You are on page 1of 37

Contents

1 Introduction to Rings

1.1

Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2

Subrings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2 Some Special Types of Rings

2.1

Integral Domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.2

Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.3

Characteristic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.4

The importance of the Zero Product Property in solving equations . . . . . . . . . . 13

3 Ring Homomorphisms and Ideals

14

3.1

Ring Homomorphisms and Isomorphisms

3.2

Ideals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

3.3

Quotient Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

3.4

The First Isomorphism Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

3.5

Prime and Maximal Ideals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

4 Polynomial Rings

. . . . . . . . . . . . . . . . . . . . . . . . 14

23

4.1

Basic Definitions and Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

4.2

The Division Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

4.3

Principal Ideal Domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

5 Factorization of Polynomials
5.1

Reducibility Tests

26

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

5.2

Reducibility over Z . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

5.3

Mod p Test and Eisensteins Criterion . . . . . . . . . . . . . . . . . . . . . . . . . . 29

5.4

Applications of irreducible polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . 31

6 Factorization in Integral Domains

31

6.1

Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

6.2

Examples in Z[ d] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

6.3

Unique Factorization Domains and Euclidean Domains . . . . . . . . . . . . . . . . . 34

6.4

Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

Index of Definitions

37

Introduction to Rings

1.1

Definitions

Example. Consider the integers. We know that Z with + is an abelian group. We cannot use
multiplication as a group operation because we dont have inverses for everything. However, were
used to using both addition and multiplication in the integers, and we even have a rule for how
they interact (the distributive law). In group theory we restricted ourselves to one operation. With
rings we get to add a second operation.
Definition 1. A ring R is a nonempty set with two binary operations, addition (denoted a + b)
and multiplication (denoted ab), such that for all a, b, c R:
(Adding is commutative) (1) a + b = b + a
(Adding is associative) (2) (a + b) + c = a + (b + c)
(Additive identity) (3) There is an element 0 R such that a + 0 = 0 + a = a.
(Additive inverses) (4) There is an element a R such that a + a = 0.
(Multiplying is associative) (5) a(bc) = (ab)c.
(Distributive Laws) (6) a(b + c) = ab + ac and (b + c)a = ba + ca
Note: The definition of binary operation includes closure. So saying that + and are binary
operations means that rings are closed under + and .
Note: A ring is an abelian group under the operation + (with additive identity 0) such that
multiplication is closed and associative and the distributive property holds.
Note: While rings must have an additive identity and every element must have an additive inverse,
rings do not need to have a multiplicative identity or multiplicative inverses. Moreover, while
addition in rings must be commutative, the multiplication need not be.
Definition 2. Let R be a ring.
The ring R is commutative if ab = ba for all a, b R.
An element 1 R is a unity if 1b = b1 = b for all b R.
If 1 is a unity and a R, then a is a unit if there is an element a1 R such that aa1 =
a1 a = 1.
We write a b to mean a + (b).

Note: The word abelian is not used for rings.


Note: If a ring R has a unity 1, then R is called a ring with identity or ring with unity. This will
often also be written as R is a ring with 1.
Example. Z, Q, R, and C with ordinary addition and multiplication are all rings. These are all
commutative rings with unity. In Z the units are 1 and 1. In Q, R, and C all nonzero elements
are units.
Example. The set Zn = {0, 1, . . . , n 1} under addition and multiplication modulo n is a commutative ring with unity. The units are the elements in U (n), i.e. they are the numbers that are
relatively prime to n.
Example. Let S be any of the rings Z, Q, R, C, Zm and n > 1. Then the set Mn (S) of n n
matrices with entries from S is a ring under standard matrix addition and multiplication. These
are noncommutative rings with unity. (The unity is the identity matrix with 1s down the diagonal
and 0s elsewhere.) The units of Mn (S) are the elements of GL(n, S).
Example. The set 2Z of even integers is a commutative ring that does not have unity.
Example. Let R1 , R2 , . . . , Rn be rings. The set R1 R2 Rn = {(a1 , a2 , . . . , an ) : a1
R1 , a2 R2 , . . . , an Rn } is a ring under component wise addition and multiplication. For example,
Z Z is a commutative ring with unity (1, 1). The units of Z Z are (1, 1), (1, 1), (1, 1), and
(1, 1). What are the units of Q Q?
Example. Let F (R) be the set of all functions f : R R. We can define addition and multiplication on F (R) as follows. For two functions f and g we define f + g to the be the function such that
(f + g)(x) = f (x) + g(x) and we define f g to the be the function such that (f g)(x) = f (x)g(x).
This is a commutative ring with unity. What is the unity? What are the units?
Example. The set Z[i] = {a + bi : a, b Z} is called the Gaussian Integers. This is a ring under
addition and multiplication of complex numbers. This is a commutative ring with unity 1. What
are the units?
Example. Define the set Z[x] to be the set of all polynomials with variable x and coefficients
from Z. Then Z[x] is a ring with ordinary addition and multiplication of polynomials. This is a
commutative ring with unity. What are the units?
Example. Lets compare the multiplication tables for Z5 and Z6 .
Z5 ,
0
1
2
3
4

0
0
0
0
0
0

1
0
1
2
3
4

2
0
2
4
1
3

3
0
3
1
4
2

4
0
4
3
2
1

Z6 ,
0
1
2
3
4
5

0
0
0
0
0
0
0

1
0
1
2
3
4
5

2
0
2
4
0
2
4

3
0
3
0
3
0
3

4
0
4
2
0
4
2

5
0
5
4
3
2
1

Notice that in Z5 if we multiply two nonzero elements the result is always nonzero. However, in
Z6 it is possible to multiply two nonzero elements and get 0. Also in Z5 every nonzero element is
a unit while in Z6 only 1 and 5 are units.
5

Definition 3. Let R be a ring.


An element a R is called a zero divisor if a 6= 0 and for some b 6= 0, ab = 0 or ba = 0.
An element a R is called idempotent if a2 = a.
An element a R is called nilpotent if ak = 0 for some positive integer k.
If a, b R and a 6= 0, then a divides b, denoted a|b, if b = ac for some c R.
Note: We will usually only consider zero divisors in commutative rings, so we will be able to say
that a is a zero divisor if a 6= 0 and ab = 0 for some b 6= 0.
So in Z5 , there are no zero-divisors, while in Z6 , the numbers 2 and 3 are zero-divisors. In Z6 the
idempotents are 0, 1, 3, 4 and in Z5 the idempotents are 0, 1. Neither ring has any nilpotents. Some
examples of nilpotents are 3 in Z9 or 2 Z8 or 6 Z12 . In Z6 the number 4 divides both 2 and 4,
but does not divide 3. The rings Z, Q, R, C do not have any zero divisors.
Note: The fact that x and y are not zero is a very important part of the definition of zero divisor.
So if youre trying to prove that an element is a zero divisor, the first thing you need to do is show
that the element is nonzero.
Notation:
For general rings we will use 0 to denote the additive identity and 1 to represent the multiplicative
identity if there is one.
We will use a to represent the additive inverse of a.
We will abbreviate a
| +a+
{z + a} by n a or na when it is clear. Note that n a is not the product
n times

of n and a as n is not an element of most rings.


Theorem 1.1. Let a, b, c R. Then
1. a0 = 0 and 0a = 0.
2. a(b) = (ab) and (a)b = (ab).
3. (a)(b) = ab
4. a(b c) = ab ac and (b c)a = ba ca
Furthermore, if R has a unity element 1, then
5. The unity is unique.
6. (1)a = a
7. (1)(1) = 1
6

Proof.
1. Let a R. Then a0 = a0 + 0, by axiom (3). Also, a0 = a(0 + 0) = a0 + a0 by axiom
(3) and (6). By transitivity, a0 + 0 = a0 + a0. Adding a0 to both sides will cancel one a0
from each side, so we get 0 = a0. Similarly, we can show 0a = 0.
2. Let a, b R. Then ab + ab = 0 = a0 = a(b + b) = ab + a(b). By cancelling ab from each
side, we get ab = a(b). Similarly, we can show (a)b = ab.
3. Let a, b R. Then
0 = (a)(b b) = (a)b + (a)(b) = ab + (a)(b),
and hence ab = (a)(b).
4. Let a, b R. Then a(b c) = a(b + c) = ab + a(c) = ab + ac = ab ac
5. Suppose 1 and e are both unities. Then e1 = e and e1 = 1, so 1 = e.
6. Use property (2).
7. Use property (3).

Some dangers regarding how rings are different from groups: You cant always use things like a1
because you dont know if multiplicative inverses exist. The ring might not even have a unity, and
even if it does, only the units have inverses. In particular, there is no more cancellation law for
multiplication. You can cancel by subtracting the same thing from both sides, but not by dividing.
For example, in R = Z10 , if 4 + x = 6, you can subtract 4 from both sides to get x = 2. But if
you have 4x = 8, you cant necessarily cancel 4 and say x = 2. For example, here it could be that
x = 7 as well. To summarize:
a + b = a + c implies b = c
ab = ac does NOT imply b = c.
If a is a unit, then you can multiply on both sides by a1 , so you can cancel the a. But in general,
unless you are given that a is a unit, proceed with caution!!

1.2

Subrings

In groups we looked at subsets that were also groups; these were called subgroups. Similarly, we
can look at subrings.
Definition 4. A nonempty subset S of a ring R is a subring of R if S is a ring with the same
operations as R.
Theorem 1.2 (Subring Test). A nonempty subset S of a ring R is a subring if S is closed under
subtraction and multiplication. In other words, a nonempty subset S of R is a subring if

1. a b S for all a, b S and


2. ab S for all a, b S.
The Subring Test works because clause (1) implies that S under addition is a subgroup of R, using
the 1-step test for subgroups. And clause (2) implies that multiplication is a binary operation on
S. Axioms (5) and (6) of rings will follow, just because they are true in R.
Example.
1. {0} and R are always subrings of R. (The ring {0} is called the trivial subring.)
Thus every ring has at least two subrings.
2. R = Z8 . Let S = {0, 2, 4, 6}. The difference or product of even numbers mod 8 is again even,
so S is a subring by the Subring Test. Notice that R has a unity element 1, but S does not
contain any unity.
3. R = Z9 . Let S = {0, 2, 4, 6, 8}. Then S is not a subring, because it is not closed under
addition or multiplication. For example, 6 + 4 = 1 and 6 2 = 3.
4. S = {0, 2, 4, 6, 8} is a subring of Z10 . Is this a ring with unity? Yes! 6 is the unity of S. This
shows that the unity of a subring need not be the same as the unity of the larger ring.
5. 3Z is a subring of Z. So a ring with unity can have a subring that does not have unity.
6. Z[i] is a subring of C.
Example. Ring or not a Ring?

1. Q[ 3] = {a + b 3 : a, b Q}
We could check all of the ring properties; however, its much easier to noticethat this is
a subset of R, so we just
need to
check if its
a subring of R. First, 0 Q[ 2], so it is
nonempty.
Now
let
a
+
b
2,
c
+
d
2

Q[
2] for some rational numbers a, b, c, d. Then

a + b 2 c d 2 =
(a c) +
(b d) 2. Since a c, b d Q, thedifference is an element
of Q[ 2].And (a + b 2)(c + d 2) = (ac + 2bd) + (ad + bc) 2 Q[ 2]. So the Subring Test
shows Q[ 2] is a subring of R.
2. R = {ri : r R, i2 = 1}.
Notice that this is a subset of C. However, this is not a ring since 3i and 5i are in R, but
15 = (3i)(5i) is not in R. Thus R is not closed under multiplication and is not a ring.

2
2.1

Some Special Types of Rings


Integral Domains

Recall that a zero-divisor is a nonzero element a in a ring R such that ab = 0 for some nonzero
element b R. We will focus almost entirely on zero-divisors in commutative rings, so that if
ab = 0, where a, b 6= 0, then both a and b are zero-divisors.
8

Example. In Z6 , we have 2 3 = 0, so 2 and 3 are both zero-divisors.


Definition 5. An Integral Domain is a commutative ring with unity that contains no zero divisors.
The following proposition gives an equivalent characterization of Integral Domains. The proposition
also gives us a way to prove that a ring is an Integral Domain.
Proposition 2.1 (Zero Product Property). A commutative ring with unity is an integral domain
if and only if whenever ab = 0, then a = 0 or b = 0.
This characterization is called the Zero Product Property.
Example. Integral Domain or Not an Integral Domain?

1. Z, Q, R, C
Integral Domain
2. M2 (R)
Not an integral domain since it is not commutative. Moreover, M2 (R) contains zero divisors.
For example,

 


0 0
0 0
1 0
.
=
0 0
0 1
0 0
3. Zp where p is prime.
For any prime p, the ring Zp is an integral domain. Why is this? Well suppose that ab = 0 in
Zp . This means that p divides ab. However, since p is prime, this means that p divides a or
p divides b, so either a or b is actually 0 in Zp . This shows that the only way that a product
can be 0 is if at least one of the factors is zero, which is the same as saying that there are no
zero divisors. Clearly the ring is commutative with 1.
4. Zn where n is not a prime.
Not an integral domain since there are zero divisors. If n is not prime, then n = st where
1 < s, t < n. Then s, t Zn and st = 0 even though s 6= 0 and s 6= 0.
5. 3Z
This is a commutative ring with no zero divisors, but it is not an integral domain since there
is no unity.
6. Z[i]
This is an integral domain. First note that the ring is commutative with 1. Also, it is a
subset of C which does not have any zero divisors.
7. Z Z
This is not an integral domain. Notice that (0, 5) is a zero divisor since (0, 5)(1, 0) = (0, 0)
even though neither (0, 5) nor (1, 0) is the zero element.

In groups we were allowed to use left and right cancelation. Can we cancel (multiplicatively) in
rings?
Example. Find an example of a ring R and elements a, b, c R such that ab = ac but b 6= c.
Consider R = Z10 and a = 5, b = 2, c = 4. Then ab = 0 = ac but b 6= c. This shows that, in general,
cancelation does not work in rings.
There is, however, a case in which cancelation always works.
Theorem 2.2 (Cancellation in an ID). Suppose that R is an integral domain and a, b, c R with
a 6= 0. Then ab = ac implies b = c.
Proof. Suppose that R is an integral domain, a 6= 0, and ab = ac. We need to show that b = c.
Note that ab ac = 0 and by the distributive property, a(b c) = 0. Since R does not contain zero
divisors and a 6= 0 we must have b c = 0 which implies that b = c.

2.2

Fields

Definition 6. A division ring is a ring with unity in which every nonzero element is a unit.
Definition 7. A commutative ring with unity in which every nonzero element is a unit is called a
field.
Note: A field is just a commutative division ring.
In a field, every nonzero element has a multiplicative inverse. One way to think of a field is that it
is an algebraic system that is closed under +, , , (except dividing by zero). The three classic
examples of fields are R, Q, and C.
Example. C, Q, R are all fields. Also Z5 is a field. (Check!) In fact, Zp is a field for any prime p.
The ring Z6 is not a field since the element 2 does not have a multiplicative inverse.

Example. Consider the set Q( 3) = {a + b 3 : a, b Q}. It is not hard to show that this
set is closed under subtraction and multiplication and hence it is a subring of R. Is it an integral
domain? Well it is commutative with unity. What about zero divisors? Since this is a subset
of R
it cannot have zero divisors. Is it a field?
We just need
to
check
for
inverses.
Suppose
a
+
b
3 6= 0.

1
1
ab 3
ab 3
a
b
=

= 2
Now consider
= 2

3 Q( 3). Thus
a 3b2
a 3b2 a2 3b2
a+b 3
a+b 3 ab 3
every nonzero element has an inverse and this is a field.
You may have noticed that all of our examples of fields are also integral domains. This is not a
coincidence.
Theorem 2.3. Every field is also an integral domain.

10

Proof. Suppose that R is a field. This means that it is commutative and has unity, so the only
thing we need to show is that R does not have any zero divisors. In other words, we need to show
that the only way that ab can equal 0 is if a or b is 0. So suppose that a, b R with ab = 0.
Suppose that a 6= 0. Then since R is a field, a is a unit and a1 R. Thus b = a1 ab = a1 0 = 0.
This shows that either a or b is 0, and hence R has no zero divisors and is an integral domain.
Is the converse true? Is every integral domain a field? No! Notice that Z is an integral domain,
but it is not a field since 2 does not have an inverse.
There is a case, however, in which Integral Domain does imply Field. Notice that for Zn , we have
a field if and only if n is prime. This is also exactly when Zn is an integral domain. The key here
is that Zn is finite.
Theorem 2.4. Let R be a finite integral domain. Then R must also be a field.
Proof. If R is an integral domain, then it is commutative with unity, so we just need to show that
every nonzero element is a unit. Let a R with a 6= 0. Also let R represent the set of nonzero
elements in R. Consider the map a : R R defined by a (r) = ar. First note that this is
actually a map from R to R since if a and r are not 0, then ar 6= 0 since R has no zero divisors.
Now we show that a is one-to-one. Suppose that a (r) = a (s). Then ar = as and since a 6= 0
and R is an integral domain, we can cancel which yields r = s. Thus a is a one-to-one map from
the finite set R to itself. Any one-to-one map from a finite set to itself must also be onto. Since
1 R , there must be some x R such that a (x) = 1, i.e. ax = 1. This means that a is a unit,
and hence R is a field.
An alternative proof:
Proof. Let R be a finite integral domain. Then it is commutative with unity, so we just need to show
that every nonzero element of R is a unit. Let a R with a 6= 0. Now consider the list of elements
a, a2 , a3 , a4 , . . .. By closure of multiplication, all of these elements must be in R, and since R is
finite, they cannot all be different. Thus ak = am for some k > m. Now akm am = ak = am = 1am .
Since R is an integral domain and a 6= 0, then am 6= 0. Thus by Theorem 2.2, we can cancel to get
akm = 1. If k m = 1, then a = 1, so a is a unit. If k m > 1, then akm1 a = 1, and hence a
is a unit. Thus every nonzero element of R is a unit and R is a field.
Since we have shown that Zp is an integral domain whenever p is prime, theorem 2.4 shows that
Zp is always a field when p is prime.
Corollary 2.5. Let p be a prime number. Then Zp is a field.

2.3

Characteristic

Definition 8. Let R be a ring. We define the characteristic of R to be the least positive integer
n such that nr = 0 for all r R. If no such n exists, then we say that the characteristic of R is 0.
11

We will denote the characteristic of R by char(R).


Example. char(Z) = 0 and char(Z7 ) = 7.
In the case that R has a unity, it is a lot easier to find the characteristic.
Theorem 2.6 (Characteristic of a ring with unity). Let R be a ring with unity 1. Then
1. charR = 0 if 1 has infinite order under addition.
2. charR = n if 1 has order n under addition.
Proof. Suppose R has unity 1.
1. Assume 1 has infinite order under addition. Then for all n, n 1 6= 0. Thus charR = 0.
2. Assume 1 has order n under addition. Let a R. Then
n a = a + a + + a

(n times)

= a1 + a1 + + a1

(n times)

= a(1 + 1 + + 1)

(n times)

= a(n 1)
= a0 = 0.
Thus n is a positive integer such that n a = 0 for all a R. Furthermore, no smaller integer
will work for 1, since the order of 1 under addition is n. So charR = n.

Example. Consider the characteristic of integral domains we know. For instance, the characteristics of Z, Q, R, and C are all 0. The characteristics of Z5 , Z7 , Z11 are 5, 7, and 11 respectively.
Example. Lets look at rings of the form Zn [i] = {a + bi : a, b Zn }.
First consider Z2 [i] = {0, 1, i, 1 + i}. Note that (1 + i)(1 + i) = 1 + 2i 1 = 0. So Z2 [i] contains
zero divisors and is hence neither an Integral Domain nor a field.
Now well look at Z3 [i] = {0, 1, 2, i, 2i, 1 + i, 1 + 2i, 2 + i, 2 + 2i}. Both 1 and 2 are their own inverses.
The elements i and 2i are inverses of each other, 1 + i and 2 + i are inverses of each other, and
1 + 2i and 2 + 2i are inverse of each other. Thus Z3 [i] is a field and hence an integral domain. The
characteristic of this ring is 3.
Clearly Z4 [i] is not an integral domain since it contains the zero divisor 2. Thus Z4 [i] is not a field.
What about Z5 [i]. Consider the elements 1+2i and 1+3i. Note that (1+2i)(1+3i) = 1+2i+3i+4 =
0. Thus Z5 [i] contains zero divisors, which means that it is not an integral domain, which means
that it is not a field.
12

You may have noticed that for all of the integral domains we have looked at, the characteristic has
been either 0 or prime. This is not a coincidence.
Theorem 2.7. Let R be an integral domain. Then the characteristic of R is either 0 or p where p
is a prime.
Proof. Let R be an integral domain and suppose that the characteristic of R is not 0. Then
char(R) = n for some positive integer n. Suppose that n is not prime. Then n = st for some
positive integers s, t with s < n and t < n. Since R is an integral domain, it must have a 1. Then
(s 1)(t 1) = (1 + 1 + + 1) (1 + 1 + + 1)
|
{z
}|
{z
}
2

s times
2

t times

= (1 + 1 + + 1 )
|
{z
}
st times

= st 1 = st 1 = n 1 = 0
So (s 1)(t 1) = 0. Since R is an integral domain, either s 1 = 0 or t 1 = 0. Without loss of
generality, s 1 = 0. Thus the additive order of 1 is less than or equal to s. However, by Theorem
2.6, the order of 1 is n, so we have a contradiction. Thus we cannot factor n as st with s and t less
than n. Consequently, n must be prime.

2.4

The importance of the Zero Product Property in solving equations

You use the ZPP all the time in high school algebra.
Example. Consider the equation x2 x 6 = 0. Find all solutions
1. in the integers Z.
x2 x 6 = 0
(x 3)(x + 2) = 0
x 3 = 0 or x + 2 = 0

because Z is an ID

x = 3, 2
2. in the integer mod 13 Z13 .
x2 x 6 = 0
(x 3)(x + 2) = 0
x 3 = 0 or x + 2 = 0

because Z13 is an ID

x = 3, 2
x = 3, 11

13

3. in the integers Z12 .


x2 x 6 = 0
(x 3)(x + 2) = 0

Here we get stuck. We cant break it into two simpler equations, because Z12 is not an ID.
For example, instead of having one factor be zero, it could be that one factor is 3 and the
other is 4, because 3 and 4 are zero-divisors. In the end, we just have to try all 12 numbers
and list the ones that work: x = 3, 6, 7, 10.
Example. Find all elements in an integral domain which are their own inverses.
Suppose D is an ID and x = x1 . Then x2 = 1, so x2 1 = 0. Factoring gives us (x 1)(x + 1) = 0.
By the ZPP, x 1 = 0 or x + 1 = 0. Thus x = 1. The only elements which are their own inverses
in an ID are 1.

3
3.1

Ring Homomorphisms and Ideals


Ring Homomorphisms and Isomorphisms

Definition 9. Suppose that R and S are rings. A map : R S is called a ring homomorphism
if (a + b) = (a) + (b) and (ab) = (a)(b) for all a, b R. A bijective ring homomorphism is
called a ring isomorphism. The kernel of a ring homomorphism : R S is the set Ker = {r
R : (r) = 0}. If an isomorphism : R S exists, we say R is isomorphic to S, which we
denote by R
= S.
Example. For any n N, we can define a ring homomorphism : Z Zn by (x) = x (mod n).
Note that (a + b) = a + b mod n = a mod n + b mod n = (a) + (b)
and (ab) = ab mod n = (a mod n)(b mod n) = (a)(b).
The kernel of this map is nZ.
Example. Consider : C C defined by (a + bi) = a bi.
((a + bi) + (c + di)) = (a + c + (b + d)i) = a + c (b + d)i = a + c bi di = a bi + c di =
(a + bi) + (c + di)
((a + bi)(c + di)) = (ac bd + (ad + bc)i) = ac bd (ad + bc)i = ac bd bci adi =
(a bi)(c di) = (a + bi)(c + di)
So is a ring homomorphism.
Example. Consider : Z[x] Z defined by (f (x)) = f (0).
(f (x) + g(x)) = ((f + g)(x)) = (f + g)(0) = f (0) + g(0) = (f (x)) + (g(x))
(f (x)g(x)) = ((f g)(x)) = (f g)(0) = f (0)g(0) = (f (x))(g(x))
14

Thus is a ring homomorphism. The kernel of is the set of polynomials in Z[x] that have no
constant term. There is nothing special about plugging 0 in. We could have replaced zero with any
integer and we would still have a homomorphism.
In general if R[x] is the ring of polynomials whose coefficients come from R and a R, then the
map a : R[x] R defined by a (f (x)) = f (a) is a ring homomorphism. Maps of this form are
called evaluation homomorphisms or evaluation maps.
Example. Consider : Q[x] Q given by (an xn + an1 xn1 + + a1 x + a0 ) = an + an1 +
+ a1 + a0 . It would be really annoying to check that this is a homomorphism. However, notice
that (f (x)) is actually just f (1), so this is actually an evaluation homomorphism.
Example. Consider : Z 2Z defined by (x) = 2x. This is a group isomorphism; is it a ring
isomorphism? Note that (2)(3) = 4(6) 6= 12 = (2 3) so it is not a ring homomorphism. Is it
possible that there is another map from Z 2Z that is a ring isomorphism? There cant be since
Z has unity and 2Z does not. In particular, if : Z 2Z were a ring isomorphism, then (1) = 2a
for some a Z. Thus 2a = (1) = (1 1) = (1)(1) = (2a)(2a) = 4a2 . Consequently, 2a = 4a2
or 2a(2a 1) = 0 which means that a = 0 or a = 21 . However, 12
/ Z, so we must have that a = 0.
If this were true, then (1) = 0, which means that (x) = (1x) = (1)(x) = 0(x) = 0 for all
x Z. Thus the only ring homomorphism from Z to 2Z is the zero map, which is not onto and
therefore not an isomorphism.
Theorem 3.1. Let : R S be a ring homomorphism. Then
1. If R is commutative, then (R) is commutative.
2. (0) = 0
3. If R and S both have unity and is onto, then (1R ) = 1S
4. For any r R and n N, (nr) = n(r) and (rn ) = ((r))n
5. is one-to-one if and only if the kernel of is {0}.
6. If a, b R, then (a b) = (a) (b).
Proof. Let : R S be a ring homomorphism.
1. Suppose that R is commutative and let x, y (R). Then x = (a) and y = (b) for some
a, b R. Now xy = (a)(b) = (ab) = (ba) = (b)(a) = yx. Thus (R) is commutative.
2. We know that group homomorphisms map the identity to the identity. Both R and S are groups
under addition with identity 0.
3. Suppose that R and S have unities 1R and 1S respectively. Suppose also that is onto. Let
x = (1R ) and let s S. Then since is onto, there is an element r R such that (r) = s. Now
xs = (1R )(r) = (1R r) = (r) = s. Similarly xs = s. Thus x is a multiplicative identity for S,
and since unities are unique, we must have that x = 1S .
4. Exercise. (Use induction.)
5. See proof from Group Theory.
6. This follows from the fact that is a group homomorphism (when we view R as an abelian
group under addition).
FACT: If R and S are isomorphic as rings then
1. R is commutative if and only if S is commutative.
2. R has unity if and only if S has unity.
15

3. R is a field if and only if S is a field.


4. R is an integral domain if and only if S is an integral domain.
etc.

3.2

Ideals

In groups, the kernel of a homomorphism was not only a subgroup of the domain, it was a normal
subgroup. What is the normal subgroup analog for rings? Well start with an example.
Example. Consider Z and the subring 3Z. Note that like all subrings, 3Z is closed under addition
and multiplication. However, there is actually a stronger version of multiplicative closure. Notice
that if we take any element of 3Z and any element of Z (in 3Z or not), the product is always back
in 3Z. This type of subring is called an ideal.
Definition 10. Let R be a ring and I a nonempty subset of R. Then I is called an ideal of R if I
is a subring of R and rx, xr I for all r R and all x I.
Example.

1. For any ring R, {0} and R are always ideals.

2. nZ is an ideal of Z for all n Z.


3. Let R = Z[x], I = {f (x) Z[x] : f (0) = 0} (the set of polynomials with no constant term),
and J = {g(x) Z[x] : g(0) 2Z} (the set of polynomials with even constant term). Both
I and J are ideals for R.
4. Z is a subring of Q. Is it an ideal? No! Notice that 2 Z and

1
3

Q but 2

1
3

/ Z.

Theorem 3.2 (Ideal Test). A nonempty subset I of a ring R is an ideal if


1. a b I for all a, b I and
2. ra, ar I for all r R and all a I.
Proof. Exercise
Theorem 3.3. Let : R S be a ring homomorphism. Then the kernel of is an ideal of R.
Proof. Let K represent the kernel of . We know that (0) = 0 and hence 0 K. Thus K 6= .
Now suppose that a, b K. Then (a) = 0 = (b). Now (a b) = (a) (b) = 0 0 = 0, and
hence a b K. Also if x R and a K, then (xa) = (x)(a) = (x)0 = 0, which shows that
xa K. Similarly, ax K. Thus K is an ideal of R.
Theorem 3.4. Let R be a ring with unity and let I be an ideal of R. If 1 I, then I = R.
Proof. Suppose that R is a ring with unity, I is an ideal of R, and 1 I. Let x R. Then
x = 1x I, and hence R I. We already know that I R, and consequently I = R.
Corollary 3.5. Suppose that R is a ring with unity, I is an ideal of R, and I contains a unit of
R. Then I = R.
16

Proof. Suppose that R is a ring with unity, I is an ideal of R, and a I for some unit a of R.
Then a1 R exists with aa1 = 1. Thus 1 = aa1 I, and by Theorem 3.4, I = R.
Example. What are the ideals of Q?
Let I 6= {0} be a nontrivial ideal of Q. (Note a nontrivial ideal is one that is not equal to {0}.)
Then there is some a I such that a 6= 0. Then a1 is in Q. Since I is an ideal, 1 = a a1 I.
However, now that 1 is in I, that means that everything in Q must be in I since if x Q, then
x = x 1 I. Thus {0} and Q are the only ideals of Q.
Theorem 3.6. Let F be a field. Then the only ideals of F are {0} and F .
Proof. Let F be a field and I 6= {0} an ideal of F . Then there exists an element a F with a 6= 0.
Since F is a field, a is a unit. Thus I contains a unit and hence I = F . Thus {0} and F are the
only ideals of F .
Example. Let R be a commutative ring with unity and a R. Consider the set hai = {ar : r
R}. Lets check that this is an ideal of R. First note that hai is nonempty since a = a1 hai. Now
let b, c hai. Then b = ar1 and c = ar2 for some r1 , r2 R. Now bc = ar1 ar2 = a(r1 r2 ) hai
since r1 r2 R. Also if x R, then xb = bx = ar1 x = a(r1 x) hai since r1 x R. Thus by the
ideal test, hai is an ideal of R.
Definition 11. Let R be a commutative ring with unity. Then the ideal hai = {ar : r R} is
called the principal ideal generated by a. An ideal I is R is said to be principal if there is some
b R such that I = hbi.
Theorem 3.7. Let R be a commutative ring with unity and a R. Then hai is an ideal of R.
Proof. We proved this above.
Theorem 3.8. Every ideal of Z is principal.
Proof. When we just look at Z as a group under addition, we know that Z is cyclic, and hence all
of its subgroups must be cyclic. Thus all subgroups of Z look like nZ for some n Z. Now every
ideal of Z is a subring of Z and therefore must be a subgroup of Z. Thus the ideals must all be of
the form nZ = hni.
Example. We have looked at another principal ideal. In Z[x] consider I = {f (x) Z[x] : f (0) =
0}. Since these are the polynomials with no constant term, they all have the property that we can
factor an x out of them. Thus I is actually equal to hxi. So I is a principal ideal. What about
the ideal J of polynomials whose constant term is even; is that a principal ideal? Suppose that
J = hg(x)i for some polynomial g(x) Z[x]. We know that 2 is in J so we must have 2 = g(x)k(x)
for some k(x) Z[x]. The only way this can happen if g(x) and k(x) are each 1 or 2. Since
g(x) J, it must have an even constant term. So we must have g(x) = 2. Now the polynomial x is
also in J so we must have x = 2r(x) for some r(x) Z[x]. This is impossible since the coefficients
of r(x) must be integers and the coefficient of x is not even. Thus J cannot be a principal ideal.

17

3.3

Quotient Rings

In groups we used normal subgroups to create factor/quotient groups. Similarly for rings, we will
use ideals to create factor or quotient rings.
Let R be a ring and I an ideal of R. Notice that Since R is an abelian group under addition and
all ideals are subgroups, I must be a normal subgroup of R. Thus in terms of groups we already
know what the set R/I looks like. In particular, R/I is the set of cosets of the form a + I where
a R. To make this into a ring we need to define a multiplication on the set of cosets. The natural
multiplication is (a + I)(b + I) = ab + I. Is this well defined? In other words if a + I = a0 + I and
b + I = b0 + I, are (a + I)(b + I) and (a0 + I)(b0 + I) equal?
Suppose that a + I = a0 + I and b + I = b0 + I. Then a0 = a + x and b0 = b + y for some x, y I.
Now a0 b0 = (a + x)(b + y) = ab + ay + xb + xy. Note that since I is an ideal, ay + xb + xy I.
Thus a0 b0 ab + I and hence a0 b0 + I = ab + I (recall that if two cosets have even one element in
common then they must be equal) and hence (a + I)(b + I) = (a0 + I)(b0 + I). It is not difficult to
show that this multiplication is associative and satisfies the distributive property. Thus R/I is a
ring with addition (a + I) + (b + I) = a + b + I and multiplication (a + I)(b + I) = ab + I.
When working with quotient rings, it is helpful to recall some facts about cosets from group theory.
Lemma 3.9 (Facts about Cosets). Let R be a ring and I an ideal of R. Let r, s R. Then
1. r + I = I if and only if r I;
2. r r + I;
3. If s r + I then r + I = s + I; and
4. r + I = s + I if and only if r s I.
Example. Z/5Z = {0 + 5Z, 1 + 5Z, 2 + 5Z, 3 + 5Z, 4 + 5Z}. If you make tables for the multiplication
and addition youll see that both of them are essentially addition and multiplication modulo 5. It
seems reasonable that Z/5Z is ring isomorphic to Z5 . Is there a reasonable way to prove this?
Example. Consider the ring Z Z and the ideal Z {0}. What can we say about the ring
(Z Z)/(Z {0})?
The elements look like (a, b) + Z {0}. Remember that (a, b) + Z {0} = (c, d) + Z {0} if
(a, b) (c, d) Z {0} which happens when a b Z and c d {0}. For this to happen, a and b
can be anything, but we must have c = d. Thus for any fixed integer n, the cosets (y, n) + Z {0}
are the same for all y Z. So every coset can be written as (0, n) + Z {0} for some n Z. Adding
two of these cosets: ((0, n) + Z + {0}) + ((0, m) + Z {0}) = (0, n + m) + Z + {0}. Multiplying
two of these cosets: ((0, n) + Z + {0})((0, m) + Z {0}) = (0, nm) + Z {0}. This looks a lot like
addition and multiplication in Z. It seems reasonable that (Z Z)/(Z {0}) is ring isomorphic to
Z. How can we prove this?
Example. Consider the ring R[x] and the ideal I = hx 1i. The elements in I look like f (x)(x 1)
where f (x) is a polynomial with real coefficients. So everything in I has 1 as a root. In fact we
know that if a polynomial has 1 as a root then we can factor an x 1 out of it, so in fact I is exactly
18

the polynomials that have 1 as a root. What does R[x]/I look like? The elements are cosets of
the form g(x) + I. Using long division of polynomials we can write g(x) as g(x) = q(x)(x 1) + r
where the remainder is a constant (since the degree of the remainder will need to be less than the
degree of (x 1). So g(x) + I = q(x)(x 1) + r + I. Recall that if J is an ideal and w J, then
w + J = J. We know that q(x)(x 1) I and hence g(x) + I = q(x)(x 1) + r + I = r + I. Thus
the elements in R[x] can all be written as r + I where r R. It seems reasonable to guess that
R[x]/I
= R.
Example. Consider the ring R[x] and the ideal I = hx2 + 1i. The elements in R[x]/I look like
f (x)+I where f (x) R[x]. By using long division of polynomials, f (x) = q(x)(x2 +1)+r(x) where
q(x), r(x) R[x] and r(x) has degree 1 or 0. Then f (x) + I = q(x)(x2 + 1) + r(x) + I = r(x) + I.
So we only have to look at cosets of the form ax + b + I where a, b R. Also, if we multiply
two cosets ax + b + I and cx + d + I we get (ax + b)(cx + d) + I = acx2 + (bc + ad)x + bd + I =
ac(x2 + 1) + (ad + bc)x + bd ac + I = (ad + bc)x + (bd ac) + I. Notice that if we multiply the
complex numbers b + ai and d + ci we get bd ac + (ad + bc)i. So it seems like the cosets behave
a lot like the elements of C and we guess that R[x]/I
= C.

3.4

The First Isomorphism Theorem

In general showing that a quotient ring is isomorphic to another ring can be difficult because
defining homomorphisms on the quotient ring can be awkward. The First Isomorphism theorem
helps with this.
Theorem 3.10 (The First Isomorphism Theorem). Suppose that : R S is an onto ring
homomorphism and let K represent the kernel of . Then R/K is ring isomorphic to S.
Proof. Suppose that : R S is an onto ring homomorphism and let K be the kernel of . We
need to define a map from R/K to S and then show that map is an isomorphism. Define by
: R/K S
r + K 7 (r)
Before we can check whether is a homomorphism we must check that it is well defined. What
does well defined mean? Remember that cosets can have lots of different names. For to be a
function we need for it to send any element in R/K to a single element of S. This means that we
need to make sure that no matter what name we use for an element of R/K that of that element
is the same.
well defined:
Suppose that r + K = w + K. We need to show that (r + K) = (w + K). Since r + K = w + K
we know that w = r + k for some k K. Now (w + K) = (w) = (r + k) = (r) + (k) =
(r) + 0 = (r) = (r + K), so is well defined.
homomorphism:
Let r + K, v + K R/K. Then ((r + K) + (v + K)) = (r + v + K) = (r + v) = (r) + (v) =
(r + K) + (v + K) and
((r + K)(v + K)) = (rv + K) = (rv) = (r)(v) = (r + K)(v + K). Thus is a homomorphism.
19

one-to-one:
Suppose that (r + K) = (v + K). Then (r) = (v) and (r) = (v) = 0. From this we see that
(r v) = 0 and hence r v K. Thus r + K = v + K and is one-to-one.
onto:
Let s S. Since is onto, there is an element r R such that (r) = s. Thus (r + K) = (r) = s
and is onto.
This shows that is a ring isomorphism and that R/K
= S.

Example. Let R = Z Z and I = Z {0}. We will use the first isomorphism theorem to show
that R/I
= Z. First we need to find an onto homomorphism from Z Z to Z. To help guide us in
finding this homomorphism, recall that we want the kernel to be Z {0}. Define
:ZZ Z
(m, n) 7 n
First we need to make sure that this is an onto homomorphism. Note that ((m, n) + (s, t)) =
(m + s, n + t) = n + t = (m, n) + (s, t) and ((m, n)(s, t)) = (ms, nt) = nt = (m, n)(s, t).
Thus is a ring homomorphism. Is it onto? Let d Z. Then (0, d) Z Z and (0, d) = d. Thus
is onto. What is the kernel of ? Note that (m, n) = 0 if and only if n = 0, so elements of the
kernel look exactly like (m, 0) where m Z. Thus K = Z {0}. By the first isomorphism theorem,
(Z Z)/(Z {0})
= Z.
Example. Now lets show that R[x]/hx 1i
= R.
Consider : R[x] R given by (f (x)) = f (1). This is an evaluation map so we know that
it is a ring homomorphism. Is it onto? Given a R consider the polynomial f (x) = a. Then
(f (x)) = f (1) = a so is onto. What is the kernel of ? Well (f (x)) = 0 if and only if f (1) = 0
if and only if 1 is a root of f (x) if and only if (x 1) is a factor of f (x). Thus the kernel of is
hx 1i. Now by the first isomorphism theorem, R[x]/hx 1i
= R. This shows that R[x]/hx 1i is
a field. Is R[x] a field? It is not. Note that the element x is nonzero but does not have an inverse
since when we multiply two nonzero polynomials the degree of the result is the sum of the degrees
of the two polynomials in the product.

3.5

Prime and Maximal Ideals

Example. Lets compare the ideals 3Z and 4Z of Z.


Suppose that a, b Z and ab 3Z. This means that 3 divides ab. Since 3 is prime, we must have
3|a or 3|b, and hence a 3Z or b 3Z. If ab 4Z must we have that a or b is in 4Z? No! Consider
2 and 6. Neither 2 nor 6 is in 4Z but 2 6 = 12 is in 4Z.
Definition 12. Let R be a commutative ring. An ideal P of R is called a prime ideal if P is a
proper ideal of R and ab P implies that a P or b P .
Example. 3Z is a prime ideal of Z and 4Z is not a prime ideal of Z. In general, if p is prime, pZ
is a prime ideal. Why? Well if ab pZ then p|ab. This means that p|a or p|b and hence a pZ or
20

b pZ. On the other hand if n N is not prime, then n = st for some positive integers s and t
such that neither is 1 or n. Then n does not divide s or t and neither s nor t is in nZ. However,
st = n is in nZ. Thus nZ is not a prime ideal.
Example. Suppose that I is an ideal of Z and 2Z I Z. If I 6= 2Z then I contains an odd
number 2n + 1. Since 2Z I, 2n I as well. As I is an ideal, it is closed under subtraction, and
hence 1 = 2n + 1 2n is in I. Consequently I = Z. This shows that there are no ideals between
2Z and Z. On the other hand, there is an ideal between 4Z and Z. In particular, 4Z 2Z Z.
Definition 13. Let R be a commutative ring. An ideal M of R is called a maximal ideal if M is
a proper ideal and if whenever I is an ideal of R with M I R, either I = M or I = R.
So a maximal ideal has the property that there arent any ideals between it and the whole ring.
Example. 2Z is a maximal ideal of Z and 4Z is not a maximal ideal of Z.
Example. Consider the ideals I = {f (x) Z[x] : f (0) = 0} and J = {f (x) Z[x] : f (0) 2Z}
of Z. Which of these is prime? maximal?
Suppose that f (x)g(x) I. Then 0 = (f g)(0) = f (0)g(0). Since f (0), g(0) Z and Z has no zero
divisors, we must have f (0) = 0 or g(0) = 0. Thus f (x) I or g(x) I, which shows that I is a
prime ideal of Z[x]. However, I is not a maximal ideal since I J Z[x].
Suppose that f (x)g(x) J. Then (f g)(0) = f (0)g(0) is even. The only way a product of two
integers can be even is if at least one of them is even. Thus either f (0) 2Z or g(0) 2Z, and
hence f (x) J or g(x) J. This shows that J is prime. Now suppose that K is an ideal of Z[x]
and J is a proper subset of K. Then there is a polynomial s(x) K such that s(x)
/ J. This
means that s(x) has an odd constant term. Now, s(x) + 1 J K. Since K is closed under
subtraction 1 = s(x) + 1 s(x) K. Thus K = Z[x], and J is maximal.
Theorem 3.11. Let R be a commutative ring with unity and let I be an ideal of R. Then R/I is
an integral domain if and only if I is prime.
Proof. Suppose that R is a commutative ring with unity. Now assume that R/I is an integral
domain. Now we need to show that I is prime. First we need to know that I is proper. Note that
R/I has at least two elements (a zero element and a unity), and hence I must not equal all of R.
Suppose that a, b R and ab I. Now a + I, b + I R/I. Note that (a + I)(b + I) = ab + I = I.
Since R/I has no zero divisors, either a + I = I or b + I = I. Thus a I or b I and I is prime.
Now assume that I is a prime ideal of R. Since R is a commutative ring with unity, R/I is
also a commutative ring with unity. We just need to show that R/I has no zero divisors. Let
a + I, b + I R/I and suppose that (a + I)(b + I) = I. Then ab + I = I and ab I. Since I is
prime, either a I or b I. Thus a + I = I or b + I which shows that R/I has no zero divisors.
Thus R/I is an integral domain.
Theorem 3.12. Let R be a commutative ring with unity and let I be an ideal of R. Then R/I is
a field if and only if I is maximal.
Proof. Let R be a commutative ring with unity and let I be an ideal of R. First suppose that R/I
is a field. Thus R/I has at least two elements and hence I 6= R. Now suppose that J is an ideal
and I J R. Suppose further that I 6= J. Then there is an element a J such that a
/ I.
21

Then a + I 6= I. Since R/I is a field, a + I has an inverse r + I. Now ar + I = (a + I)(r + I) = 1 + I.


Thus ar 1 I J. Since a J, ar J. Thus ar (ar 1) = 1 must be in J. This shows that
J = R and hence I is maximal.
Now suppose that I is maximal. Since R is commutative with unity, R/I is also commutative with
unity. We just need to show that every nonzero element of R/I is a unit. Suppose that a + I 6= I.
Then a
/ I. Consider the set J = {ar + s : r R, s I}. First we show that J is an ideal of
R. Since 0 I R and 0 = a0 + 0 J, J 6= . Now suppose that ar1 + s1 , ar2 + s2 J. Then
ar1 + s1 (ar2 + s2 ) = a(r1 r2 ) + (s1 s2 ) J since r1 r2 R and s1 s2 I. Also if ar + s J
and x R, then (ar + s)x = a(rx) + sx J since rx R and sx I. Thus J is an ideal. Now
notice that for any i I, i = a0 + i J, and hence I J. Also since a = a1 + 0 J and a
/ I,
I 6= J. Since I is maximal, we must have J = R. Thus 1 J which means that 1 = ar + s for
some r R and some s I. Now (a + I)(r + I) = ar + I = 1 s + I = 1 + I since s I. Thus
a + I is a unit and R/I is a field.
Example. We have seen that Z[x]/hxi
= Z. Since Z is an integral domain, hxi is a prime ideal of
Z[x]. Likewise, since Z is not a field, hxi is not a maximal ideal of Z[x].
Example. We know that Z/nZ
= Zn . Since Zn is a field if and only if n is prime, we see that nZ
is a maximal ideal if and only if n is prime.
Example. We have seen that (Z Z)/(Z {0})
= Z, so Z {0} is prime but not maximal.
Corollary 3.13. Let R be a commutative ring with unity. Then every maximal ideal is prime.
Proof. Suppose that R is a commutative ring with unity and A is a maximal ideal of R. Then R/A
is a field, and hence R/A is an integral domain. Thus A is prime.
We have seen in the examples that the converse is not true. There are many ideals that are prime
but not maximal. However, there are some cases where the converse holds.
Example. Suppose that R is a finite commutative ring with unity. Prove that every prime ideal
is maximal.
Suppose that R is a finite commutative ring with unity and A is a prime ideal. Then R/A is a
finite integral domain and hence a field. Thus A is maximal.
Example. Prove or disprove that I = {a + bi : a, b 2Z} is a prime ideal of Z[i].
Note that 1 + i, 1 i
/ I but (1 + i)(1 i) = 2 I. Thus I is not a prime ideal of Z[i].
Example. Prove or disprove that I = {(3x, y) : x, y Z} is a maximal ideal of Z Z.
Suppose that J is an ideal of Z Z and I J. Then there is an element (a, b) J such that
(a, b)
/ I. Thus a = 3k + 1 or a = 3k + 2 for some k Z.
Case 1: a = 3k + 1. The element (3k, b 1) I J. Thus (a, b) (3k, b 1) is in J. This shows
that (1, 1) J, and hence J = Z Z, which shows that I is maximal.
Case 2: a = 3k + 2. In this case look at the element (3k, b 2) which is in I and therefore
in J. Now (a, b) (3k, b 2) J, which means that (2, 2) J. Since (3, 3 I J, then
(1, 1) = (3, 3) (2, 2) J, which again shows that J = Z Z, and hence I is maximal.
22

Example. Prove or disprove: the intersection of two prime ideals is a prime ideal.
Example. Let A = {a + bi : a, b Z, a mod2 = b mod2}. Show that A is a maximal ideal of Z[i].
How many elements does Z[i]/A have?

4
4.1

Polynomial Rings
Basic Definitions and Results

Definition 14. Let R be a commutative ring. The set R[x] = {an xn + an1 xn1 + + a1 x +
a0 : ai R} is called the ring of polynomials over R with indeterminate x. The elements
a0 , a1 , . . . , an are called coefficients, and an is called the leading coefficient of the polynomial an xn +
an1 xn1 + + a1 x + a0 . If the leading coefficient of a polynomial is 1, the polynomial is called
monic. If n is the greatest nonnegative number for which an 6= 0, we say that the degree of
an xn + an1 xn1 + + a1 x + a0 is n. If no such n exists (i.e. the polynomial is 0), we say the
polynomial has no degree.
Note: Two polynomials an xn + an1 xn1 + + a1 x + a0 and bm xm + b1 x + b0 are equal if and
only if m = n and ai = bi for all i.
It is left as an exercise to show that R[x] is a ring with standard addition and multiplication of
polynomials.
Theorem 4.1. Suppose that R is an integral domain. Then R[x] is an integral domain. Also, if
degf (x) = n and degg(x) = m, then degf (x)g(x) = m + n.
Proof. It should be clear that if R is commutative then R[x] is also commutative. Similarly, if 1
is the unity of R, then 1 is also the unity of R[x]. We just need to show that R[x] has no zero
divisors. Suppose that f (x) = an xn + an1 xn1 + + a1 x + a0 and g(x) = bm xm + b1 x + b0 are
nonzero polynomials. Then the leading coefficients an and bm are nonzero. The leading coefficient
of f (x)g(x) is an bm which cannot be 0 since R has no zero divisors. This shows that the degree of
f (x)g(x) is m + n. It also shows that the product of f (x) and g(x) is nonzero and R[x] has no zero
divisors. This shows that R[x] is an integral domain.
Example. Find all units in Z3 [x]. The ring Z3 is a field, so deg(f g) = deg(f ) deg(g). Now if
f (x) g(x) = 1, then deg(f ) + deg(g) = 0. This means f and g both have degree 0, so they are
constant polynomials. The constants that are units are 1 and 2.
Example. If D is an ID, then D[x] is also an ID. Is this true for fields? If R is a field, is R[x] a
field? [Hint: look at the last example.]

23

4.2

The Division Algorithm

We know a division algorithm for integers: If m, n Z, and n > 0, then there exist unique integers
q, r Z such that m = qn + r and 0 r < n. We generally find q and r by long division. We can
also use long division on polynomials.
Theorem 4.2. Let F be a field and let f (x), g(x) F [x] with g(x) 6= 0. Then there exist unique
polynomials q(x), r(x) F [x] such that f (x) = q(x)g(x)+r(x) and either r(x) = 0 or deg r(x) <deg
g(x).
Proof. First we show that q(x) and r(x) exist. Note that if f (x) = 0 or if deg f (x) < deg g(x),
then we can just use q(x) = 0 and r(x) = f (x). So we will assume that deg f (x) deg g(x). Let
f (x) = an xn + an1 xn1 + + a1 x + a0
and
g(x) = bm xm + bm1 xm1 + + b1 x + b0
with an 6= 0 6= bm . By assumption, n m.
Consider all functions of the form f (x) g(x)s(x) such that s(x) F [x].
Case 1 There is a q(x) F [x] such that 0 = f (x) g(x)q(x). Taking r(x) = 0 we are done.
Case 2 There is no q(x) F [x] such that 0 = f (x) g(x)q(x) Then for all s(x) F [x], f (x)
g(x)s(x) 6= 0. Consider all functions of this form, and choose one, say r(x), that has minimal
degree. We know that r(x) = f (x) q(x)g(x) for some q(x) F [x]. We can rewrite this as
f (x) = q(x)g(x) + r(x).
Now r(x) is of minimal degree, so any other function of the form f (x) g(x)s(x) must have
degree deg(r(x)).
We need to show that deg r(x) < m. We know that
r(x) = ct xt + + c1 x + c0 .
By way of contradiction, suppose that t m. Then
tm
tm
f (x) q(x)g(x) ct b1
g(x) = r(x) ct b1
g(x)
m x
m x

= r(x) (ct xt + terms of lower degree than t).


Notice that this polynomial has degree lower than that of r(x). Also, we can write this as
tm which is in S. This contradicts the minimality of the degree of
f (x) g(x)[q(x) + ct b1
m x
r(x). Thus deg r(x) < m.

We have now shown that there exist q(x) and r(x) that satisfy the conditions of the theorem.
24

Next we need to show uniqueness. Suppose that


f (x) = q1 (x)g(x) + r1 (x) = q2 (x)g(x) + r2 (x)
where qi (x) and ri (x) have the appropriate properties. Then
g(x)(q1 (x) q2 (x)) = r2 (x) r1 (x) .
{z
} |
{z
}
|
degree m

degree < m

If r2 (x) r1 (x) 6= 0, then it is a polynomial whose degree is less than the degree of g(x). However,
g(x)(q1 (x) q2 (x)) is a polynomial that is either 0 or has degree greater than or equal to the degree
of g(x). We cant have two polynomials be equal but have different degrees, so we must have both
sides equal to 0. So r2 (x) r1 (x) = 0 which implies that r2 (x) = r1 (x). This also means that
g(x)(q1 (x) q2 (x)) = 0 and since g(x) 6= 0, we must have q1 (x) = q2 (x).
Example. Divide f (x) = 3x3 + 2x + 1 by g(x) = 2x2 + 1 in Z7 [x]. We should get q(x) = 5x and
r(x) = 4x + 1.

4.3

Principal Ideal Domains

Definition 15. Let R be an integral domain. If f (x), g(x) R[x], we say that g(x) divides f (x)
if there is a polynomial h(x) R[x] such that f (x) = g(x)h(x). In this case we call g(x) a factor
of f (x). An element R is called a root or zero of f (x) if f () = 0. When F is a field, F ,
and f (x) F [x], we say that is a root of multiplicity k (k 1) if (x )k is a factor of f (x) but
(x )k+1 is not a factor of f (x).
Corollary 4.3 (The Factor Theorem). Let F be a field, F , and f (x) F [x]. Then is a root
of f (x) if and only if x is a factor of f (x).
Proof. First suppose that is a root of f (x). By the division algorithm, f (x) = q(x)(x ) + r(x)
where either r(x) = 0 or the degree of r(x) is less than the degree of x . Since the degree of
x is 1, this means that either r(x) = 0 or r(x) is constant, so either way, r(x) is constant. Now
0 = f ()()+r() = r(). Since r(x) is constant, r(x) = r() = 0. Thus the remainder is 0 and
f (x) = q(x)(x ), which shows that x is a factor of f (x). Now suppose that x is a factor
of f (x). Then f (x) = (x a)q(x) for some q(x) F [x]. Now f () = ( )q() = 0q() = 0, so
is a root of f (x).
Example. Consider the polynomial f (x) = x2 + 3x + 2. How many roots does this have?
in Z: x2 + 3x + 2 = (x + 1)(x + 2), so 1 and 2 are only roots.
in Z2 : f (0) = 0 and f (1) = 0 so 0 and 1 are the roots.
in Z6 : f (0) = 0, f (1) = 0, f (2) = 0, f (3) = 2, f (4) = 0, and f (5) = 0. So 0, 1, 2, 4, and 5 are
all roots. So it turns out that a polynomial of degree n can have more than n roots. When is the
standard theorem about a polynomial of degree n having at most n roots true?
Corollary 4.4. Let F be a field. A nonzero polynomial p(x) F [x] of degree n can have at most
n roots in F counting multiplicity.

25

Proof. We will use induction on the degree n. A polynomial of degree 0 is a nonzero constant
polynomial and has not roots.
Induction Hypothesis: Suppose that any polynomial of degree m with m < n has at most m roots
counting multiplicity.
Now let f (x) have degree n and let a F be a root of multiplicity k. Then f (x) = (x a)k q(x) and
q(a) 6= 0. Also, n = deg f (x) = deg (x a)k q(x) = k + deg q(x). Thus k n, and since k 1, deg
q(x) < n. If f (x) has no roots other than a, then counting multiplicity, f (x) has k roots and since
k n, were done. So suppose that f (x) has another root b 6= a. Then 0 = f (b) = (b a)k q(b).
Since we are in an integral domain and (b a) 6= 0, (b a)k 6= 0, and hence q(b) = 0. This shows
that b is a root of q(x). Since n k = deg q(x) is less than n, we know that q(x) has at most n k
roots counting multiplicity (by the induction hypothesis). Every root of f (x) other than a must be
a root of q(x). So f (x) can have at most the root a of multiplicity k plus the at most n k roots
of q(x). Thus f (x) has at most n k + k = n roots.
Definition 16. A principal ideal domain is an integral domain in which every ideal is principal,
i.e. every ideal has the form hai = {ar : r R} for some a R.
Example. We know that Z is a principal ideal domain, since all ideals look like nZ for some n Z.
We also know that Z[x] is not a PID since we have seen before that the ideal J = {f (x) Z[x] :
f (0) 2Z} is not principal.
Theorem 4.5. If F is a field, then F [x] is a principal ideal domain.
Proof. Since fields are integral domains, we know that F [x] is an integral domain. Now let I be
an ideal of F [x]. We need to show that I is principal. If I = {0}, then I = h0i. So suppose that
I 6= {0}. Then there is an element g(x) I of minimal degree. That is to say that if f (x) I
and f (x) 6= 0, then deg f (x) deg g(x). We want to show that I = hg(x)i. Since g(x) I, we
know that hg(x)i I by the sucking property. Now let f (x) I. By the division algorithm,
f (x) = g(x)q(x) + r(x) for some q(x), r(x) F [x] with either r(x) = 0 for deg r(x) <deg g(x).
Then r(x) = f (x) g(x)q(x) I (since f (x) and g(x) are in I). Since g(x) has minimal degree in
I, this tells us that r(x) must be 0. Thus f (x) = g(x)q(x) hg(x)i, which shows that I hg(x)i,
and hence I = hg(x)i. This shows that all ideals in F [x] are principal.

Factorization of Polynomials

Definition 17. Let R be an integral domain. A polynomial f (x) R[x] that is neither the zero
polynomial nor a unit is called irreducible over R if whenever f (x) can be written as a product
f (x) = g(x)h(x) with g(x), h(x) R[x], then g(x) or h(x) must be a unit. A nonzero, nonunit
element of R[x] that is not irreducible over R is called reducible over R.
Note: In a field (but not necessarily in an integral domain), F , a nonconstant polynomial f (x)
F [x] is irreducible if and only if f (x) cannot be written as a product of polynomials of lower degree.
(Remember that all nonzero constant polynomials are units in F [x].)

26

Example. Consider the polynomial f (x) = 3x2 + 3.


f (x) is reducible over Z since we can factor is as 3(x2 + 1) and neither 3 nor x2 + 1 is a unit in Z[x].
f (x) is irreducible over Q since the only way to factor it using rational coefficients is 3(x2 + 1) and
3 is a unit in Q[x]. (Well have a more thorough proof that it is irreducible after we have proved
Theorem 5.1.) f (x) is irreducible over R for the same reason.
f (x) is reducible over C since we can factor it as (3x + 3i)(x i)

5.1

Reducibility Tests

Theorem 5.1. Let F be a field. If f (x) F [x] and the degree of f (x) is 2 or 3, then f (x) is
reducible over F if and only if f (x) has a root in F .
Proof. Let F be a field and suppose that f (x) F has degree 2 or 3. Suppose further that f (x)
is reducible. Then f (x) = g(x)h(x) for some g(x), h(x) F [x] such that both g(x) and h(x) have
degrees that are less than the degree of f (x) and both are nonunits (nonconstant). Thus at least
one of g(x) or h(x) has degree 1. Without loss of generality, g(x) = ax + b for some a, b F . Then
g(a1 b) = a(a1 b) + b = 0, so a1 b is a root of g(x) and is hence a root of f (x).
Conversely, suppose that f (x) has a root in F ; say f (a) = 0. Then we know that x a is a factor
of f (x), and therefore f (x) is reducible over F .
Example. f (x) = x2 + 2 is irreducible over Q since it has no roots.
What about over Z3 ?
In Z3 : f (0) = 2, f (1) = 0, so 1 is a root, and f (x) is reducible over Z3 .
Example. For polynomials of degree 4 or higher, this theorem does not apply! For example,
f (x) = x4 + 10x2 + 9 is reducible over R even though it has no roots. It can be factored as
(x2 + 1)(x2 + 9)

5.2

Reducibility over Z

Definition 18. The content of a polynomial an xn + an1 xn1 + + a1 x + a0 Z[x] is the greatest
common divisor of the coefficients. A primitive polynomial in Z[x] is one whose content is 1.
Lemma 5.2 (Gausss Lemma). The product of two primitive polynomials is primitive.
Proof. Suppose that f (x) and g(x) are primitive but f (x)g(x) is not primitive. Then the content of f (x)g(x) is not 1 and consequently there is some prime p that divides the content. Let
f (x), g(x), f (x)g(x) Zp [x] be the polynomials obtained by reducing the coefficients of f (x), g(x),
and f (x)g(x) modulo p. Since the content of f (x)g(x) is divisible by p, 0 = f (x)g(x) = f (x)g(x).
Since Zp [x] is an integral domain, either f (x) = 0 or g(x) = 0. So p either divides all of the
coefficients of f (x) or it divides all of the coefficients of g(x). Thus at least one of f (x) and g(x)

27

has content divisible by p, which contradicts both of them being primitive. Thus f (x)g(x) must be
primitive.
We know that many polynomials are irreducible over Z but reducible over C or Q. But it turns
out that if f (x) is irreducible over Z then it is also irreducible over Q. So if we are concerned with
finding irreducible polynomials over Z or Q, we can limit ourselves to just looking at Z[x].
Theorem 5.3. Let f (x) Z[x]. If f (x) is reducible over Q, then it is reducible over Z.

Proof. Let f (x) Z[x] and suppose that f (x) is reducible over Q. Then because Q is a field,
f (x) = g(x)h(x)
for some g(x), h(x) Q[x] and both g(x) and h(x) have lower degree than the degree of f (x).
Without loss of generality, we may assume that f (x) is primitive. (Otherwise, divide f (x) and
g(x)h(x) by the content of f (x).) We want to write
f (x) = g1 (x)h1 (x)
where g1 , h1 Z[x]. Let a be the least common multiple of the denominators of g(x) and b the
least common multiple of the denominators of h(x). Then ag(x), bh(x) Z[x]. Also,
abf (x) = ag(x)bh(x).
Let c1 be the content of ag(x) and let c2 be the content of bh(x). Then ag(x) = c1 g1 (x) and bh(x) =
c2 h1 (x) for some primitive polynomials g1 (x) and h1 (x) in Z[x]. Now abf (x) = c1 c2 g1 (x)h1 (x). By
Gausss Lemma, g1 (x)h1 (x) is primitive, so the content of c1 c2 g1 (x)h1 (x) is c1 c2 . Likewise the
content of abf (x) is ab. Thus ab = c1 c2 , and hence
f (x) = g1 (x)h1 (x).
Moreover, both g(x) and g1 (x) have the same degree, and h(x) and h1 (x) have the same degree,
so we have written f (x) as the product of two polynomials in Z[x] whose degrees are less than the
degree of f (x). Thus f is reducible over Z.
Example. 60x2 4x 1 = (10x 5/3)(6x + 3/5). By following the algorithm above, we can show
that there is another way to factor it that uses only integers. a = 3, b = 5.
60x2 4x 1 = (10x 5/3)(6x + 3/5)
3 5(60x2 4x 1) = 3(10x 5/3)5(6x + 3/5)
15(60x2 4x 1) = (30x 5)(30x + 3)
15(60x2 4x 1) = 5(6x 1)3(10x + 1)
60x2 4x 1 = (6x 1)(10x + 1)

28

5.3

Mod p Test and Eisensteins Criterion

Theorem 5.4 (Mod p Irreducibility Test). Let p be a prime and suppose that f (x) Z[x] with the
degree of f (x) greater than or equal to 1. Let f (x) be the polynomial in Zp [x] obtained by reducing
all coefficients of f (x) modulo p. If f (x) is irreducible over Zp and the degree of f (x) is the same
as the degree of f (x), then f (x) is irreducible over Q.
Proof. Let p be a prime and suppose that f (x) Z[x] with the degree of f (x) greater than or equal
to 1. Let f (x) be the polynomial in Zp [x] obtained by reducing all coefficients of f (x) modulo p.
Suppose that f (x) is irreducible over Zp and the degree of f (x) is the same as the degree of f (x).
Now, by way of contradiction, suppose that f (x) is reducible over Q. The f (x) = g(x)h(x) for
some g(x), h(x) Q[x] with deg g(x) <deg f (x) and deg h(x) <deg f (x). By Theorem 5.3, we may
actually assume that g(x) and h(x) are in Z[x]. Let f , g, h be the polynomials in Zp [x] corresponding
to f, g, h. Since deg f (x) =deg f (x), we have that deg g(x) deg g(x) <deg f (x) =deg f (x) and deg
h(x) deg h(x) <deg f (x) =deg f (x). But since f (x) = g(x)h(x), this contradicts our assumption
that f (x) is irreducible over Zp .
Theorem 5.5 (Eisensteins Criterion). Let f (x) = an xn + an1 xn1 + + a0 Z[x]. If there is
a prime p such that p 6 |an , p|an1 , p|an2 , . . . , p|a0 , and p2 6 |a0 , then f (x) is irreducible over Q.
Proof. Suppose that hypotheses of Eisensteins Criterion hold but that f (x) is reducible over Q.
Then by Theorem 5.3, we can actually write f (x) = g(x)h(x) for some g(x), h(x) Z[x] such
that both g(x) and h(x) have lesser degree than f (x). Suppose that g(x) = bs xs + b0 and
h(x) = ct xt + + c0 . Note that a0 = b0 c0 and by hypothesis, p|a0 and p2 6 |a0 . Thus p must divide
one of b0 and c0 , but it cannot divide both. Without loss of generality, say that p|b0 and p 6 |c0 .
Also, an = bs ct , and since p 6 |an , we must have that p 6 |bs and p 6 |ct . Let r be the least integer such
that p 6 |br . Now ar = br c0 + br1 c1 + + b0 cr . We know that p divides ar , and by our choice of r,
p also divides b0 cr , b1 cr1 , . . . , br1 c1 . Consequently p must also divide br c0 . Since p is prime, this
implies that p|br or p|c0 , which is a contradiction. Thus f (x) must be irreducible over Q.
Example. These problems can be solved either by the Mod p Test or Eisensteins Criterion.
1. Prove that f (x) = 10x3 3x2 + 2x + 18 is irreducible over Q.
Eisensteins Criterion is not applicable here, so well try the Mod p test. Note that mod 2
wont word as the degree of f would be less than the degree of f . So well try mod 3.
In Z3 [x], f = x3 + 2x and f (0) = 0, so f (x) has a root in Z3 . Thus f (x) is reducible in Z3
and the test is inconclusive.
We cant use mod 5 since again the degree of f would be less than the degree of f . So we try
mod 7.
In Z7 [x], f (x) = 3x3 + 4x2 + 2x + 4.
f (0) = 4
f (1) = 6
f (2) = 6
f (3) = 1

29

f (4) = 2
f (5) = 6
f (6) = 3
Thus f (x) has no roots in Z7 . Since the degree of f (x) is 3, this means that f (x) is irreducible
over Z7 . Thus f (x) is irreducible over Q.
2. Prove that f (x) = 10x9 + 14x7 49x3 + 21x + 35 is irreducible over Q. Use Eisensteins with
p = 7. Note that 7 does not divide 10, 7 does divide 14, 49, 21, and 35, and 72 = 49 does
not divide 35. Thus f (x) is irreducible over Q by Eisenstein.
3. Prove that f (x) = x4 + 3x3 + 5x2 + x + 9 is irreducible over Q. Well try the mod p test with
p = 2.
In Z2 , f (x) = x4 +x3 +x2 +x+1. Note that f (0) = 1 and f (1) = 1, so f (x) has no roots in Z2 .
Thus f (x) has no linear factors in Z2 , which also means that it has no cubic factors. So the
only way that f (x) could possibly factor is as the product of two irreducible quadratic factors.
So what are the irreducible quadratic factors in Z2 ?
The quadratics in Z2 [x] are f1 (x) = x2 , f2 (x) = x2 + x, f3 (x) = x2 + x + 1, f4 (x) = x2 + 1.
Note that f1 (0) = 0, f2 (0) = 0, and f4 (1) = 0, so f1 (x), f2 (x), and f4 (x) all have roots and
are therefore reducible over Z2 . Also note that f3 (0) = 1 and f3 (1) = 1, so f3 (x) has no roots.
Since f3 (x) has degree 2, this means that f3 (x) is irreducible over Z2 .
So we just need to check whether x2 + x + 1 is a factor of f (x). Dividing x4 + x3 + x2 + x + 1
by x2 + x + 1 in Z2 [x] yields a remainder of x + 1. Thus x2 + x + 1 is not a factor of f (x) and
hence f (x) has no irreducible quadratic factors. Thus f (x) is irreducible over Z2 , and hence
f (x) is irreducible over Q.
9
4. Prove that f (x) = 73 x4 72 x2 + 35
x + 35 is irreducible over Q. We start by multiplying through
by 35 to get h = 35f . Then h Z[x] so we can apply our tests. If f were reducible over Q
then h would be as well, so if h is irreducible, then so is f . Apply the Mod 2 test to h.

5. The pth cyclotomic polynomial p (x) =

xp 1
x1

is irreducible over Q. To see this, we consider

(x + 1)p 1
(x + 1) 1
1 p
= [x + pxp1 + p(p 1)xp2 + px + 1 1]
x
1
= [xp + pxp1 + p(p 1)xp2 + px]
x
= xp + pxp1 + p(p 1)xp2 + px.

p (x + 1) =

We apply Eisensteins Criterion to show that p (x + 1) is irreducible. Now if p (x) were


reducible, then p (x) = g(x)h(x). But then we could factor p (x + 1) = g(x + 1)h(x + 1),
which would be a contradiction. So p (x) is irreducible.

30

5.4

Applications of irreducible polynomials

Theorem 5.6. Let F be a field and let p(x) F [x]. Then hp(x)i is a maximal ideal of F [x] if and
only if p(x) is irreducible over F .
Proof. First suppose that hp(x)i is a maximal ideal of F [x]. Then hp(x)i is a proper ideal, and
hence p(x) is not a unit. Also, p(x) 6= 0 since {0} is not a maximal ideal of F [x]. (Note that
{0} hxi F [x].) Now suppose that p(x) = a(x)b(x) for some a(x), b(x) F [x] and suppose
that a(x) is not a unit. Then hp(x)i ha(x)i. Since hp(x)i is maximal, either ha(x)i = F [x]
or ha(x)i = hp(x)i. Since a(x) is not a unit, hp(x)i 6= F [x]. Thus ha(x)i = hp(x)i, and hence
a(x) = p(x)q(x) for some q(x) F [x]. Now p(x) = a(x)b(x) = p(x)q(x)b(x). Since F [x] is an
integral domain and p(x) 6= 0, we can cancel p(x), which yields 1 = q(x)b(x). Thus b(x) is a unit
and p(x) is irreducible.
Now suppose that p(x) is irreducible and suppose that hp(x)i J F [x] for some ideal J. Since
F is a field, F [x] is a principal ideal domain. Thus J = hj(x)i for some j(x) F [x]. Now
p(x) hp(x)i J = hj(x)i. Thus p(x) = j(x)k(x) for some k(x) F [x]. Since p(x) is irreducible,
either j(x) or k(x) must be a unit. If j(x) is a unit, then J = F [x]. If k(x) is a unit, then
j(x) = p(x)k(x)1 hp(x)i. By the sucking property, hj(x)i hp(x)i, and hence J = hp(x)i. Thus
J is either F [x] or hp(x)i, which shows that hp(x)i is maximal.
This gives us an easy way to construct fields with pk elements for some prime p. For example,
suppose we want to find a field with 25 elements. Note that Z25 is not a field. The idea is to try
to construct a field of polynomials of the form ax + b in Z5 [x]. Since there are 5 choices for each of
a and b, there are 25 elements. How do we construct such a field? We start with Z5 [x] and form
the factor ring by modding out by hp(x)i. In order to eliminate all powers of x greater than 1, the
polynomial p(x) must be quadratic. To make it a field, p(x) must be irreducible.
Consider the polynomial p(x) = x2 + 2 in Z5 [x]. Note that p(0) = 2, p(1) = 3, p(2) = 1, p(3) =
1, p(4) = 3, so p(x) has no roots. Since degp(x) = 2, this means that p(x) is irreducible over Z5 .
Thus F = Z5 [x]/hp(x)i is a field. To see that F has 25 elements, note that elements of F look like
h(x)+hx2 +2i. By the Division Algorithm, h(x) = q(x)(x2 +2)+r(x) where r(x) = 0 or degr(x) < 2.
Thus r(x) is of the form ax+b where a, b Z5 . Now h(x)+hx2 +2i = ax+b+q(x)(x2 +2)+hx2 +2i =
ax + b + hx2 + 2i. There are 5 choices for a and 5 choices for b, and hence F has 25 elements.

6
6.1

Factorization in Integral Domains


Definitions

Definition 19. Let R be an integral domain. Elements a and b in R are called associates if a = ub
for some unit u R. A nonzero element a R is called irreducible if a is not a unit and if whenever
a = bc for some b, c R, then b or c is a unit. A nonzero element a R is called prime if a is not
a unit and if a|bc with b, c R implies a|b or a|c.

31

Theorem 6.1. Let R be an integral domain and let a, b, p R. Then


(1) a|b if and only if b hai.
(2) p 6= 0 is prime if and only if hpi is a prime ideal.
(3) a and b are associates if and only if hai = hbi.

Proof. (1) Note that a|b if and only if b = ac for some c R if and only if b hai.
(2) Suppose that p is prime and suppose that xy hpi. Then p|xy and since p is prime, either p|x
or p|y. Thus x or y is in hpi, which shows that hpi is prime. Now suppose that hpi is prime and
that p|xy. Then xy hpi. Since hpi is prime, either x or y is in hpi, and hence p|x or p|y. Thus p
is prime.
(3) Suppose that a and b are associates. Then a = ub for some unit u. Now a hbi. Also b = au1
and hence b hai. By the sucking property, we have hai hbi and hbi hai and hence hai = hbi.
Conversely suppose that hai = hbi. Then a = bc for some c R and b = ad for some d R. Now
a = bc = adc. If a = 0, then b = 0d = 0 and hence hai = h0i = hbi. If a 6= 0, then since we are in
an integral domain, we may cancel, which yields 1 = dc. Thus c is a unit, which shows that a and
b are associates.
Theorem 6.2. In an integral domain every prime is irreducible.
Proof. Let R be an integral domain and suppose that a R is prime. Now suppose that a = xy.
Then a|xy and since a is prime, either a|x or a|y. If a|x, then x = aq for some q R. Then
a = xy = aqy. Since a is prime, a 6= 0, and hence we may cancel a (were in an integral domain).
Thus 1 = qy. This shows that y is a unit. A similar argument shows that if a|y then x is a unit.
Thus either x or y is a unit, and hence p is irreducible.
Theorem 6.3. In a principal ideal domain, an element is prime if and only if it is irreducible.
Proof. We already know that in an integral domain every prime is irreducible, so we just need to
show that every irreducible element is prime. Suppose that R is a principal ideal domain and a R
is irreducible. We will show that the ideal hai is maximal. Suppose that hai J R for some ideal
J. Since R is a PID, J = hji for some j R. Now p J = hji so p = jq for some q R. Since
p is irreducible, either j or q must be a unit. If j is a unit, then J = hji = R. If q is a unit, then
j = pq 1 hpi, which shows that hji hpi which shows that J = hpi. Thus J = R or J = hpi, and
hence hpi is maximal. Since all maximal ideals are prime, hpi is prime, and hence p is prime.

6.2

Examples in Z[ d]

We will now look at rings of the form Z[ d] = {a + b d : a, b Z} where d is a square free integer.
(By square free we mean that p2 6 |d for all primes p.)
We begin by defining the following function N , which we will call a norm.

32


N : Z[ d] N {0}

a + b d 7 |a2 db2 |
Lemma 6.4. The function N satisfies the
following properties:
(1) N (xy) = N (x)N (y) for all x, y Z[ d].
(2) x is a unit if and only if N (x) = 1.

(3)If N (x) is prime then x is irreducible in Z[ d].


(4) N (x) = 0 iff x = 0.
(5) If x and y are associates, then N (x) = N (y).

Proof. (1) Suppose x = a + b d and y = c + e d. Then N (xy) = N (ac + bed + (ae + bc) d) =
|(ac + bed)2 d(ae + bc)2 | = |a2 c2 + 2abcde+ b2 d2 a2
e2 d 2abcde b2 c2 d| = |a2 c2 + b2 d2 a2 e2 d
b2 c2 d| = |(a2 db2 )(c2 de2 )| = N (a + b d)N (c + e d)
(2) Suppose that x is a unit. Then xy = 1 for some y. Now N (x)N (y) = N (xy) = N (1) = 1. Since
N (x) and N (y) are natural numbers,
we must have that both N (x) and N (y) are 1.

2
2
2
2
Now suppose that x = a + b d and N (x) = 1. Then
|a db | =1. Thus a db = 1 or
2
2
2
2
a db = 1. If a db =1, then (a +
a + b d is a unit. Alternatively,
b d)(a b d) = 1, so
2
2
if a db = 1, then (a + b d)(a + b d) = 1, and again a + b d is a unit.
(3) Suppose that N (x) is prime and suppose that x = yz. Then N (x) = N (yz) = N (y)N (z). Since
N (x) is prime, one
be 1 and hence y or z is a unit.
Thus x is irreducible.
of N (y) and N (z) must

a
2
2
(4) Let x = a + b d. If N (x) = 0 then a db = 0, which means that b = d or b = 0. Since d
is irrational, this can only happen if b = 0. And then a = 0 as well.
(5) Suppose that x and y are associates. Then x = yu for some unit u. Now N (x) = N (yu) =
N (y)N (u) = N (y) 1 = N (y).

Example. Consider Z[ 3].


1. Find a unit.

1. A little playing
An element a + b 3 is a unit iff a2 3b2 = 1. In other words a2 = 3b2

around with lists of perfect squares finds at least two candidates: 4 3 and 7 4 3.

2. Find an irreducible element.

We can find an element such that N (a + b 3) is prime. One example is 3 + 2 3.

3. If N (x) is not prime, that doesnt necessarily


mean that x is irreducible. For example,

N (3 + 3) = 6, but I to figure out if 3 + 3 is reducible or not, Iwould just


have to tryto
factor it. After a large amount of trial and error, I find that 3 + 3 = (3 + 2 3)(1 + 3)
and neither factor is a unit.

Example. Consider 3 Z[ 5].


1. Show that 3 is irreducible.

33

Suppose that 3 = xy
Then 9 = N (3) = N (xy) = N (x)N (y)

So N (x) and N (y) must be 1, 3, or 9. Note that for any a + b 5, N (a + b 5) = a2 + 5b2 ,


which can never be 3 since a and b are integers. Thus either N (x) or N (y) must be 1, which
means that either x or y is a unit. Thus 3 is irreducible.
2. Show that 3 is not prime.

Note that 9 = (2 + 5)(2 5),so 3|(2 + 5)(2


5)

Note that 3 6 |(2 + 5) since if (2 + 5) =


3(a
+
b
5)
=
3a
+
3b
5, then 3a = 2, which

has no integer solutions. Similarly 3 6 |(2 5). Thus 3 is not prime.

Example. In Z[ 5] show that 7, 1 + 2 5, and 1 2 5 are all irreducible.


First suppose that 7= xy. Then 49 = N (xy) = N (x)N (y) and hence N (x) = 1, 7, or 49. The
norm function in Z[ 5] looks like a2 + 5b2 which can never be 7 (since a, b
Z), so either N (x)
or N (y) must be 1 and
either
x
or
y
must
be
a
unit.
Thus
7
is
irreducible
in
Z[
5].

Note that both 1 + 2 5 and 1 2 5 have norm 21, so if either


of
them
is
written
as xy, then

21 = N (xy) = N (x)N (y). Since we have seen that nothing in Z[ 5] canhave norm 3 or
7, either
N (x) or N (y) must be 1 and either x or y is a unit. Thus both 1 + 2 5 and 1 2 5 are
irreducible.
This
shows that 21 can be factored into irreducibles in two different ways: 3 7 and

(1 + 2 5)(1 2 5).

6.3

Unique Factorization Domains and Euclidean Domains

Definition 20. An integral domain R is a unique factorization domain (UFD) if (i) every nonzero
nonunit element r R can be written in the form r = c1 c2 cn with each ci irreducible.
(ii) if r = c1 c2 cn = d1 d2 dm with ci and dj irreducible, then n = m and each ci is an associate
of exactly one dj .
Example.
Z is a unique factorization domain.

Z[ 5] is not a unique factorization domain


we can factor 21 into irreducibles in two
because
different ways. To see that 3 7 and (1 + 2 5)(1

2
5) arenot equivalent factorizations we

need to show that 3 is not an associate of 1 + 2 5 or 1 2 5. Suppose 3 is an associate


of one of them, then 3u is equal to one of them for some unit u. Then N (3u) = 21 and hence
21 = N (3)N (u) = 9(1) which is a contradiction.
Definition 21. An integral domain D is called a Euclidean Domain if there is a function v called
the valuation map from D (the nonzero elements of D) to the nonnegative integers such that
(1) v(a) v(ab) for all a, b D . (2) If a, b D and b 6= 0, then there exist q, r D such that
a = bq + r and either r = 0 or v(r) < v(b).
Example. Z is a Euclidean domain with v(x) = |x|.
Example. If F is a field, then F [x] is a Euclidean domain with v(f (x)) = degf (x).
Example. Z[i] is a Euclidean domain with v(a + bi) = N (a + bi) = a2 + b2 :
(1) v(xy) = N (xy) = N (x)N (y) = v(x)v(y) v(x).
(2) Let x, y Z[i] with y 6= 0.
34

Consider xy Q[i]. Then xy = s + ti for some s, t Q. Let m, n Z such that |m s| 12 and


|n t| 12 . Now xy = s + ti = (m m + s) + (n n + t)i = (m + ni) + (s m) + (t n)i.
Thus x = (m + ni)y + ((s m) + (t n)i)y. Let q = (m + ni) and r = ((s m) + (t n)i)y =
x (m + ni)y Z[i]. Note that x = qy + r. Also, if r 6= 0, then v(r) = N (((s m) + (t n)i)y) =
((s m)2 + (t n)2 )N (y) ( 14 + 14 )v(y) < v(y).

6.4

Theorems

Theorem 6.5. Every principal ideal domain is a unique factorization domain.


Lemma 6.6 (Ascending Chain Condition for PIDs). In a Principal Ideal Domain, any strictly
increasing chain of ideals I1 I2 I3 must be finite in length.
Proof. Let R be a principal ideal domain and suppose that I1 I2 I3 is a strictly increasing
chain of ideals. Let I = Ij . First we show that I is an ideal. Note that since all of the Ij s are
nonempty, I is nonempty. Now suppose that a, b I. Then a Ik and b Im for some k and
m. Without loss of generality, m k. Then a, b Im and since Im is an ideal, a b Im I.
Similarly if a I and r R, then a Im for some m and since Im is an ideal, ra Im I. Thus
I is an ideal. Now since R is a principal ideal domain, I = hxi for some x R. Then x In for
some n, and hence I = hxi In . Since In is clearly a subset of I, we must have that I = In . Thus
the chain of ideals must terminate at In .
Proof of Theorem 6.5: Suppose that R is a PID and let a0 be any nonzero nonunit in R.
We start by showing that a0 is a product or irreducibles (the product might have only one factor).
First we show that a0 has at least one irreducible factor. If a0 is irreducible were done. Thus we
can assume that a0 = b1 a1 , where neither b1 nor a1 is a unit and a1 is nonzero. If a1 is irreducible,
were done. If not, a1 = b2 a2 where neither a2 nor b2 is a unit and a2 is nonzero. Continuing in this
fashion, we get a sequence b1 , b2 , . . . of nonunits and a sequence a1 , a2 , . . . of nonzero elements such
that an = bn+1 an+1 for each n. Hence (a0 ) (a1 ) is a strictly increasing chain of ideals. By
the previous lemma, the chain must terminate at some (ar ), and ar must be an irreducible factor
of a0 . Thus a0 has at least one irreducible factor.
Now we can write a0 = p1 c1 where p1 is irreducible and c1 is not a unit. If c1 is not irreducible,
then we can write c1 = p2 c2 where p2 is irreducible and c2 is not a unit. Continuing in this way, we
obtain a strictly increasing chain of ideals (a0 ) (c1 ) (c2 ) , which must terminate at some
(cs ). Then cs is irreducible and a0 = p1 p2 p2 cs where each pi is irreducible. Thus every nonzero
nonunit is a product of irreducibles.
To see that the factorization is unique, suppose that some element a D can be written as
a = p1 p2 pr = q1 q2 qs where the pi s and qj s are irreducible and repetition is permitted. We
use induction on r. If r = 1, then a is irreducible and, clearly, s = 1 and p1 = q1 . We may now
assume that any element that can be expressed as a product of fewer than r irreducibles can be so
expressed in only one way (up to order and associates). Since p1 divides q1 q2 qs , it must divide
35

some qi (irreducible in a PID implies prime). Without loss of generality, we may say that p1 |q1 .
Then q1 = p1 u where u R is a unit. Since q1 p2 pr = up1 p2 pr = uq1 q2 qs = q1 (uq2 ) qs ,
we have by cancellation that p2 pr = uq2 qs . The induction hypothesis now tells us that these
two factorizations are identical up to associates and the order in which the factors appear. Hence
the same is true about the two factorizations of a.
Corollary 6.7. If F is a field, then F [x] is a unique factorization domain.
Proof. If F is a field, then F [x] is a principal ideal domain and hence a unique factorization
domain.
Theorem 6.8. Every Euclidean Domain is a Principal Ideal Domain.
Proof. Let R be a Euclidean Domain and let I be an ideal of R. If I = {0} then I is principal, so
we may assume that I is a nonzero ideal. Let g I be such that v(g) v(x) for all x I . Since
g I, we see that hgi I. To see containment the other way, let x I. Then x = qg + r for some
q, r R and either r = 0 or v(r) < v(g). Note that r = x qg I and hence v(r) v(g). Thus
r = 0 and x = qg hgi. Hence I = hgi and every ideal of R is principal.

36

Index
associates, 31

subring, 7

characteristic of a ring, 12
coefficients, 23
commutative ring, 4
content of a polynomial, 27

unique factorization domain, 34


unit, 4
unity, 4
valuation map, 34

degree of a polynomial, 23
divides, 6, 25
division ring, 10

zero divisor, 6
zero product property, 9

euclidean domain, 34
evaluation homomorphism, 15
factor, 25
field, 10
ideal, 16
idempotent, 6
integral domain, 9
irreducible, 26, 31
isomorphic, 14
kernel, 14
leading coefficient, 23
maximal ideal, 21
monic polynomial, 23
nilpotent, 6
norm, 32
polynomial ring, 23
prime, 31
prime ideal, 20
principal ideal, 17
principal ideal domain, 26
principal ideal generated by a, 17
reducible, 26
ring, 4
ring homomorphism, 14
ring isomorphism, 14
ring with identity, 5
ring with unity, 5
root, 25
root of multiplicity k, 25
37

You might also like