22
n1
= 2
n
, and the construction is complete. (): Assume L = K(a
1
, . . . , a
n
)
for some elements a
1
, . . . , a
n
L (n 0) algebraic over K. If n = 0, then L = K
and dim
K
L = 1, so let us assume n 1. We have dim
K(a1,...,an1)
L <
0
since
a
n
is algebraic over K(a
1
, . . . , a
n1
), and dim
K
K(a
1
, . . . , a
n1
) <
0
by induc-
tion, hence dim
K
L = (dim
K(a1,...,an1)
L)(dim
K
K(a
1
, . . . , a
n1
)) <
0
.
Denition. Let K be a eld and f K[X] K.
1. Let a be a root of f (in some extension of K). The multiplicity of the root a
is the largest n Z
+
with (Xa)
n
[ f in K(a)[X]. (This exists since those
ns are bounded above by deg f by the additive degree formula. We will
expand on this observation in the next proposition.) We could equivalently
replace K(a) in this denition with any eld L K(a). (Exercise.) A
repeated root is a root of multiplicity at least 2.
2. We say f splits over K if it can be written as a product of linear (i.e., degree
1) polynomials in K[X]. Of course, any product of polynomials that split
over K splits over K. Also, any non-constant divisor of a polynomial that
splits over K splits over K. (This follows quickly from considering the
uniqueness of atomic factorizations in K[X].)
Example.
1. The polynomial X
3
+ X
2
X 1 = (X 1)(X + 2)
2
splits over 1. Its
roots are 1 and 2, which have multiplicities 1 and 2, respectively.
2. The polynomial X
4
1 = (X +1)(X 1)(X +i)(X i) splits over C but
not over 1. Its roots in 1 are 1, which each have multiplicity 1. In C it
has additional roots i, which also have multiplicity 1.
These examples suggest the following proposition relating splitting and multi-
plicities of roots.
146
Proposition 147. Let K be a eld and f K[X] K. Let a
1
, . . . , a
k
be the
roots of f in K, and let m
1
, . . . , m
k
be their multiplicities, respectively. Then
m
1
+ +m
k
deg f. Moreover, the following are equivalent.
1. f splits over K.
2. f = c(X a
1
)
m1
(X a
k
)
m
k
for some c K

.
3. m
1
+ +m
k
= deg f.
Proof. Noting that the (X a
i
)s are non-associate primes and that each (X
a
i
)
mi
[ f, we obtain (f) ((X a
1
)
m1
) ((X a
k
)
m
k
) = ((X
a
1
)
m1
(Xa
k
)
m
k
) from a past exercise, hence f = (Xa
1
)
m1
(Xa
k
)
m
k
g
for some g K[X]

, and deg f = m
1
+ +m
k
+ deg g m
1
+ +m
k
.
(1) (2): Assume f splits over K. Writing a linear factorization of f
and adjusting the factors by constant multiples as appropriate, we obtain a
factorization of the form f = c(Xa
1
)
n1
(Xa
k
)
n
k
for some c K

, where
by the denition of multiplicity each n
i
m
i
. In fact each n
i
= m
i
, since if some
n
i
< m
i
, then dividing both sides of (Xa
i
)
mi
[ f by (Xa
i
)
mi
yields Xa
i
[
(Xa
i
)
mini
[ c(Xa
1
)
n1
(Xa
i1
)
ni1
(Xa
i+1
)
ni+1
(Xa
k
)
n
k
, where
X a
i
is prime yet does not divide any of the factors on the right-hand side,
a contradiction. (2) (3): Clear. (3) (1): If m
1
+ + m
k
= deg f, then
from the above we see that deg g = 0, and f = ((X a
1
)g)(X a
1
)
m11
(X
a
2
)
m2
(X a
k
)
m
k
is a linear factorization.
Denition. Let K be a eld and f K[X] K. A splitting eld for f over K
is a minimal (with respect to inclusion) eld containing K over which f splits.
(In other words, a eld L K is a splitting eld for f over K if f splits over L
but not over any smaller eld extension of K.) If f splits over L and a
1
, . . . , a
n
are the roots of f in L, then K(a
1
, . . . , a
n
) is a splitting eld for f over K.
(Exercise.) So if L is a splitting eld for f over K, then L = K(a
1
, . . . , a
n
),
where a
1
, . . . , a
n
are the roots of f in L. Thus any splitting eld for f over K is
a nite dimensional (hence algebraic) extension of K. From the aforementioned
exercise it also follows that any eld extension of K over which f splits can be
shrunk to a splitting eld of f over K.
Theorem 148 (Existence and Uniqueness of Splitting Fields). Let K be a eld
and f K[X] K. Then there is a splitting eld for f over K, and all splitting
elds for f over K are isomorphic.
Proof. If deg f = 1, then K is a splitting eld for f over K, so let us assume
deg f 2. There is an extension of K in which f has a root a
1
, and we may
write f = (Xa
1
)g in K(a
1
)[X]. We have deg g < deg f, so by induction there
is a splitting eld L of g over K(a
1
). Let a
1
, . . . , a
n
be the roots of f in L. Then
the roots of g in L that are not already in K(a
1
) consist of some subcollection
of a
2
, . . . , a
n
, so L = K(a
1
)(a
2
, . . . , a
n
) = K(a
1
, . . . , a
n
), showing that L is a
splitting eld for f over K.
The fact that all splitting elds for f over K are isomorphic is a special case
of Lemma 150 part (3) below.
147
Lemma 149. Let : K L be a eld homomorphism. Extend to a map
K[X] L[X] by dening (f) to be the polynomial in L[X] obtained by applying
to the coecients of f. Then : K[X] L[X] is a ring homomorphism. It
is an epimorphism (resp., monomorphism) if and only if : K L is.
Proof. This follows very easily from Proposition 101.
Lemma 150. Let : K
1
K
2
be a eld isomorphism and f K
1
[X] K
1
.
1. The polynomial f splits over K
1
if and only if (f) splits over K
2
.
2. The polynomial f is irreducible if and only if (f) is. In this case, for any
roots c of f and d of (f) (in some extensions of K
1
and K
2
, respectively)
we have K
1
(c)

= K
2
(d) as rings via g(c) ((g))(d), where g K
1
[X].
3. Assume L
1
and L
2
are splitting elds for f and (f) over K
1
and K
2
,
respectively. Then can be extended to an isomorphism L
1
L
2
.
Proof.
1. (): If f = f
1
f
n
is a product of linear polynomials in K
1
[X], then so
is (f) = (f
1
) (f
n
) in K
2
[X]. (): Apply with
1
.
2. The rst statement is equivalent to proving that f is reducible if and only
if (f) is, which an be shown by replacing linear with non-constant
in the proof of (1). Now assume f is irreducible and c and d are roots of f
and (f), respectively. Recall the isomorphisms
c
: K[X]/(f) K(c) :
g + (f) g(c) and
d
: L[X]/((f)) L(d) : g + ((f)) g(d). Then

d

1
c
: K(c) L(d) is the stated isomorphism.
3. If deg
K1
L
1
= 1, then L
1
= K
1
and f splits over K, and takes a linear
factorization of f over K
1
to one of (f) over K
2
, hence L
2
= K
2
and
the result is clear. So let us assume deg
K1
L
1
2. Then L
1
,= K
1
, so
there is some root c of f in L
1
K
1
. Then m
c,K1
[ f in K
1
[X], say
f = m
c,K1
h. Because (f) splits over L
2
, the same holds for (m
c,K1
); let
d be one of its roots. Then by part (2) we may extend to an isomorphism
K
1
(c) K
2
(d). Because h splits over L
1
and the roots of f in L
1
are c
together with the roots of h, we conclude that L
1
is the splitting eld for
h over K
1
(c), and similarly L
2
is the splitting eld for (h) over K
2
(d).
Since 2 dim
K1(c)
L
1
(dim
K1
K
1
(c))(dim
K1(c)
L
1
) = dim
K1
L
1
is nite,
we conclude that dim
K1(c)
L
1
< dim
K1
L
1
, so by induction we may extend
: K
1
(c) K
2
(d) to an isomorphism L
1
L
2
.
Theorem 151 (Structure of Finite Fields). Every nite eld has prime-power
order. For a given prime power p
n
, the elds of order p
n
are precisely the
splitting elds of X
p
n
X over Z
p
. Thus there is a unique (up to isomorphism)
eld of order p
n
, which we denote F
p
n. Every element of this eld is a root of
X
p
n
X.
148
Proof. We have already proven the rst sentence.
Let f = X
p
n
X and F
p
n be a splitting eld for f over Z
p
. By a past
exercise a root of a polynomial is repeated if and only if it is also a root of the
derivative. In F
p
n[X], we have f

= p
n
X1 = 1 (recall that p
n
= 0 in Z
p
), so
f has no repeated roots, and by the multiplicity identity it must have exactly
p
n
distinct roots in F
p
n. To show that F
p
n has no more elements besides these
roots, it suces to show that the set of these roots forms a eld containing
Z
p
. We have 0
p
n
= 0, and for m Z

p
= Z

p
Lagranges Theorem shows that
m
p1
= 1, hence m
p
n
= mm
p
n
1
= m(m
p1
)
p
n1
+p
n2
++1
= m. The only
nontrivial detail left to check is that for roots a, b F
p
n of X
p
n
X with b ,= 0,
the elements ab, b
1
, and a b are roots of X
p
n
X. In this case a
p
n
= a and
b
p
n
= b, and we compute (ab)
p
n
= a
p
n
b
p
n
= ab, (b
1
)
p
n
= (a
p
n
)
1
= a
1
, and
(a b)
p
n
= a
p
n
+(1)
p
n
b
p
n
= a b (using the Freshmans Dream), as desired.
Now let K be any eld of order p
n
. To show that K is a splitting eld
for X
p
n
X over Z
p
, it will suce to show that each element of K is a root
of X
p
n
X. (This immediately shows that K is the eld generated by Z
p
and the roots of X
p
n
X in K, and the fact that this polynomial splits will
follow by adding multiplicities.) Certainly 0 is a root of X
p
n
X, and for each
a K

= K

we have a
p
n
1
= 1 by Lagranges Theorem, hence a
p
n
= a 1 = a,
as desired.
Example. We consider the simplest of the nite elds not of the form Z
p
,
namely F
4
. To describe it, we completely factor X
4
X = X(X
3
1) =
X(X1)(X
2
+X+1) over Z
2
. The elements of F
4
are roots of this polynomial,
so we have 0 and 1 and two roots of X
2
+ X + 1. Let denote one of these
roots. Then F
4
= 0, 1, , +1. (It can quickly be veried that these are four
distinct elements, so they must be all of F
4
.) We can add or multiply any two
elements of F
4
if we know what + and
2
are. The former is 0 since the
characteristic of F
4
is 2, and the latter is +1 since
2
++1 = 0. (Note that
the + and operations are the same in a eld of characteristic 2.)
Exercises.
1. Let K L be elds and f K[X] K. Let a L be a root of f and m
be its multiplicity. Show that m is the largest power of Xa that divides
f in L[X].
2. Let K L be elds and f K[X] K. Prove that the following are
equivalent.
(a) L is a splitting eld for f over K.
(b) f splits over L and L = K(a
1
, . . . , a
n
), where a
1
, . . . , a
n
are the roots
of f in L.
(c) f splits over some eld extension of L and L = K(a
1
, . . . , a
n
), where
a
1
, . . . , a
n
are the roots of f in that extension.
3. Let K L be elds. Show that K = L if and only if dim
K
L = 1.
Bibliography
[1] J.R. Durbin. Modern Algebra: An Introduction. 6 edition.
149

You might also like